This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

thanks: These authors contribute equally to this work.thanks: These authors contribute equally to this work.

Bulk superconductivity and Pauli paramagnetism in nearly stoichiometric CuCo2S4

Yu-Ying Jin Department of Physics, Zhejiang Province Key Laboratory of Quantum Technology and Devices, Interdisciplinary Center for Quantum Information, and State Key Lab of Silicon Materials, Zhejiang University, Hangzhou 310027, China    Shi-Huai Sun Department of Physics, Zhejiang Province Key Laboratory of Quantum Technology and Devices, Interdisciplinary Center for Quantum Information, and State Key Lab of Silicon Materials, Zhejiang University, Hangzhou 310027, China    Yan-Wei Cui Department of Physics, Zhejiang Province Key Laboratory of Quantum Technology and Devices, Interdisciplinary Center for Quantum Information, and State Key Lab of Silicon Materials, Zhejiang University, Hangzhou 310027, China School of Science, Westlake Institute for Advanced Study, Westlake University, Hangzhou 310064, China    Qin-Qing Zhu Department of Physics, Zhejiang Province Key Laboratory of Quantum Technology and Devices, Interdisciplinary Center for Quantum Information, and State Key Lab of Silicon Materials, Zhejiang University, Hangzhou 310027, China School of Science, Westlake Institute for Advanced Study, Westlake University, Hangzhou 310064, China    Liang-Wen Ji Department of Physics, Zhejiang Province Key Laboratory of Quantum Technology and Devices, Interdisciplinary Center for Quantum Information, and State Key Lab of Silicon Materials, Zhejiang University, Hangzhou 310027, China    Zhi Ren School of Science, Westlake Institute for Advanced Study, Westlake University, Hangzhou 310064, China    Guang-Han Cao ghcao@zju.edu.cn Department of Physics, Zhejiang Province Key Laboratory of Quantum Technology and Devices, Interdisciplinary Center for Quantum Information, and State Key Lab of Silicon Materials, Zhejiang University, Hangzhou 310027, China Collaborative Innovation Centre of Advanced Microstructures, Nanjing University, Nanjing 210093, China
Abstract

It has long remained elusive whether CuCo2S4 thiospinel shows bulk superconductivity. Here we clarify the issue by studying on the samples of sulfur-deficient CuCo2S3.7 and sulfurized CuCo2S4. The sample CuCo2S3.7 has a smaller lattice constant of a=9.454a=9.454 Å, and it is not superconducting down to 1.8 K. After a full sulfurization, the aa axis of the thiospinel phase increases to 9.475 Å, and the thiospinel becomes nearly stoichiometric CuCo2S4, although a secondary phase of slightly Cu-doped CoS2 forms. Bulk superconductivity at 4.2 K and Pauli paramagnetism have been demonstrated for the sulfurized CuCo2S4 by the measurements of electrical resistivity, magnetic susceptibility, and specific heat.

pacs:
74.70.Xa; 72.80.Ga; 61.66.Fn

I Introduction

The discoveries of superconductivity (SC) in the complex copper oxide [1] and the iron-based pnictide [2] stimulate enthusiasm to search for SC especially in late 3d-transition-metal (Fe, Co, Ni, and Cu) compounds [3, 4, 5, 6, 7]. Among them, the Co-based superconductors are very limited so far. One example is the cobalt oxyhydrate NaxCoO2y{}_{2}\cdot yH2O (x0.35,y1.3x\approx 0.35,y\approx 1.3), which shows SC at TcT_{\mathrm{c}}\approx 4.5 K [8]. The Co-based thiospinel CuCo2S4 shows similarities with NaxCoO2y{}_{2}\cdot yH2O in the Co coordination, geometrical frustration, and formal Co valence. However, it is not clear whether CuCo2S4 superconducts or not. Besides, the magnetism of CuCo2S4 also remains elusive up to present.

Earlier studies in 1960s suggested Pauli paramagnetism in CuCo2S4 [9, 10], and no superconducting transition was observed down to 0.05 K [11]. In 1990s, however, it was reported that CuCo2S4 shows a Curie-Weiss (CW) paramagnetism with an effective magnetic moment of 0.89 μB\mu_{\mathrm{B}} per formula unit (f.u.) [12]. A cusp in the magnetic susceptibility appears at TN=T_{\mathrm{N}}= 18 K, which was attributed to an antiferromagnetic spin ordering. In a multiphasic sample with the nominal composition of Cu1.5Co1.5S4, SC or superconductor-like behavior was observed with an onset transition temperature of Tconset=T_{\mathrm{c}}^{\mathrm{onset}}= 4.0 K and a zero-resistance temperature of Tczero=T_{\mathrm{c}}^{\mathrm{zero}}= 2.3 K. Investigations on the 63Cu and 59Co NMR suggested a gapless superconducting state as well as antiferromagnetic spin correlations, and SC was considered to be in line with the growth of antiferromagnetic spin correlation [13]. Contrastingly, later NMR study on the Co-rich series samples of (CuxCo1-x)Co2S4 indicated a full superconducting gap without long-range magnetic ordering for CuCo2S4 [14]. It was concluded that SC and the antiferromagnetic spin correlation are associated with the Co-3d and Cu-3d holes, respectively.

One of the present authors (G.-H.C.) and coworkers [15] attempted to reproduce SC in CuCo2S4 in 2003. However, no SC was observed above 1.8 K in the single-phase sample of CuCo2S4, although signature of SC at Tc=T_{\mathrm{c}}= 3.5 K was detected in a multiphasic sample. Aito and Sato [16] reported the resistivity data of five CuCo2S4 samples from different batches. Two of them showed a superconducting transition, and the higher TcT_{\mathrm{c}} value is 3.8 K. Fang et al. [17] synthesized an unusual K-doped sample Cu1.3K0.2Co1.5S4 which showed SC at 4.4 K together with a CW-like susceptibility but without any antiferromagnetic transition down to 9 K. A very recent work [18] showed absence of SC with a weak antiferromagnetic transition at about 4 K in CuCo2S4. In a word, the previous reports on CuCo2S4 appear to be highly dispersive and even contradictive. To our knowledge, evidence of bulk SC with specific-heat measurements has not been reported so far in the Cu-Co-S system.

The contradictive results above strongly suggest that the physical properties are sensitive to the synthesis of samples, and a controlled preparation of nearly stoichiometric samples of CuCo2S4 is crucial to clarify the intrinsic properties. Previous studies showed difficulties in obtaining desired samples of CuCo2S4 [12, 19, 20, 21, 15, 16]. They were commonly synthesized by direct reacting copper and cobalt powders with sulfur in a sealed evacuated silica tube at an elevated temperatures. While relatively low reaction temperatures (500 C - 600 C) were suggested for the preparation of monophasic CuCo2S4 [22], however, a following-up work [19] failed to obtain the single-phase sample. The synthesized sample tends to form CoS2 impurity, as revealed by the phase-relation study in the Cu-Co-S system [20, 21]. Indeed, later studies [12, 14, 15, 16] also showed presence of the CoS2 impurity for the reaction temperatures from 500 C to 800 C. Note that CoS2 is a ferromagnet with a Curie temperature of \sim120 K [23, 24, 25], which makes it more easily to be detected by the magnetic measurement [15].

Here we report a novel two-step strategy for the controlled synthesis of stoichiometric CuCo2S4. First, to minimize the formation of CoS2 impurity, we prepared sulfur-deficient CuCo2S4-δ with δ=\delta= 0.3. Then, the S-deficient sample was sulfurized by annealing in the presence of appropriate amount of sulfur. As a result, the main phase of the annealed sample was found to be nearly stoichiometric CuCo2S4. Bulk SC at 4.2 K and Pauli paramagnetism in the normal state were demonstrated in the S-compensated CuCo2S4.

II Experimental methods

Polycrystalline sample of S-deficient CuCo2S3.7 was first prepared by high-temperature reactions of the constituent elements in a sealed evacuated silica tube. The source materials were powders of copper (99.997%), cobalt (99.998%), and sulfur (99.999%). The homogenized mixture with the composition of CuCo2S3.7 was allowed to fire at 750 C for 72 hours. This procedure was repeated to improve the quality of the sample. In the second step, the synthesized CuCo2S3.7 was sulfurized in the presence of compensatory sulfur (0.35 S/f.u.) by annealing at 450 C for 144 hours in a sealed evacuated silica ampoule. Note that excess of sulfur was necessary to ensure a full sulfurization. This is because, during the sulfurization, a side reaction that forms CoS2 always takes place, which additionally consumes sulfur. Besides, in order to quantize the amount of the CoS2 impurity in the sulfurized sample, we additionally prepared CoS2 by reacting Co with S in an evacuated silica tube. The sample is of single phase with the lattice constant of a=5.535a=5.535 Å (consistent with the previous report [23]), as determined by the powder x-ray diffractions (XRD).

Powder XRD were carried out using a PANalytical diffractometer (Empyrean Series 2) with a monochromatic Cu-Kα1K_{\alpha 1} radiation. The crystal structure were refined by a Rietveld analysis using the GSAS+EXPGUI package [26]. The sulfur content in the crystallites was examined by energy-dispersive x-ray spectroscopy (EDS, Oxford Instruments X-Max) equipped in a scanning electron microscope (SEM, Hitachi S-3700N).

The electrical resistivity and specific heat were measured on a Quantum Design Physical Properties Measurement System, and the magnetic properties were measured on a Quantum Design Magnetic Property Measurement System. The resistivity measurement employed a standard four-terminal method. The heat-capacity measurement utilized a thermal relaxation technique. In the magnetic measurements, the applied magnetic fields were set to be 20 Oe and 10,000 Oe, respectively, to detect SC and to study the normal-state magnetism. In the case of the low-field measurements, both zero-field-cooling and field-cooling protocols were employed.

III Results and discussion

Refer to caption
Figure 1: Powder X-ray diffractions with the Rietveld refinement profiles for samples of sulfur-deficient CuCo2S3.7 (a) and sulfurized CuCo2S4 (b). The insets (with a logarithmic scale for the intensity) are a close-up of the marked area, which shows presence of the secondary phases of Cu2S and CoS2, respectively, in CuCo2S3.7 and sulfurized CuCo2S4.

Figure 1(a) shows the XRD profile for the sulfur-deficient sample of CuCo2S3.7. Most of the reflections can be well indexed with a face-centered cubic unit cell of the thiospinel. As is seen in the inset, no reflections associated with CoS2 are detectable, while tiny amount of Cu2S is possibly presented. Therefore, with a lack of sulfur we succeeded in avoiding the appearance of the CoS2 secondary phase. The Rietveld refinement (RwpR_{\mathrm{wp}} = 2.3% and χ2\chi^{2} = 1.31) confirms the normal spinel structure with a=9.4544(1)a=9.4544(1) Å and u=0.3865(1)u=0.3865(1) for the main phase. Note that the lattice constant is the smallest among those reported previously for CuCo2S4 [9.461(2)9.461(2) Å [22], 9.478(5)9.478(5) Å [19], and 9.472(1)9.472(1) Å [20]]. This may be attributed the apparent sulfur deficiency and/or the partial substitution of Cu by Co (hereafter denoted as Co/Cu substitution) [14]. The latter is implied by the presence of small amount of Cu2S. As a matter of fact, the Rietveld refinement does not support a significant sulfur vacancy.

Refer to caption
Figure 2: Typical scanning electron microscope images of sulfur-deficient CuCo2S3.7 (a) and sulfurized CuCo2S4 (b). The lower-right insets are the energy-dispersive x-ray spectrum (EDS) collected with the electron beam focused on the spots marked. Round-shape grains (indicated by arrows) can be seen in panel (b), which are identified to be lightly Cu-doped CoS2. The atomic ratios are given by the EDS analysis.

The XRD pattern of the sulfurized CuCo2S4 is displayed in Fig. 1(b). The main phase remains to be the cubic thiospinel, although small amount of CoS2-like phase emerges. With the two-phase Rietveld refinement (RwpR_{\mathrm{wp}} = 2.2% and χ2\chi^{2} = 1.13), the weight percentage of the CoS2-like impurity was determined to be 14.8(6)%. The lattice constant of the pyrite phase is refined to be 5.538(1) Å, which is slightly larger than that of CoS2 (5.534 Å [23]), suggesting that Cu is slightly incorporated. The structural parameters of the main phase were fitted to be a=9.4750(2)a=9.4750(2) Å and u=0.3851(1)u=0.3851(1). The aa axis is remarkably larger than that of the sulfur-deficient CuCo2S3.7, suggesting a successful sulfurization.

The two samples above were examined by SEM observations in combination with the EDS measurements. As shown in Fig. 2(a), the crystallites of S-deficient CuCo2S3.7 are similar in shape. The sulfur content, measured on the basis of the Co content, is consistent with the nominal composition. However, the Cu content is substantially lower than the nominal one. The result suggests that the real composition of the thiospinel phase is something like (Cu1-x-yCoyx{}_{x}\Box_{y})Co2S4-δ. The SEM image of the sulfurized sample [Fig. 2(b)] shows additional round-shape crystallites which were identified to be slightly Cu-doped CoS2 (1-2% Cu) by the EDS analysis. Furthermore, the sulfur deficiency is fully compensated, and the Cu content is also increased, as is indicated by the atomic ratio measured. Therefore, we conclude that the sulfurized sample mainly (\sim85% by weight) contains nearly stoichiometric CuCo2S4.

Refer to caption
Figure 3: Temperature dependence of magnetic susceptibility or magnetization for sulfur-deficient CuCo2S3.7 (a) and sulfurized CuCo2S4 (b). The inset of (a) is a close-up of the high-temperature data, indicating a positive-temperature-coefficient behavior (dashed line). In panel (b), the magnetization of CoS2 (multiplied by a factor of 18.8%) is plotted for comparison. The inset of (b) compares the magnetic susceptibilities at high temperatures.

Figure 3(a) shows the temperature dependence of magnetic susceptibility under a magnetic field of H=H= 10 kOe for the sulfur-deficient CuCo2S3.7. The magnetic susceptibility is nearly temperature independent at high temperatures. No anomaly at \sim120 K can be seen, indicating free of the ferromagnetic impurity of CoS2. There is an upturn tail at low temperatures. Fitting of the data with the CW formula, χ=χ0+C/(TθCW)\chi=\chi_{0}+C/(T-\theta_{\mathrm{CW}}), yields a temperature-independent term of χ0\chi_{0} = 0.00047 emu mol-f.u.-1, a Curie constant of CC = 0.0043 emu K mol-f.u.-1, and a paramagnetic CW temperature of θCW=1.9\theta_{\mathrm{CW}}=-1.9 K. Such a small value of the Curie constant (corresponding to 0.13 μB\mu_{\mathrm{B}}/Co) is commonly originated from tiny paramagnetic impurities. Additionally, the positive temperature coefficient of susceptibility at high temperatures, shown in the inset of Fig. 3(a), also rules out the possible CW-type paramagnetism in CuCo2S3.7.

Figure 3(b) shows the temperature dependence of magnetization (in the unit of μB\mu_{\mathrm{B}}/f.u.) of the sulfurized CuCo2S4 under the same magnetic field of H=H= 10 kOe. A ferromagnetic transition is seen at about 120 K, which is attributed to the ferromagnetic impurity of slightly Cu-doped CoS2 that was identified by the XRD experiment above. To quantify the amount of (Co,Cu)S2 independently, the magnetization data of pure CoS2 are shown for comparison. One sees that the Curie temperature of (Co,Cu)S2 is slightly lower than that of CoS2 due to the Cu incorporation. The low-temperature saturation magnetization is about 19% of that CoS2. At the same time, the high-temperature magnetic susceptibility basically coincides. Since the Cu content in (Co,Cu)S2 is only 1-2% according to the EDS measurement, the amount of the (Co,Cu)S2 impurity should be also around 19%, basically consistent with the XRD result above.

The high-temperature magnetic susceptibility data are highlighted in the inset of Fig. 3(b), which shows a CW-type paramagnetism. The CW paramagnetism is attributed to the (Co,Cu)S2 impurity, because the magnetic susceptibility of the sulfurized CuCo2S4 shows a similar temperature dependence with that of 18.8% CoS2. The magnetic susceptibility of S-compensated CuCo2S4 phase can be roughly obtained by a simple substraction. The result indicates a small value of magnetic susceptibility that is almost temperature independent. Therefore, CuCo2S4 should be intrinsically Pauli paramagnetic. Nevertheless, the accurate value of the Pauli-paramagnetic susceptibility cannot be reliably extracted not only because of the influence of the magnetic impurity, but also due to the possible Van Vleck paramagnetism involved [10]. According to the bandstructure calculation of CuCo2S4 which gives the density of states at Fermi level of 31.88 states/eV/f.u. [27], the calculated Pauli-paramagnetic susceptibility is derived to be χP=μB2N(EF)=1.03×103\chi_{\mathrm{P}}=\mu_{\mathrm{B}}^{2}N(E_{\mathrm{F}})=1.03\times 10^{-3} cm3/mol.

Refer to caption
Figure 4: Temperature dependence of magnetic susceptibility for sulfur-deficient CuCo2S3.7 as well as sulfurized CuCo2S4, measured under a magnetic field of 20 Oe in both field-cooling (FC) and zero-field-cooling (ZFC) modes. The inset shows field dependence of magnetization at 2 K for sulfurized CuCo2S4.

Figure 4 shows the low-temperature susceptibility data for the samples of CuCo2S3.7 as well as sulfurized CuCo2S4. The S-deficient CuCo2S3.7 exhibits low values of magnetic susceptibility, and no signal of SC can be detected down to 1.8 K. By contrast, the sulfurized sample shows a steep decrease in the magnetic susceptibility at 4.2 K, suggesting a superconducting transition. Note that the high value of the susceptibility above TcT_{\mathrm{c}} is due to the ferromagnetic impurity (Co,Cu)S2. The large magnitude of the ZFC diamagnetism (exceeding 100-100%) below TcT_{\mathrm{c}} could also be due to the extra magnetic field generated by the ferromagnetic (Co,Cu)S2. The inset shows the field dependence of magnetization at 2 K for the superconducting sample. An extremely type-II superconductivity with Hc2Hc1H_{\mathrm{c2}}\gg H_{\mathrm{c1}} can be concluded. As expected also, the ferromagnetic signal from (Co,Cu)S2 is superposed on the superconducting loop.

Figure 5(a) shows the temperature dependence of resistivity for the sulfur-deficient CuCo2S3.7 and the sulfurized CuCo2S4. Both samples show a metallic behavior, yet the sulfurized CuCo2S4 sample exhibits a lower room-temperature resistivity with a higher residual resistivity ratio (RRR). The RRR values are 1.4 and 6.4 for CuCo2S3.7 and CuCo2S4, respectively. Although there are about 19% (Co,Cu)S2 impurity in the sulfurized CuCo2S4 sample, no anomaly at \sim120 K associated with the ferromagnetic transition can be detected. At lower temperatures, while no superconducting transition appears down to 1.8 K for CuCo2S3.7, a sharp superconducting transition is seen at Tconset=T_{\mathrm{c}}^{\mathrm{onset}}= 4.3 K for the S-compensated CuCo2S4. The observation of SC in relation with a high RRR value was also reported previously [16]. This could suggest that the nonmagnetic scattering, measured by the residual resistivity, may destroy SC in the system, resembling the scenario in Sr2RuO4 [28] and K2Cr3As3 [29]. Besides, the low-temperature resistivity of CuCo2S4 essentially shows a T2T^{2} temperature dependence (see the inset), suggesting dominant electron-electron scattering in the system.

Refer to caption
Figure 5: (a) Temperature dependence of electrical resistivity (ρ\rho) of the polycrystalline samples of sulfur-deficient CuCo2S3.7 and sulfurized CuCo2S4. The inset plots ρ\rho versus T2T^{2} in the temperature range from 4.5 to 50 K. (b) Resistive superconducting transitions under increased magnetic fields from which the upper critical fields Hc2H_{\mathrm{c2}} were obtained. The inset plots the resultant Hc2H_{\mathrm{c2}} as a function of temperature.

The resistive superconducting transitions are more clearly shown in Fig. 5(b). One sees that the superconducting transition shifts to lower temperatures with increasing magnetic fields. Using the criterion of 50% normal-state resistivity just above TcT_{\mathrm{c}} for determining Tc(H)T_{\mathrm{c}}(H), the upper critical magnetic fields Hc2H_{\mathrm{c2}} can be extracted. The resultant Hc2(T)H_{\mathrm{c2}}(T) data are shown in the inset of Fig. 5(b), which shows an essentially linear temperature dependence down to 0.42TcT_{\mathrm{c}}. This result suggests dominant orbital pair-breaking mechanism over a paramagnetic pair-breaking mechanism. The zero-temperature upper critical field is estimated to be Hc2(0)=H_{\mathrm{c2}}(0)= 24.6 kOe from the linear extrapolation, far below the Pauli-paramagnetic limit HPH_{\mathrm{P}}\approx 77 kOe. The coherence length can thus be derived as ξ0=\xi_{0}= 11.6 nm using the relation Hc2(0)=Φ0/[2πξ(0)2]H_{\mathrm{c2}}(0)=\Phi_{0}/[2\pi\xi(0)^{2}], where Φ0(=2.07×1015\Phi_{0}(=2.07\times 10^{-15} Wb) denotes a magnetic flux quantum.

Figure 6(a) shows the temperature dependence of specific heat for the sulfurized CuCo2S4 sample. The specific heat tends to approach the value of 3NR=3NR= 174.6 J K-1 mol-1 at high temperatures, in accordance with the Dulong-Petit law. No obvious anomaly is seen at around 120 K where the CoS2 impurity undergoes a ferromagnetic transition. This observation verifies that the CoS2 impurity is the minor phase. As is seen in the inset of Fig. 6(a), at low temperatures, a remarkable specific-heat jump is observable at around 4 K, confirming bulk SC in the sulfurized sample which dominantly contains nearly stoichiometric CuCo2S4.

Refer to caption
Figure 6: Temperature dependence of specific heat for sulfurized CuCo2S4. The inset of (a) shows a close-up in the low-temperature region. Panel (b) plots C/TC/T as a function of T2T^{2}, in which a linear fit (C/T=γ+βT2C/T=\gamma+\beta T^{2}) is presented for the normal state. Panel (c) shows Ce/(γT)C_{\mathrm{e}}/(\gamma T), where Ce=CβT3C_{\mathrm{e}}=C-\beta T^{3} denotes the electronic specific heat, as a function of the reduced temperature, T/TcT/T_{\mathrm{c}}. The data basically agree with a full-gap BCS α\alpha model [αΔ(0)/(kBTc)\alpha\equiv\Delta(0)/(k_{\mathrm{B}}T_{\mathrm{c}}), where Δ(0)\Delta(0) is the anisotropic superconducting gap at zero temperature] [30] assuming 77% superconducting phase and a residual electronic specific-heat coefficient of γ0=0.25γ\gamma_{0}=0.25\gamma.

Figure 6(b) shows the plot of C/TC/T versus T2T^{2}, from which the low-temperature electronic specific heat can be separated out. The linear fit gives an intercept of γ\gamma = 32.2 mJ K-2 mol-f.u.-1, corresponding to a bare density of states of N(EF)=3γ/(π2kB2)=N(E_{\mathrm{F}})=3\gamma/(\pi^{2}k_{\mathrm{B}}^{2})= 13.6 states/eV/f.u., consistent with the electronic structure calculation [27]. Note that the Sommerfeld constant of CoS2 is 21 mJ K-2 mol-1 [24, 25], somewhat smaller than the above γ\gamma value, yet it turns out to be larger on the basis of Co content. Furthermore, the CoS2 impurity is the minor phase after all. Therefore, the Sommerfeld coefficient of the CuCo2S4 phase will not change very much even if corrections due to the existence of CoS2 impurity could be reliably made. Note that the

Assuming the γ\gamma value of 32.2 mJ K-2 mol-f.u.-1 and with the electronic specific heat of Ce=CβT3C_{\mathrm{e}}=C-\beta T^{3}, figure 6(c) was plotted using Ce/(γT)C_{\mathrm{e}}/(\gamma T) and T/TcT/T_{\mathrm{c}} as the coordinates. Under the constrain of entropy conservation, i.e. 0Tc[(CeγT)/T]dT=0\int_{0}^{T_{\mathrm{c}}}[(C_{\mathrm{e}}-\gamma T)/T]\mathrm{d}T=0, a full-gap BCS α\alpha model [30] can basically fit the data with αΔ(0)/(kBTc)=\alpha\equiv\Delta(0)/(k_{\mathrm{B}}T_{\mathrm{c}})= 1.5 if a residual electronic specific-heat coefficient of γ0=0.25γ\gamma_{0}=0.25\gamma due to the existence of non-superconducting impurity phase of CoS2 is taken into account. In this circumstance, the superconducting fraction is fitted to be 77(1)%, which is conversely consistent with \sim19% non-superconducting phase.

Here we note that the single-gap BCS model does not account for the data exclusively. Other models with line energy-gap nodes are also applicable. However, the present limited data cannot distinguish which model applies. Interestingly, previous NMR investigations concluded contrasting superconducting properties in the Cu-Co-S system: one suggested a gapless superconducting state [13], the other indicated a full superconducting gap [14]. This discrepancy seems to be due to the big difference in the sample’s quality. Our present specific-heat result excludes the possibility of gapless superconductivity in the nearly stoichiometric sample of CuCo2S4. We expect that future measurements of specific-heat, NMR, and other techniques down to lower temperatures with using better samples (with less impurity) will be able to clarify the issue of the superconducting gap.

Above we have clarified that the nearly stoichiometric CuCo2S4 thiospinel is a superconductor with Pauli paramagnetism. Now let us comment on the previous dispersive results about “CuCo2S4” [11, 12, 15, 16, 18]. They can be accounted for in terms of the deviations from the stoichiometry. The actual composition of the synthesized thiospinel phase should be written as (Cu1-xCox)Co2S4-δ (because impurity phases such as CoS2 and Cu2S appeared). The S deficiency obviously decreases the hole concentration in (Cu1-xCox)Co2S4-δ, which suppresses SC. The Co/Cu substitution (Co2+ partially substitutes Cu+) not only decreases the hole concentration, but also possibly induces magnetic impurity of Co2+ at the Cu site, both of which are detrimental to SC. This could be the main reason for the difficulty to observe SC in the sample with nominally stoichiometric composition. In the Cu-rich sample of “Cu1.5Co1.5S4” [13, 14], however, the Co/Cu substitution at the Cu site is greatly reduced because Co is poor. The possible Cu occupation at the Co site may not destroy SC because of nonmagnetic Cu+. Thus SC is easily observed in the Cu-rich samples.

IV Concluding remarks

To summarize, with a novel two-step synthesis strategy, we were able to prepare nearly stoichiometric CuCo2S4 phase which shows bulk SC at 4.2 K with Pauli paramagnetism in the normal state. We have also revealed that sulfur deficiency and Co/Cu substitution is detrimental to SC, which may explain the contradictive results in previous reports. The result calls for further investigations on the rare Co-based superconductor by optimizing the sample quality (with the CoS2 impurity as less as possible) and with various measurements down to lower temperatures.

SC in Co-based compounds is very rare. This work corroborates that CuCo2S4 is another Co-based superconductor in addition to NaxCoO2y{}_{2}\cdot yH2[8]. Albeit of different crystal structures, interestingly, the two systems show many similarities including the TcT_{\mathrm{c}} value, Co coordination, formal Co valence, and the geometrical frustration. It is of great interest to clarify whether CuCo2S4 is an unconventional superconductor [31]. On the other hand, SC is not frequently found in the thiospinel compounds. However, the CuM2M_{2}S4 (M=M= Co, Rh, or Ir) family seem to be the only exception. CuRh2S4 was first discovered to be a superconductor in 1967 with TcT_{\mathrm{c}} = 4.35-4.8 K [11, 32], which was confirmed in 1990s [33]. CuIr2S4 [34] itself is not a superconductor, yet it undergoes a metal-insulator transition at 230 K accompanied with a charge ordering as well as a spin dimerization [35]. SC with TcT_{\mathrm{c}} up to 3.4 K can be induced by the suppression of the metal-insulator transition via Zn/Cu substitution [36, 37]. For the spinel selenides, SC was reported in CuRh2Se4 (Tc=T_{\mathrm{c}}= 3.5 K [11, 32]) and Cu(Ir0.8Pt0.2)2Se4 (Tc=T_{\mathrm{c}}= 1.76 K [38]). Therefore, one may expect that CuCo2Se4 could be also a superconductor if it can be synthesized with the stoichiometric composition.

Acknowledgements.
This work was supported by the National Natural Science Foundation of China (12050003), National Key Research and Development Program of China (2017YFA0303002) and the Fundamental Research Funds for the Central Universities of China.

References

  • Bednorz and Muller [1986] J. G. Bednorz and K. A. Muller, Possible high Tc\mathrm{T_{c}} superconductivity in the Ba-La-Cu-O system, Z. Phys. B 64, 189 (1986).
  • Kamihara et al. [2008] Y. Kamihara, T. Watanabe, M. Hirano,  and H. Hosono, Iron-based layered superconductor LaO1xFxFeAs\mathrm{La}\mathrm{O}_{1-x}\mathrm{F}_{x}\mathrm{Fe}\mathrm{As} (xx = 0.05-0.12) with TcT_{\mathrm{c}} = 26 K\mathrm{K}, J. Am. Chem. Soc. 130, 3296 (2008).
  • Li et al. [2019] D. Li, K. Lee, B. Y. Wang, M. Osada, S. Crossley, H. R. Lee, Y. Cui, Y. Hikita,  and H. Y. Hwang, Superconductivity in an infinite-layer nickelate, Nature 572, 624 (2019).
  • Hu et al. [2015] J. Hu, C. Le,  and X. Wu, Predicting unconventional high-temperature superconductors in trigonal bipyramidal coordinations, Phys. Rev. X 5, 041012 (2015).
  • Hu and Le [2017] J. Hu and C. Le, A possible new family of unconventional high temperature superconductors, Sci. Bull. 62, 212 (2017).
  • Le et al. [2018] C. Le, J. Zeng, Y. Gu, G.-H. Cao,  and J. Hu, A possible family of Ni-based high temperature superconductors, Sci. Bull. 63, 957 (2018).
  • Zhang et al. [2020] Q. Zhang, K. Jiang, Y. Gu,  and J. Hu, Unconventional high temperature superconductivity in cubic zinc-blende transition metal compounds, Sci. China Phys., Mech. & Astron. 63, 277411 (2020).
  • Takada et al. [2003] K. Takada, H. Sakurai, E. Takayama-Muromachi, F. Izumi, R. A. Dilanian,  and T. Sasaki, Superconductivity in two-dimensional CoO2 layers, Nature 422, 53 (2003).
  • Lotgering and van Stapele [1967] F. Lotgering and R. van Stapele, Magnetic and electrical properties of copper containing sulphides and selenides with spinel structure, Solid State Commun. 5, 143 (1967).
  • Lotgering and van Stapele [1968] F. K. Lotgering and R. P. van Stapele, Magnetic Properties and Electrical Conduction of Copper-Containing Sulfo- and Selenospinels, J. Appl. Phys. 39, 417 (1968).
  • Maaren et al. [1967] N. V. Maaren, G. Schaeffer,  and F. Lotgering, Superconductivity in sulpho- and selenospinels, Phys. Lett. A 25, 238 (1967).
  • Miyatani et al. [1993] K. Miyatani, T. Tanaka, S. Sakita, M. Ishikawa,  and N. Shirakawa, Magnetism and Superconductivity in Copper Spinels, Jpn. J. Appl. Phys. 32, 448 (1993).
  • Furukawa et al. [1995] Y. Furukawa, S. Wada, K. Miyatani, T. Tanaka, M. Fukugauchi,  and M. Ishikawa, Nuclear spin-lattice relaxation and antiferromagnetic spin correlations in superconducting thiospinel Cu1.5{\mathrm{Cu}}_{1.5}Co1.5{\mathrm{Co}}_{1.5}S4{\mathrm{S}}_{4}Phys. Rev. B 51, 6159 (1995).
  • Wada et al. [2001] S. Wada, H. Sugita, K. Miyatani, T. Tanaka,  and T. Nishikawa, Weak antiferromagnetism and superconductivity in pseudo-binary spinel compounds (Cu,Co)Co2S4 investigated by 59Co and 63Cu magnetic resonance, J. Phys.: Condens. Matt. 14, 219 (2001).
  • Wen et al. [2003] Z.-L. Wen, G.-H. Cao, C.-M. Feng, Z.-A. Xu, M.-H. Fang,  and Z.-K. Jiao, Appearance of superconductivity by doping in CuCo2S4 thiospinel system, Chin. J. Low Temp. Phys. 25, 193 (2003).
  • Aito and Sato [2004] N. Aito and M. Sato, Superconducting transition temperatures of Cu(Co1-xMx)2S4J. Phys. Soc. Jpn. 73, 1938 (2004).
  • Fang et al. [2005] L. Fang, P. Y. Zou, X. F. Lu, Z. Xu, H. Chen, L. Shan,  and H. H. Wen, Superconductivity in thiospinel Cu1.3K0.2Co1.5S4{\mathrm{Cu}}_{1.3}{\mathrm{K}}_{0.2}{\mathrm{Co}}_{1.5}{\mathrm{S}}_{4}Phys. Rev. B 71, 064505 (2005).
  • Feng et al. [2021] Z. Feng, Q. Hou, T. Li, K. Wang, H. Wu, J. Si, W. Lv, J.-Y. Ge, S. Cao, J. Zhang,  and N.-C. Yeh, Selenium doping induced two antiferromagnetic transitions in thiospinel compounds CuCo2S4-xSex (0x0.8)(0\leq x\leq 0.8)J. Am. Cer. Soc. 104, 1806 (2021).
  • Williamson and Grimes [1974] D. P. Williamson and N. W. Grimes, An x-ray diffraction investigation of sulphide spinels, J. Phys. D: Appl. Phys. 7, 1 (1974).
  • Craig et al. [1979a] J. Craig, D. Vaughan,  and J. Higgins, Phase-equilibria in the copper-cobalt-sulfur system, Mater. Res. Bull. 14, 149 (1979a).
  • Craig et al. [1979b] J. Craig, D. Vaughan,  and J. Higgins, Phase-relation in the Cu-Co-S system and mineral associations of the carrollite (CuCo2S4)-linnaeite (Co3S4) series, Economic Geology 74, 657 (1979b).
  • Bouchard et al. [1965] R. J. Bouchard, P. A. Russo,  and A. Wold, Preparation and electrical properties of some thiospinels, Inorg. Chem. 4, 685 (1965).
  • Adachi et al. [1969] K. Adachi, K. Sato,  and M. Takeda, Magnetic properties of cobalt and nickel dichalcogenide compounds with pyrite structure, J. Phys. Soc. Jpn. 26, 631 (1969).
  • Otero-Leal et al. [2008] M. Otero-Leal, F. Rivadulla, M. García-Hernández, A. Piñeiro, V. Pardo, D. Baldomir,  and J. Rivas, Effect of spin fluctuations on the thermodynamic and transport properties of the itinerant ferromagnet CoS2{\text{CoS}}_{2}Phys. Rev. B 78, 180415 (2008).
  • Teruya et al. [2016] A. Teruya, F. Suzuki, D. Aoki, F. Honda, A. Nakamura, M. Nakashima, Y. Amako, H. Harima, M. Hedo, T. Nakama,  and Y. Ōnuki, Large Cyclotron Mass and Large Ordered Moment in Ferromagnet CoS2 Compared with Paramagnet CoSe2J. Phys. Soc. Jpn. 85, 064716 (2016).
  • Toby [2001] B. H. Toby, EXPGUI, a graphical user interface for GSAS, J. Appl. Crystallography 34, 210 (2001).
  • Oda et al. [1995] T. Oda, M. Shirai, N. Suzuki,  and K. Motizuki, Electronic band structure of sulphide spinels CuM2S4 (M=Co, Rh, Ir), Journal of Physics: Condensed Matter 7, 4433 (1995).
  • Mackenzie et al. [1998] A. P. Mackenzie, R. K. W. Haselwimmer, A. W. Tyler, G. G. Lonzarich, Y. Mori, S. Nishizaki,  and Y. Maeno, Extremely Strong Dependence of Superconductivity on Disorder in Sr2RuO4{\mathrm{Sr}}_{2}{\mathrm{RuO}}_{4}Phys. Rev. Lett. 80, 161 (1998).
  • Liu et al. [2016] Y. Liu, J.-K. Bao, H.-K. Zuo, A. Ablimit, Z.-T. Tang, C.-M. Feng, Z.-W. Zhu,  and G.-H. Cao, Effect of impurity scattering on superconductivity in K2Cr3As3Sci. China Phys., Mech. & Astron. 59, 657402 (2016).
  • Johnston [2013] D. C. Johnston, Elaboration of the α\alpha-model derived from the BCS theory of superconductivity, Supercond. Sci. Technol. 26, 115011 (2013).
  • Hirsch et al. [2015] J. Hirsch, M. Maple,  and F. Marsiglio, Superconducting materials classes: Introduction and overview, Physica C: Superconductivity and its Applications 514, 1 (2015), superconducting Materials: Conventional, Unconventional and Undetermined.
  • Robbins et al. [1967] M. Robbins, R. Willens,  and R. Miller, Superconductivity in the spinels CuRh2S4 and CuRh2Se4Solid State Commun. 5, 933 (1967).
  • Hagino et al. [1995] T. Hagino, Y. Seki, N. Wada, S. Tsuji, T. Shirane, K.-i. Kumagai,  and S. Nagata, Superconductivity in spinel-type compounds CuRh2{\mathrm{CuRh}}_{2}S4{\mathrm{S}}_{4} and CuRh2{\mathrm{CuRh}}_{2}Se4{\mathrm{Se}}_{4}Phys. Rev. B 51, 12673 (1995).
  • Nagata et al. [1994] S. Nagata, T. Hagino, Y. Seki,  and T. Bitoh, Metal-insulator transition in thiospinel CuIr2S4Physica B: Condens. Matt. 194-196, 1077 (1994).
  • Radaelli et al. [2002] P. G. Radaelli, Y. Horibe, M. J. Gutmann, H. Ishibashi, C. H. Chen, R. M. Ibberson, Y. Koyama, Y.-S. Hor, V. Kiryukhin,  and S.-W. Cheong, Formation of isomorphic Ir3+ and Ir4+ octamers and spin dimerization in the spinel CuIr2S4Nature 416, 155 (2002).
  • Cao et al. [2001] G. Cao, T. Furubayashi, H. Suzuki, H. Kitazawa, T. Matsumoto,  and Y. Uwatoko, Suppression of metal-to-insulator transition and appearance of superconductivity in Cu1xZnxIr2S4{\mathrm{Cu}}_{1-x}{\mathrm{Zn}}_{x}{\mathrm{Ir}}_{2}{\mathrm{S}}_{4}Phys. Rev. B 64, 214514 (2001).
  • Suzuki et al. [1999] H. Suzuki, T. Furubayashi, G. Cao, H. Kitazawa, A. Kamimura, K. Hirata,  and T. Matsumoto, Metal-insulator transition and superconductivity in spinel-type system Cu1-xZnxIr2S4J. Phys. Soc. Jpn. 68, 2495 (1999).
  • Luo et al. [2013] H. Luo, T. Klimczuk, L. Müchler, L. Schoop, D. Hirai, M. K. Fuccillo, C. Felser,  and R. J. Cava, Superconductivity in the Cu(Ir1-xPtx)2Se4 spinel, Phys. Rev. B 87, 214510 (2013).