You are on page 1of 10

Article

Cite This: Chem. Mater. 2019, 31, 1784−1793 pubs.acs.org/cm

Unraveling the Growth Mechanism Forming Stable γ‑In2S3 and


β‑In2S3 Colloidal Nanoplatelets
Faris Horani and Efrat Lifshitz*
Schulich Faculty of Chemistry, Solid State Institute, Russell Berrie Nanotechnology Institute, Nancy and Stephen Grand Technion
Energy Program, TechnionIsrael Institute of Technology, Haifa 3200003, Israel
*
S Supporting Information

ABSTRACT: Inorganic two-dimensional semiconductor nano-


Downloaded via UNIV NACIONAL AUTONOMA MEXICO on November 6, 2023 at 18:48:43 (UTC).
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

structures intrigue the scientific community because of their


tunable and sparse electronic states and their high conductivity.
The current work deals with colloidal nanoplatelets (NPLs)
based on In2S3 compound, focusing on the growth mechanism
that leads to the formation of two different phases, trigonal γ-
In2S3 and defect spinel β-In2S3, both stabilized at room
temperature and characterized by ordered metal voids. In
particular, we substantiate the experimental factors (e.g.,
temperature, reaction duration, and surface ligands) that
control the growth progress. The results indicated the
formation of hexagonal NPLs of the γ-phase at an elevated
temperature and dodecagon NPLs of the γ-phase at a lower
temperature. A long-reaction duration time transformed the
hexagons/dodecagons into truncated triangular shapes. Furthermore, the analysis of thermodynamic and kinetic factors
indicated a phase transformation from the γ-phase to the β-phase. All phases were produced by a new colloidal procedure based
on a single precursor. The structures created were verified by X-ray diffraction, high-resolution transmission electron
microscopy analyses, and Raman measurements. Elementary optical properties were identified by absorption and emission
measurements. The nanoplatelets discussed offer low toxicity and optical activity in the UV and visible spectral regimes and an
option for electrical or magnetic doping, enabled by the existing voids.

■ INTRODUCTION
Chalcogenide semiconductors have attracted considerable
conditions has been an obstacle since 1976.9,12−14 The α-In2S3
phase was prepared at 420−700 °C and showed a partial
scientific and technological interest because of their combina- chemical stability.15,16 Nevertheless, In2S3 bulk or thin-film
tion of optical and electrical properties. The most commonly forms have already demonstrated a practical use as a buffer
explored chalcogenides are based on the II−VI and IV−VI layer material replacing the CdS thin films in solar cells17−19
groups, such as CdS, CdTe, PbSe, and PbS. However, the and showed improvement of the open circuit voltage and
introduction of these materials in modern industrial current collection in the blue wavelength region.18,20 More-
applications, including photonics, optoelectronics, and bio- over, In2S3 has been used as efficient photocatalysts,21−23
logical imaging, is limited due to toxicity impact on the photodetectors,24,25 electrochemical storage cells, and display
environment.1,2 Therefore, there has been an increased panels.26−29
demand to develop indium-based chalcogenide materials New efforts have been demonstrated in recent years for the
from the III−VI group of elements.3 Indium sulfide (In2S3) development of In2S3 two-dimensional (2D) nanostructures,
is a promising candidate because of its low toxicity, high for example, thin films or nanoplatelets (NPLs), using either
absorption coefficient,4 photoelectrical properties,5−7 long- chemical vapor deposition or colloidal chemical proce-
term stability, and wide band gap energy in its bulk form.8 dures.30−35 It is anticipated that these 2D materials should
An In2S3 bulk semiconductor has a large exciton Bohr radius possess a tunable electronic band structure, high conductivity,
of 33.8 nm and exists in three crystalline polymorphs: α-In2S3 and efficient optical transitions.36,37 Moreover, the unique
(defect cubic), β-In2S3 (defect spinel), and γ-In2S3 (layered defect cubic and defect spinel structures of the In2 S3
hexagonal).9 Among the three phases, β-In2S3 is the most compound include voids, which can accommodate functional
chemically stable (up to 420 °C) and the most studied phase electrical or magnetic dopants, useful attribute for practical
with n-type conductivity and a wide band gap of 1.9−2.4
eV.10,11 γ-In2S3 was found to be thermally unstable, existing Received: January 2, 2019
only at elevated temperatures (7751045 °C) or high Revised: February 14, 2019
pressures (35 kbar), and therefore, its isolation at ambient Published: February 15, 2019

© 2019 American Chemical Society 1784 DOI: 10.1021/acs.chemmater.9b00013


Chem. Mater. 2019, 31, 1784−1793
Chemistry of Materials Article

Figure 1. (a) Schematic illustration of a colloidal synthesis setup (top) and the chemical reactions used for the formation of a single molecule
precursor and its decomposition product, the In2S3 compound (bottom). (b) Representative TEM images of In2S3 NPLs produced at various
temperatures and reaction durations, revealing the formation of hexagon, dodecagon, and truncated triangular-shaped NPLs. The lateral length
values of each shape are given on the diagram. (c) Scheme reproducing the NPL morphologies in correlation with the TEM observations shown in
(b), illustrating the morphology dependence on reaction temperature and on evolution of the reaction stages (nucleation, growth, and morphology
change) with progress in time. (d) Photographs of aliquots withdrawn from a reaction produced at 230 °C, exhibiting a color change as extended
reaction time.

applications. Indeed, recent studies used Cu- or Mn-doped β-


In2S3 nanostructures in prototype solar cells and light-emitting
■ EXPERIMENTAL SECTION
Chemicals and Materials. Indium (III) acetate (In(OAc)3,
diodes.38−41 The structure of shape stability was discussed only 99.99%), 1-dodecanthiol (DDT, ≥98%), oleylamine (OLAm;
in a single case, showing transformation of wrinkled technical grade, 70%). Methanol (absolute), toluene (analytical),
nanostructures into nanodisk shapes.41 Despite the efforts hexane (analytical), and isopropanol (cp) were purchased from Bio-
Lab Ltd. These chemicals were used without further purification.
mentioned in preparation and characterization of nanoscale
Synthesis of Hexagonal and Dodecagonal γ-In2S3 NPLs.
In2S3 phases, the fundamental growth mechanism and the In2S3 NPLs were prepared using a colloidal procedure, involving
factor inducing structural stabilization or transformation are either injecting or dissolving precursors in a mother solution,
inadequately understood. including stabilizing ligands. All synthesis procedures were undertaken
The current work focuses on colloidal 2D In2S3 NPLs, using standard Schlenk line techniques assisted by a nitrogen-filled
having merits beyond the corresponding thin films, when glovebox. The growth of In2S3 NPLs was initiated by preparation of a
precursor consisting of both the indium and chalcogenide elements,
prepared as free-standing species in solutions, covered by noted here as a single precursor source. The precursor was prepared
organic ligands, produced in variable sizes, and suitable for by mixing In(OAc)3 (0.292 g, 1 mmol), OLAm (5.0 mL), and DDT
incorporating in practical applications. The current work deals (5.0 mL) in a three-neck round-bottom flask, equipped with a
with colloidal nanoplatelets based on an In2S3 compound, thermocouple and a magnetic stirrer. Then, the mixture was degassed
focusing on the growth mechanism that leads to the formation at 100 °C for 1 h, leading to the formation of a colorless In thiolate
[In(DDT)3] solution, where In(DDT)3 is considered herein as a
of two different phasestrigonal γ-In2S3 and defect spinel β-
single precursor. Thereafter, the mixture was further heated under a
In2S3both stabilized at room temperature. We discuss the dry inert nitrogen gas flow until a yellowish color was developed
influence of the experimental factors (temperature and reaction (≥215 °C), accompanied by decomposition of the single precursor
duration) on the phase growth mechanism, and we show, for into In2S3 monomers. The synthesis stages are shown schematically in
the first time, a structural transformation from the γ-In2S3 to Figure 1a. The growth process, which was continued up to 32 min at
the β-In2S3 phase. The results revealed a variable shape for the 250 °C, produced hexagon shapes, whereas reaction duration up to 64
min at 230 °C produced NPLs with dodecagon shapes. The reaction
different phases: (a) a dodecagon shape in the γ-In2S3 phase was terminated by removal of the heating mantle and prompt cooling.
when using a relatively low temperature and short reaction Then, toluene, isopropanol, and methanol were added to the reaction
duration, (b) a γ-In2S3 hexagon shape that was prepared at an solution, and the NPLs were separated by centrifugation via two
elevated temperature and short reaction duration, and (c) a β- cycles. Afterward, the NPLs were dried and then redissolved in
In2S3 truncated triangular shape using low/high temperatures organic solvent such as hexane or toluene for further characterization.
Extracentrifugation and drying rounds are responsible for significant
after long reaction duration. All phases were produced by a
desorption of surface-bound ligands and aggregation of the NPLs.
new colloidal procedure based on a single precursor, which Structural characterizations as discussed below confirmed the
lead to creation of structural stability at room temperature. The formation of hexagon and dodecagon NPLs of the γ-phase.
phases produced were verified by X-ray diffraction, high- Synthesis of Truncated Triangular β-In2S3 NPLs. The same
resolution transmission electron microscopy, and Raman chemical procedures used for the growth of the γ-phase were
measurements. The elementary optical properties were implemented to the growth of the β-phase; however, the reaction time
was extended up to 128 min at 230 °C or 64 min at 250 °C. The X-
identified by absorption and emission measurements. The ray and electron microscopy measurements (see Results and
knowledge accumulated indicates future potential use of the Discussion section) proved the formation of NPLs with a truncated
In2S3 phases in various optoelectronic devices. triangular shape with a β-In2S3 crystallographic structure. Hence,

1785 DOI: 10.1021/acs.chemmater.9b00013


Chem. Mater. 2019, 31, 1784−1793
Chemistry of Materials Article

increase of the reaction duration beyond 64 min at 230 °C (or after schemes of the produced NPL transformations, compatible
32 min at 250 °C) induced the γ- to β-phase transition accompanied with the TEM images given in Figure 1b. Figure 1d exhibits a
with a gradual color change from yellow to dark orange. The snapshot photo of aliquots withdrawn from the reaction
purification of the product is similar to the procedure mentioned solution after 32, 64, and 128 min. This figure reveals a change
above for γ-In2S3 NPLs.
X-ray Diffraction (XRD). Thin films of In2S3 NPLs were prepared
in a color from pale yellow through bright orange to dark
by concentrated suspension drop-casting onto clean glass microscope orange. The yellow solution is associated with the formation of
slides and drying for further characterization. The X-ray diffraction both hexagon and dodecagon morphologies, whereas the dark
patterns were then acquired on a Rigaku SmartLab 9.0 kW orange solution corresponds to the truncated triangular
diffractometer with the X-ray source operating at 45 kV and 150 morphology.
mA. The instrument was equipped with Cu Kα radiation (λ = 1.5418 In2S3 Crystallographic Structures. The previous section
Å) and was used in the 2θ scan geometry for data acquisition using discussed the correlation of the reaction conditions with the
parallel beam (PB) geometry. visual appearance, namely, crystalline morphology and
Transmission Electron Microscopy (TEM) and Elemental apparent color, of the reaction solutions. This section supplies
Analysis. Transmission electron microscopy (TEM) images, high-
resolution TEM (HR-TEM) images, and energy dispersive X-ray a detailed description of the different In2S3 crystallographic
spectroscopy (EDS) spectra were taken by using FEI Tecnai T20 structures. Figure 2a depicts a set of X-ray diffraction patterns
operated at 200 keV. High-angle annular dark-field (HAADF)
imaging in an STEM mode (HAADF-STEM) was obtained by an
FEI Titan 300keV S-TEM system. The samples for TEM measure-
ment were prepared by dropping the solution containing NPLs on an
ultrathin carbon-coated copper grid at room temperature.
Absorbance and Photoluminescence Characterization.
Optical absorbance spectra of diluted hexane dispersions of NPLs
were recorded in quartz cuvettes of 1 cm path length at different
reaction stages and times using a JASCO V-570 UV−vis−NIR
spectrometer and a Shimadzu (UV-1800) spectrophotometer. Photo-
luminescence spectra were obtained by a JobinYvon (Fluorolog-3)
fluorometer. Both absorbance and photoluminescence spectra were
collected in air and at room temperature.
Raman Spectroscopy. The sample was prepared by drop-casting
a concentrated suspension onto a clean silicon/glass substrate. The
measurements were performed with a Horiba JobinYvon (LabRAM
HR Evolution) confocal Raman microscope, using a visible laser
excitation source (532 nm). The spectra were recorded by using 50× Figure 2. (a) X-ray diffraction patterns of three aliquots withdrawn
and 100× objective lenses and a grating of 1800 gr/mm. from the reaction solution as different time intervals. The bottom
Crystal Structure Design. γ-In2S3 and β-In2S3 NPL crystal black line is attributed to the γ-phase and corresponds to an aliquot
structures were designed by Crystal Maker software. Several taken after 32 min. The top red line is attributed to the β-phase and
parameters were calculated for each unit cell structure model, for corresponds to an aliquot taken after 128 min. The middle pattern
example, filled space, void space, unit cell volume, density, atoms per (green line) is attributed to an intermediate phase and corresponds to
unit cell, and bond lengths. The calculated parameters are corrected an aliquot taken after 64 min at 230 °C (or time interval between 32
for first-nearest-neighbor sphere overlap and site visibility. and 64 min at 250 °C). Two sets of stick diagrams at the bottom


(black) and the top (red) of the frame are associated with the
database diffraction angles of the γ-In2S3 and β-In2S3 phases,
RESULT AND DISCUSSION respectively. (b) Expanded scale of a diffraction regime between 25
In2S3 NPLs with Different Morphologies. In2S3 NPLs and 32° (yellow area) from panel (a), displaying a comparison
were prepared using a colloidal procedure as described in between the intermediate pattern (middle) with those of the γ- and β-
Experimental Section above, using In(DDT)3 precursors as the phases. The orange Gaussian fit depicts a deconvolution of the green
source of In2S3 monomers for the growth of platelets (see pattern into two peaks indexed by (I) and (II). The blue vertical lines
Figure 1a). The nucleation process initiated only after increase are drawn to guide the eye.
of the temperature up to 230 or 250 °C.
Figure 1b demonstrates a set of TEM images associated with of different aliquots taken at intermediate stages during the
the produced NPLs, extracted at various stages of the reaction. NPLs growth. The observed diffraction angles of each aliquot
The differences among the images reflect the morphology and the corresponding crystallographic planes and interplanar
evolution with temperature and time. Figure 1b shows that a d-spacing are summarized in Table 1. The bottom pattern in
reaction at T = 230 °C for half an hour induces crystalline Figure 2a (black line) corresponds to an aliquot taken at 32
growth of a rounded twelve-sided polygon called a dodecagon. min, characterized by a yellow color and a hexagon or
Then, the extension of the reaction to 1 h at the same dodecagon morphology. This X-ray pattern contains two
temperature produced a dodecagon with clearly visible sides. A intense peaks at diffraction angles as listed in Table 1. The top
reaction at T = 250 °C for half an hour led to the formation of pattern in Figure 2a (red line) corresponds to an aliquot taken
hexagon-shaped In2S3 NPLs. Both hexagon and dodecagon after 128 min, characterized by a dark orange color and a
NPLs have a thickness value of ∼1.8nm (see Supporting truncated triangular morphology. The top pattern exhibits
Information, Figure S1). However, the extension of the eight different diffraction peaks, as given in Table 1. Beside the
reaction duration at each temperature, that is, 230 and 250 experimental observations, Figure 2a contains two sets of stick
°C, beyond 64 and 32 min, respectively, brought about a diagrams at the bottom and the top of the frame, associated
transformation from hexagon or dodecagon crystallites to with the database diffraction angles of the γ-In2S3 and β-In2S3
truncated triangular-shaped NPLs. The lateral dimensions of phases, respectively (see also Table 1). A comparison between
the different shapes are shown in Figure 1b. Figure 1c displays the experimental patterns with the database sets designates that
1786 DOI: 10.1021/acs.chemmater.9b00013
Chem. Mater. 2019, 31, 1784−1793
Chemistry of Materials Article

Table 1. Summary of Diffraction Angles, Interplanar d- (II). The comparison shown in Figure 2b emphasizes the
Spacing, and Corresponding Crystallographic Planes of the similarity of the green pattern to those of the identified phases
X-ray Diffractions Shown in Figure 2a (black and red patterns) and the diffractions angles lying
between those of the phases. Furthermore, the green
diffraction peaks show pronounced broadening. The observa-
tions designate the occurrence of a structural phase trans-
formation. The broadening can be related to the existence of
structural domains at an intermediary stage of phase
transformation, with an inhomogeneous domain strain and
structural reconstruction via creation of vacancies, dislocations,
or layer faults.42 Overall, the X-ray observations discussed
reveal unequivocally, for the first time, the occurrence of
structural phase transition from the γ-In2S3 to β-In2S3 phase.
γ-In2S3 Crystal Structure. The γ-phase of In2S3 accom-
modates a corundum-type structure of crystalline alumina (α-
Al2O3), crystallizing with a trigonal symmetry of D3d and can
a
Asterisk (*) denotes the intermediate planes during phase be described by a primitive right rhombic prism with lattice
transformation. bβ-In2S3 reference: (JCPDS 25-0390). cγ-In2S3 parameters of a = 6.56 Å and c = 17.57 Å (c/a = 2.67). A
reference: (JCPDS 33-0624). polyhedral representation of the γ-In2S3 structure is displayed
in Figure 3a, showing that a single unit cell is built from
the hexagon and dodecagon morphologies have a γ-In2S3 octahedral connectivity. Each octahedron is composed of an
crystal structure, whereas the truncated triangular shape has In+3 cation binds to six S−2 ions, and each S−2 anion is
a β-In2S3 crystal structure. surrounded by four In+3 cations. Furthermore, two adjacent
An additional diffraction pattern is illustrated in Figure 2a octahedra are arranged with an antiprismatic triangular face-
(green curve) and related to an aliquot taken from the reaction sharing along the [001] direction and edge-sharing with three
solution after 64 min at 230 °C (or time interval between 32 neighboring octahedra along the N[001] direction, as drawn in
and 64 min at 250 °C). The green pattern nominally includes Figure 3b. A face- and an edge-sharing induce structural
two reflection peaks at 27.27 and 47.96° with the distortions that are particularly pronounced in materials with
corresponding d-spacing values of 3.26 and 1.90 Å (see relatively large cations and valence electron volume. Such
Table 1). A blowup of a diffraction regime between 25 and 32° structural distortions stimulate displacement of cations from
(highlighted by a yellow background in Figure 2a) is illustrated the octahedral centers or migration to vacant positions, all to
in Figure 2b, enabling a comparison between the green pattern minimize the cation−cation electrostatic repulsion. A displace-
(middle) with those of the γ- and β-phases. Figure 2b depicts a ment of the cation center, a Jahn−Teller distortion, creates
deconvolution of the green pattern by a Gaussian fit, revealing unequal In−S bond lengths (2.746 and 2.513 Å) within a
the existence of two diffraction composites at that stage, unitary octahedron, as displays in Figure 3c. Considering the
emphasized by the dashed orange curves and labeled as (I) and local distortions, a cut view of a unit cell along the [112̅0]

Figure 3. (a) Right rhombic prism of the In2S3 unit cell represented by polyhedral model. (b) Blowup of a segment of a unit cell, showing face-
sharing (red trigonal plane) and edge-sharing (red line) between neighboring octahedral units. (c) Single distorted octahedral unit composed of
two different In−S bond lengths as marked on the figure. (d) Cut view of the polyhedron shown in (a), exposing the (112̅0) plane and representing
the indium (In+3) and sulfur (S−2) layer sequence along the [0001] direction. The S−2 layers are given by the indexes 1 and 3, and the In+3 layers
are depicted by the symbols A, B, and C. Pink lines represent the unit cell frame, and blue dashed lines exhibit the nonflat indium planes within two
S−2 layers. (e) Hexagonal closed-packed (HCP) crystal structure form of the In2S3 γ-phase, consisting of three right primitive rhombohedral unit
cells. The HCP presentation reveals three types of planes: the m-plane (101̅0), which is marked by the orange area, the a-plane (112̅0) marked by
the green area, and the c-plane (on top). The a- and m-planes are rotated by an angle of 30° (pink arrow). (f) A stereographic projection
representing 12 crystallographic directions, 6 growth directions of the m-planes (orange vectors), and another 6 growth directions of the a-planes
(green vectors). The threefold axis symmetry element of the γ-phase with the D3d point group symmetry is shown at the center (black triangle).

1787 DOI: 10.1021/acs.chemmater.9b00013


Chem. Mater. 2019, 31, 1784−1793
Chemistry of Materials Article

direction was derived, as illustrated by a ball-filling


presentation in Figure 3d. The unit cell structure is composed
of six layers, each consisting of close-packed S2− anions (yellow
balls). The In3+ cations (purple balls) occupy two-thirds of the
octahedral hole sites between the adjacent S2− layers. The
layers are stacked in a sequence along the [0001] direction,
when layers 1 and 3 correspond to the S−2 layers, and layer 2
represents the In+3 cations at interstitial positions. The S−2
anion layers are flat and overlap each other. In contrast,
individual In+3 layers have a nonflat appearance, where part of
the cations lies closer to atop anion layer and other cations lie
closer to the bottom anion layer (see small-framed section).
Furthermore, adjacent In+3 layers have a lateral displacement
from one another. They are indexed by A, B, and C. The filled Figure 4. (a) Tetragonal defect spinel unit cell structure of the β-In2S3
ball display also reveals a mirror-image arrangement, with a phase, represented by a polyhedral model, including two types of
mirror plane at halfway, and while each fraction has a dipole octahedra, tetrahedron, and vacancies. (b) Ball-and-stick presentation
polarity, the overall unit cell has a zero-dipole moment. of the β-In2S3 phase, emphasizing the existence of vacancies, when
The correlation between the γ-In2S3 crystal structure and the sulfur atoms appear as yellow balls, indium atoms as green/red/black
morphology of the NPLs is discussed below. It might be easier balls, and electron-rich vacancy sites (VIn) are shown by the blue
clouds, whereas the hole-rich centers (VS) are displayed by black
to follow such a discussion by viewing the γ-phase crystal circles. (c) Three types of coordination centers consisting of two
structure as three right primitive rhombohedral unit cells, distorted octahedra (left and right) and one tetrahedron (middle);
enfolded within a primitive hexagonal close-packed (HCP) each coordination geometry is composed of different In−S bond
frame. An HCP prism is enclosed with two c-planes {0001}, six lengths. (d) Schematic display of the connectivity among the
first-order m-planes {101̅0}, and six second-order a-planes octahedron/tetrahedron units of a certain fraction in the unit cell,
{112̅0}, as illustrated in Figure 3e. The angle between the m- revealing a corner-bridging connection (pink dashed circle) along the
and a-planes is 30°, and overall, the m- and a-planes are normal [001] direction and an edge-sharing bond (pink lines) along the
to those of the c-planes. Figure 3f illustrates a stereographic N[001] direction.
projection of the m- and a-planes shown by orange and green
arrows, respectively, and the direction of the planes is marked direction or through edge-sharing along the normal direction
on the circumference of the projection. N[001], without incidence of a face-sharing connectivity. It is
β-In2S3 Crystal Structure. β-In2S3 crystallizes with a worth noting that β-In2S3 is characterized by their intrinsic VIn
tetragonal crystal symmetry, described as a defect spinel centers; however, some VS can occur accidentally, and both
structure with lattice parameters of a = 7.62 Å and c = 32.36 Å defects induce charge trapping of opposing types, which can
(c/a ratio = 4.25),43 as illustrated by a polyhedron eventually lead to the donor−acceptor type of recombina-
representation in Figure 4a. The latter presentation reveals tion.44
that a defect spinel structure is composed of two kinds of Correlation between Morphology and Crystal Struc-
octahedra and one type of tetrahedron unit, as indicated in the ture of the In2S3 Phases. The previous sections discussed
inset of Figure 4a. Unlike the case of γ-In2S3, In3+ centers are separately the polymorphs and the fundamental crystal
surrounded either by six or by four S2− anions in the octahedral structures of In2S3 NPLs, whereas this section discusses the
(InS6) and tetrahedral (InS4) units, respectively. Figure 4b correlation between them by following the various observa-
illustrates a ball-and-stick presentation of the β-In2S3 phase, tions, which are illustrated in Figure 5. Figure 5a presents a
designating the elements as given in the inset. It is seen from TEM image of a hexagon with a lateral dimension of ∼120nm.
Figure 4b that S2− anions form a distorted cubic closed packing Figure 5b shows the fast Fourier transform (FFT) pattern of a
of a sublattice, within the octahedra are fully occupied; single hexagon similar to that shown in Figure 5a, exhibiting
however, one-third of tetrahedral sites remains unoccupied, three rings, each consisting of six diffraction spots. These rings
forming negatively charged vacant sites (VIn), which is marked are ascribed to the {110} planes with d = 0.33 nm, {300}
by a blue cloud. The VIn centers are ordered along the 41 screw planes with d = 0.19 nm, and {220} planes with d = 0.16 nm.
axis, parallel to the z axis, giving rise to a small distortion from The observed diffraction planes in the FFT pattern are
the conventional cubic symmetry of the regular spinel compatible with those shown in the X-ray diffraction as shown
structure, called a defect spinel. The distortions lead to the in Figure 2. Figure 5c depicts the same FFT as in Figure 5b,
formation of a tetragonal structure of β-In2S3 with the point but it includes additional analysis done by stretching bars
group D4h19 and the space group I41/amd. The discussed between adjacent diffraction points, creating the sixfold
distortions also induce variability in the In−S bond lengths. symmetry frames (color coded with the FFT rings marked in
Figure 4c presents the three different types of coordination, Figure 5b). It is seen from the frames that the diffraction points
displayed with a color code to match the polyhedrons in Figure of {110} and {220} (green and blue frames, respectively)
4a, illustrating from left to right: an octahedron with two overlay, and both are twisted with respect to the {300} points
different In−S bond lengths of 2.660 and 2.537 Å, a (yellow frame) by α = 30°. An angle of 30° is consistent with
tetrahedron with two In−S bond lengths of 2.465 and 2.467 the tilt angle between the m- and a-planes, as shown
Å, and an additional octahedron with four different In−S bond schematically in Figure 3d,e for γ-In2S3. It is important to
lengths of 2.540, 2.629, 2.643, and 2.695 Å. Figure 4d shows note that FFT patterns of the dodecagonal NPLs exhibit
connectivity among the octahedron/tetrahedron units of a identical crystallographic diffractions as shown in Figure 5c,d;
certain fraction in the unit cell, when neighboring units are hence, the dodecagon morphology is also identified with the γ-
connected either through a corner bridge along the [001] In2S3 phase, in agreement with the X-ray observations. Figure
1788 DOI: 10.1021/acs.chemmater.9b00013
Chem. Mater. 2019, 31, 1784−1793
Chemistry of Materials Article

spacing values are marked on Figure 5j, showing compatibility


with the diffractions measured by an X-ray method. Figure 5k
exemplifies a TEM image of a superposition between two or
more β-In2S3 NPLs, which formed a Moiré pattern. The
corresponding FFT pattern is shown in Figure 5l, displaying
two rings, the inner ring belongs to {116} and the outermost
ring belongs to {109} planes with d-spacing values of 0.38 and
0.32 nm, respectively. The diffractions observed in Figure 5l
confirm the first- and second-order reflection peaks in the X-
ray diffraction pattern of the β-In2S3 phase.
Kinetic and Thermodynamic Effects in the Growth
Mechanism of In2S3 NPLs. Keeping in mind the correlation
between the observed polymorphs and the crystal structures of
the In2S3 phases, the discussion in this section reviews a
plausible growth mechanism of each polymorph provided with
Figure 5. (a) TEM image of hexagon of γ-In2S3. (b) FFT pattern of a qualitative analysis. Figure 6a displays the NPLs formation
hexagonal γ-In2S3 NPL, taken from the [001] zone axis (ZA) from the
image in (a). Diffraction spots are gathered into three circumferences
emphasized by different colors, attributed to {110} (green), {300}
(yellow), and {220} (blue) planes, and the corresponding d-spacing
values are shown on the pattern. (c) FFT pattern of the image in (a),
marked by sixfold symmetry frames, showing α = 30° rotation angle
between two adjacent frames. (d) FFT pattern of two overlapping
NPLs, exhibiting a relative rotation angle of 5°. (e) Derived image
based on the FFT pattern in (d), representing the superimposition of
two or more rotated NPLs forming [ABAB] alternating atomic
arrangement. (f) TEM image of two overlapping NPLs, generating a
mutual Moiré structure (see scheme at the inset). (g) FFT image of a
Moiré pattern zone exhibits a 2−3° rotation along the {110} planes.
(h) TEM image of truncated triangular morphology of the β-In2S3
phase. (i) HR-TEM image of a representative β-In2S3 NPL and the
sharp FFT diffraction pattern (inset). (j) Blowup image of (i)
Figure 6. (a) Scheme of NPL formation mechanism including circle
represents three planes {204}, {408}, and {219} as marked in the
frames and bar above, indicating from left to right, the formation
panel. (k) TEM image including several overlapping with the β-
stages, starting from the initial reactants, generation of the In thiolate
structure, forming a Moiré pattern (see dashed frame). (l) FFT image
single precursor, formation of monomers, and growth into two
taken from the Moiré pattern area in panel (k), showing two rings of
different polymorph shapes. (b) Scheme presentation of different
the planes {109} and {116} with the corresponding d-spacing values
growth directions, leading to the formation of nanoprism, hexagon,
of 0.32nm (pink) and 0.38nm (yellow), respectively.
and dodecagon shapes (top to bottom). Control of the final shapes is
driven by the rate of monomer deposition, indicated by Ri (i = 1, 2,
5d displays an FFT pattern of two superimposed NPL 3). A potential gradient scale is shown on the right. The bar below
hexagons with a 5° rotation between them. Figure 5e shows designates symbols of chemical constituents of (a) and (b).
the derived image of the FFT shown in Figure 5d, indicating
an alternating arrangement of [ABAB]. Figure 5f displays an
image of two or more NPL hexagons overlapping along the stages from left to right, starting from the preliminary
[0001] direction, experiencing a crystalline mismatch, leading materials, generating the single precursor, formation of
to the formation of a Moiré pattern (see scheme in inset). A monomers, and the growth into two different polymorph
Moiré pattern becomes feasible only in regions in which the shapes. The bar below designates symbols of chemical
NPL surfaces accommodated a low concentration of constituents. The growth process after nucleation is analyzed
surfactants because of the sample preparation procedure (see in depth. Although the monomer formation and the nucleation
methods section), allowing a mutual orientation between stage are determined solely by thermodynamic considerations,
adjacent platelets. Figure 5g presents the FFT of a Moiré the growth stage is dictated by a kinetically controlled reaction,
pattern zone in Figure 5f, showing a 2−3° rotation angle along combined with thermodynamic effects.
the {110} planes. The hexagonal morphology of the γ-In2S3 The nucleation stage generated polyhedron-shaped seeds
NPLs and their overlapping phenomenon are further displayed with some defined faceting. Anisotropy of a polyhedron
by HAADF-STEM measurements (see Supporting Informa- induces growth along specific crystallographic directions and
tion, Figure S2). An acquired HAADF line that is stretched eventually leads to the formation of an anisotropic
over a thin γ-In2S3 NPL shows a staircase line profile, which morphology. The growth on top of a seed is dictated by the
indicates the existence of several adjacent overlapping NPLs chemical potential of the bulk solution, which is dependent on
(see Figure S2). the concentration of [In2S3] monomers, and consequently, the
Figure 5h displays a TEM image of the β-In2S3 exposing the kinetic diffusion rate of the monomers toward nucleus faceting
truncated triangular shape as discussed in Figure 1, and the can be affected. A facet surface energy depends on the atomic
corresponding high-resolution image is shown in Figure 5i. compactness and chemical reactivity of surface atoms. Figure
The FFT image is seen in the inset of Figure 5i, whereas a 6b depicts a scheme presenting a few different growth
close-up of a small region of the image is shown in Figure 5j. A directions. An exemplary polyhedron with an HCP crystal
few diffraction planes, {204}, {408}, and {219}, and their d- structure is shown at the left side of the scheme; the {0001}
1789 DOI: 10.1021/acs.chemmater.9b00013
Chem. Mater. 2019, 31, 1784−1793
Chemistry of Materials Article

Figure 7. (a) Raman spectra of γ-In2S3 and β-In2S3 phases at room temperature. The black curve corresponds to the experimental spectrum of
hexagon/dodecagon NPLs, whereas the green curves are associated with the spectrum of truncated triangular-shaped NPLs. The green and orange
peaks are the Lorentzian best fit of each spectrum. The numbers of Raman modes are marked by ascending order (left to right). (b) Absorbance
spectra of NPLs (solid curves), prepared at 230 °C with duration times as indicated in the inset. The absorbance spectrum is accompanied by
extrapolation lines (dashed curves), indicating the Tauc fit to each curve. (c) Absorbance spectra of γ-In2S3 and β-In2S3 after storage time as
indicated in the legend. (d) Representation of PL spectra (solid curves) with the same synthesis time intervals as given in (b). The PL spectra are
associated with Gaussian best fit of each spectrum.

Table 2. Summary of the Number of Vibration Modes and Their Corresponding Frequencya
number of vibration frequency number of vibration frequency number of vibration frequency
phase mode (cm−1) mode (cm−1) phase mode (cm−1)
β-In2S3 1 60.9 9 201.4 γ-In2S3 1 117.8
2 73.8 10 220.5 2 125.1
3 92.1 11 246.9 3 132.5
4 103.1 12 269.7 4 181.7
5 117.8 13 299.1 5 192.6
6 136.1 14 309.9 6 205.3
7 165.3 15 329.6 7 293.7
8 182.8 16 364.4
a
Data derived from the Raman spectra shown in Figure 7a.

facets have the highest atomic density per unit area compared the potential gradient (ΔP) between the bulk potential (Pbulk)
to the normal directions, when the c-plane is terminated by and a facet potential (Pa,m). Schematic scaling of a potential
In+3 cations (In rich), as also shown in the scheme of Figure 3f. gradient is shown at the right side of Figure 6b. Increasing the
Hence, the In2S3 NPL surface contains a large number of growth temperature to 250 °C brings about an initial
undercoordinated cations, bounded to three anions only rather monomer-rich solution with a high chemical potential,
than six, as in a bulk position. The undercoordinated surface increasing the gradient between the bulk solution and the
can probably be passivated by organic ligands, such as thiols NPL a- and m-planes, namely, ΔPbulk→a,m ≫ 0. However, as
and amines (see low bar), which bind to an In rich surface via a defined above, the deposition rate is governed by the diffusion
single or a multiple bonding,45 and accordingly, a stabilization rate per unit area of a facet, and because m-planes have smaller
of the exterior surface energy. The capping ligands screen the area than the a-planes; thus, under ΔPbulk→a,m ≫ 0, a
surface from a flux of diffusing monomers, reducing the rate of preference growth normal to the m-planes occurs to
the deposition, and avoiding any further growth along the compensate for their larger surface strain. This preference
<0001> direction. Rate of deposition is defined in following leads to the formation of hexagonally shaped γ-In2S3 NPLs. At
texts as the rate of diffusion per unit area of a facet and a temperature of 230 °C, there is a relatively low [In2S3]
indicated by Ri (i = 1, 2, 3), so the growth along the <0001> monomer concentration in the bulk solution, and although
direction is defined as R1 (see Figure 6b). Alternatively, the ΔPbulk→a > ΔPbulk→m > 0, the overall deposition rate is nearly
growth develops along the lateral dimensions, when monomer equivalent because of the difference in the plane surface area.
units [In2S3] are deposited either on the {101̅0} m-planes or Such a uniform deposition rate advances for the formation of a
on the {112̅0} a-planes, with deposition rates R2 and R3, dodecagon shape of the NPLs. Thus, Figure 6b elucidates
respectively. Under nonequilibrium kinetic growth with a high three routes: option (I) considers R1 ≫ R2, R3, a case that is
flux of monomers, preferential growth takes place along the not relevant to our study because of passivation of the c-planes;
kinetically most favorable direction with low activation energy option (II) considers R1 ≪ R3 < R2, leading to a dominant
barriers. The diffusion of the free monomer units is driven by development of a hexagon shape; and option (III) considers R1
1790 DOI: 10.1021/acs.chemmater.9b00013
Chem. Mater. 2019, 31, 1784−1793
Chemistry of Materials Article

≪ R2 ≈ R3, directing to a development of a dodecagon shape. there is an initiation of phase transformation, whereas at 128
Therefore, a minor change in the temperature (ΔT = 20 °C) min, the NPLs transformed into truncated triangular-shaped
can give rise to a new growth pathway, because the monomer NPLs. The absorbance curves are accompanied by extrap-
formation rate and the NC growth rate depend exponentially olation lines, indicating the Tauc fit to each curve. Overall, the
on the temperature.46 intersection with the x-axis designates the energy band gap of
Optical Measurements. The formation of the γ-In2S3 and each sample, ranging between 3.5 and 3.7 eV. A literature
β-In2S3 phases is further confirmed by recording their Raman report supplied information about the β-In2S3 bulk energy
spectra at room temperature, as shown in Figure 7a. The band gap between 2.0 and 2.4 eV, although to the best of our
bottom (black) curve corresponds to the experimental knowledge, the bulk band gap of γ-In2S3 has not been reported.
spectrum of hexagon/dodecagon NPLs, whereas the green In any event, band-edge optical transitions shifted by ∼1.5eV
curves are the Lorentzian best fit. The black curve consists of from bulk values can be related to the quantum confinements
seven Raman modes, marked by numbers with an ascending induced by an NPL thickness of ∼1.8nm, considerably smaller
order from left to right, whose corresponding frequencies are than the bulk exciton Bohr radius (33.8 nm). The chemical
summarized in Table 2. The top (red) curve is associated with solidity of the NPLs against air and moisture was examined by
the spectrum of truncated triangular shaped NPLs, and the recording their absorbance after couple of months, as shown in
accompanied orange curves are the Lorentzian best fit of 16 Figure 7c. It is seen from the figure that the features in the
different frequency modes, as listed in Table 2. The analysis of absorbance curves of γ-In2S3 and β-In2S3 phases remained
the Raman spectra was done by comparison with anticipated intact after air exposure over 3 months, corroborating the
modes derived from symmetry considerations. chemical stability over time.
The γ-In2S3 phase has the R3̅C space group and D3d point Figure 7d represents the PL spectra (solid curves) with the
group symmetry; the latter is sufficient for the analysis of same synthesis time intervals as given in Figure 7b. The spectra
vibration modes of the structures. The irreducible representa- are composed from a dominant band and another very weak
tion of the D3d point group is given as Γ = 2A1g + 2A1u + 3A2g + one at deeper energies (see additional details in the Supporting
2
A2u + 4Eu + 5Eg, where 2A1g + 5Eg are the Raman vibration Information, Figure S4). The dominant PL bands were fitted
active modes. The prediction of seven modes is compatible to two Gaussian curves (dashed lines). The small Stokes shift
with the appearance of the same number of resonance modes of the high-energy Gaussian from that of the band gap
in the Raman spectrum of the hexagon/dodecagon above a designates its association with an exciton recombination. The
frequency of 100 cm−1, hence corroborating the association of deep energy broad band is a characteristic of donor−acceptor
those morphologies with the γ-In2S3 phase. The Raman active or defect-to-band recombination emission. The association of
band below 100 cm−1 was identified (see Supporting vacancies in the recombination process was also suggested
Information, Figure S3) by a remnant of oleylamine ligands, before,48 proposing involvement of indium at interstitial
used during the synthesis, and in typically, is difficult to positions, and occupation substitution of sulfur sites with
remove. oxygen atoms. Indeed, the energy dispersive X-ray spectros-
The β-In2S3 phase has the I41/amd space group and the copy (EDS), designated the existence of a small percent of
D4h19 point group, used for predicting the following infrared oxygen in the examined In2S3 phases (see Supporting
and Raman active modes at the center of the Brillouin zone: Information, Figure S5). The assignment of the mid-energy
ΓIR = 10A2u + 16Eu and ΓRaman = 9A1g + 9B1g + 4B2g + 14Eg.47 band can be correlated with a bound exciton to a defect center.
Although the prediction indicates possible occurrence of 36 Obviously, further investigations are required to identify the
modes, the experimental observation exhibits 17 modes. emission components.


Previous Raman studies of the β-In2S3 phase designated a
similar spectrum to our observations of the Raman modes,47 CONCLUSIONS AND OUTLOOK
compatible with part of the theoretical prediction. It is possible
to conclude that the Raman spectrum given by the red curve is In summary, the present work extensively explored the growth
associated with the β-In2S3 structure with a truncated mechanism, which govern the formation of chemically and
triangular morphology. It is useful to note that the red curve structurally stable In2S3 semiconductor NPLs at room
does not include a resonance related to ligand impurities, and temperature. The work focused on formation of various
it is also important to mention that the frequencies of both the morphologies of two different phases of In2S3: trigonal γ-In2S3
γ- and β-In2S3 Raman vibrations might deviate slightly from a and defect spinel β- In2S3. Growth in the two phases was based
theoretical prediction of a bulk material because of the on colloidal procedures starting from single precursors and the
existence of a quantum size and surface effects. existence of ligands that templating the following morpholo-
Optical transitions were observed in the absorbance and gies: γ-hexagon, γ-dodecagon, and β-truncated triangular. The
photoluminescence (PL) spectra as shown in Figure 7b−d. structural and optical observations presented in this study
Figure 7b displays the absorbance (dashed curves) of NPLs exposed the correlation between the crystal structure, reaction
that were prepared at 230 °C with duration time as indicated conditions, and the morphology of the NPLs. In particular, this
in the inset. Because of the relatively slow reaction rate, it was work revealed the mechanism for a structural phase transition
feasible to follow intermediate stages, one of which exhibits a from the γ-In2S3 to the β-In2S3 phase. The proposed structures
dodecagon shape. It is seen from the spectra that a exhibit an efficient exciton recombination in the blue-to-green
distinguishable feature built up in the absorbance curve spectral range, with an additional minor contribution from a
already after 8 min, and it is shifted to low energies until a defect radiative recombination at deep energies. The reported
reaction time of 32 min. However, the trend is distracted after analysis and observations of this work pave the way for the
64 min. In the following stage, after 128 min, it is possible to implementation of In2S3 phases in diverse areas such as solar
see a completely different absorbance curve. It is likely to that cells, photocatalysis, doping, and more. In addition, the NPLs
the reaction of 32 min produced the dodecagons; after 64 min, discussed in this paper are based on chemical elements with
1791 DOI: 10.1021/acs.chemmater.9b00013
Chem. Mater. 2019, 31, 1784−1793
Chemistry of Materials Article

low toxicity, so they would be attractive for various (10) Kim, W.-T.; Kim, C.-D. Optical Energy Gaps of β-In2S3 Thin
applications involving environmental considerations. Films Grown by Spray Pyrolysis. J. Appl. Phys. 1986, 60, 2631−2633.


(11) Xiong, X.; Zhang, Q.; Gan, L.; Zhou, X.; Xing, X.; Li, H.; Zhai,
ASSOCIATED CONTENT T. Geometry Dependent Photoconductivity of In2S3 kinks Synthe-
sized by Kinetically Controlled Thermal Deposition. Nano Res. 2016,
*
S Supporting Information 9, 3848−3857.
The Supporting Information is available free of charge on the (12) Pistor, P.; Á lvarez, J. M. M.; León, M.; Di Michiel, M.; Schorr,
ACS Publications website at DOI: 10.1021/acs.chemma- S.; Klenk, R.; Lehmann, S. Structure Reinvestigation of α-, β- And γ-
ter.9b00013. In2S3. Acta Crystallogr. Sect. B Struct. Sci. Cryst. Eng. Mater. 2016, 72,
410−415.
Sided TEM images of γ-In2S3 NPLs, HAADF-STEM (13) Range, K.-J.; Zabel, M. ε-In2S3, a high pressure modification
images of γ-In2S3 NPLs, Raman spectra of OLAm and with corundum type structure. Z. Naturforsch. B. 1978, 33, 463−464.
DDT ligands, representative PL spectra of In2S3 NPLs, (14) Steigmann, G. A.; Sutherland, H. H.; Goodyear, J. The Crystal
and EDS elemental analysis (PDF) Structure of β-In2S3. Acta Crystallogr. 1965, 19, 967−971.


(15) Li, X.; Magnuson, C. W.; Venugopal, A.; Tromp, R. M.;
AUTHOR INFORMATION Hannon, J. B.; Vogel, E. M.; Colombo, L.; Ruoff, R. S. Large-Area
Graphene Single Crystals Grown by Low-Pressure Chemical Vapor
Corresponding Author Deposition of Methane on Copper. J. Am. Chem. Soc. 2011, 133,
*E-mail: ssefrat@technion.ac.il. 2816−2819.
ORCID (16) Hao, Y.; Meng, G.; Ye, C.; Zhang, X.; Zhang, L. Kinetics-
Driven Growth of Orthogonally Branched Single-Crystalline
Efrat Lifshitz: 0000-0001-7387-7821 Magnesium Oxide Nanostructures. J. Phys. Chem. B 2005, 109,
Author Contributions 11204−11208.
F.H. carried out experiments, performed characterizations, and (17) Khadka, D. B.; Kim, S.; Kim, J. A Nonvacuum Approach for
analyzed the data. E.L. supervised the research. F.H. and E.L. Fabrication of Cu2ZnSnSe4/In2S3 Thin Film Solar Cell and
wrote the manuscript. Optoelectronic Characterization. J. Phys. Chem. C 2015, 119,
12226−12235.
Notes (18) Sterner, J.; Malmström, J.; Stolt, L. Study on ALD In2S3
The authors declare no competing financial interest.


/Cu(In,Ga)Se2 Interface Formation. Prog. Photovoltaics Res. Appl.
2005, 13, 179−193.
ACKNOWLEDGMENTS (19) Naghavi, N.; Abou-Ras, D.; Allsop, N.; Barreau, N.; Bücheler,
The authors acknowledge the financial support from the Israel S.; Ennaoui, A.; Fischer, C.-H.; Guillen, C.; Hariskos, D.; Herrero, J.;
Council for Higher−Focal Area Technology (No. 872967), the et al. Buffer Layers and Transparent Conducting Oxides for
Chalcopyrite Cu(In,Ga)(S,Se)2 Based Thin Film Photovoltaics:
Volkswagen Stiftung (No.88116), the Israel Science Founda-
Present Status and Current Developments. Prog. Photovoltaics Res.
tion (No.914/15), the Israel Science Foundation Bikura Appl. 2010, 18, 411−433.
Program (No.1508/14), and the Neubauer Family Founda- (20) Ho, C.-H.; Wang, Y.-P. NIR and UV Enhanced Photon
tion.


Detector Made by Diindium Trichalcogenides. Opt. Mater. Express
2013, 3, 1420.
REFERENCES (21) Batabyal, S. K.; Lu, S. E.; Vittal, J. J. Synthesis, Characterization,
(1) Das, P.; Samantaray, S.; Rout, G. R. Studies on Cadmium and Photocatalytic Properties of In2S3, ZnIn2S4, and CdIn2S4
Toxicity in Plants: A Review. Environ. Pollut. 1997, 98, 29−36. Nanocrystals. Cryst. Growth Des. 2016, 16, 2231−2238.
(2) Godt, J.; Scheidig, F.; Grosse-Siestrup, C.; Esche, V.; (22) Sharma, R. K.; Chouryal, Y. N.; Chaudhari, S.; Saravanakumar,
Brandenburg, P.; Reich, A.; Groneberg, D. A. The Toxicity of J.; Dey, S. R.; Ghosh, P. Adsorption-Driven Catalytic and Photo-
Cadmium and Resulting Hazards for Human Health. J. Occup. Med. catalytic Activity of Phase Tuned In2S3 Nanocrystals Synthesized via
Toxicol. 2006, 1, 22. Ionic Liquids. ACS Appl. Mater. Interfaces 2017, 9, 11651−11661.
(3) Werner, T. T.; Mudd, G. M.; Jowitt, S. M. Indium: Key Issues in (23) Yan, T.; Wu, T.; Zhang, Y.; Sun, M.; Wang, X.; Wei, Q.; Du, B.
Assessing Mineral Resources and Long-Term Supply from Recycling. Fabrication of In2S3/Zn2GeO4 composite Photocatalyst for Degrada-
Appl. Earth Sci. 2015, 124, 213−226. tion of Acetaminophen under Visible Light. J. Colloid Interface Sci.
(4) Mane, R. S.; Lokhande, C. D. Studies on Structural, Optical and 2017, 506, 197−206.
Electrical Properties of Indium Sulfide Thin Films. Mater. Chem. Phys. (24) Xie, X.; Shen, G. Single-Crystalline In2S3 Nanowire-Based
2003, 78, 15−17. Flexible Visible-Light Photodetectors with an Ultra-High Photo-
(5) Hara, K.; Sayama, K.; Arakawa, H. Semiconductor-Sensitized response. Nanoscale 2015, 7, 5046−5052.
Solar Cells Based on Nanocrystalline In2S3/In2O3 Thin Film (25) Cansizoglu, H.; Cansizoglu, M. F.; Watanabe, F.; Karabacak, T.
Electrodes. Sol. Energy Mater. Sol. Cells 2000, 62, 441−447. Enhanced Photocurrent and Dynamic Response in Vertically Aligned
(6) Li, M.; Tu, X.; Su, Y.; Lu, J.; Hu, J.; Cai, B.; Zhou, Z.; Yang, Z.; In2S3/Ag Core/shell Nanorod Array Photoconductive Devices. ACS
Zhang, Y. Controlled Growth of Vertically Aligned Ultrathin In2S3 Appl. Mater. Interfaces 2014, 6, 8673−8682.
nanosheet Arrays for Photoelectrochemical Water Splitting. Nanoscale (26) Chen, W.; Bovin, J.-O.; Joly, A. G.; Wang, S.; Su, F.; Li, G. Full-
2018, 10, 1153−1161. Color Emission from In2S3 and In2S3:Eu3+ Nanoparticles. J. Phys.
(7) Sumi, R.; Warrier, A. R.; Vijayan, C. Visible-Light Driven Chem. B 2004, 108, 11927−11934.
Photocatalytic Activity of β-Indium Sulfide (In2S3) Quantum Dots (27) Ye, F.; Du, G.; Jiang, Z.; Zhong, Y.; Wang, X.; Cao, Q.; Jiang, J.
Embedded in Nafion Matrix. J. Phys. D. Appl. Phys. 2014, 47, 105103. Z. Facile and Rapid Synthesis of RGO-In2S3 Composites with
(8) Bhira, L.; Essaidi, H.; Belgacem, S.; Couturier, G.; Salardenne, J.; Enhanced Cyclability and High Capacity for Lithium Storage.
Barreaux, N.; Bernede, J. C. Structural and Photoelectrical Properties Nanoscale 2012, 4, 7354−7357.
of Sprayed β-In2S3 Thin Films. Phys. Status Solidi A 2000, 181, 427− (28) Xue, B.; Xu, F.; Wang, B.; Dong, A. Shape-Controlled Synthesis
435. of β-In2S3 Nanocrystals and Their Lithium Storage Properties.
(9) Diehl, R.; Nitsche, R. Vapour Growth of Three In2 S 3 CrystEngComm 2016, 18, 250−256.
Modifications by Iodine Transport. J. Cryst. Growth 1975, 28, 306− (29) Mugesh, T.; Premkumar, T.; Chitravel, T. Engineering the
310. Structural and Photoluminescence Properties of β-In2S3 Nano-

1792 DOI: 10.1021/acs.chemmater.9b00013


Chem. Mater. 2019, 31, 1784−1793
Chemistry of Materials Article

powders for Luminescence Application. Nanocompos. Mater. 2012, 50,


10525−10527.
(30) Afzaal, M.; Malik, M. A.; O’Brien, P. Indium Sulfide Nanorods
from Single-Source Precursor. Chem. Commun. 2004, 4, 334−335.
(31) Huang, W.; Gan, L.; Yang, H.; Zhou, N.; Wang, R.; Wu, W.; Li,
H.; Ma, Y.; Zeng, H.; Zhai, T. Controlled Synthesis of Ultrathin 2D β-
In2S3 with Broadband Photoresponse by Chemical Vapor Deposition.
Adv. Funct. Mater. 2017, 27, 1702448.
(32) Park, K. H.; Jang, K.; Son, S. U. Synthesis, Optical Properties,
and Self-Assembly of Ultrathin Hexagonal In2S3 nanoplates. Angew.
Chemie - Int. Ed. 2006, 45, 4608−4612.
(33) Datta, A.; Mukherjee, D.; Witanachchi, S.; Mukherjee, P. Low
Temperature Synthesis, Optical and Photoconductance Properties of
Nearly Monodisperse Thin In2S3 Nanoplatelets. RSC Adv. 2013, 3,
141−147.
(34) Yang, W.; Liu, B.; Cho, Y.; Yang, B.; Dierre, B.; Sekiguchi, T.;
Jiang, X. Ternary In2S3/In2O3 heterostructures and Their Cathodo-
luminescence. RSC Adv. 2016, 6, 51089−51095.
(35) Gqoba, S.; Airo, M.; Ntholeng, N.; Machogo, L.; Sithole, R.;
Moloto, M. J.; van Wyk, J.; Moloto, N. Evolution of In2S3 Nanoplates
with Time. Mater. Today Proc. 2015, 2, 3901−3908.
(36) McCarthy, R. F.; Schaller, R. D.; Gosztola, D. J.; Wiederrecht,
G. P.; Martinson, A. B. F. Photoexcited Carrier Dynamics of In2S3
Thin Films. J. Phys. Chem. Lett. 2015, 6, 2554−2561.
(37) Galarza Gutiérrez, U.; de Albor Aguilera, M. L.; Hernández
Vasquez, C.; Flores Márquez, J. M.; González Trujillo, M. A.; Jiménez
Olarte, D.; Aguilar Hernández, J. R.; Remolina Millán, A. Structural
and Optoelectronic Properties of β-In2S3 Thin Films to Be Applied on
Cadmium Reduced Solar Cells. Phys. Status Solidi A 2018, 215,
1700428.
(38) Ghosh, S.; Saha, M.; Ashok, V. D.; Chatterjee, A.; De, S. K.
Excitation Dependent Multicolor Emission and Photoconductivity of
Mn, Cu Doped In2S3 Monodisperse Quantum Dots. Nanotechnology
2016, 27, 155708.
(39) Chen, B.; Chang, S.; Li, D.; Chen, L.; Wang, Y.; Chen, T.; Zou,
B.; Zhong, H.; Rogach, A. L. Template Synthesis of CuInS2
Nanocrystals from In2S3 Nanoplates and Their Application as
Counter Electrodes in Dye-Sensitized Solar Cells. Chem. Mater.
2015, 27, 5949−5956.
(40) Feng, J.; Zhu, H.; Yang, X. A Controllable Growth-Doping
Approach to Synthesize Bright White-Light-Emitting Cd:In2S3
nanocrystals. Nanoscale 2013, 5, 6318−6322.
(41) Sarkar, S.; Leach, A. D. P.; Macdonald, J. E. Folded
Nanosheets: A New Mechanism for Nanodisk Formation. Chem.
Mater. 2016, 28, 4324−4330.
(42) Ungár, T. Microstructural Parameters from X-Ray Diffraction
Peak Broadening. Scr. Mater. 2004, 51, 777−781.
(43) King, G. S. D. The Space Group of β-In2S3. Acta Crystallogr.
1962, 15, 512−512.
(44) Ho, C. H.; Wang, Y. P.; Chan, C. H.; Huang, Y. S.; Li, C. H.
Temperature-Dependent Photoconductivity in β-In2S3 Single Crys-
tals. J. Appl. Phys. 2010, 108, No. 043518.
(45) Jun, Y.-w.; Lee, J.-H.; Choi, J.-s.; Cheon, J. Symmetry-
Controlled Colloidal Nanocrystals: Nonhydrolytic Chemical Syn-
thesis and Shape Determining Parameters. J. Phys. Chem. B 2005, 109,
14795−14806.
(46) van Embden, J.; Chesman, A. S. R.; Jasieniak, J. J. The Heat-up
Synthesis of Colloidal Nanocrystals. Chem. Mater. 2015, 27, 2246−
2285.
(47) Kambas, K.; Spyridelis, J.; Balkanski, M. Far Infrared and
Raman Optical Study of α- and β-In2S3 Compounds. Phys. Status
Solidi B 1981, 105, 291−296.
(48) Jayakrishnan, R.; Sebastian, T.; Sudha Kartha, C.; Vijayakumar,
K. P. Effect of Defect Bands in β-In2S3 thin Films. J. Appl. Phys. 2012,
111, No. 093714.

1793 DOI: 10.1021/acs.chemmater.9b00013


Chem. Mater. 2019, 31, 1784−1793

You might also like