Next Article in Journal / Special Issue
Estimation of the Grüneisen Parameter of High-Entropy Alloy-Type Functional Materials: The Cases of REO0.7F0.3BiS2 and MTe
Previous Article in Journal
Mixtures of Dipolar Gases in Two Dimensions: A Quantum Monte Carlo Study
Previous Article in Special Issue
Electrical Transport Measurements on Layered La(O,F)BiS2 under Extremely High Pressure
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Superconductivity in the α-Form Layer Structured Metal Nitride Halide

1
Graduate School of Engineering, Kyushu Institute of Technology, Kitakyushu 804-8550, Japan
2
Graduate School of Natural Science and Technology, Okayama University, Okayama 700-8530, Japan
3
Research Institute for Interdisciplinary Science (RIIS), Okayama University, Okayama 700-8530, Japan
*
Author to whom correspondence should be addressed.
Submission received: 7 February 2022 / Revised: 27 February 2022 / Accepted: 28 February 2022 / Published: 1 April 2022
(This article belongs to the Special Issue Layered Superconductors III)

Abstract

:
Layered metal nitride halides MNX (M = Ti, Zr, Hf; X = Cl, Br, I) have two polymorphs, including α- and β-forms, which have the FeOCl and SmSI structures, respectively. These compounds are band insulators and become metals and show superconductivity after electron doping by intercalating alkali metals between the layers. The superconductivity of β-form had been extensively characterized from decades ago, but it is not easy to consistently interpret all experimental results using conventional phonon-mediated Bardeen–Cooper–Schriefer mechanisms. The titanium compound TiNCl crystallizes only in the α-form structure. TiNCl also exhibits superconductivity as high as ~16 K after electron doping by intercalating metals and/or organic basis. It is important to compare the superconductivity of different M–N networks. However, α-form compounds are vulnerable to moisture, unlike β-form ones. The intercalation compounds are even more sensitive to humid air. Thus, there are few experimental studies on the superconducting mechanism of α-form, although it has been discussed for exotic Cooper-pairing mechanisms. This short review gathers the recent progress in experimental studies of TiNCl.

1. Introduction

Transition metal nitride halides MNX (M = Ti, Zr, Hf; X = Cl, Br, I) belong to a characteristic class of layered materials that exhibits superconductivity by carrier doping [1,2,3]. They have two polymorphs, including α- and β-forms, which have the FeOCl and SmSI structures, respectively [4,5]. These compounds are band insulators and can be intercalated with alkali or alkaline earth metals to significantly modify the physical properties without changing their basic crystal structures. Alkali atoms with/without organic molecules are intercalated into the van der Waals gap between two adjacent X layers with the realignment of MNX layers.
The β-form polymorph contains double honeycomb-like MN layers sandwiched by halogen layers, as shown in Figure 1a,b. Band-insulating β-HfNCl is electron-doped and becomes metallic through intercalation [2,4]. The compound shows superconductivity at a transition temperature (Tc) above 25.5 K in Li and tetrahydrofuran (THF) cointercalated compound of β-HfNCl [2,4,6]; this family of materials is categorized as high-Tc superconductors. The superconductivity can also be induced by carrier doping with an electronic-double-layer transistor [7,8,9,10,11]. The Tc enhancement has been reported with reducing carrier density [12,13] and the partial substitution M site [14]. Electronic structural studies have shown that alkali-metal intercalation adds quasi-two-dimensional electronic states near the Fermi level (EF), which are hybridized states of Hf 5d and N 2p orbitals [15,16,17,18,19]. The absence of isotope effect, extremely large superconducting gap value 2Δ/kBTc, enhancement of Tc for lower carrier concentration, anomalous doping dependence in anisotropy of superconducting gap, and strong suppression of the coherence peak in nuclear spin-lattice relaxation rate strongly suggest an exotic superconducting state in this material series [13,20,21,22,23,24,25,26,27,28]. With the extensively characterized superconductivity in β-form MNX, it is not easy to consistently interpret all experimental results using conventional phonon-mediated Bardeen–Cooper–Schrieffer (BCS) mechanisms [29].
The α-form polymorph has an orthogonal MN layer network separated by halogen layers, as shown in Figure 1c.d. The titanium compound (TiNX) crystallizes only in the α-form structure. TiNCl has been reported to exhibit superconductivity as high as ~16 K after electron doping by intercalating metals and/or organic basis [3,31,32]. The crystal structure of the sodium intercalated TiNCl is also shown in Figure 1e. It is important to compare the superconductivity of different M–N networks. However, the α-form polymorphs were considered less good host materials than β-forms. The α-form compounds are vulnerable to moisture, unlike β-form ones; the intercalation compounds are even more sensitive to humid air. Moreover, compared to β-form, they can easily decompose into binary M-N compounds, which also shows superconductivity. These facts were also a disadvantage for the physical properties measurements. Therefore, most studies on superconductivity deal with intercalation compounds of the β-forms. In the recent study, the α-MNX is expected to achieve new applications utilizing two-dimensional nature, such as solar-cell [33], photocatalyst [34,35], energy storage and conversion [36], topological insulator [37,38,39], electronic devices [40], and optoelectronic devices [41]. However, all these suggestions are based on computational first-principle calculations. Only a few experimental studies of MNX have been found recently, even if we expand the searching range to β-forms [10,14,42]. The lack of experimental results is somewhat due to the difficulties in synthesizing samples of both the parent and the electron-doped α-forms.
From experimental viewpoints, we have studied the superconductivity of α-form MNX, TiNCl. This article briefly reviews the preparation and intercalation reactions in TiNCl and its superconductivities. This article will mainly focus on the recent experimental studies on the electronic structure developed in the last decade. The crystal chemistry, the structural influence of the superconductivity, and the superconductivity of β-form are not included. Reviews of these topics are presented in [4,29,43], and so on.

2. Synthesis and Electron Doping of TiNCl

Juza et al. prepared various pure Ti- and Zr-nitride halides during the 1960s [44,45,46], and the preparation of TiNCl was firstly reported in 1964 [5]. To understand how it is tricky to handle, let us look back at the synthesizing process in this section.

2.1. Conventional Preparation

For the synthesis of MNX, the tetrahalides of titanium, zirconium, and hafnium are reacted with ammonia following the generic equation:
MX4 + 4NH3MNX + 3NH4X
In the case of M = Zr, Hf in the β-form, Yamanaka has efficiently developed the method using a reaction of metal or metal hydride powders with NH4X [3]. The one-step method in a vacuum-sealed glass tube can also be adopted in a small-scale synthesis of less than ~200 mg [2].
However, in the case of M = Ti, especially TiNCl, the easy process cannot be used. In the reaction between Ti and NH4Cl, TiH2 is first formed at about 400 °C or lower. Above the temperature at which TiH2 can be decomposed, TiNCl is not formed, but TiN is formed instead. Then, the ammonia (NH3) gas should be passed over titanium tetrachloride (TiCl4) in a vertical Pyrex glass reaction cell. The reactant becomes black solid after being self-heated by the heat evolved from the intense exothermic reaction. Due to the high reactivity of TiCl4 with NH3, the reaction may lead to molecules and complexes of several manners [47]. TiCl4 ammonolysis is energetically favorable [47,48]. TiCl4 forms ammonia-incorporated complexes with the TiCl4·xNH3 gross composition, and heating the complexes in ammonia-rich atmosphere results in stepwise amination of TiCl4 to form TiCl(NH2)3 in the final step [49,50,51]. The ammonia complexes and aminated TiCl4 were especially the main products at low temperatures below 400 °C [49], while TiN is obtained at higher temperatures. The resulting solids were continuously heated in the same reaction cell at 400 °C for 3–5 h under the ammonia gas stream. Ammonium chloride (NH4Cl) crystals were grown on the upper part of the reaction cell.
The TiNCl precursor and other related compounds (crude mixture) were formed at the bottom of the glass cell. The by-products necessitate a next purification step. The crude mixture was vacuum-sealed in a Pyrex glass tube with a small amount of NH4Cl and put into a horizontal furnace with a temperature gradient for one to two weeks. Then, it was purified into highly crystalline TiNCl by chemical transport using NH4Cl. The end of the glass tube containing the mixture was maintained at 380 °C in the lower temperature zone, and the other end of the glass tube was placed in the higher temperature zone. Purified TiNCl was obtained at the higher temperature zone [52]. TiNCl starts to decompose to TiN at temperatures above 400 °C [53,54]. Once it decomposes into TiN with trivalent titanium, it is hard to return to tetravalent TiNCl. Then, it should be maintained at a temperature as low as possible above 400 °C to obtain impurity-free samples.
Assuming the same reaction as the case of ZrNCl [55], the essential reaction in the chemical vapor transport of TiNCl is expressed as follows:
TiCl4(g) + NH3(g) → TiCl3(NH2) + HCl → TiNCl(s) + 3HCl(g) (>~400 °C)
Since HCl is not easy to handle in this preparation, it is used as a transport agent by utilizing the dissociation of NH4Cl into NH3 and HCl at high temperatures. The actual reaction is then expressed as
TiCl3(NH2) + NH4Cl → TiNCl(s) + NH3(g) + HCl(g)
After these reactions, we can confirm the purity of the obtained TiNCl by the Rietveld refinement and magnetization measurement with the negligible diamagnetic contribution of TiN below 5 K.

2.2. Preparation Using Sodium Amide

In a study on a superconducting mechanism, the isotope effect on Tc provides crucial evidence of whether electron–phonon interaction is responsible for superconductivity. Considering the above conventional synthesizing process of TiNCl, retaining the 15N isotope using NH3 gas as a nitrogen source is extremely difficult or almost impossible realistically. However, if the TiNCl precursor could be formed using a solid reagent other than gas NH3 as a nitrogen source in a closed reaction cell, an essential technique for synthesizing 15N substituents would be developed. It could be led not only to elucidate the superconducting mechanism of TiNCl but also to investigate the isotope effect of other metal nitride superconductors.
NaNH2 is a highly reactive solid nitrogen source used for combustion reaction with 3d transition metal chlorides to obtain metal nitrides [56,57]. The reactions are thermodynamically driven by a metathesis reaction involving the formation of stable NaCl. The NaCl does not contribute to chemical transport; if the reaction with TiCl4 can be controlled, TiNCl may be obtained in the sealed glass tube.
When the sealed glass tube containing TiCl4 and NaNH2 was gradually heated up to 150 °C in 12 h and then kept at this temperature for 1-2 days, the infrared spectroscopy, thermogravimetry, and X-ray diffraction measurements of the resulting solids suggested that it contains TiCl3(NH2), TiCl2(NH2)2, and TiCl(NH2), together with TiCl4·xNH3 [58]. Namely, the reaction at this process can be written as the equation TiCl4 + NaNH2 → TiCl3(NH2) + NaCl in the simplest case. After an important process of removing HCl from the reaction system, we can obtain TiNCl with the whole equation in the simplest case written in TiCl4 + NaNH2 → TiCl3(NH2) + NaCl → TiNCl + NaCl + 2HCl.
The procedures to obtain TiNCl using NaNH2 can be summarized as the following. (1) The sealed mixture of TiCl4 + NaNH2 is sufficiently reacted around 160 °C to obtain preliminary reacted solids. (2) Open the seal and purge unreacted TiCl4 from the glass tube. Then, evacuate the preliminary reacted solids again with the elevated temperature at 180 °C to remove HCl from TiCl3(NH2) and/or other aminated compounds. (3) The heating and evacuation should be stopped before decomposition to TiN occurs and sealed again into a glass tube. (4) The glass tube sends to chemical transport to obtain highly crystalline TiNCl [58].

2.3. Electron Doping by Intercalation

Metal intercalation into TiNCl was developed from an electrochemical process as a cathode material [59]. First, superconducting TiNCl has been achieved by an organic Lewis base pyridine (Py) and alkali metal intercalation [3]. The alkali-metal-intercalated TiNCl was prepared using reactions with alkali metal azides at elevated temperatures of 280–350 °C in a vacuum. The intercalation was carried out in the following manner:
TiNCl + (x + y) AN3 > AyTiNCl1-x + xACl + 3(x + y)/2N2
This reaction was not mild, and partial deintercalation of chlorine atoms occurred simultaneously. The superconducting volume fractions (VF) were as low as 0.1–30%. The samples may have suffered from serious structural damage.
A chemical intercalation method using alkali-metal naphthalene solutions in THF (A-Naph/THF; A = Li, Na, K, Rb) has been used for β-phase for a long time [60]. This method has dramatically improved the low VF problem [32]. Organic molecules such as THF can be cointercalated into TiNCl at room temperature. The THF-cointercalated phase was changed into uncointercalated compound AxTiNCl by evacuation. The sample thus obtained showed a shielding VF close to 100% and a clear high onset superconducting transition. By this method, the same α-phase TiNBr has also succeeded in being a superconductor [61]. On the other hand, α-HfNBr isomorphous with TiNCl did not show superconductivity in the same method. The resistivity was larger than ~107 Ὠ cm; the doped electrons seem to be localized on the reduced Hf sites [62]. It should be noted that only Py intercalated TiNCl showed superconductivity. It appears the combination of guest donors and the amount of doping are important to realize the superconductivity of the α-form MNX.
The Tc changes with the basal spacing d along the c axis of the intercalation compound. As shown in Figure 2, the Tc decreases with the increasing d [32]. This fact is in sharp contrast to the observation in β-form, where Tc increases linearly with d in electron-doped ZrNCl and HfNCl [25,63]. The data fit on a linear line with passing through the origin. This fact suggests that the Coulomb interlayer coupling in the pairing mechanism is important in this system [32]. Another study on the α-TiNBr has demonstrated a similar linear relation on the Tc vs. 1/d [61]. Harshman et al. reported that the linear dependence is readily explained, giving the interlayer Coulomb interaction model and structure [64]. It is also a future task to experimentally observe the effect of the interlayer interaction on the electronic structure.
More recently, it has been found that superconducting properties are also highly dependent on the synthesis temperature of the pristine TiNCl. Tc and volume fraction strongly depend on the higher temperature in the synthesis of the pristine TiNCl. This fact is presumed due to hydrogen incorporation by X-ray diffraction, photoemission spectroscopy, optical spectroscopy, and hydrogen desorption measurements (to be published).
In addition to the above, the mechanism and electronic structure of superconductivity on organic molecules alone without any metal cation are completely unknown. Further experimental studies on intercalated TiNCl are required for a deeper understanding.

3. Electronic Structure of TiNCl

In order to understand the pairing mechanism of this possible exotic superconductor, it is important to clarify how electronic structure evolves with carrier doping. The band structure of pristine TiNCl has been calculated by first-principle calculations [3,65]. It exhibits a two-dimensional feature, with the general flatness of the band along the Ґ-Z direction perpendicular to the layers. The top valence band primarily has a N 2px character, and the lowest conduction band has a strong Ti 3dxy character [65]. There is 3d weight in the valence bands and N weight in the conduction bands, respectively, reflecting substantial N 2p—Ti 3d hybridization in addition to the ionic character reflected in their formal charges [65]. The β-form compounds have disconnected cylindrical Fermi surfaces favorable for nesting [66,67], while the electron-doped TiNCl has a single oval Fermi surface centered at the Ґ point [65]; the possible nesting between Fermi surfaces discussed for the β-form compounds should be excluded from TiNCl.
Several theoretical ideas are considered in arguing against phonon-mediated superconductivity. Yin et al. discussed the remaining possibilities of spin and charge fluctuations for the high-temperature superconductivity by constructing a many-body extended Hubbard model based on a realistic band structure calculation [65]. They reported that charge fluctuations might play an important role in superconductivity rather than spin fluctuations [65].
On the other hand, Kusakabe showed a criterion for superconductivity by a pair-hopping mechanism for potassium-doped TiNCl [68]. The super-pair hopping process of the spin is possible by providing the connecting orbitals between the two-dimensional electron systems. The locally formed cooper pair in a superconducting layer can be scattered from the neighboring layer via the high-energy intermediate state connecting the layers [68].
However, an electron–phonon mechanism may still be relevant for explaining the high Tc [29]. Yin et al. proposed that the Tc of TiNCl may also be well accounted for the electron–phonon mechanism using hybrid functional instead of local-density approximation (LDA) or generalized-gradient approximation (GGA) functionals in the first principles calculation [69].
In order to verify these theoretical studies, it is important to determine the electronic structure experimentally. A few experimental studies on the superconducting mechanism have been reported up to date. The reduced superconducting gap value 2Δ/kBTc has been determined from scanning tunneling microscopy/spectroscopy to be approximately 15 and 12 for KxTiNCl and NaxTiNCl, respectively [22,70]. The observed value is four times larger than the mean-field BCS value.
A photoemission spectroscopy (PES) study is the most suitable probe for observations in electronic structure, such as band dispersion. However, angle-resolved photoemission spectroscopy (ARPES) has not been achieved so far for both superconducting α- and β-forms. Although there are several reports of PES for β-form [17,18,19,71], the electronic structure studies using PES of neither the parent α-form compound nor the doped one have been reported. This problem is mainly due to the difficulty of synthesizing both the parent and the electron-doped α-form samples. However, it would gradually become solved in recent years.

3.1. Micro-PES of Electron-Doped TiNCl

For the first data of the PES study in α-form samples, micro-photoemission spectroscopy (μ-PES) with highly focused synchrotron soft X-ray was performed by Kataoka et al. [30]. In the μ-PES, the angle integrated PES spectra corresponding to the electronic density of states (DOS) of the valence band of TiNCl and NaxTiNCl were observed, as shown in Figure 3 [30]. The shape of the valence band of TiNCl roughly corresponds to that of band structure calculations. The peaks around 6 eV are hybridized states of Ti 3d, N 2p, and Cl 3p orbitals, and the 4 eV shoulder structure is dominated by the Cl 3p component [30].
The valence band spectrum of NaxTiNCl exhibits a peak at 6.5 eV, accompanied by an emergence of the metallic state near EF extending to 1.5 eV. The broad-spectrum has higher binding energy by 0.5 eV than the corresponding band of TiNCl, in line with the chemical potential shift expected from electron doping. However, the new structure at EF extends to 1.5 eV, contradicting a previous study that reported the shape of the valence and conduction band is similar to that of TiNCl [3]. It is difficult to explain the observed valence band structure of NaxTiNCl, especially the spectral structure near EF, only by the rigid band shift of that of TiNCl.
An electronic band for the appearance of superconductivity is still unclear in this material. We have to wait for the result of ARPES to clarify it, but at present, the presence of the fine structure required to exhibit superconductivity has been suggested from a PES study using hard X-ray [58]. TiNCl prepared by the above-mentioned synthetic route using NaNH2 appeared highly electron-doped, albeit without superconductivity. As an investigation of the origin, it became clear that there is a slight difference in Ti 2p. In comparing the core-level spectrum of Ti 2p with Na-intercalated samples, a fine structure around 455.5 eV emerges other than the shoulder-like component of the main structure of Na0.1TiNCl. Although the area of the fine structure is small, peak fitting can be performed reliably, as shown in Figure 4 [58]. We can see a small contribution of component D in the peak fitting of Na0.1TiNCl. However, it could not be observed in pristine non-superconducting samples even if it is in a highly electron-doped state. It did not appear unless the sample was Na-doped [58]. The difference in the preparation method changes the atmosphere during the synthesis. It may also change the hydrogen content of the resulting material. The electron doping in the TiNCl prepared by NaNH2 may be attributed to the hydrogen content, affecting the oxidation state. Moreover, in the measurement of valence band spectra, the intensity of EF and the area of component D in the doping dependence are correspondent. These findings suggest that even with the same electron doping, the electronic structure may change depending on the effective carrier insertion, determining whether or not superconductivity appears. Overall, component D seems to be a necessary structure for the appearance of superconductivity.

3.2. Metalization of TiNCl Induced by Soft X-ray Irradiation

In the β-form MNX, the carrier control has also been performed by off stoichiometry or deintercalation of X [72]. Since TiNCl is easily thermally decomposed into TiN by annealing [49,54,73], such a carrier control is challenging, but irradiation of soft X-rays can achieve the metalization [74].
In the time courses of soft X-rays irradiation, the peak area of Ti4+ in the Ti 2p core-level spectrum decreased with the irradiation time, and the intensity of the shoulder structure also increased gradually (Figure 5). The intensity of the Cl 2p spectrum after the soft X-ray irradiation decreased gradually, which suggests the irradiation induces the desorption of Cl atoms from the surface [74]. As a result, a clear Fermi edge structure was observed in the valence band spectrum after 150 min irradiation suggesting the metallic nature of the measured region of the sample surface, as shown in Figure 5e [74].
Figure 6 shows the temperature dependence of magnetic susceptibility of the pristine and irradiated samples. A weak diamagnetic signal was observed around 2 K in the pristine sample. The signal is attributed to that from TiN appearing by decomposition of TiNCl. Apart from the TiN, we can see that the other small diamagnetization near 14 K in the sample irradiated with soft X-rays for 30 min in the beam diameter region of ~20 μm (denoted as A). The signal becomes large when the area is expanded to about 0.5 × 0.5 mm2 by repeating irradiation for 3 min (denoted as B). When the decomposition of TiNCl is promoted by the irradiation, the TiN content increases, which appears as a two-step transition resulting in the enhancement of the total VF.
On the other hand, it should be noted that the VF at 5 K, which is above the Tc of TiN, becomes several times larger, as shown in the inset of Figure 6. These facts suggest that the surface region irradiated with soft X-rays is electron-doped by Cl deintercalation and becomes metallic and superconducting. This technique is significant because it enables carrier control by deintercalation in α-form, which was not achieved with conventional synthetic methods. It can be expected to further the exploration of new functionality in α-form.

4. Conclusions

More than ten years have passed since discovering superconductivity in TiNCl in 2009. However, experimental research on α-form compounds has only just begun, despite the expectation of fantastic functionality utilizing two-dimensionality. We have developed synthetic methods in recent years, and the research has started reigniting. Spectroscopic measurements have gradually become possible even for α-form intercalated superconducting samples, which are far more moisture sensitive than β-form compounds. This progress has revealed that there are still unquantified questions such as the amount of hydrogen content for the superconducting properties. Observations of band dispersion by ARPES with further technological innovation would elucidate the whole picture of exotic superconductivity by comparing the electronic structure for both α- and β-forms with different M-N layers. Further development of synthetic methods, including the carrier control by X-ray irradiation, may also be expected to acquire new functionality utilizing the two-dimensionality of TiNCl.

Funding

This research was funded by the Japan Society for the Promotion of Science (JSPS) through JSPS KAKENHI Grant Number JP18K04707, JP18KK0076 and 22K05289.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Acknowledgments

The authors express their immeasurable gratitude to Emeritus Professor Shoji Yamanaka of Hiroshima University for his helpful suggestions throughout this study. The main results on TiNCl presented in this article were partly supported by the Japan Society for the Promotion of Science (JSPS) through JSPS KAKENHI Grant Number JP18K04707 and Promotion of Joint International Research (B) (JP18KK0076) from the Ministry of Education, Culture, Sports, Science and Technology of Japan (MEXT). The photoemission studies were performed at Spring-8 BL25SU under proposal numbers 2019B1454 and 2020A1670. Preliminary studies on soft X-ray irradiation-induced superconductivity were performed at BL5, HiSOR.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yamanaka, S.; Kawaji, H.; Hotehama, K.; Ohashi, M. A new layer-structured nitride superconductor. Lithium-intercalated β-zirconium nitride chloride, LixZrNCl. Adv. Mater. 1996, 8, 771–774. [Google Scholar] [CrossRef]
  2. Yamanaka, S.; Hotehama, K.I.; Kawaji, H. Superconductivity at 25.5K in electron-doped layered hafnium nitride. Nature 1998, 392, 580–582. [Google Scholar] [CrossRef]
  3. Yamanaka, S.; Yasunaga, T.; Yamaguchi, K.; Tagawa, M. Structure and superconductivity of the intercalation compounds of TiNCl with pyridine and alkali metals as intercalants. J. Mater. Chem. 2009, 19, 2573. [Google Scholar] [CrossRef]
  4. Yamanaka, S. Intercalation and superconductivity in ternary layer structured metal nitride halides (MNX: M = Ti, Zr, Hf; X = Cl, Br, I). J. Mater. Chem. 2010, 20, 2922–2933. [Google Scholar] [CrossRef]
  5. Juza, R.; Heners, J. Über Nitridhalogenide des Titans und Zirkons. Z. Für Anorg. Und Allg. Chem. 1964, 332, 159–172. [Google Scholar] [CrossRef]
  6. Zhang, S.; Tanaka, M.; Zhu, H.; Yamanaka, S. Superconductivity of layered β-HfNCl with varying electron-doping concentrations and interlayer spacings. Supercond. Sci. Technol. 2013, 26, 085015. [Google Scholar] [CrossRef]
  7. Ueno, K.; Shimotani, H.; Yuan, H.; Ye, J.; Kawasaki, M.; Iwasa, Y. Field-induced superconductivity in electric double layer transistors. J. Phys. Soc. Jpn. 2014, 83, 032001. [Google Scholar] [CrossRef]
  8. Saito, Y.; Kasahara, Y.; Ye, J.; Iwasa, Y.; Nojima, T. Metallic ground state in an ion-gated two-dimensional superconductor. Science 2015, 350, 409–413. [Google Scholar] [CrossRef] [Green Version]
  9. Zhang, S.; Gao, M.-R.; Fu, H.-Y.; Wang, X.-M.; Ren, Z.-A.; Chen, G.-F. Electric Field Induced Permanent Superconductivity in Layered Metal Nitride Chlorides HfNCl and ZrNCl. Chin. Phys. Lett. 2018, 35, 097401. [Google Scholar] [CrossRef] [Green Version]
  10. Wang, X.; Zhang, S.; Fu, H.; Gao, M.; Ren, Z.; Chen, G. Dominant role of processing temperature in electric field induced superconductivity in layered ZrNBr. N. J. Phys. 2019, 21, 023002. [Google Scholar] [CrossRef]
  11. Nakagawa, Y.; Kasahara, Y.; Nomoto, T.; Arita, R.; Nojima, T.; Iwasa, Y. Gate-controlled BCS-BEC crossover in a two-dimensional superconductor. Science 2021, 372, 190–195. [Google Scholar] [CrossRef] [PubMed]
  12. Nakagawa, Y.; Saito, Y.; Nojima, T.; Inumaru, K.; Yamanaka, S.; Kasahara, Y.; Iwasa, Y. Gate-controlled low carrier density superconductors: Toward the two-dimensional BCS-BEC crossover. Phys. Rev. B 2018, 98, 064512. [Google Scholar] [CrossRef] [Green Version]
  13. Taguchi, Y.; Kitora, A.; Iwasa, Y. Increase in Tc upon reduction of doping in LixZrNCl superconductors. Phys. Rev. Lett. 2006, 97, 107001. [Google Scholar] [CrossRef] [PubMed]
  14. Peng, J.; Zhang, S. Synthesis and Superconductivity of Electron-Doped β-ZrNCl with Partial Substitution of Ti on Zr Site. J. Supercond. Nov. Magn. 2018, 31, 61–65. [Google Scholar] [CrossRef]
  15. Tou, H.; Maniwa, Y.; Koiwasaki, T.; Yamanaka, S. Unconventional superconductivity in electron-doped layered Li0.48(THF)yHfNCl. Phys. Rev. Lett. 2001, 86, 5775–5778. [Google Scholar] [CrossRef]
  16. Taguchi, Y.; Hisakabe, M.; Iwasa, Y. Specific heat measurement of the layered nitride superconductor LixZrNCl. Phys. Rev. Lett. 2005, 94, 2–5. [Google Scholar] [CrossRef]
  17. Yokoya, T.; Ishiwata, Y.; Shin, S.; Shamoto, S.; Iizawa, K.; Kajitani, T.; Hase, I.; Takahashi, T. Changes of electronic structure across the insulator-to-metal transition of quasi-two-dimensional Na-intercalated β-HfNCl studied by photoemission and X-ray absorption. Phys. Rev. B 2001, 64, 153107. [Google Scholar] [CrossRef]
  18. Takeuchi, T.; Tsuda, S.; Yokoya, T.; Tsukamoto, T.; Shin, S.; Hirai, A.; Shamoto, S.; Kajitani, T. Soft X-ray emission and high-resolution photoemission study of quasi-two-dimensional superconductor NaxHfNCl. Physica C 2003, 392–396, 127–129. [Google Scholar] [CrossRef]
  19. Yokoya, T.; Takeuchi, T.; Tsuda, S.; Kiss, T.; Higuchi, T.; Shin, S.; Iizawa, K.; Shamoto, S.; Kajitani, T.; Takahashi, T. Valence-band photoemission study of β-ZrNCl and the quasi-two-dimensional superconductor NaxZrNCl. Phys. Rev. B 2004, 70, 193103. [Google Scholar] [CrossRef]
  20. Tou, H.; Maniwa, Y.; Yamanaka, S. Superconducting characteristics in electron-doped layered hafnium nitride: 15N isotope effect studies. Phys. Rev. B 2003, 67, 100509. [Google Scholar] [CrossRef] [Green Version]
  21. Weht, R.; Filippetti, A.; Pickett, W.E. Electron doping in the honeycomb bilayer superconductors (Zr, Hf)NCl. Europhys. Lett. 1999, 48, 320–325. [Google Scholar] [CrossRef] [Green Version]
  22. Sugimoto, A.; Sakai, Y.; Ekino, T.; Zhang, S.; Tanaka, M.; Yamanaka, S.; Gabovich, A.M. Scanning Tunnelling Microscopy and Spectroscopy of the Layered Nitride Superconductor α-NaxTiNCl. Phys. Procedia 2016, 81, 73–76. [Google Scholar] [CrossRef] [Green Version]
  23. Ekino, T.; Takasaki, T.; Muranaka, T.; Fujii, H.; Akimitsu, J.; Yamanaka, S. Tunneling spectroscopy of MgB2 and Li0.5(THF)yHfNCl. Phys. B Condens. Matter 2003, 328, 23–25. [Google Scholar] [CrossRef]
  24. Kawaji, H.; Hotehama, K.I.; Yamanaka, S. Superconductivity of Alkali Metal Intercalated β-Zirconium Nitride Chloride, AxZrNCl (A = Li, Na, K). Chem. Mater. 1997, 9, 2127–2130. [Google Scholar] [CrossRef]
  25. Takano, T.; Kishiume, T.; Taguchi, Y.; Iwasa, Y. Interlayer-spacing dependence of Tc in LixMyHfNCl (M: Molecule) superconductors. Phys. Rev. Lett. 2008, 100, 247005. [Google Scholar] [CrossRef]
  26. Kasahara, Y.; Kishiume, T.; Kobayashi, K.; Taguchi, Y.; Iwasa, Y. Superconductivity in molecule-intercalated LixZrNCl with variable interlayer spacing. Phys. Rev. B 2010, 82, 054504. [Google Scholar] [CrossRef]
  27. Kasahara, Y.; Kishiume, T.; Takano, T.; Kobayashi, K.; Matsuoka, E.; Onodera, H.; Kuroki, K.; Taguchi, Y.; Iwasa, Y. Enhancement of Pairing Interaction and Magnetic Fluctuations toward a Band Insulator in an Electron-Doped LixZrNCl Superconductor. Phys. Rev. Lett. 2009, 103, 077004. [Google Scholar] [CrossRef] [Green Version]
  28. Kotegawa, H.; Oshiro, S.; Shimizu, Y.; Tou, H.; Kasahara, Y.; Kishiume, T.; Taguchi, Y.; Iwasa, Y. Strong suppression of coherence effect and appearance of pseudogap in the layered nitride superconductor LixZrNCl: 91Zr- and 15N-NMR studies. Phys. Rev. B 2014, 90, 020503. [Google Scholar] [CrossRef] [Green Version]
  29. Kasahara, Y.; Kuroki, K.; Yamanaka, S.; Taguchi, Y. Unconventional superconductivity in electron-doped layered metal nitride halides MNX (M = Ti, Zr, Hf; X = Cl, Br, I). Physica C 2015, 514, 354–367. [Google Scholar] [CrossRef] [Green Version]
  30. Kataoka, N.; Terashima, K.; Tanaka, M.; Hosoda, W.; Taniguchi, T.; Wakita, T.; Muraoka, Y.; Yokoya, T. µ-PES Studies on TiNCl and Quasi-two-dimensional Superconductor Na-intercalated TiNCl. J. Phys. Soc. Jpn. 2019, 88, 104709. [Google Scholar] [CrossRef] [Green Version]
  31. Yamanaka, S.; Umemoto, K.; Zheng, Z.; Suzuki, Y.; Matsui, H.; Toyota, N.; Inumaru, K. Preparation and superconductivity of intercalation compounds of TiNCl with aliphatic amines. J. Mater. Chem. 2012, 22, 10752–10762. [Google Scholar] [CrossRef]
  32. Zhang, S.; Tanaka, M.; Yamanaka, S. Superconductivity in electron-doped layered TiNCl with variable interlayer coupling. Phys. Rev. B 2012, 86, 024516. [Google Scholar] [CrossRef] [Green Version]
  33. Liang, Y.; Dai, Y.; Ma, Y.; Ju, L.; Wei, W.; Huang, B. Novel titanium nitride halide TiNX (X = F, Cl, Br) monolayers: Potential materials for highly efficient excitonic solar cells. J. Mater. Chem. A 2018, 6, 2073–2080. [Google Scholar] [CrossRef]
  34. Liu, J.; Li, X.-B.; Wang, D.; Liu, H.; Peng, P.; Liu, L.-M. Single-layer Group-IVB nitride halides as promising photocatalysts. J. Mater. Chem. A 2014, 2, 6755. [Google Scholar] [CrossRef]
  35. Zhou, L.; Zhuo, Z.; Kou, L.; Du, A.; Tretiak, S. Computational Dissection of Two-Dimensional Rectangular Titanium Mononitride TiN: Auxetics and Promises for Photocatalysis. Nano Lett. 2017, 17, 4466–4472. [Google Scholar] [CrossRef]
  36. Zhang, X.; Zhang, Z.; Zhao, X.; Wu, D.; Zhang, X.; Zhou, Z. Tetragonal-structured anisotropic 2D metal nitride monolayers and their halides with versatile promises in energy storage and conversion. J. Mater. Chem. A 2017, 5, 2870–2875. [Google Scholar] [CrossRef]
  37. Wang, A.; Wang, Z.; Du, A.; Zhao, M. Band inversion and topological aspects in a TiNI monolayer. Phys. Chem. Chem. Phys. 2016, 18, 22154–22159. [Google Scholar] [CrossRef]
  38. Zhang, Z.; Zhang, R.-W.; Li, X.; Koepernik, K.; Yao, Y.; Zhang, H. High-Throughput Screening and Automated Processing toward Novel Topological Insulators. J. Phys. Chem. Lett. 2018, 9, 6224–6231. [Google Scholar] [CrossRef]
  39. Li, X.; Zhang, Z.; Yao, Y.; Zhang, H. High throughput screening for two-dimensional topological insulators. 2D Mater. 2018, 5, 045023. [Google Scholar] [CrossRef]
  40. Rostami Osanloo, M.; Saadat, A.; Van de Put, M.L.; Laturia, A.; Vandenberghe, W.G. Transition-metal nitride halide dielectrics for transition-metal dichalcogenide transistors. Nanoscale 2022, 14, 157–165. [Google Scholar] [CrossRef]
  41. Hossain, M.M.; Naqib, S.H. Structural, elastic, electronic, andoptical properties of layered TiNX (X = F, Cl, Br, I) compounds: A density functional theory study. Mol. Phys. 2020, 118, e1609706. [Google Scholar] [CrossRef] [Green Version]
  42. Guo, Y.; Peng, J.; Qin, W.; Zeng, J.; Zhao, J.; Wu, J.; Chu, W.; Wang, L.; Wu, C.; Xie, Y. Freestanding Cubic ZrN Single-Crystalline Films with Two-Dimensional Superconductivity. J. Am. Chem. Soc. 2019, 141, 10183–10187. [Google Scholar] [CrossRef] [PubMed]
  43. Schurz, C.M.; Shlyk, L.; Schleid, T.; Niewa, R. Superconducting nitride halides MNX (M = Ti, Zr, Hf; X = Cl, Br, I). Z. Krist. 2011, 226, 395–416. [Google Scholar] [CrossRef]
  44. Juza, R.; Klose, W. Die Kristallstruktur des Zirkonnitridjodids. Angew. Chem. 1959, 71, 161. [Google Scholar] [CrossRef]
  45. Juza, R.; Friedrichsen, H. Die Kristallstruktur von β-ZrNCl und β-ZrNBr. Z. Anorg. Allg. Chem. 1964, 332, 173–178. [Google Scholar] [CrossRef]
  46. Juza, R.; Klose, W. Über ein Nitridjodid des Zirkons. Z. Anorg. Allg. Chem. Anorg. Allg. Chem. 1964, 327, 207–214. [Google Scholar] [CrossRef]
  47. Cross, J.B.; Schlegel, H.B. Molecular orbital studies of titanium nitride chemical vapor deposition: Gas phase β -elimination. Chem. Phys. Lett. 2001, 340, 343–347. [Google Scholar] [CrossRef]
  48. Umanskii, S.Y.; Novoselov, K.P.; Minushev, A.K.; Siodmiak, M.; Frenking, G.; Korkin, A.A. Thermodynamics and kinetics of initial gas phase reactions in chemical vapor deposition of titanium nitride. Theoretical study of TiCl4 ammonolysis. J. Comput. Chem. 2001, 22, 1366–1376. [Google Scholar] [CrossRef]
  49. Saeki, Y.; Matsuzaki, R.; Yajima, A.; Akiyama, M. Reaction Process of Titanium Tetrachloride with Ammonia in the Vapor Phase and Properties of the Titanium Nitride Formed. Bull. Chem. Soc. Jpn. 1982, 55, 3193–3196. [Google Scholar] [CrossRef] [Green Version]
  50. Fowles, G.W.A.; Pollard, F.H. Studies on the behaviour of halides of the transition metals with ammonia. Part II. The reaction of titanium tetrachloride with ammonia. J. Chem. Soc. 1953, 22, 2588. [Google Scholar] [CrossRef]
  51. Kurtz, S.R.; Gordon, R.G. Chemical vapor deposition of titanium nitride at low temperatures. Thin Solid Films 1986, 140, 277–290. [Google Scholar] [CrossRef]
  52. Ohashi, M.; Yamanaka, S.; Hattori, M. Chemical vapor transport of layer structured crystal β-ZrNCl. J. Solid State Chem. 1988, 77, 342–347. [Google Scholar] [CrossRef]
  53. Yajima, A.; Segawa, Y.; Matsuzaki, R.; Saeki, Y. Reaction Process of Zirconium Tetrachloride with Ammonia in the Vapor Phase and Properties of the Zirconium Nitride Formed. Bull. Chem. Soc. Jpn. 1983, 56, 2638–2642. [Google Scholar] [CrossRef] [Green Version]
  54. Sosnov, E.A.; Malkov, A.A.; Malygin, A.A. Chemical transformations at the silica surface upon sequential interactions with titanium tetrachloride and ammonia vapors. Russ. J. Gen. Chem. 2015, 85, 2533–2540. [Google Scholar] [CrossRef]
  55. Ohashi, M.; Yamanaka, S.; Hattori, M. Synthesis of β-ZrClN by Thermal Decomposition of Zirconium(IV) Amide Trichloride. Bull. Chem. Soc. Jpn. 1986, 59, 2627–2628. [Google Scholar] [CrossRef] [Green Version]
  56. Odahara, J.; Sun, W.; Miura, A.; Rosero-Navarro, N.C.; Nagao, M.; Tanaka, I.; Ceder, G.; Tadanaga, K. Self-Combustion Synthesis of Novel Metastable Ternary Molybdenum Nitrides. ACS Mater. Lett. 2019, 1, 64–70. [Google Scholar] [CrossRef]
  57. Miura, A. Low-temperature synthesis and rational design of nitrides and oxynitrides for novel functional material development. J. Ceram. Soc. Jpn. 2017, 125, 552–558. [Google Scholar] [CrossRef] [Green Version]
  58. Tanaka, M.; Kataoka, N.; Matsumoto, R.; Inumaru, K.; Takano, Y.; Yokoya, T. Synthetic Route of Layered Titanium Nitride Chloride TiNCl Using Sodium Amide. ACS Omega 2022, 7, 6375–6380. [Google Scholar] [CrossRef]
  59. Kuhn, A.; Hoppe, H.; Strähle, J.; Garcia-Alvarado, F. Electrochemical Lithium Intercalation in Titanium Nitride Chloride. J. Electrochem. Soc. 2004, 151, A843. [Google Scholar] [CrossRef]
  60. Ohashi, M.; Uyeoka, K.; Yamanaka, S.; Hattori, M. Co-Intercalation of Tetrahydrofuran and Propylene Carbonate with Alkali Metals in β-ZrNCl Layer Structured Crystal. Bull. Chem. Soc. Jpn. 1991, 64, 2814–2818. [Google Scholar] [CrossRef] [Green Version]
  61. Zhang, S.; Tanaka, M.; Watanabe, E.; Zhu, H.; Inumaru, K.; Yamanaka, S. Superconductivity of alkali metal intercalated TiNBr with alpha-type nitride layers. Supercond. Sci. Technol. 2013, 26, 122001. [Google Scholar] [CrossRef]
  62. Yamanaka, S.; Okumura, H.; Zhu, L. Alkali metal intercalation in layer structured α-HfNBr. J. Phys. Chem. Solids 2004, 65, 565–569. [Google Scholar] [CrossRef]
  63. Hotehama, K.; Koiwasaki, T.; Umemoto, K.; Yamanaka, S.; Tou, H. Effect of Swelling on the Superconducting Characteristics in Electron-Doped β-ZrNCl and HfNCl. J. Phys. Soc. Jpn. 2010, 79, 014707. [Google Scholar] [CrossRef]
  64. Harshman, D.R.; Fiory, A.T. Modeling Intercalated Group-4-Metal Nitride Halide Superconductivity with Interlayer Coulomb Coupling. J. Supercond. Nov. Magn. 2015, 28, 2967–2978. [Google Scholar] [CrossRef] [Green Version]
  65. Yin, Q.; Ylvisaker, E.R.; Pickett, W.E. Spin and charge fluctuations in α-structure layered nitride superconductors. Phys. Rev. B 2011, 83, 014509. [Google Scholar] [CrossRef] [Green Version]
  66. Felser, C.; Seshadri, R. Electronic structures and instabilities of ZrNCl and HfNCl: Implications for superconductivity in the doped compounds. J. Mater. Chem. 1999, 9, 459–464. [Google Scholar] [CrossRef]
  67. Kuroki, K. Spin-fluctuation-mediated d+id’ pairing mechanism in doped β-MNCl (M=Hf, Zr) superconductors. Phys. Rev. B 2010, 81, 104502. [Google Scholar] [CrossRef]
  68. Kusakabe, K. Pair-hopping mechanism of superconductivity activated by the nano-space layered structure. J. Phys. Chem. Solids 2012, 73, 1546–1549. [Google Scholar] [CrossRef]
  69. Yin, Z.P.; Kutepov, A.; Kotliar, G. Correlation-Enhanced Electron-Phonon Coupling: Applications of GW and Screened Hybrid Functional to Bismuthates, Chloronitrides, and Other High-Tc Superconductors. Phys. Rev. X 2013, 3, 021011. [Google Scholar] [CrossRef] [Green Version]
  70. Sugimoto, A.; Shohara, K.; Ekino, T.; Zheng, Z.; Yamanaka, S. Nanoscale electronic structure of the layered nitride superconductors α-KxTiNCl and β-HfNCl y observed by scanning tunneling microscopy and spectroscopy. Phys. Rev. B 2012, 85, 144517. [Google Scholar] [CrossRef] [Green Version]
  71. Ino, A.; Yamazaki, K.; Yamasaki, T.; Higashiguchi, M.; Shimada, K.; Namatame, H.; Taniguchi, M.; Oguchi, T.; Chen, X.; Yamanaka, S. Angle-resolved-photoemission study of layer-structured nitride β-HfNCl. J. Electron Spectros. Relat. Phenom. 2005, 144–147, 667–669. [Google Scholar] [CrossRef]
  72. Zhu, L.; Ohashi, M.; Yamanaka, S. Zirconium nitride derived from layer-structured β-ZrNCl by deintercalation of chlorine layers. Chem. Mater. 2002, 14, 4517–4521. [Google Scholar] [CrossRef]
  73. Hegde, R.I.; Fiordalice, R.W.; Tobin, P.J. TiNCl formation during low-temperature, low-pressure chemical vapor deposition of TiN. Appl. Phys. Lett. 1993, 62, 2326–2328. [Google Scholar] [CrossRef]
  74. Kataoka, N.; Tanaka, M.; Hosoda, W.; Taniguchi, T.; Fujimori, S.; Wakita, T.; Muraoka, Y.; Yokoya, T. Soft X-ray irradiation induced metallization of layered TiNCl. J. Phys. Condens. Matter 2021, 33, 035501. [Google Scholar] [CrossRef]
Figure 1. Crystal structure of MNX: (a,b) the top and side views of β-MNX, (c,d) the top and side views of α-MNX, respectively, and (e) the side view of Na-intercalated TiNCl. Upon the intercalation, TiNCl shows a polytype shift. The lattice type changes from the space group Pmmn of the pristine TiNCl to Bmmb. Reproduced with permission from Reference [30]. © (2019) The Physical Society of Japan.
Figure 1. Crystal structure of MNX: (a,b) the top and side views of β-MNX, (c,d) the top and side views of α-MNX, respectively, and (e) the side view of Na-intercalated TiNCl. Upon the intercalation, TiNCl shows a polytype shift. The lattice type changes from the space group Pmmn of the pristine TiNCl to Bmmb. Reproduced with permission from Reference [30]. © (2019) The Physical Society of Japan.
Condensedmatter 07 00033 g001
Figure 2. Tc vs. 1/d for AxTiNCl (A = Li, Na, K, Rb) superconductors with and without cointercalation. Reprinted with permission from Ref. [32]. Copyright 2022 American Physical Society.
Figure 2. Tc vs. 1/d for AxTiNCl (A = Li, Na, K, Rb) superconductors with and without cointercalation. Reprinted with permission from Ref. [32]. Copyright 2022 American Physical Society.
Condensedmatter 07 00033 g002
Figure 3. Valence band photoemission spectra of TiNCl (blue open circles) and NaxTiNCl (red open circles), together with calculated total and partial DOS of TiNCl. EF of the calculation is set to the conduction band’s bottom and shifted to 1.3 eV toward the higher binding energy side. Reproduced with permission from Reference [30]. © (2019) The Physical Society of Japan.
Figure 3. Valence band photoemission spectra of TiNCl (blue open circles) and NaxTiNCl (red open circles), together with calculated total and partial DOS of TiNCl. EF of the calculation is set to the conduction band’s bottom and shifted to 1.3 eV toward the higher binding energy side. Reproduced with permission from Reference [30]. © (2019) The Physical Society of Japan.
Condensedmatter 07 00033 g003
Figure 4. A result of peak fitting of Ti 2p hard X-ray PES spectra in Na0.1TiNCl. The inset is an enlargement scale of component D. Although the area is small, it does not appear in non-superconducting TiNCl, even if it is highly electron-doped. Reprinted from Ref. [58].
Figure 4. A result of peak fitting of Ti 2p hard X-ray PES spectra in Na0.1TiNCl. The inset is an enlargement scale of component D. Although the area is small, it does not appear in non-superconducting TiNCl, even if it is highly electron-doped. Reprinted from Ref. [58].
Condensedmatter 07 00033 g004
Figure 5. Soft X-ray irradiation time dependence of PES spectra. (a) Ti 2p core-level spectra irradiated from the initial to 150 min every 25 min. (b) Difference between the initial and 150 min of the spectra in (a). (c,d) are analogous spectra to (a) for the N 1s and Cl 2p core-level, respectively. (e) The valence band PES spectra. Inset shows the enlargement of the spectra near EF. Reprinted with permission from Ref. [74]. Copyright 2022 IOP Publishing.
Figure 5. Soft X-ray irradiation time dependence of PES spectra. (a) Ti 2p core-level spectra irradiated from the initial to 150 min every 25 min. (b) Difference between the initial and 150 min of the spectra in (a). (c,d) are analogous spectra to (a) for the N 1s and Cl 2p core-level, respectively. (e) The valence band PES spectra. Inset shows the enlargement of the spectra near EF. Reprinted with permission from Ref. [74]. Copyright 2022 IOP Publishing.
Condensedmatter 07 00033 g005
Figure 6. Temperature dependence of magnetic susceptibility in the zero-field-cooling mode of pristine TiNCl compared with soft X-ray irradiated samples. The irradiation conditions are: (A) 30 min in beam spot area, (B) 0.5 × 0.5 mm2 by repeating 3 min irradiation. The pristine sample did not show any specific magnetic behavior except for a tiny contribution of TiN below 5 K. The irradiated samples showed a weak diamagnetic signal around 14 K. The inset shows the enlargement scale.
Figure 6. Temperature dependence of magnetic susceptibility in the zero-field-cooling mode of pristine TiNCl compared with soft X-ray irradiated samples. The irradiation conditions are: (A) 30 min in beam spot area, (B) 0.5 × 0.5 mm2 by repeating 3 min irradiation. The pristine sample did not show any specific magnetic behavior except for a tiny contribution of TiN below 5 K. The irradiated samples showed a weak diamagnetic signal around 14 K. The inset shows the enlargement scale.
Condensedmatter 07 00033 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Tanaka, M.; Kataoka, N.; Yokoya, T. Superconductivity in the α-Form Layer Structured Metal Nitride Halide. Condens. Matter 2022, 7, 33. https://0-doi-org.brum.beds.ac.uk/10.3390/condmat7020033

AMA Style

Tanaka M, Kataoka N, Yokoya T. Superconductivity in the α-Form Layer Structured Metal Nitride Halide. Condensed Matter. 2022; 7(2):33. https://0-doi-org.brum.beds.ac.uk/10.3390/condmat7020033

Chicago/Turabian Style

Tanaka, Masashi, Noriyuki Kataoka, and Takayoshi Yokoya. 2022. "Superconductivity in the α-Form Layer Structured Metal Nitride Halide" Condensed Matter 7, no. 2: 33. https://0-doi-org.brum.beds.ac.uk/10.3390/condmat7020033

Article Metrics

Back to TopTop