Nanomaterial Based AOPs For The Removal of Organic Pollu - 2024 - Environmental

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

Environmental Technology & Innovation 35 (2024) 103718

Contents lists available at ScienceDirect

Environmental Technology & Innovation


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

Nanomaterial-based AOPs for the removal of organic pollutants in


aqueous matrices: A systematic review of response surface
methodology (RSM) models
Ramin Nabizadeh Nodehi , Samira Sheikhi *
Department of Environmental Health Engineering, School of Public Health, Tehran University of Medical Sciences, Tehran, Islamic Republic of Iran

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

Keywords: Organic pollutants in water mainly originate from human activities and can cause numerous
Advanced oxidation process problems for human health and the environment. In order to avoid these problems, organic
Nano compounds pollutants must be removed using appropriate and effective methods. Advanced oxidation pro­
Organic pollutants
cesses (AOPs), which comprise different technologies, have attracted attention in recent years due
RSM
Aqueous solutions
to their high efficiency in water and wastewater treatment. This systematic review evaluates the
effectiveness of nanomaterial-based AOPs for removing organic pollutants from aqueous solu­
tions. The uniqueness of this study is the investigation of response surface methodology (RSM)
models. Specifically, the present study reviewed six subgroups of nanomaterial-based AOPs,
including Fe-based, TiO2-based, Zn-based, Cu-based, Mg-based, and a set of disparate processes
categorized as others. The effects of process variables (including pH, contact time, initial pollutant
concentration, and catalyst dosage) were investigated by considering the classification of
different pollutant groups. In total, 71 studies were included in this review. The data were pri­
marily analyzed with R software, and graphs and figures were drawn if necessary. The average
pooled percentages of Fe-based, TiO2-based, Zn-based, Cu-based, and Mg-based processes for
organic pollutants removal, regardless of pollutant type, were 87.53 %, 82.61 %, 80.16 %,
82.93 %, and 87.93 %, respectively. Based on this systematic review, nano-based AOPs can effi­
ciently remove organic pollutants from aqueous matrices. Notably, however, the efficiency of the
process can be changed depending on the conditions applied in the system.

1. Introduction

It is clear that the preservation of life on earth depends on the preservation of water resources (Ighalo et al., 2021). Industriali­
zation, population growth, and human environmental activities have significantly contributed to water resource pollution (Yu et al.,
2018). A wide range of contaminants, including organic pollutants such as drugs, dyes, pesticides, etc., have been identified at varying
levels in aquatic environments. These contaminants primarily stem from human activities mentioned above and have the potential to
induce various diseases (Zhang et al., 2021). Therefore, these pollutants should be properly removed in water and wastewater
treatment plants to prevent diseases related to poor water and insufficient sanitation (Aghaeinejad-Meybodi et al., 2019; Sheikhi et al.,
2021b).

* Corresponding author.
E-mail address: s_sheikhi@razi.tums.ac.ir (S. Sheikhi).

https://doi.org/10.1016/j.eti.2024.103718
Received 22 April 2024; Received in revised form 27 May 2024; Accepted 13 June 2024
Available online 14 June 2024
2352-1864/© 2024 Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
R.N. Nodehi and S. Sheikhi Environmental Technology & Innovation 35 (2024) 103718

Conventional water and wastewater treatment techniques are ineffective for some pollutants in aqueous solutions that are resistant
to some degree (Bethi et al., 2016). In recent decades, advanced oxidation processes (AOPs) have been considered a powerful method
to remove organic pollutants from aquatic environments (Sheikhi et al., 2021a). AOPs use highly active radicals to completely
decompose organic pollutants into harmless species such as CO2 and H2O (Kurian, 2021). Hydroxyl radicals, sulfate radicals, super­
oxide radicals, and singlet oxygen are reactive species that play a major role in degrading refractory organic pollutants. It is important
to remember that different reactive species react differently to organic pollutants. (Wang and Wang, 2020; Kurian, 2021). Fast reaction
rates, the possibility of simultaneous treatment of several pollutants, and the production of less toxic intermediate products during the
degradation of organic pollutants are the advantages of AOPs (Ma et al., 2021a). Nano-based AOPs double the advantage of this
method.
Studies conducted on aqueous solutions treatment using nanomaterials show the high capacity of these materials to remove various
contaminants, including organic pollutants. Nanomaterials include nanoparticles whose size is smaller than 100 nm in at least one
dimension (Wu et al., 2020). According to studies, the nanoparticles used to treat aqueous solutions mainly include TiO2 (Vaez et al.,
2012; Shaykhi and Zinatizadeh, 2014), Fe3O4 (Khammar et al., 2020; Safari et al., 2014), ZnO (Mirzaei et al., 2018a; Bakhtkhosh and
Mehrizad, 2017), CuO (Senobari and Nezamzadeh-Ejhieh, 2018a), and MgO (Ahmadi et al., 2018), among others. TiO2 nanostructures
are widely used as semiconductors in water treatment due to their unique properties, such as non-toxicity, environmental compati­
bility, cost-effectiveness, insolubility, and high chemical stability (Rahimi et al., 2019). One of the principal problems in applying these
systems is the separation of TiO2 nanoparticles after treatment (Vaez et al., 2012).
Fe3O4 nanoparticles possess desirable properties such as superparamagnetic, environmental friendliness, excellent chemical sta­
bility, affordability, non-toxicity, and abundance in nature. These characteristics make Fe3O4 nanoparticles a proper choice for water
treatment as an effective heterogeneous catalyst (Villegas et al., 2020; Tajyani and Babaei, 2018). In addition, unlike TiO2, it can be
easily separated and reused. However, because of their large surface area and strong magnetic dipole-dipole interaction, Fe3O4
nanoparticles are susceptible to irreversible agglomeration, which reduces the efficiency of the process. To prevent the accumulation of
nanoparticles, they can be fixed on different supporting materials (for example, chitosan, graphene oxide, polymer, amino acids, and
cellulose) (Popescu et al., 2019; Li et al., 2018).
In recent years, nanoscale zero-valent iron (NZVI) particles have also been used to decompose a wide range of organic pollutants in
water and wastewater treatment processes (Ma et al., 2021b). Due to their wide application area, ZnO nanoparticles can be an
attractive material for wastewater treatment in various industries, including pharmaceutical and dyeing wastewater. This metal oxide
has high strength, is non-toxic, and is available at a low cost (Liu et al., 2018). Many studies have shown that ZnO performs better in the
nanoscale and removes complex compounds that conventional methods cannot remove (Shiri et al., 2020). The abundance of Cu
resources, desirable redox potential, large surface area, and high complexation strength are also factors that contribute to the
popularity of CuO photocatalysts. However, it is difficult to separate CuO nanoparticles from aqueous solution in a practical photo­
catalytic process, which can be overcome by supporting the finely powdered photocatalysts into proper microporous materials
(Hossain et al., 2020; Iazdani and Nezamzadeh-Ejhieh, 2021). As a heterogeneous catalyst, MgO nanoparticles are an appropriate
option for treating hazardous, antibacterial, and resistant organic pollutants. This compound is economical, stable, non-toxic,
non-volatile, and eco-friendly (Jebelli et al., 2018).
Recently, various review studies have been conducted in the field of advanced oxidation processes. Wang et al. (2022) reported that
metal organic frameworks based catalysts are promising for the removal of emerging organic contaminants through sulfate

Fig. 1. The flowchart of the search and selection of papers.

2
R.N. Nodehi and S. Sheikhi Environmental Technology & Innovation 35 (2024) 103718

Table 1
Summary of results of RSM investigations in individual studies.

Study D1 D2 D3 D4 D5 Study D1 D2 D3 D4 D5

Mahmoudabadi et al., 2022 Mohagheghian et al., 2022


Omrani et al., 2020 Khataee et al., 2010
Rahimi et al., 2019 Safari et al., 2014
Ansari et al., 2020 Belachew et al., 2020
Mondal et al., 2017 Mohammadi et al., 2016
Karimi et al., 2019 Mansouri et al., 2021
Vaez et al., 2012 Mansouri et al., 2017
Aghaeinejad et al., 2018 Naimi et al., 2023
Ahmadi et al., 2017 Shariati et al., 2022
Rahmani et al., 2017 Yashni et al., 2020
Asgari et al., 2020 (1) Lotfi et al., 2016
Asgari et al., 2020 (2) Bose et al., 2021
Ma et al., 2021 Senobari et al., 2018
Shaykhi et al., 2014 Airemlou et al., 2019
Galedari et al., 2019 Hosseini et al., 2017
Gholipoor et al., 2021 Mirzaei et al., 2018
Doria et al., 2013 Golmohammadi et al., 2022
Moradi et al., 2016 Song et al., 2016
M. Tayeb et al., 2018 Wei et al., 2019
Khoshnamvand et al., 2018 Priya et al., 2018
Zhang et al., 2022 Massoudinejad et al., 2019
Joshaghani et al., 2017 Gharbani et al., 2022
Ahmadi et al., 2018 Bakhtkhosh et al., 2017
Soltani et al., 2021 Hassani et al., 2016
Samy et al., 2020 Mansouri et al., 2017
Mirzaei et al., 2018 khodadoost et al., 2017
Sun et al., 2013 Pirsaheb et al., 2018
Biglar et al., 2021 Senobari et al., 2018
Shahbazi et al., 2018 Ciğeroğlu et al., 2021
Tony et al., 2017 Rafiee et al., 2018
Pourali et al., 2022 Doumbi et al., 2021
Eskandari et al., 2022 Afshar et al., 2018
Desai et al., 2020 Karimi et al., 2020
Aeindartehran et al., 2021 Shabanloo et al., 2019
Khammar et al., 2020 Souf et al., 2022
Mukhopadhyay et al., 2021

D1: Statistical methods


D2: Validation of the model
D3: Process optimization
D4: Process kinetics
D5: Confounding

3
R.N. Nodehi and S. Sheikhi Environmental Technology & Innovation 35 (2024) 103718

Table 2
Summary of the Fe-based AOPs for the removal of organic pollutants in aqueous solution.
Compounds Pollutants classification Pollutant name AOP features Ref.

ZnO/α-Fe2O3 Drug Amoxicillin pH = 4.5 (Ansari et al., 2020)


Time = 18
Catalyst dose = 0.4
Initial pollutant conc. = 16
Removal efficiency = 99.30
Nano-Fe3O4/Fenton-like Carbamazepine pH = 7 (Sun et al., 2013)
Time = 60
Catalyst dose = 1.84
Initial pollutant conc. = 15
Removal efficiency = 86.26
US/PS/nZVI Ciprofloxacin pH = 4.5 (Rahmani et al., 2017)
Time = 60
Catalyst dose = 0.12
Initial pollutant conc. = 50
Removal efficiency = 57
nZVI/H2O2 Ciprofloxacin pH = 7 (Mondal et al., 2018)
Time = 120
Catalyst dose = 0.28
Initial pollutant conc. = 10
Removal efficiency = 99.30
Co-Fe2O4 Cytotoxic drugs pH = 4.32 (Ershadi Afshar et al., 2018)
Time = 60
Catalyst dose = 0.16
Initial pollutant conc. = 41.43
Removal efficiency = 93
nZVIPs@Ti3C2 Ranitidine pH = 4.5 (Ma et al., 2021c)
Time = 6
Catalyst dose = 0.75
Initial pollutant conc. = 10
Removal efficiency = 92.61
Magnetic ZnO@g-C3N4- Fe3O4 Sulfamethoxazole pH = 5.6 (Mirzaei et al., 2018a)
Time = 60
Catalyst dose = 0.65
Initial pollutant conc. = 30
Removal efficiency = 90.40
S-nZVI/BC Sulfonamide pH = 3.18 (Zhang et al., 2022)
Time = 30
Catalyst dose = 10
Initial pollutant conc. = 0.33
Removal efficiency = 100
TiO2/Fe2O3 Tetracycline pH = 4.7 (Galedari et al., 2019)
Time = 89.85
Catalyst dose = 0.61
Initial pollutant conc. = 13.86
Removal efficiency = 81.30
PS/nZVI Dye Acid blue 113 pH = 3 (Pourali et al., 2022)
Time = 50
Catalyst dose = 0.08
Initial pollutant conc. = 46
Removal efficiency = 100
Mg/Fe2O4 Brilliant green pH = 8.5 (Bose et al., 2021)
Time = 10
Catalyst dose = 0.75
Initial pollutant conc. = 75
Removal efficiency = 91.63
Ca-Fe2O4 Brilliant green pH = 8.1 (Mukhopadhyay et al., 2021)
Time = 15
Catalyst dose = 0.5
Initial pollutant conc. = 50
Removal efficiency = 99
nano‑hematite photo‑Fenton Methylene blue pH = 2.5 (Tayeb et al., 2018)
Time = 90
Catalyst dose = 0.041
Initial pollutant conc. = 10
Removal efficiency = 81.60
Cu-Fe2O4 Tartrazine pH = 3 (Soufi et al., 2022)
Time = 60
Catalyst dose = 0.32
(continued on next page)

4
R.N. Nodehi and S. Sheikhi Environmental Technology & Innovation 35 (2024) 103718

Table 2 (continued )
Compounds Pollutants classification Pollutant name AOP features Ref.

Initial pollutant conc. = 30.63


Removal efficiency = 97.90
nFe3O4@CNT Pesticide Malathion pH = 5 (Eskandarimakvand et al., 2022)
Time = 60
Catalyst dose = 0.4
Initial pollutant conc. = 35.15
Removal efficiency = 82
NZVI Others Benzotriazole pH = 3 (Ahmadi et al., 2017)
Time = 60
Catalyst dose = 0.1
Initial pollutant conc. = 15
Removal efficiency = 73.40
Mn-Fe2O4 4- chlorophenol pH = 4.24 (Hassani et al., 2016)
Time = 61
Catalyst dose = 0.72
Initial pollutant conc. = 23.11
Removal efficiency = 93
Ag2CrO4/Ag/Fe3O4/RGO Methylene blue pH = 2 (Shariati et al., 2022)
Time = 103
Catalyst dose = 0.01
Initial pollutant conc. = 12
Removal efficiency = 97.6
Fe3O4/ZnO MTBE pH = 7.2 (Safari et al., 2014)
Time = 20
Catalyst dose = 1.78
Initial pollutant conc. = 89.14
Removal efficiency = 70
AK-Fe3O4 4-nitrophenol pH = 3 (Belachew et al., 2020)
Time = 75
Catalyst dose = 1
Initial pollutant conc. = 20
Removal efficiency = 96.01
Fe3O4@TiO2 PCBs pH = 5.6 (Khammar et al., 2020)
Time = 16.07
Removal efficiency = 83
Fentonlike/Nano -Hematite Phenol pH = 3.25 (Tony et al., 2018)
Time = 5
Catalyst dose = 0.04
Removal efficiency = 51
P25/Fe0/ZnO P-nitrophenol pH = 75 (Joshaghani et al., 2017)
Time = 240
Catalyst dose = 0.5
Initial pollutant conc. = 20
Removal efficiency = 86.90
α-Fe2O3 Vinasse pH = 7.36 (Mahmoudabadi et al., 2022)
Time = 90
Removal efficiency = 98.64

The time is in min, the catalyst dose is in g/L and the initial pollutant concentration is in mg/L. The removal efficiency is in %.

radicals-based AOPs (Wang et al., 2022). The results of Liu et al.’s (2023) study showed that heterogeneous AOPs using bio­
char/layered double hydroxides composites is an effective method for removing organic pollutants from wastewater (Liu et al., 2023).
Also, several studies have examined the nanomaterial-based AOPs for organic pollutants removal (Mahmoudabadi et al., 2022; Omrani
and Nezamzadeh-Ejhieh, 2020; Rahimi et al., 2019; Vaez et al., 2012). The purpose of this study was to systematically analyze
published papers and compare the effectiveness of nanomaterial-based AOPs in removing organic pollutants from aqueous solutions.
This study focused on the response surface methodology (RSM) models. RSM is a reliable statistical tool used for process optimization
and analysis of the relative importance of different operational parameters, even in complex systems (Khammar et al., 2020). In fact, in
this method, the output variable (response variable), which is affected by several input variables (independent variables) is optimized
so that it reaches the desired maximum or minimum value. Response surface methods can be designed in different ways (such as CCD,
D-Optimal and Box-Behnken methods) depending on their application in the experimental design (Asgari et al., 2020b; Ciğeroğlu et al.,
2022; Naimi-Joubani et al., 2023). In each experiment, changes in the input variables are made in order to determine the causes of
changes in the response variable. When responses are influenced by multiple inputs, RSM is used to optimize, improve and develop
processes (Arslan-Alaton et al., 2009). The novelty of this study is the review of RSM models to report the effective parameters in
nanomaterial-based AOPs, the statistical analysis of these parameters for different nanomaterials, the comparison of the removal
efficiency of organic pollutants by different nanomaterials, and providing an overview of the existing trends and challenges. This study
determines the extent to which the research area has improved, and the findings will be useful to researchers, policymakers, and other
stakeholders.

5
R.N. Nodehi and S. Sheikhi Environmental Technology & Innovation 35 (2024) 103718

Table 3
Summary of the TiO2-based AOPs for the removal of organic pollutants in aqueous solution.
Compounds Pollutants classification Pollutant name AOP features Ref.

TiO2 nanoparticles Drug Amoxicillin pH = 3 (Shaykhi and Zinatizadeh, 2014)


Time = 30
Catalyst dose = 0.55
Removal efficiency = 14
BiVO4/TiO2 Dye Acid Orange 10 pH = 3 (Rahimi et al., 2019)
Time = 50
Catalyst dose = 1.7
Initial pollutant conc. = 10
Removal efficiency = 99.91
TiO2 nanoparticles Acid Red 73 pH = 3 (Vaez et al., 2012)
Time = 60
Initial pollutant conc. = 25.3
Removal efficiency = 92.17
TiO2 nanoparticles Basic Blue 3 Time = 120 (Khataee et al., 2010)
Initial pollutant conc. = 10
Removal efficiency = 71
TiO2 nanoparticles Direct Blue 71 pH = 6.5 (Massoudinejad et al., 2019)
Time = 67.5
Catalyst dose = 1.2
Initial pollutant conc. = 55
Removal efficiency = 90.20
Ag-TiO2/PW nanocomposite DR16 pH = 2 (Rafiee et al., 2018)
Time = 120
Catalyst dose = 1
Initial pollutant conc. = 25
Removal efficiency = 94
TiO2–ZnO–NiO nanoparticles Methyl orange pH = 3 (Mansouri et al., 2022)
Time = 60
Catalyst dose = 0.3
Initial pollutant conc. = 15
Removal efficiency = 63.34
Sr/Ag-TiO2@g-C3N4 Reactive Black-42 pH = 4.5 (Wei et al., 2020)
Time = 40
Catalyst dose = 0.2
Initial pollutant conc. = 20
Removal efficiency = 95.60
UV/TiO2-ZnO-Co Reactive Red 120 pH = 7 (Mansouri et al., 2017b)
Time = 90
Catalyst dose = 0.1
Initial pollutant conc. = 16.4
Removal efficiency = 88.99
N-TiO2 Others Formaldehyde pH = 5 (Mirzaei et al., 2018b)
Time = 90
Removal efficiency = 60.15
TiO2-ZnO-CuO nanoparticles MTBE pH = 7 (Mansouri et al., 2017a)
Time = 60
Catalyst dose = 1.49
Initial pollutant conc. = 31.46
Removal efficiency = 95.63
TiO2-ZnO-CoO nanoparticle MTBE pH = 7.68 (Lotfi et al., 2016)
Time = 60
Catalyst dose = 1.68
Initial pollutant conc. = 30.58
Removal efficiency = 92.36
TiO2 nanoparticles Organic contaminants pH = 5 (Desai et al., 2020)
Time = 100
Catalyst dose = 0.75
Removal efficiency = 32.19
Nano TiO2 over Graphene Phenol pH = 6.1 (Shahbazi et al., 2018)
Time = 240
Catalyst dose = 1.48
Initial pollutant conc. = 14.04
Removal efficiency = 100

The time is in min, the catalyst dose is in g/L and the initial pollutant concentration is in mg/L. The removal efficiency is in %.

6
R.N. Nodehi and S. Sheikhi Environmental Technology & Innovation 35 (2024) 103718

Table 4
Summary of the Zn-based AOPs for the removal of organic pollutants in aqueous solution.
Compounds Pollutants classification Pollutant name AOP features Ref.

UV/ZnO Drug Aspirin pH = 5.05 (Karimi et al., 2019)


Time = 90.5
Catalyst dose = 0.37
Initial pollutant conc. = 33.84
Removal efficiency = 83.11
ZnO/GO nanocomposite Cefxime trihydrate pH = 4.03 (Ciğeroğlu et al., 2022)
Time = 60
Catalyst dose = 0.53
Initial pollutant conc. = 20.13
Removal efficiency = 86
UV/ZnO Naproxen pH = 6.87 (Karimi and Malakootian, 2020)
Time = 71.68
Catalyst dose = 0.37
Initial pollutant conc. = 21.59
Removal efficiency = 71.19
US/UV/ZnO/PS Dye AB113 pH = 6.1 (Asgari et al., 2020b)
Time = 25
Catalyst dose = 0.88
Removal efficiency = 98.70
Ni-Decorated ZnO NPs Acid Blue 1 pH = 7.77 (Aeindartehran and Talesh, 2021)
Time = 30
Catalyst dose = 1.07
Initial pollutant conc. = 10
Removal efficiency = 89.13
ZnO NPs Basic Red 51 pH = 3 (Yashni et al., 2020)
Time = 300
Catalyst dose = 0.13
Initial pollutant conc. = 7.47
Removal efficiency = 72.18
ZnS nanoparticles Direct Blue 14 pH = 3.5 (Bakhtkhosh and Mehrizad, 2017)
Time = 48
Catalyst dose = 0.8
Initial pollutant conc. = 10
Removal efficiency = 86.36
nano ZnO-Nd -photocatalyst Reactive red 198 pH = 5 (Biglar et al., 2021)
Time = 90
Catalyst dose = 0.1
Initial pollutant conc. = 10
Removal efficiency = 100
Mg/ZnO nanoparticles Rhodamine B Time = 10 (Airemlou et al., 2019)
Catalyst dose = 0.24
Initial pollutant conc. = 5
Removal efficiency = 49.61
Ni-doped ZnO nanorods Pesticide Diazinon pH = 5 (Naimi-Joubani et al., 2023)
Time = 15
Catalyst dose = 1
Initial pollutant conc. = 15
Removal efficiency = 82.29
nZnC Metaldehyde pH = 10.4 (Doria et al., 2013)
Time = 10
Initial pollutant conc. = 0.5
Removal efficiency = 44.40
GO/ZnO nanocomposite Others Methylene blue pH = 4.5 (Hosseini and Babaei, 2017)
Time = 140
Catalyst dose = 2.1
Initial pollutant conc. = 10
Removal efficiency = 99

The time is in min, the catalyst dose is in g/L and the initial pollutant concentration is in mg/L. The removal efficiency is in %.

2. Methods

2.1. Search strategy

A comprehensive literature search was performed on the effectiveness of nanomaterial-based AOPs in removing organic pollutants
from aqueous solutions. The search was conducted in Scopus, Google Scholar, Science Direct, and PubMed from January 2010 to June
2023. Search terms included ["advance oxidation process" OR "AOP"] AND ["nano compounds" OR "nano-based AOPs"] AND
["response surface methodology" OR "RSM"] AND ["aqueous solution" OR "water" OR "wastewater"] AND ["organic pollutants"] AND

7
R.N. Nodehi and S. Sheikhi Environmental Technology & Innovation 35 (2024) 103718

Table 5
Summary of the Cu-based AOPs for the removal of organic pollutants in aqueous solution.
Compounds Pollutants classification Pollutant name AOP features Ref.

nCuO/UV Drug Ciprofloxacin pH = 8.17 (Khoshnamvand et al., 2018)


Time = 60
Catalyst dose = 0.08
Initial pollutant conc. = 11.2
Removal efficiency = 96.58
CuO-Kaolin nanocomposite Pesticide Diazinon pH = 7 (Mohagheghian et al., 2022)
Time = 180
Catalyst dose = 0.4
Initial pollutant conc. = 30
Removal efficiency = 87.23
CuO-NiO nanoparticles Others Methylene blue pH = 3.5 (Senobari and Nezamzadeh-Ejhieh, 2018a)
Time = 26
Catalyst dose = 0.9
Initial pollutant conc. = 3.2
Removal efficiency = 40
Cu-MOF nanocatalyst Phenol pH = 7 (Gholipoor and Hosseini, 2021)
Time = 30
Catalyst dose = 1.5
Removal efficiency = 91.87
CuO NPs TNBP Time = 90 (Golmohammadi and Sattari, 2022)
Initial pollutant conc. = 4 %(v/v)
Removal efficiency = 99

The time is in min, the catalyst dose is in g/L and the initial pollutant concentration is in mg/L. The removal efficiency is in %.

Table 6
Summary of the Mg-based AOPs for the removal of organic pollutants in aqueous solution.
Compounds Pollutants classification Pollutant name AOP features Ref.

H2O2/MgO nanoparticles Dye Reactive Blue 19 pH = 3 (Ahmadi et al., 2018)


Time = 60
Initial pollutant conc. = 80
Removal efficiency = 93.77
Nano-MgO@CNT@Gr/O3 Pesticide Diazinon pH = 10 (Asgari et al., 2020a)
Time = 15
Catalyst dose = 1.5
Initial pollutant conc. = 10
Removal efficiency = 82.43
COP/MgO nanoparticles Other Toluene pH = 12 (Mohammadi et al., 2016)
Time = 50
Catalyst dose = 0.5
Initial pollutant conc. = 10
Removal efficiency = 87.60

The time is in min, the catalyst dose is in g/L and the initial pollutant concentration is in mg/L. The removal efficiency is in %.

["treatment" OR "degradation" OR "removal"]. In order to select a larger number of studies, additional records were identified by hand
searching. All references were imported into Endnote (version 21), and the inclusion and exclusion criteria were checked. The
availability of data (including experimental design and RSM model) and the acceptable quality of the studies were among the eligi­
bility criteria. Duplicate articles indexed in the mentioned databases were removed. Moreover, many articles were removed by
reviewing their titles and abstracts. Finally, 71 unique articles were included in this study. The diagram of the search procedure and the
number of obtained studies are shown in Fig. 1.

2.2. Statistical analysis

All data were statistically analyzed using the R software (version 4.3.0) and Microsoft Excel 2016. A graphical representation was
used to provide deeper and faster insight into the characteristics of the data. Histogram plots were drawn to show the entire distri­
bution, and box plots were used to visualize potential outliers. Panel plots helped visualize several different data sets in a single plot.
The calculated descriptive statistics were mean, median, variance, standard error, standard deviation, kurtosis, skewness, maximum,
and minimum (Tables 1–8). The statistical analysis takes into account the optimal study conditions.

2.3. Study characteristics

All studies were published after 2010, and their number increased over time. For a better understanding of the studies, five domains

8
R.N. Nodehi and S. Sheikhi Environmental Technology & Innovation 35 (2024) 103718

Table 7
Summary of the other AOPs for the removal of organic pollutants in aqueous solution.
Compounds Pollutants Pollutant name AOP features Ref.
classification

Nano-γ-alumina Drug Fluoxetine pH = 7 (Aghaeinejad-Meybodi et al., 2019)


Time = 30
Catalyst dose = 1
Initial pollutant conc. =
28.56
Removal efficiency = 96.14
BiVO4/WO3 Sulfasalazine pH = 4 (Omrani and Nezamzadeh-Ejhieh, 2020)
Time = 40
Catalyst dose = 0.8
Initial pollutant conc. = 9
Removal efficiency = 87.60
ESMNPs/NCB/alginate Tetracycline pH = 4 (Soltani et al., 2021)
Time = 60
Catalyst dose = 1
Initial pollutant conc. = 51
Removal efficiency = 15.80
BWO-GO photocatalysts Tetracycline Time = 90 (Song et al., 2016)
Catalyst dose = 0.3
Initial pollutant conc. = 10
Removal efficiency = 78.43
Bi4Ti3O12 nanoparticles Tetracycline pH = 11.8 (Khodadoost et al., 2017)
Time = 114
Initial pollutant conc. = 50
Removal efficiency = 63.90
BFO/Mr-cl-poly(AAm)-IPN-poly Dye Brilliant Blue pH = 7 (Kaith et al., 2018)
(AA) Time = 240
Catalyst dose = 0.5
Initial pollutant conc. = 10
Removal efficiency = 90.40
Nano-ZrO2/UV/PS Direct Blue 71 pH = 7 (Moradi et al., 2016)
Time = 40
Catalyst dose = 0.4
Initial pollutant conc. = 50
Removal efficiency = 97.10
MnO2 nanoparticles Methyl Orange pH = 3 (Doumbi and Noumi, 2021)
Time = 336
Catalyst dose = 1
Initial pollutant conc. = 25
Removal efficiency = 77
GNPs/ZrV2O7 Pesticide Chlorpyrifos pH = 4.05 (Samy et al., 2020)
Time = 60
Initial pollutant conc. = 44.9
Removal efficiency = 99.99
Nano-sized Mn3O4 Others Furfural pH = 4.82 (Shabanloo et al., 2020)
Time = 60
Catalyst dose = 1.2
Initial pollutant conc. = 50
Removal efficiency = 91.14
CdSe nanoparticles Methylene Blue pH = 8 (Gharbani et al., 2022)
Time = 20
Catalyst dose = 0.4
Initial pollutant conc. = 20
Removal efficiency = 92.80
NiO-CdS nanoparticles Methylene Blue pH = 3.5 (Senobari and Nezamzadeh-Ejhieh,
Time = 83 2018b)
Catalyst dose = 0.9
Initial pollutant conc. = 3.2
Removal efficiency = 85
Nano carbon dots (NCDs) Phenol pH = 3 (Pirsaheb et al., 2018)
Time = 19
Catalyst dose = 1
Initial pollutant conc. = 50
Removal efficiency = 98

The time is in min, the catalyst dose is in g/L and the initial pollutant concentration is in mg/L. The removal efficiency is in %.

9
R.N. Nodehi and S. Sheikhi Environmental Technology & Innovation 35 (2024) 103718

Table 8
Descriptive statistics of the examined variables without considering the classification of pollutants and nanomaterial-based processes.
Parameter pH Time (min) Initial concentration (mg/L) Catalyst dose (g/L) Removal efficiency (%)

Mean 5.38 73.80 25.24 0.70 83.16


Standard Error 0.28 7.78 2.49 0.07 2.31
Median 5 60 20 0.55 90.20
Mode 3 60 10 1 100
Standard Deviation 2.30 65.59 19.58 0.53 19.49
Sample Variance 5.28 4301.57 383.25 0.28 379.94
Kurtosis 0.40 5.29 1.50 -0.15 3.42
Skewness 0.83 2.17 1.33 0.77 -1.87
Minimum 2 5 0.50 0.01 14
Maximum 12 336 89.14 2.10 100
Count 68 71 62 61 71
Confidence Level(95/0 %) 0.56 15.52 4.97 0.13 4.61

(including statistical methods, validation of the model, process optimization, process kinetics, and confounding) were considered and
investigated in the studies. A summary of RSM investigations in individual studies is presented in Table 1. None of the studies in the
five domains were excluded due to bias. This review summarizes the general findings of the studies and presents them in the form of
graphs and tables. An analysis was performed on the variables used in the experimental design of the eligible studies. More precisely,
the four factors of pH, time, catalyst dose, and initial pollutant concentration, which were typically considered as variables in most
studies, were extracted and analyzed.

3. Results and discussion

3.1. Different parameters in AOP processes

After reviewing the studies for a better and more focused analysis of the results, they were divided into six main groups, including
Fe-based, TiO2-based, Cu-based, Zn-based, Mg-based, and Others, based on the type of nanomaterials that were used in the advanced
oxidation process. The processes that used nanomaterials other than the mentioned nanomaterials were categorized as "Others." This
category entailed BiVO4/WO3, nano-γ-alumina, ZrO2/UV/PS, ESMNPs/NCB/alginate, GNPs/ZrV2O7, NiO/CdS, BWO/GO, BFO/Mr-cl-
poly(AAm)-IPN-poly(AA), CdSe, Bi4Ti3O12, nano carbon dots (NCDs), MnO2, and Mn3O4. We extracted the optimal values of the four
main variables in advanced oxidation processes, i.e., pH, time, initial concentration, and catalyst dosage in the six mentioned groups. A
summary of the obtained results is illustrated as a box plot in Fig. 2. The plots are arranged according to the median (The horizontal
line inside the box shows the median value.).
According to Fig. 2(a), the highest mean pH (pH = 8.33) was linked with the Mg-based process, and the lowest (pH = 4.82) to the
TiO2-based process. On the other hand, the lowest average reaction time (67.4 min) was related to the Mg-based process (Fig. 2(b)).
Also, the highest mean initial pollutant concentration and catalyst dose was associated with the Mg-based process (Fig. 2(c), (d)),
which can be attributed to the limited number of studies related to this process (n = 3) compared to other processes. The lowest mean
score for initial pollutant concentration and catalyst dose were related to Zn-based and Fe-based processes, respectively. In most of the
studies, the optimal pollutant concentration was higher than the concentration in the real aqueous environment (Aghaeinejad-Mey­
bodi et al., 2019; Asgari et al., 2020b; Khoshnamvand et al., 2018; Song et al., 2016). On the other hand, in some studies, the optimal
concentration of the pollutant was found to be the same concentration in the real aqueous environment (Mondal et al., 2018; Sha­
banloo et al., 2020). More details related to each process are given in the relevant section. Additionally, pollutants were divided into
four general groups of drugs, pesticides, dyes, and others according to frequency, and the necessary descriptive diagrams were drawn
(Section 3.3).

3.2. Nano-based AOPs for organic pollutants removal

In the last two decades, many studies have been conducted regarding the importance of using nanoparticles in AOPs for the
degradation of various organic pollutants (Wei et al., 2020). Nanomaterials are one of the most suitable options as catalysts, whose
physicochemical properties depend on the nature and structure of the materials in terms of shape and size (Mondal et al., 2018). AOP
hybrid processes have expanded over time, and nanoparticles such as ZVI, Fe3O4, TiO2, ZnO, CuO, ZnC, MgO, etc. have been added to
the process to increase the efficiency of pollutant removal. In this section, we review advanced oxidation processes that were used in
combination with nanoparticles.

3.2.1. Fe-based AOPs


Among all studies, the number of research articles pertaining to Fe-based AOPs was the highest (n = 24). Table 2 presents a
summary of Fe-based processes for the removal of various organic pollutants. As can be seen, most pollutants removed by this process
are related to drugs (amoxicillin, ciprofloxacin, ranitidine, tetracycline, sulfonamide, carbamazepine, sulfamethoxazole, and cytotoxic
drugs) and the "others" category (vinasse, benzotriazole, p-nitrophenol, phenol, PCBs, MTBE, 4-nitrophenol, methylene blue and 4-

10
R.N. Nodehi and S. Sheikhi Environmental Technology & Innovation 35 (2024) 103718

Fig. 2. The box plots of (a) pH, (b) time, (c) initial pollutant concentration and (d) catalyst dose in different groups of nanoparticles.

chlorophenol). There was only one study on pesticides in this review. The average pH values in the three groups of drugs, dyes, and
"others" were 5.03, 5.02, and 4.79, respectively. Also, the pH related to the removal of malathion pesticide was 5. The results indicate
that Fe-based processes with an acidic pH of about 5 can remove the highest amount of organic pollutants. These findings further
support the idea that the reaction system can generate additional free radicals and active species in low pH environments and lead to
increased organic pollutant removal efficiency (Rahmani et al., 2017). The average reaction times in the three groups of drugs, dyes,
and others were 55.98, 45, and 74.45 min, respectively. The optimal reaction time for the removal of malathion pesticide was 60 min.
Therefore, it can be concluded that Fe-based processes lead to the maximum removal of organic pollutants within a reaction time of

11
R.N. Nodehi and S. Sheikhi Environmental Technology & Innovation 35 (2024) 103718

about 55–75 min. It is worth mentioning that at the initial stages of the process, rapid removal of pollutants occurs, which can be
attributed to the higher number of adsorption sites on the catalyst surface (Khammar et al., 2020). The average initial concentrations of
the pollutant in the three mentioned groups were 21.81, 42.32, and 29.87 mg/L, respectively. Therefore, it can be said that a higher
initial pollutant concentration is required to remove dyes in Fe-based processes. It is worth mentioning that depending on the type of
pollutant, increasing or decreasing the initial concentration of the pollutant leads to an increase in the efficiency of the process
(Khammar et al., 2020). Also, the average doses of the catalyst in the three mentioned groups were 0.57, 0.33, and 0.59 g/L,
respectively. Therefore, these processes have the highest removal efficiency of organic pollutants in the catalyst dosage range of
0.3–0.6 g/L. According to the findings of the studies, when the catalyst dose increases (to a certain extent), the pollutant removal
efficiency also increases (Safari et al., 2014). The average efficiency values through Fe-based processes were 88.79 % for drug removal,
94.02 % for dyes, and 83.28 for the others group. Generally, the average removal efficiency of organic pollutants with this process was
87.53 %, regardless of the type of pollutants. Fe-based processes for removing organic pollutants from aqueous environments seem to
have an efficiency higher than 80 % in optimal conditions.

3.2.2. TiO2-based AOPs


Table 3 presents a summary of TiO2-based processes for removing organic pollutants. As can be seen, the most pollutants removed
by this process are related to dyes (acid orange 10, acid red 73, basic blue 3, reactive black 42, direct blue 71, reactive red 120, and
direct red 16) and the others group (such as formaldehyde, phenol, methyl orange, and MTBE). In this review, there was no study on
pesticides and only one study on drugs. The average pH values in the two groups of dyes and the others were 4.33 and 5.63,
respectively. Moreover, the pH related to the removal of the amoxicillin drug was 3. The results indicated that as long as the pH of the
solution was in the acidic range, the removal rate of organic pollutants increased. Notably, in the oxidation process, the effect of pH
depends on the type of pollution and zero-point charge (PZC) of the catalyst (Mansouri et al., 2017a). The TiO2 surface has a positive
charge in acidic conditions. When the pH of the solution is higher than pHpzc, the number of negatively charged sites on the TiO2
surface increases. The resulting electrostatic repulsion reduces pollutant absorption by the catalyst (Vaez et al., 2012; Rafiee et al.,
2018). The average reaction times in the two groups of dyes and others were 78.21 and 101.66 min, respectively. In general, regardless
of the type of pollutant, the average reaction time for TiO2-based processes was 84.82 min. Of course, depending on the conditions
applied in the process, the retention time may be higher or lower than this value. The average initial concentration of the pollutant in
the two mentioned groups was 23.10 and 22.77 mg/L, respectively. Therefore, it can be said that to increase the removal efficiency by
TiO2-based processes, dyes need a higher initial concentration than other organic pollutants. Mansouri et al. reported that by
increasing the initial concentration of methyl orange dye, the number of active sites of nanoparticles decreases due to the increase in
the amount of dye absorbed on the surface of nanoparticles. As a result, the production of hydroxyl radicals and the rate of dye
decomposition decrease (Mansouri et al., 2022). The average dose of the catalyst in the two groups of dyes and others was 0.84 and
1.14 g/L, respectively. The catalyst dose related to the removal of amoxicillin was 0.55 g/L. The catalyst dosage plays an important
role in AOP processes due to its contribution to active free radical production and their interaction. On the other hand, the catalyst dose
depends on the pollutant’s initial concentration and other process conditions (Mansouri et al., 2022). The average removal efficiency
of TiO2-based processes was 90.26 % for dyes and 73.94 % for the others group. In general, the average removal efficiency of organic
pollutants by this process without pollutant separation was 82.61 %. It can be said that TiO2-based processes have the highest removal
efficiency for dyes.

3.2.3. Zn-based AOPs


Table 4 presents a summary of Zn-based processes for removing organic pollutants. As can be seen, the pollutants primarily
removed by this process are related to dyes (acid blue113, reactive red 198, acid blue 1, basic red 51, rhodamine B, and direct blue 14).
The average pH values in the groups of dyes, drugs, and pesticides were 5.07, 5.31, and 7.7, respectively. These results indicate that Zn-
based systems require a higher pH to remove pesticides than other pollutants. In the case of drugs (such as aspirin, cefxime trihydrate
and naproxen) and dyes (such as basic red 51, direct blue 14 and reactive red 198) when the pH of the solution is lower than pHpzc, the
ZnO surface has a positive charge and can absorb anionic pollutants and the removal efficiency increases. At pH values higher than
pHpzc, the surface of ZnO becomes negatively charged, and as a result, the rate of absorption and removal efficiency decreases
(Bakhtkhosh and Mehrizad, 2017; Biglar et al., 2021; Ciğeroğlu et al., 2022, Karimi et al., 2019; Karimi and Malakootian, 2020; Yashni
et al., 2020). On the other hand, the study by Doria et al. showed that nano-sized zinc oxide/laponite composites are very stable in
alkaline conditions and can remove metaldehyde pesticide up to 50 % under optimal conditions (Doria et al., 2013). The average
reaction times in groups of dyes, drugs, and pesticides were 83.83, 74.06, and 12.50 min, respectively. Therefore, Zn-based systems
require less time to remove pesticides than other pollutants. However, in the case of all pollutants, the process efficiency increases with
the increase in contact time between the pollutant and the catalyst (Asgari et al., 2020b; Ciğeroğlu et al., 2022; Naimi-Joubani et al.,
2023). The average initial concentrations of the pollutant in the three mentioned groups were 8.49, 25.18, and 7.75 mg/L, respec­
tively. Evidently, the average initial concentration of drugs was significantly different from those of the other two groups, indicating
that a higher initial concentration of the drug is needed for the effective removal of drugs through Zn-based processes. In general, it can
be said that by increasing the concentration of the pollutant due to the saturation of the absorbent active sites on the one hand and
reducing the rate of dissolution of the pollutant (due to the constant amount of hydroxyl radicals produced) on the other hand, the
efficiency of the process decreases (Karimi et al., 2019). The average doses of the catalyst in the dyes and drugs groups were 0.53 and
0.42 g/L, respectively. Therefore, it can be said that Zn-based processes have the highest removal efficiency of dyes and drugs at a
catalyst dose of about 0.5 g/L. Of course, this value increases or decreases depending on the type of pollutant and catalyst. Regarding
the other two groups (pesticides and others), it is not possible to make a precise statement due to the small number of studies. The mean

12
R.N. Nodehi and S. Sheikhi Environmental Technology & Innovation 35 (2024) 103718

Fig. 3. The box plots of (a) pH, (b) time, (c) initial pollutant concentration and (d) catalyst dose according to the pollutants classification.

removal efficiency coefficients of dyes, drugs, and pesticides through Zn-based processes were 82.66 %, 80.10 %, and 63.34 %,
respectively. In general, the average removal efficiency of organic pollutants with this process is 80.16 %, regardless of the type of
pollutants. It seems that Zn-based processes for the removal of dyes and drugs from aqueous environments have an efficiency higher
than 80 % under optimal conditions.

3.2.4. Cu-based AOPs


Table 5 presents a summary of Cu-based processes for the removal of various organic pollutants. Owing to the limited number of
investigations pertaining to each pollutant, a combined interpretation of the results is presented in this section. The average pH,
retention time, initial pollutant concentration, and catalyst dose related to removing organic pollutants by Cu-based processes were
6.41, 77.20 min, 14.8 mg/L, and 0.72 g/L, respectively. Notably, the values of each of these parameters can vary depending on the

13
R.N. Nodehi and S. Sheikhi Environmental Technology & Innovation 35 (2024) 103718

Table 9
The average of the investigated variables considering the classification of pollutants and nanomaterial-based processes.
Average of pH Average of Time Average of Initial pollutant conc. Average of Catalyst dose Average of Removal efficiency

Drug 5.51 60.53 23.98 0.56 78.52


Cu-based 8.17 60.00 11.20 0.08 96.58
Fe-based 5.03 55.98 21.81 0.57 88.80
Other 6.70 66.80 29.71 0.78 68.37
TiO2- 3.00 30.00 0.55 14.00
based
Zn-based 5.32 74.06 25.19 0.42 80.10
Dye 4.82 88.70 27.66 0.58 88.92
Fe-based 5.02 45.00 42.33 0.34 94.03
Mg-based 3.00 60.00 80.00 93.77
Other 5.67 205.33 28.33 0.63 88.17
TiO2- 4.33 78.21 23.10 0.84 90.27
based
Zn-based 5.07 83.83 8.49 0.54 82.66
Other 5.36 74.92 24.51 0.92 82.40
Cu-based 5.25 48.67 3.20 1.20 76.96
Fe-based 4.79 74.45 29.88 0.59 83.28
Mg-based 12.00 50.00 10.00 0.50 87.60
Other 4.83 45.50 30.80 0.88 91.74
TiO2- 5.63 101.67 22.77 1.14 73.95
based
Zn-based 4.50 140.00 10.00 2.10 99.00
Pesticide 6.91 56.67 22.59 0.83 79.72
Cu-based 7.00 180.00 30.00 0.40 87.23
Fe-based 5.00 60.00 35.15 0.40 82.00
Mg-based 10.00 15.00 10.00 1/50 82.43
Other 4.05 60.00 44.90 99.99
Zn-based 7.70 12.50 7.75 1.00 63.35
Grand Total 5.38 73.80 25.24 0.70 83.16

type of pollutant, catalyst, and other process conditions. The removal efficiency of pollutants such as ciprofloxacin and diazinon
decreases at very alkaline and acidic pH due to the neutrality of the surface charge of the pollutant and CuO nanoparticles. Studies have
shown that with the increase in contact time due to the increase in the production of free radicals and also providing enough time for
pollutant molecules to contact the active sites of the catalyst, the pollutant removal efficiency increases (Khoshnamvand et al., 2018;
Mohagheghian et al., 2022). The minimum and maximum amounts of pollutant degradation were respectively related to methylene
blue with 40 % degradation and tri-n-butyl phosphate (TNBP) with 99 % degradation. The study of Senobari and Nezamzadeh-Ejhieh
(2018) showed that the presence of chloride and bicarbonate anions in the aqueous solution reduces the degradation rate of methylene
blue due to the hydroxyl radical scavenging property (Senobari and Nezamzadeh-Ejhieh, 2018a). The study of Golmohammadi and
Sattari (2022) confirmed that the catalytic oxidation of tri-n-butyl phosphate (TNBP) can convert this substance into safe by-products
(Golmohammadi and Sattari, 2022). In general, the average removal efficiency of organic pollutants by Cu-based systems was 82.93 %
regardless of the type of pollutants.

3.2.5. Mg-based AOPs


Table 6 presents a summary of Mg-based processes for the removal of organic pollutants. As it is clear from Table 6, the efficiency of
the process varies based on the use of Mg nanoparticles along with other compounds to remove organic pollutants. The highest removal
efficiency (93.77 %) was linked with reactive blue 19 and the lowest (82.43 %) with diazinon pesticide. It seems that acidic pH and
higher initial pollutant concentration are required to remove dyes through Mg-based processes. The instability of color rings at low pH
and the electrostatic attractions between functional groups in Reactive Blue 19 (negatively charged) and the surface of MgO nano­
particles (positively charged) are the two main reasons for the decomposition of Reactive Blue 19 anionic dye by MgO nanoparticles at
acidic pH (Ahmadi et al., 2018). In the case of toluene and diazinon, the pollutant removal efficiency increases with increasing pH and
catalyst dose. In alkaline conditions, the production of hydroxyl radicals increases. On the other hand, with the increase of the catalyst
dose, the catalyst surface expands, and as a result, the possibility of pollutant contact with hydroxyl radicals and the catalyst surface
increases, leading to pollutant removal (Asgari et al., 2020a; Mohammadi et al., 2016). It is important to acknowledge, however, that
drawing conclusions about all dyes based on the findings of just one study is not feasible.

3.2.6. Other AOPs


Table 7 summarizes the effectiveness of other AOPs in removing various organic pollutants. Based on this table, it can be said that
nano-based processes can remove various organic pollutants. Moreover, according to the conditions of each process, the result ob­
tained would be different from other processes. Tetracycline drug removal has been studied within three processes: ESMNPs/NCB/
alginate (Soltani et al., 2021), BWO-GO photocatalyst (Song et al., 2016), and Bi4Ti3O12 nanoparticles (Khodadoost et al., 2017). The
results indicate that the highest removal efficiency (78.43 %) of this drug has been achieved through the BWO-GO photocatalytic
process under optimal conditions of reaction time of 90 min, initial pollutant concentration of 10 mg/L, and a catalyst dose of 0.3 g/L.

14
R.N. Nodehi and S. Sheikhi Environmental Technology & Innovation 35 (2024) 103718

Fig. 4. The panel plots of (a) pH, (b) time, (c) initial pollutant concentration and (d) catalyst dose.

Song et al. reported that ⋅O-2 is the main active species in the photodegradation of tetracycline (Song et al., 2016). The removal of
methylene blue pollutants has been studied using two categories of nanoparticles: NiO-CdS nanoparticles (Senobari and
Nezamzadeh-Ejhieh, 2018b) and CdSe nanoparticles (Gharbani et al., 2022). The results show that NiO-CdS nanoparticles can remove
85 % of methylene blue under optimal conditions of pH 3.5, reaction time 83 min, initial pollutant concentration 3.2 mg/L, and
catalyst dose 0.9 g/L. On the other hand, CdSe nanoparticles remove 92.8 % of methylene blue under optimal conditions of pH 8,
reaction time of 20 min, initial pollutant concentration of 20 mg/L, and a catalyst dose of 0.4 g/L. Senobari and Nezamzadeh-Ejhieh
reported that the pHpzc of the catalyst is 3.9, so at a pH of about 4, the catalyst surface has a net-zero charge, and methylene blue is
present in its neutral form. In this condition, the interaction of methylene blue molecules with the catalyst surface is high, and the

15
R.N. Nodehi and S. Sheikhi Environmental Technology & Innovation 35 (2024) 103718

photodegradation efficiency increases (Senobari and Nezamzadeh-Ejhieh, 2018b). On the other hand, Gharbani et al. emphasized that
the presence of hydroxyl ions at higher pH causes the formation of abundant hydroxyl radicals in the solution and ultimately leads to
an increase in the degradation of methylene blue in alkaline conditions (Gharbani et al., 2022). Therefore, it can be said that the
efficiency of removing organic pollutants depends on various parameters that can be changed to achieve optimal conditions. The
minimum and maximum levels of pollutant degradation were related to tetracycline drug with 15.8 % degradation by the
ESMNPs/NCB/alginate process (Soltani et al., 2021) and chlorpyrifos pesticide with 99.99 % degradation through the GNPs/ZrV2O7
process (Samy et al., 2020).

3.3. Analysis of different parameters according to the pollutants classification

To better analyze the studies from another point of view, the pollutants in the studies were categorized into four main groups:
drugs, dyes, pesticides, and others. Among these, the highest number of pollutants was related to the others category (n = 24) and the
lowest to the pesticide group (n = 6). Fig. 3 shows the box plot of pH, time, initial pollutant concentration, and catalyst dose according
to pollutant classification.
The average pH values (Fig. 3(a)) in the four groups of drugs, dyes, pesticides, and others were 5.51, 4.82, 6.91, and 5.36,
respectively. These results indicate that the pH required to remove pesticides by nano-based processes is higher than other pollutants.
Determining pHpzc for catalysts is very important because at a pH close to pHpzc, the interaction between pollutant molecules with the
catalyst surface is high, and at pHs lower or higher than pHpzc, the pollutant is repelled by the catalyst surface (Senobari and
Nezamzadeh-Ejhieh, 2018b). According to Fig. 3(b), the average reaction time related to dyes (88.70 min) is higher than that of the
other three groups. At the start of the process, the reaction speed is higher due to the large number of active sites on the catalyst surface.
Over time, the efficiency of the process is relatively high because the contact time between the pollutant and the catalyst increases. But
with the increase of time to more than the optimal value, the efficiency of the process decreases due to the saturation of the catalyst
surface and the low production of free radicals (Ciğeroğlu et al., 2022; Song et al., 2016; Wei et al., 2020). Likewise, the average initial
concentration of the pollutant related to dyes (27.66 mg/L) is higher than that of the other groups. However, the groups have no
significant difference (Fig. 3(c)). According to the collision theory, by increasing the initial concentration of the pollutant to a certain
level, the probability of its collision with other reacting molecules increases, leading to an improvement in the reaction speed (Sheikhi
et al., 2021b). By increasing the initial concentration of the pollutant to a level higher than the optimal value, the degradation effi­
ciency decreases due to the saturation of the catalyst surface and the low production of free radicals (Liu et al., 2023). The average
doses of catalyst (Fig. 3(d)) in the four groups of drugs, dyes, pesticides, and others were 0.56, 0.58, 0.83 and 0.92 g/L, respectively.
Therefore, it can be said that the average dose of the catalyst in the two groups of dyes and drugs is lower than in the other two groups.
The results of most studies have shown that the efficiency of pollutant destruction is improved by increasing the catalyst dose due to
the increased possibility of pollutant contact with the catalyst surface and free radicals (Asgari et al., 2020a; Biglar et al., 2021; Kaith
et al., 2018). The average removal efficiency values of pollutants through nanomaterial-based AOPs in the four mentioned groups were
78.52, 88.92, 79.72, and 82.40 %, respectively. Therefore, it can be concluded that the dye removal efficiency was relatively higher.
In all groups (in all four variables, including pH, residence time, initial pollutant concentration, and catalyst dose), the median is
lower than the mean, suggesting that the distribution of values is positively skewed (to the right) (Mirzaei et al., 2018b).

3.4. Statistical analysis

In order to create an overview of nano-based advanced oxidation processes, all collected data were examined in terms of descriptive
statistics (including mean, median, standard deviation, standard error, kurtosis, skewness, and confidence interval, among others).
Table 8 shows the descriptive statistics of the investigated variables irrespective of the classification of pollutants and nano-based
processes. According to this table, only the catalyst dosage variable has negative kurtosis among the four variables, as it indicates
lighter tails and a flatter distribution. On the other hand, the removal efficiency has negative skewness, which means that the longer
tail is on the left side of the distribution. In other words, the median (90.20 %) is greater than the average (83.16 %). Furthermore, the
time kurtosis coefficient is bigger than the coefficient of other variables. Therefore, it can be said that the probability of outliers is
higher in the time variable. Table 9 shows the average of the investigated variables considering the classification of pollutants and
nano-based processes (The relevant explanations are given in the text.). Tables S1–S8 display the regression model and analysis of
variance for each variable.

3.5. Comparison of different nanomaterial-based AOPs

Fig. 4 was plotted to better compare the processes used in this work for organic pollutant removal. Among the reviewed studies, the
largest number was related to Fe-based processes (n = 24) and the lowest to Mg-based processes (n = 3). Most pollutants removed by
the Fe-based process were drugs, and the most pollutants removed by the TiO2-based and Zn-based processes were dyes.
The average pH values required to remove organic pollutants, regardless of the type of pollutant, by Fe-based, TiO2-based, Zn-
based, Cu-based, and Mg-based processes were 4.93, 4.82, 5.56, 6.41, and 8.33, respectively. The dyes group has the longest reac­
tion time required for pollutant removal, followed by the Fe- and Zn-based groups. Also, the longest reaction time required to remove
the pollutant by the Cu-based processes belonged to the group of pesticides (Fig. 4(b)). In Fe-based, TiO2-based, and Mg-based

16
R.N. Nodehi and S. Sheikhi Environmental Technology & Innovation 35 (2024) 103718

processes, the highest initial pollutant concentration required for effective removal was related to the group of dyes. In Cu-based and
Zn-based processes, the highest initial pollutant concentration was related to the groups of pesticides and drugs, respectively (Fig. 4
(c)). The highest catalyst dose required for optimal removal by all Fe-based, TiO2-based, Zn-based, and Cu-based processes belonged to
the others group. In the Mg-based process, the highest catalyst dose was associated with pesticides (Fig. 4(d)). It is worth mentioning
that depending on the other conditions applied in the process, each of the mentioned variables can accept different values.
The average removal efficiency values of pollutants by Fe-based, TiO2-based, Zn-based, Cu-based, and Mg-based processes,
regardless of the type of pollutant, were 87.53 %, 82.61 %, 80.16 %, 82.93 %, and 87.93 %, respectively. Fe-based, TiO2-based, and
Mg-based processes had the highest removal efficiency for the dyes group. Cu-based and Zn-based processes also had the highest
removal efficiency for drugs and the others group, respectively.

4. Current challenges and future perspectives

The unique properties of nanomaterials and their convergence with AOP processes have led the current research trend to the
combined use of nanomaterials and AOP processes in water and wastewater treatment. Because many studies are limited to laboratory
and experimental research and only simulate real exposure conditions, one of the important challenges in the use of nanomaterials is to
prepare them in high quantities at a reasonable price for use in the real environment. Another challenge is the possibility of using
hybrid processes such as AOP and nanoparticles in existing treatment systems, especially in developing countries. The reaction pa­
rameters that affect the behavior of nanomaterials and the removal of pollutants should be analyzed. On the other hand, the stability,
toxicity and unknown environmental effects of nanoparticles are considered key challenges that require more research in this field.
Despite the significant progress in the production of nanomaterials, it is suggested that future investigations focus on finding
multifunctional nanomaterials and improving their catalytic efficiency. There may be future prospects for optimizing the effectiveness
of factors to achieve maximum efficiency, leading to more free radicals. This review provides a valuable reference for future research
on water and wastewater treatment.

5. Conclusion

This study presents a review of the application of nano-based AOPs in removing different organic pollutants under optimal con­
ditions. In order to facilitate process comparison, data on four important parameters (pH, reaction time, initial pollutant concentration,
and catalyst dose) from various nano-based AOPs investigated in the literature were collected and critically compared. Despite the high
variability between the results of the individual processes, similarities and differences were observed between AOPs. Fe-based, TiO2-
based, and Mg-based processes had the highest removal efficiency for the dyes group. Cu-based and Zn-based processes had the highest
removal efficiency for drugs and the others group, respectively. The average removal efficiency values of pollutants by Fe-based, TiO2-
based, Zn-based, Cu-based, and Mg-based processes, regardless of the type of pollutant, were 87.53 %, 82.61 %, 80.16 %, 82.93 %, and
87.93 %, respectively. Generally, nanomaterial-based AOPs are recommended as an effective and suitable method for removing
organic pollutants from aqueous environments due to their higher removal efficiency.

Authors’ contribution

All authors contributed to the study conception and design. Material preparation, data collection, and analysis were performed by
Ramin Nabizadeh Nodehi and Samira Sheikhi. Also, all authors reviewed the results and approved the final version of the manuscript.

CRediT authorship contribution statement

Samira Sheikhi: Writing – review & editing, Writing – original draft, Visualization, Validation, Supervision, Software, Resources,
Project administration, Methodology, Investigation, Funding acquisition, Formal analysis, Data curation, Conceptualization. Ramin
Nabizadeh Nodehi: Writing – review & editing, Writing – original draft, Visualization, Validation, Supervision, Software, Resources,
Project administration, Methodology, Investigation, Funding acquisition, Formal analysis, Data curation, Conceptualization.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Data availability

No data was used for the research described in the article.

17
R.N. Nodehi and S. Sheikhi Environmental Technology & Innovation 35 (2024) 103718

Acknowledgments

The authors would like to thank the Tehran University of Medical Sciences (Iran), for supporting this study. This research did not
receive any specific grant from funding agencies in the public, commercial, or not-for-profit sectors.

Appendix A. Supporting information

Supplementary data associated with this article can be found in the online version at doi:10.1016/j.eti.2024.103718.

References

Aeindartehran, L., Talesh, S.S.A., 2021. Enhanced photocatalytic degradation of Acid Blue 1 using Ni-Decorated ZnO NPs synthesized by sol-gel method: RSM
optimization approach. Ceram. Int. 47, 27294–27304. https://doi.org/10.1016/j.ceramint.2021.06.151.
Aghaeinejad-Meybodi, A., Ebadi, A., Shafiei, S., Khataee, A., Kiadehi, A.D., 2019. Degradation of Fluoxetine using catalytic ozonation in aqueous media in the
presence of nano-γ-alumina catalyst: experimental, modeling and optimization study. Sep. Purif. Technol. 211, 551–563. https://doi.org/10.1016/j.
seppur.2018.10.020.
Ahmadi, M., Rahmani, K., Rahmani, A., Rahmani, H., 2017. Removal of benzotriazole by Photo-Fenton like process using nano zero-valent iron: response surface
methodology with a Box-Behnken design. Pol. J. Chem. Technol. 19, 104–112. https://doi.org/10.1515/pjct-2017-0015.
Ahmadi, S., Mohammadi, L., Igwegbe, C.A., Rahdar, S., Banach, A.M., 2018. Application of response surface methodology in the degradation of Reactive Blue 19 using
H2O2/MgO nanoparticles advanced oxidation process. Int. J. Ind. Chem. 9, 241–253. https://doi.org/10.1007/s40090-018-0153-4.
Airemlou, L., Behnajady, M.A., Mahanpoor, K., 2019. Photocatalytic removal of RhB by ag and MG Co-doped ZnO nanoparticles: modeling of operational parameters
using ANN based on RSM data. Russ. J. Phys. Chem. A 93, 1769–1777. https://doi.org/10.1134/S0036024419090097.
Ansari, M., Mahvi, A.H., Salmani, M.H., Sharifian, M., Fallahzadeh, H., Ehrampoush, M.H., 2020. Dielectric barrier discharge plasma combined with nano catalyst for
aqueous amoxicillin removal: performance modeling, kinetics and optimization study, energy yield, degradation pathway, and toxicity. Sep. Purif. Technol. 251,
117270 https://doi.org/10.1016/j.seppur.2020.117270.
Arslan-Alaton, I., Tureli, G., Olmez-Hanci, T., 2009. Treatment of azo dye production wastewaters using Photo-Fenton-like advanced oxidation processes:
optimization by response surface methodology. J. Photochem. Photobiol. A 202, 142–153. https://doi.org/10.1016/j.jphotochem.2008.11.019.
Asgari, G., Seidmohammadi, A., Esrafili, A., Faradmal, J., Sepehr, M.N., Jafarinia, M., 2020a. The catalytic ozonation of diazinon using nano-MgO@ CNT@ Gr as a
new heterogenous catalyst: the optimization of effective factors by response surface methodology. RSC Adv. 10, 7718–7731. https://doi.org/10.1039/
C9RA10095D.
Asgari, G., Shabanloo, A., Salari, M., Eslami, F., 2020b. Sonophotocatalytic treatment of AB113 dye and real textile wastewater using ZnO/persulfate: modeling by
response surface methodology and artificial neural network. Environ. Res. 184, 109367 https://doi.org/10.1016/j.envres.2020.109367.
Bakhtkhosh, P., Mehrizad, A., 2017. Sonochemical synthesis of Sm-doped ZnS nanoparticles for photocatalytic degradation of Direct Blue 14: experimental design by
response surface methodology and development of a kinetics model. J. Mol. Liq. 240, 65–73. https://doi.org/10.1016/j.molliq.2017.05.053.
Belachew, N., Fekadu, R., Ayalew Abebe, A., 2020. RSM-BBD optimization of Fenton-like degradation of 4-nitrophenol using magnetite impregnated kaolin. Air Soil
Water Res. 13, 1178622120932124 https://doi.org/10.1177/1178622120932124.
Bethi, B., Sonawane, S.H., Bhanvase, B.A., Gumfekar, S.P., 2016. Nanomaterials-based advanced oxidation processes for wastewater treatment: a review. Chem. Eng.
Process. 109, 178–189. https://doi.org/10.1016/j.cep.2016.08.016.
Biglar, F., Talaiekhozani, A., Aminsharei, F., Park, J., Barghi, A., Rezania, S., 2021. Application of ZnO-Nd nano-photocatalyst for the reactive red 198 dye
decolorization in the falling-film photocatalytic reactor. Toxics 9, 254. https://doi.org/10.3390/toxics9100254.
Bose, S., Tripathy, B.K., Debnath, A., Kumar, M., 2021. Boosted sono-oxidative catalytic degradation of Brilliant green dye by magnetic MgFe2O4 catalyst:
degradation mechanism, assessment of bio-toxicity and cost analysis. Ultrason. Sonochem. 75, 105592 https://doi.org/10.1016/j.ultsonch.2021.105592.
Ciğeroğlu, Z., Şahin, S., Kazan, E.S., 2022. One-pot green preparation of deep eutectic solvent-assisted ZnO/GO nanocomposite for cefixime trihydrate photocatalytic
degradation under UV-A irradiation. Biomass Convers. Bioref. 12, 73–86. https://doi.org/10.1007/s13399-021-01734-0.
Desai, N.N., Soraganvi, V.S., Madabhavi, V.K., 2020. Solar photocatalytic degradation of organic contaminants in landfill leachate using TiO2 nanoparticles by RSM
and ANN. Nat. Environ. Pollut. Technol. 19, 651–662. https://doi.org/10.46488/NEPT.2020.v19i02.019.
Doria, F.C., Borges, A., Kim, J., Nathan, A., Joo, J., Campos, L., 2013. Removal of metaldehyde through photocatalytic reactions using nano-sized zinc oxide
composites. Water Air Soil Pollut. 224, 1–9. https://doi.org/10.1007/s11270-013-1434-3.
Doumbi, R.T., Noumi, G.B., 2021. Dip coating deposition of manganese oxide nanoparticles on graphite by sol gel technique for the indirect electrochemical oxidation
of methyl orange dye: parameter’s optimization using box-behnken design. Case Stud. Chem. Environ. Eng. 3, 100068 https://doi.org/10.1016/j.
cscee.2020.100068.
Ershadi Afshar, L., Chaibakhsh, N., Moradi-Shoeili, Z., 2018. Treatment of wastewater containing cytotoxic drugs by CoFe2O4 nanoparticles in Fenton/ozone
oxidation process. Sep. Sci. Technol. 53, 2671–2682. https://doi.org/10.1080/01496395.2018.1461113.
Eskandarimakvand, M., Sabzalipour, S., Cheraghi, M., Orak, N., 2022. Evaluation of efficiency of iron oxide nanoparticles (Fe3O4@ CNT) in removal of malathion in
aqueous medium using response surface methodology (RSM). Pollution 8, 281–293. 〈https://doi.org/10.22059/POLL.2021.323942.1090〉.
Galedari, M., Ghazi, M.M., Mirmasoomi, S.R., 2019. Photocatalytic process for the tetracycline removal under visible light: Presenting a degradation model and
optimization using response surface methodology (RSM). Chem. Eng. Res. Des. 145, 323–333. https://doi.org/10.1016/j.cherd.2019.03.031.
Gharbani, P., Mehrizad, A., Mosavi, S.A., 2022. Optimization, kinetics and thermodynamics studies for photocatalytic degradation of Methylene Blue using cadmium
selenide nanoparticles. npj Clean. Water 5, 34. https://doi.org/10.1038/s41545-022-00178-x.
Gholipoor, O., Hosseini, S.A., 2021. Phenol removal from wastewater by CWPO process over the Cu-MOF nanocatalyst: process modeling by response surface
methodology (RSM) and kinetic and isothermal studies. New J. Chem. 45, 2536–2549. https://doi.org/10.1039/D0NJ04128A.
Golmohammadi, M., Sattari, M., 2022. Catalytic supercritical water oxidation of tri-n-butyl phosphate: process optimization by response surface methodology and
cytotoxicity assessment. Ceram. Int. 48, 36401–36409. https://doi.org/10.1016/j.ceramint.2022.08.200.
Hassani, G., Takdastan, A., Ghaedi, M., Goudarzi, G., Neisi, A., Babaei, A.A., 2016. Optimization of 4-chlorophenol oxidation by manganese ferrite nanocatalyst with
response surface methodology. Int. J. Electrochem. Sci. 11, 8471–8485. https://doi.org/10.20964/2016.10.36.
Hossain, S.S., Tarek, M., Munusamy, T.D., Karim, K.M.R., Roopan, S.M., Sarkar, S.M., Cheng, C.K., Khan, M.M.R., 2020. Facile synthesis of CuO/CdS heterostructure
photocatalyst for the effective degradation of dye under visible light. Environ. Res. 188, 109803 https://doi.org/10.1016/j.envres.2020.109803.
Hosseini, S.A., Babaei, S., 2017. Graphene oxide/zinc oxide (GO/ZnO) nanocomposite as a superior photocatalyst for degradation of methylene blue (MB)-process
modeling by response surface methodology (RSM). J. Braz. Chem. Soc. 28, 299–307. https://doi.org/10.5935/0103-5053.20160176.
Iazdani, F., Nezamzadeh-Ejhieh, A., 2021. Supported cuprous oxide-clinoptilolite nanoparticles: brief identification and the catalytic kinetics in the photodegradation
of dichloroaniline. Spectrochim. Acta Part A: Spectrochim. Acta A Mol. Biomol. Spectrosc. 250, 119348 https://doi.org/10.1016/j.saa.2020.119348.
Ighalo, J.O., Adeniyi, A.G., Adeniran, J.A., Ogunniyi, S., 2021. A systematic literature analysis of the nature and regional distribution of water pollution sources in
Nigeria. J. Clean. Prod. 283, 124566 https://doi.org/10.1016/j.jclepro.2020.124566.

18
R.N. Nodehi and S. Sheikhi Environmental Technology & Innovation 35 (2024) 103718

Jebelli, M.A., Maleki, A., Amoozegar, M.A., Kalantar, E., Gharibi, F., Darvish, N., Tashayoe, H., 2018. Isolation and identification of the native population bacteria for
bioremediation of high levels of arsenic from water resources. Environ. Manag. 212, 39–45. https://doi.org/10.1016/j.jenvman.2018.01.075.
Joshaghani, M., Yazdani, D., Zinatizadeh, A.A., 2017. Statistical modeling of p-nitrophenol degradation using a response surface methodology (RSM) over nano zero-
valent iron-modified Degussa P25-TiO 2/ZnO photocatalyst with persulfate. J. Iran. Chem. Soc. 14, 2449–2456. https://doi.org/10.1007/s13738-017-1179-9.
Kaith, B.S., Shanker, U., Gupta, B., Bhatia, J.K., 2018. RSM-CCD optimized In-air synthesis of photocatalytic nanocomposite: application in removal-degradation of
toxic brilliant blue. React. Funct. Polym. 131, 107–122. https://doi.org/10.1016/j.reactfunctpolym.2018.07.016.
Karimi, P., Baneshi, M.M., Malakootian, M., 2019. Photocatalytic degradation of aspirin from aqueous solutions using the UV/ZnO process: modelling, analysis and
optimization by response surface methodology (RSM). Desalin. Water Treat. 161, 354–364. https://doi.org/10.5004/dwt.2019.24317.
Karimi, P., Malakootian, M., 2020. Optimization of photocatalytic degradation of naproxen from aqueous solutions with UV/ZnO process: response surface
methodology (RSM). Glob. Nest J. 22, 369–380.
Khammar, S., Bahramifar, N., Younesi, H., 2020. Preparation and surface engineering of CM-β-CD functionalized Fe3O4@TiO2 nanoparticles for photocatalytic
degradation of polychlorinated biphenyls (PCBs) from transformer oil. J. Hazard. Mater. 394, 122422 https://doi.org/10.1016/j.jhazmat.2020.122422.
Khataee, A., Fathinia, M., Aber, S., Zarei, M., 2010. Optimization of photocatalytic treatment of dye solution on supported TiO2 nanoparticles by central composite
design: intermediates identification. J. Hazard. Mater. 181, 886–897. https://doi.org/10.1016/j.jhazmat.2010.05.096.
Khodadoost, S., Hadi, A., Karimi-Sabet, J., Mehdipourghazi, M., Golzary, A., 2017. Optimization of hydrothermal synthesis of Bismuth titanate nanoparticles and
application for photocatalytic degradation of tetracycline. J. Environ. Chem. Eng. 5, 5369–5380. https://doi.org/10.1016/j.jece.2017.10.006.
Khoshnamvand, N., Kord Mostafapour, F., Mohammadi, A., Faraji, M., 2018. Response surface methodology (RSM) modeling to improve removal of ciprofloxacin
from aqueous solutions in photocatalytic process using copper oxide nanoparticles (CuO/UV). Amb Express 8 (1), 9. https://doi.org/10.1186/s13568-018-0579-
2.
Kurian, M., 2021. Advanced oxidation processes and nanomaterials–a review. Clean. Eng. Technol. 2, 100090 https://doi.org/10.1016/j.clet.2021.100090.
Li, X., Chen, W., Ma, L., Wang, H., Fan, J., 2018. Industrial wastewater advanced treatment via catalytic ozonation with an Fe-based catalyst. Chemosphere 195,
336–343. https://doi.org/10.1016/j.chemosphere.2017.12.080.
Liu, F., Yi, P., Wang, X., Gao, H., Zhang, H., 2018. Degradation of Acid Orange 7 by an ultrasound/ZnO-GAC/persulfate process. Sep. Purif. Technol. 194, 181–187.
https://doi.org/10.1016/j.seppur.2017.10.072.
Liu, G., Zhang, X., Liu, H., He, Z., Show, P.L., Vasseghian, Y., Wang, C., 2023. Biochar/layered double hydroxides composites as catalysts for treatment of organic
wastewater by advanced oxidation processes: a review. Environ. Res., 116534 https://doi.org/10.1016/j.envres.2023.116534.
Lotfi, H., Nademi, M., Mansouri, M., Olya, M.E., 2016. Employing response surface analysis for photocatalytic degradation of MTBE by nanoparticles. Adv. Environ.
Sci. Technol. 3, 127–135.
Ma, D., Yi, H., Lai, C., Liu, X., Huo, X., An, Z., Li, L., Fu, Y., Li, B., Zhang, M., 2021a. Critical review of advanced oxidation processes in organic wastewater treatment.
Chemosphere 275, 130104. https://doi.org/10.1016/j.chemosphere.2021.130104.
Ma, Y., Lv, X., Xiong, D., Zhao, X., Zhang, Z., 2021b. Catalytic degradation of ranitidine using novel magnetic Ti3C2-based MXene nanosheets modified with nanoscale
zero-valent iron particles. Appl. Catal. B 284, 119720. https://doi.org/10.1016/j.apcatb.2020.119720.
Ma, Y., Xiong, D., Lv, X., Zhao, X., Meng, C., Xie, H., Zhang, Z., 2021c. Rapid and long-lasting acceleration of zero-valent iron nanoparticles@Ti3C2-based MXene/
peroxymonosulfate oxidation with bi-active centers toward ranitidine removal. J. Mater. Chem. A 9, 19817–19833. https://doi.org/10.1039/D1TA02046C.
Mahmoudabadi, Z.S., Rashidi, A., Maklavany, D.M., 2022. Optimizing treatment of alcohol vinasse using a combination of advanced oxidation with porous α-Fe2O3
nanoparticles and coagulation-flocculation. Ecotoxicol. Environ. Saf. 234, 113354 https://doi.org/10.1016/j.ecoenv.2022.113354.
Mansouri, M., Nademi, M., Ebrahim Olya, M., Lotfi, H., 2017a. Study of methyl tert-butyl ether (MTBE) photocatalytic degradation with UV/TiO2-ZnO-CuO
nanoparticles. J. Chem. Health Risks 7, 19–32.
Mansouri, M., Tanzifi, M., Lotfi, H., Nademi, M., 2017b. Investigation of UV/TiO2-ZnO-Co photocatalitic degradation of azo dye (reactive red 120) by response
surface methodology. Sci. Study Res. Chem. Chem. Eng. Biotechnol. Food Ind. 18, 153.
Mansouri, M., Yari, H., Kikhavani, T., Setareshenas, N., 2022. UVA/TiO2–ZnO–NiO photocatalytic oxidation process of dye: optimization and CFD simulation. Arab. J.
Sci. Eng. 1–14. https://doi.org/10.1007/s13369-021-05733-1.
Massoudinejad, M., Sadani, M., Gholami, Z., Rahmati, Z., Javaheri, M., Keramati, H., Sarafraz, M., Avazpour, M., Shiri, S., 2019. Optimization and modeling of
photocatalytic degradation of Direct Blue 71 from contaminated water by TiO2 nanoparticles: response surface methodology approach (RSM). Iran. J. Catal. 9,
121–132.
Mirzaei, A., Yerushalmi, L., Chen, Z., Haghighat, F., 2018a. Photocatalytic degradation of sulfamethoxazole by hierarchical magnetic ZnO@ g-C3N4: RSM
optimization, kinetic study, reaction pathway and toxicity evaluation. J. Hazard. Mater. 359, 516–526. https://doi.org/10.1016/j.jhazmat.2018.07.077.
Mirzaei, M., Sabbaghi, S., Zerafat, M.M., 2018b. Photo-catalytic degradation of formaldehyde using nitrogen-doped TiO2 nano-photocatalyst: statistical design with
response surface methodology (RSM). Can. J. Chem. Eng. 96, 2544–2552. https://doi.org/10.1002/cjce.23192.
Mohagheghian, A., Besharati-Givi, N., Ayagh, K., Shirzad-Siboni, M., 2022. Mineralization of diazinon by low-cost CuO-Kaolin nanocomposite under visible light
based RSM methodology: kinetics, cost analysis, reaction pathway and bioassay. J. Ind. Eng. Chem. 116, 276–292. https://doi.org/10.1016/j.jiec.2022.09.018.
Mohammadi, L., Bazrafshan, E., Noroozifar, M., Ansari-Moghaddam, A., 2016. Application of heterogeneous catalytic ozonation process with magnesium oxide
nanoparticles for Toluene degradation in aqueous environments. Health Scope 5. https://doi.org/10.17795/jhealthscope-40439.
Mondal, S.K., Saha, A.K., Sinha, A., 2018. Removal of ciprofloxacin using modified advanced oxidation processes: kinetics, pathways and process optimization.
J. Clean. Prod. 171, 1203–1214. https://doi.org/10.1016/j.jclepro.2017.10.091.
Moradi, M., Ghanbari, F., Manshouri, M., Angali, K.A., 2016. Photocatalytic degradation of azo dye using nano-ZrO2/UV/Persulfate: response surface modeling and
optimization. Korean J. Chem. Eng. 33, 539–546. https://doi.org/10.1007/s11814-015-0160-5.
Mukhopadhyay, A., Tripathy, B.K., Debnath, A., Kumar, M., 2021. Enhanced persulfate activated sono-catalytic degradation of brilliant green dye by magnetic
CaFe2O4 nanoparticles: degradation pathway study, assessment of bio-toxicity and cost analysis. Surf. Interfaces 26, 101412. https://doi.org/10.1016/j.
surfin.2021.101412.
Naimi-Joubani, M., Ayagh, K., Tahergorabi, M., Shirzad-Siboni, M., Yang, J.-K., 2023. Design and modeling of diazinon degradation in hydrous matrix by Ni-doped
ZnO nanorods under ultrasonic irradiation: process optimization using RSM (CCD), kinetic study, reaction pathway, mineralization, and toxicity assessment.
Environ. Sci. Pollut. Res. Int. 30, 3527–3548. https://doi.org/10.1007/s11356-022-21861-z.
Omrani, N., Nezamzadeh-Ejhieh, A., 2020. BiVO4/WO3 nano-composite: characterization and designing the experiments in photodegradation of sulfasalazine.
Environ. Sci. Pollut. Res. Int. 27, 44292–44305. https://doi.org/10.1007/s11356-020-10278-1.
Pirsaheb, M., Moradi, S., Shahlaei, M., Farhadian, N., 2018. Application of carbon dots as efficient catalyst for the green oxidation of phenol: kinetic study of the
degradation and optimization using response surface methodology. J. Hazard. Mater. 353, 444–453. https://doi.org/10.1016/j.jhazmat.2018.04.038.
Popescu, R.C., Andronescu, E., Vasile, B.S., 2019. Recent advances in magnetite nanoparticle functionalization for nanomedicine. Nanomaterials 9, 1791. https://doi.
org/10.3390/nano9121791.
Pourali, P., Behzad, A., Ahmadfazeli, A., Mokhtari, S.A., Rashtbari, Y., Poureshgh, Y., 2022. Dissociation of acid blue 113 dye from aqueous solutions using activated
persulfate by zero iron nanoparticle from green synthesis: the optimization process with RSM-BBD model: mineralization and reaction kinetic study. Biomass
Convers. Bioref. 1–13. https://doi.org/10.1007/s13399-022-02942-y.
Rafiee, E., Noori, E., Zinatizadeh, A.A., Zangeneh, H., 2018. Surfactant effect on photocatalytic activity of Ag-TiO2/PW nanocomposite in DR16 degradation:
characterization of nanocomposite and RSM process optimization. Mater. Sci. Semicond. Process. 83, 115–124. https://doi.org/10.1016/j.mssp.2018.04.021.
Rahimi, B., Jafari, N., Abdolahnejad, A., Farrokhzadeh, H., Ebrahimi, A., 2019. Application of efficient photocatalytic process using a novel BiVO/TiO2-NaY zeolite
composite for removal of acid orange 10 dye in aqueous solutions: modeling by response surface methodology (RSM). J. Environ. Chem. Eng. 7, 103253 https://
doi.org/10.1016/j.jece.2019.103253.
Rahmani, A., Almasi, H., Bajalan, S., Rezaei Vahidian, H., Zarei, A., Shabanloo, A., 2017. Optimization of ciprofloxacin antibiotic sonochemical degradation with
persulfate activated by nano zero-valent iron by central composite design method. J. Health 8, 231–245.

19
R.N. Nodehi and S. Sheikhi Environmental Technology & Innovation 35 (2024) 103718

Safari, M., Rostami, M.H., Alizadeh, M., Alizadehbirjandi, A., Nakhli, S.A.A., Aminzadeh, R., 2014. Response surface analysis of photocatalytic degradation of methyl
tert-butyl ether by core/shell Fe3O4/ZnO nanoparticles. J. Environ. Health Sci. 12 (1), 10. https://doi.org/10.1186/2052-336X-12-1.
Samy, M., Ibrahim, M.G., Alalm, M.G., Fujii, M., Diab, K.E., Elkady, M., 2020. Innovative photocatalytic reactor for the degradation of chlorpyrifos using a coated
composite of ZrV2O7 and graphene nano-platelets. J. Chem. Eng. 395, 124974 https://doi.org/10.1016/j.cej.2020.124974.
Senobari, S., Nezamzadeh-Ejhieh, A., 2018a. A comprehensive study on the enhanced photocatalytic activity of CuO-NiO nanoparticles: designing the experiments.
J. Mol. Liq. 261, 208–217. https://doi.org/10.1016/j.molliq.2018.04.028.
Senobari, S., Nezamzadeh-Ejhieh, A., 2018b. A pn junction NiO-CdS nanoparticles with enhanced photocatalytic activity: a response surface methodology study.
J. Mol. Liq. 257, 173–183. https://doi.org/10.1016/j.molliq.2018.02.096.
Shabanloo, A., Salari, M., Shabanloo, N., Dehghani, M.H., Pittman Jr, C.U., Mohan, D., 2020. Heterogeneous persulfate activation by nano-sized Mn3O4 to degrade
furfural from wastewater. J. Mol. Liq. 298, 112088 https://doi.org/10.1016/j.molliq.2019.112088.
Shahbazi, R., Payan, A., Fattahi, M., 2018. Preparation, evaluations and operating conditions optimization of nano TiO2 over graphene based materials as the
photocatalyst for degradation of phenol. J. Photochem. Photobiol. A 364, 564–576. https://doi.org/10.1016/j.jphotochem.2018.05.032.
Shariati, M., Babaei, A., Azizi, A., 2022. Synthesis of Ag2CrO4/Ag/Fe3O4/RGO nanocomposite as a suitable photocatalyst for degradation of methylene blue in
aqueous media: RSM modeling, kinetic and energy consumption studies. Inorg. Chem. Commun. 145, 110004 https://doi.org/10.1016/j.inoche.2022.110004.
Shaykhi, Z., Zinatizadeh, A., 2014. Statistical modeling of photocatalytic degradation of synthetic amoxicillin wastewater (SAW) in an immobilized TiO2
photocatalytic reactor using response surface methodology (RSM). J. Taiwan Inst. Chem. Eng. 45, 1717–1726. https://doi.org/10.1016/j.jtice.2013.12.024.
Sheikhi, S., Dehghanzadeh, R., Aslani, H., 2021a. Advanced oxidation processes for chlorpyrifos removal from aqueous solution: a systematic review. J. Taiwan Inst.
Chem. Eng. 19, 1249–1262. https://doi.org/10.1007/s40201-021-00674-1.
Sheikhi, S., Dehghanzadeh, R., Maryamabadi, A., Aslani, H., 2021b. Chlorpyrifos removal from aqueous solution through sequential use of coagulation and advanced
oxidation processes: by-products, degradation pathways, and toxicity assessment. Environ. Technol. Innov. 23, 101564 https://doi.org/10.1016/j.
eti.2021.101564.
Shiri, S., Rabori, M.M., Gholami, Z., Rahmati, Z., Adiban, M., Sarafraz, M., 2020. Enhanced degradation of reactive black 5 from aqueous solution over tio2
nanoparticles under uv light irradiation: optimization, experimental & theoretical approaches. J. Environ. Treat. Tech. 8, 1232–1241. https://doi.org/10.47277/
8(3)1241.
Soltani, R.D.C., Naderi, M., Boczkaj, G., Jorfi, S., Khataee, A., 2021. Hybrid metal and non-metal activation of Oxone by magnetite nanostructures co-immobilized
with nano-carbon black to degrade tetracycline: fenton and electrochemical enhancement with bio-assay. Sep. Purif. Technol. 274, 119055 https://doi.org/
10.1016/j.seppur.2021.119055.
Song, C., Li, X., Wang, L., Shi, W., 2016. Fabrication, characterization and response surface method (RSM) optimization for tetracycline photodegration by Bi3. 84W0.
16O6. 24-graphene oxide (BWO-GO). Sci. Rep. 6, 37466 https://doi.org/10.1038/srep37466.
Soufi, A., Hajjaoui, H., Elmoubarki, R., Abdennouri, M., Qourzal, S., Barka, N., 2022. Heterogeneous Fenton-like degradation of tartrazine using CuFe2O4
nanoparticles synthesized by sol-gel combustion. Appl. Surf. Sci. 9, 100251 https://doi.org/10.1016/j.apsadv.2022.100251.
Sun, S.-P., Zeng, X., Lemley, A.T., 2013. Nano-magnetite catalyzed heterogeneous Fenton-like degradation of emerging contaminants carbamazepine and ibuprofen in
aqueous suspensions and montmorillonite clay slurries at neutral pH. J. Mol. Catal. A Chem. 371, 94–103. https://doi.org/10.1016/j.molcata.2013.01.027.
Tajyani, S., Babaei, A., 2018. A new sensing platform based on magnetic Fe3O4@ NiO core/shell nanoparticles modified carbon paste electrode for simultaneous
voltammetric determination of Quercetin and Tryptophan. J. Electroanal. Chem. 808, 50–58. https://doi.org/10.1016/j.jelechem.2017.11.010.
Tayeb, A.M., Tony, M.A., Mansour, S.A., 2018. Application of Box–Behnken factorial design for parameters optimization of basic dye removal using nano-hematite
photo-Fenton tool. Appl. Water Sci. 8, 1–9. https://doi.org/10.1007/s13201-018-0783-x.
Tony, M.A., Mansour, S.A., Tayeb, A.M., Purcell, P.J., 2018. Use of a fenton-like process based on nano-haematite to treat synthetic wastewater contaminated by
phenol: process investigation and statistical optimization. Arab. J. Sci. Eng. 43, 2227–2235. https://doi.org/10.1007/s13369-017-2632-x.
Vaez, M., Zarringhalam Moghaddam, A., Alijani, S., 2012. Optimization and modeling of photocatalytic degradation of azo dye using a response surface methodology
(RSM) based on the central composite design with immobilized titania nanoparticles. Ind. Eng. Chem. Res. 51, 4199–4207. https://doi.org/10.1021/ie202809w.
Villegas, V.A.R., Ramírez, J.I.D.L., Guevara, E.H., Sicairos, S.P., Ayala, L.A.H., Sanchez, B.L., 2020. Synthesis and characterization of magnetite nanoparticles for
photocatalysis of nitrobenzene. J. Saudi Chem. Soc. 24, 223–235. https://doi.org/10.1016/j.jscs.2019.12.004.
Wang, J., Wang, S., 2020. Reactive species in advanced oxidation processes: formation, identification and reaction mechanism. J. Chem. Eng. 401, 126158 https://
doi.org/10.1016/j.cej.2020.126158.
Wang, L., Luo, D., Yang, J., Wang, C., 2022. Metal-organic frameworks-derived catalysts for contaminant degradation in persulfate-based advanced oxidation
processes. J. Clean. Prod. 375, 134118 https://doi.org/10.1016/j.jclepro.2022.134118.
Wei, X., Xu, X., Yang, X., Li, J., Liu, Z., 2020. Visible light degradation of reactive black-42 by novel Sr/Ag-TiO2@ g-C3N4 photocatalyst: RSM optimization, reaction
kinetics and pathways. Spectrochim. Acta A Mol. Biomol. Spectrosc. 228, 117870 https://doi.org/10.1016/j.saa.2019.117870.
Wu, Q., Miao, W.-s, Zhang, Y.-d, Gao, H.-j, Hui, D., 2020. Mechanical properties of nanomaterials: a review. Nanotechnol. Rev. 9, 259–273. https://doi.org/10.1515/
ntrev-2020-0021.
Yashni, G., Al-Gheethi, A., Mohamed, R., Arifin, S.N.H., Salleh, S.N.A.M., 2020. Photodegradation of basic red 51 in hair dye greywater by zinc oxide nanoparticles
using central composite design. React. Kinet. Mech. Catal. 130, 567–588. https://doi.org/10.1007/s11144-020-01792-x.
Yu, S., Wang, X., Pang, H., Zhang, R., Song, W., Fu, D., Hayat, T., Wang, X., 2018. Boron nitride-based materials for the removal of pollutants from aqueous solutions:
a review. J. Chem. Eng. 333, 343–360. https://doi.org/10.1016/j.cej.2017.09.163.
Zhang, S., Wang, J., Zhang, Y., Ma, J., Huang, L., Yu, S., Chen, L., Song, G., Qiu, M., Wang, X., 2021. Applications of water-stable metal-organic frameworks in the
removal of water pollutants: a review. Environ. Pollut. 291, 118076 https://doi.org/10.1016/j.envpol.2021.118076.
Zhang, T., Hu, C., Li, Q., Chen, C., Hu, J., Xiao, X., Li, M., Zou, X., Huang, L., 2022. Hydrogen peroxide activated by biochar-supported sulfidated nano zerovalent iron
for removal of sulfamethazine: response surface method approach. Int. J. Environ. Res. Public Health 19, 9923. https://doi.org/10.3390/ijerph19169923.

20

You might also like