Functionalization of Antimonene and Bismuthene With Lewis Acids

You might also like

Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 16

NR|d2nr03206f|10.

1039/d2nr03206f|ART

Functionalization of antimonene and bismuthene with Lewis acids†

Manaswee Barua,‡ Mohd Monis Ayyub,‡ Shashidhara Acharya and C. N. R. Rao*

New Chemistry Unit, International Centre for Material Science and School of Advanced Materials,

Jawaharlal Nehru Center for Advanced Scientific Research, Bangalore-560064, India. E-mail:

cnrrao@jncasr.ac.in

Received 10th June 2022, Accepted 19th August 2022

Elemental 2D pnictogens (group 15) are an interesting class of materials with tunable

band structures and high carrier mobilities.

GA1

Elemental 2D pnictogens (group 15) are an interesting class of materials with tunable band structures and

high carrier mobilities. Heavier pnictogens (Sb and Bi) are stable under ambient conditions compared to the

lighter members (P and, As) and are emerging as interesting candidates for various electronic and

optoelectronic applications. The reactivity of these materials is due to the presence of a lone pair which can

be effectively utilized to tune material properties via different functionalization strategies. In this work, we

have synthesized antimonene and bismuthene nanosheets by liquid exfoliation which are emissive in the

visible range and functionalized these nanosheets with group 12 and 13 Lewis acids (ZnCl2, CdCl2, BCl3,

GaCl3, AlCl3, and InCl3). Interaction of these Lewis acids with the lone pairs on Sb/Bi leads to the

formation of Lewis acid-–base adducts with the corresponding changes in the bonding environment along

with lattice distortion and rehybridization of the band structure. Interestingly, the changes in band structure

upon functionalization were realized as a blue shift in the emission of few-layered Sb and Bi. This is the

first report on the functionalization of heavier pnictogens by the formation of Lewis acid-–base adducts and

opens a path for tuning their properties for integration in electronic and optoelectronic devices.

Electronic supplementary information (ESI) available: DETAILS. See DOI:
https://doi.org/10.1039/d2nr03206f
‡‡
These authors contributed equally to this work.
1
1. Introduction

Group 15 elemental 2D pnictogens are emerging as an interesting class of materials with tunable band

structures, carrier mobilities in the range of thousands of cm2 V−1 s−1, and are semiconducting in their

monolayer/few-layered forms, making them probable candidates for various electronic and optoelectronic

applications.1,2 However, the most studied pnictogen, phosphorene is highly oxophilic and undergoes

dissociative chemisorption which limits its applications.3 In comparison, heavier pnictogens (Sb and Bi) are

stable under ambient conditions. Bulk antimony and bismuth crystallize in a rhombohedral layered

structure with the R m space group (β-phase) consisting of buckled six-membered rings.4,5 This unique

buckled structural arrangement results in anisotropic properties in individual layers, with each sp3

hybridized Sb/Bi atom having three sigma bonds and one lone pair. The presence of a lone pair of electrons

creates a reactive surface that can be utilized for functionalization.5

The growing interest in Sb and Bi is due to their interesting band structure. Bulk Bi is semi-metallic and

undergoes a transition to an indirect semiconducting state in monolayers with a bandgap of up to 1 eV.5

Although the absorption onset of bismuthene is theoretically reported to be 0.74 eV, the major absorption

peak lies in the visible range at ~1.72 eV.6 Moreover, experimental reports indicate multiple emissions in

the visible range from bulk Bi, Bi thin films, and nanosheets which could be due to Bi3+ color centers,

defect/trapped states, and band edge emission, respectively.7–9 Similarly bulk Sb is semi-metallic in nature

and undergoes a transition to a semiconducting state under three layers.10 Experimentally, the monolayer

antimonene band gap in the visible range has been observed by UV-Vis spectroscopy and

photoluminescence (PL) emission.11,12 The presence of the band gap only in few-layer limits the

applications of antimonene for optoelectronics. However, the PL of multilayer antimonene nanoribbons on

InSe is reported to have maxima at 610 nm and antimonene quantum dots show excitation-dependent PL

emission between 350– and -500 nm.13,14

The Rreactivity of 2D materials is highly dependent on its surface/interface owing to the high surface-to-

volume ratios making functionalization through surfaces a very effective/sensitive strategy to tailor a

materials’ properties. Chemical functionalization is an effective strategy to passivate the surface and

simultaneously tailor the inherent properties and can be achieved by various strategies such as covalent and
2
non-covalent functionalization, defect engineering, and heterostructure formation. 15–18 Covalent

functionalization with reactive species like diazonium salt and Grignard reagents leads to the formation of

thick and insulating layers on the surface which limits its integration into electronic devices.19,20 In

comparison, other functionalization strategies such as Lewis acid-–base adduct formation leads to the

formation of coordinate bonds with minimal lattice distortion and controllable thicknesses. 19,21 Pnictogens

have lone pairs which are highly reactive and can be utilized for chemical functionalization without

significantly disturbing the inherent structure.19,22 Surface modification of black phosphorus with group 13

Lewis acids hasve been shown to tune the electronic properties with strong p-doping and suppression of n-

type conductivity while passivating the surface.22 Similarly, non-covalent functionalization of antimonene

with PDI and TCNQ induces a change in band gap owing to charge transfer.23

In this work, we have synthesized antimonene and bismuthene by liquid exfoliation which was found to be

emissive in the visible range. Functionalization of these nanosheets with group 12 and 13 Lewis acids

(ZnCl2, CdCl2, BCl3, GaCl3, AlCl3, and InCl3) leads to the formation of Lewis acid-–base adducts. This

reaction proceeds via the transfer of electrons from Sb/Bi lone pairs to the metal center in the Lewis acid

and the corresponding changes in the bonding environment and lattice distortion were tracked using EDS,

XPS, and Raman spectroscopy. This adduct formation leads to rehybridization of the band structure with

the addition of new bands corresponding to the Lewis acid as has been reported for phosphorene.24

Interestingly, these changes in band structure were tracked by emission studies wherein we observed a blue

shift in the emission of few-layered Sb and Bi on functionalization. This is the first report on Lewis acid-–

base adduct formation in heavier pnictogens (Sb and Bi) and opens a path for tuning their properties for

integration in electronic and optoelectronic devices.

2. Experimental section

2.1 Materials:

Antimony crystals (Smart elements, 99.99995%), Bbismuth crystals (Smart elements, 99.99995%), InCl3

(Merck, 98%), AlCl3 (Merck, anhydrous 99.999%), CdCl2 (Merck, 99.99%), ZnCl2 (SRL, 97%), GaCl3

(Merck, 99.99%), BCl3 (Spectrochem, 1 M in Ddichloromethane), NN-mMmethyl-2-pyrrolidone (NMP)

(Spectrochem, 99%), Ttetrahydrofuran (THF) (Spectrochem 99,.5%), Ttoluene (Spectrochem 99%),


3
methanol (Spectrochem 99%), and an Si substrate (University wafer, <100>, thickness 525 ± 25 μm, ρρ =

0.1–-100 Ω cm) were obtained. Prior to synthesis, all solvents were degassed by purging inert gas (Ar) for

30 minutes with continuous stirring. THF, toluene, and methanol were dried using 3 Å silica gel before use.

2.2 Synthesis of Aantimony/Bbismuth nanosheets:

Antimony/Bbismuth crystals were ground in a mortar pestle to obtain a fine powder. Sb/Bi powder (400

mg) was transferred to a round bottom centrifuge tube and filled with degassed NN-methyl-2-pyrrolidone

(17.5 mL). The tube was closed with a cap fitted with the sonicator probe and packed under an inert

atmosphere. Probe sonication was carried out for 4 hours (Sb) and 5 hours (Bi) with a pulse sequence of 4 s

on and off at 25% amplitude. The resulting dispersion of exfoliated Sb/Bi was centrifuged at 2500 rpm

(3935 g) to separate the unexfoliated Sb/Bi. Further centrifugation at higher speeds was done to separate

the two grades of Sb/Bi, Sb/Bi nanosheets (sediment at 14 500 rpm), and few-layered Sb/Bi (supernatant at

15 500 rpm).

2.3 Functionalization of Sb/Bi-Nnanosheets (Sb/Bi-NS):

Reaction solvents were selected based on the solubility of Lewis acids. Tetrahydrofuran (THF) was used

for BCl3, GaCl3, InCl3, and ZnCl2; toluene was used for AlCl3; methanol was used for CdCl2. Prior to the

reaction, Sb/Bi nanosheets (Sb-NS) were washed multiple times with the solvent appropriate for the Lewis

acid by bath sonication for 2 minutes followed by centrifugation at 14 500 rpm for 5 minutes. This process

was repeated at least 3 times. Sb/Bi-NS dispersion in dry solvent was kept in a Schlenk flask under an inert

atmosphere and an equimolar amount of Lewis acid was added to this dispersion. The solutions were kept

stirring under an inert atmosphere and were monitored. BCl3 and GaCl3 reacted vigorously and formed a

colorless solution instantaneously. AlCl3 reacted relatively mildly but formed a colorless solution after 1

hour. Reactions with InCl3 and CdCl2 were considerably slower and did n’t not corrode the material. After

optimization, the reaction with AlCl3 was done for 15 min, CdCl2 for 10 days, and InCl3 for 14 days. The

solid products obtained were washed multiple times with solvents to remove any physisorbed Lewis acid

residues and thoroughly characterized.

4
2.4 Functionalization of few-layered Sb/Bi (Sb/Bi-FL):

Sb/Bi-FL was obtained as a supernatant upon centrifugation at 15 500 rpm for 1 hour. The dispersion in

NMP was drop coated on the Si substrate and vacuum dried for further characterizations and reaction with

Lewis acids. The Sb/Bi-FL coated substrate was dipped in a solution of Lewis acid in the desired solvent

under an inert atmosphere. After the reaction, the substrate was washed multiple times with the solvent to

remove any unreacted Lewis acid. These substrates were thoroughly characterized and used for FESEM

cathodoluminescence studies.

2.5 Cleaning the Si substrate:

Si substrates were first dipped in acetone and heated to 50 °C for 10 minmins. After this the substrates were

rinsed with methanol and dried under a N2 gun. The dried substrates were dipped in Piranha solution for 10

minmins and washed rigorously with water afterward. In the final step, the substrates were dipped in 2%

HF solution for 10 minutes, washed rigorously with water, and then dried under using an N2 gun.

2.6 Physical characterization:

Liquid-phase exfoliation was carried out with a Sonics Vibracell probe sonicator (750 W) equipped with a

titanium alloy (Ti-–6Al-–4V) based tapered probe (tip diameter = 6 mm). The probe sonicator assembly

was maintained at 5 °C by circulating cold water to avoid overheating of the probe. Raman spectra of Sb

were recorded on a Horiba-Jobin Yvon LabRAM HR 800 Raman spectrometer equipped with Ar + laser

with a 514 nm notch filter with a spectral resolution of 1.5 cm−1. Raman spectra of Bi samples were

collected on a Renishaw inVia microscope fitted with a 784 nm laser. Stability studies for Bi-FL were

performed on a Horiba-Jobin Yvon LabRAM HR 800 Raman spectrometer equipped with a 632 nm HeNe

laser with a spectral resolution of 1.5 cm−1. Samples for Raman measurements were prepared by drop

coating the dispersion on a glass slide and dried under vacuum. Bright-field transmission electron

microscopy (TEM) images were recorded on a JEOL 300kV HRTEM instrument. Atomic force

microscopy (AFM) analyses were carried out in contact mode on a Bruker Innova AFM instrument. Sb/Bi

nanosheets separated at different centrifugation speeds were drop coated on the Si substrate and dried under

vacuum. Nanoscope analysis software was used to analyze the data. X-ray photoelectron spectroscopy

5
studies were carried out on a Thermo K-alpha+ spectrometer using micro-focused and mono chromated Al

Kα radiation (1486.6 eV), 400 μm spot size, pass energy of 50 eV, and a step size of 0.1 eV and resolution

of 0.6 eV. Samples for X-ray photoelectron spectroscopy (XPS) were prepared by drop coating a thick

layer of samples on the Si substrate and dried under vacuum. XPS analysis of core-level spectra was done

on fityk software using VoigtA function for peak fitting with a linear background. Field emission scanning

electron microscopy (FESEM) images were recorded on Thermo fFisher FEI Quanta 3D instrument

equipped with a field emission gun and Eenergy dispersive spectroscopy (EDS) detector.

Cathodoluminescence (CL) images and spectra were recorded with a Gatan MonoCL4 accessory with a

standard photomultiplier tube (PMT) detector (185–-850 nm) attached with FESEM instrument. Samples

for FESEM-EDS-CL analysis were prepared by drop coating a dispersion of Sb-FL on the Si substrate.

3. Results and discussionResult and discussion

3.1 Material characterization

Liquid exfoliation of antimony and bismuth crystals was carried out by probe sonication in N-methyl

pyrrolidone (NMP) followed by centrifugation at different rpm to separate two different grades of

exfoliated Sb and Bi: Sb/Bi nanosheets (Sb-NS and Bi-NS) and few-layered Sb/Bi (Sb-FL and Bi-FL)

(Experimental sectionExperimental section). Figure Fig. 1a gives a schematic of the liquid exfoliation of

heavier pnictogens in NMP. X-ray diffraction of Sb-NS and Bi-NS (Fig.Figure S1†) confirms that the

exfoliated sheets stabilize in the β-phase. Bright-field transmission electron microscopy images of Bi-FL

and Sb-FL show a layered-like morphology with surface roughness (Figure Fig. 1b and 1e). Atomic force

microscopy (AFM) analysis was carried out to understand the surface morphology and thickness of the

samples (Fig.Figure S2†). AFM particle size distribution reveals that the nanosheet (NS) grade has

thicknesses in the range of 6–-10 nm with a lateral dimensions of ~500 nm and the few-layered (FL) grade

have has thicknesses in the range of < 5 nm with lateral dimension ~500 nm. Thickness calculated from

AFM is overestimated due to the presence of residual solvent layers and contributions from capillary forces

and adhesion.25–28 Therefore, we expect the NS grade to be < 10 layers and the FL grade to be ~3 layers

thin.

6
The chemical purity of exfoliated samples was confirmed by Raman spectroscopy. Theoretically, bulk

rhombohedral pnictogens (Sb and Bi) have three Raman active optical modes, a pair of degenerate Eg

modes (Sb = 88 cm−1, Bi = 70 cm−1) corresponding to in-plane transversal and longitudinal vibrations in

opposite directions and an A1g mode (Sb = 137 cm−1; Bi = 97 cm−1) corresponding to opposite-in-phase out-

of-plane vibrations.28,29 Bulk bismuth shows two phonon modes Eg (68.9 cm−1) and A1g (95.7 cm−1) and we

did not observe any prominent shift in Bi-NS whereas in Bi-FL both the Eg and A1g peaks are blue-shifted

to 70.3 and 97 cm−1 respectively (Figure Fig. 1c). Experimentally, bulk antimony shows two phonon

modes, Eg (112 cm−1) and A1g (149.8 cm−1) which are blue-shifted to 114 cm−1 and 150.8 cm−1; 116.7 and

152 cm−1 for Sb-NS and Sb-FL, respectively (Figure Fig. 1f). The blue shift in phonon modes is due to the

contraction of the in-plane lattice constant with decreasing thickness.28 Oxides were not detected in Raman

spectra, establishing the chemical purity of the exfoliated samples. Moreover, Sb-FL and Bi-FL were stable

for over a year (Fig.Figure S3†). To understand their optical properties, exfoliated Bi and Sb sheets were

mapped by FESEM-cathodoluminescence spectroscopy. Understandingly, Bi-NS and Sb-NS did not show

any emission owing to their semi-metallic nature as predicted by their thicknesses. Interestingly, Bi-FL and

Sb-FL were emissive in the visible region, which implies an opening of the band gap owing to quantum

confinement and quasi-monolayer behavior due to turbostratic stacking (Figure Fig. 1d and 1g).30,31

3.2 Functionalization

Lewis acid-–base adduct formation is an interesting and viable strategy for tailoring pnictogen properties

with minimal lattice distortion while preserving their inherent properties.21,22 Pnictogens with lone pairs on

each atom, behave as a Lewis base and can interact with Lewis acids to form Lewis acid-–base adducts

(LABA) (Scheme 1).22 The Sstrength of an acid/base depends on two factors: Lewis acidity/basicity and

hard-–soft nature. In group 15, Lewis basicity decreases down the group with Sb being more Lewis basic

than Bi. According to the hard-–soft acid-–base (HSAB) principle,32 soft bases have low electronegativity,

are highly polarizable, and easily undergo oxidation, and hence among group 15 pnictogens, we expect Bi

to be the softest base followed by Sb. The strength/stability of the Lewis acid-–base adduct depends on the

inherent Lewis acidity/basicity of the two involved species along with their hard-–soft nature with the

strongest adducts formed between species of similar nature.

7
We have functionalized bismuthene and antimonene nanosheets (Bi-NS and Sb-NS) with group 13 halides

with Lewis acidity order BCl3 > AlCl3 ≥ GaCl3 > InCl3. BCl3 reacted instantaneously with a crackling

sound and did not yield a solid product implying complete oxidation/corrosion of the Sb/Bi lattice.

Reaction with AlCl3 and InCl3 yielded a solid product, wherein the reaction with AlCl3 was faster than

InCl3 while GaCl3 gave a colorless solution. This observation is not in line with the Lewis acidity order of

group 13 halides which predict GaCl3 to be of similar Lewis acidity to AlCl3. This anomalous behavior is

due to the increased softness of GaCl3 which results in a corrosive reaction with Sb/Bi. InCl3 being the

softest and weakest Lewis acid of the group forms the most stable adduct with Sb/Bi. To validate this

observation, functionalization of Sb/Bi was carried out with group 12 halides: ZnCl2 and CdCl2. Like

GaCl3, ZnCl2 did not yield a solid product while CdCl2 forms a stable adduct since its Lewis acidity and

softness are similar to InCl3. These observations indicate that for solid product formation the interaction of

the Lewis acid-–base needs to be moderate.

The Ssuccessful functionalization of bismuthene and antimonene was confirmed by energy dispersive

spectroscopy (EDS) mapping and X-ray photoelectron spectroscopy (XPS). EDS mapping of

functionalized materials shows uniform distribution of all the elements giving an initial indication of

functionalization (Figure Fig. 2). Core level XPS shows the presence of Al : Cl, In : Cl, and Cd : Cl in the

expected stoichiometry calculated by comparison of area ratios (with the exception of Sb-AlCl3) indicating

functionalization without complexation (Figure Fig. 3; and FiguresFig. S4 and, S5, see the ESI† for

details).

Charge transfer on functionalization was probed by Raman spectroscopy (Figure Fig. 4). Significant

shifts in Raman modes were observed in functionalized Bi and Sb. Eg and A1g modes of Bi-NS are

redshifted from 68.9 and 95.7 cm−1 to 67.6 and 93 cm−1 in Bi-CdCl2; and 67.6 and 91.7 cm−1 in Bi-InCl3

while it remains almost constant for Bi-AlCl3. Similarly, the A1g mode of Sb-NS (151 cm−1) is red-shifted to

149.9, 147.7, and 146.1 cm−1 upon functionalization with AlCl3, CdCl2, and InCl3, respectively, while we

could not detect any measurable shift in the Eg mode. A more prominent red shift in A1g vibrational mode

shows that functionalization affects the out-of-plane lattice constant more strongly compared to the in-plane

lattice constant.28 Red shifts in vibrational modes arise due to the strain introduced in the lattice owing to

8
charge transfer and steric repulsions.33 The Rred shift in Raman peaks on functionalization with Lewis acid

has been previously reported for InSe-–TiCl4 wherein the strain introduced on interaction with TiCl4 forces

Se out of its lattice.34 A similar red shift in Raman peaks is also reported for graphene, phosphorene, and

other materials.33,35,36 The extent of red shift on interaction with Lewis acids depends upon the Lewis

acidity/basicity and nature. From our study, we observe the highest red shift for InCl3 signifying stronger

charge transfer and steric repulsion in InCl3 followed by CdCl2 and negligible shifts in AlCl3. Raman modes

corresponding to Lewis acids were not observed in any of the functionalized samples.

To gain further insights into the bonding nature, core level XPS patterns were deconvoluted and thoroughly

analyzed (deconvolution details in the ESI†). Bi core level XPS consists of Bi 4f7/2 and 4f5/2 peaks. Bi 4f7/2

core level spectra of Bi-NS show two peaks at 156.8 and 158.1 eV corresponding to Bi(0) and surface Bi-

oxide species (Figure Fig. 5a). On functionalization, Bi 4f7/2 peaks can be deconvoluted into two peaks

corresponding to functionalized Bi and functionalized surface Bi-oxide species. The Ffunctionalized Bi

4f7/2 peak is blue-shifted by 0.9, 2.7 and 2.6 eV upon functionalization with AlCl3, InCl3 and CdCl2,

respectively. Based on the XPS core-level shift, both InCl3 and CdCl2 interacts to give more stable adducts.

Core level spectra of Al 2p, Cd 3d, and In 3d are correspondingly red-shifted by 1, 1.8, and 1.4 eV

indicating the transfer of electrons from Bi to Lewis acid (Figure Fig. 3).

Similarly, Sb 3d core level XPS consists of Sb 3d5/2, O 1s, and Sb 3d3/2 peaks. The Sb 3d5/2 peak overlaps

with the O 1s peak and was deconvoluted based on the peak position and area of Sb 3d3/2. The Sb 3d3/2 peak

for Sb-NS can be deconvoluted in two peaks at 537.7 eV and 539.3 eV corresponding to Sb(0) and Sb

oxide, with a high oxide content (Figure Fig. 5b). Upon functionalization, the Sb 3d3/2 peak can be

deconvoluted into three peaks corresponding to unreacted Sb(0), functionalized Sb, and functionalized Sb-

oxide. The Sb 3d3/2 peak is blue-shifted by 2.5 eV and 1.9 eV for InCl3 and CdCl2, respectively. This shift

gives an indication of the extent of electron transfer/strength of adduct formation, where InCl3 forms the

most stable adduct. However, in the case of Sb-AlCl3, we observed complete oxidation of Sb due to a

corrosive reaction with AlCl3. We did not observe any significant shift in the core-level spectra of In and

Cd for the functionalized Sb samples.

9
Blue shift in the Bi metal 4f7/2 peak and along with a prominent red shift in the Lewis acid metal core level

peak proves the successful functionalization and a stronger adduct formation in the case of Bi compared to

Sb. Functionalized Sb shows the presence of significant oxides of In and Cd (Fig.Figure S4†), however, a

similar feature was not observed for functionalized Bi (Figure Fig. 3) which further validates that Bi forms

stronger adducts with these Lewis acids compared to Sb and therefore is less susceptible to oxidation.

These observations further stress on our previous inference that a moderate Lewis acid-–base interaction is

desirable for effective functionalization. Moreover, there is a decrease of Sb/Bi-oxide after

functionalization indicating surface passivation. A Bblue shift in the core-level XPS of Sb and Bi along

with a red shift in Raman modes proves the effective functionalization of Sb and Bi with Lewis acids,

wherein InCl3 forms the most stable adduct followed by CdCl2. AlCl3 forms a weak adduct with Bi and

completely oxidizes Sb. Raman spectra of Sb-NS and Bi-NS do not show any oxide signatures, however,

XPS spectra show a high oxide content which could arise due to surface oxidation during sample storage.

The concentration of surface oxides gets significantly reduced after functionalization.

3.3 Effects of functionalization

To understand the effects of functionalization on the band structure, FESEM-cathodoluminescence studies

were carried out. The emission in nano-sized systems is very sensitive to their dimensions and hence in this

study, a single nanosheet was tracked before and after functionalization. Deconvolution of CL spectra was

based on control experiments carried out on multiple Sb-FL/Bi-FL sheets and is discussed in the ESI

(Fig.Figure S6 and S7†).

The broad feature in CL spectra of Bi-FL can be deconvoluted into four emissions (A, B, C, and D)

corresponding to absorption maxima (A)6 and other possible transitions and/or defect emissions (B, C, and

D) and a broad band below 500 nm corresponding to solvent emission (Figure Fig. 6a–c). The absence of

any emission in Bi-NS rules out the contribution from Bi3+ species.7,8 Absorption maxima (emission A) of

Bi-FL sheets studied for functionalization with AlCl3, InCl3 and CdCl2 occurs at 2.37, 2.24 and 2.31 eV,

respectively. Interestingly, these emissions are blue-shifted to 2.46, 2.29, and 2.36 eV, respectively after

functionalization (Figure Fig. 6a–c). Blue shifts were also observed in emissions B, C, and D which could

10
be due to the introduction of new states in the lattice and is more pronounced for emission B, which is split

into two emissions after functionalization.

The broad feature in CL spectra of Sb-FL can be deconvoluted into three emissions (A, B, and C), along

with a broad band below 500 nm corresponding to solvent emission, where emission A (~2.2 eV) is due to

band edge while B and C are due to other possible transitions and/or defect-related emissions (Figure Fig.

6d–f). We observe a blue shift in band edge emission to 2.27, 2.30, and 2.40 eV for AlCl3, InCl3, and

CdCl2, respectively (Figure Fig. 6d–f) for Sb-FL as well. Similar blue shifts were also observed in

emissions B and C with an increase in intensity for emission B which could be due to the introduction of

new defect states in the lattice. It is interesting to note that Sb-FL retains its structural integrity and

emission after functionalization with AlCl3 which could be due to a more controlled experimental

condition. Blue shifts and the introduction/splitting of band (emission B for Bi) in emission spectra indicate

rehybridization by the incorporation of Lewis acid metal bands in the band structure of antimonene and

bismuthene.24

The shifts in the XPS core level, Raman modes, and emission spectra indicates charge transfer from Sb/Bi

to Lewis acids, lattice distortion due to strain-induced by attachment of Lewis acids, and rehybridization of

the band structure on the successful functionalization of few-layered Sb/Bi (Table 1). Splitting/addition

of a new defect state in CL spectra of Bi-FL might be due to a stronger adduct formation compared to Sb

and corroborates with the XPS and Raman results.

4. Conclusions

This work successfully demonstrates Lewis acid-–base adduct formation in antimonene and bismuthene

with group 12 and 13 Lewis acids leading to the tuning of the band structure along with surface passivation.

The shifts in Raman modes, XPS core level, and emission spectra indicate lattice distortion, charge transfer,

and rehybridization of the band structure on the successful functionalization of exfoliated Sb/Bi. The

strength/stability of the adduct and the corresponding changes in the property properties depend on the

Lewis acidity and basicity of the interacting materials and their hard-–soft nature along with changes in

band structure due to rehybridization by Lewis acid metal bands. Based on these two parameters, different

Lewis acids can be used to achieve desirable tunability and surface passivation. The present study
11
demonstrates the efficacy of Lewis acid-–base adduct formation in tuning material properties and will be

instrumental for the incorporation of pnictogens for electronic and optoelectronic applications.

Author contributions

MB‡ and MMA‡ conceived the research, designed the experiments, and prepared the manuscript draft. S.

A. assisted in FESEM-cathodoluminescence measurements. C. N. R. R. supervised and conceived the

research. All authors have given approval to the final version of the manuscript. ‡These authors contributed

equally.

Conflicts of interest

The authors declare no competing financial interest.

Acknowledgements

MMA and MB acknowledge DST and CSIR, India for fellowship. SA acknowledges ICMS and

JNCASR for fellowship. The Aauthors acknowledge ICMS, SSL, JNCASR, SAMat for facilities.

The Aauthors thank Dr. C. P. Vinod for XPS measurements and Prof. S. Rajaram for scientific

discussion.

References

1 S. Zhang, M. Xie, F. Li, Z. Yan, Y. Li, E. Kan, W. Liu, Z. Chen and H. Zeng, Angew. Chem., Int.

Ed.Angew. Chem. Int. Ed., 2016, 55, 1666.

2 J. Gusakova, X. Wang, L. L. Shiau, A. Krivosheeva, V. Shaposhnikov, V. Borisenko, V. Gusakov and B.

K. Tay, Phys. Status Solidi A, 2017, 214, 12.

3 C. R. Ryder, J. D. Wood, S. A. Wells, Y. Yang, D. Jariwala, T. J. Marks, G. C. Schatz and M. C. Hersam,

Nat. Chem., 2016, 8, 597–602.

4 G. Wang, R. Pandey and S. P. Karna, ACS Appl. Mater. Interfaces, 2015, 7, 11490–11496.

5 S. Zhang, S. Guo, Z. Chen, Y. Wang, H. Gao, J. Gómez-Herrero, P. Ares, F. Zamora, Z. Zhu and H.

Zeng, Chem. Soc. Rev., 2018, 47, 982–1021.

12
6 D. Kecik, V. O. Özçelik, E. Durgun and S. Ciraci, Phys. Chem. Chem. Phys., 2019, 21, 7907–7917.

7 L. Kumari, J. H. Lin and Y. R. Ma, J. Phys. D: Appl. Phys., 2008, 41, 025405.

8 N. Hussain, T. Liang, Q. Zhang, T. Anwar, Y. Huang, J. Lang, K. Huang, H. Wu, N. Hussain, T. Liang,

Q. Zhang, Y. Huang, J. Lang, K. Huang, H. Wu and T. Anwar, Small, 2017, 13, 1701349.

9 M. H. Ludwig, R. E. Hummel and M. Stora, Thin Solid FilmsThin Solid Films, 1995, 255, 103–106.

10 S. Zhang, Z. Yan, Y. Li, Z. Chen and H. Zeng, Angew. Chem., Int. Ed.Angew. Chem. Int. Ed., 2015, 54,

3112–3115.

11 X. Wang, J. He, B. Zhou, Y. Zhang, J. Wu, R. Hu, L. Liu, J. Song and J. Qu, Angew. Chem., Int.

Ed.Angew. Chem. Int. Ed., 2018, 57, 8668.

12 F. Zhang, J. He, Y. Xiang, K. Zheng, B. Xue, S. Ye, X. Peng, Y. Hao, J. Lian, P. Zeng, J. Qu and J.

Song, Adv. Mater., 2018, 30, 38.

13 H. S. Tsai, C. W. Chen, C. H. Hsiao, H. Ouyang and J. H. Liang, Chem. Commun., 2016, 52, 8409–

8412.

14 Y. Liu, Y. Xiao, M. Yu, Y. Cao, Y. Zhang, T. Zhe, H. Zhang and L. Wang, Small, 2020, 16, 42.

15 G. R. Bhimanapati, Z. Lin, V. Meunier, Y. Jung, J. Cha, S. Das, D. Xiao, Y. Son, M. S. Strano, V. R.

Cooper, L. Liang, S. G. Louie, E. Ringe, W. Zhou, S. S. Kim, R. R. Naik, B. G. Sumpter, H. Terrones, F.

Xia, Y. Wang, J. Zhu, D. Akinwande, N. Alem, J. A. Schuller, R. E. Schaak, M. Terrones and J. A.

Robinson, ACS Nano, 2015, 9, 11509–11539.

16 R. Balog, B. Jørgensen, L. Nilsson, M. Andersen, E. Rienks, M. Bianchi, M. Fanetti, E. Lægsgaard, A.

Baraldi, S. Lizzit, Z. Sljivancanin, F. Besenbacher, B. Hammer, T. G. Pedersen, P. Hofmann and L.

Hornekær, Nat. Mater., 2010, 9, 315–319.

17 N. Martín, N. Tagmatarchis, Q. H. Wang and X. Zhang, Chem. – Eur. J., 2020, 26, 6292–6295.

18 Z. Xie, B. Zhang, Y. Ge, Y. Zhu, G. Nie, Y. F. Song, C. K. Lim, H. Zhang and P. N. Prasad, Chem.

Rev., 2021, 122, 1127–1207.

19 C. Jellett, J. Plutnar and M. Pumera, ACS Nano, 2020, 14, 7722–7733.

13
20 V. Georgakilas, M. Otyepka, A. B. Bourlinos, V. Chandra, N. Kim, K. C. Kemp, P. Hobza, R. Zboril

and K. S. Kim, Chem. Rev., 2012, 112, 6156–6214.

21 H. Ghodrati, N. Antonatos and Z. Sofer, Small, 2019, 15, 43.

22 D. Tofan, Y. Sakazaki, K. L. Walz Mitra, R. Peng, S. Lee, M. Li and A. Velian, Angew. Chem., Int.

Ed.Angew. Chem. Int. Ed., 2021, 60, 8329.

23 G. Abellán, P. Ares, S. Wild, E. Nuin, C. Neiss, D. R.-S. Miguel, P. Segovia, C. Gibaja, E. G. Michel,

A. Görling, F. Hauke, J. Gómez-Herrero, A. Hirsch and F. Zamora, Angew. Chem., Int. Ed.Angew. Chem.

Int. Ed., 2017, 56, 14389.

24 A. Ienco, G. Manca, M. Peruzzini and C. Mealli, Dalton Trans., 2018, 47, 17243–17256.

25 P. Nemes-Incze, Z. Osváth, K. Kamarás and L. P. Biró, Carbon, 2008, 46, 1435–1442.

26 C. Backes, R. J. Smith, N. Mcevoy, N. C. Berner, D. Mccloskey, H. C. Nerl, A. O'neill, P. J. King, T.

Higgins, D. Hanlon, N. Scheuschner, J. Maultzsch, L. Houben, G. S. Duesberg, J. F. Donegan, V. Nicolosi

and J. N. Coleman, Nat. Commun., 2014, 5, 4576.

27 D. Hanlon, C. Backes, E. Doherty, C. S. Cucinotta, N. C. Berner, C. Boland, K. Lee, A. Harvey, P.

Lynch, Z. Gholamvand, S. Zhang, K. Wang, G. Moynihan, A. Pokle, Q. M. Ramasse, N. McEvoy, W. J.

Blau, J. Wang, G. Abellan, F. Hauke, A. Hirsch, S. Sanvito, D. D. O’Regan, G. S. Duesberg, V. Nicolosi

and J. N. Coleman, Nat. Commun., 2015, 6, 8563.

28 C. Gibaja, D. Rodriguez-San-Miguel, P. Ares, J. Gómez-Herrero, M. Varela, R. Gillen, J. Maultzsch, F.

Hauke, A. Hirsch, G. Abellán and F. Zamora, Angew. Chem., Int. Ed.Angew. Chem. Int. Ed., 2016, 55,

14345–14349.

29 K. Trentelman, J. Raman Spectrosc., 2009, 40, 585–589.

30 S. Latil, V. Meunier and L. Henrard, Phys. Rev. B Condens. Matter Mater. Phys.Phys. Rev. B: Condens.

Matter Mater. Phys., 2007, 76, 201402.

31 R. Negishi, C. Wei, Y. Yao, Y. Ogawa, M. Akabori, Y. Kanai, K. Matsumoto, Y. Taniyasu and Y.

Kobayashi, Phys. Status Solidi B, 2020, 257, 2.

32 R. G. Pearson, J. Chem. Ed.J. Chem. Educ., 1968, 45, 581–587.

14
33 S. B. Pillai, H. R. Soni and P. K. Jha, Springer Proc. Phys.Springer Proceedings in Physics, 2019, 236,

379–387.

34 S. Lei, X. Wang, B. Li, J. Kang, Y. He, A. George, L. Ge, Y. Gong, P. Dong, Z. Jin, G. Brunetto, W.

Chen, Z. T. Lin, R. Baines, D. S. Galv’o, J. Lou, E. Barrera, K. Banerjee, R. Vajtai and P. Ajayan, Nat.

Nanotechnol., 2016, 11, 465–471.

35 Z. H. Ni, T. Yu, Y. H. Lu, Y. Y. Wang, Y. P. Feng and Z. X. Shen, ACS Nano, 2008, 2, 2301–2305.

36 R. Fei and L. Yang, Appl. Phys. Lett., 2014, 105, 83120.

Fig.Figure 1 (a) Schematic illustration of liquid exfoliation of Sb/Bi in NN-methyl pyrrolidone (NMP) by

probe sonication; TEM images of (b) Bi, (e) Sb; Raman spectra of (c) Bi and (f) Sb; FESEM-

cathodoluminescence mapping of few-layered (d) Bi and (g) Sb.

Fig.Figure 2 FESEM Eenergy dispersive spectroscopy (EDS) mapping of functionalized Bi: (a) Bi-AlCl3,

(b) Bi-InCl3 and (c) Bi-CdCl2, and functionalized Sb: (d) Sb-AlCl3, (e) Sb-InCl3 and (f) Sb-CdCl2.

Fig.Figure 3 XPS core level spectra of (a) Al 2p Bi-AlCl3, (b) In 3d Bi-InCl3 and (c) Cd 3d Bi-CdCl2.

Fig.Figure 4 Raman spectra of (a) functionalized Bi nanosheets (Bi-NS) with AlCl3 (Bi-AlCl3), CdCl2 (Bi-

CdCl2) and InCl3 (Bi-InCl33) and (b) functionalized Sb nanosheets (Sb-NS) with AlCl3 (Sb-AlCl3), CdCl2

(Sb-CdCl2) and InCl3 (Sb-InCl3).

Fig.Figure 5 XPS core level spectra of (a) functionalized Bi nanosheets (Bi-NS) with AlCl3 (Bi-AlCl3),

CdCl2 (Bi-CdCl2) and InCl3 (Bi-InCl3), and. (b) functionalized Sb nanosheets (Sb-NS) with AlCl3 (Sb-

AlCl3), CdCl2 (Sb-CdCl2) and InCl3 (Sb-InCl3).

Fig.Figure 6 Cathodoluminescence mapping and spectra before and after functionalization of (a) Bi-AlCl 3,

(b) Bi-CdCl2 and (c) Bi-InCl3. (d) Sb-AlCl3, (e) Sb-CdCl2, and (f) Sb-InCl3.

Scheme 1 Schematic illustration of the interaction of pnictogens (Sb and Bi) with Lewis acids (MX3 and

MX2; M = B, Al, Ga and In; X = Cl) to form Lewis acid-–base adduct.

Table 1 Summary of experimental results on functionalization

15
Shifts after Bi Sb

functionalizatio
XPS Raman CL XPS Raman CL
n with
emission emission
Bi 4f7/2 (Eg, A1g) Sb 3d3/2 (Eg, A1g)
A (eV) A (eV)
(eV) (cm−1) (eV) (cm−1)

AlCl3 0.9 0, 0 -−0.09 --— 0, -−1.1 -−0.07

InCl3 2.7 -−1.3, - -−0.05 2.2 0, -−4.9 -−0.1

−4

CdCl2 2.6 -−1.3, - -−0.05 1.8 0, -−3.3 -−0.2

−2.7

16

You might also like