0% found this document useful (0 votes)
88 views9 pages

N.BANSAL, GC MOHANTA, K SiNGH, CERAM INT. 43 (2017) 7193-7201

N.BANSAL, GC MOHANTA, K SiNGH, CERAM INT. 43 (2017) 7193-7201

Uploaded by

Neetu Bansal
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
88 views9 pages

N.BANSAL, GC MOHANTA, K SiNGH, CERAM INT. 43 (2017) 7193-7201

N.BANSAL, GC MOHANTA, K SiNGH, CERAM INT. 43 (2017) 7193-7201

Uploaded by

Neetu Bansal
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 9

Ceramics International 43 (2017) 7193–7201

Contents lists available at ScienceDirect

Ceramics International
journal homepage: www.elsevier.com/locate/ceramint

Effect of Mn2+ and Cu2+ co-doping on structural and luminescent properties MARK
of ZnS nanoparticles

Neetu Bansala, Girish C. Mohantab, K. Singha,
a
School of Physics & Materials Science, Thapar University, Patiala 147004, Punjab, India
b
Central Scientific Instruments Organisation (CSIR-CSIO), Sector 30, Chandigarh 160030, India

A R T I C L E I N F O A BS T RAC T

Keywords: Undoped and transition metal (Cu, Mn, Cu:Mn) doped ZnS nanoparticles are synthesized by chemical co-
Co-doping precipitation method via an aqueous synthesis route. Synthesized samples are characterized by various
Hexagonal techniques for their structural and optical properties. Crystallite size obtained from X-Ray Diffraction (XRD)
ZnS is 1.68, 1.87, 1.50, 1.42 nm for undoped, Cu, Mn, Cu:Mn doped ZnS nanoparticles. The XRD, High Resolution
Fluorescent
Transmission Electron Microscopy, and Selected Area Electron Diffraction confirm the evolution of stable
Nanoparticles, biological labels
hexagonal phase of ZnS nanoparticles at low temperature. Energy Dispersive Spectroscopy confirms the doping
of nanoparticles. Blue shift in UV absorbance shows the increase in optical bandgap with decrease in particle
size. The Photoluminescence studies exhibit blue, yellow and red emission in visible region. Surface
functionalization of nanoparticles is confirmed from Fourier Transform Infra Red spectroscopy. The present
samples are tunable in wider range of emission and are prospective candidates for biological labels due to their
fluorescent properties.

1. Introduction mediate bands between conduction and valence bands of ZnS, which
may lead to the change in structure of the ZnS nanoparticles.
Nanostructured semiconductors like CdS, CdSe, ZnS etc. exhibit In most of the cases, nanoparticles (NPs) are being synthesized
wide applications in the field of optoelectronics, electronics, bio- using organic solvents at high temperature (300–350 °C) [10]. The NPs
medical applications [1–4], etc. These materials can be widely tuned so prepared have their surface covered by organic ligands, which makes
by doping and processing parameters [5]. In these materials, ZnS is it water insoluble. As a result these could not be used effectively for
non-toxic having excellent electroluminescent properties. Inorganic biological applications, which invariably involve reactions in aqueous
nanoparticles like ZnS have better photostability than organic dyes medium. Thus, attempts are being made to synthesize water soluble
which are being used in bio-imaging. However, the optical band gap of NPs using different synthesis techniques [11]. One of the technique is a
ZnS lies in the UV- region. The optical band gap can be tailored by simple ligand exchange step, which replaces organic ligands with that
selecting proper dopants and their chemical nature, along with reduced containing functional groups (-COO, -NH2, -OH, -SH etc.). Recently,
particle size. Particularly, nano ZnS show change in optical band, which efforts are made towards preparing NPs directly into aqueous medium
fall in visible region. Moreover, hexagonal ZnS is found to exhibit to bypass the additional ligand exchange step required in earlier
better luminescent properties than cubic structured ZnS [6]. While protocols. In this technique, NPs are simultaneously surface functio-
wurtzite (hexagonal) phase of ZnS is a metastable phase, which forms nalized for specific biomolecular interactions. It can be achieved by
at 1020 °C for bulk ZnS, Sun et al. have obtained it at low temperature attaching biomolecules that can carry out specific interactions like,
of 110 °C using microwave thermolysis [7]. Acharya et al. reported antibody-antigen interaction, ligand-receptor interaction, Avidin/
ethylenediamine-mediated phase transformation of ZnS at 170 °C [8]. Streptavidin–Biotin interaction or enzymes that can catalyze specific
Zhang et al. have reported that hexagonal phase ZnS nanoparticles with reactions like catalase, glucose oxidase etc. [4]. Therefore, preparation
small sizes ( < 3 nm) are thermodynamically more stable than sphaler- of NPs in aqueous media, along with the attachment of bio-molecules
ite phased ZnS nanoparticles [9]. Basically, the dopants create defects on the surface, rendered NPs water soluble as well as specifically
in the ZnS matrix due to size difference, and different valence states of functional. The objective of the present study is to synthesize water
dopants in base crystalline material. These defects form some inter- soluble NPs for bio-imaging applications with reduced optical band gap


Corresponding author.
E-mail address: [email protected] (K. Singh).

http://dx.doi.org/10.1016/j.ceramint.2017.03.007
Received 7 December 2016; Received in revised form 1 March 2017; Accepted 1 March 2017
Available online 06 March 2017
0272-8842/ © 2017 Elsevier Ltd and Techna Group S.r.l. All rights reserved.
N. Bansal et al. Ceramics International 43 (2017) 7193–7201

by using cost effective low temperature technique. In addition to this,


the selected dopants are isovalent transition metals. However, during
processing their oxidation states may change and create more dis-
ordering in ZnS.
In the present study, Cu, Mn- doped as well as Cu and Mn co-doped
ZnS nanoparticles are synthesized by aqueous synthesis via chemical
co-precipitation method, with varying the type of dopant ions. It will be
worthwhile to study the effect of two different oxidation state dopants
on various properties. The as-prepared samples are characterized using
various techniques to study the effect of processing parameters and
chemical nature of dopants on optical, structural and photolumines-
cence properties.

2. Materials and methods

2.1. Sample preparation


Fig. 1. XRD patterns obtained for undoped ZnS, ZnS:Cu, ZnS:Mn, ZnS:Mn:Cu QD.
Zinc acetate (dehydrated) (Zn(CH3COO)2), sodium sulfide (anhy-
drous) (Na2S), methanol (CH3OH), copper sulphate (heptahydrate)
(CuSO4·7H2O), manganese chloride (MnCl2), β- mercaptoethanol (HS-
C2H4OH) (ME), mercaptoacetic acid (HS-CH2COOH) were used to
synthesize the present samples. These chemicals are used without
further purification. Autoclaved de-ionized (di) water is used for
preparing aqueous solution of the chemicals.

2.2. Synthesis of nanoparticles

The synthesis of nanoparticles of undoped ZnS, Mn2+- doped ZnS,


Cu2+- doped ZnS and Mn2+and Cu2+ co-doped ZnS was carried out by
colloidal precipitation reaction method [12]. ZnS:Mn nanoparticles were
precipitated from a mixture of Zn(CH3COO)2 and MnCl2 with Na2S
aqueous solution. In a typical procedure, aqueous solution of zinc acetate
(1 M) and manganese chloride (0.1 M) were added to the autoclaved di-
water at temperature of 70 °C. N2 flow is provided through condenser for
maintaining inert atmosphere, along with vigorous stirring using mag-
netic stirrer. Then, 1 M ME was added to the mixture and kept for stirring Fig. 2. Gaussian fit peak obtained from XRD patterns of samples in the range 20–42°.
for 20 min. Finally, freshly prepared sodium sulfide was added to the
reaction mixture with the molar ratio 1:1 of Zn:S to obtain white 2.3. Characterization methods
precipitates. Reaction mixture was refluxed for 1 h (h) under same
conditions. The resulting precipitates were repeatedly washed (with di- The samples of NPs were characterized by X-ray diffraction (XRD)
water) and centrifuged at 15,000 rpm for 20 min. Final wash with technique using PANalytical's X′Pert Pro diffractometer irradiated with
methanol is performed to remove any organic part or any other impurities Cu-Kα radiation of wavelength λ=1.5406 Å in the diffraction angle
from the particles. After discarding the supernatant, the precipitates were range 2θ=15–80°. Elemental analysis of the samples was done using
kept overnight in tubes for drying under room temperature. Dried scanning electron microscope (FE-SEM (SE/N S-4300) equipped with
precipitates were grounded to obtain fine powder, and then stored for energy dispersive X-ray spectrometer. For SEM-EDS, methanol dis-
further characterizations. Similarly, undoped-ZnS NPs and Mn2+and persed samples were put on a cleaned silicon wafer. The structural
Cu2+-co doped NPs are synthesized using same procedure. However, no characteristics are supported by High Resolution-Transmission
manganese chloride was added for synthesis of undoped ZnS nanoparti- Electron Microscopy (HRTEM) and Selected Area Electron
cles, whereas aqueous solution of copper sulphate and manganese Diffraction (SAED) performed using FEG-TEM 300 kV (FEI/Technai
chloride were added for synthesis of co-doped ZnS: Mn:Cu NPs, in the G2, F30). Sample was taken in powder form placed over a copper grid
procedure as followed above. ZnS:Cu NPs were synthesized via an for HRTEM. The fluorescence of the samples was observed using UV
aqueous route reported by Corrado et al. [13], with some minor lamp (Syngene) working at 302 nm wavelength of the UV-light. FTIR
modifications. Mercaptoacetic acid (2 M) was added to an aqueous spectroscopy was done using FTIR spectrometer (Nicolet iS10) to
solution of 1 M Zn(CH3COO)2 and 3% CuSO4 solution. Then pH of the confirm the surface functionalization of the samples. Optical absorp-
reaction mixture was set to alkaline range (pH 9–11) using 1 M NaOH tion spectra of the samples were recorded using UV–Vis-NIR spectro-
solution and then kept for de-aeration under inert (N2) atmosphere for photometer (Cary 4000) to study the change in optical band gap of the
30 min along with vigorous stirring. Sodium sulfide (Na2S) (1 M), was samples with change in dopant ions. Photoluminescence (PL) emission
freshly prepared in de-aerated water to prevent oxidation, and then added spectra were taken using fluorescence spectrophotometer (Cary
quickly to the reaction mixture and kept for 15 min in nitrogen atmo- Eclipse) to study the emission wavelength of the samples. UV–Vis-
sphere with continuous stirring. When a clear solution was obtained, NIR spectroscopic studies are performed by dispersing the sample in
0.5 M zinc acetate was added again and kept to reflux for one hour. White autoclaved di-water.
precipitates were formed. For separation of precipitates from the solution,
a 2:1 ratio of (NP solution: ethanol) were prepared. This solution is
centrifuged at 15,000 rpm for 20 min and washed repeatedly with
ethanol. The pallets formed after discarding the supernatant were dried
in air for overnight and then stored.

7194
N. Bansal et al. Ceramics International 43 (2017) 7193–7201

Table 1
XRD and Optical studies of synthesized nanoparticles.

Sample's labels 2θ (degree) FWHM (degree) Crystallite size (nm) Optical bandgap (eV)

XRD UV–Vis TEM

ZnS 28.6773 5.3722 1.68 2.70 1.93 4.05


ZnS:Mn 29.1772 6.0392 1.50 2.42 – 4.15
ZnS:Cu 29.1478 4.8627 1.87 3.90 – 3.86
ZnS:Mn:Cu 29.3754 6.4416 1.42 1.87 1.41 4.70

Fig. 3. TEM micrographs of undoped ZnS nanoparticles (a) at 10 nm, (b) at 2 nm representing particle size, (c) at 2 nm representing D-spacing, (d) SAED pattern representing
diffraction planes.

3. Results and discussion diffraction angle has resulted in an overall compressive stress in the
base ZnS lattice, which may lead to lowering of D-spacing.
3.1. Structural analysis Broader and diffused peaks make it difficult to identify the crystal
structure of the present system from XRD, since there is very less
Fig. 1 shows the XRD pattern of the synthesized undoped, doped variation in XRD peaks of hexagonal and cubic zinc blende phase from
and co-doped ZnS samples. Three different broad and diffused peaks XRD. Moreover, with addition of dopants, the XRD peaks become more
are observed in all the samples centred around 28.85°, 47.70°, and diffused. Thus, HRTEM with SAED is used to analyse the structure and
56.52°, respectively. The XRD data is Gaussian fit to obtain the FWHM morphology of the samples. The representative SAED pattern of the
of the most intense peak (see Fig. 2). The FWHM was used to calculate pure ZnS and co-doped ZnS nanoparticles is shown in Figs. 3 and 4(d),
the crystallite size using Scherrer formula [14]. It is found to vary from respectively. Presence of clear rings in SAED pattern shows the
1.41 to 1.87 nm for different samples as given in Table 1. These formation of homogeneous nanoparticles of very small size. The D-
calculations confirm the decrease in the crystallite size of nanoparticles values and particle size is calculated using AxioVision Rel. 4.9.1
due to co-doping. XRD data of the samples rules out the formation of software and reported in Table 2. The obtained D-values match well
MnS and CuS phases. Thus, all the dopants might have occupied the with the hexagonal zinc sulfide structure as given in ICDD No. 01–080-
substitutional sites, since secondary phase is not observed in XRD. 0007. Thus hexagonal structure of nanoparticles is indexed with three
Fig. 2 shows the right shift in the most intense peak in the region of most intense peaks corresponding to (100), (110) and (103), which are
20°−40° with doping and co-doping. Diffraction angles for various broad and diffused as observed in XRD patterns as well as in SAED
samples corresponding to this peak are given in Table 1. The shift in pattern. The variation in D-value of the samples as compared with that

7195
N. Bansal et al. Ceramics International 43 (2017) 7193–7201

Fig. 4. TEM micrographs of Cu and Mn co-doped ZnS nanoparticles: (a) at 20 nm representing agglomeration, (b) at 5 nm representing particle size, (c) at 2 nm representing D-spacing,
(d) SAED pattern representing diffraction planes.

Table 2 nucleation theory, the reduction of particle size to nanoscale results in


Lattice parameters as calculated from TEM and XRD. the increase in the surface to volume ratio, which results in the
diminishing of volume free energy in comparison to the surface free
Sample's hkl D-value Lattice parameter Volume of unit
labels (nm) (nm) cell (×106
energy. This imbalance of surface to volume energies acts as a driving
pm3) force for the exothermicity of the lattice formation, which leads to
a c spontaneous transformation in phase of the nanoparticles [16]. Zhao
et al. [17] have reported low- temperature (150 °C) synthesis of
ZnS 100 0.333 0.400 0.602 83.41
110 0.205
hexagonal ZnS nanoparticles, with very diffused XRD peaks, whereas,
103 0.173 Porta et al. [18] have demonstrated the ZnS phase control at 140 °C
mediated using tetrabutylammonium hydroxide. Phase transition from
ZnS:Mn:Cu 100 0.326 0.384 0.612 78.02 cubic to hexagonal, after annealing the sample at 250 °C, has also been
110 0.203
reported by Kole et al. [6]. In the present study, Hexagonal ZnS phase
103 0.173
is formed for doped and co-doped nanoparticles at a lower temperature
of 70 °C due to decrease in size of nanoparticles.
reported in ICDD card is attributed to the aforementioned compressive Figs. 3 and 4(a) represents the agglomeration of the particles, which
stress which is observed due to the shift (100) peak in Fig. 2. This may might have occurred during co-precipitation synthesis. Figs. 3 and 4(b)
be because of the disorderness in the structure that arises due to represents the HRTEM micrographs with visibility of nano-sized
introduction of dual dopants of different ionic radii. Most probably, the particles at a scale of 2 nm and 5 nm, respectively. Figs. 3 and 4(c)
reaction kinetics has resulted in the change in oxidation state of Mn2+ shows the formation of fringes at 2 nm scale, where fringe width
to Mn4+, which causes Sulfide vacancies in the a-b plane. Alongside, corresponds to the D-spacing values for (100) plane. The particle size as
Mn4+ ion occupies the position along c-axis, creating a distortion in the observed from HRTEM is given in Table 2.
lattice structure. Similar phase transitions due to change in oxidation Along with this, the lattice parameters and volume of the unit cell is
state of the dopant atom is studied by Samita et al. [15]. The increase in also given, which is calculated for hexagonal structured undoped and
ionicity of the ZnS bonds due to presence of dopants stabilizes the co-doped ZnS nanoparticles [14]. It matches well in proximity to the
hexagonal structure as compared to the cubic structure. This transition values reported in ICDD No. 010–080-0007.
from cubic zinc blende to hexagonal structure imitates the cis-trans
transition due to change in bond length and bond angles. This lattice
distortion along with the above explained compressive stress, mediates 3.2. SEM-EDS analysis
a phase transition from cubic zinc blende to hexagonal structure, even
at a low temperature of 70 °C. Moreover, according to the classical Fig. 5 depicts the SEM micrographs of all the samples. The average
particle size is higher than that observed in XRD. It is clearly

7196
N. Bansal et al. Ceramics International 43 (2017) 7193–7201

Fig. 5. SEM micrographs of individual undoped and doped ZnS nanocrystals.

Fig. 6. SEM-ED Spectrum of undoped-ZnS, ZnS:Mn, ZnS:Cu and ZnS:Mn:Cu QDs.

manifested that some agglomeration has taken place in the present 3.3. Absorption spectra
samples. The largest particle size is obtained for ZnS:Cu sample.
Similar results are obtained in XRD data. The nanoparticles are The aqueous solution of ZnS, ZnS:Mn, ZnS:Cu, ZnS:Mn:Cu NPs are
spherical in shape. However, in doped system, two types of particles observed under UV-lamp (see Fig. 7). ZnS:Mn and ZnS:Mn:Cu NPs
are clearly observed which may be secondary phase. To check this exhibit pink and orange color emission, respectively. However, emis-
possibility, EDS analysis has been performed at different points. sion from ZnS:Mn NPs is less intense, which may be because of
However, the elemental composition, at all the three points, is more luminescence quenching due to higher (10%) content of Mn in ZnS:Mn
or less the same. Thus, it seems that secondary phase is not present in NPs [19]. In comparison to this, 5% content of Mn2+ and Cu2+ in
the present sample, which is also supported by the XRD results. ZnS:Mn:Cu and 3% in ZnS:Cu NPs is taken, which gives a more intense
Moreover, the EDS data of the samples confirm the doping of Mn, emission. ZnS:Cu NPs showed fluorescence in blue-green region. The
Cu, Mn:Cu (see Fig. 6). The peaks of Si and C originated from the luminescent property of these semiconductor nanoparticles confirms
silicon wafer on which sample was poured for carrying out SEM study. the formation of nanoparticles and occurrence of quantum confine-
In addition to this, conducting carbon tape was also used as an ment effect at nanoscale level [5]. Undoped ZnS did not show
adhesive between the SEM stub and Si wafer. fluorescence due to its excitation and emission in UV region, so, it
may be the reason it is not appearing to be fluorescent in the image

7197
N. Bansal et al. Ceramics International 43 (2017) 7193–7201

valence band to conduction band. For obtaining the absorption


characteristics of all the samples, at first the absorbance (A) at different
wavelengths (λ) are measured, and then absorption coefficients (α) at
the corresponding wavelengths λ are calculated using the Beer–
Lambert's relation [23,24]. For calculating the direct bandgap (Eg)
value of all the samples, Tauc's Plot [25] [(αhν)2 vs. hν] is obtained (see
Fig. 9). The obtained values of the optical bandgap (see Table 1) of the
samples are higher than that of the bulk value of ZnS (3.68 eV) and are
in close proximity with the reported values of ZnS NPs [25]. The
formation of hexagonal phase in ZnS:Mn:Cu sample is showing an
exceptionally increased bandgap compared to the bulk ZnS (Wurtzite),
which is 3.91 eV.
The optical band gap of any nanoparticle varies with its size and is
governed by Brus equation [26].

h2 ⎡ 1 1 ⎤ 1. 8e2
Enp = Eg + ⎢ + ⎥−
Fig. 7. (a) ZnS:Mn, (b) ZnS:Mn:Cu, (c) ZnS:Cu and (d) undoped-ZnS QDs as observed 8r ⎣ me*
2
mh* ⎦ εr (1)
under UV lamp.
Enp and Eg are the optical band gap energy of nanoparticle and
ZnS:Mn:Cu bulk in Joules, respectively.
ZnS:Mn h is the planck's constant, r is radius of nanoparticle; me* and mh*
ZnS:Cu are effective mass of electrons and holes, respectively; e is charge on
ZnS
electron; ε is the dielectric constant for material.
In case of ZnS, [2,27], considering the band gap value of bulk as
Absorbance (a.u.)

3.68 eV, effective mass of electrons and holes as 0.34 me* and
0.23 me*(me*=mass of an electron), respectively and dielectric con-
stant ε is 8.76, the size of nanoparticles (R) is calculated using Brus
equation and compared with XRD results (see Table 1). Enp is the
optical band gap of the nanoparticles. The obtained particle size follows
the similar trend as observed in all the techniques. However, crystallite
size/ particle sizes are different from each technique, probably, because
all the methods are based on different principles. Moreover, XRD and
optical method are indirect criteria, whereas, TEM is a direct technique
200 300 400 500 600 700 800
to obtain the particle size. However, the particle sizes are comparable
Wavelength (nm) when calculated/measured by three different techniques viz. XRD,
Fig. 8. Absorption characteristics of undoped-ZnS, ZnS:Mn, ZnS:Cu, ZnS:Mn:Cu QDs TEM and UV–vis Spectroscopy. According to crystallite size and
over UV–Vis range of 200–800 nm. particle size of the nanoparticles, undoped and doped systems follow
the following sequence as ZnS:Mn:Cu, ZnS:Mn, ZnS, and ZnS:Cu in
increasing order. With the decrease in size, the optical band gap has
increased.

3.4. Photoluminescence spectroscopy

Observed excitation values for each sample were 270 nm for ZnS,
300 nm for ZnS:Cu and 330 nm for ZnS:Mn and ZnS:Mn:Cu NP
samples, at which the emission spectra is obtained (see Fig. 10).
Undoped ZnS shows three significant emission peaks. One major
peak centred around 428 nm and another two minor peaks at 367 nm
and 488 nm are observed, respectively. It is reported in literature that
the emission spectrum of ZnS can be tuned in a very broad range from
300 nm to 550 nm. Several synthesis parameters are responsible for
these wide emissions peaks, like refluxing time, capping agent and its
concentration, reaction temperature, aging etc. [28–32]. The emission
peak observed at 428 nm is attributed to defect trap states present in
Fig. 9. Calculation of optical bandgap from the UV–Vis absorption spectra. ZnS nanoparticles due to sulfur vacancies [28,29]. The mercapto group
(-SH) of the passivating ligands gets detached from the organic part
[20]. and then organic part gets attached to Zn ions. These organic groups
The absorption peaks of the undoped and doped ZnS NPs appeared are responsible in removing Zn dangling orbitals. However, the
in the spectral range of 270–290 nm wavelengths (see Fig. 8). The unsaturated sp3 hybridized orbitals of the surface dangle out of the
absorbance peak of samples of ZnS, ZnS:Mn, ZnS:Cu, ZnS:Mn:Cu is crystals surface causing deep hole trap states. This cause red shifted
observed at 265, 268, 286 and 270 nm, respectively. Present samples emission at longer wavelengths, as observed at 488 nm in our case, and
show the absorption peaks at lower wavelength than bulk ZnS also has been reported by others [30,32]. The observed 367 nm (UV)
(340 nm). The blue shift confirms particle size in nanoscale region emission of undoped ZnS is due to near-band-edge emission originat-
which lead to increase in optical band gap [21,22]. The nature and the ing from the recombination of free excitons of ZnS nanoparticles size
value of optical bandgap can be determined from the fundamental regime [28]. The position of emission peaks represents the type of
absorption values, which corresponds to electron excitation from defects present in the ZnS nanoparticles, whereas, intensity represents

7198
N. Bansal et al. Ceramics International 43 (2017) 7193–7201

Fig. 10. PL emission spectrum for the 4 samples viz undoped ZnS (λex=270 nm), ZnS:Mn (λex=330 nm), ZnS:Cu (λex=300 nm), ZnS:Mn:Cu (λex=330 nm) over complete visible
spectrum range from 400 to 700 nm.

tunable emissions usually centred in blue-green region (470–500 nm).


The emission is tunable based on dopant concentration, electronic state
of copper (Cu2+ or Cu1+) and particle size of the sample [36]. The
emission peak of ZnS:Cu is centred at 481 nm (blue green region),
which is characteristic of copper dopant emission, occurs possibly due
to recombination of an electron from the shallow delocalized donor
level to the T2 level of Cu2+.
Emission spectrum of co-doped ZnS:Mn:Cu sample shows a broad
range spectrum. The deconvolution of the spectra represents three
major peaks centred at 520, 600 and 660 nm, respectively, shown in
Fig. 11. Peak around 520 nm clearly indicates the suppressed emission
from Cu states. The ZnS:Mn:Cu sample exhibits emissions from
Cu2+and Cu1+ states, in the range of 470–550 nm, masked in presence
of Mn2+ or other states. It is also reported by Jana et al. [36] that in
presence of Mn and Cu dopant in ZnSe nanoparticles, the emission
from Mn D-states predominated over emission from Cu states.
Fig. 11. The deconvolution of ZnS:Mn:Cu PL spectra using Gaussian fitting. Therefore, co-doped ZnS nanoparticles retained the Mn-emission at
596 nm. However, the presence of a peak around 660 nm indicates the
the relative abundance to each other. The abundance of defect traps red tuned Mn d-d transitions which are originating from the con-
states at 428 nm and band-edge emission at 367 nm could be the strained 4T1-6A1 levels. [37] The presence of Cu along with Mn brings a
possible reason because of which synthesized undoped ZnS appeared strain in the lattice. Along with this, the presence of functional group
non-fluorescent when directly observed under UV-light (Fig. 7). tends to interfere with the splitting field of the D-orbitals, which results
In general, the Mn doping in ZnS produces a short range tunable in the lowering of band gap between 4T1-6A1 states. This increases the
emission while Cu is known for its wide range tunable emission broadness of PL spectra resulting in broad range emission in visible
dependent on the band gap of ZnS. The ZnS:Mn shows a broad region. The transitions corresponding to the emission states are
emission spectrum with characteristic emission peak at 603 nm. This represented in the form of energy band diagram (see Fig. 12). The
emission is attributed to 4T1-6A1 transition of Mn2+ impurity and is not distribution of energy band, in case of co-doped ZnS, into discrete
related to any defect trap emission of ZnS [27,33–36]. states is an evidence for the broadness in emission spectrum. This
Doping of copper in host semiconductors produces wide range of broad range emission is highly suitable for bio-imaging application of

7199
N. Bansal et al. Ceramics International 43 (2017) 7193–7201

Fig. 12. Transitions leading to PL emission in ZnS and ZnS:Mn:Cu sample, respectively; CB-Conduction Band, VB-Valence Band, IZn- Interstitial Zinc, VS- Sulfur Vacancy.

present only in ZnS, ZnS:Mn and co-doped ZnS:Mn:Cu nanoparticles,


which corresponds to C-O stretching in primary alcohols suggests the
particle surface oxidation. The thiol/mercapto groups are chemically
susceptible to oxidation and forms disulfide and polysulfide products.
The presence of 654 cm−1 and 480 cm−1 band corresponds to disulfide
(C-S) and polysulfide (s-S) bonds. ZnS:Cu nanoparticles sample were
directly functionalized with mercaptoacetic acid, therefore, no C-O
stretching peaks of primary alcohol is observed. The FTIR peak
corresponding to bicarboxylates (O˭C=O) at 2953 cm−1 suggests the
oxidation of surface groups [38,39,44]. Apart from these, the FTIR
spectra of each nanoparticles are also closely resembled to the FTIR
spectra of its respective capping agent.
So, the FTIR spectra of each nanoparticles sample confirm the
presence of respective capping agents with functionalization of –OH
and –COOH functional groups. The comparative study also shows that
the capping agent has affected the size of the nanoparticles. Where
Fig. 13. The FTIR spectra of the powdered sample at room temperature. mercaptoacetic acid is used for Cu doped ZnS nanoparticles, it has
resulted in bigger sized nanoparticles. The use of mercaptoethanol as a
Table 3 capping agent for the rest of the samples has proved to be a better
FTIR peaks assignment corresponding to the present functional groups. candidate to inhibit the growth of particle size and resulted in smaller
sized nanoparticles. A similar effect on particle size has been reported
Peak position Corresponding functional group [45,46].
−1
Broad, 3300 − 3400 cm O-H Stretching [38,40]
~2925 cm−1 -CH2- methylene group Stretching (Asymmetric)
[40]
~2830 − 2870 cm−1 -CH2- methylene group Stretching (symmetric.)
4. Conclusion
[40]
~2359 cm−1 Bicarboxylates (O˭C=O) [43] Undoped and transition metal doped fluorescent ZnS nanoparticles
~1590–1600 cm−1 Carboxylate Salt [41] are synthesized by low temperature aqueous, co-precipitation method.
~1350 – 1400 cm−1 Carboxylate Salt [41]
All the samples are monophasic as conferred by XRD and HRTEM. Co-
~1050 cm−1 C-O Stretching in Primary Alcohols [42]
~654 cm−1 Disulfides (C-S) [40] doping of Mn2+ and Cu2+ in ZnS has lead to stable hexagonal phase
~480 cm−1 Polysulfides (S-S) [40] formation of ZnS as confirmed by SAED at a low temperature (70 °C),
probably due to nucleation kinetics taking place at nanoscale. The UV–
Vis absorption study shows the blue shift in optical band gap for doped
the nanoparticles. NPs. The shift of emission from UV to visible region is confirmed by the
photoluminescence spectroscopic studies. The present nanoparticles
3.5. FTIR spectroscopy are dispersible in aqueous medium and have –SH functional group on
its surface. The presence of functional group along with co-doping has
Fig. 13 shows the FTIR spectra of as synthesized nanoparticles. The supported the decrease in size of particles. Therefore, transition metal
FTIR assigned peaks are given in Table 3. The peaks observed around doped NPs can be used as biological labels.
3300–3400 cm−1 appears due to -OH group stretching. In addition to
this, characteristic asymmetric and symmetric stretching of methylene
group (-CH2) can be observed around ~2925 cm−1 and ~2830– Acknowledgement
2850 cm−1, respectively, suggesting the presence of either mercap-
toethanol or mercaptoacetic acid. It is interesting to notice that the Authors are thankful to SAI labs Thapar University for carrying out
FTIR peaks near ~1600 cm−1 and 1350 cm−1 in all samples indicating the XRD of the samples and SAIF facility at IIT, Bombay for FEG-TEM
the presence of carboxylate ions. Although ZnS, ZnS:Mn and characterisation. The author is also grateful to authorities of CSIO-
ZnS:Mn:Cu nanoparticles are surface functionalized with mercap- CSIR to provide facilities to carry out the research work. One of the
toethanol (primary alcohol), presence of peaks corresponding to author (N. B.) is really thankful to Dr. Devinder Mehta, Dr. Gurbinder
carboxylates, suggests the oxidation of surface mercaptoethanol mole- Kaur and Mr. Satwinder Singh for their consistent guidance and
cules to mercaptoacetic acid. Moreover, the peak ~1050 cm−1 is support.

7200
N. Bansal et al. Ceramics International 43 (2017) 7193–7201

References [24] R. Sarkar, C.S. Tiwary, P. Kumbhakar, S. Basu, A.K. Mitra, Yellow-orange light
emission from Mn2+-doped ZnS nanoparticles, Physica E: Low.-Dimens. Syst.
Nanostruct. 40 (2008) 3115–3120.
[1] R. Bhadra, V.N. Singh, B.R. Mehta, P. Datta, Studies on some aspects of ZnS [25] J. Tauc, Amorphous and Liquid Semiconductors, Plenum Press, London, 1974.
nanocrystals for possible applications in electronics, Chalcogenide Lett. 6 (2009) [26] L.E. Brus, Electronic wave functions in semiconductor clusters: experiment and
189–196. theory, J. Phys. Chem. 90 (1986) 2555–2560.
[2] J.P. Borah, K.C. Sharma, Optical and optoelectronic properties of ZnS nanostruc- [27] L. Peng, Y. Wang, Effects of the template composition and coating on the
tures thin film, Acta Phys. Pol. A 114 (2008) 713–719. photoluminescence properties of ZnS:Mn nanoparticles, Nanoscale Res. Lett. 5
[3] D. Graham-Rowe, From dots to devices, Nat. Photonics 3 (2009) 307–309. (2010) 839–845.
[4] M. Bruchez, M. Moronne, P. Gin, S. Weiss, A.P. Alivisatos, Semiconductor [28] P. Sana, L. Hashmi, M.M. Malik, Luminescence and morphological kinetics of
nanocrystals as fluorescent biological labels, Science 281 (1998) 2013–2016. functionalized ZnS colloidal nanocrystals, ISRN Opt. 8 (2012).
[5] A.P. Alivisatos, Semiconductor clusters, nanocrystals, and quantum dots, Science [29] C.S. Pathak, D.D. Mishra, V. Agarwal, M.K. Mandal, Blue Light emission from
27 (1996) 1933–1937. barium doped zinc sulphide nanoparticles, Ceram. Int. 38 (2012) 5497–5500.
[6] A.K. Kole, P. Kumbhakar, Cubic-to-hexagonal phase transition and optical prop- [30] S. Jinjun, Motlan, K. Drozdowicz-Tomsia, G. Zhu, E.M. Goldys, International
erties of chemically synthesized ZnS nanocrystals, Results Phys. 2 (2012) 150. Conference On Nanoscience and Nanotechnology, Brisbane, Qld., July, 2006.
[7] J.Q. Sun, X.P. Shen, K.M. Chen, Q. Liu, W. Liu, Low-temperature synthesis of [31] M. Sharma, S. Kumar, O.P. Pandey, Studies of 2 - mercaptoethanol passivated ZnS
hexagonal ZNS nanoparticles by a facile microwave-assisted single-source method, core-shell nanoparticles, Optoelectron. Adv. Mat. Rapid Comm. 2 (2008) 881–885.
Solid State Commun. 147 (2008) 501–504. [32] B. Bahmani, F. Moztarzadeh, M. Rabiee, Synthesis of zinc sulfide semiconductor
[8] S.A. Acharya, N. Maheshwari, L. Tatikondewar, A. Kshirsagar, S.K. Kulkarni, nanoparticles by co-precipitation method for biological diagnostics, J.
Ethylenediamine-mediated wurtzite phase formation in ZnS, Cryst. Growth Des. 13 Optoelectron. Adv. Mat. 9 (2007) 3336–3339.
(2013) 1369–1376. [33] X. Liu, X. Chen, X. Cui, R. Yu, The structure and malfunctional behaviors of Mn-
[9] H. Zhang, F. Huang, B. Gilbert, J.F. Banfield, Molecular dynamics simulations, ZnO/Mn-ZnS nanocomposites, Ceram. Int. 40 (2014) 13847–13854.
thermodynamic analysis, and experimental study of phase stability of zinc sulfide [34] H. Yang, S. Santra, P.H. Holloway, Syntheses and Applications of Mn-Doped II-VI
nanoparticles, J. Phys. Chem. B. 107 (2003) 13051–13060. Semiconductor Nanocrystals, J. Nanosci. Nanotechnol. 5 (2005) 1364–1375.
[10] D. Bera, L. Quain, T.K. Tseng, P.H. Holloway, Quantum Dots and Their Multimodal [35] P. Yang, C. Song, M. Lu, G. Zhou, Z. Yang, D. Xu, D. Yuan, Photoluminescence of
Applications: a Review, Materials 3 (2010) 2260–2345. Cu+-doped and Cu2+-doped ZnS nanocrystallites, J. Phys. Chem. Solids 63 (2002)
[11] H. Hu, W. Zang, Synthesis and properties of transition metals and rare-earth 639–643.
metals doped ZnS nanoparticles, Opt. Mater. 28 (2006) 536–550. [36] S. Jana, B.B. Srivastava, N. Pradhan, Correlation of Dopant States and Host
[12] Z. L. Wang, Y. Liu and Z. Zhang (Eds.), Handbook of Nanophase and Bandgap in Dual-Doped Semiconductor Nanocrystals, J. Phys. Chem. Lett. 2
Nanostructured Materials Synthesis, Kluwer Academic Plenum Publishers, (2011) 1747–1752.
Tsinghua University Press, 2002. [37] N. Pradhan, Red-tuned Mn d-d emission in doped semiconductor nanocrystals,
[13] C. Corrado, Y. Jiang, F. Oba, M. Kozina, F. Bridges, J.Z. Zhang, synthesis, structural Chem. Phys. Chem. 17 (2016) 1087–1094.
and optical properties of stable ZnS:Cu,Cl nanocrystals, J. Phys. Chem. A 113 [38] S.U. Rege, R.T. Yang, A Novel FTIR Method for studying mixed gas adsorption at
(2009) 3830–3839. Low Concentrations: H2O and CO2 on NaX zeolite and γ-alumina, Chem. Eng. Sci.
[14] B.D. Cullity, Elements of X-ray Diffraction, second ed., Addison–Wesley Company, 56 (2001) 3781–3796.
Reading, Massachusetts, USA, 1978. [39] J. Coates, Interpretation of infrared Spectra, A Practical Approach, Encyclopedia of
[15] S. Thakur, O.P. Pandey, K. Singh, Structural and optical properties of Bi1−xAxFeO3 Analytical Chemistry, John Wiley & Sons, Ltd, Newtown, USA, 2006.
(A = Sr, Ca; 0.40≤x≤0.55), J. Mol. Struct. 1074 (2014) 186–192. [40] N.V. Hullavard, S.S. Hullavarad, Optical properties of organic and inorganic capped
[16] J.D. Bryan, D.R. Gamelin, Doped Semiconductor Nanocrystals: synthesis, CdS nanoparticles and the effects of x-ray irradiation on organic capped CdS
Characterization, Physical Properties, and Applications, Prog. Inorg. Chem. 54 nanoparticles, J. Vac. Sci. Technol., A 26 (2008) 1050–1057.
(2005) 47–126. [41] M. Mohammadikish, F. Davar, M.R. Loghman-Estarki, Z. Hamidi, Synthesis and
[17] Y. Zhao, Y. Zhang, H. Zhu, C.H. George, J.Q. Xiao, Low-Temperature Synthesis of characterization of hierarchial ZnS architectures based nanoparticles in the
Hexagonal (Wurtzite) ZnS Nanocrystals, J. Am. Chem. Soc. 126 (2004) 6874–6875. presence of thiglycolic acid, Ceram. Int. 39 (2013) 3173–3181.
[18] F.A. La Porta, J. Andres, M.S. Li, J.R. Sambrano, J.A. Varela, E. Longo, Zinc blende [42] C.B. Mendive, D. Hansmann, T. Bredow, D. Bahnemann, New insights into the
versus wurtzite ZnS nanoparticles: control of the phase and optical properties by mechanism of TiO2 photocatalysis: thermal processes beyond the electron-hole
tetrabutylammonium hydroxide, Phys. Chem. Chem. Phys. 16 (2014) creation, J. Phys. Chem. C 115 (2011) 19676–19685.
20127–20137. [43] S.R. Husain, F. Ahmad, M. Ahmad, Synthesis of sulphur-containing derivatives
[19] P.H. Borse, D. Srinivas, R.F. Shinde, S.K. Date, W. Vogel Fritz, S.K. Kulkarni, Effect from olefinic fatty esters, J. Am. Oil Chem. Soc. 60 (1983) 1340–1344.
of Mn2+ concentration in ZnS nanoparticles on photoluminescence and electron- [44] S.K. Arya, T. Vats, S.N. Sharma, K. Singh, A.K. Narula, Effect of mercaptopropionic
spinresonance spectra, Phys. Rev. B 60 (1999) 8659–8664. acid as linker on structural, thermal, and optical properties of TiO2–CdSe
[20] A. Chatterjee, A. Priyam, S.C. Bhattacharya, A. Saha, Differential growth and nanocomposites, J. Therm. Anal. Calorim. 107 (2012) 555–560.
photoluminescence of ZnS nanocrystals with variation of surfactant molecules, [45] A.K. Singh, V. Viswanath, V.C. Janu, Synthesis, effect of capping agents, structural,
Colloids Surf., A: Physicochem. Eng. Asp. 297 (2007) 258–266. optical and photoluminescence properties of ZnO nanoparticles, J. Lumin. 129
[21] M. Dhanam, B. Kavitha, N. Jose, D.P. Devasia, Analysis of ZnS Nanoparticles (2009) 874–878.
prepared By Surfactant Micelle- Template Inducing Reaction, Chalcogenide Lett. 6 [46] B. Kinkaed, T. Hegmann, Effects of size, capping agent, and concentration of CdSe
(2009) 713–722. and CdTe quantum dots doped into a nematic liquid crystal on the optical and
[22] R.N. Bhargava, D. Gallaghar, X. Hong, A. Nurmikko, Optical properties of electro-optic properties of the final colloidal liquid crystal mixture, J. Mater. Chem.
manganese-doped nanocrystals of ZnS, Phys. Rev. Lett. 72 (1994) 416–419. 20 (2010) 448–458.
[23] D.F. Swinehart, The Beer-Lambert Law, J. Chem. Educ. 39 (1962) 333.

7201

You might also like