Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Journal of Environmental Chemical Engineering 9 (2021) 106590

Contents lists available at ScienceDirect

Journal of Environmental Chemical Engineering


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

Kinetic analysis of nitrophenol reduction and colourimetric detection of


hydrogen peroxide based on gold nanoparticles catalyst biosynthesised
from Cynomorium songaricum
Van-Dat Doan a, Quoc-Huy Pham a, Bao-An Huynh a, Thi-Lan-Huong Nguyen b,
Anh-Tien Nguyen c, Thanh-Danh Nguyen d, e, *
a
Faculty of Chemical Engineering, Industrial University of Ho Chi Minh City, 12 Nguyen Van Bao, Ho Chi Minh City, Vietnam
b
Institute of Biotechnology and Food Technology, Industrial University of Ho Chi Minh City, 12 Nguyen Van Bao, Ho Chi Minh City, Vietnam
c
Faculty of Chemistry, Ho Chi Minh City University of Education, 280 An Duong Vuong, Ho Chi Minh City, Vietnam
d
Institute of Chemical Technology, Vietnam Academy of Science and Technology, 1A, TL29, Thanh Loc Ward, District 12, Ho Chi Minh City, Vietnam
e
Graduate University of Science and Technology, Vietnam Academy of Science and Technology, 18 Hoang Quoc Viet, Cau Giay District, Hanoi, Vietnam

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

Editor: Despo Kassinos The pseudo-first-order reaction kinetics of metallic nanocatalyst for the model reaction of nitrophenols/NaBH4
were early reported on low reaction temperatures. Here, we investigated reduction of nitrophenols in a wide
Keywords: range of reaction temperature with gold nanoparticles (AuNPs) catalyst that was fabricated via a low-cost and
Nitrophenol reduction effective method using aqueous extract of Cynomorium songaricum (CS). The biosynthesised CS-AuNPs were well
Kinetic
characterised by the analytic techniques such as X-ray diffraction (XRD), transmission electron microscope
Colourimetric detection
(TEM), scanning transmission electron microscopy (STEM). The TEM images showed multiple shapes and size in
Catalysis
Hydrogen peroxide a range of 3–30 nm with an average size of 16 nm. The colloidal solutions were stored at the room temperature
Gold nanoparticles for 30 days with a stability of 63%. The kinetic study of the NiPs reduction showed that the reaction occurred via
Langmuir–Hinshelwood mechanism with a half-order kinetic model for all tested reaction temperatures (30–70

C). The thermodynamic parameters including activation energy, enthalpy, entropy and Gibbs free energy were
calculated. The activation energies in the reduction of 2-nitrophenol and 3-nitrophenol were found to be as low
as 4.88 kJ mol− 1 and 8.66 kJ mol− 1, respectively. Moreover, CS-AuNPs were used as a nanozyme for the
detection of hydrogen peroxide (H2O2) via peroxidase-like reaction with 3,3′ ,5,5′ -tetramethylbenzidine (red-
TMB) that found the lowest limit of detection (LOD) of 2.25 µM.

1. Introduction them with bioactive agents [11,12]. Among these natural precursors,
plant extract as abundant sources is attracted particular attention due to
Metallic nanoparticles have gained great interest due to the nano­ enabling the scalable and cheap synthesis of AuNPs [13–15].
particles possessing special physicochemical properties with large Catalytic reduction of nitrophenols (NiPs) to relative aminophenols
number of active positions, which can significantly enhance their cata­ (APs) has been explored widely as a model reaction for evaluating cat­
lytic activity [1–3]. Gold nanoparticles (AuNPs) with a high specific alytic activity of metallic nanoparticles since NiPs exhibit high stability
surface area is less prone to self-poisoning which can make them very and toxicity in water, causing harm to living environment. Notably, the
active in the catalytic applications [4]. In the recent years, green product of NiPs reduction, APs, are well known to be important in­
approach utilising the biosources such as fungi [5], bacteria [6], plant termediates in the fabrication of several antipyretic and analgesic drugs
embryos [7], plants [8,9] and waste biomass [10] has frequently [16,17]. In the industrial applications, APs are also used as
investigated. The biosynthesised nanoparticles can improve biocom­ anticorrosion-lubricating agents in fuels, corrosion inhibitors in paints
patibility and stability as well as possess a low toxicity due to coating and dyeing agents for fur and feathers [18,19]. Thus, several research

* Corresponding author at: Institute of Chemical Technology, Vietnam Academy of Science and Technology, 1A, TL29, Thanh Loc Ward, District 12, Ho Chi Minh
City, Vietnam.
E-mail address: ntdanh@ict.vast.vn (T.-D. Nguyen).

https://doi.org/10.1016/j.jece.2021.106590
Received 1 September 2021; Received in revised form 30 September 2021; Accepted 13 October 2021
Available online 16 October 2021
2213-3437/© 2021 Elsevier Ltd. All rights reserved.
V.-D. Doan et al. Journal of Environmental Chemical Engineering 9 (2021) 106590

groups have explored the catalytic kinetic of NiPs reduction using obtained solution was filtered with Whatman filter paper No.1 to
NaBH4 in the presence of the metallic nanocatalyst [20–22]. In these remove the solid, and the extract was stored in a refrigerator at 4–10 ◦ C
studies, the kinetic analysis indicates that the reaction, which was for further studies.
almost performed around room temperature, following a
pseudo-first-order rate law via Langmuir-Hinshelwood model for het­ 2.2. Synthesis of biogenic AuNPs
erogeneous catalyst. Recently, Fedorczyk et al. [23] showed that cata­
lytic reduction of 4-NiP with excessive amount of borohydride was CS extract as a green reducing and stabilising agent for the synthesis
followed by either pseudo-zero-order or pseudo-first-order kinetics, of AuNPs was added into HAuCl4 solution under dark condition. The
depended on concentrations of 4-NiP. Meanwhile, Li and Chen [24] optimal parameters including concentration of gold ions (0.50–1.25
found a wide range of the reaction orders (1.5–6.7) for the reduction of mM), temperature (40–80 ◦ C) and time (15–75 min) were investigated
4-NiP in the presence of metallic nanoparticles coated on poly(allyl­ using UV-Vis measurements (UV-Vis Evolution 300 spectrophotometer).
amine hydrochloride)/poly(glycidyl methacrylate) composite. The re­ The change in physicochemical property and concentration of the
action orders depended on the coated metallic nanoparticles coverage. nanoparticles was determined by change in absorbance values and
Thus, it is clear that the kinetic model of NiPs reduction is strongly characteristic of surface plasmon resonance (SPR) absorption peaks. The
affected by the reaction conditions. CS-AuNPs biosynthesised under the optimal condition were used for the
H2O2 is used in various fields, including the manufacture of dis­ further studies. The solid CS-AuNPs were centrifuged for 30 min at 4000
infecting and cleaning agents, and antibacterial agents [25,26]. Due to rpm, washed and dried overnight at 60 ◦ C.
high stability in abiotic environments at ambient condition and neutral
pH, H2O2 is popularly environmental toxins which can kill rapidly ani­ 2.3. Characterisation of biogenic AuNPs
mal cells by producing highly-reactive hydroxyl radicals [27]. More­
over, many enzyme-catalysed reactions such as xanthine oxidase, Fourier transform infrared spectroscopy (FTIR) spectra of solid
5-lipoxygenase, and NADPH oxidases, generate H2O2 as the samples were recorded in range of 4000–500 cm− 1 on a Bruker Tensor
by-product. In inflammation behaviour, when oxidative degradation 27 (Germany) for identification of possible functional groups present in
and modification of base pairs of DNA occur in cells, H2O2 molecules can the CS extract and CS-AuNPs. Crystalline structure of CS-AuNPs was
accumulate locally, leading to inflammation [28]. Thus, accurate analysed for powder XRD on the Shimadzu 6100 X-ray diffractometer
detection of H2O2 level has been the subject of particular attention by (Japan) operating at the voltage of 40 kV, the current of 30 mA with
the researchers. Among numerous techniques, chromogenic red-TMB CuKα radiation at the wavelength of 1.5406 nm, scanning speed of
can be effectively used for detection of H2O2. In this system, red-TMB is 0.05◦ /s and step size of 0.02◦ in the range 2θ from 10◦ to 80◦ .
oxidised by H2O2 in the presence of peroxide-like nanocatalyst such as Morphological characterisation, size distribution and crystal property of
noble metallic nanoparticles (AgNPs, AuNPs, PdNPs, etc.) to afford a CS-AuNPs were evaluated by TEM and high resolution transmission
blue oxidised TMB form (ox-TMB) [29]. electron microscope (HR-TEM), STEM-EDX mapping and the selected
Cynomorium songaricum is a popular plant species living in the area electron diffraction pattern (SAED) (JEOL JEM-2100, Japan).
mountainous regions of Vietnam and also found in China, Northern Colloidal solution was used to determine zeta potential at the sliding
Africa and the Mediterranean region. C. songaricum possesses various plane and the size distribution of the particles using dynamic light
bioactivity such as antioxidant, immunity-improving and improving scattering (DLS) (nanoPartica Horiba SZ-100, Japan).
physiological status in men. Moreover, C. songaricum contains many
important ingredients such as flavonoids, phenolic, steroid, protein, 2.4. Reduction of nitrophenols
lignans and polysaccharides which can be responsible for reduction and
stabilisation of metallic nanoparticles [30–33]. For instance, the hy­ To evaluate the catalytic activity, the hydrogenation of 2-NiP and 3-
droxyl group of flavonoids in the plant plays a principal role as a NiP was carried out by using excess amount of NaBH4 in the presence of
reductant because the tautomeric transformation of enol form into keto the solid CS-AuNPs. Briefly, a mixture of NiP (2.5 mL, 0.1 mM) and
form may result reactive hydrogen atoms that can reduce metallic ions NaBH4 (0.5 mL, 0.1 M) was prepared in a cuvette with a length of 1.0
to metallic nanoparticles [34]. To the best of our knowledge, the extract cm. Then, the CS-AuNPs catalyst (3 mg) was added into the solution. The
of C. songaricum (CS) has not yet been used in the biosynthesis of catalytic performance via kinetic studies were investigated using UV-Vis
nanoparticles. spectroscopy at the wavelength of 410 nm for 2-NiP and 390 nm for 3-
In this study, we synthesised AuNPs using CS extract as an eco- NiP and thermodynamic parameters were calculated via reactions per­
friendly reducing and stabilising agent. The synthetic conditions of formed at wide range of temperatures (30–70 ◦ C).
AuNPs were optimised to determine the best effective reaction condi­
tion. Physicochemical properties were investigated by analytic tech­ 2.5. Peroxidase-like activity and detection of H2O2
niques. For their applications, AuNPs were evaluated for catalytic
kinetics of the nitrophenols reduction in a wide range of reaction tem­ Peroxidase-like activity of CS-AuNPs was based on the reaction be­
peratures and their peroxidase-like activity was used to detect H2O2 tween red-TMB and H2O2 that was carried out as follows: 200 μL of red-
levels via an H2O2-TMB sensing system. TMB + 300 μL of H2O2 + CS-AuNPs mixed with acetate buffer in a 2 mL-
Eppendorf and left for 30 min. Then, the mixture was transferred to a
2. Materials and methods standard cuvette with a 1 cm length for measurement of UV-Vis spectra.
The colourimetric detection conditions were explored by keeping
2.1. Materials and chemicals three factors and changing the other. The pH values, reaction temper­
ature, volume of CS-AuNPs solution and red-TMB concentrations were
All chemicals used were of analytical grade and utilised without investigated in the ranges of 1–8, 20–80 ◦ C, 10–80 μL and 0.5–3.5 mM,
further purification. Chemicals including hydrogen tetrachloroaurate respectively. The reaction performance was evaluated via the absor­
(III) hydrate (HAuCl4.3H2O), 2-nitrophenol (2-NiP), 3-nitrophenol (3- bance values at the peak 654 nm in UV-Vis spectra.
NiP), Glacial acetic acid (CH3COOH), sodium acetate (CH3COONa), For the detection of H2O2, the reaction was performed in 2 mL-
H2O2 solution 30% and red-TMB were purchased from Acros Co Eppendorf as follows: 200 μL of red-TMB (5.0 mM) + 60 μL of CS-AuNPs
(Belgium). The distilled water was used thoroughly in all experiments. + 1 mL of acetate buffer at pH 5.0. After that, 300 μL of different H2O2
For the preparation of CS extract, harvested C. songaricum was dried. concentrations (10–400 μM) were added and kept for 30 min at 50 ◦ C.
The CS powder (2.5 g) was boiled in distilled water (300 mL) for 1 h. The Then, UV–Vis spectra were recorded and the absorbance values at 654

2
V.-D. Doan et al. Journal of Environmental Chemical Engineering 9 (2021) 106590

Fig. 1. Schematic illustration for the biosynthesis of CS-AuNPs and their applications.

nm were evaluated to determine linear curve and LOD value calculated of around 540 nm in UV-Vis spectra confirmed good conversion of gold
via a formula 3σ/s where s denotes the slope of linearity and σ represents ions to CS-AuNPs. The physicochemical properties and structures of the
the standard deviation, respectively. nanoparticles were well characterised using analytic techniques. Their
applications involved catalytic activity for reduction of NiPs and
3. Results and discussion peroxidase-like activity for colourimetric detection of H2O2 levels.
In order to gain insight of AuNPs formation, optimal parameters of
3.1. Biosynthesis of CS-AuNPs the reduction including concentration of gold ion, temperature and time
were explored with the aid of UV-Vis measurement. Changes in the
The phytochemicals of C. songaricum showed the presence of absorbance and λmax values at the SPR band can provide important in­
numerous active organic compounds such as flavonoids, fatty acids, formation on size, morphology and yield for the formation of CS-AuNPs.
phenolic acids, polysaccharides [35,36] that can be responsible in the The absorption results showed that the formation of CS-AuNPs was
green synthesis of metallic nanoparticles without the toxic reagents strongly dependent on all of the investigated parameters as can be
used. The formation of the metallic nanoparticles in colloidal solution obviously seen in the different trends (Fig. 2).
can relate to the mechanism involved three different stages [37,38]: (1) Higher gold ion concentrations (0.5–1.25 mM) induce increased the
the conversion of metallic ions into metallic nanoparticles occurs via an formation of CS-AuNPs. An increase in absorbance values was observed
oxidoreduction process by the organic reductants present in the plant in the range 0.5–1.0 mM, indicating that the concentration of the
which can be affected by redox potential of organic compositions and nanoparticles in the colloidal solution increases with concentration of
metal. (2) The metallic nanoparticles are clustered via an aggregation gold ions and achieved a maximum value at 1.0 mM. In addition, λmax
process that can be strongly affected by the diffusion model. (3) The values were almost no change in increase of gold ion concentrations that
growth of the nanoparticles is carried out via the combination of the confirmed stable morphology of the colloidal solution in range of the
clusters. The growth rates can affect the geometry of metal nano­ investigated concentrations. Thus, the most effective CS-AuNPs forma­
particles. In general, the reducing and stabilising agents are principal tion was found at a gold ion concentration of 1.0 mM.
factors inducing the difference in their physicochemical properties and For study on the effect of temperature, the reduction was carried out
catalytic activities. In this work, we utilised the CS extract as a green at a range of 40–80 ◦ C. The formation of CS-AuNPs was observed at all
source to biosynthesise AuNPs for catalytic and sensing applications investigated temperatures that the SPR band of CS-AuNPs appeared at
(Fig. 1). The dried CS was refluxed with the distilled water, affording a 536–540 nm. The result shows that absorbance values increased rapidly
light yellow solution after the filtration process. Due to geometry and with increasing temperature and λmax values were stable at region of
size of AuNPs strongly affected their catalytic activity [39,40], the re­ high temperature (> 60 ◦ C), indicating that the yield and morphology of
action conditions including concentration of gold ion, temperature and CS-AuNPs formed is highly dependent on the reaction temperature. The
time were optimised for the production of CS-AuNPs. The changes in the most effective reduction of gold ions was determined to take place at
solution colour from yellow to pink and absorption intensity at SPR band 70 ◦ C.

3
V.-D. Doan et al. Journal of Environmental Chemical Engineering 9 (2021) 106590

The time-dependent reduction of gold ions was investigated between CS extract reduces gold ions more rapidly than that it does with the
15 and 110 min. Increases in both absorbance and λmax values confirmed aqueous extract of other plants [41–43]. For further studies on their
increasing concentration and larger size of nanoparticles in the colloidal physicochemical properties and applications, CS-AuNPs were carried
solutions with extended reaction times. The formation of CS-AuNPs out under optimal condition with gold ion concentration of 1.0 mM at
occurred rapidly after 30 min of the stirring, and a stable solution was 70 ◦ C for 60 min.
afforded at 60 min. It is clear that the presence of organic components in

Fig. 2. UV-Vis spectra (left) and colour change (upper) and plots of investigated parameters versus absorption intensity and λmax values (right): concentration of gold
ion (A and B), reaction temperature (C and D), and time (E and F) for the biosynthesis of CS-AuNPs. (For interpretation of the references to colour in this figure
legend, the reader is referred to the web version of this article.)

4
V.-D. Doan et al. Journal of Environmental Chemical Engineering 9 (2021) 106590

Fig. 3. (A) Zeta potential, (B) DLS distribution of colloidal solution of CS-AuNPs, (C) FTIR spectra of CS extract (a) and CS-AuNPs (b) and XRD patterns of CS-AuNPs.

Fig. 4. (A) TEM images and distribution of particles (inset); (B) HR-TEM and SEAD images (inset); (C) EDX patterns and elemental composition (inset) and (D and E)
STEM-mapping of CS-AuNPs.

3.2. Characterisations of CS-AuNPs showed a monodispersed distribution in range of 20–150 nm with an


average hydrodynamic diameter of 45 nm. The data reveals that the
To confirm successful biosynthesis of the colloidal solution, CS- bioactive compounds present in the CS extract well stabilised AuNPs
AuNPs solution was used to measure the zeta potential and DLS distri­ after the reduction.
bution at 25 ◦ C as shown in Fig. 3A and B. The results showed that the The functional groups of the bioactive compositions in the CS extract
colloidal solution of CS-AuNPs possesses a good stability with a greatly responsible as the stabilising agents of CS-AuNPs were identified by
negative zeta potential (− 45.9 mV). The size distribution of CS-AuNPs using FTIR measurements and the CS extract was also used as a reference

5
V.-D. Doan et al. Journal of Environmental Chemical Engineering 9 (2021) 106590

Fig. 5. (A) UV-Vis spectra and (B) stability over time of CS-AuNPs.

sample for analysis of CS-AuNPs (Fig. 2C). Similar absorption peaks are (nm) is the average crystal size, β (radian) is the full width at half
obviously observed from both spectra. The peaks of CS extract centred at maximum (FWHM), λ (0.1540 nm) is the wavelength of used CuKα X-ray
wavenumbers of 3350, 2924, 2850, 1725, 1587, 1405 and 1091 cm− 1, radiation and ‘θ’ (radian) is the Bragg diffraction angle. The crystal
while the CS-AuNPs peaks are slightly shifted to new positions at 3376, diameter was estimated to be 25.04, 19.05, 21.28 and 23.85 nm for the
2922, 2852, 1726, 1559, 1406 and 1043 cm− 1 due to the presence of fcc planes of (1 1 1), (2 0 0), (2 2 0), and (3 1 1), respectively and a mean
metal component. The hydroxyl (OH) and amino (N–H) groups repre­ value was calculated to be 22.30 nm. The mean lattice parameter and
sent at a broad band of 3350 cm− 1 and – – C–H groups are assigned to cell volume of AuNPs could be determined from interplanar spacing size
the bands of 2924 and 2850 cm− 1 that can be related to the composition that confirms the presence of AuNPs in the sample with 4.07552 Å and
of flavonoids, polyphenols, polysaccharides and proteins [33,44]. A 67.69482 Å3, respectively [46,47].
shoulder that centred at 1725 cm− 1 is assigned to the C– – O stretching The morphology and crystal structure of CS-AuNPs were evaluated
vibration of ketones or amides. The sharp peaks at 1587 cm− 1 and by TEM, HRTEM and SAED analysis (Fig. 4A and B). TEM images of CS-
1091 cm− 1 corresponded to stretching vibrations of C– – C and C–O, AuNPs indicated an uneven geometry with triangular, pentagonal,
respectively [45]. The observation indicates that the bioactive com­ hexagonal and spherical shapes and a narrow size distribution of CS-
pounds in CS extract relate to the stabilising agents of AuNPs. AuNPs in a range of 2–30 nm with an average size of 16 nm. The mul­
The crystalline structure of CS-AuNPs was evaluated using XRD tiple shapes of formed AuNPs can relate to various phytochemicals of
analysis (Fig. 3D). The XRD pattern of CS-AuNPs shows sharp diffraction stabilising agents in the CS extract that is similarly observed in the
peaks at 2θ angles of 38.18◦ , 44.36◦ , 64.59◦ and 77.55◦ correlated to the previous reports [48,49]. The HRTEM image and SAED pattern showed
face-centred cubic (fcc) structure of AuNPs (1 1 1), (2 0 0), (2 2 0) and (3 the crystalline structure of CS-AuNPs. The HRTEM image revealed a
1 1), respectively (ICDD PDF card number 00-004-0784). The XRD pa­ fringe lattice of the crystalline phase with a spacing of 0.22 nm that
rameters calculated are shown in Table S1. The crystal size of AuNPs is correlates to the (1 1 1) plane of AuNPs. The SAED pattern displayed
determined using the Debye-Scherrer equation D = 0.9λ/βcosθ, where D bright circular rings corresponding to the (1 1 1), (2 0 0), (2 2 0) and (3 1

Fig. 6. Schematic representation of heterogeneous CS-AuNPs catalyst for the reduction of NiPs via the Langmuir–Hinshelwood mechanism. The abbreviations, kNiP,
kBH4, kde and kapp are rate constants of NiP adsorption, BH4 adsorption, detachment of AP product and apparent kinetic respectively.

6
V.-D. Doan et al. Journal of Environmental Chemical Engineering 9 (2021) 106590

1) reflection planes of AuNPs. The results are in agreement with the


earlier data of the XRD pattern. The elemental composition of CS-AuNPs
was measured by EDX analysis (Fig. 4C). Strong signals at around 2.2,
9.8, 11.5 and 13.4 keV confirm the presence of elemental gold in the
sample [50] and the appearance of carbon (0.4 keV), oxygen (0.5 keV),
chloride (2.6 keV) can relate to the composition of CD extract. The mean
percentage of gold in CS-AuNPs was estimated to be 17.0% (w/w). The
scanning transmission electron microscopy (STEM) mapping images
(Fig. 4D and E) confirm the presence of well-dispersed AuNPs without
the agglomeration. The results show that the active composition in the
CS extract effectively reduced and stabilised AuNPs.

3.3. Stability

Although the result of zeta potential showed a highly stable colloidal


solution of CS-AuNPs (Fig. 3 A) the agglomeration of nanoparticles over
time can strongly affect their catalytic activity. Thus, their stability
should be investigated over the storage time. The samples were stored in
a vial with a cover at room temperature. UV-Vis spectroscopy of the
colloidal solutions were measured after various times of 3, 7, 10, 14, 17,
20, 24, 27 and 30 days. The stability was determined by decrease of the
absorption intensity at SPR peak (Fig. 5). The result shows that the
nanoparticles rapidly agglomerate in an initial week with decrease in
stability of 20%. However, a slow decrease is observed in longer time
and their stability is remained 66% after 30 days of the exposition. Thus,
the AuNPs biosynthesised from CS extract display higher stability at
room temperature in comparison with that synthesised from the other
extract [51].

3.4. Kinetic modelling

NiPs are known to be a toxic and persistent organic substance that is


emitted from dyes, pesticides and explosives [52]. Hence, the reduction Fig. 7. Changes in concentrations of 2-NiP (left) and 3-NiP (right) as a function
of NiPs into APs using NaBH4 is widely investigated. On earlier reports, of time via various kinetic models: zero-order (A and B), half-order (C and D),
the kinetic model of NiPs reduction in heterogeneous metallic nano­ first-order (E and F), one-half-order (G and H) and second-order (I and J).
particle catalysts under room temperature region was found to be
pseudo-first-order [39,53,54]. However, several recent reports showed by the NiP substrates [24] and the rate equation can be expressed as
the NiPs reduction followed zero-order or nth-order mechanism (n > 1) follows.
that depended on reaction conditions such as initial concentration of
dCNiP
NiPs [23] or coated metallic nanoparticles coverage [24]. In the present rt = − = KSθNiP = kapp CnNiP (1)
dt
work, the conversion of 2-NiP and 3-NiP into corresponding APs was
performed using BH4‾ as a reducing agent in the presence of CS-AuNPs (kNiP CNiP )x
catalyst. The reaction was conducted under a wide range of temperature θNiP = (2)
1 + (kNiP CNiP )x
from 30 to 70 ◦ C. Therefore, the reaction might undergo a reaction order
differing from pseudo-first-order. Where θNiP is the surface coverage value of NiPs, kNiP is the adsorption
With an excess of NaBH4, the catalytic reduction occurred at the constant of NiPs, x refers to the exponent related to the heterogeneity of
surface of AuNPs following Langmuir–Hinshelwood model of hetero­ the sorbent, n is the apparent order of reaction with respect to NiPs, and
geneous catalyst without leaching atom in the solution [55] and both kapp is the apparent kinetic constant which can be identified as follows:
reactants, nitrophenolate and BH4‾ ion, absorbed onto the surface of
KSkxNiP Cx−NiPn
nanoparticles before the NiPs reduction. Several mechanisms have been kapp = (3)
1 + (kNiP CNiP )x
reported for the reduction of NiPs using metal nanocatalyst that includes
surface-mediated hydrogen transfer for catalysts e.g. PdNPs, and bi­
metals [55–57] or surface-mediated electron transfer for catalysts of 3.4.1. Pseudo-zero-order kinetic
AuNPs and AgNPs [24,53,58] or combination of both these mechanisms The reaction rate based on zero-order kinetic can be expressed as
[59]. In this work, the mechanism of NiPs reduction in the presence of follows:
heterogeneous CS-AuNPs catalyst refers to be described via the electron dCNiP
transfer as illustrated in Fig. 6. The reaction mechanism can be assumed rt = − = k0 (4)
dt
via two main stages. In the first stage, the adsorption of BH4‾ ion and
electron transfer into the surface of catalyst. Due to the excess of NaBH4 Where k0 is rate constant of pseudo-zero-order reaction and has units of
used, the BH4‾ ion adsorption is assumed to be an irreversible process. mol s− 1. Integrating Eq. (4), we obtain Eq. (5).
The second stage includes the equilibrated adsorption of nitrophenolate, CNiP0 − CNiP = k0 t (5)
transferring electron from the surface of the nanoparticles into nitro­
phenolate and formation of the product via the detachment of amino­ Thus, a plot of CNiP0 − CNiP as a function of time may give a straight
phenolate. The adsorption of NiPs on the surface of the AuNPs is line, and the slope value can be used to estimate the pseudo-zero-order
reversible and modelled by the Langmuir isotherm. Thus, the reaction rate constant k0.
rate (rt) is mainly dependent on the fraction of catalyst surfaces covered

7
V.-D. Doan et al. Journal of Environmental Chemical Engineering 9 (2021) 106590

Table 1
Rate constants derived from the kinetic experiments conducted at different reaction temperatures.
Comp. Kinetic model 303 K 313 K 323 K 333 K 343 K
− 3 2 − 3 2 − 3 2 − 3 2 3
kx10 R kx10 R kx10 R kx10 R kx10− R2

2-NP zero 2.86 0.94 3.00 0.93 3.18 0.96 3.29 0.96 3.37 0.92
half 3.32 0.99 3.59 0.98 3.84 0.97 3.96 0.99 4.20 0.98
first 4.27 0.98 4.89 0.97 5.41 0.95 5.72 0.92 6.45 0.95
one-half 6.15 0.91 7.73 0.88 9.24 0.82 10.39 0.76 12.93 0.80
second 9.92 0.80 14.2 0.74 19.12 0.67 23.6 0.61 33.56 0.63
3-NP zero 1.82 0.97 2.05 0.93 2.30 0.96 2.35 0.94 2.39 0.96
half 2.09 0.99 2.42 0.99 2.72 0.99 2.97 0.99 3.09 0.99
first 2.48 0.98 3.02 0.99 3.46 0.98 4.20 0.94 4.53 0.94
one-half 3.06 0.96 4.01 0.95 4.76 0.91 6.79 0.81 7.69 0.81
second 3.92 0.91 5.65 0.87 7.09 0.81 12.52 0.67 15.09 0.67

3.4.2. Pseudo-first-order kinetic times are shown in Fig. 7 and kapp and correlation coefficients (R2) are
The reaction rate based on first-order kinetic with respect to NiPs can summarised in Table 1. The result reveals that the reaction orders more
be expressed as follows: than first (n = 1.5 and 2) possess low correlation coefficients in all
temperatures. Thus, these kinetic models are unsuitable to describe the
dCNiP
rt = − = k1 CNiP (6) reduction reaction of NiPs. It is noteworthy that linear plots of CNiP vs
dt
time with high correlation coefficients at low temperatures (R2
Where k1 is rate constant of pseudo-first-order reaction and has units of = 0.98–0.99) confirm a pseudo-first-order but these values are signifi­
s− 1. Integrating Eq. (6), we obtain Eq. (7). cantly decreased at high temperatures (R2 = 0.92–0.95). It indicates that
model of pseudo-first-order reaction is appropriate to describe the ki­
CNiP0 netic of NiPs reduction at low reaction temperature region as well
ln( ) = k1 t (7)
CNiP known in the literature [53,60,61]. However, this kinetic model is
Thus, a plot of ln(CNiP0 /CNiP ) as a function of time may give a straight inaccurate to be employed under high reaction temperatures that can
line, and the slope value can be used to estimate the pseudo-first-order induce high disorder of NiPs molecules, led to adsorption of nitro­
rate constant k1. phenolate ions on the surface of the catalyst increasing independence of
NiPs concentration. The kinetic model of zero-order reaction shows a
3.4.3. Pseudo-nth-order kinetic low correlation between CNiP vs time whereas the half-order kinetic
The reaction rate based on nth-order kinetic with respect to NiP can indicates a good linearity of CNiP vs time plots for all tested reaction
be expressed as follows: temperatures, confirming that the behaviour of CS-AuNPs catalyst for
reductions of NiPs follows half-order reaction.
dCNiP
rt = − = kn CnNiP (8) Moreover, increasing slopes along with increase of reaction tem­
dt perature can be related to collision theory which described the move­
Where kn is rate constant of pseudo-nth-order reaction and has units of ment of NiPs molecule being more vigorous in high temperature, leading
mol1− n s− 1. Integrating Eq. (8), we obtain Eq. (9). to a better chance for the collision and more rapid reaction rate.
Therefore, this data can be used to calculate the thermodynamic pa­
1
(C(1− n) (1− n)
(9) rameters. Apparent rate constant of 2-NiP and 3-NiP were found in range
NiP0 − CNiP ) = kn t
1− n of 3.32 × 10− 3 to 4.20 × 10− 3 mol1/2 s− 1 and 2.09 × 10− 3 to
Thus, a plot of (1/(1 − n))(C(1− n) (1− n) 3.09 × 10− 3 mol1/2 s− 1, respectively. The significant difference in
NiP0 − CNiP ) with n ∕
= 1, as a function of
time gives a straight line, and the slope value can be used to estimate the reactivity of NiPs can be related to the adsorption ability of their
pseudo-nth-order rate constant kn. different molecular structures.
The apparent reaction order and rate constants can be determined
using a time-dependent manner of NiPs concentration which is moni­
3.5. Thermodynamic parameters
tored by UV-Vis measurement. By fitting the time dependence of the
corresponding NiPs concentration curves using Eqs. (5), (7) and (9), the
The values kapp of half-order reaction at different temperatures was
apparent rate constant could be obtained for corresponding reaction
used to calculate thermodynamic parameters of the hydrogenation re­
orders and an appropriate kinetic modelling could be derived from
action including activation energy (Ea), enthalpy (ΔH‡), entropy (ΔS‡),
comparing among correlation coefficients. The reactions occurred under
and Gibbs energy (ΔG‡).
reaction temperature in range of 303–343 K were confirmed by discol­
The values Ea are calculated via the Arrhenius equation [62,63].
oration of NiPs and decrease in absorbance values at 410 and 390 nm in
the UV–Vis spectra of 2-NiP and 3-NiP, respectively (Table S1). The kapp = Ae(−
Ea
RT 2
)
(10)
results showed the reactions completed in 10 min for all tests.
The value kapp of both 2-NiP and 3-NiP reductions with reaction or­ Or
ders of zero, half, first, one-half and second was calculated for different Ea
reaction temperatures. The plots of the concentrations versus reaction lnkapp = lnA − (11)
RT

Where A is the pre-exponential factor and R (8.314 J K− 1 mol− 1) is the


Table 2
Thermodynamic parameters derived from Arrhenius and Eyring equations for identical gas constant. The value exp(− Ea/RT2) represents the fraction
kinetic model of half-order reaction. of collisions which have enough energy to overcome the activation
barrier at temperature T(K).
Comp. Ea (kJ mol− 1) ΔH‡ (kJ mol− 1) ΔS‡ (J mol− 1
K− 1) ΔG‡ (kJ mol− 1)
In this work, the reduction of NiPs is proposed as a unimolecular
2-NP 4.88 2.20 -176.83 54.90 reaction. Thus, the other thermodynamic parameters (ΔH‡, ΔS‡ and
3-NP 8.60 5.93 -168.26 56.07
ΔG‡) can be estimated using the Eyring and Gibbs equations.

8
V.-D. Doan et al. Journal of Environmental Chemical Engineering 9 (2021) 106590

the adsorption ability of different molecular structures of NiPs onto the


surface of nanocatalyst [53,64,65]. It also confirms the reduction of NiPs
following the Langmuir–Hinshelwood mechanism. The values Ea are
more comparable than those of AuNPs catalysts reported in the litera­
ture [53,66].
The plot of value ln(kapp/T) vs 1/T showed positive values of ΔH‡ for
the reduction of both NiPs, indicating that the NiPs reduction is endo­
thermic because reaction products, APs, possess higher enthalpy than
that of the reactant, NiPs [59]. The value ΔH‡ for reduction of 2-NiP
(2.2 kJ mol− 1) is much lower than the value of 3-NiP reduction
(5.93 kJ mol− 1) indicated a good correlation with relative activation
energy. The values ΔS‡ provide an insight of the molecularity of the
rate-determining step. The negative value of ∆S‡ indicates that the
transition state possessed a more ordered structure than that of the re­
actants in the ground state which can be attributed to an associative
mechanism of the adsorption process of NiPs and BH4‾ onto the surface
Fig. 8. UV-Vis spectra and corresponding colour (upper photographs) of of AuNPs to form the activated state [67]. The reduction of 2-NiP with a
AuNPs + H2O2; AuNPs + red-TMB; red-TMB + H2O2 and AuNPs + red- more negative ΔS‡ (more ordered structure) confirms better adsorption
TMB + H2O2 incubated for 30 min at 40 ◦ C. (For interpretation of the refer­ ability of 2-NiP in good agreement with value Ea. The Gibbs free energy
ences to colour in this figure legend, the reader is referred to the web version of of activation (ΔG‡) identified the barrier to move from the molecular
this article.) state to the activated state, was calculated at 298 K [68]. The
free-energy was found to be 54.90 kJ mol− 1 and 56.07 kJ mol− 1 for the
Eyring equation: reduction of 2-NiP and 3-NiP, respectively, indicating that reduction of
2-NiP into 2-AP required a slightly lower energy than the conversion of
kapp ∆H ‡ ∆S‡ kB 3-NiP into 3-AP. The result reveals that the reactivity of NiPs is signifi­
ln = − + + ln (13)
T RT R h cantly dependent on the substitution position of hydroxyl group as
Gibbs equation: confirmed from the kinetic data.

∆G = ‡
∆H‡ − T∆S‡ (14)
3.6. Peroxidase-like activity of CS-AuNPs nanozyme and detection of
Where kB is the Boltzmann constant (1.38 × 10− 23 J K− 1) and h is hydrogen peroxide
Planck’s constant (h = 6.63 × 10− 34 J s− 1).
The plot of value ln(kapp) vs 1/T can be used to estimate the activa­ To evaluate the peroxidase-like activity of the CS-AuNPs nanozyme
tion energy (Ea) and the thermodynamic parameters (ΔH‡ and ΔS‡) are in the oxidation of red-TMB using H2O2 as an oxidant, the solutions
defined by the slope of linear plot of ln (kapp/T) vs 1/T as described in containing different components were measured by UV-Vis spectroscopy
Fig. S3. The value ΔG‡ is derived from the Gibbs Eq. (14). The results are as shown in Fig. 8. The solutions containing H2O2 + AuNPs and red-
summarised in Table 2. The values Ea were found to be 4.88 kJ mol− 1 TMB + AuNPs were colourless without the peak at 654 nm corre­
and 8.60 kJ mol− 1 for the reduction of 2-NiP and 3-NiP, respectively, sponding to ox-TMB absorption while the solution containing red-
indicating the barrier energy of 2-NiP reduction is lower than that of 3- TMB + H2O2 was slightly blue with a tiny absorption band at 654 nm,
NiP. The significant difference in values Ea between NiPs can relate to reflecting that H2O2 slightly oxidises red-TMB into ox-TMB without the

Fig. 9. UV-Vis spectra for peroxidase-like activity of CS-AuNPs and plots of absorbance in changes of (A and B) pH range from 1–8; (C and D) incubation temperature
from 20 ◦ C to 80 ◦ C; (E and F) nanozyme volume from 10 to 70 μL; and (G and H) concentration of red-TMB from 0.5 to 3.5 mM.

9
V.-D. Doan et al. Journal of Environmental Chemical Engineering 9 (2021) 106590

Fig. 10. (A) UV-Vis spectra of ox-TMB with H2O2 concentration of 10–400 μM with the colour change of the corresponding solutions (upper photographs); and (B)
the linear calibration plot of the absorbance at 654 nm and H2O2 concentrations based on the peroxidase-like oxidation of red-TMB. (For interpretation of the
references to colour in this figure legend, the reader is referred to the web version of this article.)

catalyst. In addition, a strong absorption peak centred at 654 nm was community. Thus, the oxidation of red-TMB in the presence of the CS-
observed in the sample containing red-TMB + H2O2 + AuNPs, which AuNPs catalyst dependent on concentration of H2O2 can be used to
confirmed the catalytic activity of AuNPs. It is noteworthy that signifi­ sense H2O2 levels. In this work, various concentrations of H2O2
cant decrease in absorbance at peak around 550 nm in the absorption (10–400 μM) were added to the mixture of CS-AuNPs + acetate buffer
spectrum of H2O2 + AuNPs in comparison with red-TMB + AuNPs in­ (pH 5) and incubated at 50 ◦ C for 30 min and absorption spectra were
dicates oxidation of Au(0) into Au (I) by H2O2, which play a role as the then measured. The result shows that the blue colour of ox-TMB solu­
active site on AuNPs in the oxidation process of red-TMB into ox-TMB tions was changed from light to dark, along with increase in absorbance
[69,70]. The result confirms CS-AuNPs behaving as a nanozyme for values at 654 nm in the range of tested concentrations ( Fig. 10). The
peroxidase-like activity with red-TMB. plot of ox-TMB absorbance vs H2O2 concentrations from 10 to 400 μM
The catalytic performance of the CS-AuNPs for peroxidase-like ac­ shows a good linearity with a regression equation of Abs. = 0.0023
tivity can be affected by changes in reaction conditions. In this work, the CH2O2 + 0.897 and a correlation coefficient of R2 = 0.999. The limit of
effect of reaction factors including the pH, incubation temperature, detection (LOD) value calculated from straight line of linear calibration
concentration of CS-AuNPs, and concentration of red-TMB was explored was found to be 2.25 μM, which is better than the previously reported
via UV-Vis measurement. For effect of pH values, the reaction was values [72–74].
performed in a range of 1–8. The absorption spectra showed that the
peaks at 654 nm only appeared in pH range of 3–7. The absorbance 4. Conclusions
values were increased with increase of medium pH and a maximum
value achieved at pH 5 ( Fig. 9A and B). Greater pH values induced a This work successfully applied a low-cost and effective method for
significant decrease in concentration of ox-TMB, which might relate to green synthesis of gold nanoparticles from aqueous extract of Cyn­
the stability of ox-TMB in these media. The similar trend was observed omorium songaricum. Gold nanoparticles were formed in multiple shapes
for the optimisation of the incubation temperature that was carried out with a mean size of 16 nm which displayed effective catalytic ability in
in a range of 20–80 ◦ C at pH 5. The absorption intensity at 654 nm reduction of nitrophenols. The previous reports showed that the
increased with high temperature and found a maximum absorbance reduction of nitrophenols using NaBH4 in region of room reaction
value at 50 ◦ C (Fig. 9C and D). Higher temperatures significantly temperature via Langmuir–Hinshelwood mechanism was followed by
decreased in absorption intensity of ox-TMB, indicating low thermal the pseudo-first-order. However, the finding showed that a suitable
stability of ox-TMB in H2O2 in acidic medium [71]. For the quantity of model for the nitrophenol reduction was pseudo-half-order even at high
catalyst, different volumes of CS-AuNPs nanozyme (10–70 μL) was temperatures. The activation energies in the reaction of 2-nitrophenol
added to the solution. The best efficiency of ox-TMB formation was and 3-nitrophenol in the presence of gold nanoparticles were found to
found at 60 μL of CS-AuNPs solution, which achieved the highest be 4.88 kJ mol− 1 and 8.66 kJ mol− 1, respectively. The thermodynamic
absorbance value at peak 654 nm (Fig. 9E and F). The optimisation of parameters including ΔH‡, ΔS‡ and ΔG‡ were determined from Erying
the red-TMB concentrations was investigated in a range of 0.5–3.5 mM and Gibbs equations. Gold nanoparticles were successfully investigated
(Fig. 9G and H). The absorbance values increased with increase of for detection of the hydrogen peroxide with LOD of 2.25 µM.
red-TMB concentrations and the best catalytic activity was found at
2.5 mM. Therefore, the most effective oxidation of red-TMB in the CRediT authorship contribution Statement
presence of CS-AuNPs was at pH 5, temperature of 50 ◦ C, 60 μL of
CS-AuNPs, and red-TMB concentration of 2.5 mM. Van-Dat Doan: Conceptualization, Investigation, Writing – original
Because H2O2 is extensively employed in numerous fields and had draft. Quoc-Huy Pham: Investigation, Visualization. Bao-An Huynh:
environmental impact and high toxicity, a low-cost and effective sensor Investigation, Visualization, Writing – original draft. Thi-Lan-Huong
for the detection of H2O2 is particularly considered by research Nguyen: Visualization. Anh-Tien Nguyen: Visualization. Thanh-Danh

10
V.-D. Doan et al. Journal of Environmental Chemical Engineering 9 (2021) 106590

Nguyen: Methodology, Supervision, Writing – review & editing. [20] F.D. Souza, H. Fiedler, F. Nome, Zwitterionic surfactant stabilized palladium
nanoparticles as catalysts in aromatic nitro compound reductions, J. Braz. Chem.
Soc. 27 (2016) 372–381.
Declaration of Competing Interest [21] S. Gu, S. Wunder, Y. Lu, M. Ballauff, R. Fenger, K. Rademann, B. Jaquet,
A. Zaccone, Kinetic analysis of the catalytic reduction of 4-nitrophenol by metallic
The authors declare that they have no known competing financial nanoparticles, J. Phys. Chem. C 118 (32) (2014) 18618–18625.
[22] S. Wunder, Y. Lu, M. Albrecht, M. Ballauff, Catalytic activity of faceted gold
interests or personal relationships that could have appeared to influence nanoparticles studied by a model reaction: evidence for substrate-induced surface
the work reported in this paper. restructuring, ACS Catal. 1 (8) (2011) 908–916.
[23] A. Fedorczyk, J. Ratajczak, O. Kuzmych, M. Skompska, Kinetic studies of catalytic
reduction of 4-nitrophenol with NaBH 4 by means of Au nanoparticles dispersed in
Data availability a conducting polymer matrix, J. Solid State Electrochem. 19 (9) (2015)
2849–2858.
The data used to support the findings of this study are included [24] M. Li, G. Chen, Revisiting catalytic model reaction p-nitrophenol/NaBH 4 using
metallic nanoparticles coated on polymeric spheres, Nanoscale 5 (23) (2013)
within the article. 11919–11927.
[25] D.J. Barrington, A. Ghadouani, Application of hydrogen peroxide for the removal
Appendix A. Supporting information of toxic cyanobacteria and other phytoplankton from wastewater, Environ. Sci.
Technol. 42 (23) (2008) 8916–8921.
[26] R. Guan, X. Yuan, Z. Wu, L. Jiang, Y. Li, G. Zeng, Principle and application of
Supplementary data associated with this article can be found in the hydrogen peroxide based advanced oxidation processes in activated sludge
online version at doi:10.1016/j.jece.2021.106590. treatment: a review, Chem. Eng. J. 339 (2018) 519–530.
[27] T. Mahaseth, A. Kuzminov, Potentiation of hydrogen peroxide toxicity: from
catalase inhibition to stable DNA-iron complexes, Mutat. Res./Rev. Mutat. Res. 773
References (2017) 274–281.
[28] C. Wittmann, P. Chockley, S.K. Singh, L. Pase, G.J. Lieschke, C. Grabher, Hydrogen
[1] M.J. Ndolomingo, N. Bingwa, R. Meijboom, Review of supported metal peroxide in inflammation: messenger, guide, and assassin, Adv. Hematol. 2012
nanoparticles: synthesis methodologies, advantages and application as catalysts, (2012).
J. Mater. Sci. 55 (15) (2020) 6195–6241. [29] Y. Shang, F. Liu, Y. Wang, N. Li, B. Ding, Enzyme mimic nanomaterials and their
[2] N. Narayan, A. Meiyazhagan, R. Vajtai, Metal nanoparticles as green catalysts, biomedical applications, ChemBioChem 21 (17) (2020) 2408–2418.
Materials 12 (21) (2019) 3602. [30] G. Shi, W. Jiang, L. Cai, G. Sui, Molecular characteristics and antitumor capacity of
[3] T.S. Rodrigues, A.G. da Silva, P.H. Camargo, Nanocatalysis by noble metal glycan extracted from Cynomorium songaricum, Int. J. Biol. Macromol. 48 (5)
nanoparticles: controlled synthesis for the optimization and understanding of (2011) 788–792.
activities, J. Mater. Chem. A 7 (11) (2019) 5857–5874. [31] H.-C. Meng, C.-M. Ma, Flavan-3-ol-cysteine and acetylcysteine conjugates from
[4] M.-C. Daniel, D. Astruc, Gold nanoparticles: assembly, supramolecular chemistry, edible reagents and the stems of Cynomorium songaricum as potent antioxidants,
quantum-size-related properties, and applications toward biology, catalysis, and Food Chem. 141 (3) (2013) 2691–2696.
nanotechnology, Chem. Rev. 104 (1) (2004) 293–346. [32] R. Tao, L. Miao, X. Yu, J.O. Orgah, O. Barnabas, Y. Chang, E. Liu, G. Fan, X. Gao,
[5] M. Kitching, M. Ramani, E. Marsili, Fungal biosynthesis of gold nanoparticles: Cynomorium songaricum Rupr demonstrates phytoestrogenic or phytoandrogenic
mechanism and scale up, Microb. Biotechnol. 8 (6) (2015) 904–917. like activities that attenuates benign prostatic hyperplasia via regulating steroid 5-
[6] S. Menon, S. Rajeshkumar, V. Kumar, A review on biogenic synthesis of gold α-reductase, J. Ethnopharmacol. 235 (2019) 65–74.
nanoparticles, characterization, and its applications, Resour.-Effic. Technol. 3 (4) [33] Y.-B. Zhou, R.-R. Ye, X.-F. Lu, P.-C. Lin, S.-B. Yang, P.-P. Yue, C.-X. Zhang, M. Peng,
(2017) 516–527. GC–MS analysis of liposoluble constituents from the stems of Cynomorium
[7] M.-T. Tran, L.-P. Nguyen, D.-T. Nguyen, T. Le Cam-Huong, C.-H. Dang, T.T.K. Chi, songaricum, J. Pharm. Biomed. Anal. 49 (4) (2009) 1097–1100.
T.-D. Nguyen, A novel approach using plant embryos for green synthesis of silver [34] S. Alzharani, S. Astudillo-Calderón, B. Pintos, E. Pérez-Urria, J.A. Manzanera,
nanoparticles as antibacterial and catalytic agent, Res. Chem. Intermed. (2021) L. Martín, A. Gomez-Garay, Role of synthetic plant extracts on the production of
1–21. silver-derived nanoparticles, Plants 10 (8) (2021) 1671.
[8] P.K. Dikshit, J. Kumar, A.K. Das, S. Sadhu, S. Sharma, S. Singh, P.K. Gupta, B. [35] X. Li, M. Sdiri, J. Peng, Y. Xie, B.B. Yang, Identification and characterization of
S. Kim, Green synthesis of metallic nanoparticles: applications and limitations, chemical components in the bioactive fractions of Cynomorium coccineum that
Catalysts 11 (8) (2021) 902. possess anticancer activity, Int. J. Biol. Sci. 16 (1) (2020) 61.
[9] T.J. Jayeoye, F.N. Eze, O.O. Olatunde, S. Benjakul, T. Rujiralai, Synthesis of silver [36] J.-L. Cui, Y. Gong, X.-Z. Xue, Y.-Y. Zhang, M.-L. Wang, J.-H. Wang,
and silver@ zero valent iron nanoparticles using Chromolaena odorata phenolic A phytochemical and pharmacological review on Cynomorium songaricum as
extract for antibacterial activity and hydrogen peroxide detection, J. Environ. functional and medicinal food, Nat. Prod. Commun. 13 (4) (2018),
Chem. Eng. 9 (3) (2021), 105224. 1934578×1801300428.
[10] C. Mondelli, G. Gözaydın, N. Yan, J. Pérez-Ramírez, Biomass valorisation over [37] M. Iatalese, M. Coluccio, V. Onesto, F. Amato, E. Di Fabrizio, F. Gentile, Relating
metal-based solid catalysts from nanoparticles to single atoms, Chem. Soc. Rev. 49 the rate of growth of metal nanoparticles to cluster size distribution in electroless
(12) (2020) 3764–3782. deposition, Nanoscale Adv. 1 (1) (2019) 228–240.
[11] S. Gurunathan, J. Han, J.H. Park, J.-H. Kim, A green chemistry approach for [38] B.R. Karimadom, H. Kornweitz, Mechanism of producing metallic nanoparticles,
synthesizing biocompatible gold nanoparticles, Nanoscale Res. Lett. 9 (1) (2014) with an emphasis on silver and gold nanoparticles, using bottom-up methods,
1–11. Molecules 26 (10) (2021) 2968.
[12] R.A. Zayadi, F.A. Bakar, Comparative study on stability, antioxidant and catalytic [39] P. Suchomel, L. Kvitek, R. Prucek, A. Panacek, A. Halder, S. Vajda, R. Zboril,
activities of bio-stabilized colloidal gold nanoparticles using microalgae and Simple size-controlled synthesis of Au nanoparticles and their size-dependent
cyanobacteria, J. Environ. Chem. Eng. 8 (4) (2020), 103843. catalytic activity, Sci. Rep. 8 (1) (2018) 1–11.
[13] P. Elia, R. Zach, S. Hazan, S. Kolusheva, Ze Porat, Y. Zeiri, Green synthesis of gold [40] T. Ishida, T. Murayama, A. Taketoshi, M. Haruta, Importance of size and contact
nanoparticles using plant extracts as reducing agents, Int. J. Nanomed. 9 (2014) structure of gold nanoparticles for the genesis of unique catalytic processes, Chem.
4007. Rev. 120 (2) (2019) 464–525.
[14] M. Noruzi, Biosynthesis of gold nanoparticles using plant extracts, Bioprocess [41] S.M. Pourmortazavi, M. Taghdiri, V. Makari, M. Rahimi-Nasrabadi, H. Batooli,
Biosyst. Eng. 38 (1) (2015) 1–14. Reducing power of Eucalyptus oleosa leaf extracts and green synthesis of gold
[15] S.A. Akintelu, S.C. Olugbeko, A.S. Folorunso, A review on synthesis, optimization, nanoparticles using the extract, Int. J. Food Prop. 20 (5) (2017) 1097–1103.
characterization and antibacterial application of gold nanoparticles synthesized [42] C. Botteon, L. Silva, G. Ccana-Ccapatinta, T. Silva, S. Ambrosio, R. Veneziani,
from plants, Int. Nano Lett. (2020) 1–12. J. Bastos, P. Marcato, Biosynthesis and characterization of gold nanoparticles using
[16] Y. Fu, P. Xu, D. Huang, G. Zeng, C. Lai, L. Qin, B. Li, J. He, H. Yi, M. Cheng, Au Brazilian red propolis and evaluation of its antimicrobial and anticancer activities,
nanoparticles decorated on activated coke via a facile preparation for efficient Sci. Rep. 11 (1) (2021) 1–16.
catalytic reduction of nitrophenols and azo dyes, Appl. Surf. Sci. 473 (2019) [43] T.-T. Vo, T.T.-N. Nguyen, T.T.-T. Huynh, T.T.-T. Vo, T.T.-N. Nguyen, D.-T. Nguyen,
578–588. V.-S. Dang, C.-H. Dang, T.-D. Nguyen, Biosynthesis of silver and gold nanoparticles
[17] J.P. Peesa, P.R. Yalavarthi, A. Rasheed, V.B.R. Mandava, A perspective review on using aqueous extract from Crinum latifolium leaf and their applications forward
role of novel NSAID prodrugs in the management of acute inflammation, J. Acute antibacterial effect and wastewater treatment, J. Nanomater. 2019 (2019).
Dis. 5 (5) (2016) 364–381. [44] M. Hao-Cong, W. Shuo, L. Ying, Y.-Y. Kuang, M. Chao-Mei, Chemical constituents
[18] M.J. Vaidya, S.M. Kulkarni, R.V. Chaudhari, Synthesis of p-aminophenol by and pharmacologic actions of Cynomorium plants, Chin. J. Nat. Med. 11 (4) (2013)
catalytic hydrogenation of p-nitrophenol, Org. Process Res. Dev. 7 (2) (2003) 321–329.
202–208. [45] L.R. Camelo Caballero, A. Wilches-Torres, A. Cárdenas-Chaparro, J.A. Gómez
[19] S. Khan, M. Naushad, A. Al-Gheethi, J. Iqbal, Engineered nanoparticles for removal Castaño, M.C. Otálora, Preparation and physicochemical characterization of
of pollutants from wastewater: Current status and future prospects of softgels cross-linked with cactus mucilage extracted from cladodes of Opuntia
nanotechnology for remediation strategies, J. Environ. Chem. Eng. (2021), ficus-indica, Molecules 24 (14) (2019) 2531.
106160. [46] M. Cudalbeanu, D. Peitinho, F. Silva, R. Marques, T. Pinheiro, A.C. Ferreira,
F. Marques, A. Paulo, C.F. Soeiro, S.A. Sousa, Sono-biosynthesis and

11
V.-D. Doan et al. Journal of Environmental Chemical Engineering 9 (2021) 106590

characterization of AuNPs from Danube Delta Nymphaea alba root extracts and [60] C. Lin, K. Tao, D. Hua, Z. Ma, S. Zhou, Size effect of gold nanoparticles in catalytic
their biological properties, Nanomaterials 11 (6) (2021) 1562. reduction of p-nitrophenol with NaBH4, Molecules 18 (10) (2013) 12609–12620.
[47] V. Grasmik, C. Rurainsky, K. Loza, M.V. Evers, O. Prymak, M. Heggen, K. Tschulik, [61] H. Rodríguez Molina, J.L. Santos Muñoz, M.I. Domínguez Leal, T.R. Reina,
M. Epple, Deciphering the surface composition and the internal structure of alloyed S. Ivanova, M.Á. Centeno Gallego, J.A. Odriozola, Carbon supported gold
silver–gold nanoparticles, Chem. Eur. J. 24 (36) (2018) 9051–9060. nanoparticles for the catalytic reduction of 4-nitrophenol, Front. Chem. 7 (2019)
[48] T.-T. Vo, C.-H. Dang, V.-D. Doan, V.-S. Dang, T.-D. Nguyen, Biogenic synthesis of 548.
silver and gold nanoparticles from Lactuca indica leaf extract and their application [62] E. Preklet, L. Tolvaj, L. Bejo, D. Varga, Temperature dependence of wood
in catalytic degradation of toxic compounds, J. Inorg. Organomet. Polym. Mater. photodegradation. Part 2: evaluation by Arrhenius law, J. Photochem. Photobiol.
30 (2) (2020) 388–399. A: Chem. 356 (2018) 329–333.
[49] T.T.-N. Nguyen, T.-T. Vo, B.N.-H. Nguyen, D.-T. Nguyen, V.-S. Dang, C.-H. Dang, [63] D. Michel, Simply conceiving the Arrhenius law and absolute kinetic constants
T.-D. Nguyen, Silver and gold nanoparticles biosynthesized by aqueous extract of using the geometric distribution, Phys. A: Stat. Mech. Appl. 392 (19) (2013)
burdock root, Arctium lappa as antimicrobial agent and catalyst for degradation of 4258–4264.
pollutants, Environ. Sci. Pollut. Res. 25 (34) (2018) 34247–34261. [64] Y.M. Magdy, H. Altaher, E. ElQada, Removal of three nitrophenols from aqueous
[50] L. Qian, W. Su, Y. Wang, M. Dang, W. Zhang, C. Wang, Synthesis and solutions by adsorption onto char ash: equilibrium and kinetic modeling, Appl.
characterization of gold nanoparticles from aqueous leaf extract of Alternanthera Water Sci. 8 (1) (2018) 1–15.
sessilis and its anticancer activity on cervical cancer cells (HeLa), Artif. Cells, [65] T. Sismanoglu, S. Pura, Adsorption of aqueous nitrophenols on clinoptilolite,
Nanomed., Biotechnol. 47 (1) (2019) 1173–1180. Colloids Surf. A: Physicochem. Eng. Asp. 180 (1–2) (2001) 1–6.
[51] V.-D. Doan, M.-T. Phung, T.L.-H. Nguyen, T.-C. Mai, T.-D. Nguyen, Noble metallic [66] T.G. Duong, T.L. Phan, T.L.H. Nguyen, T.P. Chau, V.-D. Doan, Effective reduction
nanoparticles from waste Nypa fruticans fruit husk: biosynthesis, characterization, of nitrophenols and colorimetric detection of Pb (ii) ions by Siraitia grosvenorii
antibacterial activity and recyclable catalysis, Arab. J. Chem. 13 (10) (2020) fruit extract capped gold nanoparticles, RSC Adv. 11 (25) (2021) 15438–15448.
7490–7503. [67] J.-H. Noh, R. Meijboom, Catalytic evaluation of dendrimer-templated Pd
[52] S.M. Albukhari, M. Ismail, K. Akhtar, E.Y. Danish, Catalytic reduction of nanoparticles in the reduction of 4-nitrophenol using Langmuir–Hinshelwood
nitrophenols and dyes using silver nanoparticles@ cellulose polymer paper for the kinetics, Appl. Surf. Sci. 320 (2014) 400–413.
resolution of waste water treatment challenges, Colloids Surf. A: Physicochem. [68] Y. Li, X.-Q. Zhu, Theoretical prediction of activation free energies of various
Eng. Asp. 577 (2019) 548–561. hydride self-exchange reactions in acetonitrile at 298 K, ACS Omega 3 (1) (2018)
[53] S.R. Thawarkar, B. Thombare, B.S. Munde, N.D. Khupse, Kinetic investigation for 872–885.
the catalytic reduction of nitrophenol using ionic liquid stabilized gold [69] H. Liao, G. Liu, Y. Liu, R. Li, W. Fu, L. Hu, Aggregation-induced accelerating
nanoparticles, RSC Adv. 8 (67) (2018) 38384–38390. peroxidase-like activity of gold nanoclusters and their applications for colorimetric
[54] A. Verma, N. Jain, S. Singha, M. Quraishi, I. Sinha, Green synthesis and catalytic Pb 2+ detection, Chem. Commun. 53 (73) (2017) 10160–10163.
application of curcumin stabilized silver nanoparticles, J. Chem. Sci. 128 (12) [70] C.-P. Liu, K.-C. Chen, C.-F. Su, P.-Y. Yu, P.-W. Lee, Revealing the active site of gold
(2016) 1871–1878. nanoparticles for the peroxidase-like activity: the determination of surface
[55] M.A. Mahmoud, B. Garlyyev, M.A. El-Sayed, Determining the mechanism of accessibility, Catalysts 9 (6) (2019) 517.
solution metallic nanocatalysis with solid and hollow nanoparticles: homogeneous [71] D. Hoare, J. Protheroe, A. Walsh, The thermal decomposition of hydrogen peroxide
or heterogeneous, J. Phys. Chem. C 117 (42) (2013) 21886–21893. vapour, Trans. Faraday Soc. 55 (1959) 548–557.
[56] S. Wunder, F. Polzer, Y. Lu, Y. Mei, M. Ballauff, Kinetic analysis of catalytic [72] N.R. Nirala, S. Pandey, A. Bansal, V.K. Singh, B. Mukherjee, P.S. Saxena,
reduction of 4-nitrophenol by metallic nanoparticles immobilized in spherical A. Srivastava, Different shades of cholesterol: gold nanoparticles supported on
polyelectrolyte brushes, J. Phys. Chem. C 114 (19) (2010) 8814–8820. MoS2 nanoribbons for enhanced colorimetric sensing of free cholesterol, Biosens.
[57] S. Gu, Y. Lu, J. Kaiser, M. Albrecht, M. Ballauff, Kinetic analysis of the reduction of Bioelectron. 74 (2015) 207–213.
4-nitrophenol catalyzed by Au/Pd nanoalloys immobilized in spherical [73] S. Singh, K. Mitra, R. Singh, A. Kumari, S.K.S. Gupta, N. Misra, P. Maiti, B. Ray,
polyelectrolyte brushes, Phys. Chem. Chem. Phys. 17 (42) (2015) 28137–28143. Colorimetric detection of hydrogen peroxide and glucose using brominated
[58] Z. Zhang, C. Shao, Y. Sun, J. Mu, M. Zhang, P. Zhang, Z. Guo, P. Liang, C. Wang, graphene, Anal. Methods 9 (47) (2017) 6675–6681.
Y. Liu, Tubular nanocomposite catalysts based on size-controlled and highly [74] Z. Liu, D. Ye, S. Wang, X. Zhu, R. Chen, Q. Liao, Single-stream H2O2 membraneless
dispersed silver nanoparticles assembled on electrospun silica nanotubes for microfluidic fuel cell and its application as a self-powered electrochemical sensor,
catalytic reduction of 4-nitrophenol, J. Mater. Chem. 22 (4) (2012) 1387–1395. Ind. Eng. Chem. Res. 59 (35) (2020) 15447–15453.
[59] A. Iben Ayad, D. Luart, A. Ould Dris, E. Guénin, Kinetic analysis of 4-nitrophenol
reduction by “water-soluble” palladium nanoparticles, Nanomaterials 10 (6)
(2020) 1169.

12

You might also like