Huge Absorption Edge Blueshifts of Layered α-MoO3 Crystals upon

5 days ago - User Resources. About Us · ACS Members · Librarians · Authors & Reviewers · Website Demos · Privacy Policy · Mobile Site ...
0 downloads 0 Views 2MB Size
Subscriber access provided by UNIV OF DURHAM

C: Physical Processes in Nanomaterials and Nanostructures 3

Huge Absorption Edge Blueshifts of Layered #-MoO Crystals upon Thickness Reduction Approaching 2D Nanosheets Hongfei Liu, Coryl Jing Jun Lee, Yunjiang Jin, Jing Yang, Chengyuan Yang, and Dongzhi Chi J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.8b03340 • Publication Date (Web): 15 May 2018 Downloaded from http://pubs.acs.org on May 16, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Huge Absorption Edge Blueshifts of Layered α-MoO3 Crystals upon Thickness Reduction Approaching 2D Nanosheets Hongfei Liu*, Coryl J. J. Lee, Yunjiang Jin, Jing Yang, Chengyuan Yang, Dongzhi Chi Institute of Materials Research and Engineering (IMRE), A*STAR (Agency for Science, Technology and Research), 2 Fusionopolis Way, Singapore 138634, Singapore Abstract: Recent theoretical studies suggest none or minor changes in the bandgap of two-dimensional (2D) α-MoO3 nanosheets as compared with that of the bulk due to the weak interlayer electronic interactions. Unfortunately, this suggestion is lacking positive support in the literature. Herein, we report experimental observations of huge blueshifts in the absorption edge of layered α-MoO3 as its thickness t is reduced approaching atomic layers. When t > 10 nm, every order of magnitude of thickness reduction gives rise to a blueshift of ~0.29 eV without causing any Raman mode shifts. This blueshift, in terms of finite-difference-time-domain (FDTD) calculations, is attributable to optical interferences at the crystal surfaces. However, when t is further reduced below ~10 nm, an even larger blueshift, accompanied by a mode softening of the most strengthened MoO-Mo stretching phonon (Ag), has been observed. This observation is consistent with those of 2D α-MoO3 nanosheets produced by aqueous exfoliations and, based on our calculations of the electronic structures, can be explained as anisotropic in-plane strain relaxations/redistributions. A gas-phase layer-by-layer etching of the layered α-MoO3 single crystals has also been demonstrated for consequent fabrications of novel electronic devices, as well as their integrations, based on α-MoO3 and other 2D nanosheets.

*Author to whom correspondence should be addressed; electronic mail: [email protected]

1

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

I. Introduction Orthorhombic MoO3, i.e., α-MoO3 (JCPDF: 05-0508) with conventional cell axes and angles of a = 3.9628 Å, b = 13.855 Å, c =3.6964 Å and α = β = γ = 90°,1 has been widely investigated for applications in catalytic chemistry as well as in electronic/optoelectronic devices. In recent years, α-MoO3 has been attracting increasing research interest concerning its two-dimensional (2D) applications driven by ever-growing demand for transparent, flexible, and wearable electronics as well as energy storage devices.2-6 The van der Waals lattice structure of α-MoO3 along its [010] crystal axis makes it easily exfoliated into atomic layers. Carrier mobilities up to 1100 cm2/Vs have been obtained for α-MoO3 with its van der Waals lattice intercalated by hydrogen,7 which may have important consequences when fabricating high speed nanoelectronics based on 2D materials.8-9

Compared with the tremendous research effort given to transition metal dichalcogenide (TMDC) based semiconducting 2D materials such as MoS2 and WS2,10-15 researches on 2D MoO3 are still at the beginning stage.4, 9, 16 It has been well known that when reducing the thickness of TMDC crystals to a few atomic layers, their electronic bandgap increases and eventually converts from indirect to direct at the thickness of single layer due to quantum confinements.10,

17

Similar thickness-reduction-induced

bandgap increases have also been observed and well understood in 2D blackphosphorous (BP) and -arsenic (BA).18-20 In comparison, recent theoretical calculations suggest a minor or absence of such layer-thickness-dependent bandgap changes for 2D MoO3 due to its weaker van der Waals interactions than those of TMDCs and BP/BA.21

2

ACS Paragon Plus Environment

Page 2 of 30

Page 3 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

However, experimental evidence that can support this theoretical suggestion is missing in the literature.21-22 On the contrary, the absorption spectra collected from α-MoO3 nanosheets prepared by bovine serum albumin (BSA) assisted aqueous exfoliations or by oxidation of few-layer MoS2 have shown remarkable blueshifts in the absorption edge with the decrease in the crystal thickness.4, 23-24 On the other hand, in contrast to the absence of thickness-reduction-induced phonon shifts of 2D α-MoO3 claimed by Wang et al.,22 a redshift has been observed by Cai et al. for the Raman feature at ~818 cm-1 of 2D α-MoO3 as compared with that of α-MoO3 bulk and the redshift was attributed to an inplane lattice expansion of the few-layer α-MoO3.3

In this paper, we provide direct evidence that the optical bandgap energy, EOpt, derived from the absorption edge of α-MoO3 crystals synthesized by thermal vapor transport (TVT), increases by up to 1.5 eV as the layer thickness decreases from a few microns approaching a few nanometers. The bandgap increments have further been discussed in terms of two mechanisms. One is the interference effect caused by light interactions with the layered crystal at the surfaces, which is more appearance at larger layer thickness (i.e., t > 10 nm) and the other is related to lattice relaxations that occur at the thickness of a few atomic layers. The much larger bandgap changes upon thickness reductions of the obtained 2D α-MoO3 than those reported by theoretical calculations are attributable to anisotropic in-plane strain relaxations/redistributions that manifested themselves as the formation of regular [001]-oriented micro-cracks. We also demonstrate that high-quality α-MoO3 belt crystals can be thinned down via a gas-phase layer-bylayer etching, which provides an alternative routine to prepare 2D α-MoO3 nanosheets for their fundamental studies and device applications.

3

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

II. Materials and method The α-MoO3 crystals, in the form of layered belts, were synthesized by TVT in a quartz tube reactor of a tube-furnace without employing intentional substrates. The source material, i.e., MoO3 powder, was loaded in an alumina crucible located at the zone center. At 1000 °C, the vaporized MoO3 was carried by pure nitrogen, transferred to the downstream areas, and crystallized in the form of layered belts at the areas of ≤ 600 °C.4 The belt crystals are collected from the inner wall of the quartz tube reactor and disassembled on sapphire substrates for optical studies. Because of their belt shapes, the disassembled α-MoO3 crystals tend to lie on their belt surfaces, i.e., with their (010) atomic planes lying on the substrates. To thin down the disassembled α-MoO3 belt crystals, argon- and oxygen-plasma treatments as well as gas-phase XeF2- and HFetching were carried out at room temperature. The details of the crystal disassembling, plasma treatments and gas-phase etching are supplied in the Supporting Information.

Single-crystal X-ray diffractometer (SCXRD, Kappa APEX) was used to identify the ‘single-crystallinity’ characteristic of an individual α-MoO3 belt while a general-area detector diffraction system (GADDS, Bruker-D8) was used to identify the crystal phases of gathered α-MoO3 belts that were pressed on a sample holder. Optical studies, i.e., absorbance and Raman scattering, were carried out with the incident light along the surface normal direction of the individual α-MoO3 belt crystals. Their thicknesses were estimated from the interference fringes in the absorbance spectra employing the refractive index of α-MoO3 available in the literature. The estimated values were further validated by step-profiler and/or atomic-force microscopy (AFM). The absorbance spectra were

4

ACS Paragon Plus Environment

Page 4 of 30

Page 5 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

collected using an UV-VIS-NIR micro-spectrophotometer (QDI 2010, CRAIC Technologies) while the Raman spectra were collected in a backscattering configuration under a confocal micro-Raman system (Witec alpha 300 system equipped with a 532-nm wavelength argon ion laser). Morphological and spectral evolutions of the α-MoO3 belts upon plasma treatments and gas-phase chemical etching were also evaluated by optical microscopy, scanning-electron microscopy (SEM), AFM, absorbance and Raman scattering spectroscopies.

III. Results and discussion Figure 1 summarizes the synthesis [Fig. 1(a)], morphological [Figs. 1(b)-1(d)] and structural properties [Figs. 1(e) and 1(f)] of the MoO3 belt crystals. The typical SEM image recorded from the gathered crystals [Fig. 1(b)] and the microphotographs recorded from the disassembled crystals [Figs. 1(c) and 1(d)] show that the MoO3 crystals are of belt structures and they prefer to lie with their belt surfaces parallel to that of the substrate. The belt-to-belt contrasts in Fig. 1(c) and the location-to-location ones of the single belt in Fig. 1(d) provide clues for thickness comparisons, especially those of the single belt that have the advantages of being verified by step-profiler and/or AFM. Figure 1(e) presents the XRD patterns collected by SCXRD from a single belt and those collected by GADDS from gathered belts [see Fig. 1(b)]; their comparisons indicate that the MoO3 belts are typically lying on their (010) atomic planes. Figure 1(f) shows the lattice structure worked out from the SCXRD measurements, which is indeed α-MoO3 that consists of ‘monolayers’, stacking along their [010]-axis via van der Waals forces. The lattice parameters are a = 3.9527 Å, b = 13.8146 Å, c = 3.6903 Å and α = β = γ = 90°.

5

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Their comparisons with the cell parameters reported in JCPDF 05-0508 reveal that the obtained α-MoO3 belt crystals are compressively stressed in their lattices; the inherent lattice strains are εa = -0.25%, εb = -0.29%, and εc = -0.17%, i.e., anisotropic.

Because of the parallel lying and the feasible thickness measurements of the disassembled α-MoO3 belt crystals, their absorption and Raman spectra were collected to address their thickness dependencies. Figure 2(a) presents the absorbance spectra collected from the belt crystals with their thickness monotonically decreases from A to G. The fringes below the absorption edges originate due to light interferences at the crystal surfaces, from which the crystal thickness can be estimated with the optical refractive index of α-MoO3 available from the literature.25 The estimated thicknesses were further validated by step-profiler and/or AFM measurements, especially for the thinner α-MoO3 layers, e.g., samples F and G, where the interference fringes are not enough for the thickness estimations. One sees that the absorption edge in Fig. 2(a) shifts to larger photon energies as the layer thickness decreases. To quantify the optical bandgap energies (EOpt) from the absorption edges, we have replotted the absorbance spectra in the form of (Abs. × hν)2 as a function of photon energy, hν, in Fig. 2(b); where, the EOpt can be obtained by linear fitting to the increasing edge of the absorption features for the individual α-MoO3 belts; they are plotted as a function of their respective thicknesses in Fig. 2(c). It is seen that the EOpt increases by ~0.29 eV when the layer thickness decreases by one order of magnitude in the range of 10-104 nm. In this thickness range, the quantum confinement effect, in terms of 2D-TMDCs, -BP/BA and oxide quantum structures (e.g., MgO/ZnO/MgO quantum wells),10, 17-19, 26 is negligible and plays a minor

6

ACS Paragon Plus Environment

Page 6 of 30

Page 7 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

role in the thickness-reduction-induced EOpt blueshifts of the α-MoO3 belt crystals observed in Fig. 2(c).

Presented in Fig. 2(d) are the Raman spectra collected from the α-MoO3 samples AG at the locations where the absorbance spectra were measured. One sees that as the layer thickness decreases from samples A to G, the Raman features monotonically decrease in their intensities while their mode frequencies do not show any distinguishable shifts. These observations indicate that the lattice strains of the α-MoO3 belts are independent of their thickness in the studied range. Figure 2(e) presents a detailed spectral comparison, addressing the B2g mode, at about 282 cm-1, of the belt crystals A-G. It has been known that a B3g Raman active mode, at about 290 cm-1, is usually enhanced in its intensity in oxygen deficient α-MoO3 crystals. The intensity ratio of I282/I290, which has a linear relationship with respect to the O-to-Mo ratio, has been employed to evaluate the oxygen deficiencies of α-MoO3.27-29 In this regard, the comparisons in Fig. 2(e) indicate an absence of detectable variations in the oxygen deficiencies if they are present in the studied α-MoO3 belt crystals.

Energy-dispersive X-ray spectroscopy (EDX) installed in the SEM chamber was further employed to evaluate the stoichiometric variations among the α-MoO3 crystals. To minimize the influence of the oxide substrate, we have reduced the acceleration voltage of the electron beam to 5 keV from that of 15 keV for normal operations, so that the penetration depth of the electron beam in α-MoO3 is significantly reduced.5 The EDX spectra collected from the α-MoO3 belt crystals A-E at the locations where the absorbance and Raman spectra were collected are presented in Fig. 2(f) with the spectra

7

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

normalized to the peak intensity of oxygen; their survey spectra, shown in the inset, indicate that the effect of substrate is indeed excluded for the belts thicker than ~100 nm. These comparisons reveal that the α-MoO3 belt crystals are oxygen deficient (i.e., MoO2.93) but their composition stoichiometry is constant and independent of layer thicknesses, supporting the Raman spectroscopy results in Fig. 2(e).

A combination of the Raman spectroscopy and the EDX measurements provides evidence that both the lattice strain and the composition stoichiometry of the studied αMoO3 belt crystals are constant and independent of their layer thicknesses. As a result, they play a minor role in the layer-thickness-dependent EOpt shifts in Fig. 2(c). Such thickness-dependent EOpt shifts have also been observed in other semiconductors, e.g., γIn2Se3, but were attributed to the quantum confinement effect due to the poly-crystallinity and limited grain sizes of the layered materials (Supporting Information, Figure S1).30-31 To understand the EOpt shifts of our single crystal α-MoO3 belts, we have carried out transmission/absorption calculations for layered α-MoO3 on sapphire using the finitedifference-time-domain (FDTD) method with the optical refractive index available from the literature (Supporting Information, Figure S2).25 The FDTD calculations reveal a clear semi-linear relationship of EOpt as a function of the layer thickness in the range of 103-104 nm. However, the increase of EOpt tends to saturate as the layer thickness decreases below 103 nm. Taking the scattered refractive index values of α-MoO3 in the literatures into account, we can attribute the observed thickness-reduction-induced EOpt blueshifts to light interferences that occurred at the surfaces of the belt crystals rather than changes in the electronic structures of α-MoO3.

8

ACS Paragon Plus Environment

Page 8 of 30

Page 9 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Next, we have attempted to thin down the disassembled α-MoO3 belt crystals via plasma treatments and gas-phase etching at room temperature. Detailed procedures are supplied in the Supporting Information. In general, the surface of the [010]-oriented αMoO3 single crystals could have uncompleted layer steps/edges due to their van der Waals lattice structure. Figures 3(p) and 3(q) show a schematic comparison of the typical belt crystals with and without uncompleted surface steps/edges and, for the sake of brevity, they are referred to as surface-defective (SD) and defect-free (DF) crystals. Morphological evolutions recorded by optical microscopy from the SD and DF crystals upon the plasma treatments are presented in Figs. 3(a)-3(f) and 3(g)-3(l), respectively. For the plasma treatments, oxygen- and argon-plasma (referred to as OP and AP) were used in this study and the process consists of OP for 5 min (step 1), OP for 30 min (step 2), AP for 30 min (step 3), OP for 30 min (step 4), and AP for 30 min (step 5). The inplane orientations of the belt crystals, revealed by SCXRD, are indicated in Fig. 3(m). The comparisons in Figs. 3(a)-3(f) and 3(g)-3(l) reveal that the DF crystals are more resistive than the SD ones to the plasma treatments with the same process parameters. Optical absorbance spectra taken from the SD and DF crystals upon the sequential plasma treatments are presented in Figs. 3(n) and 3(o), respectively. They, together with the Raman spectra (Supporting Information, Figure S3), reveal two dominant effects of the plasma treatments. One is the bombardment of the energetic species and the other is the release of oxygen from the lattice sites, both occurred at the uncompleted surface steps/edges where the terminal oxygen atoms are readily knocked out.32 The former leads to a broadening and softening of the Raman features while the latter results in apparent redshifts in the absorption edges.27 The Raman spectra (Supporting Information, Figure

9

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

S3) also show that the AP-treatment-induced lattice disorders, which manifested themselves as the appearance of the feature at 447 cm-1, during step 3 can be more or less recovered by the OP-treatment during step 4 and regenerated by the AP-treatment during step 5. This behavior suggests that these lattice disorders are most likely associated with oxygen vacancies (VO).33-34 So-generated VO tends to diffuse into the asymmetric bridging oxygen sites (i.e., in the asymmetric O-Mo-O bonds along the [100]-axis of αMoO3), leading to the [001]-oriented dark linear micro-structures initiated at the surface steps/edges.32, 35-36 A typical microphotograph and a schematic diagram of such plasmatreatment-induced surface micro-structures are shown in the insets in Fig. 3(n).

The plasma-treated α-MoO3 belt crystals were further treated by gas-phase XeF2and HF-etching at room temperature, detailed procedures are supplied in the Supporting Information. The morphological evolutions recorded by optical microscopy in Figs. 4(a)4(e), and the evolutions of the interference fringes in Figs. 4(f)-4(g), as well as those of the Raman spectra in Figs. 4(h)-4(i), provide evidence that the [010]-oriented α-MoO3 crystal is more reactive to HF than to XeF2 at room temperature, especially for the DF crystals. Similar to the plasma treatments, the XeF2-etching tends to redshift the absorption edge and broaden/soften the Raman features of the SD crystal [see Figs. 4(f) and 4(h)] but has a minor effect on the absorption and Raman spectra of the DF crystal [see Figs. 4(g) and 4(i)]. In comparison, the absorption edge is somewhat shifted back to larger energies (i.e., blueshift) by the HF-etching for both the SD and the DF crystals. This can be attributed to an effective thickness reduction induced by the HF-etching, which is supported by the disappearance of the interference fringes from the absorption spectrum in Fig. 4(g); the non-shifted Raman features in Fig. 4(i); as well as the changed

10

ACS Paragon Plus Environment

Page 10 of 30

Page 11 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

color of the belt in Fig. 4(e) from those in Figs. 4(a)-4(d). These comparisons also indicate that the gas-phase HF-etching can effectively thin down the high-quality [010]oriented α-MoO3 (i.e., DF crystals) while keeping their crystal qualities undeteriorated. Presented in Figs. 4(j) and 4(k) are typical AFM images, together with their surface height profiles, recorded from the high-quality [010]-oriented α-MoO3 crystal before and after the HF-etching, i.e., corresponding to the samples imaged in Figs. 4(a) and 4(e), respectively. A comparison of the height profiles indicates that atomic steps of ≤ 2 MLs have been formed on the HF-etched surface. This result, together with the thickness reduction of > 200 nm [derived from the disappearance of the interference fringes in the absorbance spectrum in Fig. 4(g)], provides evidence that the gas-phase HF-etching of αMoO3 was in a layer-by-layer mode. This gas-phase layer-by-layer etching method may serve as a compensation to the wet chemical route and have important consequence when producing 2D α-MoO3 nanosheets from their high-quality [010]-oriented bulk crystals.16 In fact, our polarized Raman spectroscopy revealed that the rotation symmetry of the αMoO3 nanosheets produced by the gas-phase layer-by-layer etching method (Supporting Information, Figure S4) remains intact when compared with those of the thick α-MoO3 belt crystal.

Figure 5 presents the results collected from a high-quality [010]-oriented α-MoO3 crystal after an extended HF-etching (see the Supporting Information for the process). Shown in Fig. 5(a) is a microphotograph with its contrasts corresponding to the layer thickness variations. Figure 5(b) shows an AFM image taken from the edge area indicated by the box in Fig. 5(a), which exhibits regular linear micro-cracks along the

11

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

direction close to the [001]-axis as indicated by the arrows. The presence linear microcracks in the high-quality α-MoO3 layer upon the gas-phase etching, similar to those observed for the SD crystals upon plasma treatments discussed above (see Fig. 3), suggests occurrences of anisotropic in-plane strain relaxations and/or redistributions in the thinned crystals approaching their 2D nanosheets. The height profile in Fig. 5(c), derived from the locations indicated by the straight line in Fig. 5(b), shows that the HFetched α-MoO3 layer is about 6.4 nm (~8 MLs) thick and has a thinner area of 3.2 nm (~4 MLs) thick at the edge. Raman mappings recorded from the edge are shown in Figs. 5(d)5(f), addressing the distributions of the intensity, mode frequency, and linewidth of the Ramam feature at ~818 cm-1, respectively. One sees that Raman scattering from the thinner area at the edge of the α-MoO3 layer exhibits lowered intensity [Fig. 5(d)], softened frequency [Fig. 5(e)], and broadened linewidth [Fig. 5(f)]. In Fig. 5(e), we have further plotted the mode frequencies as a function of their respective intensities, derived from the boxed areas indicated in Figs. 5(d)-5(e), together with a nonlinear fitting (i.e., the solid line), which clearly shows redshifts at lower scattering intensities. The redshift is more clearly seen in the spectral comparisons at ~818 cm-1 of the 8 MLs thick layer and its 4 MLs thick edge presented in the inset of Fig. 5(g). These results provide evidence for the mode softening of the 818 cm-1 phonon when the thickness of α-MoO3 is reduced approaching 2D nanosheets. Likewise, the absorption spectra collected from the 8 MLs thick layer, the 4 MLs thick edge, and a bulk α-MoO3 are shown in Fig. 5(h). One sees that the Eopt of the 4 MLs thick α-MoO3 is significantly blue shifted to ~4.8 eV. This blueshift is much larger than that derived from the thickness-dependent semi-linear relationship in Fig. 2(c). However, both the Raman mode softening and the huge

12

ACS Paragon Plus Environment

Page 12 of 30

Page 13 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

absorption edge blueshift of the 2D α-MoO3 (~4 MLs) obtained by the gas-phase HFetching are consistent with those of α-MoO3 nanosheets exfoliated by the BSA-assisted aqueous method.4

Since the SCXRD measurements indicate that our α-MoO3 single crystals are compressively stressed and the lattice strains are anisotropic. In this regard, anisotropic strain relaxations and/or redistributions might occur as the α-MoO3 belt crystal is thinned down approaching 2D nanosheets. In fact, the Cai et al. have already observed a slight redshift of the 818 cm-1 mode in few-layer MoO3 as compared with that of the bulk crystal and they attributed the redshift to an in-plane lattice expansion of the 2D nanosheet.3 It is thus reasonable that the observed Raman mode softening of our 2D αMoO3 nanosheets processed from belt crystals is due to an overall effect of anisotropic relaxation/redistribution of the coherent compressive strains. Such anisotropic strain relaxations/redistributions could tune the electronic structure of the 2D α-MoO3 nanosheets, leading to the absorption edge blueshifts. In this light, we have further studied the electronic structures of 2D α-MoO3 nanosheets concerning the strains relaxations and/or redistributions.

Figures 6(a)-6(e) present the electronic structures of a 4 MLs thick α-MoO3 with and without uniaxial in-plane strains (i.e., ε = ± 6%). They were calculated based on the periodic density-functional-theory (DFT) and the details can be found in the Supporting Information. One sees that the electronic structures exhibit the typical indirect bandgap character with the valence band maximum at the R point while the conduction band minimum at the Γ point except for the case of ε = 6% along the [100]-axis [see Fig. 6(d)]. 13

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Plotted in Fig. 6(f) are the direct bandgap and the indirect bandgap energies as a function of in-plane strain variations up to ε = ± 6%. It is seen that the bandgap changes induced by the strain variations along the [100]-axis are larger than those induced by the same amount of strain variations along the [001]-axis; a strain variation from ε = -6% to 6% along the [100]-direction can result in a blueshift of ~1.5 eV in the direct bandgap.

Taking the regular linear micro-cracks observed in Fig. 5(b) into account, we believe that in-plane strain relaxations/redistributions occurred during the layer’s thinning down approaching 2D nanosheets. The [001]-oriented linear micro-cracks are most likely created due to the relaxation of tensile strains along the [100]-axis. In this regard, the creak-free area of the thinner 2D nanosheets could be tensile stressed along their [100]axis and that, in turn, resulted in the larger bandgap energies. It has to be noted that the inherent VO and their diffusions upon the gas-phase etching of α-MoO3 towards 2D nanosheets may also affect their in-plane strain relaxations and/or redistributions; unfortunately, the detailed mechanism is unclear at this stage and, therefore, not taken into account in the current calculations. Nevertheless, the combination between the anisotropic strain relaxations/redistributions and the calculated electronic structures provide clear clues to the bandgap widening of the 2D α-MoO3 nanosheets and thus the observed absorption edge blueshifts.

IV. Conclusion In conclusion, we have synthesized high-quality layered α-MoO3 single crystals by TVT of MoO3 powders. SCXRD revealed that the layered crystals are compressively stressed

14

ACS Paragon Plus Environment

Page 14 of 30

Page 15 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

and the lattice strains are anisotropic. EDX indicated that the layered crystals are oxygen deficient; however, the oxygen deficiency (i.e., MoO2.93) is constant and independent of the layer thickness in the range of 10-104 nm. In this thickness range, we observed apparent absorption edge blueshifts of the layered crystals as their thickness is reduced; however, their Raman features do not exhibit any frequency shifts at all. The former, based on our theoretical FDTD calculations, is due to light interferences at the crystal surfaces while the latter, together with the constant intensity ratios of I282/I290, supports the constant oxygen deficiency in the layered single crystals. The α-MoO3 layered crystal without surface defects is more resistive than the surface defective ones to plasma treatments at room temperature. The plasma-treatment-induced linear structures along the [001]-axis at the edge of the layer steps, accompanied by redshifts in the absorption edge and softening/broadening of the Raman features, of the surface defective α-MoO3 belts rather than the defect-free ones suggest that the formation of the linear structures is related to the knocking-out of oxygen atoms as well as the diffusions of VO. This conclusion is further supported by the creation/recover/recreation of the Raman feature at ~447 cm-1 of the α-MoO3 single crystal upon the sequential argon/oxygen/argon plasma treatments. A novel gas-phase layer-by-layer etching of the layered α-MoO3 single crystals has been demonstrated and the HF-etching is more effective than that of XeF2etching at room temperature. By extending the HF-etching, an α-MoO3 nanosheet of about 4 MLs thick has been obtained at the edge of a ~8 MLs thick α-MoO3 layer; this 2D nanosheet exhibits apparent phonon softening of the Ag mode at ~818 cm-1 and a huge EOpt blueshift. The phonon softening provides evidence for the in-plane strain relaxations/redistributions while the huge EOpt blueshift, in terms of our periodic DFT

15

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

calculations, is attributable to the anisotropic in-plane strain relaxations/redistributions. Although the effect of VO (either inherent or created during the film thinning), as well as their diffusions, along with the strain relaxations/redistributions was not taken into account, they could also play roles in varying the electronic structures of the 2D α-MoO3.

Supporting Information Available Optical bandgap of γ-In2Se3 as a function layer thickness; theoretical finite-differencetime-domain (FDTD) calculations of the transmissions, and thus the absorptions and optical bandgaps, of layered α-MoO3 single crystals as a function of their thicknesses in the range of 10-104 nm; Raman spectra collected from a surface-defective (SD) and a defect-free (DF) α-MoO3 single crystals upon sequential oxygen- and argon-plasma treatments; disassembling of α-MoO3 belt crystals on sapphire substrate; polarized Raman spectra of α-MoO3 thick layer and thin nanosheet; plasma treatment procedures; gas-phase etching procedures; and details of the periodic density-functional-theory (DFT) calculations.

Acknowledgement The authors would like to acknowledge S. Q. Bai for his help in SCXRD measurements. This research is supported by A*STAR Science and Engineering Research Council Pharos 2D Program (SERC Grant No. 152-70-00012).

16

ACS Paragon Plus Environment

Page 16 of 30

Page 17 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Reference: 1.

Berthumeyrie, S.; Badot, J. C.; Pereira-Ramos, J. P.; Dubrunfaut, O.; Bach, S.; Vermaut, P., Influence of lithium lnsertion on the electronic transport in electroactive MoO3 nanobelts and classical powders: Morphological and particle size effects. J. Phys. Chem. C 2010, 114, 19803-19814.

2.

Guo, H.; Lan, C.; Zhou, Z.; Sun, P.; Wei, D.; Li, C., Transparent, flexible, and stretchable WS2 based humidity sensors for electronic skin. Nanoscale 2017, 9, 6246-6253.

3.

Cai, L.; McClellan, C. J.; Koh, A. L.; Li, H.; Yalon, E.; Pop, E.; Zheng, X., Rapid flame synthesis of atomically thin MoO3 down to monolayer thickness for effective hole doping of WSe2. Nano Lett. 2017, 17, 3854-3861.

4.

Liu, H.; Cai, Y.; Han, M.; Guo, S.; Lin, M.; Zhao, M.; Zhang, Y.; Chi, D., Aqueous and mechanical exfoliation, unique properties, and theoretical understanding of MoO3 nanosheets made from free-standing α-MoO3 crystals: Raman mode softening and absorption edge blue shift. Nano Res. 2018, 11, 1193-1203.

5.

Liu, H.; Yang, R. B.; Yang, W.; Jin, Y.; Lee, C. J. J., Atomic layer deposition and post-growth thermal annealing of ultrathin MoO3 layers on silicon substrates: Formation of surface nanostructures. Appl. Surf. Sci. 2018, 439, 583-588.

6.

Sreedhara, M. B.; Santhosha, A. L.; Bhattacharyya, A. J.; Rao, C. N. R., Composite of few-layer MoO3 nanosheets with graphene as a high performance anode for sodium-ion batteries. J. Mater. Chem. A 2016, 4, 9466-9471.

7.

Balendhran, S.; Deng, J.; Ou, J. Z.; Walia, S.; Scott, J.; Tang, J.; Wang, K. L.; Field, M. R.; Russo, S.; Zhuiykov, S.; et al., Enhanced charge carrier mobility in twodimensional high dielectric molybdenum oxide. Adv. Mater. 2013, 25, 109-114.

8.

Balendhran, S. Devices and systems based on two dimensional MoO3 and MoS2. RMIT University, 2013.

17

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

9.

Balendhran, S.; Walia, S.; Nili, H.; Ou, J. Z.; Zhuiykov, S.; Kaner, R. B.; Sriram, S.; Bhaskaran, M.; Kalantar-zadeh, K., Two-dimensional molybdenum trioxide and dichalcogenides. Adv. Fun. Mater. 2013, 23, 3952-3970.

10. Liu, H. F.; Wong, S. L.; Chi, D. Z., CVD growth of MoS2-based two-dimensional materials. Chem. Vap. Dep. 2015, 21, 241-259. 11. Novoselov, K. S.; Jiang, D.; Schedin, F.; Booth, T. J.; Khotkevich, V. V.; Morozov, S. V.; Geim, A. K., Two-dimensional atomic crystals. Proc. Nat. Acad. Sci. USA 2005, 102, 10451-10453. 12. Manzeli, S.; Ovchinnikov, D.; Pasquier, D.; Yazyev, O. V.; Kis, A., 2D transition metal dichalcogenides. Nat. Rev. Mater. 2017, 2, 17033. 13. Liu, H.; Ansah-Antwi, K. K.; Ying, J.; Chua, S.; Chi, D., Towards large area and continuous MoS2 atomic layers via vapor-phase growth: Thermal vapor sulfurization. Nanotechnology 2014, 25, 405702. 14. Liu, H.; Ansah-Antwi, K. K. A.; Chua, S.; Chi, D., Vapor-phase growth and characterization of Mo1-xWxS2 (0 ≤ x ≤1) atomic layers on 2-inch sapphire substrates. Nanoscale 2014, 6, 624-629. 15. Liu, H.; Chi, D., Dispersive growth and laser-induced rippling of large-area singlelayer MoS2 nanosheets by CVD on c-plane sapphire substrate. Sci. Rep. 2015, 5, 11756. 16. Fahmida, R.; Taimur, A.; Sumeet, W.; Edwin, M.; Sharath, S.; Madhu, B.; Sivacarendran, B., Two-dimensional MoO3 via a top-down chemical thinning route. 2D Mater. 2017, 4, 035008. 17. Mak, K. F.; Lee, C.; Hone, J.; Shan, J.; Heinz, T. F., Atomically thin MoS2: A new direct-gap semiconductor. Phys. Rev. Lett. 2010, 105, 136805. 18. Cai, Y.; Zhang, G.; Zhang, Y.-W., Layer-dependent band alignment and work function of few-layer phosphorene. Sci. Rep. 2014, 4, 6677.

18

ACS Paragon Plus Environment

Page 18 of 30

Page 19 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

19. Osters, O.; Nilges, T.; Bachhuber, F.; Pielnhofer, F.; Weihrich, R.; Schöneich, M.; Schmidt, P., Synthesis and identification of metastable compounds: Black arsenic— science or fiction? Angew. Chem. Internation. Ed. 2012, 51, 2994-2997. 20. Liu, B.; Köpf, M.; Abbas, A. N.; Wang, X.; Guo, Q.; Jia, Y.; Xia, F.; Weihrich, R.; Bachhuber, F.; Pielnhofer, F.; et al., Black arsenic–phosphorus: Layered anisotropic infrared semiconductors with highly tunable compositions and properties. Adv. Mater. 2015, 27, 4423-4429. 21. Molina-Mendoza, A. J.; Lado, J. L.; Island, J. O.; Niño, M. A.; Aballe, L.; Foerster, M.; Bruno, F. Y.; López-Moreno, A.; Vaquero-Garzon, L.; van der Zant, H. S. J.; et al., Centimeter-scale synthesis of ultrathin layered MoO3 by van der Waals epitaxy. Chem. Mater. 2016, 28, 4042-4051. 22. Wang, D.; Li, J.-N.; Zhou, Y.; Xu, D.-H.; Xiong, X.; Peng, R.-W.; Wang, M., Van der Waals epitaxy of ultrathin α-MoO3 sheets on mica substrate with single-unit-cell thickness. Appl. Phys. Lett. 2016, 108, 053107. 23. Sreedhara, M. B.; Matte, H. S. S. R.; Govindaraj, A.; Rao, C. N. R., Synthesis, characterization, and properties of few-layer MoO3. Chem. Asian J. 2013, 8, 24302435. 24. Ji, F.; Ren, X.; Zheng, X.; Liu, Y.; Pang, L.; Jiang, J.; Liu, S., 2D-MoO3 nanosheets for superior gas sensors. Nanoscale 2016, 8, 8696-8703. 25. Lajaunie, L.; Boucher, F.; Dessapt, R.; Moreau, P., Strong anisotropic influence of local-field effects on the dielectric response of α-MoO3. Phys. Rev. B 2013, 88, 115141. 26. Sun, C. W.; Xin, P.; Liu, Z. W.; Zhang, Q. Y., Room-temperature photoluminescence of ZnO/MgO multiple quantum wells on Si (001) substrates. Appl. Phys. Lett. 2006, 88, 221914.

19

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

27. Taka-aki, Y.; Keisuke, Y.; Yuhei, H.; Tomohiro, H.; Fumio, O.; Masahiko, H., Probing edge-activated resonant Raman scattering from mechanically exfoliated 2D MoO3 nanolayers. 2D Mater. 2015, 2, 035004. 28. Dieterle, M.; Weinberg, G.; Mestl, G., Raman spectroscopy of molybdenum oxides Part I. Structural characterization of oxygen defects in MoO3-x by DR UV/VIS, Raman spectroscopy and X-ray diffraction. Phys. Chem. Chem. Phys. 2002, 4, 812821. 29. Yan, B.; Zheng, Z.; Zhang, J.; Gong, H.; Shen, Z.; Huang, W.; Yu, T., Orientation controllable growth of MoO3 nanoflakes: Micro-Raman, field emission, and birefringence properties. J. Phys. Chem. C 2009, 113, 20259-20263. 30. Ho, C.-H., Amorphous effect on the advancing of wide-range absorption and structural-phase transition in γ-In2Se3 polycrystalline layers. Sci. Rep. 2014, 4, 4764. 31. Ho, C.-H.; Chen, Y.-C., Thickness-tunable band gap modulation in γ-In2Se3. RSC Adv. 2013, 3, 24896-24899. 32. Smith, R. L. The structural evolution of the MoO3 (010) surface during reduction and oxidation reactions. Carnegie Mellon University, 1998. 33. Eda, K., Infrared spectra of hydrogen molybdenum bronze, H0.34MoO3. J. Solid State Chem. 1989, 83, 292-303. 34. Chen, D.; Liu, M.; Yin, L.; Li, T.; Yang, Z.; Li, X.; Fan, B.; Wang, H.; Zhang, R.; Li, Z.; et al., Single-crystalline MoO3 nanoplates: topochemical synthesis and enhanced ethanol-sensing performance. J. Mater. Chem. 2011, 21, 9332-9342. 35. Hsu, Z. Y.; Zeng, H. C., Generation of double-layer steps on (010) surface of orthorhombic MoO3 via chemical etching at room temperature. J. Phys. Chem. B 2000, 104, 11891-11898. 36. Zeng, H. C., Chemical etching of molybdenum trioxide:  A new tailor-made synthesis of MoO3 catalysts. Inorg. Chem. 1998, 37, 1967-1973.

20

ACS Paragon Plus Environment

Page 20 of 30

Page 21 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure Captions Figure 1: Synthesis and morphological and structural properties of α-MoO3 single crystals. (a) Schematic diagram of the thermal vapor transport (TVT) setup; (b) typical SEM image recorded from the as-grown belt crystals lying on their belt surfaces; (c) and (d) microphotographs taken from the belt crystals disassembled on sapphire substrates with their color variations corresponding to their thickness changes; (e) XRD patterns collected by SCXRD and GADDS systems from a single belt and gathered belts, respectively, their comparisons indicate that the belts are typically lying on their (010) atomic planes; and (f) the lattice structure calculated from the SCXRD measurements, which is indeed α-MoO3 that consists of ‘monolayers’ stacking along their [010]-axis via van der Waals forces, the SCXRD measurements also revealed that the length direction of the belts is along their [001]-axis. Figure 2: Absorption, Raman scattering, and EDX spectra collected from the α-MoO3 belt crystals as a function of their thicknesses. (a) Absorbance spectra; (b) absorption spectra in the form of (abs. × hν)2; (c) optical bandgap energy EOpt derived from the absorption edge of (abs. × hν)2; (d) Raman spectra; (e) normalized Raman spectra addressing the constant intensity ratio of I282/I290; and (f) EDX spectra collected with reduced electron-beam accelerated voltage, the inset are the survey spectra showing the absence of aluminum from the sapphire substrate.

21

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 3: Morphological (a-l) and absorption spectral (n and o) evolutions of a surfacedefective (SD, a-f) and a defect-free (DF, g-l) α-MoO3 belt crystals upon sequential oxygen- and argon-plasma (referred to as OP and AP) treatments. (m) shows the crystal orientations of the crystals in (a-l). The insets in (n) show a microphotograph and its schematic diagram of the plasma-treatment-induced [001]-oriented linear structures on the surface of the SD crystals. (p) and (q) are schematic diagrams showing the SD and DF α-MoO3 belt crystals. Figure 4: Morphological and absorption and Raman spectral evolutions of a surfacedefective (SD) and a defect-free (DF) α-MoO3 belt crystals upon sequential gas-phase XeF2- and HF-etching. (a)-(e) microphotographs; (f)-(g) absorbance spectra; (h)-(i) Raman spectra; (j)-(k) AFM images and their surface height profiles. (f) and (h) were collected from the SD crystal while (g) and (i) were collected from the DF crystal; (j) and (k) were recorded from the DF crystal before and after the extended HF-etching, demonstrating a layer-by-layer etching mode. Figure 5: Morphology, Raman mapping, and absorption spectra of gas-phase etched αMoO3 belt crystal. (a) Microphotograph; (b) AFM image recorded at the boxed area in (a), the arrows indicate the regular linear defect structures, along the direction close to the [001]-axis, created during the thinning down of the layered crystal approaching 2D nanosheets; (c) surface height profile derived from the locations indicated by the straight line in the (b) that shows a thinner area at the edge of the α-MoO3 layer on the sapphire substrate; (d)-(f) Raman

22

ACS Paragon Plus Environment

Page 22 of 30

Page 23 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

mappings of the Mo-O-Mo (Ag) mode at 818 cm-1 addressing its intensity, mode frequency, and linewidth distributions, respectively; (g) mode frequencies as a function of their respective scattering intensities derived from the boxed areas in (d) and (e), the inset shows a spectral comparison derived from the thicker inner area and the thinner edge area of the α-MoO3 layer; and (h) absorption spectra collected from the thicker inner area and the thinner edge area the α-MoO3 layer together with that of a bulk crystal. Figure 6: Electronic structures and bandgap energies of a 4 MLs thick α-MoO3 calculated using the periodic density-functional-theory (DFT). For these calculations, inplane uniaxial strain of up to ε = ± 6% was introduced along either the [100]- or the [001]-axis to evaluate the effect of strain relaxations/redistributions on the electronic structures and thus the bandgap energies of the 4 MLs thick α-MoO3 nanosheet, which provides clear clues to the thickness-reduction-induced huge absorption edge blueshifts. (a) Fully relaxed; (b) 6% compressive strain along the [100]-axis; (c) 6% compressive strain along the [001]-axis, (d) 6% tensile strain along the [100]-axis; and (e) 6% tensile strain along the [001]-axis. (f) Bandgap energies of the α-MoO3 nanosheet as a function of in-plane strains.

23

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

TOC Graphic

24

ACS Paragon Plus Environment

Page 24 of 30

Page 25 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 1: Synthesis and morphological and structural properties of α-MoO3 single crystals. (a) Schematic diagram of the thermal vapor transport (TVT) setup; (b) typical SEM image recorded from the as-grown belt crystals lying on their belt surfaces; (c) and (d) microphotographs taken from the belt crystals disassembled on sapphire substrates with their color variations corresponding to their thickness changes; (e) XRD patterns collected by SCXRD and GADDS systems from a single belt and gathered belts, respectively, their comparisons indicate that the belts are typically lying on their (010) atomic planes; and (f) the lattice structure calculated from the SCXRD measurements, which is indeed α-MoO3 that consists of ‘monolayers’ stacking along their [010]-axis via van der Waals forces, the SCXRD measurements also revealed that the length direction of the belts is along their [001]-axis. 80x77mm (300 x 300 DPI)

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 2: Absorption, Raman scattering, and EDX spectra collected from the α-MoO3 belt crystals as a function of their thicknesses. (a) Absorbance spectra; (b) absorption spectra in the form of (abs. × hν)2; (c) optical bandgap energy EOpt derived from the absorption edge of (abs. × hν)2; (d) Raman spectra; (e) normalized Raman spectra addressing the constant intensity ratio of I282/I290; and (f) EDX spectra collected with reduced electron-beam accelerated voltage, the inset are the survey spectra showing the absence of aluminum from the sapphire substrate. 110x86mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 26 of 30

Page 27 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 3: Morphological (a-l) and absorption spectral (n and o) evolutions of a surface-defective (SD, a-f) and a defect-free (DF, g-l) α-MoO3 belt crystals upon sequential oxygen- and argon-plasma (referred to as OP and AP) treatments. The insets in (n) show a microphotograph and its schematic diagram of the plasmatreatment-induced [001]-oriented linear structures on the surface of the SD crystals. (p) and (q) are schematic diagrams showing the SD and DF α-MoO3 belt crystals. 115x95mm (300 x 300 DPI)

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 4: Morphological and absorption and Raman spectral evolutions of a surface-defective (SD) and a defect-free (DF) α-MoO3 belt crystals upon sequential gas-phase XeF2- and HF-etching. (a)-(e) microphotographs; (f)-(g) absorbance spectra; (h)-(i) Raman spectra; (j)-(k) AFM images and their surface height profiles. (f) and (h) were collected from the SD crystal while (g) and (i) were collected from the DF crystal; (j) and (k) were recorded from the DF crystal before and after the extended HF-etching, demonstrating a layer-by-layer etching mode. 111x88mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 28 of 30

Page 29 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 5: Morphology, Raman mapping, and absorption spectra of gas-phase etched α-MoO3 belt crystal. (a) Microphotograph; (b) AFM image recorded at the boxed area in (a), the arrows indicate the regular linear defect structures, along the direction close to the [001]-axis, created during the thinning down of the layered crystal approaching 2D nanosheets; (c) surface height profile derived from the locations indicated by the straight line in the (b) that shows a thinner area at the edge of the α-MoO3 layer on the sapphire substrate; (d)-(f) Raman mappings of the Mo-O-Mo (Ag) mode at 818 cm-1 addressing its intensity, mode frequency, and linewidth distributions, respectively; (g) mode frequencies as a function of their respective scattering intensities derived from the boxed areas in (d) and (e), the inset shows a spectral comparison derived from the thicker inner area and the thinner edge area of the α-MoO3 layer; and (h) absorption spectra collected from the thicker inner area and the thinner edge area the α-MoO3 layer together with that of a bulk crystal. 139x139mm (300 x 300 DPI)

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 6: Electronic structures and bandgap energies of a 4 MLs thick α-MoO3 calculated using the periodic density-functional-theory (DFT). For these calculations, in-plane uniaxial strain of up to ε = ± 6% was introduced along either the [100]- or the [001]-axis to evaluate the effect of strain relaxations/redistributions on the electronic structures and thus the bandgap energies of the 4 MLs thick αMoO3 nanosheet, which provides clear clues to the thickness-reduction-induced huge absorption edge blueshifts. 139x90mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 30 of 30