Graciela
Villalpando
a,
Jiaze
Xie
a,
Nitish
Mathur
a,
Guangming
Cheng
a,
Nan
Yao
b and
Leslie M.
Schoop
*a
aDepartment of Chemistry, Princeton University, Princeton, NJ 08544, USA. E-mail: lschoop@princeton.edu
bPrinceton Materials Institute, Princeton, NJ 08544, USA
First published on 25th July 2024
Magnetic two-dimensional (2D) materials are a unique class of quantum materials that can exhibit interesting magnetic phenomena, such as layer-dependent magnetism. The most significant barrier to 2D magnet discovery and study lies in our ability to exfoliate materials down to the monolayer limit. Therefore designing exfoliation methods that produce clean, monolayer sheets is crucial for the growth of 2D material research. In this work, we develop a facile chemical exfoliation method using lithium naphthalenide for obtaining 2D nanosheets of magnetic van der Waals material CrOCl. Using our optimized method, we obtain freestanding monolayers of CrOCl, with the thinnest measured height to date. We also provide magnetic characterization of bulk, intercalated intermediate, and nanosheet pellet CrOCl, showing that exfoliated nanosheets of CrOCl exhibit magnetic order. The results of this study highlight the tunability of the chemical exfoliation method, along with providing a simple method for obtaining 2D CrOCl.
New conceptsMagnetic two-dimensional (2D) materials are a relatively new and integral class of quantum materials that are highly sought after for both research and real-life application purposes. Despite the significance, the current library of existing air-stable magnetic monolayers is extremely limited. CrOCl, an antiferromagnet rich in magnetic features, is one of many magnetic materials yet to be exfoliated to produce freestanding monolayers. Chemical exfoliation, a top-down solution based method, has not only shown promise in producing freestanding monolayers, but has produced air-stable monolayers of magnetic materials in previous studies. In this work we provide a facile chemical exfoliation method using lithium naphthalene to produce freestanding air-stable monolayers of CrOCl for the first time. Our results provide infrastructure for future studies on monolayer CrOCl. We also provide the thinnest atomic force microscopy measurements of monolayer CrOCl obtained thus far and show how the chemically exfoliated sheets can be cleaned. We hope that this work highlights the adaptability of chemical exfoliation for difficult materials, and bolsters research into chemical exfoliation as a tool to expand the number of existing magnetic monolayers. |
Recently, chemical exfoliation, an alternative to mechanical exfoliation, has been proved a useful method for obtaining 2D or even 1D (one-dimensional) materials where traditional mechanical methods fail.5,9–11 Similar to Scotch tape exfoliation, chemical exfoliation is a top-down method typically used for layered bulk materials. One of the common chemical exfoliation approaches takes advantage of redox chemistry to intercalate (or deintercalate) ions or molecules between the layers of the materials of interest, weakening the interlayer interactions.5,12,13 Thereafter, shear forces such as shaking and/or sonication are applied to delaminate the layers, creating a 2D nanosheet suspension. The merit of this method lies in the tunability of each step, allowing for exfoliation of more complicated materials. For example, the choice of intercalent and the involved chemistry for insertion between the layers of the host material has been shown to have an immense affect on the success of the exfoliation.14,15 It is highly dependent on the reduction potential of the intercalent compared to the host material.15,16 Other factors, such as the solvents used for shaking, timing, and sonication parameters have been shown to impact the quality of exfoliation as well.10,14,17,18
Transition metal oxychlorides (TMOCl) are van der Waals layered compounds in the space group Pmmn. This family has proven valuable as cathode materials for batteries, in addition to exhibiting interesting electronic and magnetic properties such as magnetic anisotropy and Mott insulator behaviour.19–21 Unlike other van der Waals materials, the conventional mechanical methods are not efficient for the exfoliation of TMOCl compounds to the monolayer limit. Recently, by using wet chemical treatments, our lab has reported the successful isolation of monolayers or nanosheets of FeOCl and VOCl, which retained their magnetic order.10,22 The methodology developed from the previous work should be also suitable for other members of TMOCl family, allowing us to continues the exploration of more interesting low-dimensional magnetism.
CrOCl has captured interest in the 2D materials community due to the complex magnetic features of the bulk phase. This compound is an insulating antiferromagnet with in-plane anisotropy and highly complex magnetic interactions. Previous studies have shown that CrOCl has a Néel temperature of approximately 13.6 K associated with a monoclinic distortion to P21/m and an additional phase transition below 30 K due to an incommensurate magnetic state, as found by magnetic, heat capacity, and neutron measurements.23,24 Additionally, recent work using dynamic cantilever magnetometry reports a rich magnetic phase diagram of bulk CrOCl with four total magnetic phase transitions: antiferromagnetic, ferrimagnetic, incommensurate, and paramagnetic.25 The intriguing magnetism found in the bulk has cultivated significant interest in using monolayer CrOCl for strain tuned and spintronic devices.26,27 Furthermore, theoretical predictions on the magnetic character of monolayer CrOCl are ambiguous, having been calculated to be both ferromagnetic and antiferromagnetic in separate studies, making experimental studies on monolayer CrOCl valuable for refining theoretical studies as well.28,29 In literature, although this compound has been exfoliated mechanically down to the monolayer limit and monolayer edges have been obtained, freestanding monolayers of CrOCl have yet to be obtained.30 The lack of studies on freestanding monolayers makes it difficult to corroborate the ferromagnetic theoretical predictions.28,31 Taken all together, an alternative exfoliation method is necessary for decent studies of interesting 2D magnets.
While in previous studies, monolayers of other members of the TMOCl family are successfully obtained by treating with n-butyllithium,10,22 a common reducing agent, we found that n-butyllithium was not reducing enough to be effective for CrOCl. This is most likely due to the stable oxidation state of Cr3+ (low Cr3+/Cr2+ redox potential). Thus, we shift our attention to lithium naphthalenide, one of the most reducing organic agents, as a lithiation reagent, as shown in Fig. 1. In fact, alkali naphthalenide compounds are widely used in organometallic chemistry, and have historically been tested for solid-state intercalation chemistry.32,33 Additionally, previous studies by Zheng et al. have had success in chemically exfoliating MoS2 down to the monolayer level using alkali naphthalenides as an intercalant. Compared to n-butyllithium, naphthalenide salts are not only stronger reductants but also provide reversible redox reactions, namely leading to innocent and soluble naphthalene as final products. Despite these advantages, the use of lithium naphthalenide in a chemical exfoliation based synthesis of 2D transition metal oxychlorides has not been reported to our knowledge. In this study, by using lithium naphthalenide-based intercalation, followed by shaking and sonication, we successfully obtained freestanding CrOCl monolayers. Furthermore, we establish a simple route to cleaning chemically exfoliated nanosheets, addressing the issue of chemical exfoliation producing “dirty”surfaces.5 The resulting product contains monolayers with a height of 0.75 nm after cleaning, which are the thinnest sheets of CrOCl reported to date and the only freestanding monolayers to our knowledge. In addition, some magnetic characterizations are carried out for both fresh-made bulk materials and restacked nanosheet pellets.
The crystallinity of both the parent and the lithiated intermediate were assessed with powder X-ray diffraction (PXRD) as shown in Fig. 2. The parent pattern shows pure, crystalline CrOCl. The lithiated intermediates PXRD pattern indicates successful intercalation of lithium. The large peaks at low angles and the smaller peak at 4.43° (labeled with purple asterisks in Fig. 2) represent the expansion of the interlayer distance attributed to intercalation of both lithium and solvated lithium.34 The larger peak labeled with a purple asterisk at approximately 2.23° represents an increase of the interlayer gallery distance from approximately 0.78 nm in the parent to 1.12 nm, which is consistent with other Li-THF intercalated materials.35 The reflection located at 4.43° is possibly a disparity in solvent intercalation in the interlayer spacing throughout the crystal, and more closely resembles the intercalation of a non-solvated lithium ion based on the much smaller increase of interlayer distance (interlayer distance equal to 0.92 nm). However, the main peaks at 11.8° and 21.24° are still retained. Additional reflections (labeled with green triangles in Fig. 2) are consistently in the lithiated batches, however cannot be easily explained based on the quality of the data. The drastic shifting of the (001) peak and peak broadening are characteristic of successful intercalation of lithium. This is further supported by inductively coupled plasma optical emission measurements (ICP-OES) (Table S1 in the ESI†), indicating approximately a 1:1 molar ratio of chromium to lithium in the lithiated intermediate.
Nanosheet pellets were obtained from nanosheet suspensions as described in the “Methods” section below. The large low-angle peak (labeled with a purple asterisk, Fig. 2) is slightly shifted towards higher angles compared to that of the lithiated intermediate, indicating a decrease in interlayer stacking distance. This peak also contains the Warren-type tailing towards higher angles often seen in turbostratically restacked nanosheets.36 However, a large background still exists at all angles below this peak. This is common for restacked nanosheets and stems from the large spacing between the randomly restacked nanosheets.34 Additionally, a peak not found in the parent (labeled with a green triangle, Fig. 2) is retained from the lithiated intermediate along with a peak at approximately 21.24° that matches both the intermediate and parent compound. However, the quality of the PXRD from the nanosheet pellet prevents accurate structural analysis. Further structural information of the nanosheets can be found in the diffraction obtained from transmission electron microscopy (TEM) (Fig. 3 and Fig. S1 and S2, ESI†) and the high resolution scanning transmission electron microscopy (HRSTEM) (Fig. S3, ESI†).
TEM of these chemically exfoliated CrOCl nanosheets reveals large amounts of clean and thin sheets (Fig. 3, Fig. S1 and S2, ESI†). Energy-dispersive X-ray spectroscopy (EDX) demonstrates that these sheets contain chromium, oxygen, and chlorine as indicated by Fig. 3D–F below. Selected area electron diffraction (SAED) reveals that these nanosheets are also highly crystalline (Fig. 3B). Compared to that of the simulated bulk CrOCl pattern, the CrOCl nanosheets contain similar d-spacing distances (Fig. 3C). However, CrOCl nanosheet diffraction patterns also contain additional peaks that are forbidden for the parent space group (Pmmn) (as shown in Fig. S2, ESI†). This is very similar to our previous study on the nanosheets of chemically exfoliated VOCl, which showed the same diffraction pattern when exfoliated down to lower dimensions.10 These additional spots could mark a shift from orthorhombic to monoclinic, a requirement for in-plane antiferromagnetic ordering seen at lower temperatures in this compound.23,37 Alternatively, this could be a consequence of the low dimensionality of the sheets, a phenomenon that has been observed in multiple low-dimensional systems.38,39 Diffraction patterns were also obtained with increasing temperature from 25 °C to 500 °C (Fig. S3, ESI†), but no changes in the pattern were observed. This implies that the structure of these sheets is extremely robust. Additional TEM and diffraction of CrOCl nanosheets can be found in Fig. S1–S4 (ESI†).
Thickness of nanosheets obtained from the chemically exfoliated CrOCl were measured using atomic force microscopy. Atomic force microscopy measurements of a CrOCl monolayer can be found in Fig. 4 showing that a layer of CrOCl is approximately 1.15 nm per layer, which is further corroborated by the folded portion showing a step of approximately 1.16 nm and a roughness of 154.4 pm. The height from substrate to sheet of the monolayer, not being the expected height of a single layer CrOCl, is due to solvent and water absorption on the surfaces of both the sheet and substrate.36 Our sheet is thinner than previous studies of mechanically exfoliated monolayer CrOCl edges on SiO2 substrates (1.40 nm), further supporting evidence that we achieved the monolayer limit.30 A additional height measurement of a bilayer (Fig. S5, ESI†) is measured to be 1.80 nm, agreeing with previous studies of the CrOCl layer to layer distance being approximately 0.80 nm.30 The optical microscope images of these nanosheets along with additional sheets can be found in Fig. S6 (ESI†). A histogram showing the typical lateral size distribution of nanosheets made with this method are also included in Fig. S7 (ESI†).
Fig. 4 Atomic force microscopy measurement of the monolayer CrOCl shown in (A). Panel (B) and (C) show the height of a monolayer to be approximately 1.15 nm. |
In order to show the stability of the chemically exfoliated CrOCl sheets, another height measurement was conducted on the same monolayer left in ambient conditions nineteen days later. The height of this monolayer was measured to be 1.68 nm (Fig. 5) with a roughness of 107.4 pm. The increase in height and decrease in roughness is due to additional absorption of water and hydrocarbons on the surface of the sheet over time. To support this, an additional measurement was taken after plasma cleaning the same monolayer for one minute. The plasma-cleaned monolayer has a height of 0.75 nm and a roughness of 146.3 pm, the thinnest monolayer measurement of CrOCl obtained thus far and close to the approximated layer to layer distance of 0.80 nm. This is also significantly closer to the measured monolayer distance obtained from the crystal model of CrOCl (0.64 nm). HRSTEM images of pre-plasma cleaned sheets also show an amorphous layer on the surface of the sheets (Fig. S3B, ESI†). Overall, these measurements conclude that the chemically exfoliated CrOCl sheets are not air sensitive, and plasma cleaning can be a valuable tool in obtaining clean sheet surfaces and accurate atomic force microscopy height measurements without hydrocarbon or water contribution.
Fig. 5 Height measurements for the same monolayer shown in Fig. 4 after 19 days in air (A) and (B) and after 19 days in air with 1 minute of plasma cleaning with O2 (C) and (D). |
The lithiated intermediate shows a drastic change in the magnetic character compared to the bulk parent. The susceptibility measurement shows a single transition at 8 K, indicating the loss of the higher-temperature transition associated with the structural change. The field versus moment plots only show a soft “S-shape” at low temperatures (Fig. 6D inset), in addition to the loss of the metamagnetic steps found in the bulk.
The restacked NS pellet magnetic data is complementary to the lithiated intermediate, with a single magnetic transition and only a soft “S-shape” at low temperatures for the field versus moment data. However, the magnetic transition is further decreased to 2.74 K (Fig. 6E inset). The Curie–Weiss fit (Fig. S8, ESI†) indicates a Curie–Weiss temperature of −42.26 K, implying antiferromagnetic interactions and a calculated μeff of 3.35μB. The latter value is close to the expected value of 3.87μB for spin-only Cr3+.
The absence of the second, higher transition temperature in both the lithiated intermediate and restacked NS pellet most likely stems from the lack of a structural transition. Based on the similar XRD patterns for both the lithiated intermediate and restacked nanosheet pellet (Fig. 2), and the SAED for the nanosheets (Fig. 3B and C), we can infer that the intermediate and nanosheets have monoclinic symmetry, thereby eliminating the structural transition usually seen in bulk CrOCl, thus destroying the higher temperature magnetic transition. The lack of hysteresis and metamagnetic steps in both the lithiated intermediate and the restacked NS pellet could be a consequence of these measurements being performed with no field orientation. The magnetic properties of bulk CrOCl are highly anisotropic, therefore the lack of field orientation could suppress hysteresis in the lithiated and restacked NS pellet measurements. Another possibility is the decrease in interlayer interactions in the lithiated intermediate and the restacked NS pellet could have broken the magnetic ordering that leads to these features in the bulk. Since the bulk only shows these metamagnetic features with the field aligned along the stacking axis, either of these arguments could contribute to the loss of the characteristic metamagnetic steps and hysteresis loops. Although our measurements were not performed on CrOCl monolayers, the magnetic characterization of the restacked NS pellet still shows a magnetic moment high enough to imply that the whole sample is contributing to the magnetism. Additionally, the magnetic transition obtained appears to be highly air stable, even after measuring another restacked nanosheet pellet exposed to air and bromine for over 24 hours (Fig. S9, ESI†). The results support that the exfoliated sheets are still magnetic. Certainly, in order to obtain a more comprehensive characterization of the magnetic nature of 2D/monolayer CrOCl, careful measurements on a single monolayer of CrOCl will be required, such as magneto optical Kerr effect (MOKE) measurements.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4nh00137k |
This journal is © The Royal Society of Chemistry 2024 |