Anushree 2021

You might also like

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

Journal of Molecular Liquids 337 (2021) 116429

Contents lists available at ScienceDirect

Journal of Molecular Liquids


journal homepage: www.elsevier.com/locate/molliq

Efficient Dye Degradation via Catalytic Persulfate Activation using


Iron Oxide-Manganese Oxide Core-Shell Particle Doped with Transition
Metal Ions
C. Anushree, D. Nanda Gopala Krishna, John Philip ⇑
Corrosion Science and Technology Division, Metallurgy and Materials Group, Indira Gandhi Centre for Atomic Research, HBNI, Kalpakkam, Tamil Nadu 603102, India

a r t i c l e i n f o a b s t r a c t

Article history: Organic-pollutants released into water bodies pose serious concerns because of its detrimental effects in
Received 25 February 2021 aquatic biota, environment and economy. Water remediation using persulfate based advanced oxidation
Revised 23 April 2021 process is one of the emerging techniques. In this study, we have developed iron oxide-manganese oxide
Accepted 4 May 2021
core–shell nanoparticles doped with transition metal ions for the efficient degradation of bromophenol
Available online 11 May 2021
blue dye, for the first time. Here, the transition metal ion doped MnOx shell enables efficient persulfate
activation while the magnetic core facilitate in the effective recovery of used particles. The synthesized
Keywords:
particles were characterized by X-ray powder diffraction, transmission electron microscopy, X-ray pho-
Magnetic nanoparticles
Persulfate activation
toelectron spectroscopy and vibrating sample magnetometer. The presence of dopants is found to signif-
Amorphous manganese oxide icantly change the ratio of Mn3+ to Mn4+, reactivity and mobility of lattice oxygen in doped samples. Silver
Dye removal ion doped MnOx showed the highest dye degradation efficiency of ~91% with 1.8 wt% nanoparticles, due
to the enhanced reactivity and mobility of lattice oxygen activity of the core–shell nanoparticle in pres-
ence of Ag ions in persulfate activation. From the experimental results, the mechanism of activation of
persulfate by nanoparticles is elucidated. Our findings provide a new platform for efficient water reme-
diation using dual functional core shell magnetic nanoparticles.
Ó 2021 Elsevier B.V. All rights reserved.

1. Introduction eration capability without secondary pollution [4,13,14]. Further,


superparamagnetic particles and their dispersions have been a
Magnetic nanoparticles, particularly superparamagnetic nano- wonderful model system probe fundamental phenomena [15–17]
particles and their dispersions exhibit interesting magnetic and molecular conformational changes at interfaces [18].
stimulus dependent physio-chemical properties that are widely Among various water treatments, advanced oxidation pro-
exploited for applications in heat transfer [1–4], oil removal [5], cesses such as Fenton, photochemical, electrochemical, ultrasonic
defect sensors [6], cancer therapy [7], mechanical sealants [8], and persulfate (peroxymono sulfate and peroxydisulfate) based
cation sensors [9], MRI contrast agents [10] magnetic emulsions oxidation, ozonation, gamma ray/electron beam etc. are efficient
[11] and tunable dampers [12]. For the aforementioned applica- approaches in oxidizing or degrading organic effluents from
tions, the long term stability of dispersions and the compatibility industrial wastes [19]. Every year ~0.8 million ton of different
of surface functionalization moieties with dispersing media are dyes are produced worldwide, of which 10–15% are discharged
important. Superparamagnetic iron oxide nanoparticles, exhibiting in water bodies [20]. These effluents pose a serious threat to
high saturation magnetization, zero or negligible coercivity and human health, environment and mammalians, if untreated. Any
remanence are ideal candidate material for high gradient magnetic physiochemical method that increases the efficiency of oxidation
separation technology for waste water treatment. The superpara- capacity is termed as persulfate based advanced oxidation pro-
magnetic particles do not settle under gravitational field owing cess, where the main oxidant involved is insitu generated sulfate
to their ultra small size, non-aggregation under strong fields, avail- radical or singlet oxygen [21]. There are also reports on reaction
ability of a variety of surface functionalization options and regen- pathways, without any radical for oxidation. Persulfate based oxi-
dation is found to be more economical due to the lower material
cost, availability of a wide variety of metal oxides suitable for
⇑ Corresponding author at: Head, Smart Materials Section & Corrosion Science
persulfate activation, higher yield and less stringent operational
and Technology Division.
E-mail address: philip@igcar.gov.in (J. Philip).
parameters.

https://doi.org/10.1016/j.molliq.2021.116429
0167-7322/Ó 2021 Elsevier B.V. All rights reserved.
C. Anushree, D. Nanda Gopala Krishna and J. Philip Journal of Molecular Liquids 337 (2021) 116429

In most of the magnetic systems used for persulfate activa- 2.1.1. Synthesis of iron oxide-MnOx core–shell nanoparticles
tion, till date, have been utilizing iron oxide or iron oxide with The stock solution of carboxymethyl cellulose sodium salt was
a combination of mesoporous silica or carbon allotropes prepared by dissolving 3.75 g of polymer in 500 ml of Milli Q water.
[22–26]. Hung et al. [24] synthesized a Fe3O4-carbon black For the synthesis of magnetic nanoparticle coated with CMC,
nanostructures of particle size ranging 10–50 nm using precipi- 400 ml of stock CMC solution was transferred to 1 l flask. The solu-
tation method and used for remediation of polycyclic aromatic tion was heated to 85 °C under constant stirring at 600 rpm. Then,
hydrocarbons (PAHs) from marine sediments using persulfate. a solution containing 3.6 g of FeSO47H2O in 160 ml of water was
Song et al. [22] prepared persulfate activated Fe3O4 nanoparti- added to the reaction medium, thereby generating a yellow Fe
cles through ultrasonic-assisted reverse co-precipitation for (II)-CMC complex. In a separate 100 ml beaker, 3.36 g of KOH
removal of Cr(VI) and triclosan from aqueous solutions. Liang and 0.15 g of KNO3 were solubilized in water and heated to
et al. [27] synthesized the Fe3O4 particles loaded on mesoporous 85 °C. The hot solution was quickly transferred to the reaction
silica spheres for degradation of methylene blue using persulfate medium containing the Fe(II)-CMC complex, and kept stirring at
activation. Xie et al. [25] developed mesoporous CoFe2O4/SiO2 600 rpm for 30 min. Finally, the suspension was cooled to the room
for degradation of chlorpyrifos by combining ultrasonic treat- temperature to obtain magnetic Fe3O4@CMC particles that were
ment and persulfate activation. Liu et al. [23] investigated the separated with an external magnet. The isolated nanoparticles
degradation of azo dye using Fe3O4@activated carbon along with were dried at room temperature.
heat and ultrasound treatment. One of the major problems with Iron oxide coated MnOxnanoparticles were synthesized using
these approaches is that on activation of persulfate (without hydrothermal approach [37]. A homogeneous dispersion of Fe3-
shell), Fe2+ is converted to Fe3+ and leaches into the solution, O4@CMC was made by adding 1.25 g of dried nanoparticles in
thereby loosing the magnetic properties which makes retrieving 10 ml of Milli-Q water, followed by sonication. Then, an aqueous
the particles from the water bodies difficult. Further, the excess solution of KMnO4 was added into the reaction solution. To this
of Fe2+ quenches SO 4 radicals that reduces the efficiency of 0.15 ml of HCl was added drop wise, to produce a dark brown solu-
degradation [27]. tion. The entire solution was then transferred to a Teflon lined
The main objective of this work was to develop an innovative stainless steel autoclave and was heated at 100 °C for two hours.
strategy to overcome the drawbacks associated with the current After the reaction was complete, the obtained black precipitate
approaches and to enable persulfate activation at room tempera- was collected using an external magnet and washed with Milli-Q
ture to make the process more energy efficient. In this context, water. The iron oxide coated MnOx nanoparticles obtained after
we have developed an iron oxide-MnOx core–shell system for drying in an oven at 85 °C is named as S0 hereafter.
persulfate activation at 30 °C and demonstrated its efficient dye
degradation. The dye used in the study is bromophenol blue 2.1.2. Synthesis of iron oxide-transition metal ion doped manganese
which is an anionic dye, comes under the class of triarylmethane oxide nanoparticles
dye (TAM). TAM dyes are synthetic dyes, extensively used in tex- Fe3O4@CMC (1.25 g), AgNO3 (0.102 g) and KMnO4 (0.425 g)
tile, paper and leather industries due to its low cost [28]. It is a were dissolved in 200 ml Milli Q water. HCl (0.15 ml) was added
teratogen, carcinogenic and hence banned in aquaculture applica- to the solution drop wise and was transferred to a para polyphenol
tions [29]. liner. The resulting solution was heated in an autoclave at 100 °C
MnOx was chosen because manganese oxide based magnetic for 2 h to obtain a brown colored suspension, which was trans-
materials have been found to have applications in treatment of ferred into a beaker and was washed with Milli Q water. The
dyes [30,31], gaseous pollutants [32], heavy metal ion extraction nanoparticles was separated using a magnet and dried at 60 °C
[33], Lithium storage [34] and microwave absorption [35]. to obtain maghemite core with manganese oxide shell doped with
Although the use of transition metal ion doped manganese oxide Ag ion (hereafter referred as SA). The same procedure was used to
in advanced oxidation process was studied [36], the controlling prepare Zn2+ doped manganese oxide shell (hereafter referred as
of MnOx thickness over iron oxide, without loosing of magnetic SZ) where 0.18 g ZnNO3 is used instead of AgNO3.
properties of the core, remained as a challenge, which is very
important to separate particles after the catalysis. In the dual func- 2.2. Characterization
tional nanoparticles, MnOx is used for activation of persulfate
while the iron oxide core is used for imparting magnetic property Powder XRD measurements were performed using Rigaku
for the easy separation of particles after the use. Further, transition Ultima IV instrument, in the h range of 20 to 80° at 2°/min in the
metal ions such as Ag+ and Zn2+ were doped in MnOx to achieve Bragg-Brentano geometry with a Cu Ka source (k = 1.5418 Å).
better dye degradation efficiency through increased oxygen The acquisition and preliminary analysis of the data were per-
activity. formed using the PDXL software. High Resolution Transmission
Electron Microscope (HRTEM) images of magnetic nanoparticles
were performed by drop casting nanoparticles on top of a
2. Materials and methods carbon-coated 300 mesh copper grid using JEOL JEM 2100 with a
field emission gun operating at 200 kV. MH loop was recorded
2.1. Materials at room temperature in the field range of 1T to 1T, using a Cryo-
gen free VSM (M/s Cryogenics, UK). Surface chemical and phase
Carboxymethyl cellulose sodium salt (CMC) (Mw = 90000, composition characterization was carried out using XPS (SPECS
degree of substitution = 0.7) was obtained from Sigma Aldrich Surface Nano Analysis GmbH, Germany) with Al Ka X-ray
(Merck Co., Germany). Bromophenol blue (BPB) and potassium (1486.7 eV) source operated at 315 W power. Zeta nanosizer (Mal-
nitrate (KNO3) were purchased from Himedia (Himedia Laborato- vern ZEN3600) was used for zeta potential measurement of the
ries, Mumbai, India). Ferrous sulfate (FeSO47H2O), potassium suspension.
hydroxide (KOH), potassium permanganate (KMnO4) and
hydrochloric acid (HCl) were obtained from Lobachemie (Loba 2.3. Catalytic degradation experiments
Chemie Pvt. Ltd., Mumbai, India). All reactants were analytical
grade, and used without any treatment. Milli-Q water with resis- Dye degradation studies were performed using a Synergy HTX
tivity of 18 MXcm was used. multi-mode microplate reader. The stock solution was prepared
2
C. Anushree, D. Nanda Gopala Krishna and J. Philip Journal of Molecular Liquids 337 (2021) 116429

by the addition of 0.05 g of dye in 100 ml of Milli Q water and 0.1 g intense peak), were 25.3 ± 2, 25.7 ± 2 and 29 ± 2 nm for S0, SA
of persulfate in 3 ml of Milli Q water. For all the degradation exper- and SZ, respectively. The slight change in the crystallite size of SZ
iments, the dye concentration used is 200 mg/l, based on the detec- may be due to the lowering of edge free energy due to doping that
tion limits. S0/SA/SZ concentration was used in the steps of one can promote crystal growth [38]. The metal ions can affect the for-
from 1 to 9 g/l. Persulfate concentration is 16.5 mg/l and was kept mation of structure and surface, and when ions adsorbed to the
constant [25]. The dye solution was diluted with appropriate con- high-energy surface it can promote growth [39].
centration from the stock solution for the calibration purpose. The size and morphology of the synthesized particles were
From the stock solution, dye, persulfate and the desired concentra- studied using TEM. Fig. 2(a–b) shows TEM and HRTEM images of
tion of S0/SA/SZ were added into 10 ml beakers. After 10 min, the S0, respectively. The morphology of the particles were roughly
S0/SA/SZ was separated and collected using a magnet. The super- square shaped. The observed lattice fringe spacing of 0.243 nm,
natant solution containing left over bromophenol blue was mea- corresponding to the (3 1 1) plane of c-Fe2O3. The HTREM does
sured using a UV–Vis spectrometer. Using the calibration graph, not show a sharp contrast of presence of MnOx over c-Fe2O3 since
the concentration of residual bromophenol blue was calculated. the difference in the atomic number Z of the two materials is only
The percentage of degradation was determined using the following one unit. To further quantify the core–shell structure of the parti-
equations: cle, image processing of HRTEM of Fig. 2(b) was carried out. Fig. 2
(c) shows the pseudo-colored HRTEM image of S0, where MnOx
Co  C Ao  A
Degradation % ¼  100 ¼  100 ð1Þ shell of thickness of ~2.9 nm was clearly distinct from the crys-
Co Ao
talline core of c-Fe2O3. The MnOx shell was seen with a diffuse
where Co is the initial concentrations (mg/l), C is the concentration diffraction pattern, without any sharp fringes, which is a character-
after time t (mg/l), Ao is the initial absorbance, A is the absorbance istic of amorphous materials. Fig. 2(d) shows intensity variation in
after a time interval t at maximum absorption (kmax) of bromophe- gray scale across the length of the cross sections of c-Fe2O3/MnOx
nol blue dye. regions (shown by pink and blue lines of Fig. 2(c)), where the thick-
ness of the MnOx layer was ~2.9 nm. Fig. 2(e) shows the EDX spec-
3. Results and discussion tra of S0 with its elemental composition. The five elements present
in S0 are C, O, Mn, Fe and Cu. The 2.90 atom% of Mn confirm that
3.1. Crystal structure and morphology of iron oxide-MnOx the presence of formation of MnOx over c-Fe2O3. The Cu and C orig-
nanoparticles inate from the carbon coated Cu grid (Cu  10.25 atom %) and the
oxidized fragments of carboxymethyl cellulose, respectively. The
Fig. 1 shows the XRD pattern of maghemite core with man- average particle size obtained from the histogram distribution of
ganese oxide shell without doping (S0), maghemite core with man- S0 was found to be 36 ± 4 nm (Fig. 2(f)). The larger particle size
ganese oxide shell doped with Ag ion (SA) and Zn2+ ions (SZ) doped obtained from TEM as compared to the crystallite size is due to
nanoparticles. The diffraction peaks at 2h values of 30.1, 35.6, 43.2, presence of MnOx over the surface of c-Fe2O3.
53.4, 57.1 and 62.9° corresponds to the reflection from the planes
(2 0 0), (3 1 1), (4 0 0), (4 2 2), (5 1 1), and (4 4 0), of maghemite 3.2. X-ray photoelectron spectroscopy analysis
(c-Fe2O3) which has a cubic spinal structure (JCPDS card no.
39-1346). Sample S0 showed a broad peak at 21.4°, indicating In order to confirm the incorporation of Ag and Zn dopants into
the existence of an amorphous MnOx layer on iron oxide [34]. In the MnOx, X-ray photoelectron spectroscopy (XPS) analyses were
silver and Zn doped samples (SA and SZ), the MnOx peak at 21.4° performed on S0, SA and SZ. The elements observed from the sur-
was significantly lower. c-Fe2O3-MnOx core–shell nanoparticles vey spectrum (Fig. 3) of S0 are iron (Fe 2p  710 eV) manganese
were further characterized using HRTEM and XPS (discussed later). (Mn 2p  640 eV), carbon (C 1s  285 eV) and oxygen (O 1s 
The crystallite size (d) of c-Fe2O3 core of nanoparticles, calculated 530 eV) with atomic concentration in the ratio of 5.3 : 9.4 : 39.4 :
using Scherrer equation (d ¼ bcosh 0:9k
, where k = 1.54 Ao, h is the 45.9, respectively. In the case of SA, the elements observed in
the survey spectrum are: iron, manganese, silver (Ag 3d 
diffraction angle and b is the full width half maximum of highest
370 eV), oxygen and carbon with atomic concentration ratio of
3.6:10.4:4.2:44.3:37.5, respectively. For SZ, the elements observed
from the survey spectrum of the sample are: iron, manganese, zinc
(Zn 2p3/2–1020 eV), oxygen and carbon with atomic concentration
in the ratio of 5.2:8.9:4.5:48.0:33.4, respectively.
The high resolution spectra of S0 (Fig. 4) showed that the man-
ganese is present in both Mn3+ and Mn4+ states while the iron is
found only in the Fe3+ state. The relative concentration of Mn3+:
Mn4+ states is 6:4. The O 1s region showed the bonding states of
OAFe/Mn (67%), OAH (16%), OAC (13%) and H2O (4%), while the
C 1s region showed the bonding states of CAC/H (77%), CAO
(11%) and C@O (12%), respectively. The carbon in the nanoparticles
originate from the oxidized fragments of the carboxymethyl cellu-
lose polymer
The deconvolution of the Mn and Fe 2p peaks in SA nanoparti-
cles (Fig. 5), with multiplet structure shows that the manganese is
present in both Mn3+ (63%) and Mn4+ (37%) states while the iron is
present only in the Fe3+ state. The Ag 3d region is deconvoluted
with a single broad symmetric doublet peak structure at 367.2
(3d5/2) and 373.2 eV (3d3/2), suggesting that the silver is present
as silver oxide. The O 1s region showed the bonding states of
Fig. 1. Room temperature XRD pattern of undoped, Ag and Zn2+ ion doped OAM (64.8%), OAH (21.2%), OAC (11.4%) and H2O (2.6%), where
maghemite-manganese oxide core–shell nanoparticles, S0, SA and SZ, respectively. ‘‘M” corresponds to Fe/Mn/Ag, respectively. In the case of SZ
3
C. Anushree, D. Nanda Gopala Krishna and J. Philip Journal of Molecular Liquids 337 (2021) 116429

Fig. 2. (a) TEM and (b) HRTEM image of S0 showing the c-Fe2O3 with a lattice fringe spacing of 0.243 nm, corresponding to the (3 1 1) plane and an amorphous layer of MnOx
of ~2.9 nm (c) Pseudo-colored image processing of HRTEM image of figure (b),(d) intensity variation in gray scale across the length (pink and blue lines) of core–shell regions
of figure (c), (e) Energy dispersive X-ray spectrum and (f) particle size distribution of S0.

nanoparticles (Fig. 6), the manganese and iron are present in only that indicate the doping with Ag had increased the Mn3+ content
Mn3+ and Fe3+ states, respectively. The Zn 2p3/2 region is deconvo- in SA as compared to S0. Ran et al. [40] studied the structural
luted with a single broad symmetric peak at 1020.9 eV, suggesting and magnetic properties of pure and Co-doped Fe3O4 films and
zinc is present as Zn2+. The O 1s region showed the bonding states found that the reduction of Fe2+/Fe3+ ratio from 0.5 (for Fe3O4) to
of OAM (63.3%), OAH (21.5%), OAC (10.7%) and H2O (4.5%), where 0.2, indicate the incorporation of Co2+ in Fe3O4 lattice. Deng et al.
‘‘M” corresponds to Fe/Mn/Zn. The change in the carbon bonding [41] studied the role of structural defects in MnOx by Ag doping
states in SA and SZ, as compared to S0 was very meager. Table. 1 and observed that incorporation of silver via hydrothermal route
shows the summary of the results from XPS analysis with details caused an increase in the Mn3+ concentration by 7.5 to 20%. The
of the percentage of Fe/Mn, OAH, OAC and H2O of S0, SA and SZ. binding energy peak Oad (~530.9 eV) indicate the presence of
In all the core–shell nanoparticles (S0, SA and SZ), Fe was in +3 surface adsorbed oxygen (O 
2 or O ), OH groups and oxygen
oxidation state, indicating that the phase is c-Fe2O3. The absence of vacancies [42]. Oad was ~5.2% more in SA, as compared to S0. The
any shift in the binding energy of Fe peaks in doped samples indi- ~5.2% increase in Oad in SA, as compared to S0 is due to replace-
cates that the dopants were not present in the c-Fe2O3 core. How- ment of Mn4+ by Mn3+, which causes the removal of O2– in the
ever, Mn showed distinct variation in its oxidation state on doping: lattice (Olatt) and formation of vacancy or OH on the surface of
the ratio Mn3+/Mn4+ in S0 and SA were 1.5 and 1.7, respectively the nanoparticles, which can be represented by the equation

4
C. Anushree, D. Nanda Gopala Krishna and J. Philip Journal of Molecular Liquids 337 (2021) 116429

capping on c-Fe2O3 was amorphous, obtaining further insight into


its structure was not possible.

3.3. Magnetic properties of maghemite- manganese oxide core–shell

The room temperature magnetic properties of nanoparticles S0,


SA and SZ are shown in Fig. 7(a). The room temperature magneti-
zation curves do not show significant magnetic hysteresis. The
remanence (and coercivity) of sample S0, SA and SZ are 5.76
(0.0075), 7.1 (0.0077) and 8.17 (0.01) emu/g (T), respectively. The
saturation magnetization values of the S0, SA and SZ were found
to be ~58 ± 0.6, 49 ± 0.5 and 59 ± 0.6 emu/g, respectively, as shown
in Fig. 7(b). Addition of dopants caused a change in the vacancies
and structural distortions in the nanoparticles, which influences
the magnetic properties slightly. Mandal et al. [45] studied the
effect of vacancy concentration on the magnetic properties of
NiO core–shell system and reported that the magnetic moment
Fig. 3. XPS survey spectrum of undoped, Ag and Zn2+ion doped maghemite-
manganese oxide core–shell S0, SA and SZ, respectively. was affected by the interacting vacancies in the material and distri-
bution of vacancies in the particle volume. This was attributed to
the vacancy migration from the bulk, static disorder and change
in coordination number near the surface region. It was reported
3þ 2 
Mn4þ1ab Mnb O24ab OH4aþb , where ’a’ corresponds to fraction of that the oxygen vacancies and structural distortions causes the
Mn missing in the lattice and ’b’ is the fraction of Mn4+ replaced
4+
weakening of hybridization of oxygen and the transition metal,
by Mn3+[43]. This was further evident in the binding energy thereby weakening the double exchange interaction [46]. SA
peak ~ 529.8–530.1 eV, which is ascribed to lattice oxygen Olatt showed the lowest magnetic moment, among the three samples,
(O2
2 ), where the metal–oxygen bond is reduced from 67 to 64.8% suggesting that Ag doping caused a change in Oad and structural
in SA as compared to S0. Similarly Zn doping in SZ causes Olatt to distortion (due to the large size of silver ion) that resulted in a
drop to 63.3% and Oad to increase to 21.5%. The absence of Mn4+ weakening the double exchange interaction. The addition of
state, indicate the possibility of formation of amorphous Zn doped dopants also caused a change in the Mn3+/Mn4+ ratio as described
MnOOH/Mn2O3. Grangeon et al. [44] studied Zn sorption on dMnO2 earlier. The outer shell electronic configuration of Mn3+ and Mn4+
and reported that the amount of Zn loading impacts the coordina- ions consists of unpaired electrons i.e.3d4 and 3d3, respectively.
tion environment as well as crystallinity of MnOx. Since MnOx The larger the Mn3+/Mn4+, the higher the magnetic moment, due

Fig. 4. High resolution spectra of S0 with deconvolution after Shirley type background subtraction for (a) Fe 2p3/2 (with multiplet structure), (b) Mn 2p3/2 (with multiplet
structure), (c) O 1s and (d) C 1s core levels.

5
C. Anushree, D. Nanda Gopala Krishna and J. Philip Journal of Molecular Liquids 337 (2021) 116429

Fig. 5. High resolution spectra of SA with deconvolution after Shirley type background subtraction for (a) Fe 2p3/2 (with multiplet structure), (b) Mn 2p3/2 (with multiplet
structure), (c) O 1s and (d) Ag 2p3/2 core levels.

Fig. 6. High resolution spectra of SZ with deconvolution after Shirley type background subtraction for (a) Fe 2p3/2 (with multiplet structure), (b) Mn 2p3/2 (with multiplet
structure), (c) O 1s and (d) Zn 2p3/2 core levels.

6
C. Anushree, D. Nanda Gopala Krishna and J. Philip Journal of Molecular Liquids 337 (2021) 116429

Table 1 clinoptilolite enriched with iron to study the photocatalytic degra-


Summary of the results from XPS analysis with the percentage of Fe/Mn, OAH, OAC dation of Rhodamine B and observed gradual blue shift of peak
and H2O of S0, SA and SZ.
from 553 nm, due to rupture of the chromophore rings in
Catalyst S0 SA SZ Rhodamine B. In the current case, the loss of 592 peak and
Fe/Mn% 67 64.8 63.3 hypsochromic shift followed by decolourization unambiguously
OAH% 16 21.2 21.5 confirm the successful degradation of BPB. Similar behavior was
OAC% 13 11.4 10.7 observed for S0 and SZ as shown in Fig. 8(b). From Fig. 8(c), it
H2O% 4 2.6 4.5
can be seen that among the three samples, SA showed the highest
degradation efficiency, indicating the better persulfate activation
efficiency of SA. The degradation percentage of SA, S0 and SZ with
the one extra electron of Mn3+ [47]. However, no clear trend was 1.8 wt% particles are 91, 61 and 51%, respectively. With increase in
observed with regards to the ratio of Mn3+/Mn4+ or vacancies in the weight percentage of particle, the degradation efficiency also
the three systems, which may be due to the amorphous nature of increased in all the cases, due to increase in the activation sites.
MnOx. The highest magnetic moment was 58.5 emu/g in SZ, which SZ showed a lowest degradation efficiency with weight percentage
was slightly more than that of S0. Due to the doping with Zn in SZ, because of absence of multiple oxidation state of Mn. The catalytic
the oxidation state of manganese in the core–shell nanoparticles activity of Mn oxides depends on the oxygen mobility in the lattice
was +3 alone. c-Fe2O3 and Mn2O3 are ferrimagnetic and antiferro- (Olatt) which allows a better oxidative transformation between var-
magnetic, respectively. At the interface of c-Fe2O3 and MnOx, there ious Mn oxidation states, density of exposed reactive adsorption
exists a non uniform intermixed layer, which may contains broken sites like vacancies and facets (Oad), presence of multiple valence
bonds and uncompensated spins and hence the exchange interac- of Mn and its corresponding redox potential, specific surface area,
tions at the interfacial layer can lower the magnetization value and degree of crystallinity [53]. In the case of S0, SA and SZ, the rel-
[48]. Wang et al. [49] studied the magnetic properties in Mn2O3/ ative surface oxygen composition (Oad/Olatt) were 23.9, 32.7 and
Mn3O4 nanoclusters and reported that magnetization originates 33.9, respectively. The higher relative surface oxygen composition
from both the phases Mn2O3 and Mn3O4, and uncompensated sur- in SA and SZ is attributed to the change in coordination, vacancies
face moments of antiferromagnetic Mn2O3. Therefore, the presence and variation in oxidation state in Mn due to Ag and Zn doping.
of uncompensated interfacial moments in SZ, may be responsible Mountapmbeme et al. [54] studied the effect of addition of Co in
for a slightly higher magnetic moment compared to SA and S0. NiACo oxide to study the oxidation of unsaturated hydrocarbons
and found that increase in Co content increased the relative surface
4. Dye degradation studies oxygen composition (Oad/Olatt) in the oxide. The absence of multi-
ple oxidation state in SZ (only +3) can be the reason for the lowest
The dye degradation and removal capability of S0, SA and SZ degradation efficiency in this sample. Fig. 8(d) shows the plot of ln
were studied using BPB dye. The degradation of BPB via persul- (A/Ao) as a function of time. Langmuir–Hinshelwood mechanism
phate activator using the samples S0, SA and SZ were studied using explains the kinetics of heterogeneous catalytic processes, repre-
UV–Vis spectroscopy by varying the wt% of the catalyst (S0, SA and sented by [55]
SZ). Fig. 8(a) shows two peaks 436 and 592 nm for 0.1ɸ and 0.5ɸ of dC kKC
SA (ɸ = 0.33 wt%) i.e. at very low concentration of SA catalyst. The r¼ ¼ ð2Þ
dt 1 þ KC
peaks at 436 nm and 592 nm were due to chromophore of the BPB
dye. With increase in the weight percentage, the peak at 592 disap- where r is the degradation rate of dye (mg/l min), concentration of
pears and the maximum absorption peak was shifted to a lower dye (mg/l), t is time, K is the adsorption coefficient of the dye (l/mg)
wavelength ~400 nm (hypsochromic shift), due to the breakage and k is the reaction rate constant (mg/l min). Under limiting con-
of the BPB structure. Li et al. [50] studied the electochemical degra- ditions, when KC1, the rate of degradation follows pseudo first
dation of BPB using boron doped diamond anode and reported a order kinetics and the equation is given by
faster decrease of 592 nm, peak intensity as compared to other
dC
peak during electrolysis due to the rupture of benzene rings. Jones ¼ kKC ¼ kapp C ð3Þ
dt
et al. [51] studied the Rhodamine B dye degradation by Ta3N5/
W18O49 nanostructures and observed a hypsochromic shift from where kapp = kK represents apparent first order rate constant. Fur-
554 to 532 nm, which was attributed to the removal of ethyl ther absorbance (A) is proportional to concentration (C) given by
groups of Rhodamine B molecules. Sydorchuk et al. [52] developed the equation

Fig.7. (a) Room temperature magnetization curves of maghemite- MnOx nanoparticles S0, SA and SZ and (b) Enlarged view of the boxed region.

7
C. Anushree, D. Nanda Gopala Krishna and J. Philip Journal of Molecular Liquids 337 (2021) 116429

Fig. 8. (a) Absorbance of BPB containing different wt% of SA (ɸ= 0.33 wt%) (b) Absorbance spectra of S0, SA and SZ at 1 and 3ɸ (c) percentage of degradation of S0, SA and SZ
with different wt% at equilibrium and (d) plot of ln(A/Ao) as a function of time; Inset of (d) shows the structure of bromophenol blue.

A sis of doping with silver) in SA to determine the maximum


ln ¼ kapp t ð4Þ
Ao percentage of degradation efficiency of BPB. Fig. 9(a) shows that
the percentage of degradation for various KMnO4:AgNO3 reactant
The kapp value obtained from the best fit is 0.0364 min1 with
ratio of SA nanoparticles. The dye percentage of degradation using
R2 = 0.95. The obtained kapp value is comparable with other reports
1.6 wt% of SA nanoparticles containing 1:0.1, 1:0.25, 1:0.5, 1:0.75
[23,56]. The structure of BPB is shown in Fig. 8(d) inset. BPB con-
of KMnO4:AgNO3 (0.2, 0.6, 1.3, 1.9 mmol of AgNO3, respectively)
tains two hydroxyl groups and at low pH, the hydroxyl groups
were 75, 90.6, 92.4 and 64%, respectively. The degradation percent-
remain neutral. With increase in pH, one of the hydroxyl groups
age of BPB is found to increase with SA when doped with silver
is deprotanated and becomes anionic in nature. At very high pH,
using 1:0.5 KMnO4:AgNO3 reactant. Further increase in silver
both OH groups exist as dianionic form [57]. The BPB stock solution
nitrate led to a reduction of degradation percentage. Fig. 9(b)
was prepared at neutral pH. After the addition of persulfate and the
shows that with increase in weight percentage using SA nanopar-
nanoparticles, the pH has changed to ~3.82, which indicate the
ticles, containing increasing concentration of silver nitrate reac-
attributed to activation of persulfate. This causes the production
tant, the percentage of degradation also increases. However, SA
of SO +
4 which reacts with the solvent (water) and releases H there
nanoparticles containing 1:0.5 and 1:0.75 of KMnO4:AgNO3 reac-
by reducing the pH,[23] as shown by the equation (5)
tant does not follow a regular trend with increase in wt%, which
can be due to reduced dispersion stability as compared to lower
SO  þ 2
4 þ H2 O ! OH þ H þ SO4 ð5Þ
doping of silver. Fig. 9(c) showed the zeta potential of SA of differ-
No additional pH adjustment was done and the solution was ent doping of silver. The zeta potential of SA containing 1:0.1,
used as such. In Fenton and in most of the advanced oxidation pro- 1:0.25, 1:0.5, 1:0.75 of KMnO4:AgNO3 were 35, 38, 20 and
cesses, the optimal pH used in the reaction ranges from 2 to 4. Liu 19 mV, respectively. The system with higher the zeta potential
et al. [23] studied the degradation of acid red 73 by activated per- values are better electrostatically stabilized against the attractive
sulfate using Fe3O4@activated carbon in an ultrasonic bath at 50 °C forces where agglomeration and flocculation are minimal. In addi-
and reported that the decolourization reaction rate was fast when tion, magnetic nanoparticles have a tendency to aggregate due to
pH was 3. At pH > 7, more hydroxyl radicals are generated from dipolar interactions and zeta potential greater than ± 30 mV are
sulfate radicals, which is followed by the scavenging reaction considered stable. The highest zeta potential (-37.7 mV) was seen
between hydroxyl radical and sulfate radicals. for SA prepared with 1:0.25 KMnO4:AgNO3 reactant. It should be
Experiments were also performed using SA nanoparticles con- noted that SA nanoparticles used in all other experimental analysis
taining different KMnO4:AgNO3 reactant ratio (used in the synthe- contained 0.6 mmol of silver nitrate reactant i.e. 1:0.25 of KMnO4:
8
C. Anushree, D. Nanda Gopala Krishna and J. Philip Journal of Molecular Liquids 337 (2021) 116429

Fig. 9. (a) Percentage of degradation of BPB with 1.6 wt% of SA catalyst using various KMnO4:AgNO3 reactant ratio used in the synthesis (b) percentage of degradation of BPB
with wt% using SA nanoparticles containing different concentration of silver nitrate reactant (c) zeta potential of SA and (d) percentage of degradation on recycling using
1.8 wt% of SA (1:0.25 KMnO4:AgNO3).

AgNO3 reactant because of its excellent stability. To study the bon with atomic concentration in the ratio of 6.2:8.3:1.0:47.7:36.8,
reusability of the SA nanoparticles for BPB degradation, recycling respectively. The Zn 2p3/2 region is deconvoluted with a single
experiments were performed. The SA nanoparticles were reused broad symmetric peak at 1020.9 eV is shown in Fig. 10(c). The
by washing nanoparticles using hexane and ethanol. Fig. 9(d) amount of Zn was reduced from 4.5 to 1% in SZ nanoparticles
showed the BPB degradation efficiency for different cycles using due to leaching after the dye degradation through persulfate
1.8 wt% of SA (0.6 mmol of silver nitrate). BPB degradation effi- activation.
ciency after cycles 1–5 was ~91, 58.3, 35.6, 30.9 and 29.5%, respec- Fig. 11(a&b) shows the high resolution spectra of SA after dye
tively. This reduction degradation efficiency was attributed to the degradation through activation of persulfate of Mn 2p3/2 (with
leaching of silver ions on repeated use. Degradation studies were multiplet structure) and O 1s core levels, respectively. The man-
also performed using other dyes like Rhodamine B and crystal vio- ganese was present in both Mn3+ (72%) and Mn4+ (28%) states.
let. But on degradation, multiple peaks were observed indicating The O 1s region showed the bonding states of OAM (58.5%),
the formation of complex intermediates. Further studies are OAH (25.1%), OAC (13.4%) and H2O (3.0%), where ‘‘M” corresponds
required to understand the intermediates. to Fe/Mn. Fig. 11(c&d) showed that the manganese is present in
both Mn3+ (79%) and Mn4+ (21%) states while zinc is found to be
4.1. Mechanism of dye degradation only in Zn2+ states. The O 1s region showed the bonding states of
OAM (60.3%), OAH (23.9%), OAC (12.0%) and H2O (3.8%), where
To obtain further insight into the mechanism of activation of ‘‘M” corresponds to Fe/Mn/Zn. The oxidation state of iron was
persulfate by the nanoparticles, XPS analysis was performed on retained Fe3+ state in both the cases.
the core–shell nanoparticles (SZ and SA) after the degradation of Table. 2 shows the summary of the XPS analysis of peak
BPB dye. SZ and SA were recycled 4 times before performing the ~529.8–530.1 eV, where the percentage of lattice oxygen Olatt
analysis. (O2
2 ), Oad (~530.9 eV), percentage of Mn
3+/4+
of SA and SZ, before
The elements observed from the survey spectrum (Fig. 10(a)), and after dye degradation is listed. SA and SZ showed a change of
after the degradation of BPB dye through activation of persulfate Olatt of ~6.3 and 3%, respectively. This findings suggest an
of SA are: iron, manganese, oxygen and carbon with atomic con- enhanced reactivity and mobility of lattice oxygen activity in SA
centration in the ratio of 6.1:7.9:44.9:41.1, respectively. The ele- as compared to SZ and Oad corresponding to surface adsorbed
ment silver was not observed, which indicates that silver was oxygen (O 
2 or O ), OH groups and oxygen vacancies in SA,
fully leached into the solution after the experiment. In the case (3.9% as compared to 2.4%) in SZ. This indicates that surface oxy-
of SZ shown in Fig. 10(b), the elements observed from the survey gen species in the vicinity of AgAOAMn entities are more active
spectrum of the sample are: iron, manganese, zinc, oxygen and car- than those near ZnAOAMn entities [41]. Deng et al. [41] studied
9
C. Anushree, D. Nanda Gopala Krishna and J. Philip Journal of Molecular Liquids 337 (2021) 116429

(c)

2.0x104

1.9x104

1023 1020
Binding energy (eV)

Fig. 10. XPS survey spectrum of (a) SA (b) SZ after the degradation of bromophenol blue dye through persulfate activation and (c) high resolution spectra of Zn 2p3/2 core
levels of SZ with deconvolution after Shirley type background.

Fig. 11. High resolution spectra of SA and SZ after dye degradation through activation of persulfate with deconvolution after Shirley type background subtraction for (a&c)
Mn 2p3/2 (with multiplet structure) and (b&d) O 1s core levels, respectively.

Table 2
Results of XPS analysis of peak ~ 529.8–530.1 eV. The percentage of lattice oxygen Olatt (O2
2 ), Oad (~530.9 eV), percentage of Mn
3+/4+
of SA and SZ, before and after dye degradation
are given.

Pristine nanoparticles After dye degradation


Catalysts Olatt(%) Oad(%) Mn3+(%) Mn4+(%) Olatt(%) Oad(%) Mn3+(%) Mn4+(%)
SA 64.8 21.2 63 37 58.5 25.1 72 28
SZ 63.3 21.5 100 0 60.3 23.9 79 21

the O3 decomposition by Ag doped MnOx and reported that silver in lattice oxygen activity. This was because the addition of dopant
addition can activate the lattice oxygen mobility due to decrease reduces the energy for the removal of lattice oxygen compared to
in the coordination number with MnAO. Hou et al. [58] studied undoped ones. Therefore, it is concluded that the observed higher
the effect of benzene oxidation with and without doping of K+ catalytic activity of SA, as compared to SZ, is due to better redox
cryptomalene sieves and observed that doping caused an increase cycling of Mn3+/4+, higher activity of lattice oxygen and surface
10
C. Anushree, D. Nanda Gopala Krishna and J. Philip Journal of Molecular Liquids 337 (2021) 116429

oxygen species. Further, the single oxidation state of Mn (3+) of 5. Conclusion


SZ and the ZnAO bond in MnOx (thermodynamically favorable)
would have reduced the overall mobility and active surface sites Iron oxide capped MnOx core–shell type nanoparticles, doped
on the MnOx surface in SZ [59]. Therefore, the possible mecha- with transition metal ions(Ag, Zn) were synthesized using
nism of activation of persulfate on S0, SA and SZ nanoparticles oxidation-precipitation followed by hydrothermal synthesis. The
is adsorption of persulfate ion on Oad site. The presence of Oad iron oxide core with diameter of ~25–29 nm with an amorphous
sites is one of the prerequisites for the proximity of persulfate MnOx shell of ~2.9 nm is prepared using this approach. The core–
to adsorb on the surface [60]. Oad sites are formed due to insuffi- shell structure without doping (S0), doping with Ag (SA) and
cient coordination in the metal oxide lattice [43]. Such sites have Zn2+ ion (SZ) were tested for their efficacy in dye degradation
lower lattice stabilization energy and higher chemical reactivity. through activation of persulfate. XPS analysis revealed that doping
In the nanoparticles, Mn3+ oxidizes to Mn4+ and transfer the elec- of transition metal ions caused the change in the ratio Mn3+:Mn4+
tron to persulfate ion as shown schematically in Fig. 12(a). The with the order S0 < SA < SZ, where SZ showed only Mn3+ state. Sim-
peroxide bond and the hydroxyl bond of persulfate is cleaved to ilarly the saturation magnetization values was highest for SZ
form SO 
4 and OH radicals. To maintain the charge balance, the (58.5 emu/g) because of exchange interactions at the interfacial
lattice oxygen converts to O2 by donating electrons to two Mn4+ layer. All the nanoparticles showed superparamagnetic behavior
to form Mn3+. This results in decrease in Olatt, formation of with negligible coercivities and remanence, making them ideal
vacancy or OH, which in described by the equation by the equa- for magnetic separation. Efficiency of dye degradation was tested
3þ 2 
tion Mn4þ 1ab Mnb O24ab OH4aþb , where ’a’ corresponds to fraction
using bromophenol blue dye. SA, S0 and SZ showed degradation
of Mn missing in the lattice and ’b’ is the fraction of Mn4+
4+ percentage of 91, 61 and 51% with 1.8 wt% of nanoparticles. SZ
replaced by Mn3+. This causes an increase in the Oad site, which showed the lowest degradation efficiency due to lack of multiple
is in agreement with the XPS results. oxidation state of Mn. The kinetics of degradation was determined
Fig. 12(b-d) shows the photographs of samples showing dye using Langmuir-Hinshelwood plot where the apparent first order
degradation steps using iron oxide-Ag doped MnOx(SA) magnetic rate constant value is found to be 0.0364 min1. The XPS analysis
nanoparticles. Fig. 12(b) shows the photograph of BPB at pH 7. revealed that the activity of SA catalyst is due to enhanced reactiv-
Addition of magnetic nanoparticles and persulfate turns the solu- ity and mobility of lattice oxygen in doped samples. This approach
tion to dark brown as shown in Fig. 12(c). The formation of bubbles provides a new platform for efficient water remediation using dual
on top of the vial is due to gaseous products formed due to BPB dye function magnetic nanoparticles.
degradation. Fig. 12(d) shows the complete discoloration of dye,
indicating the efficient degradation of dye by SA nanoparticles. Funding
These results suggest that the present approach is a facile and
robust way for degradation of BPB dye and easy separation of par- This research did not receive any specific grant from funding
ticles using a magnet. agencies in the public, commercial, or not-for-profit sectors.

(a) 2- CO2+ byproducts


S2O8

SO4
.

O2-

O2

(b) (c) (d)

Fig. 12. (a) Schematic depicting the core–shell structure and the dye degradation mechanism. Photographs of the samples showing different steps involved in the dye
degradation:(b) BPB solution alone (c) dispersion of SA nanoparticles and persulfate into dye solution (d) separation of nanoparticles after dye degradation using a magnet.

11
C. Anushree, D. Nanda Gopala Krishna and J. Philip Journal of Molecular Liquids 337 (2021) 116429

CRediT authorship contribution statement [22] X. Song, C. Ren, Q. Zhao, B. Su, Simultaneous removal of Cr(VI) and triclosan
from aqueous solutions through Fe3O4 magnetic nanoscale-activated
persulfate oxidation, Chem. Eng. J. 381 (2020) 122586.
C. Anushree: Conceptualization, Investigation, Methodology, [23] L. Liu, C. Yang, W. Tan, Y. Wang, Degradation of acid red 73 by activated
Writing - original draft. D. Nanda Gopala Krishna: Investigation, persulfate in a heat/Fe3O4@AC system with ultrasound intensification, ACS
Omega 5 (2020) 13739–13750.
Formal analysis. John Philip: Conceptualization, Supervision, Writ-
[24] C.M. Hung, C.P. Huang, S.S. Lam, C.W. Chen, C.D. Dong, The removal of
ing - review & editing. polycyclic aromatic hydrocarbons (PAHs) from marine sediments using
persulfate over a nano-sized iron composite of magnetite and carbon black
activator, J. Environ. Chem. Eng. 8 (2020) 104440.
Declaration of Competing Interest [25] H. Xie, W. Xu, Enhanced activation of persulfate by meso-CoFe2O4/SiO2 with
ultrasonic treatment for degradation of chlorpyrifos, ACS Omega 4 (2019)
17177–17185.
The authors declare that they have no known competing finan-
[26] R.M. Hassan, Prospective mechanistic on electron-transfer nature for
cial interests or personal relationships that could have appeared reduction of permanganate ion by ascorbic acid in aqueous acidic solutions,
to influence the work reported in this paper. J. Mol. Liq. 309 (2020) 113154.
[27] S. Liang, Z. Ziyu, J. Han, D. Xiaoyan, Facile synthesis of magnetic mesoporous
silica spheres for efficient removal of methylene blue via catalytic persulfate
Acknowledgement activation, Sep. Purif. Technol. 256 (2021) 117801.
[28] D.F. Duxbury, The photochemistry and photophysics of triphenylmethane dyes
in solid and liquid media, Chem. Rev. 93 (1993) 381–433.
The authors wish to thank Dr. A. K. Bhaduri, Dr. Shaju K. Albert [29] N. López Gutiérrez, R. Romero González, J.L. Martínez Vidal, A.G. Frenich,
and Dr. S. Raju for their support and encouragement. Analysis of triphenylmethane dyes in seafood products: a review of extraction
methods and determination by liquid chromatography coupled to mass
spectrometry, Anal. Methods 5 (2013) 3434–3449.
References [30] Y.G. Kang, H. Yoon, C.S. Lee, E.J. Kim, Y.S. Chang, Advanced oxidation and
adsorptive bubble separation of dyes using MnO2-coated Fe3O4
nanocomposite, Water Res. 151 (2019) 413–422.
[1] P.D. Shima, J. Philip, Tuning of thermal conductivity and rheology of nanofluids
[31] Q. Yang, H. Song, Y. Li, Z. Pan, M. Dong, F. Chen, Z. Chen, Flower-like core-shell
using an external stimulus, J. Phys. Chem. C 115 (2011) 20097–20104.
Fe3O4@MnO2 microspheres: synthesis and selective removal of congo red dye
[2] H. Hong, B. Wright, J. Wensel, S. Jin, X.R. Ye, W. Roy, Enhanced thermal
from aqueous solution, J. Mol. Liq. 234 (2017) 18–23.
conductivity by the magnetic field in heat transfer nanofluids containing
[32] F. Arena, R. Di Chio, A. Palella, L. Spadaro, L. Frusteri, B. Fazio, S. Morandi, G.
carbon nanotube, Synth. Met. 157 (2007) 437–440.
Martra, Tailoring manganese oxide catalysts for the total oxidation of
[3] M.J.P. Gallego, L. Lugo, J.L. Legido, M.M. Pineiro, Enhancement of thermal
pollutants in gas and liquid phase, Appl. Catal. A: Gen. 610 (2021) 117917.
conductivity and volumetric behavior of FexOy nanofluids, J. Appl. Phys. 110
[33] S. Song, S. Zhang, S. Huang, R. Zhang, L. Yin, Y. Hu, T. Wen, L. Zhuang, B. Hu, X.
(2011) 014309.
Wang, A novel multi-shelled Fe3O4@MnOx hollow microspheres for
[4] R. Lenin, P.A. Joy, Role of primary and secondary surfactant layers on the
immobilizing U(VI) and Eu(III), Chem. Eng. J. 355 (2019) 697–709.
thermal conductivity of lauric acid coated magnetite nanofluids, J. Phys. Chem.
[34] C. Zeng, W. Weng, T. Lv, W. Xiao, Low-temperature assembly of ultrathin
C 120 (2016) 11640–11651.
amorphous MnO2 nanosheets over Fe2O3 spindles for enhanced lithium
[5] M.V.G. Paixão, R.C.S. da Luz, R.d.C. Balaban, Thermoresponsive polymer
storage, ACS Appl. Mater. Interfaces 10 (2018) 30470–30478.
brushes on magnetic chitosan microspheres: synthesis, characterization and
[35] M. Qiao, X. Lei, Y. Ma, L. Tian, K. Su, Q. Zhang, Dependency of tunable
application in oily water of high salinity, J. Mol. Liq. 286 (2019).
microwave absorption performance on morphology-controlled hierarchical
[6] V. Mahendran, J. Philip, Spectral response of magnetic nanofluid to toxic
shells for core-shell Fe3O4@MnO2 composite microspheres, Chem. Eng. J. 304
cations, Appl. Phys. Lett. 102 (2013) 163109.
(2016) 552–562.
[7] A. Farzin, S.A. Etesami, J. Quint, A. Memic, A. Tamayol, Magnetic nanoparticles
[36] J. Huang, H. Zhang, Mn-based catalysts for sulfate radical-based advanced
in cancer therapy and diagnosis, Adv. Healthcare Mater. 9 (2020) 1901058.
oxidation processes a review, Environ. Int. 133 (2019) 105141.
[8] B.K. Sreedhar, R.N. Kumar, P. Sharma, S. Ruhela, J. Philip, S.I. Sundarraj, N.
[37] C. Anushree, D.N.G. Krishna, J. Philip, Oil-absorbent MnOx capped iron oxide
Chakraborty, M. Mohana, V. Sharma, G. Padmakumar, B.K. Nashine, K.K. Rajan,
nanoparticles: synthesis, characterization and applications in oil recovery, J.
Development of active magnetic bearings and ferrofluid seals towards oil free
Mol. Liq. 320 (2020) 114324.
sodium pumps, Nucl. Eng. Des. 265 (2013) 1166–1174.
[38] H.E. Lundager Madsen, Influence of foreign metal ions on crystal growth and
[9] V. Mahendran, J. Philip, Influence of Ag+ interaction on 1D droplet array
morphology of brushite (CaHPO4, 2H2O) and its transformation to octacalcium
spacing and the repulsive forces between stimuli-responsive nanoemulsion
phosphate and apatite, J. Cryst. Growth 310 (2008) 2602–2612.
droplets, Langmuir 30 (2014) 10213–10220.
[39] R. Liu, Y. Jiang, H. Fan, Q. Lu, W. Du, F. Gao, Metal ions induce growth and
[10] T. Neuberger, B. Schopf, H. Hofmann, M. Hofmann, B.V. Rechenberg,
magnetism alternation of a-Fe2O3 crystals bound by high-index facets, Chem.
Superparamagnetic nanoparticles for biomedical applications: Possibilities
Eur. J. 18 (2012) 8957–8963.
and limitations of a new drug delivery system, J. Magn. Magn. Mater. 283
[40] F.Y. Ran, Y. Tsunemaru, T. Hasegawa, Y. Takeichi, A. Harasawa, K. Yaji, S. Kim, A.
(2005) 483–496.
Kakizaki, Valence band structure and magnetic properties of Co-doped
[11] J. Philip, G.G. Prakash, T. Jaykumar, P. Kalyanasundaram, B. Raj, Stretching and
Fe3O4(100) films, J. Appl. Phys. 109 (2011) 123919.
collapse of neutral polymer layers under association with ionic surfactants,
[41] H. Deng, S. Kang, J. Ma, L. Wang, C. Zhang, H. He, Role of Structural Defects in
Phys. Rev. Lett. 89 (2002) 268301.
MnOx promoted by Ag doping in the catalytic combustion of volatile organic
[12] K.H. Song, B.J. Park, H.J. Choi, Effect of magnetic nanoparticle additive on
compounds and ambient decomposition of O3, Environ. Sci. Technol. 53 (2019)
characteristics of magnetorheological fluid, IEEE Trans. Magn. 45 (2009) 4045–
10871–10879.
4048.
[42] D. Chen, D. He, J. Lu, L. Zhong, F. Liu, J. Liu, J. Yu, G. Wan, S. He, Y. Luo,
[13] S. Mirshahghassemi, A.D. Ebner, B. Cai, J.R. Lead, Application of high gradient
Investigation of the role of surface lattice oxygen and bulk lattice oxygen
magnetic separation for oil remediation using polymer-coated magnetic
migration of cerium-based oxygen carriers: XPS and designed H2-TPR
nanoparticles, Sep. Purif. Technol. 179 (2017) 328–334.
characterization, Appl. Catal. B Environ. 218 (2017) 249–259.
[14] A. Ditsch, S. Lindenmann, P.E. Laibinis, D.I.C. Wang, T.A. Hatton, High-gradient
[43] H. Abbas, S.A. Nasser, Hydroxyl as a defect of the manganese dioxide lattice
magnetic separation of magnetic nanoclusters, Ind. Eng. Chem. Res. 44 (2005)
and its applications to the dry cell battery, J. Power Sources 58 (1996) 15–21.
6824–6836.
[44] S. Grangeon, A. Manceau, J. Guilhermet, A.C. Gaillot, M. Lanson, B. Lanson, Zn
[15] A.O. Ivanov, P.J. Camp, Theory of the dynamic magnetic susceptibility of
sorption modifies dynamically the layer and interlayer structure of vernadite,
ferrofluids, Phys. Rev. E 98 (2018) 050602.
Geochim. Cosmochim. Acta 85 (2012) 302–313.
[16] A.Y. Solovyova, E.A. Elfimova, A.O. Ivanov, P.J. Camp, Modified mean-field
[45] S. Mandal, S. Banerjee, K.S.R. Menon, Core-shell model of the vacancy
theory of the magnetic properties of concentrated, high-susceptibility,
concentration and magnetic behavior for antiferromagnetic nanoparticle,
polydisperse ferrofluids, Phys. Rev. E 96 (2017) 052609.
Phys. Rev. B 80 (2009) 214420.
[17] R.F.G. Apóstolo, G. Tsagkaropoulou, P.J. Camp, Molecular adsorption, self-
[46] Q. Zhang, Z.F. Xu, L.F. Wang, S.H. Gao, S.J. Yuan, Structural and electromagnetic
assembly, and friction in lubricants, J. Mol. Liq. 277 (2019) 606–612.
properties driven by oxygen vacancy in Sr2FeMoO6d double perovskite, J.
[18] A.W. Zaibudeen, J. Philip, Adsorption of bovine serum albumin at oil-water
Alloys Compd. 649 (2015) 1151–1155.
interface in the presence of polyelectrolytes and nature of interaction forces,
[47] C.Y. Ouyang, Ž. Šljivančanin, A. Baldereschi, Transition from Mn4+ to Mn3+
Colloids Surf. A Physicochem. Eng. Asp. 566 (2019) 38–47.
induced by surface reconstruction at k-MnO2(001), J. Chem. Phys. 133 (2010)
[19] J. Wang, S. Wang, Reactive species in advanced oxidation processes: formation,
204701.
identification and reaction mechanism, Chem. Eng. J. 401 (2020) 126158.
[48] E. Skoropata, R.D. Desautels, C.C. Chi, H. Ouyang, J.W. Freeland, J. van Lierop,
[20] O.S. Bayomie, H. Kandeel, T. Shoeib, H. Yang, N. Youssef, M.M.H. El Sayed,
Magnetism of iron oxide based core-shell nanoparticles from interface mixing
Novel approach for effective removal of methylene blue dye from water using
with enhanced spin-orbit coupling, Phys. Rev. B 89 (2014) 024410.
fava bean peel waste, Sci. Rep. 10 (2020) 7824.
[49] Z.H. Wang, D.Y. Geng, W.J. Hu, W.J. Ren, Z.D. Zhang, Magnetic properties and
[21] J. Lee, U. von Gunten, J.H. Kim, Persulfate-based advanced oxidation: critical
exchange bias in Mn2O3∕Mn3O4 nanoclusters, J. Appl. Phys. 105 (2009)
assessment of opportunities and roadblocks, Environ. Sci. Technol. 54 (2020)
07A315.
3064–3081.

12
C. Anushree, D. Nanda Gopala Krishna and J. Philip Journal of Molecular Liquids 337 (2021) 116429

[50] W. Li, B. Li, W. Ding, J. Wu, C. Zhang, D. Fu, Response surface methodology as a [56] Y. Huang, X. Tian, Y. Nie, C. Yang, Y. Wang, Enhanced peroxymonosulfate
tool to optimize the electrochemical incineration of bromophenol blue on activation for phenol degradation over MnO2 at pH 3.5–9.0 via Cu(II)
boron-doped diamond anode, Diam. Relat. Mater. 50 (2014) 1–8. substitution, J. Hazard. Mater. 360 (2018) 303–310.
[51] D.R. Jones, V. Gomez, J.C. Bear, B. Rome, F. Mazzali, J.D. McGettrick, A.R. Lewis, [57] M. Chattopadhyaya, N.A. Murugan, Z. Rinkevicius, Origin of the absorption
S. Margadonna, W.A. AlMasry, C.W. Dunnill, Active removal of waste dye band of bromophenol blue in acidic and basic pH: insight from a combined
pollutants using Ta3N5/W18O49 nanocomposite fibres, Sci. Rep. 7 (2017) 4090. molecular dynamics and TD-DFT/MM study, J. Phys. Chem. A 120 (2016)
[52] V. Sydorchuk, V. Vasylechko, O. Khyzhun, G. Gryshchouk, S. Khalameida, L. 7175–7182.
Vasylechko, Effect of high-energy milling on the structure, some [58] J. Hou, L. Liu, Y. Li, M. Mao, H. Lv, X. Zhao, Tuning the K+ concentration in the
physicochemical and photocatalytic properties of clinoptilolite, Appl. Catal. tunnel of OMS-2 nanorods leads to a significant enhancement of the catalytic
A: Gen. 610 (2021) 117930. activity for benzene oxidation, Environ. Sci. Technol. 47 (2013) 13730–13736.
[53] W.D. Oh, Z. Dong, T.T. Lim, Generation of sulfate radical through [59] T. Xiong, Z.G. Yu, H. Wu, Y. Du, Q. Xie, J. Chen, Y.W. Zhang, S.J. Pennycook, W.S.
heterogeneous catalysis for organic contaminants removal current V. Lee, J. Xue, Defect engineering of oxygen-deficient manganese oxide to
development, challenges and prospects, Appl. Catal. B Environ. 194 (2016) achieve high-performing aqueous zinc ion battery, Adv. Energy Mater. 9
169–201. (2019) 1803815.
[54] P. Mountapmbeme Kouotou, M. Waqas, A. El Kasmi, Z. Atour, Z.Y. Tian, [60] W.D. Oh, Z. Dong, T.T. Lim, Generation of sulfate radical through
Influence of Co addition on Ni-Co mixed oxide catalysts toward the deep heterogeneous catalysis for organic contaminants removal: current
oxidation of low-rank unsaturated hydrocarbons, Appl. Catal. A: Gen. 612 development, challenges and prospects, Appl. Catal. B Environ. 194 (2016)
(2021). 169–201.
[55] R.J. Baxter, P. Hu, Insight into why the Langmuir-Hinshelwood mechanism is
generally preferred, J. Chem. Phys. 116 (2002) 4379–4381.

13

You might also like