Introduction

Charge-density-wave (CDW) and conventional superconductivity (SC) are both caused by strong electron-phonon coupling and Fermi surface (FS) instabilities, and thus the interplay between these two collective electronic phenomena has been a subject of extensive investigations over the past decades1. The former usually occurs in the low-dimensional materials and involves a periodic modulation of the conduction electron density and the underlying lattice; this will open a gap over part of the FS and thus lead to reduced conductivity. In contrast, bulk SC is a macroscopic quantum phenomenon that requires phase coherence in three dimensions and it features an FS gap that results in infinite conductivity. As such, these two electronic orders are usually competing with each other. It is frequently observed in the low-dimensional CDW conductors that suppression of CDW order leads to a bulk SC. Here, we report on a rare case that an intrinsic quasi-one-dimensional (Q1D) filamentary SC emerges and persists in a relatively wide temperature/pressure range when the CDW order of HfTe3 along the same crystallographic direction is completely suppressed by pressure.

As shown in Fig. 1a, HfTe3 crystallizes in the TiS3-type structure with monoclinic space group P21/m, featured by two kinds of chain-like units, i.e., the infinite prismatic (HfTe3) chains along the b-axis and the almost linear chains of -Te2-Te3- with alternating short (2.80 Å) and long (3.07 Å) distances along the a-axis2. The two-dimensional (2D) prism slabs in the ab plane are stacked along the c-axis through weak van der Waals (vdW) interactions. These quasi-1D and 2D structural characters have a profound impact on the physical properties of HfTe3 at ambient pressure. Resistivity measurements on high-quality HfTe3 single crystals reveal pronounced anisotropic behaviors and obvious anomalies in both ρa(T) and ρc(T) but a very weak kink-like anomaly in ρb(T) at around 93 K2. In analog with the sister compound ZrTe33,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19, the transport anomaly at TCDW = 93 K has been ascribed to the formation of a CDW order within the ac-plane6. Since the ρc is about two orders of magnitude higher than ρa, the CDW should have a dominant 1D character along the -Te2-Te3- infinite chains. However, the wavevector of the CDW order has not been determined experimentally.

Fig. 1: Crystal structure and transport property at ambient pressure.
figure 1

a Crystal structure of MTe3 (M = Zr, Hf). b X-ray diffraction pattern of the single-crystal HfTe3 showing only the (00 l) diffraction peaks. The inset shows the FWHM of the (006) Bragg peak. c Temperature dependence of the a- and b-axis normalized resistance R(T)/R300 K measured on the same piece of a sample at ambient pressure. Displayed in the inset of (c) is the photo of the crystal, which shows fringes extending along the b-axis.

For high-quality single-crystalline HfTe3 samples, no SC was detected down to 50 mK at ambient pressure2. In contrast, the polycrystalline HfTe3 sample shows the coexistence of the CDW order below TCDW ≈ 80 K and filamentary SC with Tc ≈ 2 K at ambient pressure2. The resistivity anomaly around TCDW is much reduced in comparison with that of a single crystal, implying that the emergence of SC should be correlated with the suppression of CDW, presumably due to the presence of chemical disorders or defects in the polycrystalline samples. However, the nature of the filamentary SC without magnetic and thermodynamic signatures and its relationship with CDW remains elusive so far. For the polycrystalline sample, Denholme et al.20 have performed preliminary high-pressure (HP) resistivity measurements up to ~1 GPa; they observed pressure-induced enhancement of the CDW transition, but the evolution of SC and its relationship with respect to CDW was not reported20. Considering the strong anisotropy of the physical properties associated with the Q1D nature of the crystal structure, it is imperative to carry out a comprehensive HP study on single-crystalline samples in order to unveil the nature of the filamentary SC and its interplay with CDW.

In this work, we first determined experimentally the CDW vector q = 0.91(1) a* + 0.27(1) c* of HfTe3 by low-temperature transmission electron microscope (TEM) at ambient pressure, and then performed a comprehensive HP transport study on the HfTe3 single crystals along three major crystallographic axes. Intriguingly, we observed the emergence of an intrinsic filamentary SC in Ra(T) at T ≤ 4–5 K when the CDW of HfTe3 is suppressed completely by P ≥ 5 GPa, whereas no clear sign of SC is observed down to 1.4 K in Rb(T) and Rc(T), demonstrating an extraordinary case of Q1D SC rarely seen in real materials. We have analyzed the excess conductivity according to the theory of Aslamazov and Larkin (A-L)21 and compared our results with those of ZrTe33,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,22. Finally, we discussed the observed Q1D SC in terms of the anisotropic FSs and the local pairs formed along the -Te2-Te3- chains parallel to the a-axis based on the first-principles electronic structure calculations.

Results

Characterizations at ambient pressure

The single-crystal XRD pattern of HfTe3 shown in Fig. 1b displays only the sharp (00 l) diffraction peaks, which confirm that the single crystals are grown along the ab plane and stacked along the c-axis. The c-axis lattice constant is determined to be c = 10.0519 Å, in consistent with the previous report2. As shown in the inset of Fig. 1b, the full width at half maximum (FWHM) of the (006) Bragg peak is about 0.2°, implying a high quality of the crystals. Within the plate-like crystal surface, the a and b axes are determined to be perpendicular and parallel to the crystal edges, respectively, by the Laue back diffraction. As shown in the inset of Fig. 1c, the HfTe3 single crystal shows fringes extending along the b-axis, which is parallel to the crystal edge. Thus, the a- and b-axis that are orthogonal with each other within the ab plane can be distinguished visually for the as-grown crystals. For this sample, we measured the temperature dependence of anisotropic resistance R(T) along the a- and b-axis, respectively. As seen in Fig. 1c, the a-axis resistance exhibits a pronounced hump anomaly at TCDW = 93 K associated with the CDW transition, while the b-axis resistance shows a very weak kink-like anomaly at TCDW without any trace of superconducting transition down to 1.8 K. These results agree well with the previous report2. The residual resistivity ratio, RRR ≡ R(300 K)/R(2 K) along the b-axis is calculated to be 62, which also highlights the high quality of the studied single-crystal sample. We have thus selected samples with similar RRR values for the HP measurements as shown in Supplementary Fig. 1.

Although it is expected that HfTe3 should form a similar CDW as that of ZrTe3 with q = 0.93 a* + 0.33 c*, the q-vector of HfTe3 has not been determined experimentally so far. Here we employed a low-temperature TEM to directly determine the CDW vector of the HfTe3 single crystal. Figure 2a, b show the selected area electron diffraction (SAED) patterns of HfTe3 taken along the [001] zone axis at room temperature and 26 K, respectively. The pristine lattice at room temperature has a monoclinic structure with space group P21/m, and the satellite spots can only be identified below the CDW phase transition temperature. As shown in the inset of Fig. 2b, the intensity of the satellite spot goes up as the scattering angle increases and reaches a maximum at 17.5 mrad, indicating that the structural modulation vector q has both a* and c* components, i.e., q = qa a* + qc c*. The extra Laue zones (indicated by blue arrows) presented in the [001] CBED pattern, Fig. 2c, further confirm that the modulation vector does not lie in the ab plane.

Fig. 2: Electron diffraction patterns of HfTe3.
figure 2

Typically selected area electron diffraction (SAED) patterns taken along the [001] zone axis at a 295 K and b 26 K, respectively. The satellite spots (indicated by blue arrows) appear at low temperatures around the Bragg spots (green arrows), indicating the formation of structural modulation. c A convergent beam electron diffraction (CBED) pattern taken along the [001] zone axis, showing extra Laue zones (blue arrows) due to the structural modulation. The radial intensity distribution is plotted as an inset, with the positions of the first/second-order Laue zone (FOLZ/SOLZ) indicated by green arrows. Based on (b) and (c), the modulation vector can be determined to be q = 0.91(1) a* + 0.27(1) c*. d Intensity line profile from spot \((\bar 600)\) to \((\overline {10} 00)\) extracted from the SAED pattern (b). The intervals between different spots are labeled. e Upper panel: the geometric configuration showing the Ewald sphere intercepts points in the Laue zones. Lower panel: the intensity radial profile extracted from CBED pattern (c), from which the L values can be obtained.

In order to determine the value of qa and qc, the intensity distributions of CDW spots/rings in Fig. 2b, c are summarized in Fig. 2d, e, respectively. Figure 2d shows the line intensity profile from spot \((\bar 600)\) to \((\overline {10} 00)\), in which the intervals between adjacent Bragg spots (LBragg) and the distance from one Bragg spot to the next-nearest CDW neighbor (LCDW) are labeled. Using qa = LCDW/LBragg, qa is calculated to be 0.91(1). On the other hand, the value of qc can be estimated with the aid of the geometric configuration presented in Fig. 2e. The positions of the rings in Fig. 2c are directly related to the interception of the Ewald sphere with the Laue zone layers. For high-energy electrons, we can write an expression for the angle θ in Fig. 2e: \({{{\mathrm{sin}}}}\theta = \frac{{{\Delta}{{{\mathrm{K/}}}}2}}{r} = \frac{h}{{{\Delta}{{{\mathrm{K}}}}}}\). Using the equation\(({\Delta}{{{\mathrm{K}}}})^2 = h^2 + L^2\), we have \(h = r - \sqrt {r^2 - L^2}\), where r is the radius of the Ewald sphere (r = 398.74 nm−1 when the electron energy is 200 keV). The L values for different rings can be obtained from the radial profile plotted in the low panel of Fig. 2e. Then, by calculating h for different Laue zones and the ratios (for example qc = hCDW/hFOLZ, where hCDW and hFOLZ are the h values for the innermost ring and the FOLZ ring, respectively), qc is estimated to be 0.27(1). Thus, we can write q = 0.91(1) a* + 0.27(1) c*.

To further understand the temperature-dependent behavior of the structural modulation, we acquired SAED patterns at several different temperatures, as shown in Supplementary Fig. 2. It is evident that the superlattice spots (indicated by blue arrows) become dim at 81 K and totally disappear at 88 K, showing good agreement with the CDW phase transition temperature determined from the resistivity measurements. The result suggests that the superlattice spots observed in diffraction patterns are directly related to the formation of the CDW phase.

It is worth noting that the satellite spots appeared in the electron diffraction patterns of ZrTe3 at room temperature. Eaglesham et al.6 believed that the satellite spots at room temperature indicated the composition modulation of ZrTe3 caused by off-stoichiometry. This is also considered to be a strong evidence that the quality of ZrTe3 crystals is difficult to control. For the HfTe3 single crystal, we can see that the SAED patterns do not show any satellite spots at room temperature, which indicates that the crystal structure of the HfTe3 single crystal is perfect and closer to the stoichiometry.

Effect of pressure on the a-axis resistance

Figure 3a shows the temperature dependence of the a-axis resistance Ra(T) of the HfTe3 single crystal (#1) under various pressures up to 2.0 GPa measured with a PCC. Details about the experimental setup for anisotropic resistance measurements can be found in Supplementary Fig. 4. At ambient pressure, Ra(T) displays a hump anomaly due to the CDW formation at TCDW = 93 K, which can be determined from the minimum of dRa/dT as shown in the inset of Fig. 3a. With increasing pressure, TCDW first moves to higher temperatures, reaching the maximum of ~128 K at 1.5 GPa, and then decreases gradually to 122.4 K at 2.01 GPa. The jump of Ra(T) around TCDW is reduced gradually accompanying the suppression of CDW transition. Such a non-monotonic variation of TCDW(P) has also been observed in ZrTe312, and thus should have an origin in common. The initial enhancement of TCDW at low pressures indicates that the lattice compression first prompts the CDW formation, presumably due to the enhancement of electron-lattice coupling or the optimization of the nesting condition for the parallel FSs along the a*-axis. Nonetheless, the CDW should be eventually suppressed by pressure as observed in ZrTe312.

Fig. 3: The a-axis resistance under high pressure.
figure 3

a Temperature dependence of the a-axis resistance Ra(T) for sample #1 under various pressures up to 2.01 GPa measured with a piston-cylinder cell (PCC); the inset shows the dRa/dT of the data in (a) to determine the charge-density-wave transition temperature TCDW. b Ra(T) for sample #2 under various pressures from 2.4 to 12.1 GPa measured with a palm cubic anvil cell (CAC); the upper panel shows the dRa/dT curves at 2.4 and 4.0 GPa to determine the values of TCDW. c The low-temperature Ra(T) for sample #2 under various pressure from 5.4 to 12.1 GPa. d Ra(T) for sample #3 under various pressures from 2.5 to 12.5 GPa; the upper panel shows the dRa/dT curve at 2.5 GPa to determine the values of TCDW. e The low-temperature Ra(T) for sample #3 under various pressure from 5 to 12.5 GPa highlighting the evolution of the possible superconducting transition.

The evolution of the CDW transition at P > 2 GPa is further investigated by measuring Ra(T) of a second HfTe3 crystal (#2) with a CAC. The obtained Ra(T) and the selected dRa/dT curves are displayed in Fig. 3(b). The Ra(T) at 2.4 GPa is similar to that measured in PCC, and the TCDW has been further reduced to 93.8 K. When the pressure is increased to 4 GPa, the TCDW is quickly lowered to 43.3 K and the resistance anomaly at TCDW is significantly suppressed, as shown in Fig. 3b. At 5.4 GPa and above, no resistance anomaly can be discerned anymore at T > 5 K, implying the complete suppression of the CDW transition in HfTe3. In this pressure range, the magnitude of Ra(T) at high temperatures decreases gradually with increasing pressure while that at low temperatures keeps almost constant.

Interestingly, we observed a gradual resistance drop at low temperatures for P > 5.4 GPa, signaling the possible emergence of SC accompanying the elimination of CDW in HfTe3. As shown in Fig. 3c, the enlarged view of Ra(T) at T ≤ 6 K evidences a slight downturn feature below Tconset ≈ 3 K at 5.4 GPa. Upon further increasing pressure, the drop of resistance becomes stronger and the Tconset moves gradually to higher temperatures, reaching a nearly constant value of ~5 K at P ≥ 9.1 GPa. The zero resistance is finally achieved at Tczero ≈ 2-2.5 K at P ≥ 11.4 GPa. We found that the resistance drop in Ra(T) can be suppressed by applying either high electric currents and/or external magnetic fields, as seen in Fig. 4a, b). These observations are consistent with the occurrence of SC in HfTe3 at P ≥ 5.4 GPa when the CDW transition along the a-axis is suppressed completely. It should be noted that the resistance data became too noisy under magnetic fields higher than 0.3 T, which prevented us from determining precisely the upper critical field.

Fig. 4: The low-temperature a-axis resistance under high pressure.
figure 4

The low-temperature Ra(T) for sample #2 at 11.4 GPa was measured with a different currents and b magnetic fields. c The low-temperature Ra(T) for #3 at 11 GPa was measured with different currents. d The low-temperature Ra(T) for #3 at 12.5 GPa measured with different currents and/or magnetic fields.

In order to ensure the reproducibility of the experimental results, we performed HP resistance measurements on a third HfTe3 (#3) single crystal along the a-axis. As shown in Fig. 3d, e, the CDW transition at 93 K is clearly observed for Ra(T) at 2.5 GPa and is completely suppressed at 5 GPa; the SC transition is also confirmed at low temperature for P ≥ 5 GPa. It is noted that the observed Tc for this sample is slightly lower than that of sample #2, implying that the pressure-induced SC in HfTe3 is sensitive to the sample quality. As shown in Supplementary Fig. 5, the RRR values of sample #2 are indeed larger than those of sample #3 in the pressure range when CDW is suppressed completely. Accordingly, the Tconset of sample #2 is also higher than that of sample #3. Nevertheless, the results of sample #3 are in general agreement with those of the #2 sample. Furthermore, we also measured the low-temperature resistance at 11 and 12.5 GPa under different electric currents and external magnetic fields. As shown in Fig. 4c, d, the gradual suppression of the Ra(T) drop upon applying high electric currents and/or external magnetic fields further substantiates the emergence of pressure-induced SC in HfTe3 along the a-axis.

To verify if the observed SC is bulk or not, we have attempted to measure the ac magnetic susceptibility χ′(T) of HfTe3 together with a piece of Pb as a superconducting reference in the CAC up to 12 GPa. As shown in Supplementary Fig. 6, no obvious diamagnetic signal was observed for HfTe3 down to 1.4 K, except for the superconducting transition of Pb, which shifts to lower temperatures progressively with increasing pressure. The observed sharp transitions of Pb in Supplementary Fig. 6 elaborate an excellent hydrostatic pressure condition in the CAC at least up to 12 GPa. The absence of obvious diamagnetic response thus indicates a non-bulk or filamentary nature for the observed SC in HfTe3 at least down to 1.4 K. This observation motivated us to further investigate the effect of pressure on the b- and c-axis resistances, which lead to very surprising results as shown below.

Effect of pressure on the b-axis resistance

Because the samples #2 and #3 measured above are always disassembled into many filaments after decompression from high pressures, presumably due to the special sample microstructure and the presence of residual stress upon compression. This issue prevents the subsequent measurements along the b-axis direction on the identical samples. Under such a circumstance, we selected another two HfTe3 samples #4 and #5 for measurements of the b-axis resistance Rb(T) in two separated runs in the pressure range 7–12.5 GPa, where zero resistance can be achieved in Ra(T). The results of Rb(T) for these two samples are shown in Fig. 5a. As seen clearly, the Rb(T) curves of both samples do not exhibit any anomaly at the temperatures of ~4–5 K where Ra(T) starts to show the superconducting transition. For sample #5, we observed a slight downturn of Rb(T) starting at T < 2 K under P ≥ 10 GPa, Fig. 5b, which is likely caused by some misalignment of the electrical contact so that some a-axis component was picked up.

Fig. 5: The b- and c-axis resistance under high pressure.
figure 5

a The temperature dependence of b-axis resistance Rb(T) under different pressures from 7 to 12.5 GPa for two samples (#4 and #5) and b their low-temperature data. c The temperature dependence of c-axis resistance Rc(T) under different pressures from 2.5 to 11 GPa for sample #6 and d their low-temperature parts below 5 K. The inset of (c) shows the photo of the sample for the resistance measurement along c-axis.

These results are striking. First of all, they immediately exclude the possibility that the pressure-induced resistivity drop in Ra(T) arises from a secondary phase. Secondly, the observed SC in Ra(T) is very likely a Q1D phenomenon. As shown below, our simulations with a 2D resistor matrix can well reproduce the above Rb(T) results. To substantiate this conclusion, we further measured the c-axis resistance Rc(T) under various pressures.

Here, we want to point out that it is quite challenging experimentally to collect merely the b-axis resistance without any a-axis contribution for the resistance measurements on a bulk sample. This is particularly difficult given the fact that the a-axis resistance exhibits a pronounced drop around 4–5 K whereas the b-axis resistance almost levels off. Therefore, a slight misalignment of the voltage leads V+/V− with respect to the a-axis, i.e., they should be perfectly parallel with the a-axis, will inevitably introduce the a-axis feature, overshadowing the intrinsic response of b-axis resistance. Under such a circumstance, the chance to observe a resistivity drop along the b-axis is high, but it should be convincing if we can observe no obvious resistivity drop by chance. To overcome this problem, we tried to first pattern the electrical contacts by evaporating the gold pads on the sample surface so as to make them as parallel to the a-axis as possible, see Supplementary Fig. 4b, then we attached the electrical leads with silver paste. In this way, we are able to observe no any resistance drop for sample #4 and a tiny drop for sample #5 as shown in Fig. 5b.

Effect of pressure on the c-axis resistance

We employed the configuration shown in the inset of Fig. 5c to measure precisely the Rc(T) of the plate-like HfTe3 crystal (#6). The electric current flows through the sample via the U-shaped I+/I− pads on the opposite sides of the crystal, while the voltage leads V+/V− are attached to the center pads on both sides. At ambient pressure, Rc(T) exhibits an upturn upon cooling through TCDW2. As seen in Fig. 5c, the resistance upturn below TCDW is quickly suppressed by pressure, and the Rc(T) at 2.4 GPa displays a metallic behavior in the whole temperature range with a clear kink-like anomaly at TCDW ≈ 96.5 K, which is consistent with that determined from Ra(T) in Fig. 3b. The CDW transition cannot be discerned anymore at P ≥ 4.2 GPa, implying the complete suppression of CDW by pressure. In the investigated pressure range, we observed no emergence of possible SC down to 1.4 K. Interestingly, the Rc(T) curves at P ≥ 5.8 GPa even exhibit an unusual upturn below around 2.5 K, as shown in Fig. 5d. The upward trend of Rc(T) below 2.5 K is more obvious when the pressure increases to 11.2 GPa. As shown below, these results can be well accounted for by turning on a Q1D SC in the resistor-matrix model. The above results thus unambiguously demonstrate the emergence of an extremely anisotropic SC in HfTe3 under pressures.

T-P phase diagram of HfTe3

The pressure dependencies of TCDW and Tc determined from Ra(T) are shown in Fig. 6, which depicts clearly the competition between CDW and the possible Q1D SC along the a-axis. With increasing pressure, TCDW initially increases, reaching a broad maximum at 1.5 GPa, then decreases quickly and vanishes completely at about 5 GPa, above which the Q1D SC emerges abruptly. The Tconset increases gradually and reaches a nearly constant value of ~5 K at least up to 12 GPa. The zero resistance is eventually realized at Tczero ≈2 K under P > 11 GPa, however, the diamagnetic response cannot be obtained due to the lack of phase coherence along the b-and c-axis.

Fig. 6: Temperature-pressure phase diagram of HfTe3 single crystal.
figure 6

Pressure dependences of the chage-density-wave transition temperature TCDW and the superconducting transition temperatures Tconset and Tczero determined from the resistance measurements along the a-axis.

Therefore, our present work not only demonstrates unambiguously a competitive nature between CDW and SC but also reveals an intriguing anisotropic SC in the pressurized HfTe3, i.e., the Q1D SC emerges along the a-axis in a relatively broad temperature/pressure range after the CDW is completely suppressed, whereas the phase coherence between the superconducting chains cannot be realized down to 1.4 K along the other directions.

One noteworthy feature of the phase diagram shown in Fig. 6 is that SC emerges along the a-axis only after the CDW is suppressed completely by pressure without showing any coexistence region. In contrast, the polycrystalline HfTe3 sample with at a lower TCDW ≈80 K shows the coexistence of CDW and filamentary SC at ambient pressure2. Although both the chemical disorders/defects and the physical pressure can suppress CDW and then induce SC, the above comparison actually highlights their distinct roles in tuning the CDW and SC states. In comparison with the “clean” tuning of physical pressure, the chemical disorders/defects may introduce extra charge carriers. According to the theoretical study by Imada et al.23, the relationship between CDW and SC in a Q1D system depends sensitively on the band filling, i.e., they are separated for a half-filled band but can coexist for quarter filling. They also pointed out that the coexistence of SC and CDW in a non-half-filled system is a general feature of anisotropic systems. It is plausible that the Q1D bands of HfTe3 along the a-axis is close to half-filling, leading to a complete competitive relationship between SC and CDW, while the deviation from the half-filling in the polycrystalline samples gives rise to the coexistence of CDW and SC. Our simple calculations on the density of states (DOS) distribution of the px/py/pz type Wannier orbits on the -Te2-Te3- atoms seem to support such a scenario, i.e., the filling numbers of the px-type Wannier orbits along the 1D atom chain direction are about 0.55, Supplementary Fig. 14. It should be noted that the valence state of Te ions in HfTe3 is complex, and more experimental and theoretical investigations are needed to understand the interplay between CDW and SC.

Crystal structure under high pressure

To study the structural stability of HfTe3 under compression, we also measured the high-pressure XRD on the pulverized HfTe3 crystals at room temperature and found no structure transition up to at least 14 GPa. As shown in Fig. 7, the lattice parameters a, b, c, and volume V contract smoothly with increasing pressure, implying that the crystal structure is stable in the studied pressure range. The pressure dependence of volume V(P) can be well fitted with the Birch–Murnaghan (B–M) equation of state as shown by the solid line in Fig. 7b. By fixing B1 = 4, we obtained a bulk modulus B0 = 45.3(7) GPa and V0 = 228.7(4) Å3. In addition, there is no obvious anomaly for the lattice parameter a at Pc ≈5 GPa, where the CDW disappears and the Q1D SC emerges at low temperatures. The axial ratios c/a and b/a decrease monotonically with increasing pressure, Supplementary Fig. 7, indicating that the b and c axes are more compressible than a-axis. The smooth evolutions of the axial ratios c/a and b/a up to ~10 GPa also indicate that the crystal structure is stable within the studied pressure range. Therefore, we can safely conclude that pressure-induced SC around 5 GPa should be correlated with an electronic transition rather than a structural transition, i.e., it is linked directly with the elimination of the CDW order. In this sense, the pressure-induced SC along the a-axis can be regarded as a new electronic phase, which shares a similar crystal structure as the ambient phase but has no CDW instability.

Fig. 7: Crystal-structural under high pressure.
figure 7

Pressure dependences of a the lattice parameters and b volume of HfTe3 at room temperature. The solid line in (b) represents the fit with the Birch–Murnaghan (B–M) equation of state to extract the bulk modulus B0 as given in the figure.

Discussions

In order to simulate the above anisotropic transport behaviors at low temperatures under HP, we designed a simplified 2D resistor matrix consisting of 45 resistors as shown in Supplementary Figs. 811 We first simulate the change of b-axis resistance Rb with a perfect alignment upon entering the superconducting state along the a-axis, manifested by a reduction of the resistors’ resistance from 5 to 0.1 Ω along this direction. As shown in Supplementary Fig. 8, the voltage drop across V+/V− along the b-axis (or equivalently Rb) remains unchanged when the resistance of resistors along the a-axis are reduced from 5 to 0.1 Ω, consistent with the experimental results for sample #4 in Fig. 5b. However, when the alignment of voltage leads is not perfect, some a-axis component will be picked up as simulated in Supplementary Fig. 9. In this case, the voltage drop across V+/V− is found to decrease slightly when the resistors along the a-axis are reduced from 5 to 0.1 Ω. This reproduces well the observed resistance drop of sample #5 shown in Fig. 5b, confirming that the change of Ra will affect the Rb sensitively if the alignment of voltage leads is imperfect.

We further simulated the influence of resistance drop of Ra and the misalignment of V+/V− on the c-axis resistance Rc. As illustrated in Supplementary Figs. 10 and 11, interestingly, the voltage drop across V+/V− along the c-axis (or equivalently Rc), no matter aligned perfectly or not, was both found to rise progressively when the resistance of the resistors along the a-axis is reduced gradually, in excellent agreement with the experimental results shown in Fig. 5d. In contrast, the voltage drop will decrease if the superconducting transition takes place in the c-axis direction. The above simulations not only reproduce the experimental observations but also support strongly the occurrence of a Q1D SC in the pressurized HfTe3 single crystal.

To further substantiate the occurence of Q1D SC along the a-axis, we then analyzed the excess conductivity \(\left( {{\Delta}\sigma = \sigma - \sigma _0} \right)\) around Tc according to A-L model21, viz.

$${\Delta}\sigma \propto \left[ {\left( {T - T_c} \right)/T_c} \right]^{ - (4 - D)/2}$$
(1)

where D is the dimensionality and Tc is the mean-field transition temperature, which equals to Tczero. For D = 1, the excess conductivity due to 1D superconducting fluctuations is given by:

$${\Delta}\sigma = \left( {\pi e^2/16\hbar d^2} \right)\xi (0)\left[ {\left( {T - T_c} \right)/T_c} \right]^{ - 3/2}$$
(2)

Here, e is the electron charge, d is the diameter of a superconducting filament, and \(\xi \left( 0 \right)\) is the coherence length at T = 0 K. As illustrated in Fig. 8, the excess conductivity\({\Delta}\sigma\) from 9 to 12.5 GPa for sample #3 measured at a smaller current 0.1 mA indeed follows a nice linear relationship as a function of \(\left[ {\left( {T - T_c} \right)/T_c} \right]^{ - 3/2}\); here the \({\Delta}\sigma\) is calculated from data shown in Fig. 3e and \(\sigma _0\) is the conductivity at 6 K for I//a with 0.1 mA current. These observations indicate that strong 1D superconducting fluctuations persist even up to 12.5 GPa. It is also noteworthy that the slope of the fitting curve decreases gradually with increasing pressure with a ratio of \(k_{9\;{{{\mathrm{GPa}}}}}:k_{11\;{{{\mathrm{GPa}}}}}:k_{12.5\;{{{\mathrm{GPa}}}}} = 2.4:1.4:1\). As shown in Fig. 3e, the Tc in the investigated pressure range only changes slightly, and thus the change of \(\xi \left( 0 \right)\) should not dominate the significant decrease of the slope or the pre-factor of the above Eq. (2). Instead, the reduction of the slope should be mainly attributed to the increase of the diameter d of the superconducting filaments under pressure. This is conceivable considering the fact that the superconducting filaments are getting closer under HP. It can be expected that the superconducting state will gradually evolve into three dimensions under higher pressure or lower temperatures where the Josephson coupling between filaments is strengthened.

Fig. 8: Analysis with the Aslamazov and Larkin (A-L) model.
figure 8

The reduced temperature dependence of excess conductivity around Tc along the a-axis.

Before discussing the possible origins for the observed Q1D SC in the pressurized HfTe3, it is imperative to understand the nature of the CDW transition and its interplay with SC. A side-by-side comparison between HfTe3 and ZrTe3 is instructive. At ambient pressure, ZrTe3 and HfTe3 exhibit a CDW transition at TCDW = 63 and 93 K, respectively. The determined CDW vector of HfTe3, i.e., q = 0.91(1) a* + 0.27(1) c* is very similar to that of ZrTe3, i.e., q = 0.93 a* + 0.33 c*,6. In general opinion, the CDW vector is related to both the FS nesting vector and the electron-phonon coupling distribution. As shown in Fig. 9, the calculated electronic susceptibility χq′ of HfTe3 demonstrates an obvious peak at 0.90 a*, which is in excellent agreement with the experimental result determined from the TEM measurements. Nevertheless, there is no obvious peak along the c* direction, so it is difficult to determine the c* component of q merely based on the information of FS nesting. We suspect that the electron-phonon coupling in HfTe3 may play a dominant role in the c* component of its CDW vector. The calculated electronic susceptibilities of ZrTe3 under 0 and 5 GPa are shown in Supplementary Fig. 13 for comparison. The red regions were drawn by the contour line of calculated electronic susceptibility χq′, while the experimentally determined CDW vectors q were also labeled. Obviously, the c* components of experimental q do not correspond to the peaks of calculated χq′, while their a* components seem to be well matched. This is similar to the above case of HfTe3.

Fig. 9: The calculated electronic susceptibility and localization function of HfTe3.
figure 9

The calculated real part of the electronic susceptibility χq′ for HfTe3 in the a*-c* plane at a 0 GPa and b 5 GPa. The components along the b* direction have been integrated. The calculated electronic localization function of HfTe3 at c 0 GPa and d 5 GPa. The isosurface was set as 0.5.

Below TCDW, an anisotropic superconducting transition was discovered in ZrTe35,13. In specific, ρa starts to decrease at Tconset ≈4 K and reaches zero at Tczero ≈2 K with a relatively broad superconducting transition, while ρb exhibits a much narrower transition from Tconset ≈2 K to Tczero ≈1.5 K. Based on the specific-heat measurements, the observed anisotropic superconducting transition has been ascribed to a gradual crossover from a filamentary SC (between 4 and 2 K) induced by local pairs along the a-axis to a bulk SC (below 1.5 K) when the superconducting filaments form phase coherence13. Such a mixed bulk-filament SC observed in ZrTe3 has been attributed to the peculiar electronic structures characterized by the presence of Q1D and 3D FSs10,13,19. In striking contrast to ZrTe3, no SC was detected down to 50 mK for the high-quality HfTe3 single crystals2. Their distinct low-temperature behaviors can be rationalized by comparing to the effects of physical pressure on the CDW and SC of ZrTe3. At ambient conditions, the cell volume of HfTe3 (228.70 Å3) is about 1.35% smaller than that of ZrTe3 (231.83 Å3)8,20, which corresponds to the application of 0.84 GPa hydrostatic pressure on ZrTe3 given compressibility of 0.016 GPa−1 14 At this pressure, it was found that the CDW transition of ZrTe3 is enhanced to ca. 100 K while the SC is suppressed completely19,22, which are similar to the situations seen in HfTe3 at ambient pressure.

The above simple arguments indicate that the coexistent CDW and SC in ZrTe3 at ambient pressure are actually competing with each other, which becomes more evident under pressure. For ZrTe3, a reentrant SC emerges above 5 GPa when the CDW disappears abruptly; the observed SC was believed to be bulk in nature17, but the diamagnetic evidence is still lacking to date as mentioned above12. In the present case, we observed similar pressure-induced SC in Ra of HfTe3, but the absences of SC along the other directions and the absence of diamagnetic signals point to an intrinsic filamentary character for the observed SC. Our work thus calls for a reexamination of the anisotropic properties of ZrTe3 under high pressures. Our DFT calculations also provide some clues for the suppression of CDW under pressure. As shown in Fig. 9c, d, the distances between the Te2 and Te3 atoms in the -Te2-Te3- chain along the a-axis becomes more uniform at high pressures. In addition, there are more localized electrons between the Te2 and Te3 atoms from different prismatic (HfTe3) chains.

Finally, we discuss the possible origins for the observed extremely anisotropic or Q1D SC in the pressurized HfTe3 according to the proposal by Yamaya et al.13 Similar to ZrTe38,9, our electronic structure calculations show that HfTe3 owns both 3D FSs around the BZ center and Q1D FSs around the BZ boundary, as shown in Fig. 10a. The former ones are dominated by Hf 5d orbitals in the (HfTe3) prism, while the latter ones are mainly contributed by unidirectional Te-5px orbitals within the -Te2-Te3-Te2-Te3- chain along the a-axis, Fig. 1a. At the intersection area between these two types of FSs, there are also Q1D + 3D FSs as indicated by the black arrows in Fig. 10a. On these FSs, the Fermi velocities are displayed with the color scales, where the red and blue colors represent the highest and zero Fermi velocities, respectively. In particular, the low Fermi velocities near the Q1D + 3D FSs with the narrow-band feature may result in van Hove singularity (vHs) in the density of states. It is known that the presence of a vHs at EF can induce FS instabilities and lead to SC or CDW states. In the narrow bands that construct the Q1D + 3D FSs, electrons can form local pairs that interact attractively with each other over short distances13. With increasing pressure, as shown in Fig. 10b, the 3D FSs of HfTe3 around the Brillouin zone center expand, which enhance their overlapping with the Q1D FSs and reduce the composition of the Q1D bands at the Fermi level. As a result, the FS nesting of Q1D FSs along the a* direction is weakened under pressure, which may suppress the CDW order and induce the SC. Since the narrow-band electron of HfTe3 also has a dominant Te-5px characteristic originating in the -Te2-Te3- chains as in ZrTe3, the local pairs must be dominantly formed along the a-axis, leading to the Q1D SC transition in the a-axis direction19. In comparison, the 3D FSs of ZrTe3 are larger than those of HfTe3 at 0 GPa and even become the open-orbital pockets along the a* direction at 5 GPa, as shown in Fig. 10c, d. This indicates that the CDW order due to the Q1D FSs nesting in ZrTe3 is not so strong as in HfTe3, which is consistent with the tendency of the experimental TCDW of ZrTe3 (63 K)6 versus HfTe3 (93 K). Therefore, the anisotropic SC property in HfTe3 can maintain in a wider temperature/pressure range, which makes HfTe3 an ideal platform to study the local pairs in the Q1D superconducting system.

Fig. 10: Pressure effect on the Fermi surfaces of HfTe3 and ZrTe3.
figure 10

The calculated Fermi surface sheets of a, b HfTe3 and c, d ZrTe3 at 0 GPa and 5 GPa, respectively. The color scales on the Fermi surface represent the Fermi velocities, where the pure red and pure blue correspond to the highest and zero Fermi velocities, respectively. The bottom figures in each panel are the sections of Fermi surface sheets at the c* = 0 plane.

In summary, we have measured the anisotropic resistances of the CDW conductor HfTe3 under high pressures. We found that a Q1D SC emerges in Ra(T) at T < 4–5 K when the CDW is suppressed completely by P > 5 GPa, whereas no clear sign of SC is observed down to 1.4 K in Rb(T) and Rc(T). Our results demonstrate an intimate interplay between CDW and SC in HfTe3, which are particularly interesting given their Q1D characters. We have compared our results with those of ZrTe3 and discussed the pressure-induced Q1D SC in pressurized HfTe3 in terms of the peculiar electronic structures and the local pairs formed along the -Te2-Te3- chains parallel to the a-axis.

Methods

Single-crystal growth and characterization

High-quality HfTe3 single crystals used in the present study were grown with the iodine chemical-vapor-transport method as described elsewhere2. Before HP measurements, the crystals have been carefully characterized at ambient pressure by single-crystal X-ray diffraction (XRD), energy dispersive spectroscopy (EDS), TEM, and resistivity measurements. The CDW vector q at ambient pressure was determined by TEM measurements at 26 K. The specimens for TEM observation were prepared by mechanical exfoliation at room temperature and then transfer onto the Cu grid with the aid of crystal bond and acetone. In situ, low-temperature TEM experiments were performed using a liquid-helium-cooled specimen holder (GATAN, HCHDT3010) in a JEM-2100F electron microscope equipped with a charge-coupled device (CCD) camera system and operated at 200 keV.

Details of high-pressure measurements

The temperature dependences of resistance R(T) under various pressures were measured with the standard four-probe method by using either a self-clamped piston-cylinder cell (PCC)24 in the pressure range 0 < P ≤ 2 GPa or a palm-type cubic anvil cell (CAC)25 in the pressure range 2 < P ≤ 12.5 GPa. In order to improve the electrical contact, we first evaporated four gold pads on the sample surface in a vacuum and then attached four gold wires of 20 μm in diameter on top of the gold pads with the silver paste. AC magnetic susceptibility under pressure has been measured with the mutual induction method. The excitation and pick-up coils are made of enameled copper wires of 25 μm in diameter, and they were manually wound around a Teflon sheet with a rectangular cross-section of 0.3 mm × 0.5 mm. The Teflon sheet can be removed easily and the rectangular space ensures a large filling factor for the susceptibility measurements. Each coil has two layers and about 40 turns in total. The HfTe3 sample and a piece of lead (Pb) with a similar size of about 0.3 mm × 0.3 mm × 0.25 mm were put inside the pick-up coil. An excitation current of 1 mA with a frequency of 1117 Hz was applied to the excitation coil and the output signal was picked up with a Stanford Research SR830 lock-in amplifier. For these HP measurements, the Daphne 7373 or glycerol was employed as the pressure transmitting medium. The pressure in PCC was determined from the superconducting transition temperature of Pb, while the pressure in CAC was estimated from the calibration curve at low temperatures26. High-pressure synchrotron powder XRD (λ = 0.6199 Å) was performed at room temperature at 4W2 beamline, Beijing Synchrotron Radiation Facility (BSRF). Silicone oil was used as the transmitting medium. The Le Bail method was employed to fit the XRD data with the General Structure Analysis System (GSAS) package program27. The ruby fluorescence method was used to determine the pressure.

Theoretical calculations

To investigate the electronic structures of HfTe3 under pressure, we carried out density functional theory (DFT) calculations with the projector augmented wave method28,29 as implemented in the VASP package30,31,32. The generalized gradient approximation of Perdew–Burke–Ernzerhof (PBE) type33 was adopted for the exchange-correlation functional. The kinetic energy cutoff of the plane wave basis was set to 300 eV. A 16 × 16 × 8 k-point mesh was used for the Brillouin zone (BZ) sampling. The Fermi surface was broadened by the Gaussian smearing method with a width of 0.05 eV. The spin–orbit coupling effect was included as both Hf and Te are heavy elements. The vdW interaction was described by the optB86b-vdW functional34. In the structural optimization, both cell parameters and internal atomic positions were allowed to relax until the forces on all atoms were smaller than 0.01 eV/Å. The Fermi surface (FS) and Fermi velocity of HfTe3 were studied with the Wannier9035 and FermiSurfer36 packages, respectively. To describe the Fermi surface nesting, the real part of electronic susceptibility \(\chi _{{{\boldsymbol{q}}}}\prime\) (\(\chi _{{{\boldsymbol{q}}}}\prime = \mathop {\sum}\nolimits_{{{{\boldsymbol{k}}}},ij} {\frac{{f\left( {\varepsilon _{{{{\boldsymbol{k}}}} + {{{\boldsymbol{q}}}},j}} \right) - f\left( {\varepsilon _{{{{\boldsymbol{k}}}},i}} \right)}}{{\varepsilon _{{{{\boldsymbol{k}}}},i} - \varepsilon _{{{{\boldsymbol{k}}}} + {{{\boldsymbol{q}}}},j}}}}\)) was computed, where \(f\left( {\varepsilon _{{{{\boldsymbol{k}}}},i}} \right)\) is the Fermi-Dirac distribution function and \(\varepsilon _{{{{\boldsymbol{k}}}},i}\) is the energy of band i at vector \({{{\boldsymbol{k}}}}\).