1 s2.0 S240584402031447X Main

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Heliyon 6 (2020) e04603

Contents lists available at ScienceDirect

Heliyon
journal homepage: www.cell.com/heliyon

Research article

Green synthesis of iron oxide nanoparticle using Carica papaya leaf extract:
application for photocatalytic degradation of remazol yellow RR dye and
antibacterial activity
Md. Shakhawat Hossen Bhuiyan a, Muhammed Yusuf Miah a, Shujit Chandra Paul a, *,
Tutun Das Aka b, c, Otun Saha d, Md. Mizanur Rahaman d, Md. Jahidul Islam Sharif e,
Ommay Habiba a, Md. Ashaduzzaman e, **
a
Department of Applied Chemistry and Chemical Engineering, Noakhali Science and Technology University, Sonapur 3814, Noakhali, Bangladesh
b
Department of Pharmacy, Noakhali Science and Technology University, Sonapur 3814, Noakhali, Bangladesh
c
Department of Pharmacy, Atish Dipankar University of Science and Technology, Uttara, Dhaka 1230, Bangladesh
d
Department of Microbiology, University of Dhaka, Dhaka 1000, Bangladesh
e
Department of Applied Chemistry and Chemical Engineering, University of Dhaka, Dhaka 1000, Bangladesh

A R T I C L E I N F O A B S T R A C T

Keywords: Synthesis of iron oxide nanoparticles by the recently developed green approach is extremely promising because of
Materials science its non-toxicity and environmentally friendly behavior. In this study, nano scaled iron oxide particles (α-Fe2O3)
Materials chemistry were synthesized from hexahydrate ferric chloride (FeCl3.6H2O) with the addition of papaya (Carica papaya) leaf
Nanotechnology
extract under atmospheric conditions. The synthesis of iron oxide nanoparticles was confirmed by systematic
Iron oxide nanoparticles
Carica papaya
characterization using FTIR, XRD, FESEM, EDX and TGA studies. The removal efficiency of remazol yellow RR dye
Photocatalytic activity with the synthesized iron oxide nanoparticles as a photocatalyst was determined along with emphasizing on the
Remazol yellow RR parameters of catalyst dosage, initial dye concentration and pH. Increasing the dose of iron oxide nanoparticles
Antibacterial activity enhanced the decolorization of the dyes and a maximum 76.6% dye degradation was occurred at pH 2 after 6 h at
Cytotoxicity a catalyst dose of 0.8 g/L. Unit removal capacity of the photocatalyst was found to be 340 mg/g at dye con-
centration of 70 ppm and at a catalyst dose of 0.4 g/L. The synthesized nanoparticles showed moderate anti-
bacterial activity against Klebsiella spp., E.Coli, Pseudomonas spp., S.aureus bacterial strains. Although the cytotoxic
effect of nanoparticles against Hela, BHK-21 and Vero cell line was found to be toxic at maximum doses but it can
be considered for tumor cell damage because it showed excellent activity against the Hela and BHK-21 cell lines.

1. Introduction types of reactive azo dyes, are widely being used for coloring the textile
fibers, yarns, fabrics, etc. Various kinds of toxicities have been reported
Quality of surface water is getting deteriorated day by day due to the for such types of remazol dyes like teratogenicity in frog embryos,
release of various industrial effluents like dye in recent years. More than enzymic degradation metabolites toxicity, genotoxicity, carcinogenicity,
10,000 dyes are now being used in various industries and most of them and phytotoxicity (Birhanli and Ozmen, 2005; Silva et al., 2013; Jadhav
are used in textile dying purposes. It is estimated that about 15% of the et al., 2011).
total textile dyes are getting released into the surroundings as effluents Purification of textile waste effluents particularly dye is a crucial
and are doing harm to our environments directly or indirectly (Lachheb demand for the sake of our mankind, and environment as well. Various
et al., 2002). Most of the dyes are even entering our life cycle through kinds of methods like biological methods, filtration, adsorption, sedi-
water, and animals and causing skin irritation, respiratory diseases, and mentation, ion-exchange, UV treatment, ozonation, etc. are available for
even cancer development (Khan and Malik, 2014; Immich et al., 2009). purification textile dyes (Hassan and Carr, 2018; Bhatia et al., 2017;
Remazol yellow RR, remazol red RR, and remazol blue RR dye-different Katheresan et al., 2018; Yagub et al., 2014; Robinson et al., 2001; Lau and

* Corresponding author.
** Corresponding author.
E-mail addresses: [email protected] (S.C. Paul), [email protected] (Md. Ashaduzzaman).

https://doi.org/10.1016/j.heliyon.2020.e04603
Received 26 May 2020; Received in revised form 24 July 2020; Accepted 28 July 2020
2405-8440/© 2020 The Author(s). Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
Md.S.H. Bhuiyan et al. Heliyon 6 (2020) e04603

Ismail, 2009). But most of them are very costly, unavailable and inef- then filtered using Whatman no 42 filter paper. The filtrate was
fective in some cases. Currently, photodegradation is attracting the sci- concentrated using rotary evaporator and stored at 40 C for further use.
entist to be used for such purposes because this process possesses high
efficiency and feasibility as compared to the other traditional methods 2.3. Green synthesis of α-Fe2O3 nanoparticles
(Lopez Cisneros et al., 2002). However, these advantages cannot be
gained properly without a photocatalyst having high surface area, sta- Ferric chloride hexahydrate (FeCl3.6H2O) was used as the precursor
bility, photocatalytic activity and biocompatibility. Only nanomaterials for the synthesis of the α-Fe2O3 nanoparticles. 50 mL of the papaya leaves
based photocatalyst found to be possessed such properties and hence are extract was added dropwise with 50 mL of 0.1M FeCl3.6H2O solution in
now considering for degradation of textile dyes proficiently. Various 1:1 ratio at room temperature. Following this, 1 M NaOH was added till
types of metal (Ag, Au) and metal oxide (ZnO, TiO2) based nanoparticles the pH became 11. The resultant mixture was stirred using a magnetic
(NPs) have been studied for utilizing as a photocatalyst for the degra- stirrer for 30 min and the formation of intense black colored solution
dation of textile dyes (Hamalo glu et al., 2017; Ahammed et al., 2020). confirmed the synthesis of iron oxide nanoparticles (Bibi et al., 2019).
Among various types nanomaterials, iron oxide nanoparticles (FO NPs) The nanoparticles were separated by centrifugation at 8000 rpm for 20
have excellent catalytic and reductive properties to be used for waste- min and cleansed by subsequent washing with ethanol and water for 2–3
water treatment and it has the advantage of the ease of separation as times. The NPs were finally dried in a hot air oven at 80  C for 3 hr and
compared to the other nanomaterials requiring highly expensive centri- stored in a seal tight container for further use.
fugation for separation (Oliveira et al., 2002). FO NPs are usually used for
wide range applications from removal heavy metals, dyes, antibiotics 2.4. Characterization
from water sources to the biomedical field like site-specific drug delivery
and damaging tumor cell (Vasantharaj et al., 2019; Devi et al., 2019). FT-IR spectra of sample was recorded on a FT-IR 8400S spectropho-
Again, iron-based nanoparticles found to be effective against various tometer (Shimadzu corporation, Japan) in the wavenumber range of
pathogenic bacterial strains and fungi effectively as they can produce 4000–400 cm1. XRD patterns were recorded by an x-ray diffractometer
highly reactive oxygen species (ROS) (Muthukumar et al., 2019). (U1tima IV, Rigaku Corporation, Japan) by using Cu kα radiation (λ ¼
There are various available methods for synthesis of nanomaterials 0.154) from a broad focus Cu tube operated at 40 KV and 40 MA. The
like sol-gel method, chemical reduction method, co-precipitation, hy- morphology of nanoparticles was analyzed by means of field emission
drothermal synthesis etc (Ahmed, 2015). The chemicals used in these scanning electron microscope (JEOL JSM-7600F, Japan) run at a voltage
methods are considered as harmful for the environment. Therefore, in the of 5.0 KV. The UV-VIS spectra of the sample to measure the absorbance of
recent times green synthesis of nanomaterials has gained importance due the dye which were done by using the double beam UV-1700 Series
to its low cost, simplicity and environmentally friendly nature (Kumar Spectrophotometer (Shimadzu corporation, Japan).
et al., 2014). Green synthesis of nanoparticles utilizes plant extract as
both reducing and capping agent eliminating the necessity of harmful 2.5. Photocatalytic activity
reducing agents (Peralta-Videa et al., 2016). Plant extract contains
various phytochemicals such as polyphenols, flavonoids, terpenoids, To study the photocatalytic activity of nanoparticles, a series of
phenolic acids, which are responsible for the reduction and formation of remazol yellow RR dye solution having different concentrations (10 ppm,
stabilized nanoparticles (Izadiyan et al., 2020). Magnetic nanoparticle 30ppm, 50 ppm and 70ppm) were prepared. 100ml of the dye solution
synthesized by green methods are non-toxic as compared to nano- was taken in a beaker along with different quantity of nanoparticles (0.2
particles synthesized using sodium borohydride (Mahdavi et al., 2013). g/L, 0.4/g/L, 0.6 g/L and 0.8 g/L) and kept under sunlight along with
Recently, several studies were carried out for green synthesis of continuous stirring. The absorbance was then taken after definite time
iron-based nanoparticles from various plants parts like fruit extract of intervals and the degradation efficiency was calculated by Eq. (1).
Cynometra ramiflora, rind of Persea americana, seeds extract of Punica
granatum, flower extract of Avicennia marina, etc. for the degradation of C0  Ct
Degradation efficicency ð%Þ ¼ X 100 (1)
various textile dyes (Bishnoi et al., 2018; Kamaraj et al., 2019; Bibi et al., C0
2019; Karpagavinayagam and Vedhi, 2019).
The goal of this study was to synthesize the iron oxide (α-Fe2O3) NPs C0 is the initial concentration and Ct is the concentration at time t of
by using papaya plant leaf extract (Carica papaya) as reducing/stabilizing remazol yellow RR dye.
agent and studying its photocatalytic efficiency for the degradation of The effect of pH on dye degradation was determined by maintaining a
reactive azo dye (remazol yellow RR), antibacterial activity against different pH environment (pH 2, 4, 6 and 8) using buffer solutions at
bacterial strains and its probable in-vitro cytotoxicity. optimum dye concertation and nanoparticles doses.
The time required for 25%, 50% and 75% degradation of dyes (T25,
2. Materials and methods T50 and T75) were calculated from Eqs. (2), (3), and (4).

0:288
2.1. Chemicals T25 ¼ (2)
k

Analytical grade ferric chloride hexahydrate (FeCl3.6H2O), sodium 0:693


hydroxide pellets (NaOH) were purchased from Merck, India. All T50 ¼ (3)
k
chemicals are used without further purification. Remazol yellow RR dye
was collected from local textile of Bangladesh. 1:386
T75 ¼ (4)
k
2.2. Collection and preparation plant extract
Where, and k is the rate constant of the photocatalytic dye degradation
The papaya plant (Carica papaya) leaves were collected from the reaction.
premises of Noakhali Science and Technology University, Bangladesh.
The fresh leaves were then washed multiple times with tap water fol- 2.6. Antibacterial activity
lowed by deionized water. The leaves were then dried in oven for an hour
and then grinded to from fine powder. 20 grams of fine powders are Well diffusion method by agar plates was used for calculating the
boiled with 1L of deionized water at 80  C for 30 min and the extract is zone of inhibition (Shamaila et al., 2016). Clinical pathogenic bacteria

2
Md.S.H. Bhuiyan et al. Heliyon 6 (2020) e04603

Figure 1. (a) Synthesis route of α-Fe2O3 nanoparticles and (b) possible reaction mechanism for synthesis of α-Fe2O3 nanoparticles.

Klebsiella spp.strain KH15, E.Coli strain EH9, Pseudomonas spp. strain 2.7. Cytotoxicity study
PsI1strain PsI1, S.aureus strain 6s were developed on nutrient agar
plate and maintained at 37  C for whole night. The overnight culture Cytotoxic effect of α-Fe2O3 nanoparticle was observed against Hela,
of bacteria in nutrient broth (Oxoid Limited, United Kingdom) was BHK-21 and Vero cell line that were collected from Centre for Advanced
used for the experiment. In this method, sterilized nutrient agar plate Research in Sciences (CARS) of University of Dhaka. Hela, a human
was equipped for each bacterium. All bacterial culture was adjusted in cervical carcinoma cell line, BHK-21, a baby hamster kidney fibroblast
optical density (OD) 0.1 using UV–Visible Spectrophotometer (Spec- cell line and Vero cell line, a kidney epithelial cells of African green
trumlab 1200RS, Japan). The spectrophotometer was first made auto monkey, were maintained in DMEM (Dulbecco's Modified Eagles' me-
zero using blank to eliminate the effect of assay reagents. These three dium) containing 1% penicillin-streptomycin (1:1) and 0.2% gentamycin
bacterial pathogens were then coated over the agar plate with the help and 10% fetal bovine Serum (FBS). Hela cells (2104/100 μl) and BHK-
of sterile swab of cotton. Then these plates were permitted to dry. 21 cells (1.5104/100 μl) were seeded onto 96-well plate and incubated
After that, one wells were bored by a sterile well cutter (7.0 mm at 37  C in CO2 incubator (Nuaire, USA). 25 μl (30 mg/mL) sample of
diameter) in each agar plate. Subsequently, the suspension of NPs (5 nanoparticles (autoclaved) was added each well. After 48 h of incuba-
mg/ml, 20 mg/ml and 30 mg/ml) was poured into individual wells for tion, insoluble samples were washed out with fresh media and cytotox-
each strain. The plates were permitted to put for 1 h for complete icity was examined under a Trinocular microscope with camera (Optika,
diffusion followed by incubation at 37  C for 24 h (hr) and measured Italy). Duplicate wells were used for each sample.
the diameter of inhibitory zones in mm.

Figure 2. Vibrational properties of plant extract mediated synthesized iron oxide nanoparticles.

3
Md.S.H. Bhuiyan et al. Heliyon 6 (2020) e04603

Figure 3. XRD data of α-Fe2O3 nanoparticles.

Figure 4. (a–b) FESEM image showing morphology; (c) Size distribution histogram for α-Fe2O3 nanoparticle (d) EDX data of green synthesized α-Fe2O3 nanoparticles.

4
Md.S.H. Bhuiyan et al. Heliyon 6 (2020) e04603

et al., 2020; Liu et al., 2009). The peaks at position of 3357.57 cm1
Table 1. EDX and zeta potential data of α-Fe2O3 nanoparticles. represent the –OH bond stretching from aqueous phase. Again, the peaks
EDX data Zeta potential (ζ) at 1433.11 cm1, and 3691.57 cm1 also represents the –OH bond
stretching and bending from various phenolic and carboxylic group
Element Mass% Day mV
respectively present the plant extract (Qasim et al., 2020). Moreover, the
C 25.05 0 -15.1
peaks at 2927.94 cm1, 1625.99 cm1 and 1101.35 cm1 denotes C–H
O 31.03 15 -14.8
stretching, C¼C stretching and C–O stretching ensuring the presence of
Fe 32.70 30 -14.4
alkane, conjugated alkene and secondary alcohol in the plant extract
correspondingly as observed by other study (Aisida et al., 2020). The
2.8. Ethical statement shift in peak position in the range of 400–4000 cm1 ensure that these
functional groups containing compounds bound to the iron oxide surface.
This research work did not include any animal. We carried out anti- The XRD data as shown in Figure 3 indicates that the crystal planes of
bacterial activity on different bacterial reference strains that we (012), (104), (110), (113), (024), (116) and (018) corresponding to the
mentioned. We also carried out cytotoxic activity using 3 different kind of 25.16 , 35.12 , 36.63 , 40.64 , 49.97 , 57.08 , and 59.42 indicates
of cell lines that are commercially available. There was no necessity for the formation of α-Fe2O3 nanoparticles. The intense and sharp peaks
ethical approval consideration in this study. undoubtedly revealed that Fe2O3 nanoparticles formed by the reduction
method using Carica papaya leaf extract were crystalline in nature. The
3. Results and discussions results are almost similar to the results obtained for iron oxide nano-
particles by other researchers (Ahmmad et al., 2013; Suresh et al., 2016;
3.1. Preparation of α-Fe2O3 nanoparticles Lassoued et al., 2018). The average crystallite size as determined using
the Debye–Scherrer equation was found to be 4.58 nm.
The synthesis scheme of iron oxide NPs is shown in Figure 1a. The The morphology of the synthesized nanoparticle is shown in Figure 4
various phytochemicals like polyphenols, flavonoids, glycosides, and (a-b). The figure indicates that the synthesized nanoparticles are not
tannins present in the leaf extract equally act as reducing and stabilizing uniform in nature and get agglomerated in some cases. The size of NPs
agents for the synthesis of NPs (Juarez-Rojop et al., 2014). The formation was determined by selecting 100 particles and their average diameter
of black color precipitates occurred due to the interaction between these was found to be 21.59 nm (Figure 4c). The big agglomerated clusters
phytochemicals and metal ions ensuring the formation of α-Fe2O3 were formed due to accumulation of tiny building blocks of various
nanoparticles (Anchan et al., 2019). After mixing of iron salt with leaf bioactive reducing agents of plant extract or this might be due to the
extract at definite reaction condition, it is not be able to reduce Fe3þ to lower capping ability of the plant extract and agglomeration tendency of
Fe0; rather, the phytochemicals react with the iron ions to give iron oxide the iron-based nanoparticles due to magnetic interactions.
NPs, as it is prone to oxidation (Figure 1b) (Devi et al., 2019). Furthermore, the elemental composition of the sample was analyzed
by EDX analysis. The EDX analysis reported in Figure 4c, clearly shows
3.2. Characterization the presence of the K-α at 6.4 keV due to Fe atoms present in the nano-
particle and two K-α lines at 0.28 keV and 0.6 keV coming from the C and
The FTIR analysis (400-4000 cm1) of the synthesized sample O atoms respectively. Similar results were also obtained by other study
ensured the synthesis of α-Fe2O3 nanoparticles as well the existence of (Vasantharaj et al., 2019). The percentage of mass present under the
various reducing agents functional groups presents in the papaya plant irradiated area is 25.05%, 31.03%, 32.70% for Carbon, Oxygen, and iron
leaf extract (Figure 2). The peaks at 474.49 cm1, 621.08 cm1 and respectively (Table 1). The presence of carbon is from the plant extract
678.94 cm1 ensure the presence of Fe–O bond in the sample (Aisida components existing in the NPs surface (also shown in FTIR data) and the

Figure 5. TGA data of green synthesized α-Fe2O3 nanoparticles.

5
Md.S.H. Bhuiyan et al. Heliyon 6 (2020) e04603

Figure 6. Degradation of remazol yellow RR dye (a) in absence of photocatalyst; in presence of (b) 0.2 g/L, (c) 0.4 g/L, (d) 0.6 g/L, (e) 0.8 g/L; (f) extent of
degradation with time at different catalyst doses with time; (g) degradation efficiency at different time intervals (h) influence of catalyst loading on degradation of dye.

6
Md.S.H. Bhuiyan et al. Heliyon 6 (2020) e04603

Figure 7. The (a)first-order and (b) second-order linear plot of ln (C0/Ct) vs. time for remazol yellow RR dye degradation in the presence of nanoparticles as catalyst.

presence of Cl as impurities is usually observed during iron oxide NPs degradation reached to about 2, 4, 14, 25, 40, 57% after1 2, 3, 4, 5, and 6
synthesis from ferric chloride as observed by others (Qasim et al., 2020). h respectively (Figure 6g). Again, the results in Figure 6h, demonstrate
Furthermore, the negative zeta potential data (-15.1 mV) revealed that the increase in catalyst dose is directly proportional to the photocatalytic
the synthesized NPs is stable and its stability remains almost constant dye degradation percentage. More specifically, without photocatalyst the
over thirty days after its synthesis (Table 1). degradation was only 6.38%, for 0.6 g/L doses the degradation was 74%
TGA analysis was done to know the thermal stability, decomposition and for maximum catalyst dose (0.8 g/L) the degradation almost reached
temperature, and also decomposition rate of the nanoparticles. TGA to 77% after 6 hr as shown in Figure 6h. The lower degradation rate at
measurements were performed with a heating rate of 10  C/min in the lesser catalyst dose was due to the less amount of photon absorption by
temperature range of 24.38  C to 800  C. The TGA curve of the iron oxide the catalyst and the subsequent decrease in ROS production. While with a
nanoparticle is shown in Figure 5. The TGA plot depicted three weight continuous increase in catalyst (for 0.6 and 0.8 g/L), the degradation
loss steps in the tested temperature range of 24.38  C to 798.70  C. The efficiency became almost constant (Figure 6g-h) which might be due to
first step 25.41% weight loss occurred in the temperature range of 24.38 the opacity of solution, agglomeration of nanoparticles and light scat-

C to 301.70  C indicating the removal of water, organic solvent or re- tering ability of nanoparticles at higher catalyst dose (Reza et al., 2017).
sidual solvent, physiosorbed and chemisorbed H2O molecules in the Therefore, it can be concluded that neither a higher nor a lower catalyst
sample. In second and third steps of TGA curve, total of 34.43% weight dose is suitable for the degradation of reactive dyes, therefore we
loss occurred at 301.70  C to 798.70  C suggesting the elimination or selected a dose of 0.4 g/L for further studies.
decomposition of the capping biomolecules. Further, there was no weight
loss observed above 798.70  C and 40.16% weight residue of the iron 3.3.2. Degradation kinetics
oxide nanoparticles remained at 798.70  C. The kinetic constants of dye photodegradation are usually estimated
by applying a pseudo-first-order and second-order reaction rate Eqs. (5)
3.3. Degradation study of remazol yellow RR dye and (6).
 
Co
To study the efficiency of synthesized nanoparticles as a photo- ln ¼ K1 t (5)
Ct
catalyst, we used remazol yellow RR as a model dye. The experiment was
conducted under sunlight irradiation during 8 am to 4 pm every day. To
1 1
complete the degradation process time, we kept the sample of current  ¼ K2 t (6)
Ct C0
day in dark place to keep it again under sunlight in the next day if
necessary. For that reason, we also experimented the degradation under Where t (hr) is the reaction time, C0 is the initial concentration, Ct is the
dark place and we did not observe any degradation. However, to opti- concentration at time t, k1 (hr1) and k2 (Lmg1hr1) represents the first-
mize the dye degradation we maintained the degradation process with order and second-order reaction rate constant.
different doses of photocatalyst, time, initial dye concentration and pH The rate constant (k) of first-order reaction was obtained by plotting
value. ln (C0/Ct) against the reaction time t and 1/Ct- 1/Co against time (t) for
second-order reaction as shown in Figure 7(a-b). Figure showed that the
3.3.1. Influence of contact time and catalyst loading degradation of dye is completely dose dependent and fits perfectly with
To study the effect of contact time and catalyst loading, a series of first-order kinetics (R2 value of first-order reaction is higher than second-
experiments were carried out for 6 hr by varying the nanoparticles doses order reaction for all doses). The lowest degradation rate of 0.008 
from 0.2 to 0.8 g/L keeping the dye concentration at 50 ppm and pH of 2 0.0009 hr-1 (R2 ¼ 0.99) was observed for blank sample (without NPs)
as shown in Figure 6(b-e). To evaluate the nanoparticle activity a blank while the rate constant increased from 0.06 to 0.33 hr 1 for 0.2–0.8 g/L
experiment was also done represented in Figure 6a. From Figure 6(a-f), it of catalyst doses which indicates that the reaction rate is almost forty
is very clear that without using nanoparticles there was no significant times higher in case of 0.8 g/L of catalyst as compared to the absence of
degradation over a period of 6 hr, whereas with the increasing amount of NPs (Table 2). However, the dye degradation reached its half value by
catalyst loading the there was a sharp increase in degradation rate with 3.47 h for 0.4 g/L doses while it took only 3.47 hr and 2.57 hr for 0.6 g/L
reaction time. As shown in the figure, for all catalyst doses the degra- and 0.8 g/L of catalyst doses correspondingly. Furthermore, although 0.2
dation increased with time and for 0.4 g/L of catalyst dose, the

7
Md.S.H. Bhuiyan et al. Heliyon 6 (2020) e04603

further get united with water to give oxygen reactive species and OH‫־‬
Table 2. Comparison of degradation kinetics of dye solution by nanoparticles for that help in degradation of the dye (Kansal et al., 2009). Furthermore, the
6 hr.
various biomolecules present in plant extract and even in the NPs surface
Nanoparticle Catalyst Parameters Remazol yellow RR dye acts like catalysts to boost the photocatalytic activity and the subsequent
Zero loading First-order rate constant (k) (hr1) 0.008  0.0009 enriched degradation of dye molecules (Haritha et al., 2016).
T25 (hr) -
T50 (hr) - 3.3.3. pH effect
T75 (hr) - To study the effect of pH effect over dye degradation a number of
0.2 g/L NPs First-order rate constant (k) (hr1) 0.06  0.004 experiments were done within a pH range 2–8 maintaining 0.4 g/L of
T25 (hr) 4.79 catalyst doses and 50 ppm of dye concentration (Figure 9). The pH of the
T50 (hr) - solution was adjusted before sunlight irradiation and is not controlled
T75 (hr) - during the degradation reaction. From Figure 9 it is clear that the
0.4 g/L NPs First-order rate constant (k) (hr1) 0.20  0.03
degradation was favored at lower pH while it was found to be decreased
T25 (hr) 1.44
with the increase in pH value. More specifically, the highest degradation
of almost 99% was observed at pH 2, followed by 68% degradation at pH
T50 (hr) 3.47
4 while for the rest of the pH region the degradation almost remained
T75 (hr) -
same with the highest value of about 15% within 10 hr. The degradation
0.6 g/L NPs First-order rate constant (k) (hr1) 0.27  0.03
of dye usually depends on the adsorption of dye molecules on the catalyst
T25 (hr) 1.07
surface and formation of hydroxyl free-radical which in turn depends on
T50 (hr) 2.57
the availability of hydroxyl ion in the reaction medium. The zero-point
T75 (hr) 5.13
charge (pHzpc) of iron oxide is between 6-7 which means that below
0.8 g/L NPs First-order rate constant (k) (hr1) 0.33  0.02
this range it becomes positively charged and above this range it becomes
T25 (hr) 0.87 negatively charged (Meng et al., 2016). So, at lower pH value (particu-
T50 (hr) 2.10 larly at 2) attractive forces between the NPs surface (positively charged)
T75 (hr) 4.20 and the anionic remazol yellow RR dye favors adsorption resulting in the
enhanced degradation. The lower degradation rate at neutral pH was
observed because, close to zero-point charge (pHzpc) region the iron
g/L of catalyst could not cause 75% of dye degradation within 6 h period
oxide nanoparticles get agglomerated and resulting in the decreased
but it was completed only in 5.13 and 4.27 hr by 0.6 g/L and 0.8 g/L of
amount hydroxyl radical formation (OH⋅) and a subsequent decrease in
catalyst doses respectively.
degradation rate. Furthermore, although higher pH value could provide a
Iron oxide NPs produce hydroxyl radical (OH) under sunlight radi-
higher concentration of hydroxyl ions to react with valence band holes
ation which are reposnsible for the degradation of remazol yellow RR dye
(hþ) to form more OH⋅, but the negative surface charge of NPs resulting
and the probable degradation mechanism is shown in Figure 8. Gener-
in electrostatic repulsion, which keep away the hydroxyl ion and oxygen
ally, when NPs were irradiated under sulight, an electron (e) and hole
molecule from adsorbing on its surface and thus decreases the avail-
(hþ) pair is produced (Varadavenkatesan et al., 2019). The produced
ability of hydroxyl and superoxide radical for dye degradation (Choud-
electron is excited from the valence band to the conduction band, leaving
hury et al., 2012). Again, at higher pH deterioration of H2O2 into oxygen
the hþ in the valence band. This hole (hþ) is responsible for the con-
and water rather than hydroxyl radical (OH⋅) become favorable resulting
version of water into hydroxyl radical, which is responsible for oxidative
in a decrease in dye degradation (Shu et al., 2004). However, it is shown
degradation of dye. On the other hand, electron combines with molecular
form the figure that, although more than 75% of the dye (at pH 2) was
oxygen and converted into superoxide radical. The superoxide radical is
degraded within 6 hr but it took four more hours for its complete
further converted in to hydroxyl radical which is a strong oxidizing agent
degradation, which means that the degradation become slow after 6 hr.
and degrades the dye to harmless end products (Kamaraj et al., 2019).
This has happened because the surface of NPs might get occupied with
Again, the highly oxidizing hole generated by NPs after absorbing the
the dye molecules which restrains the NPs from further degradation of
sunlight causes the direct oxidation of dyes and release Hþ ions which
dye molecules.

Figure 8. Probable degradation mechanism of remazol yellow RR dye by iron oxide NPs under sunlight irradiation.

8
Md.S.H. Bhuiyan et al. Heliyon 6 (2020) e04603

Figure 9. Effect of pH on photodegradation of remazol yellow RR dye.

Figure 10. Influence of initial dye concentration on the degradation efficiency


of dye by NPs.
3.3.4. Initial dye dose variation
The effect of initial dye dose variation on the photocatalytic degra- method showed activity against safranin dye with degradation (68.8%)
dation efficiency of dye by iron oxide NPs was examined by changing the within 180 min (Qasim et al., 2020). In case of degradation of remazol
dye concentration from 10 to 70 mg/L at constant catalyst loading of 0.4 yellow RR dye by TiO2 NPs in presence of H2O2 it took about 15 min for
g/L and pH 2. As shown in Figure 10, it is very clear that the degradation 98% degradation with a catalyst doses of 1 g/L (Bibi et al., 2017). The
efficiency decreases with the increase in initial due concentration. As NPs synthesis in this study showed excellent activity without any addi-
indicated in figure, the 10ppm solution reached its maximum degrada- tives and with a lower catalyst dose as well as compared to other studies
tion within 4 h, 30 ppm dye reached its maximum degradation within 6 (Kumar et al., 2018).
hr and it took about 14 hr for almost complete degradation of 70 ppm dye
solution for the same amount of catalyst doses. This is because, at higher 3.4. Antimicrobial activity of α-Fe2O3 NPs
dye concentration, dye molecules might adsorb on the catalyst surface,
hence a significant amount of UV light is absorbed by the dye molecules The antibacterial activity of green synthesized nanoparticles suspen-
themselves rather than by the nanoparticles which in turns results in the sions of different concentrations was done against three gram negative
decrease in the generation of hydroxyl radicals and dye degradation bacteria such as Klebsiella spp., E.Coli, Pseudomonas spp., and one gram
(Daneshvar et al., 2003). Also, the intermediates formed during photo- positive bacteria which is S.aureus in aqueous lysogeny broth (LB). The
catalytic degradation process might compete with the intact dye mole- well diffusion method was used to test the ability of the antibacterial
cules for the available active sites on the catalyst surface (So et al., 2002). agent (NPs) to rupture the bacterial cells. The antibacterial activity
Actually, with the increase in initial dye concentration the necessity of studied against gram-positive and gram-negative bacteria at different
catalyst surface for degradation also increases. Since both the sunlight concentrations of samples are shown in Figure 12.
irradiation time and catalyst dose are constant, the hydroxyl radical From Figure 12 it is very clear that the synthesis nanoparticles
(OH) formed on the surface of the catalyst is also constant and insuffi- showed antibacterial properties and the highest effect was observed for
cient for the degradation of excess dye. S. aureus while the lowest effect was for the Klebsiella spp. whereas the
activities are completely dose dependent. For S. aureus the zone of in-
3.3.5. Unit removal capacity (URC) hibition was 7  1 mm for 5 mg/ml doses while it gets almost doubled to
The unit removal capacity of photocatalyst was calculated to deter- an inhibition zone of 12.5  0.5 mm for 30 mg/ml. On the other hand,
mine the removal capacity of catalyst per unit of weight. From Figure 11 Klebsiella spp. showed resistance to 5 mg/ml and 20 mg/ml doses but at
(a-b) it is shown that the URC value was increased sharply with the in- 30 mg/ml doses it grown an inhibition zone of about 9  1 mm. More-
crease in initial dye concentration and it showed an inverse trend in case over, the NPs exhibited moderate effect on both E.Coli and Pseudomonas
of catalyst dose variation for the same initial dye concentration. Each spp. with an inhibition zone of about 9  1 mm and 10  0.5 mm
gram of nanoparticles removed 50, 140, 245 and 340 mg of dye in case of respectively at 30 mg/ml dose. The as possessed antibacterial properties
10, 30, 50 and 70 ppm of initial dye concentration respectively. On the of nanoparticles is due to its nanoscale size allowing to accumulate or
other hand, for the same initial dye concentration (50 ppm) each gram of deposit on the surface of studied bacterial strains which is reported by
photocatalyst remove 240, 120, 80 and 60 mg for 0.2, 0.4, 0.6 and 0.8 g/ other researchers (Varadavenkatesan et al., 2019; Jagathesan and Rajiv,
L of catalyst doses correspondingly. The increase in URC with the in- 2018; Groiss et al., 2017).
crease in initial dye concentration was due to the availability of dye for Apart from the NPs, the plant extracts might also possess antibacterial
degradation while the decrease in URC with increasing catalyst dose activity due to the presence of phytochemical components (Vasantharaj
might be due to the unsaturation of active sites of catalyst. et al., 2019). However, there are a number of hypotheses involved to
explain the exact mechanism of NPs against bacterial strains. Generally,
3.3.6. Comparison with other works the iron oxide nanoparticles showed its antibacterial properties due to
A comparison on degradation of various reactive dyes by several NPs the production of reactive oxygen species (ROS), oxidative stress caused
is summarized in Table 3. Iron oxide NPs synthesized by green reduction by ROS, the reaction of ions released by nanoparticles with thiol groups

9
Md.S.H. Bhuiyan et al. Heliyon 6 (2020) e04603

Figure 11. Unit removal capacity of photocatalyst with respect to (a) initial dye concentration and (b) catalyst doses. Date are presented as mean  SD.

Table 3. Comparison of reactive dyes degradation efficiency by various NPs.

Nanoparticles Synthesis route Catalyst Dose (g/L) Dye Dye dose (ppm) Degradation (%) Time (min) Reference
FeO nanorods Green reduction 0.005 Safranin 10 68.8 180 (Qasim et al., 2020)
FeO nanorods Chemical Reduction 0.005 Safranin 10 24.82 180 (Qasim et al., 2020)
FeO NPs Green reduction - Crystal violet 10 78.78 150 (Vasantharaj et al., 2019)
α-Fe2O3 NPs Green reduction 0.8 Remazol yellow RR 50 75 250 This study
TiO2 NPs with H2O2 Chemical method 1 Remazol yellow RR 75 98 15 (Soutsas et al., 2010)
Cobalt Oxide NPs Green reduction 0.005 Remazol brilliant orange 3R 150 78.45 50 (Bibi et al., 2017)
Tin Oxide NPs Green reduction 1 Reactive yellow 186 40 90 180 (Kumar et al., 2018)
Silver NPs Green reduction 2.5 Methylene blue in presence of NaBH4 40 80 20 (Paul et al., 2020)

(–SH) of the bacterial cell, therefore, interrupt the DNA replication and increasing concentration of nanoparticles might cause excessive
protein synthesis process of microorganism by altering their structure ROS-mediated oxidative stress to the cell leading to the DNA damage.
(Arakha et al., 2015; Luo et al., 2015). Again, the positively charged This means that cancer cells may be more sensitive to ROS challenge
metal ions released from NPs might interact with negatively charged than normal cells, thus it may be possible to target cancer cells by ROS-
bacterial strains surface with a result in the disruption and destabilization mediated mechanisms (Mitra et al., 2019). As iron-based nanoparticles
of microorganism surface protein and subsequent cell death (Cardillo
et al., 2016). However, this study also suggested that the nanoparticles
showed better activity against gram positive bacteria as compared to the
gram-negative strains. This has happened because the gram-negative
bacteria usually own an extra outer layer of lipopolysaccharide and
peptidoglycan which helps the gram-negative bacteria to reduce the
damage that might cause by nanoparticles.

3.5. Cytotoxicity study

Cytotoxic effect was examined against Hela, a human cervical carci-


noma cell line, BHK-21, a baby hamster kidney fibroblast cell line and
Vero-isolated from kidney epithelial cells extracted from African green
monkey. The cytotoxicity effect of green synthesized α-Fe2O3 nano-
particle are shown in Table 4 and Figure 13.
From the Table and Figure, it is elucidated that without solvent (deion-
ized water) 100% of the cell lines can survive without any damage. But in
presence of DI water about 5% of the cell was damaged after 48 h of intu-
bation. However, the green synthesized α-Fe2O3 nanoparticles showed a
high toxicity towards the Hela and BHK-21 cell lines and it is reported that
only 5% of the cell lines were survived after 48 h. Moreover, in the case of
Vero cell, about 10–20% of the cell were survived for the same doses of
nanoparticle. In this study, higher toxicity was observed because of the Figure 12. Antibacterial inhibition zone of α-Fe2O3 nanoparticles at different
utilization of higher doses of nanoparticles as observed by other study also concentration. Date are presented as mean  SD and analyzed with one-way
analysis of variance. Different letters represent statistical significance among
(Kamaraj et al., 2019). These types of results are observed because the
different treatments (p < 0.05). The symbol “R” indicates that those are resistant
to studied bacterial strains.

10
Md.S.H. Bhuiyan et al. Heliyon 6 (2020) e04603

Table 4. Cytotoxicity effect of green synthesized α-Fe2O3 nanoparticle on various cell lines.

Sample Dose Survival of cells

Hela BHK-21 Vero (Non-tumoral)


In absence of solvent 25 μL 100% 100% 100%
Presence of solvent 25 μL >95% >95% >95%
α-Fe2O3 nanoparticles 25 μL (30 mg/mL) <5% <5% 10–20%

Figure 13. (a) and (b) represents the cell line effect in absence and presence of solvent in the medium respectively; (c) represents the cell line before washing with
fresh media. In vitro cytotoxicity of NPs on (d) Hela cell line; (e) BHK-21 cell line; (f) Vero cell line.

can serve as a strong reactive oxygen species (ROS) inducer therefore an 4. Conclusions
effective amount can selectively kill tumor cells and inhibit the growth of
the tumor cells. However, Hela cells are derived from cervical cancer cells The iron oxide nanoparticle was synthesized effectively through a
therefore such types of nanoparticles could be used for cancer treatment as green synthesis route by using the leaf extract of Carica papaya plant. The
it has been reported that cancer cells have higher ROS levels as wells as as synthesized α-Fe2O3 nanoparticles showed efficient degradation abil-
more oxidative DNA damage than normal cells in the same tissues. ity against remazol yellow RR dye in the presence of sunlight and about

11
Md.S.H. Bhuiyan et al. Heliyon 6 (2020) e04603

77% of the dye was degraded within 6 h for 0.8 mg/L of dosage. Syn- Arakha, M., Pal, S., Samantarrai, D., Panigrahi, T.K., Mallick, B.C., Pramanik, K.,
Mallick, B., Jha, S., 2015. Antimicrobial activity of iron oxide nanoparticle upon
thesized NPs exhibited moderate antibacterial efficiency on specific
modulation of nanoparticle-bacteria interface. Sci. Rep.
gram-negative and gram-positive bacterial strains. Although nano- Bhatia, D., Sharma, N.R., Singh, J., Kanwar, R.S., 2017. Biological methods for
particles showed toxicity at high doses but it showed excellent activity textile dye removal from wastewater: a review. Crit. Rev. Environ. Sci. Technol.
(almost 95% of cells destroyed) against BHK-21 and HELA cell line 1836–1876.
Bibi, I., Nazar, N., Iqbal, M., Kamal, S., Nawaz, H., Nouren, S., Safa, Y., Jilani, K.,
indicating that they could be a promising alternative for damaging tumor Sultan, M., Ata, S., Rehman, F., Abbas, M., 2017. Green and eco-friendly synthesis of
cells at optimum doses. However, a detailed study is necessary to find out cobalt-oxide nanoparticle: characterization and photo-catalytic activity. Adv. Powder
the exact doses and reaction conditions for utilizing the nanoparticles for Technol. 29 (9), 2035–2043.
Bibi, I., Nazar, N., Ata, S., Sultan, M., Ali, A., Abbas, A., Jilani, K., Kamal, S., Sarim, F.M.,
these purposes. Khan, M.I., Jalal, F., Iqbal, M., 2019. Green synthesis of iron oxide nanoparticles
using pomegranate seeds extract and photocatalytic activity evaluation for the
Declarations degradation of textile dye. J Mater Res Technol 8 (6), 6115–6124.
Birhanli, A., Ozmen, M., 2005. Evaluation of the toxicity and teratogenity of six
commercial textile dyes using the frog embryo teratogenesis assay-xenopus. Drug
Author contribution statement Chem. Toxicol. 28 (1), 51–65.
Bishnoi, S., Kumar, A., Selvaraj, R., 2018. Facile synthesis of magnetic iron oxide
nanoparticles using inedible Cynometra ramiflora fruit extract waste and their
Shakhawat H. Bhuiyan: Performed the experiments; Wrote the paper. photocatalytic degradation of methylene blue dye. Mater. Res. Bull. 97, 121–127.
Muhammed Y. Miah, Otun Saha, Mizanur Rahaman: Analyzed and Cardillo, D., Weiss, M., Tehei, M., Devers, T., Rosenfeld, A., Konstantinov, K., 2016.
interpreted the data; Contributed reagents, materials, analysis tools or Multifunctional Fe2O3/CeO2 nanocomposites for free radical scavenging ultraviolet
protection. RSC Adv.
data; Wrote the paper. Choudhury, B., Borah, B., Choudhury, A., 2012. Extending photocatalytic activity of TiO 2
Shujit C. Paul: Conceived and designed the experiments; Performed nanoparticles to visible region of illumination by doping of cerium. Photochem.
the experiments; Analyzed and interpreted the data; Wrote the paper. Photobiol. 88 (2), 257–264.
Daneshvar, N., Salari, D., Khataee, A.R., 2003. Photocatalytic degradation of azo dye acid
Tutun D. Aka: Conceived and designed the experiments; Analyzed red 14 in water: investigation of the effect of operational parameters. J. Photochem.
and interpreted the data; Wrote the paper. Photobiol., A 157, 111–116.
Jahidul I. Sharif, Ommay Habiba: Analyzed and interpreted the data; Devi, H.S., Boda, M.A., Shah, M.A., Parveen, S., Wani, A.H., 2019. Green synthesis of iron
oxide nanoparticles using Platanus orientalis leaf extract for antifungal activity. Green
Wrote the paper.
Process. Synth. 8 (1).
Ashaduzzaman: Conceived and designed the experiments; Analyzed Groiss, S., Selvaraj, R., Varadavenkatesan, T., Vinayagam, R., 2017. Structural
and interpreted the data; Contributed reagents, materials, analysis tools characterization, antibacterial and catalytic effect of iron oxide nanoparticles
or data; Wrote the paper. synthesized using the leaf extract of Cynometra ramiflora. J. Mol. Struct. 1128,
572–578.
Hamalo € Sa
glu, K.O., g, E., Tuncel, A., 2017. Bare, gold and silver nanoparticle decorated,
monodisperse-porous titania microbeads for photocatalytic dye degradation in a
Funding statement newly constructed microfluidic, photocatalytic packed-bed reactor. J. Photochem.
Photobiol. Chem. 332, 60–65.
This research did not receive any specific grant from funding agencies Haritha, E., Roopan, S.M., Madhavi, G., Elango, G., Al-Dhabi, N.A., Arasu, M.V., 2016.
Green chemical approach towards the synthesis of SnO2 NPs in argument with
in the public, commercial, or not-for-profit sectors. photocatalytic degradation of diazo dye and its kinetic studies. J. Photochem.
Photobiol. B Biol. 162, 441–447.
Hassan, M.M., Carr, C.M., 2018. A critical review on recent advancements of the removal
Competing interest statement of reactive dyes from dyehouse effluent by ion-exchange adsorbents. Chemosphere
201–209.
The authors declare no conflict of interest. Immich, A.P.S., Ulson de Souza, A.A., Ulson de Souza SM de, A.G., 2009. Removal of
Remazol Blue RR dye from aqueous solutions with Neem leaves and evaluation of
their acute toxicity with Daphnia magna. J. Hazard Mater. 164, 1580–1585.
Izadiyan, Z., Shameli, K., Miyake, M., Hara, H., Mohamad, S.E.B., Kalantari, K.,
Additional information Taib, S.H.M., Rasouli, E., 2020. Cytotoxicity assay of plant-mediated synthesized iron
oxide nanoparticles using Juglans regia green husk extract. Arab. J. Chem. 13,
No additional information is available for this paper. 2011–2023.
Jadhav, S.B., Phugare, S.S., Patil, P.S., Jadhav, J.P., 2011. Biochemical degradation
pathway of textile dye Remazol red and subsequent toxicological evaluation by
Acknowledgements cytotoxicity, genotoxicity and oxidative stress studies. Int. Biodeterior. Biodegrad. 65
(6), 733–743.
Jagathesan, G., Rajiv, P., 2018. Biosynthesis and characterization of iron oxide
The author thanks to the laboratory and staffs of Department of nanoparticles using Eichhornia crassipes leaf extract and assessing their antibacterial
Applied Chemistry and Chemical Engineering, University of Dhaka, activity. Biocatal Agric Biotechnol 13, 90–94.
Bangladesh, for providing all the facilities to carry out this study. The Juarez-Rojop, I.E., Tovilla-Zarate, C.A., Aguilar-Domínguez, D.E., Fuente, LFR de la,
Lobato-García, C.E., Ble-Castillo, J.L., L
opez-Meraz, L., Díaz-Zagoya, J.C., Bermúdez-
author also thanks to the Centre for Advanced Research in Sciences Oca~ na, D.Y., 2014. Phytochemical screening and hypoglycemic activity of carica
(CARS), University of Dhaka, Bangladesh, and Glass and Ceramics En- papaya leaf in streptozotocin-induced diabetic rats. Brazilian J Pharmacogn 24,
gineering, Bangladesh University and Engineering Technology (BUET), 341–347.
Kamaraj, M., Kidane, T., Muluken, K.U., Aravind, J., 2019. Biofabrication of iron oxide
Bangladesh, for their assistance in instrumental analysis.
nanoparticles as a potential photocatalyst for dye degradation with antimicrobial
activity. Int. J. Environ. Sci. Technol. 16, 8305–8314.
References Kansal, S.K., Kaur, N., Singh, S., 2009. Photocatalytic degradation of two commercial
reactive dyes in aqueous phase using nanophotocatalysts. Nanoscale Res. Lett.
Karpagavinayagam, P., Vedhi, C., 2019. Green synthesis of iron oxide nanoparticles using
Ahammed, K.R., Ashaduzzaman, M., Paul, S.C., et al., 2020. Microwave assisted synthesis
Avicennia marina flower extract. Vacuum 160, 286–292.
of zinc oxide (ZnO) nanoparticles in a noble approach: utilization for antibacterial
Katheresan, V., Kansedo, J., Lau, S.Y., 2018. Efficiency of various recent wastewater dye
and photocatalytic activity. SN Appl. Sci. 2, 955.
removal methods: a review. J. Environ. Chem. Eng. 6, 4676–4697.
Ahmed, E.M., 2015. Hydrogel: preparation, characterization, and applications: a review.
Khan, S., Malik, A., 2014. Environmental and health effects of textile industry
J. Adv. Res. 6 (2), 105–121.
wastewater. In: Environmental Deterioration and Human Health: Natural and
Ahmmad, B., Leonard, K., Islam, S., Kurawaki, J., Muruganandham, M., et al., 2013.
Anthropogenic Determinants, pp. 55–71.
Green synthesis of mesoporous hematite (a-Fe2O3) nanoparticles and their
Kumar, B., Smita, K., Cumbal, L., Debut, A., 2014. Biogenic synthesis of iron oxide
photocatalytic activity. Adv. Powder Technol. 24 (1), 160–167.
nanoparticles for 2-arylbenzimidazole fabrication. J. Saudi Chem. Soc. 18 (4),
Aisida, S.O., Madubuonu, N., Alnasir, M.H., et al., 2020. Biogenic synthesis of iron oxide
364–369.
nanorods using Moringa oleifera leaf extract for antibacterial applications. Appl.
Kumar, M., Mehta, A., Mishra, A., Singh, J., Rawat, M., Basu, S., 2018. Biosynthesis of tin
Nanosci. 10, 305–315.
oxide nanoparticles using Psidium Guajava leave extract for photocatalytic dye
Anchan, S., Pai, S., Sridevi, H., Varadavenkatesan, T., Vinayagam, R., Selvaraj, R., 2019.
degradation under sunlight. Mater. Lett. 215, 121–124.
Biogenic synthesis of ferric oxide nanoparticles using the leaf extract of Peltophorum
Lachheb, H., Puzenat, E., Houas, A., Ksibi, M., Elaloui, E., et al., 2002. Photocatalytic
pterocarpum and their catalytic dye degradation potential. Biocatal Agric Biotechnol
degradation of various types of dyes (Alizarin S, Crocein Orange G, Methyl red,
20, 101251.

12
Md.S.H. Bhuiyan et al. Heliyon 6 (2020) e04603

Congo red, methylene blue) in water by UV-irradiated titania. Appl. Catal. B Environ. Qasim, S., Zafar, A., Saif, M.S., Ali, Z., Nazar, M., Waqas, M., Haq, A.U., Tariq, T.,
39 (1), 75–90. Hassan, S.G., Iqbal, F., Shu, X.G., Hasan, M., 2020. Green synthesis of iron oxide
Lassoued, A., Lassoued, M.S., Dkhil, B., Ammar, S., Gadri, A., 2018. Photocatalytic nanorods using Withania coagulans extract improved photocatalytic degradation and
degradation of methylene blue dye by iron oxide (α-Fe2O3) nanoparticles under antimicrobial activity. J. Photochem. Photobiol. B Biol. 204, 111784.
visible irradiation. J. Mater. Sci. Mater. Electron. 29, 8142–8152. Reza, K.M., Kurny, A., Gulshan, F., 2017. Parameters affecting the photocatalytic
Lau, W.J., Ismail, A.F., 2009. Polymeric nanofiltration membranes for textile dye degradation of dyes using TiO2: a review. Appl. Water Sci. 7, 1569–1578.
wastewater treatment: preparation, performance evaluation, transport modelling, Robinson, T., McMullan, G., Marchant, R., Nigam, P., 2001. Remediation of dyes in textile
and fouling control - a review. Desalination 245, 321–348. effluent: a critical review on current treatment technologies with a proposed
Liu, H., Li, P., Lu, B., Wei, Y., Sun, Y., 2009. Transformation of ferrihydrite in the presence alternative. Bioresour. Technol. 77, 247–255.
or absence of trace Fe(II): the effect of preparation procedures of ferrihydrite. J. Solid Shamaila, S., Zafar, N., Riaz, S., Sharif, R., Nazir, J., Naseem, S., 2016. Gold nanoparticles:
State Chem. 182, 1767–1771. an efficient antimicrobial agent against enteric bacterial human pathogen.
L
opez Cisneros, R., Gutarra Espinoza, A., Litter, M.I., 2002. Photodegradation of an azo Nanomaterials 6 (4), 71.
dye of the textile industry. Chemosphere 48, 393–399. Shu, H.Y., Chang, M.C., Fan, H.J., 2004. Decolorization of azo dye acid black 1 by the UV/H2O2
Luo, Z., Qin, Y., Ye, Q., 2015. Effect of nano-TiO2-LDPE packaging on microbiological and process and optimization of operating parameters. J. Hazard Mater. 113, 201–208.
physicochemical quality of Pacific white shrimp during chilled storage. Int. J. Food Silva, M.C., Torres, J.A., Vasconcelos De Sa, L.R., Chagas, P.M.B., Ferreira-Leit~ao, V.S.,
Sci. Technol. 50 (7), 1567–1573. Corr^ea, A.D., 2013. The use of soybean peroxidase in the decolourization of Remazol
Mahdavi, M., Namvar, F., Ahmad, M Bin, Mohamad, R., 2013. Green biosynthesis and Brilliant Blue R and toxicological evaluation of its degradation products. J. Mol.
characterization of magnetic iron oxide (Fe 3O4) nanoparticles using seaweed Catal. B Enzym. 89, 122–129.
(Sargassum muticum) aqueous extract. Molecules 18, 5954–5964. So, C.M., Cheng, M.Y., Yu, J.C., Wong, P.K., 2002. Degradation of azo dye Procion Red
Meng, X., Ryu, J., Kim, B., Ko, S., 2016. Application of iron oxide as a pH-dependent MX-5B by photocatalytic oxidation. Chemosphere 46 (6), 905–912.
indicator for improving the nutritional quality. Clin. Nutr. Res. Soutsas, K., Karayannis, V., Poulios, I., Riga, A., Ntampegliotis, K., Spiliotis, X.,
Mitra, S., Nguyen, L.N., Akter, M., Park, G., Choi, E.H., Kaushik, N.K., 2019. Impact of Papapolymerou, G., 2010. Decolorization and degradation of reactive azo dyes via
ROS generated by chemical, physical, and plasma techniques on cancer attenuation. heterogeneous photocatalytic processes. Desalination 250, 345–350.
Cancers 11, 1030. Suresh, S., Karthikeyan, S., Jayamoorthy, K., 2016. Effect of bulk and nano-Fe2O3
Muthukumar, H., Mohammed, S.N., Chandrasekaran, N.I., Sekar, A.D., Pugazhendhi, A., particles on peanut plant leaves studied by Fourier transform infrared spectral studies
Matheswaran, M., 2019. Effect of iron doped Zinc oxide nanoparticles coating in the Effect of Fe2O3 particles on peanut plant leaves. J. Adv. Res. 7, 739–747.
anode on current generation in microbial electrochemical cells. Int. J. Hydrogen Varadavenkatesan, T., Lyubchik, E., Pai, S., Pugazhendhi, A., Vinayagam, R., Selvaraj, R.,
Energy 44, 2407–2416. 2019. Photocatalytic degradation of Rhodamine B by zinc oxide nanoparticles
Oliveira, L.C.A., Rios, R.V.R.A., Fabris, J.D., Garg, V., Sapag, K., Lago, R.M., 2002. synthesized using the leaf extract of Cyanometra ramiflora. J. Photochem. Photobiol.
Activated carbon/iron oxide magnetic composites for the adsorption of contaminants B Biol. 199, 111621.
in water. Carbon N Y 40 (12), 2177–2183. Varadavenkatesan, T., Selvaraj, R., Vinayagam, R., 2019. Dye degradation and
Paul, S.C., Bhowmik, S., Nath, M.R., Islam, M.S., Paul, S.K., Neazi, J., Monir, T.S.B., antibacterial activity of green synthesized silver nanoparticles using Ipomoea digitata
Dewanjee, S., Salam, M.A., 2020. Silver nanoparticles synthesis in a green approach: Linn. flower extract. Int. J. Environ. Sci. Technol. 16, 2395–2404.
size dependent catalytic degradation of Cationic and anionic dyes. Orient. J. Chem. Vasantharaj, S., Sathiyavimal, S., Senthilkumar, P., LewisOscar, F., Pugazhendhi, A.,
36 (3). 2019. Biosynthesis of iron oxide nanoparticles using leaf extract of Ruellia tuberosa:
Peralta-Videa, J.R., Huang, Y., Parsons, J.G., Zhao, L., Lopez-Moreno, L., Hernandez- antimicrobial properties and their applications in photocatalytic degradation.
Viezcas, J.A., Gardea-Torresdey, J.L., 2016. Plant-based green synthesis of metallic J. Photochem. Photobiol. B Biol. 192, 74–82.
nanoparticles: scientific curiosity or a realistic alternative to chemical synthesis? Yagub, M.T., Sen, T.K., Afroze, S., Ang, H.M., 2014. Dye and its removal from aqueous
Nanotechnol. Environ. Eng. 1 (1), 1–29. solution by adsorption: a review. Adv. Colloid Interface Sci. 209, 172–184.

13

You might also like