You are on page 1of 17

Journal Pre-proof

nMusa AAA and Jatropha curcas L. sap mediated TiO2


nanoparticles: synthesis and characterization

Agatha W. Wagutu , Kohei Yano , Kohei Sato , Eugene Park ,


Yoshiki Iso , Tetsuhiko Isobe

PII: S2468-2276(19)30764-1
DOI: https://doi.org/10.1016/j.sciaf.2019.e00203
Reference: SCIAF 203

To appear in: Scientific African

Received date: 23 March 2019


Revised date: 22 September 2019
Accepted date: 10 October 2019

Please cite this article as: Agatha W. Wagutu , Kohei Yano , Kohei Sato , Eugene Park ,
Yoshiki Iso , Tetsuhiko Isobe , nMusa AAA and Jatropha curcas L. sap mediated
TiO2 nanoparticles: synthesis and characterization, Scientific African (2019), doi:
https://doi.org/10.1016/j.sciaf.2019.e00203

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2019 Published by Elsevier B.V. on behalf of African Institute of Mathematical Sciences / Next
Einstein Initiative.
This is an open access article under the CC BY-NC-ND license.
(http://creativecommons.org/licenses/by-nc-nd/4.0/)
Highlights

 TiO2 nanoparticles are biosynthesized using Musa AAA and Jatropha sap

 Raman spectroscopy and X-Ray diffraction confirm formation of anatase Phase

 Presence of organic molecules capping nanoparticles is identified with FT-IR and


Raman
Musa AAA and Jatropha curcas L. sap mediated TiO2 nanoparticles: synthesis
and characterization
Agatha W. Wagutu*1, 2, Kohei Yano3, Kohei Sato3 Eugene Park2, Yoshiki Iso3, Tetsuhiko Isobe3,Email:
awagutu@gmail.com
1
Department of Pure and Applied Sciences, Kirinyaga University, P. O. Box 134-10300 Kerugoya, Kenya
2
Department of Materials and Energy Science and Engineering (MESE), Nelson Mandela African Institution of
Science and Technology (NM-AIST), P.O.BOX 447, Arusha, Tanzania
3
Department of Applied Chemistry, Faculty of Science and Technology, Keio University, 3-14-1 Hiyoshi,
Yokohama 223-8522, Japan

ABSTRACT
Titanium dioxide (TiO2) nanoparticles were synthesized using East African highland banana (Musa
AAA) and Jatropha curcas L. sap. The nanoparticles were characterized with X-ray diffractometry
(XRD), Raman spectroscopy, Fourier transform infrared (FT-IR) spectroscopy and scanning electron
microscopy (SEM). FT-IR peaks show presence of amino, carboxyl and hydroxyl groups on the
crystal surface of the nanoparticles. XRD and Raman characteristic indicate anatase phases. The
average crystallite size calculated from XRD peak width using Scherrer equation are 5.5 nm and 6.1
nm for Jatropha and Musa AAA sap TiO2 particles respectively. SEM images show polydispersed
spherical clusters of agglomerated nanoparticles. The nanoparticles showed antimicrobial activity
against gram-negative bacteria when irradiated in natural sunlight for 2 h. The sap extracts provide a
simple, cheaper and more environmentally benign reducing and stabilizing agents for transforming
transition metal salts into nano-scale particles, which have wide range of industrial applications.

Key words: Biosynthesis; Fourier transform infrared; Musa AAA; Raman spectroscopy; TiO2
nanoparticles; X-ray diffraction

1.0 Introduction
The growing field of nanotechnology involves development and characterization of functional
materials at nano-scale. Properties of such materials are significantly enhanced compared to
their bulk size, even though they have same crystal structure [1]. Such properties include
electrical, optical characteristics, surface area, structural orientation and chemical reactivity
[2]. Nanoparticles of different inorganic materials have been envisaged to provide sustainable
solutions in areas of catalysis, medicine, water treatment, electronic and energy conversion
[3]. Due to the many applications, their demand is anticipated to grow exponentially in the
near future. Manufacture and assembly of such large quantities of nano-sized functional
materials calls for shift from the conventional chemical routes requiring utilization of toxic
solvents, corrosive concentrated acids and alkaline media to aqueous based greener routes [3].
Interaction between biological entities and inorganic materials has received significant
attention in biosynthesis of nanoparticles. Maaza and co-workers have demonstrated efficient
bio-mediated synthesis of various metal oxide nanoparticles such as CdO, ZnO, CeO2, Sm2O3,
SnO2 and PbO using strictly plant extract as reducing, capping and stabilizing agents in
aqueous media without addition of hazardous and expensive chemicals [4-9]. The inherent
antioxidant and stabilizing properties of biological extracts offers a bottom up approach

2
similar to chemical reduction of metal ions, but more environmental friendly and economical
[10, 11].

TiO2 is among of the most studied transition metal oxide, possibly because of its unique
optoelectrical properties, photocatalytic ability, biocompatibility, lower toxicity, chemical
inertness and mechanical stability. These traits make TiO2 very versatile for wide range of
applications related to science and development [12]. In bulk form, TiO2 is mostly used in
paints and coatings, plastics, papers, inks, medicines, pharmaceuticals, food products,
cosmetics and toothpaste due to its brightness, high refractive index and resistance to
discoloration [13]. Various investigations have however, established that TiO2 is more
effective in its nanoparticle size rather than the bulk crystalline form [14]. Several physical
and chemical methods have been developed for synthesis of TiO2 nanoparticle [15]. These
include; sol-gel techniques [16], micro-emulsion [17], precipitation [18], hydrothermal
methods [19], solvothermal methods [20] and electrochemical processes [21]. However, most
of these methods are highly polluting, energy intensive and very expensive.

Biological methods have thus been explored as alternative greener routes involving
application of natural plant extracts or microbes in reduction of Ti4+ and stabilizing the zero
valence nanoparticles formed [22]. Use of plant extract is more favorable than using microbes
because it does not require isolation and culturing of the bacteria or fungi. The challenge with
this route for titanium and other transition metals like niobium, tantalum, silicon and bismuth
is the absence of readily and stable water soluble salts precursors that can easily provide
cations for bioreduction [3]. Titanium salts mostly studied in biosynthesis include, titanium
(IV) chloride (TiCl4) titanium (IV) oxo sulfate (TiOSO4) and titanium (IV) tetraisopropoxide
(TTIP), all which readily hydrolyze in water at room temperature releasing toxic fumes and
precipitates, thus requiring acid stabilization in aqueous solution [3]. Few works have
reported on pure bioreduction of simple water insoluble bulk TiO2 to yield much-valued
nanoparticles. More research is therefore needed to establish the ability of different biological
extracts to synthesize TiO2 nanoparticles from abundant and stable bulk TiO2 precursor.
Various plants extracts reported in literature for biosynthesis of TiO2 nanoparticles are listed
in Table 1, together with Ti precursor used.

Table 1: Plants extracts reported in green synthesis of TiO2 nanoparticles

SN Plant extract Precursor Crstallite Crystal Use Ref


Used size (nm) Phase
1 Trigonella foenum graceum leaf TiOSO4 20-90 anatase antimicrobial [23]
2 Diospyros ebenum leaf TTIP 24-33 anatase catalysis/antibacterial [24]
3 Water soluble starch TTIP 9-15 anatase catalysis [25]
4 Euphorbia heteradena Jaub root TiO(OH)2 20 rutile - [26]
5 Jatropha curcas L. leaf TiCl4 13 anatase catalysis [27]
6 Jatropha curcas L. Latex TiO(OH)2 20-50 anatase - [28]
7 Cicer arietinum L. waste water TiCl4 13-24 anatase/ Li-ion battery [29]
rutile
8 Justicia gendrarussa leaf TTIP - anatase catalysis/ Anticancer [30]
9 Lippia citriodora leaf TiCl4 20-40 anatase/ anti-brain injury [31]
rutile
10 Tamarindus indica leaf TTIP 20-40 - catalysis [32]

3
The active metabolites in the plants acts as biogenic reducing and capping agents [33]. The choice
of the plant extract may be informed by phytochemical characteristic of the extract, availability,
sustainability and other beneficial properties such as medicinal and antimicrobial activity. This work
reports biosynthesis of TiO2 nanoparticles using green banana (Musa AAA) and Jatropha curcas L.
sap. The plants are locally abundant in Northern Tanzania. Their sap is characterized by high reducing
(antioxidant) ability attributed to presence of major bioactive compounds reported in Table 2.

Table 2: Major photochemical compound of Jatropha Curcas and Musa AAA sap

Jatropha Curcas sap [34-36] Banana (Musa AAA) sap [37-40]


Curcian (proteolytic enzyme) Dopamine
Curcacyline A (cyclic octapeptide) Apigenin glycoside I –flavone
Curcacycline B (cyclic nonapeptide) Apigenin glycoside 2- flavone
Myricetin-3-O- ritinoside -flavanol
Quecertin-3-O- rutinoside – flavanol

2.0 Experimental

2.1 Preparation of TiO2 nanoparticles


Analytical grade bulk titanium (IV) oxide monohydrate (H2TiO3) was purchased from Mitsuwa
Chemicals. Musa AAA sap was obtained from fresh green bananas purchased at the local market
(Tengeru) in Northern Tanzania, 3° 22’ 66‖ S and 36° 49’ 60‖ E, with elevation 1169 m above sea
level. The variety of banana selected is known to produce the highest quantity of sap when fresh and is
called ―Ndizi Ng’ombe‖ in local language. The bananas were washed and dried to remove dirt, tips
(both ends) were then cut and the oozing sap collected in a small clean 15 ml falcon tube. Jatropha sap
was obtained from life fence of Jatropha curcas L. plants around the same market place. Sap was
collected as fluid exudate from freshly cut stalk leaves and young stems.

TiO2 nanoparticles were synthesis following procedure reported in [26, 41] with slight modification.
For each plant’s sap, exactly 4 ml were diluted to 400 ml of solution with distilled water. The solution
was filtered to remove any particles, then 2.0 g of the bulk H2TiO3 powder was dispersed in the
solution while stirring continuously at 60 °C for 3 h. The mixture was then cooled at room temperature
for 12 h to allow particle growth. The precipitate deposited at the bottom of reaction flask was filtered
under vacuum using Whatman filter paper (GF/A, 55 mm), then rinsed with distilled water and oven
dried at 80 °C for 24 h. The powdered products obtained were packed in sample bottles for further
analysis.

2.2 Characterization of TiO2

All characterization was done at the Department of Applied Chemistry, Faculty of Science and
Technology, Keio University in Japan. Crystallographic data were determined from XRD patterns of
the samples obtained using Rigaku Rint-2200 X-ray diffractometer with monochromatized CuKα
radiation (λ= 0.154 nm) over 2θ range of 10 – 60° with a step size of 0.05°, beam voltage and current
at 30 kV and 40 mA, respectively. Raman vibrational analysis was performed using a Renishaw inVia
spectrometer. The power of the laser at 0.1%, laser excitation 532 nm and the exposure time of 30 s.
Surface morphology was examined using JEOL JSM-7600F and FEI Inspect F50 field emission
4
scanning electron microscopes Fourier transform infrared (FT-IR) analysis was conducted using a
JASCO-4200 FTIR spectrometer in absorbance mode with spectral range of 4000-400 cm-1.

2.3 Evaluation of Microbial activity


Musa AAA-TiO2 nanoparticles were tested for photocatalytic purification of water
contaminated with E.coli bacteria. Membrane filtration and culturing in HiCrome media
described in literature [42] was used for bacteria detection and enumeration, with slight
modification. Water samples were collected from local stream; Physical properties were
tested at the point of collection. Microbial analysis before and after treatment was done within
12 h. Approximately 30 mg of nanoparticles were dispersed in 150 ml of raw water. Mixture
was stirred gently under natural sunlight (solar irradiance, 800 Wm-1 at 26 ° C) for 3 h.
Samples of 50 ml were drawn hourly, filtered through 0.45 micron membrane filters (47 mm
diameter) under vacuum. The filters were then placed petri dishes containing the culture
media and incubated for 18 h at 37.0 °C. Colonies forming units (CFU) were counted
manually. Another set of experiment was kept in dark under similar treatment as an
unexposed control.
3.0 Results and Discussion

3.1 Preparation of sap coated TiO2

White colored H2TiO3 powder reacted with sap solution (initial pH = 4.5) to form a yellowish
mixture which turned orange after short time. Precipitate recovered after filtration and drying
retained some orange color.

3.2 XRD analysis


XRD analysis show that biosynthesized particles exhibited broad diffraction pattern (Fig. 1)
characteristic to nanocrystalline material with very small crystallite sizes [43]. Peak analyzer
function within Origin Pro 2015 data analysis software (OriginLab, USA) [44] was used for
peak location, peak fitting into Gaussian distribution and determination of full width at the
half maximum intensity (FWHM), (Table 3). Plane orientations of identified peaks are; (101),
(004), (200) and (211) which occurred at 2θ values 25.4°, 38.0°, 48.0° and 54.7° respectively.
Principle peak (101) at 2θ = 25.4° for both samples confirmed formation of anatase phase
(ICCD: 01-084-1285).The calculated lattice constants a = b ≠ c for tetragonal crystal
structure; a = 3.78 Å and c = 9.27 Å are in agreement with reported value for anatase TiO2 (
a = 3.73 Å and c = 9.37 Å) [12].

5
Fig 1. X-ray diffraction patterns of biosynthesized Musa AAA sap capped and Jatropha sap capped TiO2
nanoparticles.

The crystallite sizes were calculated using the Scherrer’s equation; L=0.9 λ/ βcosθ. Where L is the
average size of the crystallite, λ is the wavelength of X-ray radiation, β is FWHM and θ is the angle of
diffraction. The calculated average crystallite sizes are 5.5 and 6.1 nm for Jatropha and Musa AAA sap
TiO2 respectively. The sizes are slightly lower than literature value, 7 nm obtained using Hibiscus
flower extract [45].

Table 3: XRD peak analysis for Jatropha and Musa AAA sap capped TiO2 nanoparticles (CuKα1 λ = 1.540 Å, β
= 0.9)

2θ peak hkl d spacing β (FWHM) βcosθ crystallite size ( L)


maxi (°) (Å) radians (Å)
JC sap-TiO2
25.39 101 3.50 0.02517 0.02455 56.5
38.01 004 2.37 0.02720 0.02572 53.9
48.05 200 1.89 0.02497 0.02281 60.8
54.69 211 1.68 0.03252 0.02890 48.0
AVE 2.36±0.8 54.8±5
Musa sap-TiO2
25.34 101 3.51 0.02328 0.02271 61.0
37.98 004 2.37 0.02166 0.02048 67.7
48.00 200 1.89 0.02265 0.02069 67.0
54.68 211 1.68 0.03172 0.02818 49.2
AVE 2.36±0.8 61.2±7

3.3. Raman analysis


Raman scattering for the biosynthesized nanoparticles (Fig. 2) was measured to verify the crystalline
phase of TiO2 and identify organic phases originating from sap extracts. Peaks were fitted by
Lorenzian function to determine FWHM. Intense peak at 152 cm-1 and minor peaks at 404, 524 and
640 cm-1 match well with the characteristic peaks of bulk anatase TiO2 (Table 4). Prominent broad
peaks at 1370 cm-1 for Musa-NP and 1594 cm-1 for JC-Np correspond to D-bands for disordered
graphitic lattice and G-band for graphitic sp2 hybridized bond vibrations, respectively [46, 47].
Absence of G-band in Musa AAA indicated presence of dominantly aromatic rings capping the
nanoparticles. Low intensity D-bands observed in JC–NP (D1214, D1319 and D1428) in addition to

6
G-band is presumed to indicate presence of sp2 sites of both olefinic and aromatic origins on surface of
nanoparticle [48].
Table 4: Raman shift in cm-1 and FWHM (in brackets) of biosynthesized TiO2 nanoparticles compared with
bulk TiO2 in anatase crystalline phase [49]

Phase of peak Eg B1g A1g or B1g Eg A1g - D E2g - G

Reference-TiO2 144.0 (8.6) 396 (18.34) 514 (21.96) 637 (23.36) - -


bulk

Musa sap –TiO2 151.9 (21.0) 403.2 (26.8) 522.8 (36.2) 640.5 (43.1) 1370 -

JC sap-TiO2 152.0 (21.8) 404.6 (27.2) 523.9 (37.3) 639.6 (44.4) 1428 1595

Raman principle peak for anatase TiO2 which is usually very intense and sharp at 144 cm-1 [50] [51],
shifted to longer wavenumbers (152 cm-1) and is well resolved for the biosynthesized samples. Similar
observations were made for minor peaks involving O-Ti-O vibrational modes. FWHM increased
significantly for the sap-processed samples, indicating peak broadening. This can be attributed to
phonon confinement in the nano-sized particles [52] and also strain produced by compressive stress on
atoms by organic molecules, evidenced by presence of D and G-bands, on the nanoparticle surface,
causing them to pack more closely. This phenomenon has been identified earlier by Xu et al, when
they studied the possible causes of blue shift of Raman bands of TiO2 nanoparticles coated with
different organic molecules [53]. They observed larger blue shift and FWHM at the Eg mode for
dodecylbenzenesulfonic acid (DBS) capped nanoparticles compared to stearic acid (St) capped
nanoparticles of the same size (8 nm). They attributed the difference to larger compressive forces in
DBS capped nanoparticles compared to St capped nanoparticles.

7
Fig 2. Raman spectra of biosynthesized Musa AAA and Jatropha sap capped TiO2 nanoparticles (200-2000
cm-1)

3.4 SEM Analysis


The SEM images with magnifcation of 100,000 and 1,000,000 times respecively are shown in Fig. 3.
The nanoparticles show spherical clusters of polydispersed aggregated and individual nanoparticles.
This observation agrees with results for TiO2 nanoparticles capped with Eclipta prostrata leaf extract
[54].

Fig 3. FESEM images of Musa AAA and Jatropha sap synthesized TiO2

3.5 FT-IR analysis


FT-IR analysis (Fig. 4) was performed to identify fundamental organic functional groups capping the
TiO2 nanoparticles. The absorption peaks identified in both Jatropha and Musa AAA samples are very
close and agree with those reported in literature denoting presence of OH, C-O, carbonyl and amino
groups from peptide, dopamine, flavanones and flavanol molecules found in the plants sap [37] [28].
Major absorption frequencies are shown in Table 5 in comparison with peaks of bacteria synthesised
TiO2 nanoparticles [55].

8
Fig 4. FT-IR spectra of Musa AAA and Jatropha curcas sap capped TiO2 nanoparticles.

The broad, high intensity peak at 530-500 cm-1 is attributed to vibration of Ti-O bond in the TiO2
lattice [45]. Minor peak at 1250-1000 cm-1 is assigned to the stretching mode of C-O and C-N bonds
from esters, amines and aromatic compound present in the plant extracts. Peak at 1650-1600 cm-1 is
associated with the bending mode of O-H bond and absorption at 1700 cm-1 is assigned to the
stretching mode of C=O group [56]. Another minor peak between 1400-1300 cm-1 is assigned to COO
symmetrical stretching [57]. Absorption at 2350 cm-1 originates from CO2 in air. The strong peak at
around 3200 cm-1 stands for presence of amide groups (NH+ triple bond, NH2+, C=NH+) or medium
stretching of N-H bonds [58]. The high broad intensity peak at 3500-3300 cm-1 is recognized as
stretching of O-H groups present in phenols and alcohols in the sap extracts. In views of identified
functional groups in FT-IR spectrum, the presence of D-bands in the Raman spectra and major
bioactive components in plants sap (Table1), mechanism for bioreduction Ti4+ by Musa AAA sap is
proposed (Fig. 5). It was presumed that, one of the lone pairs from oxygen of H2TiO3 takes up a
proton from OH functional group of the organic molecules to form protonated moiety, [H3Ti-O3]+ and
proton defficient organic molecule. The protonated moiety then rearranges by loss of water molecule
to form [Ti-OOH]3+. The negatively charged organic molecules then takes a proton from [Ti-OOH]3+,
leaving TiO2 nanoparticles stabilized by organic molecules on the surface. Similar pathway for
biosynthesis of TiO2 nanoparticles are reported in literature [59, 60]. It is thus concluded that
OH functional groups were mainly responsible for reduction and other groups such as COO
and NH groups were responsible for stabilization and capping the nanoparticles.

Table 5: Main absorption peaks wavenumbers (cm-1) identified in Musa AAA and Jatropha sap capped TiO2
nanoparticles in comparison with Moraxella Osloensis bacteria capped TiO2 nanoparticles [55]

Peak No. P1 P2 P3 P4 P4 P6

Musa AAA sap –TiO2 527.4 1123.5 1405.6 1637.5 2350.4 3412.5

Jatropha sap-TiO2 523.3 1091.7 1352.3 1631.3 2362.8 3459.4

Moraxella Osloensis-TiO2 516.0 1117.0 - 1638.0 2333.0 3412.0

9
Qucertin-O-rutinoside Aspigenin glycosede
Dopamine
OH O
O H
H O NH2
O
HO O Sugar

O O
OH O Rurinoside OH O
H OH

HO
Ti O

O H

HO + H
Ti O
O H

- H 2O

OH
O- OH NH2

HO O
-O
Ti3+
O
OH O OH
Rurinoside
O

OH NH2
O
H
HO O TiO2 -NPs HO

O OH
OH O Rurinoside

Fig 5. Proposed mechanism for bio-reduction of Ti 4+ by Musa AAA sap

3.6 Antimicrobial analysis


The presence of E.coli in drinking water indicates faecal contamination and potential risk of
dangerous pathogens such as salmonella, pseudomonas, viruses and intestinal parasites [61].
TiO2 nanoparticles synthesized using Musa AAA sap were tested for phototoxicity against
E.coli bacteria in water samples. The physical properties of water analyzed were as follows,
10
pH (7.5), dissolved oxygen (5.3 ppm), electro-conductivity (812 μS/cm), total dissolved solids
(416 mg/l) and total hardness (101 mg/l of CaCO3).
Results showed that no organisms survived after 2 h exposure in natural sunlight, colonies-
forming units decreased from 68 CFU/50 ml of waters sample to zero (Fig 6). Control test
done in the dark indicated a small reduction of CFU < 10%. This was associated with
presence of organic molecules on the surface of TiO2 nanoparticles. Pure TiO2 nanoparticles
are reported to exhibit extremely low toxicity towards E.coli bacteria in the dark. This
suggests that physical damage, due to contact of TiO2 nanoparticles with the bacteria cell
wall, is not sufficient to kill cells in the dark [62, 63]. The mechanism of killing bacteria was
thus tied to oxidative damage of the cell wall, induced by reactive oxygen species (ROS) such
as hydroxyl radical and superoxide ions produced when TiO2 NPs absorbed UV/visible
radiation from the sun [64]. Also, the nano-spherical morphology of the nanoparticles could
have enhanced the cytotoxicity, following evidence presented by Tong et al, showing that
spherical morphology of anatase TiO2 NPs had higher activity against bacteria as compared to
lower dimensions such as nanotubes and nano sheets [65]. Since the gram-negative cell wall
of E.coli is more complex than gram-positive [66], it is expected that the biosynthesized
nanoparticles would also be effective in photo killing gram positive bacteria.

c d
a b

Fig 6. E.coli colonies isolated from natural water, a) before treatment, b) treatment with Musa-NPs with 1 h solar
exposure c) 2 h exposure treatment d) 3h exposure treatment

3.7 Conclusion
Overall, XRD and Raman profiles reveal that the green synthesis protocol adopted was able to
produce nanosized TiO2 particles in anatase phase. The Raman peak of Eg in the nanoparticles
coincided with that of crystalline TiO2. FT-IR spectra show particles surrounded by aromatic,
amino, carbonyl and hydroxyl functional groups, which act as capping agent/stabilizers on the
surface. This was further confirmed by presence of D and G-bands in Raman spectra
indicating presence aromatic and olefinic sites on the surface of the particles. Absence of G-
bands in Musa-NPs indicated presence of mostly aromatic molecules on the surface. This is in
agreement with bioactive molecules reported in the sap, mainly dopamine. Antimicrobial
analysis indicate that the nanoparticles are effective in photocatalytic treatment of water
contaminated by gram-negative bacteria with 100 % reduction of microbial load in two hours
when irradiated with natural sunlight. To the best of our knowledge, this is the first time that
Musa AAA sap has been considered for biosynthesis of TiO2 nanoparticles. This is particularly
important as the sap is a very common waste in food processing (especially in East and
Central Africa) and is considered a nuisance because of staining fabric, equipment and hands.
Application of its high reducing ability in the synthesis of nanoparticles is thus very
economical and efficient. The process described here is very simple, fast, cost effective and
11
environmentally benign. We have shown here that locally abundant waste plant product have
natural potential to revolutionize the nanotechnology industry at the very basic level.

Conflict of interest

All authors declare no competing interest in this work


Acknowledgements

Authors are grateful to Materials, Energy Science and Engineering laboratory (NM-AIST)
where samples were synthesized, and the Functional Materials Design laboratory (Keio
University) for sample characterization. This research did not receive any specific grant from
funding agencies in the public, commercial or not-for-profit sectors.

References

[1] C. Rao, A. Cheetham, Science and technology of nanomaterials: current status and future
prospects, Journal of Materials Chemistry, 11 (2001) 2887-2894.
[2] A.P. Alivisatos, Semiconductor clusters, nanocrystals, and quantum dots, Science 271 (1996)
933-937.
[3] M. Kakihana, M. Kobayashi, K. Tomita, V. Petrykin, Application of water-soluble titanium
complexes as precursors for synthesis of titanium-containing oxides via aqueous solution
processes, Bulletin of the Chemical Society of Japan, 83 (2010) 1285-1308.
[4] F.T. Thema, P. Beukes, A. Gurib-Fakim, M. Maaza, Green synthesis of Monteponite CdO
nanoparticles by Agathosma betulina natural extract, Journal of Alloys and Compounds, 646
(2015) 1043-1048.
[5] F.T. Thema, E. Manikandan, M.S. Dhlamini, M. Maaza, Green synthesis of ZnO nanoparticles
via Agathosma betulina natural extract, Materials Letters, 161 (2015) 124-127.
[6] N. Thovhogi, A. Diallo, A. Gurib-Fakim, M. Maaza, Nanoparticles green synthesis by
Hibiscus sabdariffa flower extract: Main physical properties, Journal of Alloys and
Compounds, 647 (2015) 392-396.
[7] B.T. Sone, E. Manikandan, A. Gurib-Fakim, M. Maaza, Sm2O3 nanoparticles green synthesis
via Callistemon viminalis' extract, Journal of Alloys and Compounds, 650 (2015) 357-362.
[8] A. Diallo, E. Manikandan, V. Rajendran, M. Maaza, Physical and enhanced photocatalytic
properties of green synthesized SnO2 nanoparticles via Aspalathus linearis, Journal of Alloys
and Compounds, 681 (2016) 561-570.
[9] A.T. Khalil, M. Ovais, I. Ullah, M. Ali, S.A. Jan, Z.K. Shinwari, M. Maaza, Bioinspired
synthesis of pure massicot phase lead oxide nanoparticles and assessment of their
biocompatibility, cytotoxicity and in-vitro biological properties, Arabian Journal of Chemistry,
(2017), https://doi.org/10.1016/j.arabjc.2017.08.009 (in press).
[10] I. Hussain, N. Singh, A. Singh, H. Singh, S. Singh, Green synthesis of nanoparticles and its
potential application, Biotechnology Letters, 38 (2016) 545-560.
[11] A.K. Mittal, Y. Chisti, U.C. Banerjee, Synthesis of metallic nanoparticles using plant extracts,
Biotechnology Advances, 31 (2013) 346-356.
[12] U. Diebold, The surface science of titanium dioxide, Surface Science Reports, 48 (2003) 53-
229.
[13] A. Weir, P. Westerhoff, L. Fabricius, K. Hristovski, N. von Goetz, Titanium dioxide
nanoparticles in food and personal care products, Environmental Science and Technology, 46
(2012) 2242-2250.
[14] H. Han and R. Bai, Buoyant photocatalyst with greatly enhanced visible-light activity prepared
through a low temperature hydrothermal method, Industrial and Engineering Chemistry
Research, 48 (2009) 2891-2898.

12
[15] S.M. Gupta, M. Tripathi, A review on the synthesis of TiO2 nanoparticles by solution route,
Central European Journal of Chemistry, 10 (2012) 279-294.
[16] Y. Bessekhouad, D. Robert, J. Weber, Preparation of TiO2 nanoparticles by sol-gel route,
International Journal of Photoenergy, 5 (2003) 153-158.
[17] V. Chhabra, V. Pillai, B. Mishra, A. Morrone, D. Shah, Synthesis, characterization, and
properties of microemulsion-mediated nanophase TiO2 particles, Langmuir, 11 (1995) 3307-
3311.
[18] H.E. Namin, H. Hashemipour, M. Ranjbar, Effect of aging and calcination on morphology and
properties of synthesized nanocrystalline TiO2, International Journal of Modern Physics B, 22
(2008) 3210-3215.
[19] N. Murakami, Y. Kurihara, T. Tsubota, T. Ohno, Shape-controlled anatase titanium(IV) oxide
particles prepared by hydrothermal treatment of peroxo titanic acid in the presence of
polyvinyl alcohol, The Journal of Physical Chemistry C, 113 (2009) 3062-3069.
[20] H.G. Yang, G. Liu, S.Z. Qiao, C.H. Sun, Y.G. Jin, S.C. Smith, J. Zou, H.M. Cheng, G.Q. Lu,
Solvothermal synthesis and photoreactivity of anatase TiO2 nanosheets with dominant {001}
facets, Journal of the American Chemical Society, 131 (2009) 4078-4083.
[21] J. Ng, J.H. Pan, D.D. Sun, Hierarchical assembly of anatase nanowhiskers and evaluation of
their photocatalytic efficiency in comparison to various one-dimensional TiO2 nanostructures,
Journal of Materials Chemistry, 21 (2011) 11844-11853.
[22] L. Castro, M.L. Blázquez, J.A. Muñoz, F. González, C. García-Balboa, A. Ballester,
Biosynthesis of gold nanowires using sugar beet pulp, Process Biochemistry, 46 (2011) 1076-
1082.
[23] S. Subhapriya, P. Gomathipriya, Green synthesis of titanium dioxide (TiO2) nanoparticles by
Trigonella foenum-graecum extract and its antimicrobial properties, Microbial Pathogenesis,
116 (2018) 215-220.
[24] S. Senthilkumar, M. Ashok, L. Kashinath, C. Sanjeeviraja, A. Rajendran, Phytosynthesis and
characterization of TiO2 canoparticles using Diospyros ebenum leaf extract and their
antibacterial and photocatalytic degradation of crystal violet, Smart Science, 6 (2018) 1-9.
[25] S. S. Muniandy, N.H. Mohd Kaus, Z.-T. Jiang, M. Altarawneh, H.L. Lee, Green synthesis of
mesoporous anatase TiO2 nanoparticles and their photocatalytic activities, RSC Advances, 7
(2017) 48083-48094.
[26] M. Nasrollahzadeh, S.M. Sajadi, Synthesis and characterization of titanium dioxide
nanoparticles using Euphorbia heteradena Jaub root extract and evaluation of their stability,
Ceramics International, 41 (2015) 14435-14439.
[27] S.P. Goutam, G. Saxena, V. Singh, A.K. Yadav, R.N. Bharagava, K.B. Thapa, Green synthesis
of TiO2 nanoparticles using leaf extract of Jatropha curcas L. for photocatalytic degradation of
tannery wastewater, Chemical Engineering Journal, 336 (2018) 386-396.
[28] M. Hudlikar, S. Joglekar, M. Dhaygude, K. Kodam, Green synthesis of TiO2 nanoparticles by
using aqueous extract of Jatropha curcas L. latex, Materials Letters, 75 (2012) 196-199.
[29] A.A. Kashale, K.P. Gattu, K. Ghule, V.H. Ingole, S. Dhanayat, R. Sharma, J.-Y. Chang, A.V.
Ghule, Biomediated green synthesis of TiO2 nanoparticles for lithium ion battery application,
Composites Part B: Engineering, 99 (2016) 297-304.
[30] S. Senthilkumar, A. Rajendran, Biosynthesis of TiO2 nanoparticles using Justicia gendarussa
leaves for photocatalytic and toxicity studies, Research on Chemical Intermediates, 44 (2018)
5923-5940.
[31] Y. Sun, S. Wang, J. Zheng, Biosynthesis of TiO2 nanoparticles and their application for
treatment of brain injury-An in-vitro toxicity study towards central nervous system, Journal of
Photochemistry and Photobiology B: Biology, (2019)1-5.
[32] S. Hiremath, A. Raj, C. Prabha, Tamarindus indica mediated biosynthesis of nano TiO2 and its
application in photocatalytic degradation of Titan yellow, Journal of Environmental Chemical
Engineering, 6 (2018) 7338-7346.
[33] H. Duan, D. Wang, Y. Li, Green chemistry for nanoparticle synthesis, Chemical Society
Reviews, 44 (2015) 5778-5792.

13
[34] S. Joglekar, K. Kodam, M. Dhaygude, M. Hudlikar, Novel route for rapid biosynthesis of lead
nanoparticles using aqueous extract of Jatropha curcas L. latex, Materials Letters, 65 (2011)
3170-3172.
[35] A. Van den Berg, S. Horsten, J. Kettenes-Van Den Bosch, B. Kroes, C. Beukelman, B.
Leeflang, R. Labadie, Curcacycline A—a novel cyclic octapeptide isolated from the latex of
Jatropha curcas L, FEBS letters, 358 (1995) 215-218.
[36] L. Nath, S. Dutta, Extraction and purification of curcain, a protease from the latex of Jatropha
curcas Linn, Journal of pharmacy and pharmacology, 43 (1991) 111-114.
[37] V. Paul, K. Kanny, G. Redhi, Formulation of a novel bio-resin from banana sap, Industrial
Crops and Products, 43 (2013) 496-505.
[38] P. Pothavorn, K. Kitdamrongsont, S. Swangpol, S. Wongniam, K. Atawongsa, J. Svasti, J.
Somana, Sap phytochemical compositions of some bananas in Thailand, Journal of
Agricultural and Food Chemistry, 58 (2010) 8782-8787.
[39] A. Pereira, M. Maraschin, Banana (Musa spp) from peel to pulp: ethnopharmacology, source
of bioactive compounds and its relevance for human health, Journal of Ethnopharmacology,
160 (2015) 149-163.
[40] C.V. Tsamo, M.-F. Herent, K. Tomekpe, T.H. Emaga, J. Quetin-Leclercq, H. Rogez, Y.
Larondelle, C. Andre, Phenolic profiling in the pulp and peel of nine plantain cultivars (Musa
sp.), Food Chemistry, 167 (2015) 197-204.
[41] R. Raliya, P. Biswas, J.C. Tarafdar, TiO2 nanoparticle biosynthesis and its physiological effect
on mung bean (Vigna radiata L.), Biotechnology Reports, 5 (2015) 22-26.
[42] G.A. Ngwa, R. Schop, S. Weir, C.G. León-Velarde, J.A. Odumeru, Detection and
enumeration of E. coli O157:H7 in water samples by culture and molecular methods, Journal
of Microbiological Methods, 92 (2013) 164-172.
[43] S. Bates, G. Zografi, D. Engers, K. Morris, K. Crowley, A. Newman, Analysis of amorphous
and nanocrystalline solids from their X-ray diffraction patterns, Pharmaceutical Research, 23
(2006) 2333-2349.
[44] E. Seifert, OriginPro 9.1: Scientific data analysis and graphing software-software review,
Journal of Chemical Information and Modeling, 54 (2014) 1552-1560.
[45] P.S. Kumar, A.P. Francis, T. Devasena, Biosynthesized and chemically synthesized titania
nanoparticles: Comparative analysis of antibacterial activity, Journal of Environmental
Nanotechnology, 3 (2014) 73-81.
[46] Z. Yang, J. Shen, L.A. Archer, An in situ method of creating metal oxide–carbon composites
and their application as anode materials for lithium-ion batteries, Journal of Materials
Chemistry, 21 (2011) 11092-11097.
[47] B. Wang, H. Xin, X. Li, J. Cheng, G. Yang, F. Nie, Mesoporous CNT@ TiO2-C nanocable
with extremely durable high rate capability for lithium-ion battery anodes, Scientific Reports,
4 (2014) 3729.
[48] A.C. Ferrari, J. Robertson, Interpretation of Raman spectra of disordered and amorphous
carbon, Physical Review B, 61 (2000) 14095-14107.
[49] H.C. Choi, Y.M. Jung, S.B. Kim, Size effects in the Raman spectra of TiO2 nanoparticles,
Vibrational Spectroscopy, 37 (2005) 33-38.
[50] T. Ohsaka, F. Izumi, Y. Fujiki, Raman spectrum of anatase, TiO2, Journal of Raman
Spectroscopy, 7 (1978) 321-324.
[51] T. Mazza, E. Barborini, P. Piseri, P. Milani, D. Cattaneo, A.L. Bassi, C. Bottani, C. Ducati,
Raman spectroscopy characterization of TiO2 rutile nanocrystals, Physical Review B, 75
(2007) 045416.
[52] A.L. Bassi, D. Cattaneo, V. Russo, C. Bottani, E. Barborini, T. Mazza, P. Piseri, P. Milani, F.
Ernst, K. Wegner, Raman spectroscopy characterization of titania nanoparticles produced by
flame pyrolysis: the influence of size and stoichiometry, Journal of Applied Physics, 98 (2005)
074305-9.
[53] C. Xu, P. Zhang, L. Yan, Blue shift of Raman peak from coated TiO2 nanoparticles, Journal
of Raman Spectroscopy, 32 (2001) 862-865.

14
[54] G. Rajakumar, A.A. Rahuman, B. Priyamvada, V.G. Khanna, D.K. Kumar, P. Sujin, Eclipta
prostrata leaf aqueous extract mediated synthesis of titanium dioxide nanoparticles, Materials
Letters, 68 (2012) 115-117.
[55] R. Vijayalakshmi, N. Kumar, M. Bavanilatha, S. Sunkar, Moraxella osloensis mediated
synthesis of TiO2 nanoparticles, International Journal of Pharmacy and Pharmaceutical
Sciences, 8 (2016) 397-400.
[56] H. Bar, D.K. Bhui, G.P. Sahoo, P. Sarkar, S.P. De, A. Misra, Green synthesis of silver
nanoparticles using latex of Jatropha curcas, Colloids and surfaces A: Physicochemical and
Engineering Aspects, 339 (2009) 134-139.
[57] B. Liu, K. Nakata, M. Sakai, H. Saito, T. Ochiai, T. Murakami, K. Takagi, A. Fujishima,
Hierarchical TiO2 spherical nanostructures with tunable pore size, pore volume, and specific
surface area: facile preparation and high-photocatalytic performance, Catalysis Science and
Technology, 2 (2012) 1933-1939.
[58] T. Yadav, A.A. Mungray, A.K. Mungray, A comparative analysis of a TiO2 nanoparticle
dispersion in various biological extracts, RSC Advances 5 (2015) 64421-64432.
[59] N.A. Órdenes-Aenishanslins, L.A. Saona, V.M. Durán-Toro, J.P. Monrás, D.M. Bravo, J.M.
Pérez-Donoso, Use of titanium dioxide nanoparticles biosynthesized by Bacillus mycoides in
quantum dot sensitized solar cells, Microbial Cell Factories, 13 (2014) 90-99.
[60] S.M. Roopan, A. Bharathi, A. Prabhakarn, A. Rahuman, K. Velayutham, G. Rajakumar, R.D.
Padmaja, M. Lekshmi, G. Madhumitha, Efficient phyto-synthesis and structural
characterization of rutile TiO2 nanoparticles using Annona squamosa peel extract,
Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy, 98 (2012) 86-90.
[61] R.D. Harmel, J.M. Hathaway, K.L. Wagner, J.E. Wolfe, R. Karthikeyan, W. Francesconi, D.T.
McCarthy, Uncertainty in monitoring E. coli concentrations in streams and stormwater runoff,
Journal of Hydrology, 534 (2016) 524-533.
[62] T.P. Dasari, K. Pathakoti, H.-M. Hwang, Determination of the mechanism of photoinduced
toxicity of selected metal oxide nanoparticles (ZnO, CuO, Co3O4 and TiO2) to E. coli bacteria,
Journal of Environmental Sciences, 25 (2013) 882-888.
[63] V.T. Nguyen, V.T. Vu, T.A. Nguyen, V.K. Tran, P. Nguyen-Tri, Antibacterial activity of
TiO2-and ZnO-decorated with silver nanoparticles, Journal of Composites Science, 3 (2019)
61 -75.
[64] Y. Li, W. Zhang, J. Niu, Y. Chen, Mechanism of photogenerated reactive oxygen species and
correlation with the antibacterial properties of engineered metal-oxide nanoparticles, ACS
Nano, 6 (2012) 5164-5173.
[65] T. Tong, A. Shereef, J. Wu, C.T. Binh, J.J. Kelly, J.-F. Gaillard, K.A. Gray, Effects of material
morphology on the phototoxicity of nano-TiO2 to bacteria, Environmental Science and
Technology, 47 (2013) 12486-12495.
[66] C. Jayaseelan, A.A. Rahuman, S.M. Roopan, A.V. Kirthi, J. Venkatesan, S.-K. Kim, M.
Iyappan, C. Siva, Biological approach to synthesize TiO2 nanoparticles using Aeromonas
hydrophila and its antibacterial activity, Spectrochimica Acta Part A: Molecular and
Biomolecular Spectroscopy, 107 (2013) 82-89.

15
Graphical Abstract

16

You might also like