The discovery of another monoclinic polymorph in the niobium trisulfide system expands the structural possibilities for quasi-1D transition metal trichalcogenide materials. We describe here NbS3-VI prepared by chemical vapor transport (CVT) using ammonium chloride as the transport agent rather than the typical iodine or excess chalcogen. This example establishes precedent for transport agent control over CVT product polymorphism, thereby opening an alternative avenue for structural engineering. The single crystal x-ray diffraction structure of NbS3-VI shows that this polymorph unexpectedly incorporates features of both NbS3-IV and NbS3-V; specifically, NbS3-VI contains corrugated chains with paired Nb–Nb and uniform chains with unpaired, equidistant Nb centers. We also use single crystal x-ray diffraction to compare NbS3-VI with (Nb0.6Ti0.4)S3, which contains solely uniform chains with slightly shorter metal–metal distances than those of uniform NbS3-VI chains.

Recent advancements in the synthesis, exfoliation, and characterization of low-dimensional materials have stimulated interest in their fundamental properties and applications. The transition metal di- and trichalcogenides, MX2 and MX3 (where M = transition metal and X = chalcogen), are two classes of low-dimensional materials at the heart of these developments. Although the 2D MX2 materials have been more widely studied, quasi-one-dimensional (quasi-1D) MX3 compositions possess many attractive properties, including high anisotropy, superconductivity, semiconducting to metallic conductivity, and charge density waves (CDWs) that originate in the structure and bonding of these materials.1–7 

Crystalline MX3 structures consist of asymmetric trigonal prismatic MX6 units (M = selected transition metal; X = S, Se, Te) assembled into infinite parallel chains of face-sharing polyhedra. These chains are arranged into bilayers that associate through van der Waals interactions, leading to the quasi-1D (rather than true 1D) properties of these materials. Structural diversity arises through differences in metal–metal bonding along the chains, variations in chalcogen–chalcogen bonding within the coordination sphere of the metal, the strength of interchain interactions, the relative positions of neighboring chains, and the stacking arrangements of bilayers.4 Specifically, for transition metal trisulfides, trigonal bipyramidal asymmetry emerges from differences in the shortest S⋯S distance in each trigonal bipyramid, i.e., within the (S2)2− moiety of the M4+(S)2−(S2)2− fragment. The number of different (X2)2− interactions within a unit cell is one differentiating characteristic of the three prototype MX3 structures, i.e., the ZrSe3-, TaSe3-, and NbSe3-types, which exhibit one, two, or three distinct (X2)2− interactions, respectively. These structures also vary in the relative arrangement of MX3 chains within the bilayers.

Niobium and tantalum trisulfides display interesting and productive polymorphism, and the resulting structural variations are distinctive of 1D materials. NbS3 is significantly more diverse than any other MX3 system.5–8 NbS3-I and NbS3-IV both contain distorted NbS3 chains featuring Nb pairs or dimers, which results in alternating short and long distances between Nb centers (where short distances correspond to Nb–Nb bonding).4,8,9 This scenario is believed to effectively confine the lone conducting electron of each Nb center, resulting in semiconducting rather than metallic properties.7 This metal pairing appears to be unique to NbS3-I and -IV among all MX3 compositions.

On the other hand, NbS3-II and NbS3-V contain chains of Nb4+ at uniform distances of ∼3.35 Å.8,10,11 The literature describing NbS3-II shows that this polymorph is metallic with unusually high CDW transitions at 450–475, 360, and 150 K, which can be discerned from the superstructures revealed by selected area electron diffraction and from nonlinear transport behavior.12–14 Zybtsev et al. have reported evidence for the existence of NbS3-II subphases or polytypes designated as “low-Ohmic” and “high-Ohmic” with different electrical properties.14,15

Polymorphs are valuable for several reasons. With sufficient control over their formation, they can be harnessed to clarify key structure–property relationships. Because structural differences can impact properties to a significant degree, polymorphism also presents an opportunity to engineer materials for specific applications. Well-known examples include the carbon allotropes (diamond, graphite, C60, and nanotubes)16 and the silicon carbide polytypes (200+ known forms).17 In this latter system, 4H-SiC is an important ceramic material and also a commercialized wide bandgap semiconductor.

The primary challenge associated with polymorphs is phase pure synthesis and isolation. In the case of NbS3, all polymorphs can be prepared by chemical vapor transport (CVT), which yields needle- or ribbon-like MX3 crystals up to several centimeters in length. The recently reported NbS3-IV and NbS3-V polymorphs were prepared using elemental transport agents (I2 or Br2 for NbS3-IV, excess S for NbS3-V).8,18 Reaction temperatures and gradients influenced which major phase was isolated (670→570 °C for NbS3-IV vs 700→670 °C for NbS3-V). Here, we find that switching the transport agent to ammonium chloride, NH4Cl, can provide phase control such that solely polymorph NbS3-VI is isolated. This result is surprising because polymorph control in CVT and chemical vapor deposition (CVD) is usually accomplished via temperature, pressure, the use of crystalline seeds/substrates (epitaxy), or the application of cooling/quenching protocols at the conclusion of crystal growth.19 We are not aware of other instances where transport agent selection determines a product's atomic structure.

NH4Cl has been used as a CVT transport agent and chemical vapor deposition (CVD) “growth promoter” or “synergistic additive” for a variety of metal chalcogenides (e.g., FeS2, TiS2, VTe2, and ZnSe), metal nitrides (e.g., ZrNCl), metal oxides (e.g., ZnO, WO2, InWO3, and Cu2OSeO3), and other materials.20–23 At temperatures beyond ∼340 °C, NH4Cl(s) sublimes and decomposes into two highly effective transport species, NH3(g) and HCl(g). Ammonia may further decompose into H2(g) and N2(g) at sufficiently elevated temperatures, especially in the presence of a catalyst. Thus, in addition to serving as an in situ source of HCl(g), NH4Cl can generate an advantageous reducing atmosphere.

To produce NbS3-VI, we conducted CVT from the elements and NH4Cl at a final temperature gradient of 670→570 °C. After two weeks of growth, we isolated silver-gray, wirelike crystals of NbS3-VI. CVT at the same temperature with I2 or Br2 instead yields NbS3-IV. Compositional analyses with energy dispersive x-ray spectroscopy (EDS) showed that the experimental S/Nb atomic ratio matches the anticipated theoretical value within error (Table S1). Elemental maps of Nb and S show even distributions of both elements throughout the crystals (Fig. S1).

FIG. 1.

Structure of NbS3-VI: (a) perspective view of a NbS3 bilayer composed of corrugated chains (green) and uniform chains (blue); (b) view of corrugated and uniform NbS3 chains along their b-axis, with distances between Nb4+ indicated; (c) bilayers of corrugated and uniform NbS3 chains down their b-axis.

FIG. 1.

Structure of NbS3-VI: (a) perspective view of a NbS3 bilayer composed of corrugated chains (green) and uniform chains (blue); (b) view of corrugated and uniform NbS3 chains along their b-axis, with distances between Nb4+ indicated; (c) bilayers of corrugated and uniform NbS3 chains down their b-axis.

Close modal

Bulk structural characterization by powder x-ray diffraction (PXRD) provided experimental patterns that match well with droplines from the single crystal x-ray structure of NbS3-VI (Fig. S2) once orientation effects are accounted for. The diffraction patterns of NbS3-IV, NbS3-V, and NbS3-VI have many common features, making the polymorphs challenging to differentiate by PXRD alone. For this reason, polymorph identification in the NbS3 system must be assessed by unit cell determination.

FIG. 2.

Crystal structure of Nb0.56Ti0.44S3 (blue = Nb/Ti, yellow = S): (a) view down the chains (b-axis) and (b) view down the c-axis. Unit cell is outlined in black.

FIG. 2.

Crystal structure of Nb0.56Ti0.44S3 (blue = Nb/Ti, yellow = S): (a) view down the chains (b-axis) and (b) view down the c-axis. Unit cell is outlined in black.

Close modal

To establish the NbS3 polymorph discovered in this work, we solved the structure of NbS3-VI from single crystal x-ray diffraction data. Key crystallographic parameters and atomic coordinates are summarized in Tables S2 and S3, and Table I makes comparisons among known NbS3 structures. We find that the combination of β angle and unit cell volume (Table S4) serves to distinguish among NbS3-IV, -V, and -VI.

TABLE I.

Structural parameters for selected NbS3 polymorphs and (Nb,Ti)S3.

Unit cell parametersCorrugated chainsUniform chains
CompoundSpace groupa, b, c (Å)α, β, γ (°)Paired Nb–Nb (Å)Nb⋯Nb (Å)Nb⋯Nb/Ti (Å)Reference
NbS3-I P1¯ 4.963, 6.730, 9.144 90, 97.17, 90 3.045 3.702 ⋯ 9  
NbS3-II P21/m 9.6509 (8), 3.3459 (2), 19.850 (1) 90, 110.695 (4), 90 ⋯ ⋯ ∼3.346 11  
NbS3-IV P21/c 6.7515 (5), 4.9736 (4), 18.1315 (13) 90, 90.116 (2), 90 3.0448 (8) 3.7087 (8) ⋯ 8  
NbS3-V P21/m 4.950 (5), 3.358 (4), 9.079 (10) 90, 97.35 (2), 90 ⋯ ⋯ 3.358 (4) 8  
NbS3-VI Pm 4.961 (2), 6.743 (2), 9.137 (2) 90, 97.05 (3), 90 3.007 (6) 3.736 (6) 3.3724 (10) This work 
(Nb0.56Ti0.44)S3 P21/m 4.9751 (8), 3.3692 (6), 9.0244 (14) 90, 97.368 (4), 90 ⋯ ⋯ 3.3692 (6) This work 
Unit cell parametersCorrugated chainsUniform chains
CompoundSpace groupa, b, c (Å)α, β, γ (°)Paired Nb–Nb (Å)Nb⋯Nb (Å)Nb⋯Nb/Ti (Å)Reference
NbS3-I P1¯ 4.963, 6.730, 9.144 90, 97.17, 90 3.045 3.702 ⋯ 9  
NbS3-II P21/m 9.6509 (8), 3.3459 (2), 19.850 (1) 90, 110.695 (4), 90 ⋯ ⋯ ∼3.346 11  
NbS3-IV P21/c 6.7515 (5), 4.9736 (4), 18.1315 (13) 90, 90.116 (2), 90 3.0448 (8) 3.7087 (8) ⋯ 8  
NbS3-V P21/m 4.950 (5), 3.358 (4), 9.079 (10) 90, 97.35 (2), 90 ⋯ ⋯ 3.358 (4) 8  
NbS3-VI Pm 4.961 (2), 6.743 (2), 9.137 (2) 90, 97.05 (3), 90 3.007 (6) 3.736 (6) 3.3724 (10) This work 
(Nb0.56Ti0.44)S3 P21/m 4.9751 (8), 3.3692 (6), 9.0244 (14) 90, 97.368 (4), 90 ⋯ ⋯ 3.3692 (6) This work 

The overall unit cell dimensions of NbS3-VI are most similar to those of NbS3-I. However, NbS3-VI crystallizes in the monoclinic space group Pm rather than the triclinic space group P1¯ of NbS3-I. Like other NbS3 polymorphs, the structure of NbS3-VI is based on infinite chains of face-sharing, niobium sulfide trigonal bipyramids arranged into layers [Fig. 1(a)] separated by a ∼3.0 Å van der Waals gap. The outstanding feature of NbS3-VI is its combination of two chain types; that is, NbS3-VI is constructed of alternating “corrugated” chains (where Nb–Nb pairing leads to alternating distances between Nb4+ along the chains) and “uniform” chains (where the absence of pairing leads to regular distances between Nb4+). These chains are illustrated in Fig. 1, and Table I includes key metrics that define the chains in each structure. In particular, the distances between Nb4+ in NbS3-VI vary from as short as 3.007(6) Å within Nb2 pairs that make up corrugated chains, to as long as 3.736(6) Å between neighboring Nb2 pairs. These values suggest that Nb pairs in NbS3-VI are more strongly bonded than in NbS3-I and NbS3-IV. In comparison, the distance between Nb ions in uniform chains of NbS3-VI is 3.3724(10) Å, just slightly longer than those in NbS3-V [3.358(4) Å].

The presence of both uniform and corrugated chains in NbS3-VI results in three different S–S pairing distances: one in the uniform chains, and two in the corrugated chains. The shortest S–S distance is slightly longer in the uniform chains (2.12 Å) than those in the corrugated chains (1.987 and 2.002 Å). The longer S–S distances (long side of the S3 triangle) in the corrugated chains of NbS3-IV and VI also differ. In NbS3-VI, these S–S distances fall in the range of 3.361–3.372 or 3.822–3.897 Å, consistent with a distorted triangular base. Thus, it appears that the presence of both uniform and corrugated chains in each bilayer causes the uniform chains to be more symmetric than those in NbS3-V, whereas the corrugated chains are more distorted than those in NbS3-IV.

An interesting aspect of NbS3-VI is that its structural components—semiconducting, corrugated chains of NbS3 and metallic, uniform chains of NbS3—have distinctly different electronic character. This feature can be related conceptually to the alternating octahedral and trigonal prismatic layers incorporated within 4H- and 6R-TaX2.24,25 We also note that the mixed chains in NbS3-VI lead to “Janus” motifs where one face of the bilayer displays corrugated chains and the other face displays uniform chains. In comparison, previously described Janus metal chalcogenide nanosheets are characterized by different chalcogen substitution on each 2D face.26–28 

We expanded this study to include NbxTix−1S3, which is known to form solid solutions across its full compositional range with isovalent substitution of the slightly smaller Ti4+ (60.5 pm ionic radius) for Nb4+ (68.0 pm ionic radius).29,30 Crystals of Nb0.56Ti0.44S3 were grown by CVT using excess sulfur as the transport agent. To minimize inhomogeneous crystal nucleation and favor crystal growth at the cooler end of the ampule rather than throughout the ampule, we first applied a reverse temperature gradient (hot growth zone → cold source zone) prior to forward transport (hot source → cold growth). This process removes crystallization seeds from the walls of the ampule in the growth zone region and previously was reported to improve CVT crystal growth.31,32 For these experiments, we initially positioned the sealed ampule over the edge of the tube furnace so that the source zone was suspended outside of the furnace at a temperature of ∼25 °C, while the empty growth zone was positioned within the furnace at 600 °C. After one week, the ampule was completely inserted into the furnace and a forward gradient of 600 → 500 °C (source zone–growth zone) was established for crystal growth. This technique significantly improved the CVT growth of (Nb,Ti)S3, leading to larger and more uniform crystals.

Scanning electron microscopy (SEM) imaging revealed flat, ribbon-like crystal morphologies (Fig. S3). EDS provided an experimental composition of Nb0.56Ti0.44S3 (Table S5) and elemental mapping verified homogeneity (Fig. S3). Single crystal x-ray diffraction showed that this material crystallizes in the monoclinic space group P21/m with unit cell parameters (Table I) similar to those of both NbS3-V and TiS3.8,33,34 As illustrated in Fig. 2, each Nb0.56Ti0.44S3 unit cell contains two uniform-type chains with metal–metal spacings of 3.3692(6) Å, a value that is intermediate between the d(Nb–Nb) of 3.358(4) Å in NbS3-V and 3.3724(10) Å in NbS3-VI.8 The isosceles triangle-based prisms are consistent with the so-called A-variant structure of TiS3.33,34 The base of each triangle contains an S2 pair at 2.038 Å separation, similar to a distance of 2.039 Å in TiS3. Bridging MS linkages help to assemble neighboring chains into bilayers; the 2.638 Å interchain distance in Nb0.56Ti0.44S3 exceeds that of both NbS3-V (2.619 Å, 2.635 Å) and TiS3 (2.416 Å), reflecting the impact of Nb4+/Ti4+ substitution on interchain interactions.

FIG. 3.

SEM images showing (a) the fibrous morphology of NbS3-VI crystals and (b) their striated surface texture at higher magnification. (c) Raman spectrum of NbS3-VI (red) in comparison with NbS3-IV (blue) and NbS3-V (black).

FIG. 3.

SEM images showing (a) the fibrous morphology of NbS3-VI crystals and (b) their striated surface texture at higher magnification. (c) Raman spectrum of NbS3-VI (red) in comparison with NbS3-IV (blue) and NbS3-V (black).

Close modal

Further characterization of NbS3-VI entailed Raman spectroscopy. A representative spectrum is shown in Fig. 3, along with comparative spectra from NbS3-IV and NbS3-V. At first glance, the spectra of NbS3-VI and -IV have numerous similarities, whereas the spectrum of NbS3-V is quite different.30–32 Several strong resonances exhibited by both NbS3-IV and -VI include ∼145, ∼155, ∼189, and ∼335 cm−1, which also have been reported for NbS3-I and -II.30,32,33 By analogy, we can tentatively assign the resonance at ∼145 cm−1 as a lattice compression mode, whereas the one at ∼155 cm−1 is most likely the intrachain ν(Nb–Nb) stretching mode. The ∼189 cm−1 peak can be associated with intrachain ν(Nb–S) stretches. The presence of the lower intensity shoulder at ∼195 cm−1 as well as the absence of the two distinct ν(Nb–S2) stretches differentiate NbS3-VI and are consistent with the reduced influence of weaker ν(Nb–S) interactions with corrugated chains. The peaks appearing between 250 and 350 cm−1 are found in both NbS3-IV and -VI spectra with similar intensities. However, the peaks at 198, 375, and 565 cm−1 are significantly reduced in intensity in NbS3-VI compared to NbS3-IV, with the 550 cm−1 peak of NbS3-IV notably absent from the spectrum of NbS3-VI. This resonance is likely the ν(S–S) stretch associated with S–S pairing.

Morphological characterization of NbS3-VI by scanning electron microscopy (SEM) is shown in Fig. 3. Previous work revealed consistent variations among NbS3 polymorphs, specifically wider, more ribbon-like crystals of NbS3-I and -IV vs thinner, more wire-like crystals of NbS3-II and V.8,9,35Figure 3(a) shows that NbS3-VI possesses a fibrous morphology, with individual crystals ranging from ∼100 nm to ∼100 μm in width, and from several micrometers to centimeters in length. Most of the NbS3-VI fibers exhibit a ribbon-like morphology similar to those of NbS3-I and IV. However, wire-like crystals resembling NbS3-II and V are present in the sample, which may be due to mixed phases or mechanical cleavage during sample preparation. A closer examination of the NbS3-VI surface reveals a striated surface texture of parallel ridges [Fig. 3(b)]. These features continue along the length of the NbS3-VI ribbons (the b-axis) with varied width and spacing. This texture, as well as the sample cleavage, indicates the strong growth anisotropy along the quasi-1D NbS3 chains and the weaker interchain interactions along the a- and c-directions. Similar features are apparent in TiS3, TaSe3, and NbSe3.36–38 

High resolution (scanning) transmission electron microscopy [HR-(S)TEM] was used to visualize NbS3 crystals at the nanometer and atomic scales. Figure 4 shows NbS3-VI crystals under three transmission imaging modes: low magnification TEM, high magnification TEM, and HR-STEM. At low magnification (panel a), the exfoliated NbS3-VI appears as multilayered nanoribbons. Selected area electron diffraction (SAED) (panel b, inset) confirms the single crystalline nature of the narrower ribbon at lower right. Indexing the observed diffraction spots shows the [001] zone axis to be parallel to the imaging direction. Notably, the SAED pattern lacks the extra rows of spots corresponding to superstructures in NbS3-II and -V caused by CDW phases.11,39 Their absence suggests that this NbS3 polymorph does not exhibit CDWs at room temperature.

FIG. 4.

(a) Low magnification and (b) high magnification brightfield BF-TEM images of NbS3-VI. The inset of panel B shows an indexed SAED pattern. (c) Atomic resolution high angle annular darkfield HAADF-STEM image showing the chain structure perpendicular to the [001] zone axis. (d) Enlarged view of the boxed area of panel (d) (from inverse FFT) with observed atomic spacings. (e) Corrugated and uniform NbS3-VI chain structures that correspond to the image in panel (d), with atomic spacings from the single crystal x-ray structure.

FIG. 4.

(a) Low magnification and (b) high magnification brightfield BF-TEM images of NbS3-VI. The inset of panel B shows an indexed SAED pattern. (c) Atomic resolution high angle annular darkfield HAADF-STEM image showing the chain structure perpendicular to the [001] zone axis. (d) Enlarged view of the boxed area of panel (d) (from inverse FFT) with observed atomic spacings. (e) Corrugated and uniform NbS3-VI chain structures that correspond to the image in panel (d), with atomic spacings from the single crystal x-ray structure.

Close modal

High resolution TEM of the NbS3-VI nanoribbons reveals the superstructure visible in Fig. 4(b): evenly spaced and perpendicular linear features crossing the nanoribbon length and width. The orientation of the NbS3 chains along the length of the nanoribbon, confirmed by SAED, can be attributed to the features parallel to the nanoribbon edge. The perpendicular lines, which have been observed for NbS3-I and -IV nanoribbons as well (but not NbS3-II and -V), are most likely a result of increased electron density between neighboring chains caused by the close chain proximity and corrugated chains within NbS3 bilayers.

The structure of NbS3-VI was further clarified by atomic resolution STEM imaging. In Fig. 4(c), the orientation of the NbS3 chains runs horizontally across the image. The linear features observed by HR-TEM are still present at higher magnification, reappearing every fifth column of atoms, yet they do not appear to directly link neighboring chains along the [100] axis. At this magnification and with the aid of the enlarged view in Fig. 4(d), it is clear that the linear features seen at lower magnification are generated by overlapping chains: uniform chains in a “top row” and corrugated chains in a “bottom row,” which can be identified by distinctive spacings between the Nb4+ centers [Fig. 4(e)].

In summary, the preparation and characterization of NbS3-VI further expand the possibilities for transition metal trichalcogenide structure and bonding. Single crystal x-ray diffraction and atomic-resolution electron microscopy establish the unique features of NbS3-VI that distinguish it from other known polymorphs of NbS3, namely, NbS3-VI contains both uniform and corrugated NbS3 chains with equidistant and paired Nb4+, respectively. Access to NbS3-VI is enabled by ammonium chloride as the crystal growth transport agent, which represents an alternative approach to polymorph engineering. Although the selection of the transport agent has always been a key aspect of CVT reactions, the results described in this contribution demonstrate an additional role that it can play in the development of low-dimensional materials, especially 1D van der Waals materials.

See the supplementary material for synthetic details, crystallographic information, atomic coordinates, and additional characterization data.

This work was supported in part by a U.S. National Science Foundation Emerging Frontiers of Research Initiative 2-DARE grant (No. NSF EFRI-1433395). Transmission electron microscopy studies were carried out at EAG Materials, Netherlands, with the help of M. A. Verheijen.

The authors have no conflicts to disclose.

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

1.
A.
Geremew
,
M. A.
Bloodgood
,
E.
Aytan
,
B.
Woo
,
W. R.
Corber
,
G.
Liu
,
K.
Bozhilov
,
T. T.
Salguero
,
S.
Rumyantsev
,
M. P.
Rao
, and
A. A.
Balandin
, “
Current carrying capacity of quasi-1D ZrTe3 van der Waals nanoribbons
,”
IEEE Electron Device Lett.
39
,
735
738
(
2018
).
2.
X.
Wen
,
W.
Lei
,
L.
Ni
,
L.
Yang
,
P.
Zhang
,
Y.
Liu
,
H.
Chang
, and
W.
Zhang
, “
Evaluating the electrical characteristics of quasi-one-dimensional ZrTe3 nanoribbon interconnects
,”
ACS Appl. Electron. Mater.
3
,
4228
4235
(
2021
).
3.
J. O.
Island
,
A. J.
Molina-Mendoza
,
M.
Barawi
,
R.
Biele
,
E.
Flores
,
J. M.
Clamagirand
,
J. R.
Ares
,
C.
Sánchez
,
H. S. J.
Van Der Zant
, and
R.
D'Agosta
, “
Electronics and optoelectronics of quasi-1D layered transition metal trichalcogenides
,”
2D Mater.
4
,
022003
(
2017
).
4.
A.
Patra
and
C. S.
Rout
, “
Anisotropic quasi-one-dimensional layered transition-metal trichalcogenides: Synthesis, properties and applications
,”
RSC Adv.
10
,
36413
(
2020
).
5.
Crystal Chemistry and Properties of Materials with Quasi-One-Dimensional Structures: A Chemical and Physical Synthetic Approach
, edited by
J.
Rouxel
(
Springer
,
Boston, MA
,
1986
).
6.
P.
Monceau
, “
Electronic crystals: An experimental overview
,”
Adv. Phys.
61
,
325
(
2012
).
7.
S.
Conejeros
,
B.
Guster
,
P.
Alemany
,
J. P.
Pouget
, and
E.
Canadell
, “
Rich polymorphism of layered NbS3
,”
Chem. Mater.
33
,
5449
5463
(
2021
).
8.
M. A.
Bloodgood
,
P.
Wei
,
E.
Aytan
,
K. N.
Bozhilov
,
A. A.
Balandin
, and
T. T.
Salguero
, “
Monoclinic structures of niobium trisulfide
,”
APL Mater.
6
(
2
),
026602
(
2018
).
9.
J.
Rijnsdorp
and
F.
Jellinek
, “
The crystal structure of niobium trisulfide, NbS3
,”
J. Solid State Chem.
25
,
325
328
(
1978
).
10.
L.
Brattås
,
A.
Kjekshus
,
J.
Krogh-Moe
,
J.
Songstad
, and
Å.
Pilotti
, “
On the properties of compounds with the ZrSe3 type structure
,”
Acta Chem. Scand.
26
,
3441
3449
(
1972
).
11.
E.
Zupanič
,
H. J. P.
van Midden
,
M. A.
van Midden
,
S.
Šturm
,
E.
Tchernychova
,
V. Y.
Pokrovskii
,
S. G.
Zybtsev
,
V. F.
Nasretdinova
,
S. V.
Zaitsev-Zotov
,
W. T.
Chen
,
W. W.
Pai
,
J. C.
Bennett
, and
A.
Prodan
, “
Basic and charge density wave modulated structures of NbS3-II
,”
Phys. Rev. B
98
(
17
),
174113
(
2018
).
12.
Z. Z.
Wang
,
P.
Monceau
,
H.
Salva
,
C.
Roucau
,
L.
Guemas
, and
A.
Meerschaut
, “
Charge-density-wave transport above room temperature in a polytype of NbS3
,”
Phys. Rev. B
40
(
17
),
11589
11593
(
1989
).
13.
T.
Cornelissens
,
G.
Van Tendeloo
,
J.
Van Landuyt
, and
S.
Amelinckx
, “
Peierls distortion, chain polytypism, and dislocation coupling in NbS3
,”
Phys. Status Solidi A
48
,
K5
K9
(
1978
).
14.
S. G.
Zybtsev
,
V. Y.
Pokrovskii
,
V. F.
Nasretdinova
,
S. V.
Zaitsev-Zotov
,
V. V.
Pavlovskiy
,
A. B.
Odobesco
,
W. W.
Pai
,
M.-W.
Chu
,
Y. G.
Lin
,
E.
Zupanič
,
H.
van Midden
,
S.
Šturm
,
E.
Tchernychova
,
A.
Prodan
,
J. C.
Bennett
,
I. R.
Mukhamedshin
,
O. V.
Chernysheva
,
A. P.
Menushenkov
,
V. B.
Loginov
,
B. A.
Loginov
,
A. N.
Titov
, and
M.
Abdel-Hafiez
, “
NbS3: A unique quasi-one-dimensional conductor with three charge density wave transitions
,”
Phys. Rev. B
95
,
035110
(
2017
).
15.
S. G.
Zybtsev
,
N. Y.
Tabachkova
,
V. Y.
Pokrovskii
,
S. A.
Nikonov
,
A. A.
Maixlakh
, and
S. V.
Zaitsev-Zotov
, “
New polytype of the quasi-one-dimensional conductor NbS3 with a high-temperature charge density wave
,”
JETP Lett.
114
(
1
),
40
44
(
2021
).
16.
P. S.
Karthik
,
A. L.
Himaja
, and
S. P.
Singh
, “
Carbon-allotropes: Synthesis methods, applications and future perspectives
,”
Carbon Lett.
15
,
219
237
(
2014
).
17.
A. A.
Lebedev
, “
Heterojunctions and superlattices based on silicon carbide
,”
Semicond. Sci. Technol.
21
,
R17
R34
(
2006
).
18.
E. V.
Formo
,
J. A.
Hachtel
,
Y.
Ghafouri
,
M. A.
Bloodgood
, and
T. T.
Salguero
, “
Thermal stability of quasi-1D NbS3 nanoribbons and their transformation to 2D NbS2: Insights from in situ electron microscopy and spectroscopy
,”
Chem. Mater.
34
,
279
287
(
2022
).
19.
K.
Hernandez Ruiz
,
Z.
Wang
,
M.
Ciprian
,
M.
Zhu
,
R.
Tu
,
L.
Zhang
,
W.
Luo
,
Y.
Fan
, and
W.
Jiang
, “
Chemical vapor deposition mediated phase engineering for 2D transition metal dichalcogenides: Strategies and applications
,”
Small Sci.
2
,
2100047
(
2022
).
20.
M.
Ohashi
,
S.
Yamanaka
, and
M.
Hattori
, “
Chemical vapor transport of layer structured crystal β-ZrNCl
,”
J. Solid State Chem.
77
,
342
347
(
1988
).
21.
Y.
Steiner
, “
Phase relations and chemical vapor transport of hexagonal indium tungsten bronze InxWO3
,”
J. Alloys Compd.
605
,
96
101
(
2014
).
22.
J. R.
Panella
,
B. A.
Trump
,
G. G.
Marcus
, and
T. M.
McQueen
, “
Seeded chemical vapor transport growth of Cu2OSeO3
,”
Cryst. Growth Des.
17
,
4944
4948
(
2017
).
23.
M.
Zhao
,
X.
Niu
,
L.
Guan
,
H.
Qian
,
W.
Wang
,
J.
Sha
, and
Y.
Wang
, “
Understanding the growth of black phosphorus crystals
,”
CrystEngComm
18
,
7737
(
2016
).
24.
E.
Bjerkelund
,
A.
Kjekshus
,
Å.
Nilsson
,
J.
Sandström
,
H.
Theorell
,
R.
Blinc
,
S.
Paušak
,
L.
Ehrenberg
, and
J.
Dumanović
, “
On the structural properties of the Ta1+xSe2 phase
,”
Acta Chem. Scand.
21
,
513
526
(
1967
).
25.
R.
Huisman
and
F.
Jellinek
, “
On the polymorphism of tantalum diselenide
,”
J. Less-Common Met.
17
,
111
117
(
1969
).
26.
A.-Y.
Lu
,
H.
Zhu
,
J.
Xiao
,
C.-P.
Chuu
,
Y.
Han
,
M.-H.
Chiu
,
C.-C.
Cheng
,
W.-W.
Yang
,
K.-H.
Wei
,
Y.
Yang
,
Y.
Wang
,
D.
Sokaras
,
D.
Nordlund
,
P.
Yang
,
D. A.
Muller
,
M.-Y.
Chou
,
X.
Zhang
, and
L.-J.
Li
, “
Janus monolayers of transition metal dichalcogenides
,”
Nat. Nanotechnol.
12
,
744
749
(
2017
).
27.
R.
Chaursiya
and
A.
Dixit
, “
Defect engineered MoSSe Janus monolayer as a promising two dimensional material for NO2 and NO gas sensing
,”
Appl. Surf. Sci.
490
,
204
219
(
2019
).
28.
A. B.
Maghirang
,
Z.-Q.
Huang
,
R.
Villaos
,
C.-H.
Hsu
,
L.-Y.
Feng
,
E.
Florido
,
H.
Lin
,
A.
Bansil
, and
F.-C.
Chuang
, “
Predicting two-dimensional topological phases in Janus materials by substitutional doping in transition metal dichalcogenide monolayers
,”
npj 2D Mater. Appl.
3
,
35
(
2019
).
29.
K.
Wu
,
M.
Blei
,
B.
Chen
,
L.
Liu
,
H.
Cai
,
C.
Brayfield
,
D.
Wright
,
H.
Zhuang
, and
S.
Tongay
, “
Phase transition across anisotropic NbS3 and direct gap semiconductor TiS3 at nominal titanium alloying limit
,”
Adv. Mater.
32
,
2000018
(
2020
).
30.
R. D.
Shannon
, “
Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides
,”
Acta Crystallogr., Sect. A
32
,
751
767
(
1976
).
31.
A.
Ubaldini
,
J.
Jacimovic
,
N.
Ubrig
, and
E.
Giannini
, “
Chloride-driven chemical vapor transport method for crystal growth of transition metal dichalcogenides
,”
Cryst. Growth Des.
13
,
4453
4459
(
2013
).
32.
M.
Binnewies
,
R.
Glaum
,
M.
Schmidt
, and
P.
Schmidt
,
Chemical Vapor Transport Reactions
(
de Gruyter
,
2012
).
33.
E.
Guilmeau
,
D.
Berthebaud
,
P. R. N.
Misse
,
S.
Hébert
,
O. I.
Lebedev
,
D.
Chateigner
,
C.
Martin
, and
A.
Maignan
, “
ZrSe3-type variant of TiS3: Structure and thermoelectric properties
,”
Chem. Mater.
26
(
19
),
5585
5591
(
2014
).
34.
N. B.
Bolotina
,
I. G.
Gorlova
,
I. A.
Verin
,
A. N.
Titov
, and
A. V.
Arakcheeva
, “
Defect structure of TiS3 single crystals of the A-ZrSe3 type
,”
Crystallogr. Rep.
61
(
6
),
923
930
(
2016
).
35.
V. Y.
Pokrovskii
,
S. G.
Zybtsev
,
M. V.
Nikitin
,
I. G.
Gorlova
,
V. F.
Nasretdinova
, and
S. V.
Zaitsev-Zotov
, “
High-frequency, ‘quantum’ and electromechanical effects in quasi-one-dimensional charge density wave conductors
,”
Phys-Usp.
56
(
1
),
29
48
(
2013
).
36.
J. O.
Island
,
M.
Buscema
,
M.
Barawi
,
J. M.
Clamagirand
,
J. R.
Ares
,
C.
Sánchez
,
I. J.
Ferrer
,
G. A.
Steele
,
H. S. J.
van der Zant
, and
A.
Castellanos-Gomez
, “
Ultrahigh photoresponse of few-layer TiS3 nanoribbon transistors
,”
Adv. Opt. Mater
2
,
641
645
(
2014
).
37.
M. A.
Stolyarov
,
G.
Liu
,
M. A.
Bloodgood
,
E.
Aytan
,
C.
Jiang
,
R.
Samnakay
,
T. T.
Salguero
,
D. L.
Nika
,
S. L.
Rumyantsev
,
M. S.
Shur
,
K. N.
Bozhilov
, and
A. A.
Balandin
, “
Breakdown current density in h-BN-capped quasi-1D TaSe3 metallic nanowires: Prospects of interconnect applications
,”
Nanoscale
8
(
34
),
15774
15782
(
2016
).
38.
T.
Matsuura
,
M.
Yamanaka
,
N.
Hatakenaka
,
T.
Matsuyama
, and
S.
Tanda
, “
Topologically linked crystals
,”
J. Cryst. Growth
297
,
157
160
(
2006
).
39.
F. W.
Boswell
and
A.
Prodan
, “
Peierls distortions in NbS3 and NbSe3
,”
Physica B+C
99
,
361
364
(
1980
).

Supplementary Material