Next Article in Journal
A Novel Strategy to Enhance the Photostability of InP/ZnSe/ZnS Quantum Dots with Zr Doping
Next Article in Special Issue
Study of the Oxidation Behavior of Fine-Grained Graphite ET-10 by Combining X-ray μCT with Mercury Porosimetry
Previous Article in Journal
Causes and Consequences of Ordering and Dynamic Phases of Confined Vortex Rows in Superconducting Nanostripes
Previous Article in Special Issue
Obtaining N-Enriched Mesoporous Carbon-Based by Means of Gamma Radiation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Carbon-Based Materials for Adsorptive and Photocatalytic Antibiotic Removal

State Key Laboratory of Pollution Control and Resources Reuse, College of Environmental Science and Engineering, Tongji University, Shanghai 200092, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(22), 4045; https://doi.org/10.3390/nano12224045
Submission received: 27 October 2022 / Revised: 15 November 2022 / Accepted: 16 November 2022 / Published: 17 November 2022
(This article belongs to the Special Issue Porous Carbon Nanomaterials)

Abstract

:
Antibiotics have been a primary environmental concern due to their widespread dispersion, harmful bioaccumulation, and resistance to mineralization. Unfortunately, typical processes in wastewater treatment plants are insufficient for complete antibiotic removal, and their derivatives in effluent can pose a threat to human health and aquatic communities. Adsorption and photocatalysis are proven to be the most commonly used and promising tertiary treatment methods. Carbon-based materials, especially those based on graphene, carbon nanotube, biochar, and hierarchical porous carbon, have attracted much attention in antibiotic removal as green adsorbents and photocatalysts because of their availability, unique pore structures, and superior physicochemical properties. This review provides an overview of the characteristics of the four most commonly used carbonaceous materials and their applications in antibiotic removal via adsorption and photodegradation, and the preparation of carbonaceous materials and remediation properties regarding target contaminants are clarified. Meanwhile, the fundamental adsorption and photodegradation mechanisms and influencing factors are summarized. Finally, existing problems and future research needs are put forward. This work is expected to inspire subsequent research in carbon-based adsorbent and photocatalyst design, particularly for antibiotics removal.

1. Introduction

There have been growing concerns over water pollution due to human activities and technological and industrial development. Because some synthetic organic chemicals pose a significant risk to the ecological environment or human health while being used extensively, they have attracted the attention of researchers, and are defined as emerging contaminants (ECs) [1,2]. ECs encompass a wide range of artificial chemicals, such as pharmaceutical and personal care products (PPCPs), endocrine-disrupting compounds (EDCs), surfactants, and pesticides, among others [3,4,5,6]. They are difficult to remediate by natural attenuation and conventional treatment processes, so they easily accumulate in the ecosystem [7].
The excess presence of pharmaceuticals in water is emerging contamination, and antibiotics are one of the most common treatment groups. Antibiotics refer to metabolites produced by microorganisms such as fungi, bacteria, and actinomycetes that impede the growth and reproduction of other microbes [8,9]. Antibiotics are extensively used in aquaculture and human/animal therapies, resulting in a direct discharge of antibiotic residue into water sources, causing pollution problems [10]. Due to the antibacterial ability arising from functional groups of antibiotics, such as the linear fused tetracyclic nucleus (four rings) of tetracyclines, four-membered β-lactam ring of β-lactams, and large lactone ring of macrolides, the degradation of antibiotic molecules by microorganisms is beset with difficulties [11]. It was reported that antibiotics existing in sediments, soil, and water not only bring biochemical pollution, but also hasten the formation of antibiotic-resistance genes (ARG) and antibiotic-resistant bacteria (ARB) [12]. Therefore, the removals of antibiotics and their derivatives from water environments are urgent. Research showed that sulfonamides are the most frequently detected in all kinds of water environments, followed by fluoroquinolones, macrolides, and tetracyclines, while other categories of antibiotics are not as frequently detected [13]. Therefore, the removal of these four groups of antibiotics has attracted most researchers’ attention.
Various techniques have been devised to remove or reduce the concentration of antibiotics in the water environment, ranging from advanced oxidation such as ozonation, Fenton and photo-Fenton processes, and electrochemical oxidation to biological, photocatalysis, chlorination, membrane separation and adsorption [11,14,15,16]. Among them, adsorption is especially attractive for its facile operation, high efficiency, and potential scaling up [17]. As a green AOP technology, photocatalysis is expected to use the strong oxidation ability of photogenerated holes to mineralize antibiotic pollutants completely, without forming intermediates [18,19].
Both the efficiency of the adsorption and the photocatalysis process is strongly influenced by the type of materials and their characteristics. Compared to other commonly used adsorbents and photocatalysts, carbon-based materials show more advanced properties, including large specific surface areas (SSAs), developed pore structures, abundant functional groups, and high chemical stability [20]. Until now, many efforts have been made in pore structure engineering that can be classified as hierarchical porous structures, interconnected pores, carbonaceous defects, and graphene assemblies. In addition, more researchers are focusing on doping carbonaceous materials with heterogeneous atoms such as B, N, S, O, and P, and loading functional compounds like metals (oxides) to enhance the performance of adsorbents/catalysts for targeted pollutant removal [21,22,23]. As a result, carbon-based materials have been widely applied in both treatment processes in the past few years, showing great potential in applications of antibiotic removal.
In this review, we focus on removing antibiotics via adsorption and photodegradation by typical carbon-based materials. We consider graphene, carbon nanotube, biochar, and hierarchical porous carbon as the four primary carbonaceous materials, discussing their preparation and remediation properties regarding each target contaminant (Figure 1). Meanwhile, the basic adsorption and photodegradation mechanisms and influencing factors are summarized. Finally, existing problems and future research needs are put forward. We expect this review to inspire subsequent research in carbon-based adsorbent and photocatalyst design, particularly for antibiotics removal.

2. Characteristics of Carbonaceous Materials

The main carbonaceous materials used for antibiotic removal are graphene, carbon nanotube (CNT), biochar (BC), and hierarchical porous carbon (HPC), which is a novel, precisely designed carbonaceous material. They have many advantages, such as simple synthesis methods, cheap and readily available raw materials, large porosity, high SSA, well-developed and adjustable pore structure consisting of micropores, mesoporous, and macropores, and a large number of oxygen-containing functional groups, so they have many potential application prospects in antibiotic removal.

2.1. Graphene and Its Derivatives

Graphene is a two-dimensional nanomaterial in which carbon atoms connected by sp2 hybridization are closely packed into a single atomic layer of honeycomb-like crystal lattice which is organized through σ- and π-bonds. This material is the basis for all forms of graphite: folded-up graphene becomes fullerene (0D), a rolled-up graphene layer forms a carbon nanotube (1D), and graphite (3D) sheets become graphene when taken apart. It has the characteristics of good stability, conductivity, optical properties, and adsorption capacity, which makes graphene-based materials a new type of environmentally friendly catalysts with more active sites.
Synthesis approaches have a crucial impact on the performance and quality of graphene, and there are different synthetic routes such as chemical vapor deposition, the exfoliation method, and the organic synthesis approach [24]. These synthesis methods can be divided into two approaches: the top-down approach, based on reducing van der Waals forces of attraction between the graphite layers, and the bottom-up approach, based on combining unique molecular building blocks. In its pure form, graphene cannot perform well in water due to its hydrophobic nature. Thus, materials with hydrophilic properties, such as graphene oxide (GO) and reduced graphene oxide (rGO), are prepared for water treatment purposes. GO is graphene modified with oxygen-containing functional groups, showing various active sites (hydroxyl, carboxyl, epoxy, and carbonyl) in the carbon lattice. rGO is a type of GO with lesser oxygen content and higher interaction with the π bond of organic molecules, making it more efficient to remove the contaminants in the water. The graphene can be functionalized through covalent and noncovalent bonding to augment its loading capacity, specificity, biocompatibility and solubility (Figure 2), and graphene-based composites were invented to improve its binding affinity for anionic compounds and ability to be detached from treated wastewater.
Graphene-based materials are widely used in wastewater treatment via separation, membrane, and degradation processes. They are proven to be incredibly effective for the adsorption of inorganic contaminants such as heavy metal ions and rare earth metal ions, as well as organic contaminants, especially dyes, antibiotics [25], and AOPs such as the Fenton process, electro-Fenton process, photocatalysis, photoelectrocatalysis, ozonation, and sonolysis [26].

2.2. Carbon Nanotube (CNT)

A carbon nanotube, several microns in length, is a hollow cylindrical tube with side walls composed of sp2 hybridized carbon atoms, including two types: single-walled carbon nanotubes (SWCNTs) and multi-walled carbon nanotubes (MWCNTs). SWCNTs with a diameter between 0.4 and 3 nm are formed by curling a single layer of graphene. MWCNTs, whose outer and inner diameters are 2–100 nm and 1–3 nm, respectively, are made of two or more graphene sheets rolling into cylindrical shapes [27]. Due to of the attractive properties of CNTs, such as large surface area-to-volume ratio, high thermal and chemical stability, well-defined adsorption sites, easy attachment of functional groups, and nanoparticle loading [28], much research has been conducted to develop its potential in wastewater treatment applications.
CNTs are produced by using carbon sources and energy. Generally, there are three main synthesis methods: (1) the arc discharge method between graphite electrodes, (2) the laser ablation method catalyzed by transition metals, and (3) the chemical vapor deposition method for bulk fabrication [27]. Due to the defects of traditional CNTs, such as easy entanglement and agglomeration, surface inertness, and dispersing difficulty in the polymer matrix, it is hard to exert their decontamination performances, so their application is greatly restricted [29]. Therefore, similarly to graphene, functional methods based on covalent and noncovalent bonding are applied to equip CNTs with the required functional groups and make them more dispersible, increasing their affinity for water contaminants (Figure 2). Surface-modified CNTs have been studied extensively to remove heavy metals and organics from wastewater by adsorption. Furthermore, CNT composites can be prepared by introducing nanoparticles, high molecular polymers, etc., as photocatalysts and electrocatalysts (electrochemical reaction electrodes, microbial dye battery electrodes).

2.3. Biochar (BC)

Biochar is the solid carbonaceous residue from various agricultural and industrial biomasses via thermal methods. It is a low-cost, porous, and stable carbonaceous material with high SSA, high surface activity, and high ion exchange capacity that is well suited for use as an adsorbent and as a support matrix for metals and metal oxides [30,31].
As a biomass-derived material, BC is fabricated from different biomass feedstocks, including agricultural residues, forestry waste, organic fraction of municipal solid waste, industrial biomass by-products, and animal manures, via pyrolysis (300–700 °C) under oxygen-free or oxygen-limited conditions [32]. Critical physical properties of BC, surface area, and porosity play crucial roles in the wastewater treatment process. Moreover, biomass characteristics, including morphology structure, particle size, and pyrolysis parameters, predominantly temperature, are key influencing factors [33]. Higher pyrolysis temperature (>500 °C) results in lower oxygen content, greater hydrophobicity, higher surface area, and higher micropore volume, making the produced BC more suitable for removing organic pollutants. Conversely, BC pyrolyzed under lower temperatures (<500 °C) is better fit for the removal of inorganic pollutants [34].
To enable BC to better fulfill the requirements of practical applications in pollutant removal, modification is widely used to improve its adsorption or catalysis performance and alleviate secondary pollution. Physical modification methods, including steam/gas activation (air, CO2), ball milling, and microwave pyrolysis, have been applied to develop the pore structure of BC, and enhance its hydrophobicity and stability [35]. Chemical modification methods involving the action of chemical compounds can increase BC’s surface functional groups, polarity, and porosity. Acids, H2O2, and KMnO4 are common oxidants, while alkalis are reducers [35]. BC-based composites prepared by combining BC with clay minerals, metals, metal oxides, carbon materials, and polymers via impregnation, co-pyrolysis, or co-precipitation offer enhanced adsorption capabilities and conduct particular functionalities in contaminants removal [36]. In general, loaded magnetic materials enable BC to have magnetic separation properties, loaded carbonaceous materials improve their porosity, loaded TiO2 makes it catalytic, and loaded nano zero-valent iron can make it highly reactive.

2.4. Hierarchical Porous Carbon (HPC)

Hierarchical porous carbon is a novel carbonaceous material that contains two, or all three, types of interconnected pores (micropores (<2 nm), mesopores (2–50 nm), and macropores (>50 nm)), which serve different functions. Micro and smaller mesopores can provide a large surface area to promote active site dispersion, larger mesopores can minimize diffusive resistance to mass transport and allow easy access to the micropores, and macropores shorten mass-transfer pathways as reactant reservoirs. Moreover, the interconnected hierarchical pore structures can facilitate reactant entry into micropores and expedite the reaction rate [37,38]. In conclusion, because the space in HPC is effectively utilized, HPC with multiple levels of pores has higher SSA and pore volume than carbon materials with single-type pores [39,40]. It exhibits excellent performance in applications like adsorption, catalysis, and energy storage due to the functional synergy of the above structures.
The typical preparation strategies of HPC are the hard/soft template method and the template free method [41]. The hard template method synthesizes HPC materials by using silica, metal oxides, inorganic salts, and sublimable organic compounds as hard templates, incorporating them into carbons or carbon precursors, and removing them by either chemical etching or dissolution. Unlike hard templates, soft templates are mainly organic molecules or block copolymers that can be deposited or evaporated during annealing, avoiding the template removal step after pyrolysis. The template free method fabricates HPC by pyrolyzing select carbon precursors, mainly biomass and pre-engineered organic matter with intrinsic macropores, and constructing hierarchical porous structures by chemical activation. There are also some modification methods for HPC to develop additional micropores and mesopores, such as a combination of template carbonization and chemical activation (i.e., KOH or NaOH) to optimize the synthesis process and treat precursors with different chemicals [39]. Changing the feedstock to heteroatom containing materials to prepare heteroatom-doped HPC can change its physicochemical and electronic properties, and enhance its adsorption, catalysis, and electrochemistry performance [37].
Some researchers have paid attention to the application of HPC in removing organic contaminants from wastewater by adsorption. They studied organic contaminants included dye [42,43], antibiotic [44,45], hydrocarbon [46,47], bilirubin [48], etc.

3. Adsorption Removal of Antibiotics by Carbon-Based Materials

There has been extensive research on using carbonaceous materials to remove antibiotics by adsorption. Adsorption is an exothermic process that occurs on a surface, and involves the accumulation of adsorbate molecules in a gaseous or liquid form. Adsorption on carbonaceous materials includes four steps: (1) transport of solute from the bulk, (2) diffusion through the liquid film around the porous carbon particle, (3) diffusion in the liquid contained in the pore space, and (4) interaction between adsorbate and adsorbent [49]. The adsorption interactions between carbon-based materials and antibiotics can be summarized into physical and chemical adsorption. Weak Van der Waal interactions, π-π interactions, electrostatic contacts, and pore-filling/size-selective adsorption are the main mechanisms of physical adsorption. For chemical adsorption, surface chemical bonds or inner sphere coordination complexes are formed on an adsorbate and adsorbent surface either by electron pairing or electron transfer via Lewis acid-base interaction, ion/ligand exchange, and oxidation or reduction [50]. The existence of functional groups (–COOH, –OH, –NH2) in the carbonaceous materials may also have an impact on the adsorption process and efficiency [51]. The adsorption of antibiotics from the aqueous phase by carbonaceous materials is mainly influenced by three factors: the functional groups in the carbon materials, the pore size and pore uniformity of carbon materials, and environmental parameters, including pH, temperature, the concentration of inorganic salt, and other organic substances in the aqueous solution [52].

3.1. Graphene-Based Materials Applied in Antibiotics Adsorption

Several excellent properties make graphene an effective adsorbent for removing antibiotics.: (1) highly exposed atoms lead to easy contact with antibiotics (mainly by π-π interaction), (2) porous structure and high surface area ensure faster diffusion or surface reactions of antibiotics, resulting in rapid and effective adsorption, and (3) lower cost of large-scale production [53]. The adsorption of GO, rGO, functionalized graphene, and graphene-based composites on antibiotics will be presented and discussed below. Table 1 lists recent studies of graphene-based adsorbents on antibiotics.
GO has been investigated extensively for antibiotic adsorption removal, particularly in aqueous environments due to its high hydrophilicity. It has been reported that GO can effectively sorb sulfamethoxazole (SMX), ciprofloxacin (CIP) [54], and various tetracycline antibiotics [55] via π-π interaction and electrostatic attractions. Recently, Salihi et al. (2020) reported on the adsorption behaviors of trimethoprim (TMP) and isoniazid (INH) onto GO. The maximum adsorption capacities of TMP and INH were 204.08 mg g−1 (pH = 8) and 13.89 mg g−1 (pH = 2), respectively [56]. Moreira et al. (2020) verified the simultaneous adsorption degradation of norfloxacin (NOR) by GO, with a removal capacity of 374.9 mg g−1 [57]. Moreover, GO also exhibited good adsorption properties for cephalexin (CFX) (164.35 mg g−1)[58] and azithromycin (AZM) (55.55 mg g−1) [59]. Fewer studies were conducted on rGO. For example, Liu et al. (2016) examined the adsorption of two sulfonamide antibiotics by two rGOs, which are less affected by dissolved organic matter than CNTs and graphite [60].
For better adsorption properties, hydrophilic functional groups (–OH, epoxide, –COOH) are often used for the covalent modification of graphene, and graphene-based composites were prepared by heteroatom doping [61], nanoparticle loading, and organic molecular and polymeric modification [62]. Recently, different types of components are often used simultaneously in modifications to achieve a synergistic effect, among which magnetic nanoparticles such as Fe3O4 are most commonly used to obtain good solid-liquid separation properties [62,63]. For instance, Lin et al. (2021) fabricated an ethanolamine and Fe3O4 nanoparticle; a co-functionalized, graphene-based adsorbent with magnetic properties by co-precipitation. The introduction of ethanolamine into GO can decrease the possibility of agglomeration, while Fe3O4 nanoparticles can make the adsorbent rapidly separate using a magnetic field. Due to the synergy, in only five minutes, the adsorption capacities of the material for TC and levofloxacin (LVX) were 315.25 mg g−1 and 229.53 mg g−1 [64]. In addition, natural materials like chitosan [65] and attapulgite [66] have also been used to modify graphene to adsorb antibiotics.
To overcome the shortcomings of the 2D nanostructure, three-dimensional (3D) graphene materials were synthesized, including aerogels and hydrogels. A ring defects-rich and pyridinic N-doped graphene aerogel was fabricated by Wang et al. (2022), and exhibited an outstanding performance for TC removal (adsorption capacity of 607.1 mg g−1) as a floating adsorbent [67]. Moreover, ultra-thin g-C3N4 modified graphene oxide hydrogels were prepared via a one-step hydrothermal reaction by Wang et al. (2022), which exhibited super high co-adsorption capabilities for TC and Cu(II) [68]. These aerogels and hydrogels have large adsorption capacity and can be easily modified and separated from water.
MOFs/GO composites are also becoming promising adsorbents for wastewater treatment, as the metal ions in MOFs can interact with the epoxy and hydroxyl functional groups on both sides of the GO flakes. Chen et al. (2022) synthesized Alg-Cu@GO@MOF-525 by dispersing GO with copper alginate and in situ growth of MOF-525, which helped reduce aggregation. The SSA of the material was as high as 807.3 m2 g−1, which is favorable for the adsorption of TC (adsorption capacity of 533 mg g−1) [69].
Table 1. Adsorption characteristics of antibiotics by graphene-based materials.
Table 1. Adsorption characteristics of antibiotics by graphene-based materials.
Antibiotic ClassAntibiotic CompoundsAdsorbentTemperature
(K)
pHEquilibrium
Time (Min)
Adsorption
Capacity
Qm (mg g−1)
Refs.
TetracyclineTetracyclineGO2983.6190313.48[55]
TetracyclineMAEGO30341440487.82[64]
TetracyclineGO@ATP3085120 [66]
TetracyclineDNGA298420607.1[67]
TetracyclineUCN-GH2985100 [68]
TetracyclineAlg-Cu@GO@MOF-5253187900533.2[69]
OxytetracyclineGO2983.690212.31[55]
OxytetracyclineGO293560130.4[61]
OxytetracyclineB-rGO29356083.9[61]
DoxycyclineGO2983.690398.41[55]
DoxycyclineGO@Fe3O4@β-cyclodextrin298745204.5[70]
SulfonamideSulfamethoxazoleGO29851440240[54]
SulfamethoxazoleGO@β-cyclodextrin@ dopamine hydrochloride308290144[62]
SulfadiazineGO@β-cyclodextrin@ dopamine hydrochloride308290152[62]
TrimethoprimGO298860204.08[56]
IsoniazidGO29826013.89[56]
QuinoloneCiprofloxacinGO29852880379[54]
NorfloxacinGO303730374.9[57]
levofloxacinMAEGO30341440330.71[64]
β-lactamsCephalexinGO2987420164.35[58]
MacrolideAzithromycinGO29870.2555.55[59]

3.2. Carbon Nanotube-Based Materials Applied in Antibiotics Adsorption

The large SSA, biological harmlessness, and an abundance of oxygen-containing functional groups of CNTs offered various potential applications for the adsorption of antibiotics in an aqueous environment. Table 2 illustrates data on antibiotic removal by various CNTs.
Earlier research focused on using pristine and oxidized MWCNTs to adsorb different kinds of antibiotics, such as sulfonamides [71], amoxicillin [72], fluoroquinolone [73], and ciprofloxacin [74], and comparing the adsorption effect of different types of CNTs towards antibiotics. In most instances, SWCNT showed the highest adsorption capacity for its large surface area and unique basic surface [74,75].
CNT-based composites have been studied extensively for the adsorption removal of antibiotics, especially for TC and CIP. Various magnetic carbon nanotube adsorbents were developed via coprecipitation [76,77,78] or hydrothermal [79] methods for their magnetic separation ability, which could improve the recyclability of CNTs. For example, Sereshti et al. (2022) developed Fe3O4 doped and CdS functionalized MWCNTs, used for TC and cefixime removal. Due to the introduction of Fe3O4, the material was magnetized and could be easily separated by an external magnet [78]. However, adsorption site occupation by ferrites may lead to a decrease in adsorption capacity. Zhao et al. (2021) used three high crystalline magnetic nano spinel ferrites (MFe2O4; M = Co, Cu, Mn) to modify carbon nanotubes and compared their adsorption mechanisms. They found that the highest adsorption capacity (63.32 mg g−1) and removal efficiency (99.3%) of CIP were obtained by using CoFe2O4/CNTs, because Co atoms were most favorable for O adsorption according to their d-band states [79]. Besides, layered double hydroxides (LDHs) [80] and ionic liquids (ILs) proved to be ideal materials to be loaded into CNTs for better removal efficiency.
Fe-doped MOF-loaded CNTs [81,82] attracted many researchers’ attention, as the introduction of high-valence Fe3+ metal ions could improve the water stability of MOFs, therefore enhancing their applications in the liquid phase. The CNT/GO/sodium alginate triple-network nanocomposite hydrogel was a 3D material that was prepared by Ma et al. (2020). It achieved high-efficiency removal of CIP, with an adsorption capacity of 200 mg g−1 at 288 K [83].
Table 2. Adsorption characteristics of antibiotics by CNT-based materials.
Table 2. Adsorption characteristics of antibiotics by CNT-based materials.
Antibiotic ClassAntibiotic CompoundsAdsorbentTemperature
(K)
pHEquilibrium
Time (Min)
Adsorption
Capacity
Qm (mg g−1)
Refs.
TetracyclineTetracyclineM-MWCNT3084-7240494.91[76]
TetracyclineFe3O4/MWCNT-CdS298560116.27[78]
TetracyclineMWCNT/ZIF-8(Fe)2986360589.42[82]
Tetracycline hydrochlorideLDH@CNT2988480756.2[80]
Tetracycline hydrochlorideMWCNT/MIL-53(Fe)2987 364.37[81]
Oxytetracycline hydrochlorideMWCNT/MIL-53(Fe)2987 325.59[81]
Chlortetracycline hydrochlorideMWCNT/MIL-53(Fe)2987 180.68[81]
SulfonamideSulfamethoxazoleMWCNT2983720 [71]
SulfamethazineFe3O4/MWCNTs29871440 [77]
QuinoloneFluoroquinoloneO-MWCNT29831440 [73]
CiprofloxacinCoFe2O4/CNTs2986–730063.32[79]
CiprofloxacinCNTs/L-cys@GO/SA2885.43600200[83]
Ciprofloxacin4.7%O-MWCNT298460177.8[84]
β-lactamsAmoxicillinMWCNT333775159.4[72]
NitroimidazoleMetronidazoleSWCNT29877200101[75]
DimetridazoleSWCNT2987720084[75]
CephalosporinCefiximeFe3O4/MWCNT-CdS298560105.26[78]

3.3. Biochar-Based Materials Applied in Antibiotics Adsorption

Due to its excellent properties, e.g., high porosity, aromaticity, multiple anionic functional groups, large surface area, and a wide range of sources, biochar has been considered an effective carbonaceous sorbent for antibiotics. Table 3 lists recent advances in BC-based materials for the sorptive removal of antibiotics.
The adsorption of TC on BC derived from auricula dregs [85] and cow manure [86], obtained at different pyrolysis temperatures, was investigated. Results showed that the BCs prepared at a high temperature had a greater capacity for removing TC, as they have better pore distribution, larger SSA, and greater adsorption affinity. Stylianou et al. (2021) assessed the adsorption capacity of three different derived biochar on seven antibiotics. They found that the removal efficiency of these antibiotics dramatically depends on the pH value of the medium and, therefore, on the antibiotics’ speciation, and the physicochemical and structural characteristics of the adsorbents (primarily SSA, pore size, functional groups, dosage) [87]. Zhao et al. (2022) used a concentrated sulfuric acid-based, one-step carbonization method to synthesize BC for enrofloxacin (ENR) adsorption. The obtained BC was highly graphitized with more active functional groups on the surface and showed significantly superior adsorption capacity for ENR (142.3 mg g−1), which was 13.7 times that of pyrolytically synthesized BC (10.4 mg g−1) [88]. In conclusion, antibiotics adsorption by pristine BCs mainly depends on their feedstock type, synthesis method, and synthesis condition.
Various modification methods have been used to improve the unsatisfactory adsorption capacities of pristine BCs. Ball milling, a physical method, has received enormous attention from researchers recently. For example, urea N-doped biochar was facilely prepared by Wu et al. (2021) via wet ball-milling to adsorb and degrade NOR. It was proved that ball-milling enhanced biochar for NOR adsorption (adsorption capacity of 11.48 mg g−1) by H-bonds, π-π EDA, and pore-filling [89]. At the same time, different kinds of chemical reagents were used to change the chemical nature of BCs, such as H3PO4 [90,91], citric acid [92], and ZnCl2 [93,94], all of which confirmed that chemical modification could increase the functional groups, SSA, and porosity of BC. Specifically, Sheng et al. (2022) synthesized a critical acid-modified BC (CABRC) as an efficient adsorbent for TC. Higher surface area, more pore structures, and stronger electron donor capacity provided by generated oxygen functional groups (e.g., –COO, –COOH, and –OH) synergistically improved the adsorption performance of the CABRC [92].
FeCl3 was widely used to synthesize magnetic BCs via the impregnation method, and some special chemical reagents were added for better adsorption capacities and higher chemical stability, such as humic acid for coating [95], dicyandiamide for the introduction of N [96], MgCl2·6H2O [97], and NiSO4·6H2O [98] for the formation of ferrites. Apart from FeCl3, K2FeO4 was also found suitable for BC magnetization, as ferrates could not only serve as the precursor of iron oxide (loaded on the biochar surface) but also enlarge the surface area of the resulting biochar due to their solid oxidizing capability [99,100]. In addition, Qu et al. (2021) studied a novel magnetization process. They used two-step Microwave (MW) assisted processes to provide the obtained BC with high SSA and favorable magnetism, whose adsorption capacity of TC was 202.62 mg g−1 [101].
Furthermore, clay mineral BC composites are known to have improved overall performance [102,103]. For instance, Arif et al. (2022) investigated the removal of CIP using biochar treated with clay minerals and subsequently activated with carbon dioxide. The material exhibited a large adsorption capacity (50.32 mg g−1) due to enhanced active sites and functional groups [103]. In another research, Xiang et al. (2022) improved the SSA and pore volume of two biochar by lignin impregnation and enhanced their adsorption capacities of tetracycline hydrochloride (TCH). The results illustrated that biochar with porous modulation by lignin impregnation could be significantly applied to the removal of antibiotics [104].
Table 3. Adsorption characteristics of antibiotics by BC-based materials.
Table 3. Adsorption characteristics of antibiotics by BC-based materials.
BiomassEngineering MethodAntibioticPyrolysis
Temp
(°C)
Adsorption Temp
(K)
pHEquilibrium
Time (Min)
Adsorption
Capacity
Qm (mg g−1)
Refs.
Auricula dregs Tetracycline70029876011.9[85]
Cow manure Tetracycline7002986144011.80[86]
Camellia oleifera shellsH3PO4Tetracycline6002986240451.6[91]
Biogas residueCitric acidTetracycline800298760058.25[92]
Aerobic granular sludgeZnCl2Tetracycline7003085288093.44[93]
Flueggea suffruticosaZnCl2Tetracycline500303750188.7[94]
Walnut shellFeCl3·6H2O, dicyandiamideTetracycline6002987.2 238.9[96]
Water hyacinthFeCl3·6H2OTetracycline700318 200202.62[101]
Wheat StrawLigninTetracycline hydrochloride6002987 31.48[104]
Flueggea suffruticosaZnCl2Chlortetracycline50030310 200.0[94]
Flueggea suffruticosaZnCl2Oxytetracycline5003037 129.9[94]
Coffee groundsH3PO4Sulfadiazine700298 180139.2[90]
Wheat stalkK2FeO4Sulfadiazine700298654047.85[100]
Garlic peelConcentrated H2SO4 carbonizationEnrofloxacin 2987T1/2 = 34.13142.3[88]
Apple branchesFeCl3, humic acidEnrofloxacin700308572048.3[95]
Apple branchesFeCl3, humic acidMoxifloxacin700308872061.5[95]
Corn stalkBall-milling, ureaNorfloxacin6002985 11.48[89]
SludgeBentoniteNorfloxacin5502986108089.36[102]
Pomelo peelMgFe2O4Levofloxacin7002985240115[97]
VinasseNiFe2O4Levofloxacin70029861080172[98]
Rice huskMontmorillonite, CO2Ciprofloxacin350295772050.32[103]
Penicillin fermentation dregsAcetic acid, K2FeO4Penicillin40030811 322.58[99]

3.4. Hierarchical Porous Carbon-Based Materials Applied in Antibiotics Adsorption

Recently, HPC with uniform and interconnected pore structure for antibiotics adsorption has received increasing research attention (Table 4), which is mostly due to the special characteristics of HPC, including its fine and well-defined pore systems, good mass transport through larger pores, and the substantial number of adsorption sites at smaller pores [39].
Most HPC for antibiotics adsorption were synthesized via the template free method, using a two-step thermal treatment involving low-temperature carbonization, which uses biomass as carbon precursor, and high-temperature activation, which includes acid-alkaline treatment, metal ions, steam/gas purging, or oxidizing agents [44,105,106,107]. Wang et al. (2022) chose five waste biomasses to prepare hierarchically micro/mesoporous carbons with a high BET surface area of 1687-2003 m2 g−1 and large adsorption capacity for chloramphenicol (CAP) (>300 mg g−1) after an equilibrium of only 40 min. They established that micropore filling might be a leading adsorption mechanism and that the ratio of micropores to mesopores in biochar adsorbents was a key element in attaining rapid, high-capacity adsorption [44]. To simplify the manufacturing process and reduce preparation cost and time, Chen et al. (2021) proposed a one-step method for making layered carbons with different pore structures by adjusting the temperature and the mass ratio of activator K2CO3. The decomposed K2O and CO2 reacted with carbon, forming more micropores and improving pore volume. The obtained HPC had excellent adsorption capacities (534 mg g−1) at a lower initial concentration of CAP (120 mg L−1) [108]. Other researchers used pre-engineered organic matter as a carbon precursor. Pi et al. (2022) used glucose hydrochar as the precursor to prepare bifunctional N-doped HPC, which could adsorb TC and activate PDS as the activator, maintaining above 77.18% adsorption capacity after six cycles. The material’s large BET surface area and microporous-mesoporous structure promoted adsorption by offering more binding sites on its surface to form a stable interaction. At the same time, the mesoporous was conducive to the transfer of reactants or electrons and increased the catalytic performance of the material [45].
Earlier researchers used halloysite nanotubes as the hard template to obtain HPC through liquid phase impregnation and carbonization. This kind of HPC possessed extraordinary adsorption capacity for TC (1297.0 mg g−1) and CAP (1067.2 mg g−1) [109,110]. In another study, a soft-hard template method was used to fabricate an HPC that possessed micro/meso bimodal pores, in which tetraethyl orthosilicate was used as an inorganic precursor, and the triblock copolymer F127 as a structure-directing agent. Due to the bimodal porous structure with meso channels and connected micropores of the prepared HPC, which enabled rapid transport of TC molecules and easier access to the adsorption sites, the HPC exhibited great adsorption performance of TC in both the batch mode adsorption and the column adsorption process [111]. Recently, Zhang et al. (2022) successfully prepared N-doped HPC via simple one-step carbonization and employed it for TCH adsorption. The sodium dodecylbenzene sulfonate was introduced as a soft template, which optimized the pore structure of the carbon materials, increased the nitrogen content, and added surface functional groups, leading to a high TCH adsorption capacity [112].
Table 4. Adsorption characteristics of antibiotics by HPC-based materials.
Table 4. Adsorption characteristics of antibiotics by HPC-based materials.
Hierarchical Porous Carbon MaterialCarbon PrecursorModification MethodAntibioticAd
Temp
(K)
pHTime (Min)Ad
Capacity
Qm
(mg g−1)
Refs.
Macro-meso-micro hierarchical porous carbonWheat strawKOH + KMnO4 activationTetracycline3187 584.19[105]
Fe-doped HPCEichhornia crassipes debrisHCl activation,
Fe + amino acetic acid synergistic treatment
Tetracycline3183–1110457.85[107]
N-doped bifunctional HPCGlucose hydrocharKHCO3 activation, nitrogen dopingTetracycline3034.85 629.76[45]
Macro-meso-micro hierarchical porous carbonSodium lignin sulfonateKOH activationTetracycline29833601297.0[109]
Micro/meso bimodal porous carbonSoluble phenolic resin Tetracycline2987 701.31[111]
N-doped HPCSoft-templated ZIF-8 forNitrogen dopingTetracycline hydrochloride2984.590080.92[112]
Hierarchical micro/mesoporous carbonSoybeanKOH activationChloramphenicol298540892.9[44]
Hierarchical micro/mesoporous carbonCorncobKOH activationChloramphenicol318940662.3[44]
Oxygen-enriched HPCSodium lignosulfonateK2CO3 activationChloramphenicol3034.86720534.0[108]
Macro-meso-micro hierarchical porous carbonSodium lignin sulfonateKOH activationChloramphenicol2983–113601067.2[109]
Hierarchical micro/mesoporous carbonSodium carboxymethyl celluloseKOH activationChloramphenicol2982–660769.95[110]
Hierarchical micro/mesoporous carbonHigh-salted Spirulina residueKHCO3 activationSulfathiazole2987240218.4[106]

4. Photocatalytic Degradation of Antibiotics by Carbon-Based Materials

In addition to adsorption, photocatalysis is a green and efficient technology used to remove antibiotics. Antibiotics can be effectively degraded into a non-toxic small molecular species by a reactive species (e.g.,·OH, O2) produced by photocatalysts under sunlight, visible light, or ultraviolet (UV) light.
The predominant mechanisms of photocatalyst degrading antibiotics can be summarized into three main steps: photon absorption, excitation, and reaction. First, the photocatalyst adsorbs photons equal or superior in energy to its band gap; then, the electrons in the valance band (VB) are activated and jump into the conduction band (CB) where a hole (h+VB) is produced; finally, the highly reactive positive holes and electrons are separated and migrate to the surface of the photocatalyst, reacting with compounds such as O2, OH and H2O groups, and then generating oxidizing species which can degrade antibiotics (Figure 3) [14,113]. Generally, h+VB can also directly react with antibiotics, leading to an efficient removal [113]. As for the influencing factors in antibiotic photodegradation, there are three aspects: light source (wavelength), photocatalyst (size, shape, morphology, and dosage), and environment (solution pH, concentration of antibiotics, and coexisting ions and organics) [11].
Semiconducting materials, such as TiO2, WO3, ZnO, Fe2O3, CdS, MoS2, and graphitic carbon nitride (g-C3N4), are widely used photocatalysts for water remediation. However, most of them suffer from limitations including large band gaps, instability in an aqueous medium, and high electron-hole recombination rates [114]. To solve these problems, researchers started to combine them with carbonaceous materials to form composites for three main advantages of carbonaceous materials: (1) high surface area, which implies an improved adsorption ability; (2) property of decreasing electron-hole recombination rates, which is because they can effectively scavenge photo excited electrons from the conduction band of semiconducting photocatalysts; (3) as the carrier, co-catalyst, and photosensitizer, abilities to broaden the light absorption range and elevate the photocatalytic activity of the composite [115,116].

4.1. Graphene-Based Materials Applied in Photocatalytic Degradation of Antibiotics

Due to its large SSA for uniform dispersion, narrow band-gap energies, high electrical conductivity, and low production cost for mass production, graphene has been thoroughly investigated as a favorable photocatalyst for photodegrading antibiotic pollutants in the aqueous environment. Nevertheless, the self-aggregate process between graphene sheets makes it easy to lose catalytic activity. Therefore, efforts were made to study graphene-based photocatalysts combined with other photocatalysts to achieve a better catalytic performance of antibiotics (Table 5) [53].
Metal oxides are most commonly applied in the photocatalytic degradation of antibiotics. Graphene [117], Go [118], and rGO [119] have all been used to couple with TiO2 to increase TiO2′s visible light adsorption during photocatalytic degradation. Bi-based composite is another prominent photocatalyst. Bismuth tungsten oxide (Bi2WO6) [120], bismuth vanadate (BiVO4) [121,122], and Br-deficient bismuth oxybromide (BixOyBrz) [123,124] are the three most common Bi-based oxides used for construction of hierarchical heterojunction with graphene and its derivatives. Jiang et al. (2022) facilely synthesized (Bi)BiOBr/rGO by employing an in situ reduction strategy, which exhibited >98% degradation efficiency of TC within 20 min. The improved photocatalytic ability was because rGO was able to speed up the electron transfer, further restraining the recombination of the electron-hole pairs and promoting the photodegradation process [123]. Meanwhile, other metal oxide photocatalysts, such as magnetic metal oxides (cobalt ferrite [125], Fe2O3 [126]), delafossite oxides (AgFeO2 [127]) and pyrochlore (La2Zr2O7 [128]), were combined with graphene for their unique properties. For example, cobalt ferrite/rGO porous balls synthesized via modified microfluidic and calcination methods are to be used for efficient OTC degradation (84.7%) under visible light and could be quickly recovered by magnets. This photocatalytic performance was ascribed to a narrow band gap, outstanding light-harvesting properties, low carrier recombination, and high electron-hole separation capacity [125]. Layered double hydroxide was also identified as a promising photocatalyst, and Ce-doped NiAl LDH/rGO composite was fabricated via a one-step hydrothermal method by Gao et al. (2021) and achieved superior degradation of 94% for CIP within 60 min in visible light. The presence of rGO offered larger SSA, suppressed the agglomeration of NiAl0.85Ce0.15 nanosheets, and accelerated separation of photo-generated charges, leading to an excellent photocatalytic performance [129].
Compared to binary and single photocatalysts, ternary photocatalysts exhibited enhanced photoelectric performance and photocatalytic activity. With a large surface area and excellent conductivity, Graphene has shown great potential as the mediator in ternary Z–scheme heterojunctions. Hence, many combinations of two different semiconductors, whose energy bands could couple with each other, forming effective interfacial charge transfer channels, were studied to be anchored on graphene, with graphitic carbon nitride (g-C3N4) used most frequently [130,131,132,133,134]. A remarkable work by Wang et al. (2021) described Z-scheme phosphate-doped BiVO4/graphene quantum dots/P-doped g-C3N4 composites with enhanced photocatalytic activity for the degradation of NOR (86.3% within 120 min). The principal reason for such a noteworthy performance was the role of graphene quantum dots as the electron mediator for inhibiting the recombination of photogenerated electron-holes, boosting interfacial charge transfer between phosphate-doped BiVO4 and P-doped g-C3N4, and expanding the range of visible light response [133]. In addition to Z-scheme systems, many other ternary photocatalysts were studied for their advanced properties and facile synthesis. For example, Kumar et al. (2022) prepared ZnO/CdO/rGO nanocomposites by refluxing, which realized 99.28% removal of CIP under UV light for 75 min. The fantastic photocatalytic activity of the material was due to the combination of ZnO and CdO, which expanded the band gap energy, and the presence of rGO, which improved adsorption capacity and generated more reactive oxygen species (ROS) by enhancing electron-hole pair separation [135]. Moreover, BiVO4/FeVO4@rGO [136] with novel 3D/2D/2D heterojunction structure and bifunctional CeO2/CdS/rGO [137] for photodegradation and photoreduction were synthesized by a green and cost-effective one-step hydrothermal method.
Table 5. Degradation characteristics of antibiotics by graphene-based photocatalysts.
Table 5. Degradation characteristics of antibiotics by graphene-based photocatalysts.
PhotocatalystsAntibioticDosage (g/L)Detection Wavelength (nm)Light SourceDegradation EfficiencyRef.
(Bi)BiOBr/rGOTetracycline1.0 Visible light>98% within 20 min[123]
rGO/Bi4O5Br2Tetracycline0.5356Visible light95.2% within 60 min[124]
α-Fe2O3/rGOTetracycline5.0 Visible light99% within 140 min[126]
La2Zr2O7/rGOTetracycline1.0357Visible light82.1% within 40 min[128]
Graphene/TiO2/g–C3N4Tetracycline 357Visible light83.5% within 80 min[130]
g-C3N4/MnO2/GOTetracycline0.5 Visible light91.4% within 60 min[131]
15%AgBr/5GO/Bi2WO6Tetracycline0.4357Visible light73.3% within 15 min, up to 84%[132]
Ag2O/Bi2WO6/rGOTetracycline1.0 Visible light95.3% within 40 min[134]
BiVO4/FeVO4@rGOTetracycline0.6356Visible light91.5% within 100 min[136]
QDs-BiVO4/rGHTetracycline hydrochloride0.5357Visible light73.2% within 120 min[122]
CF/rGOOxytetracycline 354Visible light84.7%[125]
GO/TiO2Amoxicillin0.6230UV light99.84% within 60 min[118]
rGO/Bi2WO6Norfloxacin0.5 Visible light87.49% within 180 min[120]
BiVO4/GQDs/PCNNorfloxacin1.0273Visible light86.3% within 120 min[133]
W-BiVO4-x/rGOCiprofloxacin1.0 Visible light93.6% within 60 min[121]
NiAlCe LDH/rGOCiprofloxacin0.25271Visible light94% within 180 min[129]
ZnO/CdO/rGOCiprofloxacin0.5270UV light99.28% within 75 min[135]
CeO2/CdS/rGOCiprofloxacin0.5 Sunlight90% within 120 min[137]
AgFeO2/GO3Lomefloxacin0.583250–450Visible light88% within 75 min[127]

4.2. Carbon Nanotube-Based Materials Applied in Photocatalytic Degradation of Antibiotics

CNTs with multiple graphite layers not only have the electron-trapping ability to attract electron pairs that semiconductors emit and increase the efficiency of photocatalytic reactions but also exhibit photosensitization, which can release electrons into the semiconductor’s conduction band, raising the electron-hole density, resulting in efficient charge separation. Thus, CNTs should play essential roles in photocatalytic composites for antibiotics degradation [138]. Table 6 lists recent studies of CNT-based photocatalysts for antibiotics removal.
It was reported that CNTs-based binary photocatalysts have the potential to change nanocrystal structure and superior photocatalytic efficiency. Many photocatalysts, such as TiO2 [139], MOFs [140], Bi-based oxides [141,142], and lanthanum vanadate [143], were combined with CNTs for antibiotics photocatalytic degradation. Very recently, Gao et al. (2022) used MWCNTs to induce and optimize the {312}/{004} facet ratio change of Bi5O7I photocatalysts, and therefore accelerated the transformation of photogenerated charge carriers. As a result, the composites exhibited high photocatalytic oxidation efficiency (88.2%) for representative ofloxacin (OFL) [142]. Another noteworthy work presented by Khazaee et al. (2021) in which CuBi bimetallic alloy nanosheets-supported functionalized MWCNTs were synthesized via a facile one-pot process, and achieved the highest degradation rates of CIP and OFL at only 90 min reaction time. The synergistic effects of the SPR effect of Cu, the boosting of hot electron flux at the Schottky junction caused by the intimate contact between the alloy and nanotubes, and the higher oxidative activity led to a high photocatalytic oxidation rate of the nanocomposite [144]. Moreover, Zuo et al. (2022) constructed mesoporous carbon nanotube networks loaded with trace atomic Fe sites by calcining melamine–cyanuric acid complex and glucose absorbed with Fe3+ at high temperatures, which exhibited high photodegradation rates of 93.2%, 99.4% and 94.3% for TCH, CTC and OTC. The single atomic Fe with a Fe-N-C structure improved visible light absorption, and the N-doped CNT networks promoted electron and mass transfer as well as enhanced the fast transfer of degradation products along the inner and outer walls of the tubes [145].
Currently, more researchers are paying attention to the design of ternary photocatalysts based on CNTs, enhancing the removal efficiency of various antibiotics, which mainly are quinolones [138,146], sulfonamides [146], β-lactams [147,148] and tetracyclines [138,149,150]. Zuo et al. (2022) demonstrated the preparation of Z-scheme Ag–AgBr/Bi2O2CO3/CNT heterojunctions, which had an outstanding complete removal of TC within 40 min. CNTs increased visible light absorption, and their electrical conduction promoted the fast transfer of electrons, the increase in carrier density, the decrease in interface transfer resistance, and the corresponding increase in photocurrent, all of which were responsible for the significantly improved photocatalytic performance [138].
Table 6. Degradation characteristics of antibiotics by CNT-based photocatalysts.
Table 6. Degradation characteristics of antibiotics by CNT-based photocatalysts.
PhotocatalystsAntibioticDosage (g/L)Detection Wavelength (nm)Light SourceDegradation EfficiencyRef.
Ag–AgBr/Bi2O2CO3/CNTTetracycline0.4 Visible light100% within 40 min[138]
MWCNT/TiO2Tetracycline0.2360UV light100% within 100 min[139]
HPWx@Fe2O3-CNTsTetracycline0.25356Visible light100% within 40 min[149]
Fe-CNTsTetracycline hydrochloride0.05358Visible light93.2% within 100 min[145]
MWCNT/BiVO4Oxytetracycline0.25360Visible light88.7% within 60 min[141]
Fe-CNTsOxytetracycline0.05355Visible light94.3% within 100 min[145]
Fe-CNTsChlortetracycline0.05370Visible light99.4% within 80 min[145]
NiFe2O4/MWCNTs/BiOIDoxycycline1.25351UV light92.18% within 300 min[150]
CNT@MIL-101(Fe)Ciprofloxacin0.5 Visible light90% within 45 min[140]
CuBi/MWCNTsCiprofloxacin Visible light93% within 90 min[144]
MWCNTs-{312}/{004}Bi5O7IOfloxacin1.0 Visible light88.2% [142]
CuBi/MWCNTsOfloxacin Visible light91% within 90 min[144]
Bi2MoO6/Bi2WO6/MWCNTsOfloxacin2.0 Visible light91.3% within [146]
CNTs/LaVO4Sulfamethazine0.3255Visible lightUp to 95% within 90 min[143]
Bi2MoO6/Bi2WO6/MWCNTsSulfadimidine2.0 Visible light88.8%[146]
SWCNT/ZnO/Fe3O4Cefixime0.46280UV-A light94.19%[147]
Bi2WO6/CNT/TiO2Cephalexin0.75262Sunlight98.7% within 70 min[148]

4.3. Biochar-Based Materials Applied in Photocatalytic Degradation of Antibiotics

Under light irradiation, biochar can generate various reactive oxygen species (ROS) through biochar carbon matrix (BCM) and dissolved organic matter (DOM) to enhance the indirect photodegradation of pollutants and oxidize various molecules. Yang et al. (2021) revealed that BCs could photodegrade the SMX and CAP through direct and indirect pathways. Specifically speaking, in the low-concentration BC solutions, the oxygen reduction performances of BCs and ·OH radicals generated by DOMs can indirectly photodegrade SMX and CAP by promoting the oxidation process and electron transfer [151]. Furthermore, modification methods were applied to improve the light response performance of BC, such as ball milling was proven to improve the capacities of adsorption and ROS generation by increasing the SSA, pore volume, oxygen-containing functional groups, and the carbon defects of BC [152].
Moreover, BC is a good carrier for photocatalysts as it can bring some advantages: (1) facilitate the photodegradation process due to the high adsorption capacity, (2) expand the light absorption range, and (3) provide an efficient electron-transfer channel and acceptor to enhance the separation of photogenerated electron-hole pairs. So, the combination of BC and semiconductors such as metallic oxides, metallic sulfides, and graphitic carbon nitride were intensively explored for antibiotics removal. For example, (modified) TiO2/BC [153,154,155], Bi2WO6/Fe3O4/BC [156], PbMoO4/BC [157], CdS/BPC [158], and g-C3N4/BC [159,160] all have high photocatalytic efficiency and long-term use durability because of their high surface area, adequate active sites, and compact interface. Recently, layered double hydroxides/oxides [161,162] and spinel ferrites received great attention as emerging categories of photocatalysts to be combined with BC. Li et al. (2022) prepared ZnO/ZnFe-LDH loaded on BC via a facile hydrothermal method. The BC matrix offered a homogeneous distribution of LDH, narrowed band gap, expanded visible-light adsorption, and enhanced separation of electron-hole pairs, leading to significant photocatalytic degradation of TC (87.7%) [163]. Zinc ferrite (ZnFe2O4) with narrow band-gap energy of 1.9 eV was combined with B, N co-doped BC by Peng et al. (2021) to generate a large surface area and numerous vacant sites for TCH adsorption and electron capture, and form the tight interface for efficient transfer and mobility of charge carriers [164].
Additionally, it has been shown that a system based on transition metal phosphides is a wise choice for photocatalytic applications. Wang et al. (2022) constructed a photocatalytic system composed of NixP/biocarbon composite as the reactive center and erythrosine B as a sensitizer for SMX removal. The NixP mixed crystal phases presented excellent photocatalytic activity due to its faster electron transfer ability [165]. Table 7 lists the BC-based photocatalysts mentioned above for antibiotics degradation.

4.4. Hierarchical Porous Carbon-Based Materials Applied in Photocatalytic Degradation of Antibiotics

Carbon materials with a hierarchical architecture are promising photocatalysts as their special interconnected porous structures can enhance the transport of substrates/products and the exposure of active sites so that in subsequent reactions, the charge carrier transfer distance would be shortened, which will promote the separation and migration of photo-excited electron-hole carriers. However, few pieces of research have been conducted on HPC-based photocatalysts for antibiotics removal.
Zhou et al. (2021) reported a novel catalyst of N-doped magnetic three-dimensional hierarchical porous carbon microspheres@TiO2 (N-doped MCMs@TiO2), which demonstrated impressive photodegradation ability for TC (up to 99.87% within 120 min) under UV irradiation. According to the test results, the catalyst exhibited an extended continuous pore size distribution and a meso/macropore dominant hierarchical structure, which gave TiO2 a better supporting matrix and increased the number of catalytic sites. The N-doped carbon structure significantly narrowed the band gap of TiO2, increased the adsorption of visible light, and favored the electrons transferring. Above are the reasons for the improved photocatalytic performance of the material [166]. In another study, Wang et al. (2021) prepared metal-organic complex-derived 3D HPC-supported g-C3N4/TiO2 as photocatalysts for chlortetracycline hydrochloride (CTC-HCl) removal. It showed high CTC-HCl photocatalytic efficiency up to 97.8% after 60 min in the static system, which was ascribed to two structural properties: (1) the 3D porous structure with a high SSA could offer multidimensional adsorption-enrichment sites, (2) the heterojunction could promote the separation and migration of the photogenerated electrons and holes [167].

5. Conclusions and Outlooks

Along with human activities and industrial development, water pollution issues are becoming increasingly severe, threatening the ecological environment and human health. Due to the overuse of antibiotics in aquaculture and human/animal treatment in recent years, antibiotics are one of the major emerging pollutants in the aquatic environment. From one-dimensional carbon nanotubes, two-dimensional graphene, and three-dimensional biochar to hierarchical porous carbon, carbonaceous materials have been widely used as adsorbents and photocatalysts for eliminating antibiotics. This review discusses recent advances in carbon-based materials for antibiotic adsorption and photodegradation. The main adsorption mechanisms involved in the adsorptive removal process of antibiotics are π-π interactions, electrostatic interaction, and hydrogen bonding. Except for environmental parameters, the adsorption performance of carbon-based materials primarily depends on their pore size, uniformity, and functional groups. Carbonaceous materials have been modified via chemical or physical methods and loaded with functional nanoparticles or compounds to form carbon-based composites for improved adsorption properties. On the other hand, carbon-based materials have been applied in antibiotic photodegradation as photocatalyst carriers, co-catalysts, and photosensitizers. The predominant steps for antibiotic photocatalytic degradation are photon absorption, excitation, and reaction, depending on three main species, O2, h+, and OH radicals. The carbon-based photocatalysts usually show better adsorption performance, a higher ability to scavenge photo-excited electrons, and broadened light absorption range.
Even though many studies were conducted on the adsorptive and photocatalytic removal of antibiotics by carbon-based materials, many research gaps and uncertainties remain. The main challenges to overcome in future research are as follows.
  • Hastening the formation of antibiotic-resistance genes and antibiotic-resistant bacteria is the main pathway for antibiotics to harm the ecosystem. However, using carbon-based materials to disrupt antibiotic resistant genes and alter bacterial resistance has been rarely reported. Further research should be conducted to determine the impact of carbon-based materials on antibiotic resistance.
  • Some carbon-based materials have intrinsic toxicity, and others can produce toxic by-products when removing antibiotics, such as the leaching of metal ions. So, they may have a negative impact on the water environment. The long-term fate and environmental risks of carbon-based materials in the aqueous environment are still unclear, and further studies of their toxicity and biological responses are needed.
  • Despite intense research, antibiotics’ adsorption and photodegradation mechanisms remain vaguely interpreted, as conclusions based on partial characterization analysis and traditional models are not accurate or comprehensive. Issues like the synergistic effect of the components in the carbon-based nanocomposites, the effect of the antibiotic species on the removal properties, and the applicability of the relevant models need to be further investigated.
  • Currently, most studies on the removal of antibiotics by carbon-based materials are limited to batch experiments at the laboratory scale rather than the pilot scale, resulting in a gap between research and application. The experiments were usually conducted in mixed antibiotics or antibiotics-metals systems instead of a complex system with coexisted multi-pollutants. From an application point of view, it is vital to test the effectiveness of carbon-based materials for antibiotic removal in a system that resembles a natural water environment, and to investigate other pollutants’ influence on the removal process. More attention should be paid to interactive mechanisms among antibiotics, interfering substances, and carbon-based materials. Furthermore, for large-scale engineering applications, in addition to the removal properties of the materials, their mass production feasibility and economic efficiency should be considered, such as raw materials, production cost, production cycle, and yield.

Author Contributions

Conceptualization, R.M. and Y.X.; investigation, R.M., Y.X., Q.M., Y.C. and S.Y.; resources, J.F.; writing—original draft preparation, R.M.; writing—review and editing, R.M., Y.X. and J.F.; supervision, J.F.; funding acquisition, J.F. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (NSFC No. 51878472) and State Key Laboratory of Pollution Control and Resource Reuse Foundation (No. PCRRK 21003).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bolong, N.; Ismail, A.F.; Salim, M.R.; Matsuura, T. A review of the effects of emerging contaminants in wastewater and options for their removal. Desalination 2009, 239, 229–246. [Google Scholar] [CrossRef]
  2. Petrie, B.; Barden, R.; Kasprzyk-Hordern, B. A review on emerging contaminants in wastewaters and the environment: Current knowledge, understudied areas and recommendations for future monitoring. Water Res. 2015, 72, 3–27. [Google Scholar] [CrossRef] [PubMed]
  3. Rivera-Utrilla, J.; Sanchez-Polo, M.; Angeles Ferro-Garcia, M.; Prados-Joya, G.; Ocampo-Perez, R. Pharmaceuticals as emerging contaminants and their removal from water. A review. Chemosphere 2013, 93, 1268–1287. [Google Scholar] [CrossRef] [PubMed]
  4. Gavrilescu, M.; Demnerova, K.; Aamand, J.; Agathoss, S.; Fava, F. Emerging pollutants in the environment: Present and future challenges in biomonitoring, ecological risks and bioremediation. New Biotechnol. 2015, 32, 147–156. [Google Scholar] [CrossRef] [PubMed]
  5. Ahmed, M.B.; Zhou, J.L.; Huu Hao, N.; Guo, W.; Thomaidis, N.S.; Xu, J. Progress in the biological and chemical treatment technologies for emerging contaminant removal from wastewater: A critical review. J. Hazard. Mater. 2017, 323, 274–298. [Google Scholar] [CrossRef]
  6. Yang, Y.; Ok, Y.S.; Kim, K.-H.; Kwon, E.E.; Tsang, Y.F. Occurrences and removal of pharmaceuticals and personal care products (PPCPs) in drinking water and water/sewage treatment plants: A review. Sci. Total Environ. 2017, 596, 303–320. [Google Scholar] [CrossRef]
  7. Rodriguez-Narvaez, O.M.; Manuel Peralta-Hernandez, J.; Goonetilleke, A.; Bandala, E.R. Treatment technologies for emerging contaminants in water: A review. Chem. Eng. J. 2017, 323, 361–380. [Google Scholar] [CrossRef] [Green Version]
  8. Fu, J.; Zhao, Y.; Yao, Q.; Addo-Bankas, O.; Ji, B.; Yuan, Y.; Wei, T.; Esteve-Nunez, A. A review on antibiotics removal: Leveraging the combination of grey and green techniques. Sci. Total Environ. 2022, 838, 156427. [Google Scholar] [CrossRef]
  9. Bhattacharyya, P.; Basak, S.; Chakrabarti, S. Advancement towards Antibiotic Remediation: Heterostructure and Composite materials. Chemistryselect 2021, 6, 7323–7345. [Google Scholar] [CrossRef]
  10. Nguyen, L.M.; Nguyen, N.T.T.; Nguyen, T.T.T.; Nguyen, T.T.; Nguyen, D.T.C.; Tran, T.V. Occurrence, toxicity and adsorptive removal of the chloramphenicol antibiotic in water: A review. Environ. Chem. Lett. 2022, 20, 1929–1963. [Google Scholar] [CrossRef]
  11. Wu, S.; Lin, Y.; Hu, Y.H. Strategies of tuning catalysts for efficient photodegradation of antibiotics in water environments: A review. J. Mater. Chem. 2021, 9, 2592–2611. [Google Scholar] [CrossRef]
  12. Mutuku, C.; Gazdag, Z.; Melegh, S. Occurrence of antibiotics and bacterial resistance genes in wastewater: Resistance mechanisms and antimicrobial resistance control approaches. World J. Microbiol. Biotechnol. 2022, 38, 152. [Google Scholar] [CrossRef] [PubMed]
  13. Yu, C.; Pang, H.; Wang, J.-H.; Chi, Z.-Y.; Zhang, Q.; Kong, F.-T.; Xu, Y.-P.; Li, S.-Y.; Che, J. Occurrence of antibiotics in waters, removal by microalgae-based systems, and their toxicological effects: A review. Sci. Total Environ. 2022, 813, 151891. [Google Scholar] [CrossRef]
  14. Alegbeleye, O.; Daramola, O.B.; Adetunji, A.T.; Ore, O.T.; Ayantunji, Y.J.; Omole, R.K.; Ajagbe, D.; Adekoya, S.O. Efficient removal of antibiotics from water resourcchenes is a public health priority: A critical assessment of the efficacy of some remediation strategies for antibiotics in water. Environ. Sci. Pollut. Res. 2022, 29, 56948–57020. [Google Scholar] [CrossRef]
  15. Juela, D.M. Promising adsorptive materials derived from agricultural and industrial wastes for antibiotic removal: A comprehensive review. Sep. Purif. Technol. 2022, 284, 120286. [Google Scholar] [CrossRef]
  16. Ma, Q.; Sun, Y.; Zhang, C.; Xue, Y.; Chen, Y.; Teng, W.; Fan, J. Iron pyrophosphate doped carbon nanocomposite for tetracycline degradation by activation of peroxymonosulfate. New J. Chem. 2022, 46, 17985–17994. [Google Scholar] [CrossRef]
  17. Ahmed, M.B.; Zhou, J.L.; Ngo, H.H.; Guo, W. Adsorptive removal of antibiotics from water and wastewater: Progress and challenges. Sci. Total Environ. 2015, 532, 112–126. [Google Scholar] [CrossRef]
  18. Li, D.; Shi, W. Recent developments in visible-light photocatalytic degradation of antibiotics. Chin. J. Catal. 2016, 37, 792–799. [Google Scholar] [CrossRef]
  19. Yang, X.; Chen, Z.; Zhao, W.; Liu, C.; Qian, X.; Zhang, M.; Wei, G.; Khan, E.; Ng, Y.H.; Ok, Y.S. Recent advances in photodegradation of antibiotic residues in water. Chem. Eng. J. 2021, 405, 126806. [Google Scholar] [CrossRef] [PubMed]
  20. Tian, W.; Zhang, H.; Duan, X.; Sun, H.; Shao, G.; Wang, S. Porous Carbons: Structure-Oriented Design and Versatile Applications. Adv. Funct. Mater. 2020, 30, 1909265. [Google Scholar] [CrossRef]
  21. Xue, Y.; Teng, W.; Chen, Y.; Ma, Q.; Chen, X.; Sun, Y.; Fan, J.; Qiu, Y.; Fu, R. Amorphous Mn-La oxides immobilized on carbon sphere for efficient removal of As(V), Cd(II), and Pb(II): Co-adsorption and roles of Mn species. Chem. Eng. J. 2022, 429, 132262. [Google Scholar] [CrossRef]
  22. Viglasova, E.; Galambos, M.; Divis, D.; Dankova, Z.; Dano, M.; Krivosudsky, L.; Lengauer, C.L.; Matik, M.; Briancin, J.; Soja, G. Engineered biochar as a tool for nitrogen pollutants removal: Preparation, characterization and sorption study. Desalin. Water Treat. 2020, 191, 318–331. [Google Scholar] [CrossRef]
  23. Dano, M.; Viglasova, E.; Stamberg, K.; Galambos, M.; Galanda, D. Pertechnetate/Perrhenate Surface Complexation on Bamboo Engineered Biochar. Materials 2021, 14, 486. [Google Scholar] [CrossRef] [PubMed]
  24. Jilani, A.; Othman, M.H.D.; Ansari, M.O.; Hussain, S.Z.; Ismail, A.F.; Khan, I.U.; Inamuddin. Graphene and its derivatives: Synthesis, modifications, and applications in wastewater treatment. Environ. Chem. Lett. 2018, 16, 1301–1323. [Google Scholar] [CrossRef]
  25. Ali, I.; Basheer, A.A.; Mbianda, X.Y.; Burakov, A.; Galunin, E.; Burakova, I.; Mkrtchyan, E.; Tkachev, A.; Grachev, V. Graphene based adsorbents for remediation of noxious pollutants from wastewater. Environ. Int. 2019, 127, 160–180. [Google Scholar] [CrossRef]
  26. Nidheesh, P.V. Graphene-based materials supported advanced oxidation processes for water and wastewater treatment: A review. Environ. Sci. Pollut. Res. 2017, 24, 27047–27069. [Google Scholar] [CrossRef] [PubMed]
  27. Aslam, M.M.-A.; Kuo, H.-W.; Den, W.; Usman, M.; Sultan, M.; Ashraf, H. Functionalized Carbon Nanotubes (CNTs) for Water and Wastewater Treatment: Preparation to Application. Sustainability 2021, 13, 5717. [Google Scholar] [CrossRef]
  28. Mashkoor, F.; Nasar, A.; Inamuddin. Carbon nanotube-based adsorbents for the removal of dyes from waters: A review. Environ. Chem. Lett. 2020, 18, 605–629. [Google Scholar] [CrossRef]
  29. Ouni, L.; Ramazani, A.; Fardood, S.T. An overview of carbon nanotubes role in heavy metals removal from wastewater. Front. Chem. Sci. Eng. 2019, 13, 274–295. [Google Scholar] [CrossRef]
  30. Liu, J.; Jiang, J.; Meng, Y.; Aihemaiti, A.; Xu, Y.; Xiang, H.; Gao, Y.; Chen, X. Preparation, environmental application and prospect of biochar-supported metal nanoparticles: A review. J. Hazard. Mater. 2020, 388, 122026. [Google Scholar] [CrossRef]
  31. Weidner, E.; Karbassiyazdi, E.; Altaee, A.; Jesionowski, T.; Ciesielczyk, F. Hybrid Metal Oxide/Biochar Materials for Wastewater Treatment Technology: A Review. ACS Omega 2022, 7, 27062–27078. [Google Scholar] [CrossRef]
  32. Sabzehmeidani, M.M.; Mahnaee, S.; Ghaedi, M.; Heidari, H.; Roy, V.A.L. Carbon based materials: A review of adsorbents for inorganic and organic compounds. Mater. Adv. 2021, 2, 598–627. [Google Scholar] [CrossRef]
  33. Leng, L.; Xiong, Q.; Yang, L.; Li, H.; Zhou, Y.; Zhang, W.; Jiang, S.; Li, H.; Huang, H. An overview on engineering the surface area and porosity of biochar. Sci. Total Environ. 2021, 763, 144204. [Google Scholar] [CrossRef]
  34. Enaime, G.; Bacaoui, A.; Yaacoubi, A.; Luebken, M. Biochar for Wastewater Treatment-Conversion Technologies and Applications. Appl. Sci. 2020, 10, 3492. [Google Scholar] [CrossRef]
  35. Krasucka, P.; Pan, B.; Ok, Y.S.; Mohan, D.; Sarkar, B.; Oleszczuk, P. Engineered biochar—A sustainable solution for the removal of antibiotics from water. Chem. Eng. J. 2021, 405, 126926. [Google Scholar] [CrossRef]
  36. Premarathna, K.S.D.; Rajapaksha, A.U.; Sarkar, B.; Kwon, E.E.; Bhatnagar, A.; Ok, Y.S.; Vithanage, M. Biochar-based engineered composites for sorptive decontamination of water: A review. Chem. Eng. J. 2019, 372, 536–550. [Google Scholar] [CrossRef]
  37. Zhou, X.-L.; Zhang, H.; Shao, L.-M.; Lu, F.; He, P.-J. Preparation and Application of Hierarchical Porous Carbon Materials from Waste and Biomass: A Review. Waste Biomass Valorization 2021, 12, 1699–1724. [Google Scholar] [CrossRef]
  38. Liu, C.; Li, Q.; Kang, W.; Lei, W.; Wang, X.; Lu, C.; Naebe, M. Structural design and mechanism analysis of hierarchical porous carbon fibers for advanced energy and environmental applications. J. Mater. Chem. 2021, 10, 10–49. [Google Scholar] [CrossRef]
  39. Zhang, M.; Igalavithana, A.D.; Xu, L.; Sarkar, B.; Hou, D.; Zhang, M.; Bhatnagar, A.; Cho, W.C.; Ok, Y.S. Engineered/designer hierarchical porous carbon materials for organic pollutant removal from water and wastewater: A critical review. Crit. Rev. Environ. Sci. Technol. 2021, 51, 2295–2328. [Google Scholar] [CrossRef]
  40. Xue, Y.; Yu, Q.; Ma, Q.; Chen, Y.; Zhang, C.; Teng, W.; Fan, J.; Zhang, W.-X. Electrocatalytic Hydrogenation Boosts Reduction of Nitrate to Ammonia over Single-Atom Cu with Cu(I)-N3C1 Sites. Environ. Sci. Technol. 2022, 56, 14797–14807. [Google Scholar] [CrossRef] [PubMed]
  41. Liu, T.; Zhang, F.; Song, Y.; Li, Y. Revitalizing carbon supercapacitor electrodes with hierarchical porous structures. J. Mater. Chem. 2017, 5, 17705–17733. [Google Scholar] [CrossRef]
  42. Zhang, X.; Elsayed, I.; Nayanathara, R.M.O.; Song, X.; Shmulsky, R.; Hassan, E.B. Biobased hierarchically porous carbon featuring micron-sized honeycomb architecture for CO2 capture and water remediation. J. Environ. Chem. Eng. 2022, 10, 107460. [Google Scholar] [CrossRef]
  43. Zhu, S.; Xu, J.; Wang, B.; Xie, J.; Ying, G.; Li, J.; Cheng, Z.; Li, J.; Chen, K. Highly efficient and rapid purification of organic dye wastewater using lignin-derived hierarchical porous carbon. J. Colloid Interface Sci. 2022, 625, 158–168. [Google Scholar] [CrossRef]
  44. Wang, G.; Yong, X.; Luo, L.; Yan, S.; Wong, J.W.C.; Zhou, J. Structure-performance correlation of high surface area and hierarchical porous biochars as chloramphenicol adsorbents. Sep. Purif. Technol. 2022, 296, 121374. [Google Scholar] [CrossRef]
  45. Pi, Z.; Hou, K.; Yao, F.; He, L.; Chen, S.; Tao, Z.; Zhou, P.; Wang, D.; Li, X.; Yang, Q. In-situ regeneration of tetracycline-saturated hierarchical porous carbon by peroxydisulfate oxidation process: Performance, mechanism and application. Chem. Eng. J. 2022, 427, 131749. [Google Scholar] [CrossRef]
  46. Wang, T.; Zhang, H.; Liu, Y.; Zhang, L.; Xing, B. Ultrathin porous carbon nanosheet as an efficient adsorbent for the removal of bisphenol A: The overlooked role of topological defects. Chemosphere 2022, 306, 135549. [Google Scholar] [CrossRef] [PubMed]
  47. Shi, W.; Wang, H.; Yan, J.; Shan, L.; Quan, G.; Pan, X.; Cui, L. Wheat straw derived biochar with hierarchically porous structure for bisphenol A removal: Preparation, characterization, and adsorption properties. Sep. Purif. Technol. 2022, 289, 120796. [Google Scholar] [CrossRef]
  48. Ma, C.-F.; Gao, Q.; Xia, K.-S.; Huang, Z.-Y.; Han, B.; Zhou, C.-G. Three-dimensionally porous graphene: A high-performance adsorbent for removal of albumin-bonded bilirubin. Colloids Surf. B-Biointerfaces 2017, 149, 146–153. [Google Scholar] [CrossRef]
  49. Peiris, C.; Gunatilake, S.R.; Mlsna, T.E.; Mohan, D.; Vithanage, M. Biochar based removal of antibiotic sulfonamides and tetracyclines in aquatic environments: A critical review. Bioresour. Technol. 2017, 246, 150–159. [Google Scholar] [CrossRef]
  50. Mangla, D.; Annu; Sharma, A.; Ikram, S. Critical review on adsorptive removal of antibiotics: Present situation, challenges and future perspective. J. Hazard. Mater. 2022, 425, 127946. [Google Scholar] [CrossRef]
  51. Xiang, Y.; Xu, Z.; Wei, Y.; Zhou, Y.; Yang, X.; Yang, Y.; Yang, J.; Zhang, J.; Luo, L.; Zhou, Z. Carbon-based materials as adsorbent for antibiotics removal: Mechanisms and influencing factors. J. Environ. Manag. 2019, 237, 128–138. [Google Scholar] [CrossRef]
  52. Yu, F.; Li, Y.; Han, S.; Ma, J. Adsorptive removal of antibiotics from aqueous solution using carbon materials. Chemosphere 2016, 153, 365–385. [Google Scholar] [CrossRef]
  53. Li, M.-F.; Liu, Y.-G.; Zeng, G.-M.; Liu, N.; Liu, S.-B. Graphene and graphene-based nanocomposites used for antibiotics removal in water treatment: A review. Chemosphere 2019, 226, 360–380. [Google Scholar] [CrossRef] [PubMed]
  54. Chen, H.; Gao, B.; Li, H. Removal of sulfamethoxazole and ciprofloxacin from aqueous solutions by graphene oxide. J. Hazard. Mater. 2015, 282, 201–207. [Google Scholar] [CrossRef]
  55. Gao, Y.; Li, Y.; Zhang, L.; Huang, H.; Hu, J.; Shah, S.M.; Su, X. Adsorption and removal of tetracycline antibiotics from aqueous solution by graphene oxide. J. Colloid Interface Sci. 2012, 368, 540–546. [Google Scholar] [CrossRef]
  56. Salihi, E.C.; Wang, J.; Kabacaoglu, G.; Kirkulak, S.; Siller, L. Graphene oxide as a new generation adsorbent for the removal of antibiotics from waters. Sep. Sci. Technol. 2021, 56, 453–461. [Google Scholar] [CrossRef]
  57. Moreira, V.R.; Rocha Lebron, Y.A.; da Silva, M.M.; de Souza Santos, L.V.; Jacob, R.S.; Barbosa de Vasconcelos, C.K.; Viana, M.M. Graphene oxide in the remediation of norfloxacin from aqueous matrix: Simultaneous adsorption and degradation process. Environ. Sci. Pollut. Res. 2020, 27, 34513–34528. [Google Scholar] [CrossRef]
  58. Wernke, G.; Shimabuku-Biadola, Q.L.; Dos Santos, T.R.T.; Silva, M.F.; Fagundes-Klen, M.R.; Bergamasco, R. Adsorption of cephalexin in aqueous media by graphene oxide: Kinetics, isotherm, and thermodynamics. Environ. Sci. Pollut. Res. 2020, 27, 4725–4736. [Google Scholar] [CrossRef] [PubMed]
  59. Upoma, B.P.; Yasmin, S.; Shaikh, M.A.A.; Jahan, T.; Haque, M.A.; Moniruzzaman, M.; Kabir, M.H. A Fast Adsorption of Azithromycin on Waste-Product-Derived Graphene Oxide Induced by H-Bonding and Electrostatic Interactions. ACS Omega 2022, 7, 29655–29665. [Google Scholar] [CrossRef]
  60. Liu, F.-F.; Zhao, J.; Wang, S.; Xing, B. Adsorption of sulfonamides on reduced graphene oxides as affected by pH and dissolved organic matter. Environ. Pollut. 2016, 210, 85–93. [Google Scholar] [CrossRef]
  61. El Hadki, A.; Ulucan-Altuntas, K.; El Hadki, H.; Ustundag, C.B.; Kabbaj, O.K.; Dahchour, A.; Komiha, N.; Zrineh, A.; Debik, E. Removal of oxytetracycline by graphene oxide and Boron-doped reduced graphene oxide: A combined density function Theory, molecular dynamics simulation and experimental study. Flatchem 2021, 27, 100238. [Google Scholar] [CrossRef]
  62. Yu, H.; Zheng, K.; Xu, X.; Liu, X.; Zhao, B.; Ding, H.; Yu, Z.; Deng, C. Preparation of beta-cyclodextrin/dopamine hydrochloride-graphene oxide and its adsorption properties for sulfonamide antibiotics. Environ. Sci. Pollut. Res. 2022, 29, 70192–70201. [Google Scholar] [CrossRef] [PubMed]
  63. Braki, Z.A.; Sohrabi, M.R.; Motiee, F. Efficient simultaneous removal of tetracycline and cefazolin from aqueous solution using a novel adsorbent based on zero-valent iron nanoparticles supported by montmorillonite and graphene oxide: Isotherms, kinetic, and thermodynamic studies. Int. J. Environ. Anal. Chem. 2022, 1–23. [Google Scholar] [CrossRef]
  64. Lin, J.; Liu, Y.; Hu, S.; Zhang, Y.; Qian, C.; Li, A.; Zhang, S. Ultra-fast adsorption of four typical pollutants using magnetically separable ethanolamine-functionalized graphene. Sep. Purif. Technol. 2021, 271, 118862. [Google Scholar] [CrossRef]
  65. Shen, J.-W.; Li, J.; Dai, J.; Zhou, M.; Ren, H.; Zhang, L.; Hu, Q.; Kong, Z.; Liang, L. Molecular dynamics study on the adsorption and release of doxorubicin by chitosan-decorated graphene. Carbohydr. Polym. 2020, 248, 116809. [Google Scholar] [CrossRef] [PubMed]
  66. Xiaosan, S.; Boyang, S.; Yiru, W.; Jie, Z.; Sanfan, W.; Nan, W. Adsorption performance of GO-doped activated ATP composites towards tetracycline. RSC Adv. 2022, 12, 19917–19928. [Google Scholar] [CrossRef]
  67. Wang, T.; Xue, L.; Liu, Y.; Fang, T.; Zhang, L.; Xing, B. Ring defects-rich and pyridinic N-doped graphene aerogel as floating adsorbent for efficient removal of tetracycline: Evidence from NEXAFS measurements and theoretical calculations. J. Hazard. Mater. 2022, 435, 128940. [Google Scholar] [CrossRef] [PubMed]
  68. Feng, Y.; Chen, G.; Zhang, Y.; Li, D.; Ling, C.; Wang, Q.; Liu, G. Superhigh co-adsorption of tetracycline and copper by the ultrathin g-C3N4 modified graphene oxide hydrogels. J. Hazard. Mater. 2022, 424, 127362. [Google Scholar] [CrossRef]
  69. Chen, B.; Li, Y.; Du, Q.; Pi, X.; Wang, Y.; Sun, Y.; Wang, M.; Zhang, Y.; Chen, K.; Zhu, J. Effective Removal of Tetracycline from Water Using Copper Alginate @ Graphene Oxide with in-situ Grown MOF-525 Composite: Synthesis, Characterization and Adsorption Mechanisms. Nanomaterials 2022, 12, 2897. [Google Scholar] [CrossRef]
  70. Ghafoori, M.; Cheraghi, M.; Sadr, M.K.; Lorestani, B.; Sobhanardakani, S. Magnetite graphene oxide modified with beta-cyclodextrin as an effective adsorbent for the removal of methotrexate and doxorubicin hydrochloride from water. Environ. Sci. Pollut. Res. 2022, 29, 35012–35024. [Google Scholar] [CrossRef]
  71. Zhao, H.; Liu, X.; Cao, Z.; Zhan, Y.; Shi, X.; Yang, Y.; Zhou, J.; Xu, J. Adsorption behavior and mechanism of chloramphenicols, sulfonamides, and non-antibiotic pharmaceuticals on multi-walled carbon nanotubes. J. Hazard. Mater. 2016, 310, 235–245. [Google Scholar] [CrossRef] [PubMed]
  72. Balarak, D.; Mostafapour, F.; Bazrafshan, E.; Saleh, T.A. Studies on the adsorption of amoxicillin on multi-wall carbon nanotubes. Water Sci. Technol. 2017, 75, 1599–1606. [Google Scholar] [CrossRef] [PubMed]
  73. Chen, Z.; Han, R.; Zhong, C. Adsorption of fluoroquinolone by carbon nanotubes: A combined experimental and density functional theory study. Chem. Pap. 2020, 74, 3847–3856. [Google Scholar] [CrossRef]
  74. Veclani, D.; Melchior, A. Adsorption of ciprofloxacin on carbon nanotubes: Insights from molecular dynamics simulations. J. Mol. Liq. 2020, 298, 111977. [Google Scholar] [CrossRef]
  75. Carrales-Alvarado, D.H.; Leyva-Ramos, R.; Rodriguez-Ramos, I.; Mendoza-Mendoza, E.; Moral-Rodriguez, A.E. Adsorption capacity of different types of carbon nanotubes towards metronidazole and dimetridazole antibiotics from aqueous solutions: Effect of morphology and surface chemistry. Environ. Sci. Pollut. Res. 2020, 27, 17123–17137. [Google Scholar] [CrossRef] [PubMed]
  76. Zhao, W.; Tian, Y.; Chu, X.; Cui, L.; Zhang, H.; Li, M.; Zhao, P. Preparation and characteristics of a magnetic carbon nanotube adsorbent: Its efficient adsorption and recoverable performances. Sep. Purif. Technol. 2021, 257, 117917. [Google Scholar] [CrossRef]
  77. Zhuang, S.; Zhu, X.; Wang, J. Adsorptive removal of plasticizer (dimethyl phthalate) and antibiotic (sulfamethazine) from municipal wastewater by magnetic carbon nanotubes. J. Mol. Liq. 2020, 319, 114267. [Google Scholar] [CrossRef]
  78. Sereshti, H.; Beyrak-Abadi, E.; Bidhendi, M.E.; Ahmad, I.; Shahabuddin, S.; Nodeh, H.R.; Sridewi, N.; Ibrahim, W.N.W. Sulfide-Doped Magnetic Carbon Nanotubes Developed as Adsorbent for Uptake of Tetracycline and Cefixime from Wastewater. Nanomaterials 2022, 12, 3576. [Google Scholar] [CrossRef]
  79. Yao, J.; Deng, Y.; Pan, S.; Korna, R.; Wen, J.; Yuan, N.; Wang, K.; Li, H.; Yang, Y. The difference in the adsorption mechanisms of magnetic ferrites modified carbon nanotubes. J. Hazard. Mater. 2021, 415, 125551. [Google Scholar] [CrossRef]
  80. Xia, C.; Huang, H.; Liang, D.; Xie, Y.; Kong, F.; Yang, Q.; Fu, J.; Dou, Z.; Zhang, Q.; Meng, Z. Adsorption of tetracycline hydrochloride on layered double hydroxide loaded carbon nanotubes and site energy distribution analysis. Chem. Eng. J. 2022, 443, 136398. [Google Scholar] [CrossRef]
  81. Xiong, W.; Zeng, G.; Yang, Z.; Zhou, Y.; Zhang, C.; Cheng, M.; Liu, Y.; Hu, L.; Wan, J.; Zhou, C.; et al. Adsorption of tetracycline antibiotics from aqueous solutions on nanocomposite multi-walled carbon nanotube functionalized MIL-53 (Fe) as new adsorbent. Sci. Total Environ. 2018, 627, 235–244. [Google Scholar] [CrossRef] [PubMed]
  82. Li, J.; Tian, T.; Jia, Y.; Xu, N.; Yang, S.; Zhang, C.; Gao, S.; Shen, W.; Wang, Z. Adsorption performance and optimization by response surface methodology on tetracycline using Fe-doped ZIF-8-loaded multi-walled carbon nanotubes. Environ. Sci. Pollut. Res. 2022. [Google Scholar] [CrossRef] [PubMed]
  83. Ma, J.; Jiang, Z.; Cao, J.; Yu, F. Enhanced adsorption for the removal of antibiotics by carbon nanotubes/graphene oxide/sodium alginate triple-network nanocomposite hydrogels in aqueous solutions. Chemosphere 2020, 242, 125188. [Google Scholar] [CrossRef] [PubMed]
  84. Yu, F.; Sun, S.; Han, S.; Zheng, J.; Ma, J. Adsorption removal of ciprofloxacin by multi-walled carbon nanotubes with different oxygen contents from aqueous solutions. Chem. Eng. J. 2016, 285, 588–595. [Google Scholar] [CrossRef]
  85. Dai, Y.; Li, J.; Shan, D. Adsorption of tetracycline in aqueous solution by biochar derived from waste Auricularia auricula dregs. Chemosphere 2020, 238, 124432. [Google Scholar] [CrossRef]
  86. Zhang, P.; Li, Y.; Cao, Y.; Han, L. Characteristics of tetracycline adsorption by cow manure biochar prepared at different pyrolysis temperatures. Bioresour. Technol. 2019, 285, 121348. [Google Scholar] [CrossRef]
  87. Stylianou, M.; Christou, A.; Michael, C.; Agapiou, A.; Papanastasiou, P.; Fatta-Kassinos, D. Adsorption and removal of seven antibiotic compounds present in water with the use of biochar derived from the pyrolysis of organic waste feedstocks. J. Environ. Chem. Eng. 2021, 9, 105868. [Google Scholar] [CrossRef]
  88. Zhao, Y.; Liu, X.; Li, W.; Huang, K.; Shao, H.; Qu, C.; Liu, J. One-step synthesis of garlic peel derived biochar by concentrated sulfuric acid: Enhanced adsorption capacities for Enrofloxacin and interfacial interaction mechanisms. Chemosphere 2022, 290, 133263. [Google Scholar] [CrossRef]
  89. Wu, J.; Wang, T.; Liu, Y.; Tang, W.; Geng, S.; Chen, J. Norfloxacin adsorption and subsequent degradation on ball-milling tailored N-doped biochar. Chemosphere 2022, 303, 135264. [Google Scholar] [CrossRef]
  90. Zeng, X.-Y.; Wang, Y.; Li, R.-X.; Cao, H.-L.; Li, Y.-F.; Lu, J. Impacts of temperatures and phosphoric-acid modification to the physicochemical properties of biochar for excellent sulfadiazine adsorption. Biochar 2022, 4, 14. [Google Scholar] [CrossRef]
  91. Liu, Q.; Li, D.; Cheng, H.; Cheng, J.; Du, K.; Hu, Y.; Chen, Y. High mesoporosity phosphorus-containing biochar fabricated from Camellia oleifera shells: Impressive tetracycline adsorption performance and promotion of pyrophosphate-like surface functional groups (C-O-P bond). Bioresour. Technol. 2021, 329, 124922. [Google Scholar] [CrossRef] [PubMed]
  92. Sheng, X.; Wang, J.; Cui, Q.; Zhang, W.; Zhu, X. A feasible biochar derived from biogas residue and its application in the efficient adsorption of tetracycline from an aqueous solution. Environ. Res. 2022, 207, 112175. [Google Scholar] [CrossRef] [PubMed]
  93. Yan, L.; Liu, Y.; Zhang, Y.; Liu, S.; Wang, C.; Chen, W.; Liu, C.; Chen, Z.; Zhang, Y. ZnCl2 modified biochar derived from aerobic granular sludge for developed microporosity and enhanced adsorption to tetracycline. Bioresour. Technol. 2020, 297, 122381. [Google Scholar] [CrossRef] [PubMed]
  94. Wang, H.; Lou, X.; Hu, Q.; Sun, T. Adsorption of antibiotics from water by using Chinese herbal medicine residues derived biochar: Preparation and properties studies. J. Mol. Liq. 2021, 325, 114967. [Google Scholar] [CrossRef]
  95. Zou, M.; Tian, W.; Chu, M.; Gao, H.; Zhang, D. Biochar composite derived from cellulase hydrolysis apple branch for quinolone antibiotics enhanced removal: Precursor pyrolysis performance, functional group introduction and adsorption mechanisms. Environ. Pollut. 2022, 313, 120104. [Google Scholar] [CrossRef]
  96. Li, X.; Shi, J. Simultaneous adsorption of tetracycline, ammonium and phosphate from wastewater by iron and nitrogen modified biochar: Kinetics, isotherm, thermodynamic and mechanism. Chemosphere 2022, 293, 133574. [Google Scholar] [CrossRef]
  97. Yao, B.; Luo, Z.; Du, S.; Yang, J.; Zhi, D.; Zhou, Y. Sustainable biochar/MgFe(2)O4 adsorbent for levofloxacin removal: Adsorption performances and mechanisms. Bioresour. Technol. 2021, 340, 125698. [Google Scholar] [CrossRef]
  98. Xu, Z.; Xiang, Y.; Zhou, H.; Yang, J.; He, Y.; Zhu, Z.; Zhou, Y. Manganese ferrite modified biochar from vinasse for enhanced adsorption of levofloxacin: Effects and mechanisms. Environ. Pollut. 2021, 272, 115968. [Google Scholar] [CrossRef]
  99. Wang, Q.; Zhang, Z.; Xu, G.; Li, G. Magnetic porous biochar with nanostructure surface derived from penicillin fermentation dregs pyrolysis with K2FeO4 activation: Characterization and application in penicillin adsorption. Bioresour. Technol. 2021, 327, 124818. [Google Scholar] [CrossRef]
  100. Yan, J.; Zuo, X.; Yang, S.; Chen, R.; Cai, T.; Ding, D. Evaluation of potassium ferrate activated biochar for the simultaneous adsorption of copper and sulfadiazine: Competitive versus synergistic. J. Hazard. Mater. 2022, 424, 127435. [Google Scholar] [CrossRef]
  101. Qu, J.; Wang, S.; Jin, L.; Liu, Y.; Yin, R.; Jiang, Z.; Tao, Y.; Huang, J.; Zhang, Y. Magnetic porous biochar with high specific surface area derived from microwave-assisted hydrothermal and pyrolysis treatments of water hyacinth for Cr(VI) and tetracycline adsorption from water. Bioresour. Technol. 2021, 340, 125692. [Google Scholar] [CrossRef] [PubMed]
  102. Cao, X.; Meng, Z.; Song, E.; Sun, X.; Hu, X.; Li, W.; Liu, Z.; Gao, S.; Song, B. Co-adsorption capabilities and mechanisms of bentonite enhanced sludge biochar for de-risking norfloxacin and Cu2+contaminated water. Chemosphere 2022, 299, 134414. [Google Scholar] [CrossRef]
  103. Arif, M.; Liu, G.; Rehman, M.Z.U.; Yousaf, B.; Ahmed, R.; Miana, M.M.; Ashrafa, A.; Munirf, M.A.M.; Rashida, M.S.; Naeem, A. Carbon dioxide activated biochar-clay mineral composite efficiently removes ciprofloxacin from contaminated water—Reveals an incubation study. J. Clean. Prod. 2022, 332, 130079. [Google Scholar] [CrossRef]
  104. Xiang, W.; Zhang, X.; Luo, J.; Li, Y.; Guo, T.; Gao, B. Performance of lignin impregnated biochar on tetracycline hydrochloride adsorption: Governing factors and mechanisms. Environ. Res. 2022, 215, 114339. [Google Scholar] [CrossRef] [PubMed]
  105. Xu, J.; Zhang, Y.; Li, B.; Fan, S.; Xu, H.; Guan, D.-X. Improved adsorption properties of tetracycline on KOH/KMnO4 modified biochar derived from wheat straw. Chemosphere 2022, 296, 133981. [Google Scholar] [CrossRef] [PubMed]
  106. Wang, K.; Wang, Y.; Zhang, S.; Chen, Y.-D.; Wang, R.; Ho, S.-H. Tailoring a novel hierarchical cheese-like porous biochar from algae residue to boost sulfathiazole removal. Environ. Sci. Ecotechnol. 2022, 10, 100168. [Google Scholar] [CrossRef] [PubMed]
  107. Li, X.; Xu, J.; Shi, J.; Luo, X. Rapid and efficient adsorption of tetracycline from aqueous solution in a wide pH range by using iron and aminoacetic acid sequentially modified hierarchical porous biochar. Bioresour. Technol. 2022, 346, 126672. [Google Scholar] [CrossRef]
  108. Chen, A.; Pang, J.; Wei, X.; Chen, B.; Xie, Y. Fast one-step preparation of porous carbon with hierarchical oxygen-enriched structure from waste lignin for chloramphenicol removal. Environ. Sci. Pollut. Res. 2021, 28, 27398–27410. [Google Scholar] [CrossRef]
  109. Xie, A.; Dai, J.; Chen, X.; Ma, P.; He, J.; Li, C.; Zhou, Z.; Yan, Y. Ultrahigh adsorption of typical antibiotics onto novel hierarchical porous carbons derived from renewable lignin via halloysite nanotubes-template and in-situ activation. Chem. Eng. J. 2016, 304, 609–620. [Google Scholar] [CrossRef]
  110. Zhang, R.; Zhou, Z.; Xie, A.; Dai, J.; Cui, J.; Lang, J.; Wei, M.; Dai, X.; Li, C.; Yan, Y. Preparation of hierarchical porous carbons from sodium carboxymethyl cellulose via halloysite template strategy coupled with KOH-activation for efficient removal of chloramphenicol. J. Taiwan Inst. Chem. Eng. 2017, 80, 424–433. [Google Scholar] [CrossRef]
  111. Zhang, M.; Xu, L.; Qi, C.; Zhang, M. Highly Effective Removal of Tetracycline from Water by Hierarchical Porous Carbon: Batch and Column Adsorption. Ind. Eng. Chem. Res. 2019, 58, 20036–20046. [Google Scholar] [CrossRef]
  112. Zhang, Z.; Chen, Y.; Wang, P.; Wang, Z.; Zuo, C.; Chen, W.; Ao, T. Facile fabrication of N-doped hierarchical porous carbons derived from soft-templated ZIF-8 for enhanced adsorptive removal of tetracycline hydrochloride from water. J. Hazard. Mater. 2022, 423, 127103. [Google Scholar] [CrossRef] [PubMed]
  113. Bai, X.; Chen, W.; Wang, B.; Sun, T.; Wu, B.; Wang, Y. Photocatalytic Degradation of Some Typical Antibiotics: Recent Advances and Future Outlooks. Int. J. Mol. Sci. 2022, 23, 8130. [Google Scholar] [CrossRef] [PubMed]
  114. Kuvarega, A.T.; Mamba, B.B. TiO2-based Photocatalysis: Toward Visible Light-Responsive Photocatalysts Through Doping and Fabrication of Carbon-Based Nanocomposites. Crit. Rev. Solid State Mater. Sci. 2017, 42, 295–346. [Google Scholar] [CrossRef]
  115. Dai, L.; Sun, F.; Fan, Q.; Li, H.; Yang, K.; Guo, T.; Zheng, L.; Fu, P. Carbon-based titanium dioxide materials for hydrogen production in water-methanol reforming: A review. J. Environ. Chem. Eng. 2022, 10, 107326. [Google Scholar] [CrossRef]
  116. Jayaraman, T.; Murthy, A.P.; Elakkiya, V.; Chandrasekaran, S.; Nithyadharseni, P.; Khan, Z.; Senthil, R.A.; Shanker, R.; Raghavender, M.; Kuppusami, P.; et al. Recent development on carbon based heterostructures for their applications in energy and environment: A review. J. Ind. Eng. Chem. 2018, 64, 16–59. [Google Scholar] [CrossRef]
  117. Zhao, W.; Duan, J.; Ji, B.; Ma, L.; Yang, Z. Novel formation of large area N-TiO2/graphene layered materials and enhanced photocatalytic degradation of antibiotics. J. Environ. Chem. Eng. 2020, 8, 102206. [Google Scholar] [CrossRef]
  118. Balarak, D.; Mengelizadeh, N.; Rajiv, P.; Chandrika, K. Photocatalytic degradation of amoxicillin from aqueous solutions by titanium dioxide nanoparticles loaded on graphene oxide. Environ. Sci. Pollut. Res. 2021, 28, 49743–49754. [Google Scholar] [CrossRef]
  119. Karaolia, P.; Michael-Kordatou, I.; Hapeshi, E.; Drosou, C.; Bertakis, Y.; Christofilos, D.; Armatas, G.S.; Sygellou, L.; Schwartz, T.; Xekoukoulotakis, N.P.; et al. Removal of antibiotics, antibiotic-resistant bacteria and their associated genes by graphene-based TiO2 composite photocatalysts under solar radiation in urban wastewaters. Appl. Catal. B-Environ. 2018, 224, 810–824. [Google Scholar] [CrossRef]
  120. Zhao, Y.; Liang, X.; Hu, X.; Fan, J. rGO/Bi2WO6 composite as a highly efficient and stable visible-light photocatalyst for norfloxacin degradation in aqueous environment. J. Colloid Interface Sci. 2021, 589, 336–346. [Google Scholar] [CrossRef]
  121. Zhao, G.; Ding, J.; Zhou, F.; Zhao, Q.; Wang, K.; Chen, X.; Gao, Q. Insight into a novel microwave-assisted W doped BiVO4 self-assembled sphere with rich oxygen vacancies oriented on rGO (W-BiVO4-x/rGO) photocatalyst for efficient contaminants removal. Sep. Purif. Technol. 2021, 277, 119610. [Google Scholar] [CrossRef]
  122. Ma, C.; Seo, W.C.; Lee, J.; Kim, Y.; Jung, H.; Yang, W. Construction of quantum dots self-decorated BiVO4/reduced graphene hydrogel composite photocatalyst with improved photocatalytic performance for antibiotics degradation. Chemosphere 2021, 275, 130052. [Google Scholar] [CrossRef] [PubMed]
  123. Jiang, H.; Wang, Q.; Chen, P.; Zheng, H.; Shi, J.; Shu, H.; Liu, Y. Photocatalytic degradation of tetracycline by using a regenerable (Bi) BiOBr/rGO composite. J. Clean. Prod. 2022, 339, 130771. [Google Scholar] [CrossRef]
  124. Xu, M.; Wang, Y.; Ha, E.; Zhang, H.; Li, C. Reduced graphene oxide/Bi4O5Br2 nanocomposite with synergetic effects on improving adsorption and photocatalytic activity for the degradation of antibiotics. Chemosphere 2021, 265, 129013. [Google Scholar] [CrossRef] [PubMed]
  125. Wang, J.; Zang, L.; Wang, L.; Tian, Y.; Yang, Z.; Yue, Y.; Sun, L. Magnetic cobalt ferrite/reduced graphene oxide (CF/rGO) porous balls for efficient photocatalytic degradation of oxytetracycline. J. Environ. Chem. Eng. 2022, 10, 108259. [Google Scholar] [CrossRef]
  126. Shibu, M.C.; Benoy, M.D.; Shanavas, S.; Abu Haija, M.; Duraimurugan, J.; Kumar, G.S.; Ahamad, T.; Maadeswaran, P.; Van Le, Q. White LED active—Fe2O3/rGO photocatalytic nanocomposite for an effective degradation of tetracycline and ibuprofen molecules. Environ. Res. 2022, 212, 113301. [Google Scholar] [CrossRef]
  127. Yashas, S.R.; Shivaraju, H.P.; McKay, G.; Shahmoradi, B.; Maleki, A.; Yetilmezsoy, K. Designing bi-functional silver delafossite bridged graphene oxide interfaces: Insights into synthesis, characterization, photocatalysis and bactericidal efficiency. Chem. Eng. J. 2021, 426, 131729. [Google Scholar] [CrossRef]
  128. Wang, Z.; Wang, Y.; Huang, L.; Liu, X.; Han, Y.; Wang, L. La2Zr2O7/rGO synthesized by one-step sol-gel method for photocatalytic degradation of tetracycline under visible-light. Chem. Eng. J. 2020, 384, 123380. [Google Scholar] [CrossRef]
  129. Gao, Z.; Liang, J.; Yao, J.; Zhao, Y.; Meng, Q.; He, G.; Chen, H. Synthesis of Ce-doped NiAl LDH/RGO composite as an efficient photocatalyst for photocatalytic degradation of ciprofloxacin. J. Environ. Chem. Eng. 2021, 9, 105405. [Google Scholar] [CrossRef]
  130. Wu, Z.; Liang, Y.; Yuan, X.; Zou, D.; Fang, J.; Jiang, L.; Zhang, J.; Yang, H.; Xiao, Z. MXene Ti3C2 derived Z?scheme photocatalyst of graphene layers anchored TiO2/g?C3N4 for visible light photocatalytic degradation of refractory organic pollutants. Chem. Eng. J. 2020, 394, 124921. [Google Scholar] [CrossRef]
  131. Du, C.; Zhang, Z.; Tan, S.; Yu, G.; Chen, H.; Zhou, L.; Yu, L.; Su, Y.; Zhang, Y.; Deng, F.; et al. Construction of Z-scheme g-CN4/MnO2/GO ternary photocatalyst with enhanced photodegradation ability of tetracycline hydrochloride under visible light radiation. Environ. Res. 2021, 200, 111427. [Google Scholar] [CrossRef] [PubMed]
  132. Guan, Z.; Li, X.; Wu, Y.; Chen, Z.; Huang, X.; Wang, D.; Yang, Q.; Liu, J.; Tian, S.; Chen, X.; et al. AgBr nanoparticles decorated 2D/2D GO/Bi2WO6 photocatalyst with enhanced photocatalytic performance for the removal of tetracycline hydrochloride. Chem. Eng. J. 2021, 410, 128283. [Google Scholar] [CrossRef]
  133. Wang, M.; Yu, H.; Wang, P.; Chi, Z.; Zhang, Z.; Dong, B.; Dong, H.; Yu, K.; Yu, H. Promoted photocatalytic degradation and detoxication performance for norfloxacin on Z-scheme phosphate-doped BiVO4/graphene quantum dots/P-doped g-C3N4. Sep. Purif. Technol. 2021, 274, 118692. [Google Scholar] [CrossRef]
  134. Wei, J.; Chen, Z.; Tong, Z. Engineering Z-scheme silver oxide/bismuth tungstate heterostructure incorporated reduced graphene oxide with superior visible-light photocatalytic activity. J. Colloid Interface Sci. 2021, 596, 22–33. [Google Scholar] [CrossRef] [PubMed]
  135. Kumar, S.; Kaushik, R.D.; Purohit, L.P. ZnO-CdO nanocomposites incorporated with graphene oxide nanosheets for efficient photocatalytic degradation of bisphenol A, thymol blue and ciprofloxacin. J. Hazard. Mater. 2022, 424, 127332. [Google Scholar] [CrossRef] [PubMed]
  136. Yang, R.; Zhu, Z.; Hu, C.; Zhong, S.; Zhang, L.; Liu, B.; Wang, W. One-step preparation (3D/2D/2D) BiVO4/FeVO4@rGO heterojunction composite photocatalyst for the removal of tetracycline and hexavalent chromium ions in water. Chem. Eng. J. 2020, 390, 124522. [Google Scholar] [CrossRef]
  137. Yao, J.; Gao, Z.; Meng, Q.; He, G.; Chen, H. One-step synthesis of reduced graphene oxide based ceric dioxide modified with cadmium sulfide (CeO2/CdS/RGO) heterojunction with enhanced sunlight-driven photocatalytic activity. J. Colloid Interface Sci. 2021, 594, 621–634. [Google Scholar] [CrossRef]
  138. Zuo, H.; Wu, C.; Du, H.; Shi, H.; Fu, Y.; Zhang, T.; Yan, Q. Construction of Z-scheme Ag-AgBr/Bi2O2CO3/CNT heterojunctions with remarkable photocatalytic performance using carbon nanotubes as efficient electronic mediators. Chemosphere 2022, 302, 134927. [Google Scholar] [CrossRef]
  139. Ahmadi, M.; Motlagh, H.R.; Jaafarzadeh, N.; Mostoufi, A.; Saeedi, R.; Barzegar, G.; Jorfi, S. Enhanced photocatalytic degradation of tetracycline and real pharmaceutical wastewater using MWCNT/TiO2 nano-composite. J. Environ. Manag. 2017, 186, 55–63. [Google Scholar] [CrossRef]
  140. Yan, D.; Hu, H.; Gao, N.; Ye, J.; Ou, H. Fabrication of carbon nanotube functionalized MIL-101(Fe) for enhanced visible-light photocatalysis of ciprofloxacin in aqueous solution. Appl. Surf. Sci. 2019, 498, 143836. [Google Scholar] [CrossRef]
  141. Ye, S.; Zhou, X.; Xu, Y.; Lai, W.; Yan, K.; Huang, L.; Ling, J.; Zheng, L. Photocatalytic performance of multi-walled carbon nanotube/BiVO4 synthesized by electro-spinning process and its degradation mechanisms on oxytetracycline. Chem. Eng. J. 2019, 373, 880–890. [Google Scholar] [CrossRef]
  142. Gao, P.; Huang, S.; Tao, K.; Li, Z.; Feng, L.; Liu, Y.; Zhang, L. Synthesis of adjustable {312}/{004} facet heterojunction MWCNTs/Bi5O7I photocatalyst for ofloxacin degradation: Novel insights into the charge carriers transport. J. Hazard. Mater. 2022, 437, 129374. [Google Scholar] [CrossRef] [PubMed]
  143. Samy, M.; Ibrahim, M.G.; Alalm, M.G.; Fujii, M. Effective photocatalytic degradation of sulfamethazine by CNTs/LaVO4 in suspension and dip coating modes. Sep. Purif. Technol. 2020, 235, 116138. [Google Scholar] [CrossRef]
  144. Khazaee, Z.; Mahjoub, A.R.; Khavar, A.H.C. One-pot synthesis of CuBi bimetallic alloy nanosheets-supported functionalized multiwalled carbon nanotubes as efficient photocatalyst for oxidation of fluoroquinolones. Appl. Catal. B-Environ. 2021, 297, 120480. [Google Scholar] [CrossRef]
  145. Zhang, S.; Rong, F.; Huang, S.; Zhao, S.; Wang, M.; He, L.; Zhang, Z.; Du, M. Atomic Fe sites embedded within carbon nanotubes for the efficient photodegradation of multiple tetracyclines. Sep. Purif. Technol. 2022, 287, 120530. [Google Scholar] [CrossRef]
  146. Wang, W.; Tian, J.; Zhu, Z.; Zhu, C.; Liu, B.; Hu, C. Insight into quinolones and sulfonamides degradation, intermediate product identification and decomposition pathways with the assistance of Bi2MoO6/Bi2WO6/MWCNTs photocatalyst. Process Saf. Environ. Prot. 2021, 147, 527–546. [Google Scholar] [CrossRef]
  147. Erim, B.; Cigeroglu, Z.; Sahin, S.; Vasseghian, Y. Photocatalytic degradation of cefixime in aqueous solutions using functionalized SWCNT/ZnO/Fe3O4 under UV-A irradiation. Chemosphere 2022, 291, 132929. [Google Scholar] [CrossRef] [PubMed]
  148. Rabanimehr, F.; Farhadian, M.; Nazar, A.R.S.; Moghadam, M. Fabrication of Z-scheme Bi2WO6/CNT/TiO2 heterostructure with enhanced cephalexin photodegradation: Optimization and reaction mechanism. J. Mol. Liq. 2021, 339, 116728. [Google Scholar] [CrossRef]
  149. Sun, P.; Zhou, S.; Yang, Y.; Liu, S.; Cao, Q.; Wang, Y.; Wagberg, T.; Hu, G. Artificial chloroplast-like phosphotungstic acid—Iron oxide microbox heterojunctions penetrated by carbon nanotubes for solar photocatalytic degradation of tetracycline antibiotics in wastewater. Adv. Compos. Hybrid Mater. 2022, 5, 3158–3175. [Google Scholar] [CrossRef]
  150. Yan, X.; Qian, J.; Pei, X.; Zhou, L.; Ma, R.; Zhang, M.; Du, Y.; Bai, L. Enhanced photodegradation of doxycycline (DOX) in the sustainable NiFe2O4/MWCNTs/BiOI system under UV light irradiation. Environ. Res. 2021, 199, 111264. [Google Scholar] [CrossRef]
  151. Yang, F.; Zhu, Q.; Gao, Y.; Jian, H.; Wang, C.; Sun, H. Effects of biochar-dissolved organic matter on the photodegradation of sulfamethoxazole and chloramphenicol in biochar solutions as revealed by oxygen reduction performances and free radicals. Sci. Total Environ. 2021, 781, 146807. [Google Scholar] [CrossRef]
  152. Xiao, Y.; Lyu, H.; Tang, J.; Wang, K.; Sun, H. Effects of ball milling on the photochemistry of biochar: Enrofloxacin degradation and possible mechanisms. Chem. Eng. J. 2020, 384, 123311. [Google Scholar] [CrossRef]
  153. Xie, X.; Li, S.; Zhang, H.; Wang, Z.; Huang, H. Promoting charge separation of biochar-based Zn-TiO2/pBC in the presence of ZnO for efficient sulfamethoxazole photodegradation under visible light irradiation. Sci. Total Environ. 2019, 659, 529–539. [Google Scholar] [CrossRef] [PubMed]
  154. Wang, W.; Zhang, J.; Chen, T.; Sun, J.; Ma, X.; Wang, Y.; Wang, J.; Xie, Z. Preparation of TiO2-modified Biochar and its Characteristics of Photo-catalysis Degradation for Enrofloxacin. Sci. Rep. 2020, 10, 6588. [Google Scholar] [CrossRef] [Green Version]
  155. Louros, V.L.; Silva, V.; Silva, C.P.; Calisto, V.; Otero, M.; Esteves, V.I.; Freitas, R.; Lima, D.L.D. Sulfadiazine’s photodegradation using a novel magnetic and reusable carbon based photocatalyst: Photocatalytic efficiency and toxic impacts to marine bivalves. J. Environ. Manag. 2022, 313, 115030. [Google Scholar] [CrossRef]
  156. Wang, Z.; Cai, X.; Xie, X.; Li, S.; Zhang, X.; Wang, Z. Visible-LED-light-driven photocatalytic degradation of ofloxacin and ciprofloxacin by magnetic biochar modified flower-like Bi2WO6: The synergistic effects, mechanism insights and degradation pathways. Sci. Total Environ. 2021, 764, 142879. [Google Scholar] [CrossRef]
  157. Chen, Z.; He, Z.; Zhou, M.; Xie, M.; He, T.; Zhao, Y.; Chen, X.; Wu, Y.; Xu, Z. In-situ synthesis of biochar modified PbMoO4: An efficient visible light-driven photocatalyst for tetracycline removal. Chemosphere 2021, 284, 131260. [Google Scholar] [CrossRef]
  158. Huang, H.-B.; Zhang, N.; Yu, K.; Zhang, Y.-Q.; Cao, H.-L.; Lu, J.; Cao, R. One-Step Carbothermal Synthesis of Robust CdS@BPC Photocatalysts in the Presence of Biomass Porous Carbons. ACS Sustain. Chem. Eng. 2019, 7, 16835–16842. [Google Scholar] [CrossRef]
  159. Tang, R.; Gong, D.; Deng, Y.; Xiong, S.; Zheng, J.; Li, L.; Zhou, Z.; Su, L.; Zhao, J. pi-pi stacking derived from graphene-like biochar/g-C3N4 with tunable band structure for photocatalytic antibiotics degradation via peroxymonosulfate activation. J. Hazard. Mater. 2022, 423, 126944. [Google Scholar] [CrossRef]
  160. Wang, T.; Zheng, J.; Cai, J.; Liu, Q.; Zhang, X. Visible-light-driven photocatalytic degradation of dye and antibiotics by activated biochar composited with K+ doped g-C3N4: Effects, mechanisms, actual wastewater treatment and disinfection. Sci. Total Environ. 2022, 839, 155955. [Google Scholar] [CrossRef]
  161. Gholami, P.; Khataee, A.; Soltani, R.D.C.; Dinpazhoh, L.; Bhatnagar, A. Photocatalytic degradation of gemifloxacin antibiotic using Zn-Co-LDH@biochar nanocomposite. J. Hazard. Mater. 2020, 382, 121070. [Google Scholar] [CrossRef] [PubMed]
  162. Azalok, K.A.; Oladipo, A.A.; Gazi, M. Hybrid MnFe-LDO-biochar nanopowders for degradation of metronidazole via UV-light-driven photocatalysis: Characterization and mechanism studies. Chemosphere 2021, 268, 128844. [Google Scholar] [CrossRef] [PubMed]
  163. Li, M.; Li, P.; Zhang, L.; Chen, M.; Tang, J.; Qin, C.; Lee, S.L.J.; Lin, S. Facile fabrication of ZnO decorated ZnFe-layered double hydroxides @ biochar nanocomposites for synergistic photodegradation of tetracycline under visible light. Chem. Eng. J. 2022, 434, 134772. [Google Scholar] [CrossRef]
  164. Peng, H.; Li, Y.; Wen, J.; Zheng, X. Synthesis of ZnFe2O4/B,N-codoped biochar via microwave-assisted pyrolysis for enhancing adsorption-photocatalytic elimination of tetracycline hydrochloride. Ind. Crops Prod. 2021, 172, 114066. [Google Scholar] [CrossRef]
  165. Wang, X.; Peng, Y.; Xie, T.; Zhang, Y.; Wang, Y.; Yang, H. Photocatalytic removal of sulfamethoxazole using yeast biomass-derived NixP/biocarbon composites in the presence of dye sensitizer. J. Environ. Chem. Eng. 2022, 10, 107426. [Google Scholar] [CrossRef]
  166. Zhou, Y.; Cai, T.; Liu, S.; Liu, Y.; Chen, H.; Li, Z.; Du, J.; Lei, Z.; Peng, H. N-doped magnetic three-dimensional carbon microspheres@TiO2 with a porous architecture for enhanced degradation of tetracycline and methyl orange via adsorption/photocatalysis synergy. Chem. Eng. J. 2021, 411, 128615. [Google Scholar] [CrossRef]
  167. Wang, Y.; Wang, A.; Pan, J.; Xue, Z.; Li, J.; Wang, G. Metal-organic complex-derived 3D porous carbon-supported g-C3N4/TiO2 as photocatalysts for the efficient degradation of antibiotic. Crystengcomm 2021, 23, 4717–4723. [Google Scholar] [CrossRef]
Figure 1. Antibiotics removal on carbon-based materials.
Figure 1. Antibiotics removal on carbon-based materials.
Nanomaterials 12 04045 g001
Figure 2. Schematic representation of modifications of graphene and carbon nanotube by covalent and non-covalent methodology.
Figure 2. Schematic representation of modifications of graphene and carbon nanotube by covalent and non-covalent methodology.
Nanomaterials 12 04045 g002
Figure 3. Photodegradation of antibiotics by carbon-based photocatalyst.
Figure 3. Photodegradation of antibiotics by carbon-based photocatalyst.
Nanomaterials 12 04045 g003
Table 7. Degradation characteristics of antibiotics by BC-based photocatalysts.
Table 7. Degradation characteristics of antibiotics by BC-based photocatalysts.
BiomassPyrolysis
Temp (°C)
PhotocatalystsAntibioticDosage (g/L)Light SourceDegradation EfficiencyRefs.
Poplar sawdust500PbMoO4/BCTetracycline3.0Visible light61.0% within 120 min[157]
Potato stems and leaves600CdS/BPCTetracycline0.1Visible light84.6% within 120 min[158]
Enteromorpha700g-C3N4/BCTetracycline0.2Visible light88% within 60 min[159]
Caragana korshinskii650K-g-C3N4/BCTetracycline1.0Visible light90.94% within 180 min[160]
Rice husks400ZnO/ZnFe-LDH/BCTetracycline0.2Visible light87.7% within 240 min[163]
Sugarcane bagasse600ZnFe/BN-BCTetracycline hydrochloride0.33Sunlight98.19% within 120 min[164]
Rice straw700Pure BCSulfamethoxazole0.5Visible light96.28% within 720 min[151]
Reed straw500Zn-TiO2/pBCSulfamethoxazole1.25Visible light81.21% within 180 min[153]
Baker’s yeast900NixP/BCSulfamethoxazole0.4Visible light98.71% within 120 min[165]
Primary paper mill sludge800Mag-TiO2/KBCSulfadiazine0.1Sunlightt1/2= 5.6 ± 0.4 h[155]
Rice straw700Pure BCChloramphenicol0.5Visible light95.23% within 720 min[151]
Caragana korshinskii650K-g-C3N4/BCChloramphenicol1.0Visible light82.42% within 180 min[160]
Poplar woodchips300Ball-milled BCEnrofloxacin0.2Visible lightUp to 82.5% within 150 min[152]
Corn stalk500TiO2/KBCEnrofloxacin2.5UV light85.25% within 60 min[154]
Reed straw500Bi2WO6/Fe3O4/BCOfloxacin0.4Visible light83.1% within 60 min[156]
Reed straw500Bi2WO6/Fe3O4/BCCiprofloxacin0.4Visible light91.5% within 60 min[156]
Caragana korshinskii650K-g-C3N4/BCNorfloxacin1.0Visible light83.62% within 180 min[160]
husks and paper sludge500Zn-Co-LDH/BCGemifloxacin0.75UV light92.7% within 100 min[161]
Palm seeds600MnFe-LDO/BCMetronidazole0.5UV light98% within 60 min[162]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ma, R.; Xue, Y.; Ma, Q.; Chen, Y.; Yuan, S.; Fan, J. Recent Advances in Carbon-Based Materials for Adsorptive and Photocatalytic Antibiotic Removal. Nanomaterials 2022, 12, 4045. https://doi.org/10.3390/nano12224045

AMA Style

Ma R, Xue Y, Ma Q, Chen Y, Yuan S, Fan J. Recent Advances in Carbon-Based Materials for Adsorptive and Photocatalytic Antibiotic Removal. Nanomaterials. 2022; 12(22):4045. https://doi.org/10.3390/nano12224045

Chicago/Turabian Style

Ma, Raner, Yinghao Xue, Qian Ma, Yanyan Chen, Shiyin Yuan, and Jianwei Fan. 2022. "Recent Advances in Carbon-Based Materials for Adsorptive and Photocatalytic Antibiotic Removal" Nanomaterials 12, no. 22: 4045. https://doi.org/10.3390/nano12224045

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop