You are on page 1of 17

Applied Physics A (2023) 129:169

https://doi.org/10.1007/s00339-023-06463-x

Antibacterial studies of ZnO and silica capped manganese doped zinc


sulphide nanostructures
Sunil Kumar1 · Anita Jain2 · Sanjay Panwar3 · Indu Sharma4 · Suhaas Gupta5 · Milan Dopita5 · Ravi Kant Choubey6

Received: 14 October 2022 / Accepted: 28 January 2023


© The Author(s), under exclusive licence to Springer-Verlag GmbH, DE part of Springer Nature 2023

Abstract
To investigate a potential new antibacterial agent to combat increasing antimicrobial resistance, undoped and 1% manganese
doped Zinc Sulphide quantum dots (ZnS and Z ­ n0.99Mn0.01S QDs, respectively) were synthesised by co-precipitation method
and capped with increasing amounts of Zinc Oxide and Silica in aqueous media to prepare ZnS@ZnO, Z ­ n0.99Mn0.01S@
ZnO, and Z ­ n0.99Mn0.01S@SiO2 nanostructures. P-XRD analysis confirmed the cubic zinc-blende phase of the seed ZnS
QDs, ­Zn0.99Mn0.01S QDs, and ­Zn0.99Mn0.01S@SiO2 nanostructures, and the wurtzite phase of the ZnO in the ZnS@ZnO
and ­Zn0.99Mn0.01S@ZnO nanostructures, further confirmed using TEM studies, which also revealed the size of the largest
nanostructures to be in the range of a hundred nanometres. FTIR spectroscopy illustrated the quenching of characteristic ZnS
peaks with increasing capping material. UV–Visible absorption spectroscopy and subsequent Tauc analysis illustrated the
strong size confinement of the synthesised ZnS and Z ­ n0.99Mn0.01S QDs; Brus equation calculations revealed that the particle
size of the samples increases with increasing capping material. Photoluminescent emission spectroscopy illustrated the tune-
able emission properties of the prepared nanostructures; manganese doping induced the characteristic orange emission in
the ­Zn0.99Mn0.01S QDs, which was enhanced by ZnO, but quenched by S ­ iO2. The antimicrobial activity of all the prepared
samples was qualitatively evaluated using well known Agar well diffusion method against six human pathogenic bacteria:
Gram positive Bacillus subtilis and Staphylococcus aureus; Gram negative Salmonella Typhi, Escherichia coli, Klebsiella
pneumoniae and Pseudomonas aeruginosa. Qualitative antibacterial assay confirmed the high antibacterial potential of
the synthesised ZnS and ­Zn0.99Mn0.01S QDs, especially against E. coli. Increasing ZnO amount improves the antibacterial
activity of the nanostructures against different Gram-positive bacterial strains, while increasing ­SiO2 amount improves the
antibacterial activity of the nanostructures against both Gram positive strains and three of the four Gram negative bacte-
rial strains. Thus, the positive results suggest that the prepared ZnS@ZnO, ­Zn0.99Mn0.01S@ZnO, and Z ­ n0.99Mn0.01S@SiO2
nanostructures should be further studied as antimicrobial agents.

Keywords ZnS · ZnO · SiO2 · Optical properties · Antibacterial assay

4
* Suhaas Gupta Department of Biotechnology, Maharishi Markandeshwar
suhaas96@gmail.com Univeristy, Mullana, Ambala, Haryana 133207, India
5
* Ravi Kant Choubey Department of Condensed Matter Physics, Faculty
ravikantchoubey14@gmail.com of Mathematics and Physics, Charles University, Ke Karlovu
5, 12116 Prague, Czech Republic
1
Indira Gandhi University, Meerpur, Rewari, 6
Department of Applied Physics, Amity Institute of Applied
Haryana 122502, India
Sciences (AIAS), Amity University, Noida Campus,
2
Department of Physics, Maharishi Markandeshwar Sector‑125, Noida, Uttar Pradesh 201313, India
University, Mullana, Ambala, Haryana 133207, India
3
School of Basic and Applied Sciences, Maharaja Agrasen
University, Atal Shiksha Kunj, Kalujhanda, Barotiwala,
Solan, Himachal Pradesh 174103, India

13
Vol.:(0123456789)
169 Page 2 of 17 S. Kumar et al.

1 Introduction very promising candidates for the replacement or enhance-


ment of conventional antibiotic agents. Nanoparticle based
Antibacterial agents act to either kill bacterial cells (bac- antimicrobial agents make use of multiple simultaneous
tericidal) or suppress bacterial growth (bacteriostatic); mechanisms to cause cell death in a broad spectrum of
antibacterial and antimicrobial materials are relevant to bacterial strains, overcoming the problem of development
multiple industries, like food, packaging, textile, decora- of antimicrobial resistance in bacteria [11, 12].
tion, construction, etc., along with the obvious applica- The present work highlights the antibacterial activity
tions in the healthcare and biomedical fields. Their overuse of pristine and 1% manganese doped ZnS nanoparticles,
has, however, led to multiple bacterial strains expressing employed as seed material in nanostructures with increas-
increased resistance to conventional drug regimes, a phe- ing amounts of ZnO or ­SiO2, against six different human
nomenon that is both increasingly common and problem- pathogenic bacteria. ZnS, ZnO, ­SiO2 nanoparticles and a
atic; as the drug-resistance of bacterial strains increases, myriad of their nanocomposites have been shown to exhibit
necessary dose of drugs (and corresponding toxicity), antibacterial potential before. Recently, Panthi et al. showed
length and complexity of treatments, and mortality rates the light activated antibacterial activity of polyvinyl acetate
also increase. Increased antibacterial resistance can mani- nanofibers doped with rice grain-shaped ZnS nanoparticles
fest due to various different mechanisms, including but not [13]; Ananth et al. showed the shape-dependent antibacte-
limited to, gene mutation, decreased uptake of relevant rial activity of ZnS nanomaterials (rods, layers and spherical
drugs into the cell, formation of biofilm to avoid contact aggregates) prepared with an atmospheric pressure soft jet
with the antibacterial agent, or deactivation of antibiotic plasma device [14]; Kokilavani et al. showed the improved
molecule; while these problems have been addressed to antibacterial activity of ZnS/Ag2WO4 nanocomposites as
some extent through the development of new drugs or the compared to ZnS and ­Ag2WO4 individually [15]; Bhar-
modification of existing ones, the efforts have been out- athi et al. compared the antibacterial activity of S­ iO2–ZnO
paced by the persistent evolution of pathogens. A compro- nanocomposites prepared with chemical synthesis and
mised ability to treat serious infections gravely threatens green synthesis techniques [16]. Despite the vast amount
human health, and hence, research on more long-term, of research done on the antibacterial activity of ZnS, ZnO
robust and reliable methods to combat increased antibac- and/or ­SiO2 nanocomposites [17], the authors are unaware of
terial resistance is in vogue [1–4]. any investigation carried out on the antibacterial efficacy of
Nanoparticles having particle sizes in the strong-con- the nanostructure scheme specified in the present work, i.e.,
finement regime are extremely popular materials for appli- ZnS@ZnO, ­Zn0.99Mn0.01S@ZnO and ­Zn0.99Mn0.01S@SiO2
cation based research in a number of different fields (elec- nanostructures with increasing amounts of capping material;
tro-optics like in photovoltaic cells and display devices, previously, the authors of this work have investigated the
magnetics like in spintronic devices, energy storage and antibacterial efficacy of ZnS@SiO2 nanostructures, which
conversion, catalysis, building, filling and coating of nano- exhibited increased antibacterial activity against Bacillus
composite assemblies, to name a few) due to their easily- subtilis with increasing ­SiO2 shell thickness, and demon-
tailored novel electrical, optical, magnetic and chemical strated very good photo-stability [18]. The authors have also
properties that arise as a consequence of quantum confine- previously investigated the usefulness of Manganese doping
ment of charge carriers and enhanced effective surface area on the optical properties of ZnS for potential biocompat-
[5–7]. These novel material- and size-dependent properties ible applications [19, 20]. Structural and morphological
have found extensive potential and realised applications properties of the prepared samples were investigated with
in various biomedical fields like labelling, sensing and the help of X-ray diffraction, transmission electron micros-
detection of biomolecules, optical and magnetic imaging copy and Fourier transform infra-red spectroscopy; optical
of cells, targeted delivery of drugs and proteins, repair of properties were investigated with the help of UV–visible
implants and prosthetics and destruction of tumours [8, 9]. absorption and photoluminescent emission spectroscopy.
Several nanoparticles are also well known to exhibit anti- The antimicrobial activity of all the prepared samples was
microbial properties against a number of pathogenic bac- qualitatively evaluated using well known Agar well diffu-
teria, particularly silver nanoparticles which are already sion method against six human pathogenic bacteria: Gram
in use in biomedical fields in combination with organic positive Bacillus subtilis and Staphylococcus aureus; Gram
antibiotics [10]. Nanoparticles, while having high reac- negative Salmonella Typhi, Escherichia coli, Klebsiella
tivity due to enhanced surface properties, are relatively pneumoniae and Pseudomonas aeruginosa.
stable and biocompatible in aqueous media even in harsh
conditions, and do not produce harmful byproducts if
appropriately incorporated for disinfection, making them

13
Antibacterial studies of ZnO and silica capped manganese doped zinc sulphide nanostructures Page 3 of 17 169

2 Experimental methods The washed precipitate was allowed to dry in a vacuum


oven at 60 °C for 24 h, and then lightly grinded with a
2.1 Synthesis of ZnS and ­Zn0.99Mn0.01S quantum mortar-pestle to obtain a fine powder of ZnS@ZnO and
dots (QDs) ­Z n 0.99Mn 0.01S@ZnO, where the thickness of ZnO shell
was controlled by the amount of Zn(NO3)2 used during
ZnS and ­Zn0.99Mn0.01S QDs were synthesised using the synthesis; obtained powder samples were used for all
co-precipitation method reported in our earlier published characterisations.
articles, with slight modifications [19, 21]; Zinc Acetate
[Zn(CH3COO)2.H2O] was used as the ­Zn2+ source, Man-
ganese Acetate [Mn(CH 3COO) 2] was used as the ­M n 2+ 2.3 Synthesis of ­Zn0.99Mn0.01S@SiO2 samples
source, and Sodium Sulphide ­[ Na 2S] was used as the
­S2− source. Zn0.99Mn0.01S QDs, synthesised as described above, were
To synthesise ZnS QDs, carefully measured amounts (to used as the seed material for the growth of S ­ iO2 shell by
make 0.5 M solutions) of zinc source and sulphur source Stober process, with small modifications to the scheme
were dissolved separately in deionised water with vigorous reported in our earlier published articles [18, 20]. First,
magnetic stirring for 30 min at room temperature. Then, 0.8 g of the seed ­Zn0.99Mn0.01S QDs was dispersed in a
sulphur solution was added drop wise to the zinc solution; mixture of 5 mL deionised water and 20 mL ethanol using
precipitation of ZnS QDs soon starts, and was allowed ultrasonication for 1 h. Then, different amounts (0.25,
to continue for 1 h at room temperature with vigorous 0.5, 0.75 and 1 mL) of Tetraethyl Orthosilicate [TEOS,
magnetic stirring. The ZnS QD dispersion was then centri- Si(OC 2H 5) 4] were added to the dispersion under vigor-
fuged at 4000 rpm for 5 min, followed by filtration of the ous magnetic stirring at room temperature. 0.5 mL of
precipitate using Whatman 40 filter paper, and repeated 25% Ammonium Hydroxide ­[NH4OH] was then injected
washing with deionised water to remove unreacted mate- into the dispersion to act as a catalyst for the continuous
rial and adhered impurities. The washed precipitate was hydrolysis of the silica precursor and creation of ­S iO 2.
allowed to dry in a vacuum oven at 60 °C for 24 h, and After vigorous magnetic stirring for 1 h at room tempera-
then lightly grinded with a mortar-pestle to obtain a fine ture, the dispersion was centrifuged at 4000 rpm for 5 min,
powder of ZnS QDs which were used for characterisations followed by filtration of the precipitate using Whatman 40
and further synthesis. Z
­ n0.99Mn0.01S QDs were synthesised filter paper, and repeated washing with deionised water to
in the same scheme by simply replacing carefully meas- remove unreacted material and adhered impurities. The
ured amount of zinc source with the manganese source to washed precipitate was allowed to dry in a vacuum oven
obtain 1% dopant concentration. at 90 °C for 24 h, and then lightly grinded with a mortar-
pestle to obtain a fine powder of ­Z n 0.99Mn 0.01S@SiO 2,
where the thickness of the S ­ iO2 shell was controlled by
2.2 Synthesis of ZnS@ZnO and ­Zn0.99Mn0.01S@ZnO the amount of TEOS used during the synthesis; obtained
samples powder samples were used for all characterisations.

ZnS QDs and Z ­ n0.99Mn0.01S QDs, synthesised as described


above, were used as the seed material for the growth of 2.4 Characterisations
ZnO shell, synthesised using the same scheme as reported
in our earlier published articles [22, 23]. First, 0.2 g of Powder X-ray diffraction (P-XRD) was carried out on
the required seed material was dispersed in 200 mL of Rigaku D/MAX IIIC diffractometer in the Bragg–Brentano
deionised water using ultrasonication for 1 h. Then, dif- geometry with Cu–Kα radiation (1.54056Å) in the range of
ferent amounts (10, 30, 50 and 100 mL) of 0.05 M Zinc 20°–70°. Transmission electron microscopy (TEM) images
Nitrate [Zn(NO3)2] solution were added to the dispersion were recorded using Hitachi H-7500 TEM operating at
under vigorous magnetic stirring at room temperature, fol- 120 kV and equipped with a CCD camera. Fourier trans-
lowed by slow titration with 0.1 M Sodium Hydroxide form infrared (FTIR) spectroscopy was carried out on Perki-
[NaOH] solution, until the dispersion achieved 10pH, at nElmer Spectrum RX-I (40) in the range of 2000–400 ­cm−1.
which point the ZnO has nucleated and starts to grow. The UV–visible absorption spectroscopy was carried out on
dispersion was then centrifuged at 4000 rpm for 5 min, double beam PerkinElmer Hitachi 330 UV–Vis-NIR spec-
followed by filtration of the precipitate using Whatman trophotometer in the range 200–600 nm. Photoluminescent
40 filter paper, and repeated washing with deionised water emission (PLE) spectroscopy was carried out on Horiba
to remove unreacted material and adhered impurities. Jobin Yvon FluoroMax-3 spectrofluorometer using a Xenon
lamp as the excitation source.

13
169 Page 4 of 17 S. Kumar et al.

2.5 Antimicrobial assay

The antimicrobial activity of all the prepared samples


was evaluated using well known Agar well diffusion
method, which is a relatively quick and easily executable
semi-quantitative test to assess the antibacterial activity
of diffusible antibacterial agents [24]. Inhibitory action
of the prepared ZnS@ZnO, ­Z n 0.99Mn 0.01S@ZnO, and
­Z n 0.99Mn 0.01S@SiO 2 nanostructures was quantitatively
evaluated against six human pathogenic bacteria: Gram
positive Bacillus subtilis and Staphylococcus aureus;
Gram negative Salmonella Typhi, Escherichia coli, Kleb-
siella pneumoniae and Pseudomonas aeruginosa; anti-
microbial susceptibility discs (6 mm in diameter) of Tet-
racycline (30 µg/disc) and Gentamicin (10 µg/disc) were
used as the positive control, while pure ethanol was used
Fig. 2  Powder X-ray diffraction plots of the prepared Z
­ n0.99Mn0.01S@
as the negative control. Isolated colonies of the bacteria ZnO samples
were picked up from revived cultures with an inoculat-
ing loop and inoculated in saline water; inoculum den-
sity was adjusted against McFarland 0.5 standard solu- 3 Results and discussion
tion of ­BaSO4 (0.05 mL 1% ­BaCl2 + 9.95 mL 1% ­H2SO4)
which leads to an approximate bacteria cell density of 3.1 Powder X‑ray diffraction
1.5 × ­1 0 8 CFU/mL. 1 mL of the inoculum was spread
onto a Mueller Hinton Agar plate and four wells of 6 mm Figures 1, 2 and 3 show the P-XRD plots of the prepared
diameters were made, one in each of the quadrants, using ZnS@ZnO, ­Z n 0.99Mn 0.01S@ZnO and ­Z n 0.99Mn 0.01S@
a sterile cork borer. 5 mg/mL dispersions of all the syn- SiO2 samples, respectively; the P-XRD patterns of all the
thesised samples were made in ethanol, and 100 µL of the prepared samples were analysed using the software pack-
dispersions were poured into the wells. The plates were age MStruct [25–27], which is an extension of the FOX/
incubated overnight at 37 °C, and antimicrobial activity ObjCryst program [28, 29], with implementation of the
of the synthesised samples was evaluated by measuring whole powder pattern modelling (WPPM) method [30] for
the diameter of the inhibition zone. All the tests were per- real structure determination. Peak positions were determined
formed in triplicate and reported values (in mm) are the by refining unit-cell parameters and accounting for the peak
average of three results. position shift due to refraction at the sample interface; peak

Fig. 1  Powder X-ray diffraction plots of the prepared ZnS@ZnO Fig. 3  Powder X-ray diffraction plots of the prepared Z
­ n0.99Mn0.01S@
samples SiO2 samples

13
Antibacterial studies of ZnO and silica capped manganese doped zinc sulphide nanostructures Page 5 of 17 169

intensities were calculated from a known crystal structure decreased with increasing ZnO amount, until it was smaller
(as in FOX/ObjCryst) with absorption correction. Broaden- than that of the synthesised ­Zn0.99Mn0.01S QDs.
ing of the peaks was modelled using numerical convolution Figure 3 shows the P-XRD plots of the prepared
of a known pseudo-Voigt instrumental broadening function ­Zn0.99Mn0.01S@SiO2 samples. Characteristic peaks observed
and several other refinable parameters; for the sake of com- for all the samples are (111), (220) and (311), which match
putational simplicity, it was assumed that the distribution of with the bulk zinc-blende ZnS PDF No. 05-0566, indicat-
crystallite sizes takes the form of a log-normal distribution, ing that the Z­ n0.99Mn0.01S@SiO2 nanostructure adapts the
and that the nano-crystallites were spherical in shape. All the structure of the seed cubic Z ­ n0.99Mn0.01S material. For all
figures depict the observed intensity values, the calculated the prepared Z ­ n0.99Mn0.01S@SiO2 samples, obtained lat-
intensity values, and their difference. tice parameters of the cubic seed Z ­ n0.99Mn0.01S QDs were
Figure 1 shows the P-XRD plots of the prepared ZnS@ smaller than the lattice parameter reported in the ZnS PDF
ZnO samples. Characteristic peaks observed for the syn- No. 05–0566; the lattice parameter initially decreased drasti-
thesised ZnS QDs are (111), (220) and (311), which match cally in the presence of ­SiO2, but gradually increased with
with the bulk zinc-blende ZnS PDF No. 05-0566, confirm- increasing ­SiO2 shell thickness.
ing the required phase of the synthesised seed cubic ZnS
material; obtained lattice parameter a = 5.3815 Å is slightly 3.2 Transmission electron microscopy
smaller than the lattice parameter reported in the ZnS PDF
No. 05-0566 (= 5.4053 Å). In the samples containing ZnO, Figure 4a, b show the TE micrograph of the synthesised
characteristic peaks observed are (100), (002), (101), (102), ZnS and ­Zn0.99Mn0.01S QDs, respectively. Generally, ZnS
(110), (103), (200), (112) and (201), which match with the exhibits poor contrast in electron microscopy, however the
bulk wurtzite ZnO PDF No. 36–1451; as the amount of micrographs confirm the presence of spherical ZnS nano-
Zn(NO3)2 used during the synthesis is increased, character- particles with some agglomeration up to few tens of nano-
istic diffraction peaks of wurtzite ZnO become more well metres, while the ­Zn0.99Mn0.01S QDs exhibit slightly better
defined, and the intensity of the characteristic (111) peak monodispersity and smaller size due to incorporation of
of the seed cubic ZnS QDs is quenched. For all the pre- manganese in the lattice [19, 20].
pared ZnS@ZnO samples, obtained lattice parameters of the Figure 4c, d show the TE micrograph of the prepared
wurtzite ZnO were slightly larger than the lattice parameters ZnS@ZnO(30 mL) and ZnS@ZnO(100 mL) samples,
reported in the ZnO PDF No. 36-1451; lattice parameter of respectively, confirming the formation of the ZnS@ZnO
the cubic seed ZnS QDs further decreased in the presence nanocomposite from the seed ZnS QDs. Increasing the
of ZnO, but did not vary much with increasing ZnO amount. amount of Zn(NO3)2 used during synthesis increases the par-
Figure 2 shows the P-XRD plots of the prepared ticle size of the samples, and the micrograph of the ZnS@
­Zn0.99Mn0.01S@ZnO samples. Characteristic peaks observed ZnO(100 mL) sample exhibits the pure wurtzite structure
for the synthesised ­Zn0.99Mn0.01S QDs are (111), (220) and of the ZnO shell, with a size of upto few hundreds of nano-
(311), which match with the bulk zinc-blende ZnS PDF No. metres. Figure 4e, f show the TE micrograph of the pre-
05-0566, confirming the required phase of the synthesised pared ­Zn0.99Mn0.01S@ZnO(10 mL) and Z ­ n0.99Mn0.01S@
seed cubic Z ­ n0.99Mn0.01S material, and indicating that the ZnO(30 mL) samples, respectively, confirming the forma-
­Mn2+ ions used as dopant occupy (empty) sites in the ZnS tion of the ­Zn0.99Mn0.01S@ZnO nanocomposite from the
lattice without drastically changing the unit cell structure seed ­Zn0.99Mn0.01S QDs. Increasing the amount of Zn(NO3)2
of the of the host cubic material; obtained lattice parameter used during synthesis increases the particle size and agglom-
a = 5.3495 Å is smaller still than the lattice parameter of eration tendency of the samples. Figure 4g, h show the TE
the synthesised undoped ZnS QDs obtained above. In the micrograph of the prepared ­Zn0.99Mn0.01S@SiO2(0.5 mL)
samples containing ZnO, characteristic peaks observed are and ­Zn0.99Mn0.01S@SiO2(1 mL) samples, respectively. Sto-
(100), (002), (101), (102), (110), (103), (200), (112) and ber process used during synthesis produces spherical ­SiO2
(201), which match with the bulk wurtzite ZnO PDF No. particles and increasing the amount of the TEOS used dur-
36–1451; as the amount of Zn(NO3)2 used during synthesis ing synthesis increases the particle size and dispersity of
is increased, characteristic peaks of wurtzite ZnO become the samples.
more well defined, and the intensity of the characteristic
(111) peak of the seed cubic ­Zn0.99Mn0.01S QDs is quenched. 3.3 Fourier transform infrared spectroscopy
For all the prepared ­Zn0.99Mn0.01S@ZnO samples, obtained
lattice parameters of the wurtzite ZnO were slightly larger Figure 5 shows the FTIR spectra of the prepared ZnS@ZnO
than the lattice parameters reported in the ZnO PDF No. samples. Broad peaks common to all the samples around
36-1451; lattice parameter of the cubic seed ­Zn0.99Mn0.01S 1560 ­cm−1 and 1398 ­cm−1 are attributed to the antisym-
QDs initially increased in the presence of ZnO, but gradually metric and symmetric stretching of the ­COO − in the

13
169 Page 6 of 17 S. Kumar et al.

Fig. 4  Transmission electron


micrograph of the prepared:
a ZnS QD; b ­Zn0.99Mn0.01S
QD; c ZnS@ZnO(10 mL);
d ZnS@ZnO(100 mL); e
­Zn0.99Mn0.01S@ZnO(10 mL); f
­Zn0.99Mn0.01S@ZnO(30 mL); g
­Zn0.99Mn0.01S@SiO2(0.5 mL);
and h ­Zn0.99Mn0.01S@
SiO2(1 mL) samples

Zn-carboxylate group, respectively; broad band around 900- the range 495–515 ­cm−1, with increasing ZnO concentration
1500 ­cm−1 is commonly attributed to frequencies of oxygen or bulk character; minor peak exhibited around 421 ­cm−1
bending and stretching [19, 31–33]. Synthesised ZnS QDs corresponds to the Raman-active E2 (high) mode of hex-
exhibit a C–C stretching doublet peak at 1339 ­cm−1, absence agonal ZnO. The weak bands around 1100–1200 ­cm−1 are
of which in the prepared ZnS@ZnO samples can be an indi- characteristic of inorganic ions, and the weak bands around
cator of surface modification of the ZnS QDs by ZnO [34]. 600–900 ­cm−1 are attributed to vibrational frequencies aris-
Synthesised ZnS QDs also exhibit the following peaks: O–H ing due to changes in the lattice at the microstructural level
bending peak at 670 ­cm−1; characteristic Zn–S stretching [37, 38].
peak at 615 ­cm−1, which is absent in the FTIR spectra of the Figure 6 shows the FTIR spectra of the prepared
prepared ZnS@ZnO samples; and characteristic metal-oxide ­Zn0.99Mn0.01S@ZnO samples. Broad peak common to all
stretching mode at 485 ­cm−1 [35, 36]. The FTIR spectra the samples around 1623 ­cm−1 is attributed to the vibra-
of the prepared ZnS@ZnO samples also exhibit the Zn–O tion of the C=O bond in the Zn-carboxylate group; broad
bond stretching frequencies at increasing wavenumbers, in peaks around 1560 ­cm−1 and 1423 ­cm−1 are attributed to the

13
Antibacterial studies of ZnO and silica capped manganese doped zinc sulphide nanostructures Page 7 of 17 169

samples; and characteristic metal-oxide stretching mode at


486 ­cm−1 [35, 36, 40, 41]. The FTIR spectra of the prepared
­Zn0.99Mn0.01S@ZnO samples also exhibit the Zn–O bond
vibration modes in the range 491–548 ­cm−1 [36]. The weak
bands around 1100-1200 ­cm−1 are characteristic of inorganic
ions, and the weak bands around 600-900 ­cm−1 are attrib-
uted to vibrational frequencies arising due to changes in the
lattice at the microstructural level [22, 37, 38].
Figure 7 shows the FTIR spectra of the prepared
­Zn0.99Mn0.01S@SiO2 samples. Preliminary analysis reveals
that the FTIR spectra of all the samples are similar and share
a lot of common peaks: broad peaks around 1556 ­cm−1 and
1406 ­cm−1 are attributed to the antisymmetric and sym-
metric stretching of the C ­ OO− in the Zn-carboxylate group,
respectively; weak peak around 1340 ­cm−1 is attributed
Fig. 5  Fourier transform infrared spectra of the prepared: a ZnS QD; to the doublet of C–C stretching; broad band around 900-
b ZnS@ZnO(10 mL); c ZnS@ZnO(30 mL); d ZnS@ZnO(50 mL); 1500 ­cm−1 is commonly attributed to frequencies of oxygen
and e ZnS@ZnO(100 mL) samples bending and stretching; peak around 672 ­cm−1 is attributed
to O–H bending; and broad peak around 495 ­cm−1 is attrib-
uted to the metal-oxide stretching mode [18, 19, 31–36].
However, FTIR spectrum of the synthesised Z ­ n0.99Mn0.01S
QDs exhibits some key differences: broad peak at 1616 ­cm−1
is attributed to the vibration of the C=O bond in the Zn-car-
boxylate group; intense peak at 1106 ­cm−1, whose absence
in the FTIR spectra of the undoped ZnS QDs indicates
the effect the doped M ­ n2+ ions have on the host ZnS; and
characteristic Zn–S stretching peak at 617 c­ m−1; decreased
intensity of these peaks in the FTIR spectra of the prepared
­Zn0.99Mn0.01S@SiO2 samples can be an indicator of surface
modification of the Z ­ n0.99Mn0.01S QDs by S ­ iO2 [35, 36,
39–41].The weak bands around 1100-1200 ­cm−1 are char-
acteristic of inorganic ions, and the weak bands around 600-
900 ­cm−1 are attributed to vibrational frequencies arising

Fig. 6  Fourier transform infrared spectra of the prepared: a


­ n0.99Mn0.01S QD; b ­Zn0.99Mn0.01S@ZnO(10 mL); c ­Zn0.99Mn0.01S@
Z
ZnO(30 mL); d ­Zn0.99Mn0.01S@ZnO(50 mL); and e ­Zn0.99Mn0.01S@
ZnO(100 mL) samples

antisymmetric and symmetric stretching of the C ­ OO− in the


Zn-carboxylate group, respectively; broad band around 900-
1500 ­cm−1 is commonly attributed to frequencies of oxygen
bending and stretching [19, 39]. Synthesised Z ­ n0.99Mn0.01S
QDs exhibit a C–C stretching doublet peak at 1340 ­cm−1,
absence of which in the prepared Z ­ n 0.99Mn 0.01S@ZnO
samples can be an indicator of surface modification of the
­Zn0.99Mn0.01S QDs by ZnO [34]. Synthesised Z ­ n0.99Mn0.01S
QDs also exhibit the following peaks: intense peak at
1106 ­cm−1, whose absence in the FTIR spectra of the
undoped ZnS QDs indicates the effect the doped M ­ n2+ ions
Fig. 7  Fourier transform infrared spectra of the prepared:
have on the host ZnS; O–H bending peak at 671 ­cm−1; char-
a ­Zn0.99Mn0.01S QD; b ­Zn0.99Mn0.01S@SiO2(0.25 mL); c
acteristic Zn–S stretching peak at 617 ­cm−1, which is absent ­Zn0.99Mn0.01S@SiO2(0.5 mL); d ­Zn0.99Mn0.01S@SiO2(0.75 mL); and
in the FTIR spectra of the prepared Z ­ n0.99Mn0.01S@ZnO e ­Zn0.99Mn0.01S@SiO2(1 mL) samples

13
169 Page 8 of 17 S. Kumar et al.

Fig. 8  UV–Visible absorption spectra of the prepared: a ZnS@ZnO; b ­Zn0.99Mn0.01S@ZnO; and c ­Zn0.99Mn0.01S@SiO2 samples

due to changes in the lattice at the microstructural level [32, the core ZnS material for all the prepared samples were cal-
37, 38]. culated from the Eg values using the relation
hc
𝜆a = .
Eg
3.4 UV–visible absorption spectroscopy

Figures 8a–c show the UV–visible absorption spectra Particle sizes (R) of the core ZnS material for all the pre-
of the prepared ZnS@ZnO, Z ­ n 0.99Mn 0.01S@ZnO and pared samples were calculated from the Eg values using the
­Zn0.99Mn0.01S@SiO2 samples, respectively. Energy band effective mass approximation model of the Brus equation
gap (Eg) values of the core ZnS material for all the prepared [46, 47], given as
samples were determined from the UV–visible absorption
data using the Tauc relation [42–45], given as ℏ2 𝜋 2 1.8e2
Eg = Egbulk + 2
− ,
2𝜇R 4𝜋𝜀0 𝜀r R
(𝛼h𝜈)n = 𝛽 h𝜐 − Eg ,
( )

where Egbulk is the bulk energy band gap (= 3.6 eV for


where α is the absorption coefficient, h is the Planck con- cubic zinc-blende ZnS), ℏ is the reduced Planck constant,
stant, ν is the frequency of the photon, n = 2 for direct band µ is the reduced mass of the exciton, e is the elementary
gap semiconductors like ZnS, and β is an arbitrary constant. charge of an electron, ε0 is the permittivity of free space,
To obtain the values of energy band gap for all the prepared and εr is the relative permittivity of ZnS (= 8.9). Obtained
samples, linear portion of the Tauc plot between (αhν)2 on values of Eg, λa and R are summarised in Table 1, includ-
the y-axis and Eg on the x-axis is extrapolated to the x-inter- ing values of the absorption peak and energy band gap of
cept. Figures 9, 10 and 11 show the Tauc plots of the pre- the ZnO shell material that can be easily obtained from the
pared ZnS@ZnO, ­Zn0.99Mn0.01S@ZnO and ­Zn0.99Mn0.01S@ UV–visible absorption spectra of the prepared ZnS@ZnO
SiO2 samples, respectively. Absorption edge (λa) values of and ­Zn0.99Mn0.01S@ZnO samples.

13
Antibacterial studies of ZnO and silica capped manganese doped zinc sulphide nanostructures Page 9 of 17 169

Fig. 9  Tauc plots of the pre-


pared ZnS@ZnO samples

Fig. 10  Tauc plots of the


prepared ­Zn0.99Mn0.01S@ZnO
samples

Pristine ZnS QDs exhibit the largest band gap energy ­ n0.99Mn0.01S material, which is to
particle size) of the seed Z
(3.93 eV) as a result of being the most strongly confined be expected due to the optically transparent nature of ­SiO2
nanoparticles out of all the samples; pristine ZnS QDs and the small amounts of TEOS used during synthesis; for
also exhibit the smallest size (2.1 nm), and ­Zn0.99Mn0.01S the maximum amount of TEOS used during the synthesis,
QDs exhibit the next smallest particle size (2.3 nm) with a the ­Zn0.99Mn0.01S@SiO2(1 mL) sample exhibits a band gap
corresponding band gap energy of 3.84 eV. Capping with energy of 3.82 eV and corresponding particle size of 2.5 nm.
­SiO2 does not drastically affect the energy band gap (or the In the case of the prepared ZnS@ZnO sample, capping with

13
169 Page 10 of 17 S. Kumar et al.

Fig. 11  Tauc plots of the


prepared ­Zn0.99Mn0.01S@SiO2
samples

ZnO induces a more noticeable trend of decreasing band between acceptor and donor trap energy levels, and peaks
gap energy of the core ZnS material, and a corresponding at about 481 nm and 491 nm can be attributed to transi-
increase in the particle size of the same; for the maximum tions facilitated by oxygen antisite defects. Green emission
amount of Zn(NO 3)2 used during the synthesis, ZnS@ peak at about 557 nm can be attributed to transition between
ZnO(100 mL) sample exhibits a particularly decreased band donor–acceptor complexes or singly ionised oxygen vacan-
gap energy of 3.62 eV (and corresponding particle size of cies [48–51].
5.3 nm), which is close to the bulk energy band gap value Figure 12b shows the PLE spectra of the prepared
of ZnS. In comparison, increasing the amount of Zn(NO3)2 ­Zn0.99Mn0.01S@ZnO samples in the range of 350–600 nm
used during the synthesis of the Z ­ n0.99Mn0.01S@ZnO sam- with an excitation wavelength of 325 nm; prepared samples
ples also induces the same decreasing trend in the band exhibit near band edge emission at about 404 nm. Prepared
gap energy of the core ZnS material, but not as drastically samples also exhibit minor peak at about 422 nm, which
as for the prepared ZnS@ZnO samples; however, all the can be attributed to the transitions between ­Zn2+ intersti-
­Zn0.99Mn0.01S@ZnO samples exhibit a strong absorption tial level to the valence band, and a sharp peak at about
peak around 363 nm, corresponding to the direct band gap 450 nm, which can be attributed to the transitions between
energy (~ 3.4 eV) of nano-sized ZnO. doubly ionised Z ­ n2+ vacancy level and the valence band [50,
52]; these peaks exhibit very slight red-shift with increasing
3.5 Photoluminescent emission spectra ZnO amount. The prepared samples also exhibit an orange
emission peak at about 590 nm, well known to be charac-
Figure 12a shows the PLE spectra of the prepared ZnS@ teristic of the 4T1 → 6A1 transition of the doped M­ n2+ [44];
ZnO samples in the range of 325–575 nm with an excita- in the case of the prepared ­Zn0.99Mn0.01S@ZnO(100 mL)
tion wavelength of 320 nm. Synthesised ZnS QDs exhibit sample at 575 nm the emission peak of the ­Mn2+ transition
a continuous emission from 380 to 440 nm, attributed to fuses with the yellow-green emission of the ZnO, exhibited
the defect related emission band [19, 21]. The prepared at about 560 nm in the rest of the samples. The prepared
ZnS@ZnO samples exhibit a limited discretisation of this ­Zn0.99Mn0.01S@ZnO(100 mL) samples also exhibit a well-
broad band, possibly due to the increase in crystallinity as defined peak at 472 nm and a minor hump at 529 nm, which
a result of ZnO capping; the band exhibits a peak at about could be emissions arising from transitions facilitated by
400 nm, attributed to the near band-edge emission. Peaks at ­Zn2+ and ­O2− interstitials, respectively [48].
about 450 nm can be attributed to the transitions between Figure 12c shows the PLE spectra of the prepared
doubly ionised Z­ n2+ vacancy level and the valence band, ­Zn0.99Mn0.01S@SiO2 samples in the range of 350–600 nm
peaks at about 467 nm can be attributed to the transitions with an excitation wavelength of 325 nm; prepared samples

13
Antibacterial studies of ZnO and silica capped manganese doped zinc sulphide nanostructures Page 11 of 17 169

exhibit near band edge emission at about 405 nm. Prepared

3.82

2.5
324.6



samples also exhibit minor peak at about 423 nm, which can

1.0
be attributed to the transitions between ­S2− vacancy level to
the valence band, a sharp peak at about 450 nm, which can

Zn0.99Mn0.01S@SiO2 (x mL)

3.83

2.4
323.7



0.75
be attributed to the transitions between acceptor vacancies
and donor vacancies, and peak at about 557 nm (which nar-
rows and blue-shifts with increasing ­SiO2 amount) which

3.83

2.4
323.7



0.5 can be attributed to elemental sulphur species [19, 21]. Peak
at about 590 nm is well known to be characteristic orange
3.84
emission of the 4T1 → 6A1 ­Mn2+ transition [44, 52], whose
2.4
322.9



0.25

emission intensity decreases with increasing ­SiO2 amount.


From the PLE spectra of the prepared samples it is clear that
the emission intensity of the synthesised ZnS QDs is not
3.4
3.7

364.3
3.3
335.1
100

only highly tuneable by varying the material or amount of


capping or doping, but is also highly photo-stable for long-
­ n0.99Mn0.01S@ZnO, and ­Zn0.99Mn0.01S@SiO2 samples
Zn0.99Mn0.01S@ZnO (x mL)

3.73

3.4
363.8
3.0
332.4

term antibacterial application.


50

3.6 Antibacterial assay
326.3

3.41
363
3.8

2.6
30

Figures 13, 14 and 15 represent qualitative analysis of


the synthesised ZnS@ZnO, ­Z n 0.99Mn 0.01S@ZnO and
322.9

­Zn0.99Mn0.01S@SiO2 nanostructures by agar well-diffusion


3.43
3.84

361
2.4
10

method against the six human pathogenic bacteria. The Agar


well-diffusion method is a relatively quick method with eas-
Zn0.99Mn0.01S

ily executable semi-quantitative tests to assess the antibacte-


rial activity of diffusible antibacterial agents. The antibacte-
rial activity of the synthesised ZnS@ZnO, Z ­ n0.99Mn0.01S@
3.85
322
2.3


ZnO and ­Zn0.99Mn0.01S@SiO2 nanostructures; shows maxi-


mum inhibitory action against B. subtilis and E. coli with a
3.06
3.62

404.7
5.3
342.5

zone of inhibition 17 and 22 mm, respectively. The visible


100
Table 1  UV–visible absorption spectroscopy parameters of the prepared ZnS@ZnO, Z

presence of inhibition zones larger in diameter than the wells


bored in the Agar plate clearly indicate the antibacterial
3.24
3.71

381.7
3.2
334.2

activity of the synthesised nanostructures.


50

Table 2 summarises the antibacterial activity of the


ZnS@ZnO (x mL)

synthesised samples, along with the inhibition zone diam-


3.78
328
2.7
30

eters produced in the positive and negative controls, and



Fig. 16 graphically illustrates the same; synthesised ZnS and


3.86

2.3
321.2

­Zn0.99Mn0.01S QDs exhibit the maximum inhibitory action




10

against E. coli, with zones of inhibition measuring 24 mm


and 22 mm, respectively.
3.93

2.1
315.5


In the case of pristine ZnS QDs and ­Zn0.99Mn0.01S QDs,


ZnS

the antibacterial activity against Gram negative bacteria is


greater than against Gram positive bacteria, possibly as a
Particle size (nm) from Brus equation

result of the slender membrane of Gram negative bacteria


UV–visible absorption parameters

as compared to the multi-layered cell walls of Gram posi-


Absorption edge (nm) λa of ZnS
Band gap (eV) from Tauc Plot

Absorption peak (nm) of ZnO

tive bacteria; antibacterial activity against E. coli is better


than even that of the positive control. With the inclusion
Band gap (eV) of ZnO

of ZnO capping of varying thickness on the pristine ZnS


QDs, the antibacterial activity of the samples against Gram
negative bacteria decreases, while the antibacterial activ-
ity against Gram positive bacteria slightly improves. In
the case of the Z ­ n0.99Mn0.01S@ZnO samples, the antibac-
terial activity against Gram positive B. subtilis and Gram

13
169 Page 12 of 17 S. Kumar et al.

Fig. 12  Photoluminescent emission spectra of the prepared: a ZnS@ZnO; b ­Zn0.99Mn0.01S@ZnO; and c ­Zn0.99Mn0.01S@SiO2 samples

Fig. 13  Representative Agar plates for the well-diffusion investiga- ­Zn0.99Mn0.01S QDs; f ZnS@ZnO; g ­Zn0.99Mn0.01S@ZnO; and h
tion of antibacterial activity of the prepared: a ZnS and Z
­ n0.99Mn0.01S ­Zn0.99Mn0.01S@SiO2 samples against Gram positive Staphylococcus
QDs; b ZnS@ZnO; c ­Zn0.99Mn0.01S@ZnO; and d ­Zn0.99Mn0.01S@ aureus
SiO2 samples against Gram positive Bacillus subtilis; e ZnS and

negative Salmonella Typhi increases with an increase in ­SiO2 shell generally induces a trend of increasing antibac-
the thickness of ZnO shell; the maximum inhibition zone terial activity against both Gram positive and Gram nega-
exhibited is 14 mm and 16 mm, respectively. For the synthe- tive strains (maximum inhibition zone exhibited against
sised ­Zn0.99Mn0.01S@SiO2 samples, increasing thickness of B. subtilis was 17 mm); this is in agreement with reports

13
Antibacterial studies of ZnO and silica capped manganese doped zinc sulphide nanostructures Page 13 of 17 169

Fig. 14  Representative Agar plates for the well-diffusion investiga- SiO2 samples against Gram negative Salmonella Typhi; e ZnS and
tion of antibacterial activity of the prepared: a ZnS and Z
­ n0.99Mn0.01S ­Zn0.99Mn0.01S QDs; f ZnS@ZnO; g ­Zn0.99Mn0.01S@ZnO; and h
QDs; b ZnS@ZnO; c ­Zn0.99Mn0.01S@ZnO; and d ­Zn0.99Mn0.01S@ ­Zn0.99Mn0.01S@SiO2 samples against Gram negative Escherichia coli

Fig. 15  Representative Agar plates for the well-diffusion investiga- and ­Zn0.99Mn0.01S QDs; f ZnS@ZnO; g ­Zn0.99Mn0.01S@ZnO; and h
tion of antibacterial activity of the prepared: a ZnS and Z
­ n0.99Mn0.01S ­Zn0.99Mn0.01S@SiO2 samples against Gram negative Pseudomonas
QDs; b ZnS@ZnO; c ­Zn0.99Mn0.01S@ZnO; and d ­Zn0.99Mn0.01S@ aeruginosa
SiO2 samples against Gram negative Klebsiella pneumoniae; e ZnS

of ­SiO2 nanocomposites exhibiting enhanced antibacte- be broadly classified into physical, biological or chemical
rial activity [14, 53, 54] ­Zn0.99Mn0.01S@SiO2 samples also mechanisms depending on the process behind it. Physical
exhibit markedly better antibacterial activity than previously processes involve the penetration of the outer cell membrane
investigated ZnS@SiO2 nanostructures, which exhibited a of the bacteria, causing a disturbance in the cell permeability
maximum inhibition zone of 12 mm and 10 mm against B. due to cell wall decomposition, which can lead to leakage
subtilis and Escherichia coli [18]. In conclusion, an appreci- of proteins, minerals and genetic material, and eventual cell
able enhancement in antibacterial activity was observed in death. These physical processes and the antibacterial activ-
­Zn0.99Mn0.01S@ZnO samples as compared to ZnS@SiO2. ity caused by them depend on the nanoparticle size, surface
There is no common consensus, as of yet, on a single morphology and surface area, and increase with surface-to-
mechanism to explain the antibacterial activity of nanoparti- volume ratio [14, 55]. Biological processes suppress protein
cles, and the different possible mechanisms that can explain metabolism, inhibit active transport, and disturb the enzy-
the antibacterial activity of ZnS and ZnO nanoparticles can matic system of the bacteria to cause cell death. These may

13
169 Page 14 of 17 S. Kumar et al.

be facilitated by the binding or adsorption between the nano-

­ n0.99Mn0.01S@SiO2 samples, and the positive and negative controls against six human pathogenic bac-

(10 µg/disc)
+ve control
Gentamicin
particle surface and the bacteria outer membrane caused by
electrostatic or Van der Waals forces, or by the release of
­Zn2+ and direct cellular internalisation [14, 55, 56].

22
22

24
21
21
22
Tetracycline
The incorporation of ZnO in our prepared nanostructures
(30 µg/disc) suggests the possibility of chemical processes, that involve
+ve control the generation of reactive oxygen species (ROS) [57, 58]
like hydrogen peroxide and hydroxyl radicals that oxidize

15
16

16
17
17
16
cellular proteins and lipids and cause damage to the cel-
lular genetic material, leading to cell death. Nanoparticles
−ve control

exhibit defect energy levels in the forbidden band gap that


Zn0.99Mn0.01S@SiO2 (x mL) Ethanol

increase the capacity of transporting charge carriers to the


nanoparticle surface and transferring that energy to a nearby
0
0

0
0
0
0
oxygen molecule leading to the generation of ROS [14, 59].
1.0

17
12

11
14
11
10
4 Conclusion
0.75

14
10

11
12
10
10

In the present work, ZnS and Z ­ n0.99Mn0.01S QDs were syn-


0.5

10
11

10
10
10
11

thesised using the co-precipitation method, and used as


seed material for the subsequent synthesis of ZnS@ZnO,
0.25

­Zn0.99Mn0.01S@ZnO and ­Zn0.99Mn0.01S@SiO2 nanostruc-


10
10

10
10
10
10

tures, synthesised in separate wet chemical processes with


Zn0.99Mn0.01S@ZnO (x mL)

100

increasing amounts of Zn(NO3)2 and TEOS, employed as


14
11

16
10
11
11

the precursors for ZnO and S ­ iO2 shell, respectively. Soft-


ware package MStruct was used to analyse the P-XRD pat-
50

13
12

14
12
11
11

terns of all the samples; zinc-blende phase was confirmed


for the synthesised ZnS and Z ­ n 0.99 Mn 0.01 S QDs. The
30

11
11

13
10
12
12

ZnS@ZnO and ­Z n 0.99Mn 0.01S@ZnO samples exhibited


Table 2  Inhibition zone diameters of the prepared ZnS@ZnO, ­Zn0.99Mn0.01S@ZnO, and Z

increasing intensity of wurtzite ZnO phase with increas-


10

10
10

12
12
11
14

ing Zn(NO3)2 used during synthesis; Z ­ n0.99Mn0.01S@SiO2


samples exhibited the same phase as the seed zinc-blende
Zn0.99Mn0.01S

­Z n0.99Mn0.01S. TE micrographs confirmed the spherical


shape of the synthesised ZnS and ­Zn0.99Mn0.01S QDs, and
the wurtzite phase of the ZnS@ZnO(100 mL) sample.
12

12
22
17
14

FTIR spectra of the prepared samples exhibited a number


8

of characteristic peaks; peak at around 1340 ­cm−1 in the


100

11
12

11
11
11
12

FTIR spectra of the synthesised ZnS and Z ­ n0.99Mn0.01S


QDs is attributed to the C–C doublet peak, whose absence
in the FTIR spectra of the prepared ZnS@ZnO and
ZnS@ZnO (x mL)

50

12
12

11
10
12
10

­Zn0.99Mn0.01S@ZnO nanostructures could be an indica-


tion of surface modification of the seed material by ZnO.
30

12
11

12
11
11
12

The synthesised Z ­ n0.99Mn 0.01S QDs also exhibit a peak


at around 1106 ­cm−1 due to the doping of ­Mn2+ ions in
10

13
10

12
10
10
10

the host ZnS lattice, which is quenched along with the


characteristic Zn–S stretching peak 617 ­cm−1 in the FTIR
ZnS

10

14
24
12
14

spectra of the prepared Z ­ n0.99Mn0.01S@SiO2 nanostruc-


8

ture. UV–visible absorption spectra of the synthesised


Salmonella Typhi

ZnS and Z ­ n 0.99Mn 0.01S QDs exhibited strong blue-shift


K. pneumoniae
Inhibition Zone
Diameter (mm)

P. aeruginosa
Gram negative
Bacteria strain

with respect to the absorption edge of bulk ZnS; energy


Gram positive
terial strains

B. subtilis
S. aureus

band gap value (~ 3.9 eV) obtained from the Tauc rela-
E. coli

tion and particle size (~ 2.2 nm) obtained from the Brus
equation confirmed the strong size confinement of the

13
Antibacterial studies of ZnO and silica capped manganese doped zinc sulphide nanostructures Page 15 of 17 169

Fig. 16  Histogram illustrating the antibacterial activity of the prepared: a ZnS@ZnO; b ­Zn0.99Mn0.01S@ZnO; c ­Zn0.99Mn0.01S@SiO2 samples;
and d positive control(s) against six human pathogenic bacterial strains

synthesised ZnS and Z ­ n0.99Mn0.01S QDs. Band gap val- Acknowledgements S.K., A.J. and S.P. are thankful to the Department
ues generally decreased (and corresponding particle sizes of Science (DST), New Delhi, India for supporting part of this research
work (vide Project No. SR/FTP/PS-69/2008), dated 15/1/2010. Ravi
increased) with an increase in the amount of Zn(NO3) 2 Kant Choubey is thankful to the Council of Science & Technology,
and TEOS used during the synthesis of the ZnS@ZnO, Lucknow, Uttar Pradesh, India for the financial support (Vide No.
­Zn0.99Mn0.01S@ZnO, and Z ­ n0.99Mn0.01S@SiO2 samples. CST/4051).
PLE spectra of the synthesised ZnS and ­Zn0.99Mn0.01S QDs
Author contributions All the authors contributed to the present work,
exhibited surface defect mediated emission bands, which conception and design, Material preparation, data collection and analy-
was increasingly quenched with an increase in the ZnO or sis were performed by the authors. The author’s contribution is shown
­SiO 2 shell thickness. Characteristic orange emission of below. The first draft of the manuscript was written by SK, RKC and
the dopant ­Mn2+ 4T1 → 6A1 transition in the synthesised SG and all the authors commented on the previous version of the manu-
script. All the authors read and approved the final manuscript.
­Zn0.99Mn0.01S QDs is enhanced with ZnO capping, but is
increasingly quenched with increasing S ­ iO2 capping. The Availability of data and materials The data used in the manuscript can
PLE spectra not only illustrated the highly tuneable emis- be available from the corresponding author upon reasonable request.
sion properties, but also highlighted the enhanced surface
properties of the synthesised ZnS@ZnO, Z ­ n0.99Mn0.01S@ Declarations
ZnO, and ­Z n 0.99Mn 0.01S@SiO 2 nanostructures, both of Conflict of interest I hereby confirm that, the work described has not
which are incredibly useful for biomedical applications. been published before; it is not under consideration for publication
Antibacterial assay showed the incredible antibacterial anywhere else; and publication has been approved by all co-authors
potential of the synthesised ZnS and ­Zn0.99Mn0.01S against of this manuscript.
E. coli. Increasing thickness of the ZnO shell increases Research data policy and data availability statements All data gener-
the antibacterial activity of the prepared ZnS@ZnO sam- ated or analysed during this study have been deposited in this manu-
ples against S. aureus and the prepared ­Zn0.99Mn0.01S@ script. All the compared data were properly cited and included in the
ZnO samples against B. subtilis. Increasing thickness of reference section following the journal style. Also, the data will be
available from the corresponding author on reasonable request.
the ­SiO2 shell increases the antibacterial activity of the
prepared ­Zn0.99Mn0.01S@SiO2 samples against both Gram
positive bacterial strains and against E. coli.

13
169 Page 16 of 17 S. Kumar et al.

References 22. A. Jain, S. Panwar, T.W. Kang, S. Kumar, J. Mater. Sci. Mater.
Electron. 24(12), 5147–5154 (2013). https://​doi.​org/​10.​1007/​
s10854-​013-​1537-z
1. L. Zhang, J. Ma, B. Lyu, Y. Zhang, V.K. Thakur, C. Liu, Green
23. A. Jain, S. Panwar, T.W. Kang, H.C. Jeon, S. Kumar, R.K.
Chem. 23(19), 7576–7588 (2021). https://​d oi.​o rg/​1 0.​1 039/​
Choubey, J. Mater. Sci. Mater. Electron. 25(4), 1716–1723 (2014).
D1GC0​1766G
https://​doi.​org/​10.​1007/​s10854-​014-​1788-3
2. B.Y. Lu, G.Y. Zhu, C.H. Yu, G.Y. Chen, C.L. Zhang, X. Zeng,
24. M. Balouiri, M. Sadiki, S.K. Ibnsouda, J. Pharm. Anal. 6(2),
Q.M. Chen, Q. Peng, Nano Res. 14(1), 185–190 (2021). https://​
71–79 (2016). https://​doi.​org/​10.​1016/j.​jpha.​2015.​11.​005
doi.​org/​10.​1007/​s12274-​020-​3064-6
25. Z. Matej, R. Kuzel, L. Nichtova, Powder Diffr. 25(2), 125–131
3. Y. Gao, Y. Dong, S. Yang, A. Mo, X. Zeng, Q. Chen, Q. Peng,
(2010). https://​doi.​org/​10.​1154/1.​33923​71
J. Colloid Interface Sci. 617, 533–541 (2022). https://​doi.​org/​
26. Z. Matej, A. Kadlecova, M. Janecek, L. Matejova, M. Dopita, R.
10.​1016/j.​jcis.​2022.​03.​032
Kuzel, Powder Diffr. 29(S2), S35–S41 (2014). https://​doi.​org/​10.​
4. C. Yu, S. Sui, X. Yu, W. Huang, Y. Wu, X. Zeng, Q. Chen, J.
1017/​S0885​71561​40008​52
Wang, Q. Peng, Colloids Surf. B Biointerfaces 217, 112663
27. MStruct-software for Micro Structure analysis by powder diffrac-
(2022). https://​doi.​org/​10.​1016/j.​colsu​r fb.​2022.​112663
tion (xray.cz/mstruct)
5. M.M. Modena, B. Rühle, T.P. Burg, S. Wuttke, Adv. Mater.
28. V. Favre-Nicolin, R. Cerny, J. Appl. Cryst. 35(6), 734–743 (2002).
31(32), 1901556 (2019). https://​doi.​org/​10.​1002/​adma.​20190​
https://​doi.​org/​10.​1107/​S0021​88980​20152​36
1556
29. FoxWiki-FOX, Free Objects for Crystallography Wiki (fox.
6. A. Mohajerani, L. Burnett, J.V. Smith, H. Kurmus, J. Milas, A.
vincefn.net)
Arulrajah, A. Horpibulsuk, A. Abdul Kadir, Materials 12(19),
30. P. Scardi, M. Leoni, Acta Cryst. A58(2), 190–200 (2002). https://​
3052 (2019). https://​doi.​org/​10.​3390/​ma121​93052
doi.​org/​10.​1107/​S0108​76730​10212​98
7. S. Shrestha, B. Wang, P. Dutta, Adv. Colloid Interface Sci. 279,
31. G. Ghosh, M.K. Naskar, A. Patra, M. Chatterjee, Opt. Mater.
102162 (2020). https://​doi.​org/​10.​1016/j.​cis.​2020.​102162
28(8–9), 1047–1053 (2006). https://​doi.​org/​10.​1016/j.​optmat.​
8. P.L. Suarez, M. Garcia-Cortes, M.T. Fernandez-Arguelles, J.R.
2005.​06.​003
Encinar, M. Valledor, F.J. Ferrero, J.C. Campo, J.M. Costa-
32. B.S. Rema Devi, R. Raveendran, A.V. Vaidyan, Pramana 68(4),
Fernández, Anal. Chim. Acta 1046, 16–31 (2019). https://​doi.​
679–687 (2007). https://​doi.​org/​10.​1007/​s12043-​007-​0068-7
org/​10.​1016/j.​aca.​2018.​08.​018
33. N.V. Hullavarad, S.S. Hullavarad, J. Vac. Sci. Technol. A 26(4),
9. Y. Hui, X. Yi, F. Hou, D. Wibowo, F. Zhang, D. Zhao, H. Gao,
1050–1057 (2008). https://​doi.​org/​10.​1116/1.​29403​46
C.X. Zhao, ACS Nano 13(7), 7410–7424 (2019). https://​doi.​org/​
34. M. Sharma, S. Kumar, O.P. Pandey, J. Nanopart. Res. 12(7),
10.​1021/​acsna​no.​9b039​24
2655–2666 (2010). https://​doi.​org/​10.​1007/​s11051-​009-​9844-2
10. C. Bankier, R.K. Matharu, Y.K. Cheong, G.G. Ren, E. Clout-
35. A.K. Thottoli, A.K.A. Unni, J. Nanostruct. Chem. 3(1), 1–12
man-Green, L. Ciric, Sci. Rep. 9(1), 1–8 (2019). https://​doi.​org/​
(2013). https://​doi.​org/​10.​1186/​2193-​8865-3-​56
10.​1038/​s41598-​019-​52473-2
36. I. Ahemen, O. Meludu, E. Odoh, Br. J. Appl. Sci. Technol. 3(4),
11. Y. Jing, B. Mu, M. Zhang, L. Wang, H. Zhong, X. Liu, A. Wang,
1228 (2013)
Colloids Surf. A Physicochem. Eng. Asp. 600, 124965 (2020).
37. S. Muthukumaran, R. Gopalakrishnan, Opt. Mater. 34(11), 1946–
https://​doi.​org/​10.​1016/j.​colsu​r fa.​2020.​124965
1953 (2012). https://​doi.​org/​10.​1016/j.​optmat.​2012.​06.​004
12. Y. Gao, Y. Chen, Y. Cao, A. Mo, Q. Peng, Eur. J. Med. Chem.
38. K. Raja, P.S. Ramesh, D. Geetha, Spectrochim. Acta A Mol. Bio-
213, 113056 (2021). https://​d oi.​o rg/​1 0.​1 016/j.​e jmech.​2 020.​
mol Spectrosc. 131, 183–188 (2014). https://​doi.​org/​10.​1016/j.​
113056
saa.​2014.​03.​047
13. G. Panthi, R. Ranjit, S. Khadka, K.R. Gyawali, H.Y. Kim, M.
39. Y. Wang, B. Wu, C. Yang, M. Liu, T.C. Sum, K.T. Yong, Small
Park, Adv. Compos. Hybrid Mater. 3(1), 8–15 (2020). https://​
12(4), 534–546 (2016). https://​doi.​org/​10.​1002/​smll.​20150​3352
doi.​org/​10.​1007/​s42114-​020-​00141-9
40. Y. Hu, Z. Wei, B. Wu, B. Shen, Q. Dai, P. Feng, AIP Adv. 8(1),
14. A. Ananth, I. Han, M. Akter, J.H. Boo, E.H. Choi, J. Ind. Eng.
15014 (2018). https://​doi.​org/​10.​1063/1.​50108​33
Chem. 90, 389–398 (2020). https://​doi.​org/​10.​1016/j.​jiec.​2020.​
41. Y. Hu, B. Hu, B. Wu, Z. Wei, J. Li, J. Mater. Sci. Mater. Elec-
07.​042
tron. 29(19), 16715–16720 (2018). https://​d oi.​o rg/​1 0.​1 007/​
15. S. Kokilavani, S.A. Al-Farraj, A.M. Thomas, H.A. El-Serehy,
s10854-​018-​9764-y
L.L. Raju, S.S. Khan, Ceram. Int. 47(9), 12997–13006 (2021).
42. J. Tauc, A. Menth, D.L. Wood, Phys. Rev. Lett. 25(11), 749
https://​doi.​org/​10.​1016/j.​ceram​int.​2021.​01.​163
(1970). https://​doi.​org/​10.​1103/​PhysR​evLett.​25.​749
16. D.S. Bharathi, A. Boopathyraja, S. Nachimuthu, K. Kannan,
43. J. Tauc, A. Menth, J. Non. Cryst. Solids 8, 569–585 (1972).
J. Clust. Sci. 33, 2499–2515 (2021). https://​doi.​org/​10.​1007/​
https://​doi.​org/​10.​1016/​0022-​3093(72)​90194-9
s10876-​021-​02170-w
44. A. Kumar, S. Mukherjee, H. Sharma, D.K. Rana, A. Kumar,
17. S. Kumar, K.H.S. Bhatti, K. Singh, S. Gupta, S. Sharma, V.
R. Kumar, R.K. Choubey, Mater. Sci. Semicond. Process. 155,
Kumar, R.K. Choubey, Phys. Scripta 96, 125807 (2021). https://​
107226 (2023). https://​doi.​org/​10.​1016/j.​mssp.​2022.​107226
doi.​org/​10.​1088/​1402-​4896/​ac1eb3
45. A. Kumar, M. Kumar, V. Bhatt, S. Mukherjee, S. Kumar, H.
18. S. Kumar, A. Jain, S. Panwar, I. Sharma, H.C. Jeon, T.W. Kang,
Sharma, M.K. Yadav, S. Tomar, J.H. Yun, R.K. Choubey, Sens.
R.K. Choubey, Int. J. Appl. Ceram. Technol. 16(2), 531–540
Actuators A. Phys. 331, 112988 (2021). https://​doi.​org/​10.​1016/j.​
(2019). https://​doi.​org/​10.​1111/​ijac.​13145
sna.​2021.​112988
19. S. Tomar, S. Gupta, S. Mukherjee, A. Singh, S. Kumar, R.K.
46. L.E. Brus, J. Chem. Phys. 79(11), 5566–5571 (1983). https://​doi.​
Choubey, Semiconductors 54(11), 1450–1458 (2020). https://​
org/​10.​1063/1.​445676
doi.​org/​10.​1134/​S1063​78262​01102​4X
47. L.E. Brus, J. Chem. Phys. 80(9), 4403–4409 (1984). https://​doi.​
20. S. Tomar, S. Gupta, S. Mukherjee, A. Singh, S. Kumar, V.
org/​10.​1063/1.​447218
Kumar, R.K. Choubey, Phys. Scr. 96(6), 65802 (2021). https://​
48. S.K. Mishra, R.K. Srivastava, S.G. Prakash, R.S. Yadav, A.C.
doi.​org/​10.​1088/​1402-​4896/​abed7e
Panday, Optp-Electron. Rev. 18(4), 467–473 (2010). https://​doi.​
21. S. Gupta, R.K. Choubey, L.K. Sharma, M.P. Ghosh, M. Kar, S.
org/​10.​2478/​s11772-​010-​0037-4
Mukherjee, Semicond. Sci. Technol. 34(10), 105006 (2019).
49. T.K. Kundu, N. Karak, P. Barik, S. Saha, Int. J. Soft Comput. Eng.
https://​doi.​org/​10.​1088/​1361-​6641/​ab3a00
1, 19–24 (2011)

13
Antibacterial studies of ZnO and silica capped manganese doped zinc sulphide nanostructures Page 17 of 17 169

50. D. Das, P. Mondal, RSC Adv. 4(67), 35735–35743 (2014). https://​ 57. C. Hwang, M.H. Choi, H.E. Kim, S.H. Jeong, J.U. Park, NPG Asia
doi.​org/​10.​1039/​C4RA0​6063F Mater. 14, 72 (2022). https://d​ oi.o​ rg/1​ 0.1​ 038/s​ 41427-0​ 22-0​ 0420-5
51. D. Mandal, L.K. Sharma, S. Mukherjee, Appl. Phys. A 122(12), 58. V.L. Prasanna, R. Vijayaraghavan, Langmuir 31, 9155 (2015).
1–10 (2016). https://​doi.​org/​10.​1007/​s00339-​016-​0573-y https://​doi.​org/​10.​1021/​acs.​langm​uir.​5b022​66
52. R.K. Chandrakar, R.N. Baghel, V.K. Chandra, B.P. Chandra, 59. H. Labiadh, K. Lahbib, S. Hidouri, S. Touil, T.B. Chaabane, Asian
Superlattices Microstruct. 86, 256–269 (2015). https://​doi.​org/​ Pac. J. Trop. Med. 9(8), 757–762 (2016). https://d​ oi.o​ rg/1​ 0.1​ 016/j.​
10.​1016/j.​spmi.​2015.​07.​043 apjtm.​2016.​06.​008
53. Z.A. Baka, M.M. El-Zahed, Bioresour. Bioprocess. 9(1), 1–19
(2022). https://​doi.​org/​10.​1186/​s40643-​022-​00591-7 Publisher's Note Springer Nature remains neutral with regard to
54. M.M. El-Zahed, M.I. Abou-Dobara, A.K. El-Sayed, Z.A.M. Baka, jurisdictional claims in published maps and institutional affiliations.
Nova Biotechnol. Et Chim. 21(1), 1023 (2022). https://d​ oi.o​ rg/1​ 0.​
36547/​nbc.​1023 Springer Nature or its licensor (e.g. a society or other partner) holds
55. R. Kumar, P. Sakthivel, P. Mani, Appl. Phys. A 125(8), 1–12 exclusive rights to this article under a publishing agreement with the
(2019). https://​doi.​org/​10.​1007/​s00339-​019-​2823-2 author(s) or other rightsholder(s); author self-archiving of the accepted
56. S. Vijayan, C.S. Dash, G. Umadevi, M. Sundararajan, R. Mari- manuscript version of this article is solely governed by the terms of
appan, J. Clust. Sci. 32(6), 1601–1608 (2021). https://​doi.​org/​10.​ such publishing agreement and applicable law.
1007/​s10876-​020-​01923-3

13

You might also like