Next Article in Journal
Color-Stable Formulations for 3D-Photoprintable Dental Materials
Previous Article in Journal
Testing Protocol Development for the Fracture Toughness of Parts Built with Big Area Additive Manufacturing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Covalent–Organic Frameworks for Selective and Sensitive Detection of Antibiotics from Water

School of Chemistry and Chemical Engineering, Linyi University, Linyi 276005, China
*
Authors to whom correspondence should be addressed.
Shared First authorship.
Polymers 2024, 16(16), 2319; https://doi.org/10.3390/polym16162319
Submission received: 12 July 2024 / Revised: 4 August 2024 / Accepted: 9 August 2024 / Published: 16 August 2024
(This article belongs to the Section Polymer Applications)

Abstract

:
As the consumption of antibiotics rises, they have generated some negative impacts on organisms and the environment because they are often unable to be effectively degraded, and seeking effective detection methods is currently a challenge. Covalent–organic frameworks (COFs) are new types of crystalline porous crystals created based on the strong covalent interactions between blocked monomers, and COFs demonstrate great potential in the detection of antibiotics from aqueous solutions because of their large surface area, adjustable porosity, recyclability, and predictable structure. This review aims to present state-of-the-art insights into COFs (properties, classification, synthesis methods, and functionalization). The key mechanisms for the detection of antibiotics and the application performance of COFs in the detection of antibiotics from water are also discussed, followed by the challenges and opportunities for COFs in future research.

1. Introduction

Antibiotics were first discovered by A. Fleming in 1929, and since then, large quantities of antibiotics have been widely used in aquaculture, the livestock industry, and medical treatment [1]. Antibiotics refer to a class of secondary metabolites produced by microorganisms (including bacteria, fungi, and actinomycetes) or higher animals and plants in the process of life, which have anti-pathogen or other activities and can interfere with the development of other living cells [2]. The abuse of antibiotics results in a negative influence on the environment and human health, including toxic effects, reproductive disorders, and disruption of the ecological balance [3]. Receptors are found to be synergistic to antibiotics or their residues and result in severe diseases, even in very low concentrations. Moreover, one key issue is that antibiotics are highly cumulative and not easy to degrade, resulting in biological effects [4].
Thus, many countries strictly regulate the use of antibiotics and limit the maximum residue of antibiotics in different products [5]. Therefore, it is of great importance to develop sensitive and selective methods for the detection of antibiotics, especially in large antibiotic abuse settings [6]. Currently, for antibiotic identification in water environments, many methods have been established, including high-performance liquid chromatography (HPLC), gas chromatography (GC), liquid chromatography–mass spectrometry (LC-MS), surface-enhanced Raman spectroscopy (SERS), and so on [7,8]. However, these traditional methods expose many shortcomings, such as high costs, long use times, and possible secondary pollution caused by the intermediates [9]. Therefore, finding efficient, energy-saving and effective materials and methods is a key task for the detection of antibiotics.
Covalent–organic frameworks (COFs) were first discovered by the Yaghi group in 2005 [10]. Their organic building units are connected by covalent bonds to form a porous skeleton with a periodic structure through dynamic covalent bonds of light elements (B, C, N, O, and Si), thus having excellent properties [11]. The covalent bonding linkage of COFs mainly includes a boric acid ester bond, an imine-based linkage, a triazine-based linkage, and so on [12]. According to its spatial structure, COFs can be divided into two-dimensional (2D) and three-dimensional (3D) structures [13]. In 3D COFs, the organic units are connected by covalent bonds to form a three-dimensional network structure, and the 3D structure material is widely used in catalysis, gas adsorption, and other aspects [14]. In 2D COFs, organic units are connected with two-dimensional atomic layers, and atomic layers are further stacked to form layered structures through π–π interaction, and the whole frame structure is determined by covalent bonds within layers and controlled by non-covalent forces between layers, so it has a broad application prospect in the energy storage field [15]. Because of their unique advantages, such as high thermal stability, large specific surface area, rich pores, adjustable molecular structure, and many active sites, COFs have been gradually applied to the fields of catalysis, separation, photoelectricity, energy, and so on [16,17]. Figure 1 shows the number of papers about COFs and antibiotics published since 2019.
In recent years, various materials, like metal–organic frameworks (MOFs), carbon nanomaterials, nanoparticle biosensors, and molecularly imprinted polymers (MIPs), have been designed and developed for the detection of antibiotics [18]. However, these materials still have the drawbacks of expensive costs, toxic residues, complicated pretreatment processes, and so on [19]. Therefore, convenient and low-cost detection materials with high accuracy and sensitivity are desirable [20]. In comparison with these materials, COFs have higher porosity and bigger and tunable pore sizes, which speed up the diffusion of reactants and desorption of products, thereby promoting higher yield and selectivity [21]. Moreover, some excellent reviews have been published about COFs in the detection of pharmaceuticals, and in this way, there is a great need and room for the analysis of COFs in antibiotic detection [22].
This review first aims to present state-of-the-art insights into COFs (properties, classification, synthesis methods, and functionalization). Secondly, the key mechanisms for the detection of antibiotics and the application performance of COFs in the detection of antibiotics from water are also discussed. Finally, the challenges and opportunities of COFs in future research are also discussed.

2. Antibiotics

2.1. Classification and Synthetic Pathways of Antibiotics

According to the different classification standards, antibiotics can be classified as the following:
Firstly, according to the chemical structures, antibiotics can be divided into quinolone antibiotics, β-Lactam antibiotics, macrolides, aminoglycoside antibiotics, etc. [23].
Secondly, according to the function, antibiotics can be divided into antibacterial antibiotics, antifungal antibiotics, anti-tumor antibiotics, antiviral antibiotics, livestock antibiotics, agricultural antibiotics, and others [24].
Antibiotics can be produced in various ways according to their types, such as penicillin biosynthesis through microbial fermentation and sulfonamides, quinolones, etc., through chemical synthesis [25]. There are also semi-synthetic antibiotics, and they are derivatives made by modifying the molecular structure of antibiotics obtained through biosynthesis using chemicals (biological or biochemical methods) [26].

2.2. Hazards of Antibiotics

There are five major hazards of antibiotics, including intestinal microbiota imbalance, fungal infections, leukopenia, liver and kidney dysfunction, and gastrointestinal symptoms [27]. Firstly, antibiotics not only kill harmful bacteria but also kill normal gut microbiota, which can cause dysbiosis of gut microbiota and lead to diarrhea, known as antibiotic-related diarrhea [28]. Secondly, if there is long-term use of antibiotics, a decrease in the normal bacterial population in the body and fungal reproduction can be easily caused, resulting in fungal enteritis, thrush, etc. [29]. Thirdly, the use of antibiotics can also cause damage to normal cells in the body, leading to a decrease in neutrophils and white blood cells [30]. Fourthly, the use of antibiotics for a long time and in large doses can easily cause drug-induced hepatitis, liver function damage, etc., and drug excretion through the kidneys can also increase the burden on the kidneys [31]. Fifthly, the use of antibiotics may result in gastrointestinal symptoms, such as diarrhea, abdominal pain, nausea, and vomiting [32].

3. COF Structure and Classification

3.1. COF Structure

COFs are another important three-dimensionally ordered material emerging after metal–organic framework materials (MOFs) [33]. There are many important factors in determining the geometries of COFs, such as size, morphology, structure, and the dynamic nature of the bonds used to connect components [34]. COFs are composed of strong covalent bonds (C-C, C-O, C-B, and B-O), and they have high thermal stability, high surface area, and extremely low density [35]. In addition to their well-organized structure, COFs can also be designed with holes of their size and chemical characteristics according to the selection of initial structural units and adjusting its side chain length [36].
COFs can be divided into 2D and 3D COFs according to the topological structure [37,38] (Figure 2). In 2D COFs, monomers are connected by covalent bonds to form a layered structure in the plane. Conjugate systems are formed between layers through π–π interaction, and one-dimensional channels are also formed. The size and shape of the channels are closely related to the stacking mode between layers [39]. The high selectivity and permeability of 2D COFs render them with high separation performance and are then successfully utilized for purification applications [40]. For 3D COFs, the monomer extends infinitely through covalent bonds to form a framework with regular and periodic structure [41]. Different from 2D COFs, 3D COFs can easily form large, cage-like cavities when they are extended in 3D space, while monomers can continue to react in the cavities and grow outwards so that 3D COFs often have multiple interpenetrating structures, and 3D COFs with enhanced reactant and active sites accessibility exhibit potential for separation due to their large specific surface area and high porosity [42].

3.2. COF Classification

In general, COFs can be divided into boron, triazine, imine, phenyl hydrazone, ketone enamine, polyimide, phthalocyanine, and porphyrin according to the type of covalent bond formed by polycondensation [43]. Figure 3 shows the diversity of linkages for the formation of COFs. The covalent bond connecting COFs endows them with high-temperature resistance, but the other chemical stability of different covalent bonds of COFs is different [44]. As boroxines and boronic esters are water-labile bonds and rather sensitive to hydrolysis, the introduction of the imine linkage formed by the condensation between aldehydes and primary amines supports the development of more stable coupling motifs, as well as with hydrazones, azines, and imides [45]. Moreover, β-ketoenamines with further improved chemical stability forming from primary amines and 1,3,5-triformylphloroglucinol undergo a second and irreversible enol-keto-tautomerization to create networks that are stable against acids and bases [46].

3.2.1. Boron-Based COFs

The first boron-based COFs (COF-1 and COF-5) were prepared by Yaghi and coworkers, and since then, diverse boron-based COFs have been synthesized through the formation of boronate ester, boroxine, or borazine [47]. According to the synthetic strategies, most boron-based COFs can be clarified into two categories: single building block self-condensation and two or more building unit co-condensation [48]. Yaghi et al. first synthesized boron-based COF-1 with 1,4-benzene diboronic acid at 120 °C, and the obtained COF-1 had a layered structure of a hexagonal graphene-like structure formed by the interconnection of B3O3 rings [49]. Liu et al. used this reaction to obtain COF-102 and COF-103, and they were boron-containing polymers with a C-C bond and Si-C bond as the central connection points, respectively [50]. To improve the stability of boron-based COFs, Li et al. introduced anionic boron centers fabricated of a spiroborate-linked COF, slowing down its hydrolysis and improving its structural controllability while improving the stability of COFs. The research on COFs was also promoted to a new height [51].
Boron-based COFs generally possess low density, a high surface area, and high thermal stability, leading to various applications, but most of these materials have the defect of easy decomposition in water due to the electronic deficiency of the boron element. They are often unstable in a humid environment, and some of them may undergo structural transformation even under air [52].

3.2.2. Triazine-Based COFs

The first triazine-based COFs (donated as CTFs-1) were achieved by Thomas and co-workers in 2008 through cyclotrimerization of aromatic nitriles at 400 °C [53]. At present, only a few triazine-based COFs have been successfully synthesized using this method because the reaction conditions are harsh, the synthesis process is difficult to control, and the requirements for the thermal stability of the precursor are high, and it is not conducive to mass production [54]. Thus, other strategies based on milder conditions have been developed, and a low-temperature polycondensation approach was utilized to synthesize CTFs based on a broader range of building blocks via mild conditions [55].
Triazine-based COFs have many advantages, such as a high surface area and remarkable thermal and chemical stabilities. In the study by Li, L. et al., the stability of the triazine-based COFs could retain 88% of the initial capacitance after 20,000 charge−discharge cycles in the cycle life test, and such excellent cycle stability has rarely been reported in analogous materials [56]. Wang et al. showed that triazine-based COFs had a surface area of 789 m2 g−1, a total pore volume of 0.37 cm3 g−1, and a pore size of 1.23 nm because of the triazine framework. Their high specific area and specific active sites made them excellent carriers, which had broad application prospects in separation, analysis, photoelectric, heterogeneous catalysis, electrochemistry, and other fields [57].

3.2.3. Imine-Based COFs

Imine-based COFs are another representative class of COFs, which are based on the fact that the formation of an imine is a reversible reaction [58]. At present, the synthesized imine-based COFs mainly include two groups: one group is formed via the co-condensation of aldehydes and amines, and the other group is formed by the co-condensation reaction of aldehydes and hydrazides [59]. The first imine-based COF (COF-300) was developed by Yaghi and co-workers through the reaction of aldehyde and amine [60]. Since then, F. J. Uribe Romo et al. successfully prepared 3D COF-300 with good thermal stability and chemical stability and connected by an imine bond using the amine aldehyde condensation method for the first time [61]. Yan and co-workers achieved imine-based COFs with 5.3 nm wide pore loadings with Rhodamine B. Rhodamine B’s luminescence intensity was inversely proportional to temperature, and it also showed good optical sensing characteristics [62]. Jiang et al. demonstrated imine-based COFs loaded with N-heterocyclic compounds (triazole and imidazole) with high proton conductivity 2–4 orders of magnitude higher than nonporous or microporous systems [63].
Imine-based COFs have attracted large interest in many areas because they combine the advantages of boron-containing COFs and triazine-based COFs, for example, much better crystallinity and structural regularity than triazine-based COFs and better chemical stability in most organic solvents or water compared with boron-containing COFs [64].

3.2.4. Hydrazone-Based COFs

Hydrazone-based COFs are connected by hydrazone bonds formed by the dehydration and condensation of aldehyde and hydrazine and were first studied by Uribe Romo et al. [65]. Hydrazone-based COFs have good stability for hydrolysis, which is conducive to the selective resolution of heteroporous COFs due to the stability of hydrazone holding units [66]. Hydrazone-based COFs have been widely used in the field of photocatalysis, separation, and electrochemistry due to their unique photoelectric properties, good light absorption ability, and physical and chemical stability [67].

3.2.5. Azine-Based COFs

The first azine-linked COFs were synthesized by the condensation of 1,3,5,8-tetrakis(4-formylphenyl)pyrene with hydrazine, and for azine linkage, hydrazine acted as a common building block to react with diverse aldehydes [68]. Azine-based COFs could provide abundant active sites for specific guest molecules; in addition, nitrogen-containing COFs had directional and vertical one-dimensional channels and an orderly periodic network structure, which enabled small guest molecules to diffuse smoothly in the channels without blocking. Therefore, they had great advantages in separation, adsorption, and the removal of harmful molecules [69].

3.2.6. β-Ketoenamine-Based COFs

β-ketoenamine-based COFs have keto-enol tautomerism; for example, Banerjee et al. prepared a kind of β-ketoenamine-based COF through irreversible enol–keto tautomerization of ketenimine linkages in imine-based COFs [70]. β-ketoenamine-based COFs have excellent adsorption performance and catalytic activity, and at the same time, the forward reaction rate is fast, which leads to the self-repair and insufficient reassembly ability of such polymers, and it becomes very difficult to form an ordered structure [71].

3.2.7. Imide-Based COFs

Imide-based COFs were first reported by Yan’s group in 2014. They were fabricated via a reversible imidization reaction and simply by extending building molecules, and the large pore size of imide-based COFs could be tuned [72]. The imide-based COFs exhibited a high surface area and had great potential for loading dye molecules and drug delivery [73].

4. Synthetic Methods of COFs

The development of simple and effective synthesis methods is very important for the large-scale application of COFs [74]. Figure 4 shows the proposed process for the synthesis of COFs. The main synthesis methods include ionic heating, microwave-assisted solvent heating, mechanical grinding, and solvothermal methods [75,76].
The thermodynamic reversible reaction is conducive to the formation of a long-range ordered structure, so COFs are generally prepared by a thermodynamic reversible reaction [77]. A single or mixed solvent is used to dissolve some of the monomers, and COF microcrystalline aggregates are obtained by long-term reaction under high-temperature and low-pressure conditions. During the process, polymerization, crystallization, assembly, and other processes are involved [78]. The synthesis of COFs is affected by the reaction temperature, time, composition, etc. In the synthesis process, their morphology, size, surface properties, and functions can be controlled [79].

4.1. Ionothermal Synthesis

So far, the ionothermal synthesis of COFs has only been used for the synthesis of triazine COFs [80]. Figure 5 shows the synthesis of COFs using the ionothermal method.
Thomas’s research group first synthesized CTF-1 COF material using the ionothermal method at 400 °C using molten ZnCl2 as the solvent. After 40 h, when the system temperature dropped to room temperature, the amperometric flask was opened to obtain CTF-1 after a series of solid treatments [81]. Due to the strict reaction conditions and high reaction temperature of this method, the requirements for thermal stability of the constituent units are also increased, and at present, the use of ionothermal synthesis of COFs is relatively rare [82].

4.2. Microwave-Assisted Synthesis

The microwave-assisted synthesis method refers to the use of microwave heating, and compared with the traditional heating method, it does not need a heat-conduction process and can achieve uniform heating in a short time [83]. Microwave-assisted synthesis was first used to synthesize boron-based COFs in 2009. Compared with traditional solvothermal methods, microwave-assisted synthesis can be realized in a short time (within 20 min), with a higher specific surface area (2019 m2 g−1 for COF-5 and 2926 m2 g−1 for COF-102). The development of the microwave-assisted method provides a simple, rapid, and efficient method for the synthesis of stable COFs, which is expected to be a promising method for the industrial production of COFs [84]. Cooper’s group synthesized 2D COF-5 and 3D COF-102 via a rapid microwave synthesis method in 20 min with higher surface areas compared to other synthesis methods [85]. Figure 6 shows the images of COF-5 using the microwave-assisted method (gray COF-5 with purple impurities (a), the removal of impurities via the MW extraction process (b), and purified gray COF-5 (c)).
The main characteristics of microwave heating are fast heating speed, energy saving, high efficiency, easy control, etc. Therefore, using microwave heating can improve the reaction speed and shorten the reaction time, which is a very promising overall heating method for COF synthesis [86].

4.3. Room-Temperature Synthesis

Room-temperature syntheses chiefly consist of two techniques: mechanochemical grinding and the solvothermal method [87].
Compared with the traditional solvothermal method, the mechanical grinding synthesis of COFs has the characteristics of simple operation, fast speed, environmentally friendly, and high production capacity [88]. Biswal and co-workers synthesized 1,3,5-triformylphlorog lucinol-diamine COFs using the room-temperature synthesis method [89]. However, due to its low productivity and general performance, there are few examples of this method being used to synthesize COFs [90].

4.4. Solvothermal Method

The solvothermal method is a commonly used method to synthesize COFs; for example, Shiraki et al. reported the synthesis of COFs in 1,4-phenylenediamine and 1,3,5-benzenetricarbaldehyde using the solvothermal method [91]. Matsumoto et al. established a solvothermal method synthesis of 2D imine COFs with high stability and BET surface areas (2175 m2 g−1) [92]. Figure 7 shows the preparation of COFs via the solvothermal synthesis approach. The solvothermal synthesis method refers to the reaction of mixtures with organic or non-aqueous solutions as solvents under a certain temperature and pressure [93]. The general synthesis step is to add monomer and solvent into the Pyrex tube, freeze with liquid nitrogen, vacuum, thaw, repeat the above operation three times, and remove oxygen from the system. Then, seal the Pyrex tube with a flame lance, put it into a thermostatic oven, and after several reactions, a large amount of insoluble substances will be produced in the system. Finally, COFs can be obtained through a series of purifications [94].

5. Physical and Chemical Properties of COFs

5.1. Chemical and Thermal Stability

The chemical and thermal stability of COFs are significant features that distinguish them from other materials [95]. Firstly, the strong stability of COFs is determined by its constituent monomers. The monomer elements include linkages, knots, edges, hydrazone, imine, and azine, as well as some monomers with electron-donating groups (such as methoxy, hydroxyl, and amino), and these groups are usually introduced into the phenyl edge to enhance the chemical stability of COFs [96]. Secondly, the stability is related to the strong covalent bond connecting the mediating elements that form a highly sustainable structure, such as the Ce-N bond, Be-O bond, and intermolecular hydrogen bond, which can improve the stability of COFs [97]. Thirdly, they are also related to the synthesis conditions and methods, such as polycondensation, photocatalysis, and synthesis in a closed heat-resistant tube under solvothermal conditions, to ensure that H2O can be used for bond self-correction and increase their chemical and thermal stability [98].

5.2. High Porosity and Large Specific Surface Area

The pore size of COFs is directly affected by the length of the organic functional groups, and the longer the organic ligand is, the larger the pore size of the material after removing the guest molecules [99]. The COFs have the properties of high porosity, relatively small pore size, and adjustable pore size, which are conducive to rapid molecular diffusion and transport and show excellent properties in practical applications [100].
The specific surface area is another important index used to evaluate the catalytic performance and adsorption capacity of porous materials [101]. By continuously changing the monomers and covalent bonds of COFs, the specific surface area of COFs can be increased. Most of the surface areas of COFs reported in the relevant literature exceeded 1000 m2/g, and the specific surface areas of COF-102 and COF-103 reached 3472 m2/g and 4210 m2/g [102].
COFs are entirely built from organic building blocks connected by covalent bonds having π-conjugated structures, and many COFs have high sustainability, chemical stability, and modest electronic properties for the detection of antibiotics [103]. The properties of different kinds of reported COFs are tabulated in Table 1.

5.3. Rigid Topology and Good Modifiability

The topological structure of COFs extends from zero dimensions to three dimensions, and its species extend from amorphous to crystalline [104]. Dynamic covalent chemistry is the scientific basis for the directional design of COFs, and in the process of covalent assembly, it has a high degree of adjustable deformation and is also the key to obtaining crystallinity and stability [105].
There are many adjustable measures for COFs, which can not only selectively introduce functional groups but can also take inorganic particles, such as molecular sieve, activated carbon, carbon nanotubes, and porous silica, with electron pairs as active centers and graft them onto the COFs skeleton [98].
Due to the advantages of high thermal stability, high specific surface area, rigid topology, and good modifiability, COFs have great advantages in sensing, selective capture of pollutants, energy storage, separation, and catalysis, showing a broad development prospect of COFs [106]. In recent years, some research groups at home and abroad have conducted distinctive research on the applications of COFs, and COFs have been successfully applied on a large scale, becoming an important part of green chemistry [107].

6. Application of COFs for Detection of Antibiotics

COFs, due to their high stability, large specific surface area, and easy functionalization, can provide an alternative to easily detect antibiotics with high sensitivity, mainly in two approaches, namely, biosensors and adsorbents [108].

6.1. As Biosensors for Detection of Antibiotics

Biosensors have achieved vigorous and rapid development in recent decades because of their good selectivity, high sensitivity, fast analysis speed, low cost, and online continuous monitoring in complex systems, including environmental monitoring, health care, and food inspection [109]. COF-based biosensor materials show a high specific surface area and porous skeleton, which can provide a large surface interface for biomolecule immobilization, and in this way, the aptamer chain can not only be attached to the COF substrate by π–π stacking and hydrogen bonding but also be bound by the electrostatic interaction between the amino terminal functional group and the negatively charged aptamer, facilitating their applications in detecting antibiotics [110].
Yao, D.’s work was related to a new Au nanocluster@ covalent–organic framework (Au@COF) sensor and applied to detect ultatrace small molecules, and the Au@COF sensor showed excellent sensitivity and good accuracy, proving the great potential of COF materials to be used as a sensor in the detection of biological samples [111]. Wang et al. condensed 1,3,6,8-tetra (4-formylphenyl) pyrene and melamine to synthesize an imine-based COFs material. This COF material was used as an electrochemical aptamer sensor for the detection of different kinds of antibiotics. The results showed that this method had a lower detection limit (6.07 and 0.04 fg mL−1 for enrofloxacin and ampicillin, respectively) and excellent sensing performance, which effectively extended the COFs as sensor materials to the sensitive detection of antibiotics [112]. Figure 8 shows the scheme of electrochemical detection of enrofloxacin and ampicillin using COF-based aptasensors.
Sun, Y. et al. used a simple one-pot strategy to effectively graft COFs onto carbon nanotubes and used them as electrical sensing materials to detect nitrofuran antibacterial agent furazolidone. As a network coating, COFs made the sensing materials have a high specific surface area (147.3 m2 g−1) and good sensing and separation properties (7.75 × 10−8 M when the detection ranged from 0.2 μM to 100 μM) [113].
Yang, Y. et al. successfully synthesized a highly sensitive SERS immunosensor based on COFs for the simultaneous detection of multiple foodborne pathogens. By using COF as a site to load non-bio-interference Raman signals and specific antibodies for the detection of foodborne pathogens, it showed strong non-interference characteristic Raman signals (2271 and 2113 cm−1 for two different foodborne pathogenic bacteria), and it could realize the simultaneous detection of two components. This COFs-based sensor provided a new idea for the detection of pathogenic bacteria [114].
Zhang, X. et al. prepared a new photoelectrochemical aptamer sensor based on 2D porphyrin COFs and used it for rapid, label-free detection of C-reactive protein. Due to the strong and rigid covalent bonds of COFs, the COF-based sensor had good conductivity and stability. In the experiment results, it was shown that the COF-based sensor had good analytical performance, such as high stability, sensitive response, good selectivity, and a wide linear range, making the COF-based sensor an ideal candidate for biological detection [115].

6.2. As the Adsorbent Materials for Pretreatment Technology

Pretreatment procedures are one of the most important steps in extraction methods, and the fast, simple, and automatic pretreatment technology not only saves time and labor but also reduces the error caused by different personnel’s operation and sample transfer [116]. So far, the common pretreatment methods include supercritical fluid extraction, solid-phase microextraction (SPME), solid-phase extraction (SPE), liquid-phase microextraction, membrane-separation technology, etc. [117,118].
In recent years, COF materials have received more and more attention in the field of sample processing; for example, COFs show obvious advantages in SPE, magnetic solid-phase extraction (MSPE), SPME, and chromatographic separation, with a high adsorption capacity and enrichment efficiency [119,120]. Figure 9 shows the fabrication process and application of magnetic COFs in MPE. This review mainly introduces the application of COF materials in SPE and SPME for the detection of antibiotics in water solutions.

6.2.1. SPE

The SPE procedure uses a solid adsorbent to adsorb target compounds, separate them from the sample matrix and interfering compounds, and then elute or heat them with the eluent to achieve the purpose of separating and enriching target compounds [121]. A high-efficiency adsorbent is the key factor of SPE, which can improve the recovery rate and enrichment multiple and reduce the consumption of organic solvents, costs, etc. [122]. COFs not only have high crystallinity, excellent specific surface area, and high porosity but also have different pore sizes, which can be used as adsorbents for SPE to adsorb active ingredients efficiently and quickly [123].
Wang et al. used triazine-based COFs as the adsorbent of SPE and then used it in the separation of five sulfonamides and four tetracyclines. In the experiment, the triazine-based COFs showed a good extraction efficiency for the studied antibiotics (recoveries: 88.0–93.4%, and limit of detection (LOD): 0.031–0.55 μg L−1, and relative standard deviation (RSD): 2.5–4.8%), and the reusability was also tested for five cycles after absolute elution and balance of each cycle, proving the stability of adsorption characteristics. The successful application of triazine-based COFs to the separation of multiple antibiotics proved that the COF material was a promising antibiotic adsorbent [124].
Wang et al. synthesized a new type of COF using a simple electrospinning method and used it as an adsorbent for SPE for enriching tetracycline antibiotics. The LOD (range from 0.6 to 3 ng mL−1), the limit of quantitation (LOQ) (range from 2 to 10 ng mL−1), and the recoveries decreased less than 30% after five consecutive cycles of SPE operation, exhibiting acceptable reproducibility and reusability. As an SPE adsorbent, the COFs have a fast adsorption rate and high adsorption performance, expanding the potential application of COFs in sample detection [125]. Figure 10 shows the schematic diagrams of the preparation of COFs (a) and application in SPE procedure (b) for the extraction of tetracyclines.
Wen et al. designed and synthesized a spherical tribenzodimethoxy dimethyl terephthalate-based COF and used it as an SPE adsorbent to extract polar sulfonamide antibiotics from water samples. The COF had rich adsorption sites and showed excellent adsorption performance for trace polar sulfonamides in water. The obtained COF had a low detection limit (0.5–1.0 ng L−1) and good repeatability and could be reused over 20 times without evident loss in the enrichment efficiency (76–87%). These results indicated that the COF was an important tool for the study of pollutant adsorption and was of great significance [126].
Wen et al. combined Fe3O4@COF with MSPE and used this method for the extraction and adsorption of fluoroquinolones. The LOD of this method for fluoroquinolones was 0.1–1.0 μg kg −1 and had good linearity (R2 ≥ 0.9989), and the RSD values (3.8–7.7%) of inter-day precisions were repeated three different days, indicating good stability. The experimental results showed that the Fe3O4@COF in the detection of fluoroquinolones in complex liquid matrices had important application prospects [127].

6.2.2. SPME

SPME was first proposed by Pawliszyn et al. in 1989, and its principle was to melt various cross-linked and bonded stationary phases on the mandrel in the syringe with an outer sleeve [128]. When using, the mandrel was pushed out and immersed in the crude sample solution, and the components to be measured were adsorbed on the mandrel, and then the mandrel of the probe was directly inserted into the sample inlet of the detector. Finally, the components to be measured were analyzed at the sample inlet and entered the instrument for analysis [129]. COFs as SPME adsorbents can obtain a high extraction and enrichment efficiency owing to their large specific surface area, suitable pore size, and functional groups on the surface [130].
The magnetic COFs were synthesized by Zhang et al., and they showed good adsorption efficiency (RSDs less than 5.1%), high recoveries (in the range of 87.0–113.8%), and recoveries of approximately 80% after six recycles, demonstrating good stability and reusability when used as SPME adsorbent to adsorb tetracycline. These studies have opened the door to the application of COFs in the field of chromatographic separation technology, which is very beneficial to the application and promotion of effective analysis technology of COFs in trace antibiotics in water samples [131].
Yue et al. prepared COF by using 1,3,5-trimethylphloroglucinol and benzidine samples. This functionalized COF was used as the adsorbent in SPME to adsorb phthalates, and the COF fiber could be used repeatedly at least 150 times with a significantly low loss of extraction efficiency (<4.8%), limits of detection for phthalate esters (ranging from 0.001 to 0.430 μg/L), and enrichment factors (ranging from 226 to 2154). These research results strongly proved that COFs had a bright future in the adsorption of antibiotic organics [132]. Figure 11 shows the schematic diagram of the preparation of SPME COF-fiber and its application in the detection of phthalate esters in the SPME process.
Guo et al. combined 1,3,5-tris (4-formylphenyl) benzene and benzidine to prepare a COF fiber for the SPME of the selective identification and adsorption of polychlorinated biphenyls in aquatic products. The experimental results showed that the method had good recoveries (87.1–99.7%), a high enhancement factor (4471–7488), low detection limit (0.07–0.35 ng L−1), high accuracy (RSD < 100 ng L−1), and the peak areas of 100 ng L−1 had no significant changes after 180 extraction cycles. This work realized the detection of aromatic toxic organic pollutants in the extracted environmental samples of COFs and expanded the application of COFs in the adsorption field [133].
Ji et al. realized the application of the COF with dioxin linkage as an SPME coating to adsorb perfluoroalkyl substances in water. The designed COF-based SPME method showed high precision (RSD ≤ 7.9%), good linearity (0.01–1000 ng L−1) (R2 ≥ 0.9945), and low LOD (0.0020–0.0045 ng L−1), and the COF-coated fibers could be reused at least 20 times without the loss of extraction performance. These results show the prospect of the application of COFs in SPME and greatly encourage researchers to devote themselves to the application research of COFs in the field of extraction [134].

6.3. Comparison between COF-Based Materials and Other Materials for Detection of Antibiotics

By now, various kinds of materials have been studied for antibiotic detection, including MOFs, graphene, graphene oxide, quantum dots, MIPs, and other kinds of nanomaterials. However, using these materials for antibiotic detection exhibited some drawbacks; for example, the synthesis process for MIPs was complex and generated some toxic by-products [135]. Classical materials and methods that were used for the detection of antibiotics from water samples are presented in Table 2. In the table, it is shown that compared with other materials, COFs revealed better recovery, lower limits of detection, and RSD in the detection of antibiotics from water samples.

7. Adsorption Mechanism of Antibiotics by COFs

COFs have gained significant attention as a promising material for the adsorption of antibiotics. Lots of research has been conducted on their adsorption mechanisms of antibiotics, and the adsorption mechanism mainly includes the following aspects: porosity, pH value, and charge interaction; π–π stacking; hydrogen bonding; and hydrophobic interaction [152].

7.1. Porosity

COFs have high porosity and a large specific surface area, which enable them to act as carriers with adsorption activity. On the other hand, they reduce the aggregation of inorganic nanoparticles and introduce more adsorption sites to a greater extent. The characteristics of the corresponding structure greatly improve the adsorption performance of the composite materials and further efficient and sustainable adsorption [153].
Jiang et al. displayed the SO3H-anchored COFs with markedly high adsorption capacity of fluoroquinolone antibiotics (726 mg g−1, 531 mg g−1, and 9.5 mg g−1, respectively) due to the inerratic pore distribution [154]. Gendy et al. found that the specific surface area and porosity of COFs profoundly affected their adsorption capacity for organic pollutants, and this work laid a foundation for the subsequent development of COFs with better performance [155].

7.2. pH Value and Charge Interaction

The influence of the pH and charge interaction is mainly reflected in the electrostatic interaction force [156]. The pH value has an important influence on the adsorption reaction process, which affects the existence form of antibiotics in the water environment (charged or uncharged). The Zeta potential of the adsorbent can judge the positive and negative charges of the adsorbent under different pH values and then adsorb antibiotics through electrostatic effect [157].
Xu et al. investigated the influence of the pH on the adsorption capacity of TAPA-TFPB-COFs for quinolone antibiotics (QAs). In their study, to a certain extent, the adsorption performance increased with the increase in the pH value. The recoveries of QAs were found to be relatively high and stable with little change in the range of 5.0–9.0, which reflected the influence of the pH value on the adsorption efficiency [158].

7.3. π–π Interaction and Designability of Functional Group

COFs not only extend the π conjugated system to the entire three-dimensional structure but also expand the delocalization range of π electrons while inhibiting the π–π interaction between conjugated units. Therefore, COFs have high electron mobility and superior selectivity [159]. Moreover, COFs can introduce adsorption sites on the monomer by introducing charged groups or groups containing lone pair electrons on the monomer through monomer design and using their mutual use with the ions to be adsorbed to introduce the adsorption sites. The introduction of corresponding functional group ion groups on the polymerized COFs fixes COFs with a good adsorption capacity in the adsorption process [160].
Guan, S. et al. synthesized magnetic COFs and studied their adsorption of fluoroquinolone antibiotics by introducing specific functional groups. The adsorption recovery of the fluoroquinolones was found to be as high as 90.4 to 101.2%, with a low detection limit (0.05 to 0.20 μg/L), and the enrichment factor reached 115.5–127.3, which benefitted from the related functional groups [161].

8. Conclusions and Perspective

In recent years, COF has attracted more and more attention in the fields of gas storage, photoelectric, catalysis, sensors, etc. Especially, COFs have become a very popular and cutting-edge research topic, providing strong technical support for many fields and showing huge application potential.
Moreover, COFs have excellent application prospects in the detection of antibiotics from aqueous solutions due to their high and regular porosity, large specific surface area, adjustable pore size, and excellent structural stability. The COFs can better separate the target substance from the coexisting matrix interfering substance, achieve better purification and enrichment, and have the advantages of high detection sensitivity, good reproducibility, and low detection cost.
Despite the remarkable advances in COF-based applications that have been achieved, several challenges should still be addressed. Firstly, the synthetic optimization of a COF hybrid needs to aim at reducing the cost of COFs and increasing chemical compatibility, which is urgent for large-scale synthesis and practical use. Secondly, due to the complex activity of antibiotics, novel functional components of COFs should be developed, and it is necessary to systematically optimize the composition of COFs, such as organic ligands and coordination center ions, for efficient and sensitive separation.

Author Contributions

Conceptualization, G.L., C.Z. and Y.H.; methodology, G.L. and C.Z.; software, Y.H.; validation, G.L., C.Z., Y.H. and Y.X.; formal analysis, Y.X. and C.Z.; investigation, J.H. and C.Z.; resources, J.H.; data curation, G.L. and C.Z.; writing—original draft preparation, G.L.; writing—review and editing, G.L.; visualization, G.L. and C.Z.; supervision, C.Z.; project administration, C.Z.; funding acquisition, G.L. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Natural Science Foundation of Shandong Province (No. ZR2021QB060 and ZR2021QB208) and a Linyi University Doctoral Research Foundation grant funded by Linyi University (No. LYDX2020BS027 and No. LYDX2020BS016).

Institutional Review Board Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Hutching, M.I.; Truman, A.W.; Wilkinson, B. Antibiotics: Past, present and future. Curr. Opin. Microbiol. 2019, 51, 72–80. [Google Scholar] [CrossRef] [PubMed]
  2. Wu, S.; Lin, Y.; Hu, Y. Strategies of tuning catalysts for efficient photodegradation of antibiotics in water environments: A review. J. Mater. Chem. A 2021, 9, 2592–2611. [Google Scholar] [CrossRef]
  3. Chow, L.K.M.; Ghaly, T.M.; Gillings, M.R. A survey of sub-inhibitory concentrations of antibiotics in the environment. J. Environ. Sci. 2021, 99, 21–27. [Google Scholar] [CrossRef] [PubMed]
  4. Zheng, D.; Yin, G.; Liu, M.; Chen, C.; Jiang, Y.; Hou, L.; Zheng, Y. A systematic review of antibiotics and antibiotic resistance genes in estuarine and coastal environments. Sci. Total Environ. 2021, 777, 146009–146026. [Google Scholar] [CrossRef]
  5. Kim, C.; Ryu, H.D.; Chung, E.G.; Kim, Y.; Lee, J. A review of analytical procedures for the simultaneous determination of medically important veterinary antibiotics in environmental water: Sample preparation, liquid chromatography, and mass spectrometry. J. Environ. Manag. 2018, 217, 629–645. [Google Scholar] [CrossRef]
  6. Bitew, Z.; Amare, M. Recent reports on electrochemical determination of selected antibiotics in pharmaceutical formulations: A mini review. Electrochem. Commun. 2020, 121, 106863–106876. [Google Scholar] [CrossRef]
  7. Zhang, C.; Deng, Y.; Zheng, J.; Zheng, J.; Zhang, Y.; Yang, L.; Liao, C.; Su, L.; Zhou, Y.; Gong, D.; et al. The application of the QuEChERS methodology in the determination of antibiotics in food: A review. Trends Anal. Chem. 2019, 118, 517–537. [Google Scholar] [CrossRef]
  8. Ahmed, S.; Ning, J.; Peng, D.; Chen, T.; Ahmad, I.; Ali, A.; Lei, Z.; Shabbir, M.A.; Cheng, G.; Yuan, Z. Current advances in immunoassays for the detection of antibiotics residues: A review. Food Agric. Immunol. 2020, 31, 268–290. [Google Scholar] [CrossRef]
  9. Du, C.; Zhang, Z.; Yu, G.; Wu, H.; Chen, H.; Zhou, L.; Zhang, Y.; Su, Y.; Tan, S.; Yang, L.; et al. A review of metal organic framework (MOFs)-based materials for antibiotics removal via adsorption and photocatalysis. Chemosphere 2021, 272, 129501. [Google Scholar] [CrossRef]
  10. Dong, J.; Han, X.; Liu, Y.; Li, H.; Cui, Y. Metal–covalent organic frameworks (MCOFS): A bridge between metal–organic frameworks and covalent organic frameworks. Angew. Chem. Int. Ed. 2020, 132, 13826–13837. [Google Scholar] [CrossRef]
  11. Zhang, P.; Wang, Z.; Cheng, P.; Chen, Y.; Zhang, Z. Design and application of ionic covalent organic frameworks. Coordin. Chem. Rev. 2021, 438, 213873. [Google Scholar] [CrossRef]
  12. Lin, S.; Diercks, C.S.; Zhang, Y.B.; Kornienko, N.; Nichols, E.M.; Zhao, Y.; Paris, A.R.; Kim, D.; Yang, P.; Yaghi, O.; et al. Covalent organic frameworks comprising cobalt porphyrins for catalytic CO2 reduction in water. Science 2015, 349, 1208–1213. [Google Scholar] [CrossRef] [PubMed]
  13. Huang, J.; Liu, X.; Zhang, W.; Liu, Z.; Zhong, H.; Shao, B.; Liang, Q.; Liu, Y.; Zhang, W.; He, Q. Functionalization of covalent organic frameworks by metal modification: Construction, properties and applications. Chem. Eng. J. 2021, 404, 127136–127159. [Google Scholar] [CrossRef]
  14. Valenzuela, C.; Chen, C.; Sun, M.; Ye, Z.; Zhang, J. Strategies and applications of covalent organic frameworks as promising nanoplatforms in cancer therapy. J. Mater. Chem. B 2021, 9, 3450–3483. [Google Scholar] [CrossRef] [PubMed]
  15. Guan, X.; Chen, F.; Fang, Q.; Qiu, S. Design and applications of three dimensional covalent organic frameworks. Chem. Soc. Rev. 2020, 49, 1357–1384. [Google Scholar] [CrossRef]
  16. Liu, Y.; Zhou, W.; Teo, W.L.; Wang, K.; Zhang, L.; Zeng, Y.; Zhao, Y. Covalent-organic-framework-based composite materials. Chem 2020, 6, 3172–3202. [Google Scholar] [CrossRef]
  17. Keller, N.; Bein, T. Optoelectronic processes in covalent organic frameworks. Chem. Soc. Rev. 2021, 50, 1813–1845. [Google Scholar] [CrossRef]
  18. Nguyen, Q.X.T.; Manh Khong, H.; Duc La, D.; Dang, T.D. Self-Assembly of the porphyrin monomer on the surface of Fe/graphene material: A novel sensing material for the detection of chloramphenicol antibiotic in aqueous solution. ChemPhysChem 2024, e202400355. [Google Scholar] [CrossRef]
  19. Zhang, Z.; Zhang, H.; Tian, D.; Phan, A.; Seididamyeh, M.; Alanazi, M.; Xu, Z.; Sultanbawa, Y.; Zhang, R. Luminescent sensors for residual antibiotics detection in food: Recent advances and perspectives. Coord. Chem. Rev. 2024, 498, 215455–215482. [Google Scholar] [CrossRef]
  20. Tan, G.; Wang, S.; Yu, J.; Chen, J.; Liao, D.; Liu, M.; Nezamzadeh-Ejhieh, A.; Pan, Y.; Liu, J. Detection mechanism and the outlook of metal-organic frameworks for the detection of hazardous substances in milk. Food Chem. 2024, 430, 136934. [Google Scholar] [CrossRef]
  21. Ji, M.; Li, J.; Liu, A.; Ma, D. Covalent organic frameworks-based materials for antibiotics fluorescence detection. Heliyon 2024, 10, e33118–e33132. [Google Scholar] [CrossRef] [PubMed]
  22. Pilli, P.; Kommalapati, H.S.; Golla, V.M.; Khemchandani, R.; Ramachandran, R.K.; Samanthula, G. Covalent organic frameworks: Spotlight on applications in the pharmaceutical arena. Bioanalysis 2024, 16, 279–305. [Google Scholar] [CrossRef] [PubMed]
  23. Fenneman, A.C.; Weidner, M.; Chen, L.A.; Nieuwdorp, M.; Blaser, M. Antibiotics in the pathogenesis of diabetes and inflammatory diseases of the gastrointestinal tract. Nat. Rev. Gastro. Hepatol. 2023, 20, 81–100. [Google Scholar] [CrossRef] [PubMed]
  24. Li, M.; Yang, L.; Yen, H.; Zhao, F.; Wang, X.; Zhou, T.; Feng, Q.; Chen, L. Occurrence, spatial distribution and ecological risks of antibiotics in soil in urban agglomeration. J. Environ. Sci. 2023, 125, 678–690. [Google Scholar] [CrossRef] [PubMed]
  25. Tian, Y.; Li, J.; Tang, L.; Meng, J.; Li, J. Antibiotics removal from piggery wastewater by a novel aerobic-microaerobic system: Efficiency and mechanism. Chem. Eng. J. 2023, 454, 140265. [Google Scholar] [CrossRef]
  26. Fu, Y.; Zhu, Y.; Dong, H.; Li, J.; Zhang, W.; Shao, Y.; Shao, Y. Effects of heavy metals and antibiotics on antibiotic resistance genes and microbial communities in soil. Process Saf. Environ. 2023, 169, 418–427. [Google Scholar] [CrossRef]
  27. Butler, J.; Handy, R.D.; Upton, M.; Besinis, A. Review of antimicrobial nanocoatings in medicine and dentistry: Mechanisms of action, biocompatibility performance, safety, and benefits compared to antibiotics. ACS Nano 2023, 17, 7064–7092. [Google Scholar] [CrossRef] [PubMed]
  28. Nguyen, T.D.; Lee, T.; Van Tran, T.; Nguyen, V.; Nong, L.; Giang, L.; Vo, B. Multicomponent photocatalysts for synergic removal of antibiotics in aqueous media: A review. Environ. Chem. Lett. 2023, 21, 935–980. [Google Scholar] [CrossRef]
  29. Jumpertz, M.; Guilhaumou, R.; Million, M.; Parola, P.; Lagier, J.; Brouqui, P.; Cassir, N. Subcutaneously administered antibiotics: A review. J. Antimicrob. Chemother. 2023, 78, 1–7. [Google Scholar] [CrossRef]
  30. Le, V.V.; Tran, Q.G.; Ko, S.R.; Lee, S.; Oh, H.; Kim, H.; Ahn, C. How do freshwater microalgae and cyanobacteria respond to antibiotics? Crit. Rev. Biotechnol. 2023, 43, 191–211. [Google Scholar] [CrossRef] [PubMed]
  31. Gupta, A.; Vyas, R.K.; Vyas, S. A review on antibiotics pervasiveness in the environment and their removal from wastewater. Sep. Sci. Technol. 2023, 58, 326–344. [Google Scholar] [CrossRef]
  32. Wu, J.; Wang, J.; Li, Z.; Guo, S.; Li, K.; Xu, P.; Ok, Y.; Jones, D.; Zou, J. Antibiotics and antibiotic resistance genes in agricultural soils: A systematic analysis. Crit. Rev. Environ. Sci. Technol. 2023, 53, 847–864. [Google Scholar] [CrossRef]
  33. Mueller, U.; Schubert, M.; Teich, F.; Puetter, H.; Schierle-Arndt, K.; Pastré, J. Metal–organic frameworks—Prospective industrial applications. J. Mater. Chem. 2006, 16, 626–636. [Google Scholar] [CrossRef]
  34. Arqueros, C.; Zamora, F.; Montoro, C. A perspective on the application of covalent organic frameworks for detection and water treatment. Nanomaterials 2021, 11, 1651. [Google Scholar] [CrossRef] [PubMed]
  35. Guo, L.; Yang, L.; Li, M.; Kuang, L.; Song, Y.; Wang, L. Covalent organic frameworks for fluorescent sensing: Recent developments and future challenges. Coord. Chem. Rev. 2021, 440, 213957. [Google Scholar] [CrossRef]
  36. Gong, Y.N.; Guan, X.; Jiang, H.L. Covalent organic frameworks for photocatalysis: Synthesis, structural features, fundamentals and performance. Coord. Chem. Rev. 2023, 475, 214889. [Google Scholar] [CrossRef]
  37. Wang, J.; Zhuang, S. Covalent organic frameworks (COFs) for environmental applications. Coord. Chem. Rev. 2019, 400, 213046. [Google Scholar] [CrossRef]
  38. Diercks, C.S.; Yaghi, O.M. The atom, the molecule, and the covalent organic framework. Science 2017, 355, eaal1585. [Google Scholar] [CrossRef]
  39. Lohse, M.S.; Bein, T. Covalent organic frameworks: Structures, synthesis, and applications. Adv. Funct. Mater. 2018, 28, 1705553–1705624. [Google Scholar] [CrossRef]
  40. Kandambeth, S.; Dey, K.; Banerjee, R. Covalent organic frameworks: Chemistry beyond the structure. J. Am. Chem. Soc. 2018, 141, 1807–1822. [Google Scholar] [CrossRef] [PubMed]
  41. Yang, J.; Tu, B.; Zhang, G.; Liu, P.; Hu, K.; Wang, J.; Yang, Z.; Huang, Z.; Fang, M.; Hou, J.; et al. Advancing osmotic power generation by covalent organic framework monolayer. Nat. Nanotechnol. 2022, 176, 22–628. [Google Scholar] [CrossRef] [PubMed]
  42. Lu, M.; Zhang, M.; Liu, J.; Yu, T.; Chang, J.; Shang, L.; Li, S.; Lan, Y. Confining and highly dispersing single polyoxometalate clusters in covalent organic frameworks by covalent linkages for CO2 photoreduction. J. Am. Chem. Soc. 2022, 144, 1861–1871. [Google Scholar] [CrossRef] [PubMed]
  43. Huang, N.; Wang, P.; Jiang, D. Covalent organic frameworks: A materials platform for structural and functional designs. Nat. Rev. Mater. 2016, 1, 1–19. [Google Scholar] [CrossRef]
  44. Chen, L.; Furukawa, K.; Gao, J.; Nagai, A.; Nakamura, T.; Dong, Y.; Jiang, D. Photoelectric covalent organic frameworks: Converting open lattices into ordered donor–acceptor heterojunctions. J. Am. Chem. Soc. 2014, 136, 9806–9809. [Google Scholar] [CrossRef] [PubMed]
  45. Yang, X.; Li, X.; Liu, M.; Yang, S.; Xu, Q.; Zeng, G. Quantitative construction of boronic-ester linkages in covalent organic frameworks for the carbon dioxide reduction. Angew. Chem. 2024, 136, e202317785. [Google Scholar] [CrossRef]
  46. Yue, L.; Wang, X.; Chen, L.; Shen, D.; Shao, Z.; Wu, H.; Xiao, S.; Liang, W.; Yu, Y.; Li, Y. In situ interface engineering of highly nitrogen-rich triazine-based covalent organic frameworks for an ultra-stable, dendrite-free lithium-metal anode. Energy Environ. Sci. 2024, 17, 1117–1131. [Google Scholar] [CrossRef]
  47. Tilford, R.W.; Gemmill, W.R.; zur Loye, H.C.; Lavigne, J.J. Facile synthesis of a highly crystalline, covalently linked porous boronate network. Chem. Mater. 2006, 18, 5296–5301. [Google Scholar] [CrossRef]
  48. Yaghi, O.M.; O’Keeffe, M.; Ockwig, N.W.; Chae, H.K.; Eddaoudi, M.; Kim, J. Reticular synthesis and the design of new materials. Nature 2003, 423, 705–714. [Google Scholar] [CrossRef]
  49. Cote, A.P.; Benin, A.I.; Ockwig, N.W.; O’Keeffe, M.; Matzger, A.J.; Yaghi, O.M. Porous, crystalline, covalent organic frameworks. Science 2005, 310, 1166–1170. [Google Scholar] [CrossRef]
  50. Liu, Y.; Feng, Y.; Huang, Z.; Zhang, X. Thermal conductivity of 3D boron-based covalent organic frameworks from molecular dynamics simulations. J. Phys. Chem. C 2016, 120, 17060–17068. [Google Scholar] [CrossRef]
  51. Li, H.; Li, H.; Dai, Q.; Li, H.; Bredas, J.L. Hydrolytic stability of boronate ester-linked covalent organic frameworks. Adv. Theory Simul. 2018, 1, 1700015–1700024. [Google Scholar] [CrossRef]
  52. Bao, T.; Tang, P.; Kong, D.; Mao, Z.; Chen, Z. Polydopamine-supported immobilization of covalent-organic framework-5 in capillary as stationary phase for electrochromatographic separation. J. Chromatogr. A 2016, 1445, 140–148. [Google Scholar] [CrossRef] [PubMed]
  53. Kuhn, P.; Antonietti, M.; Thomas, A. Porous, covalent triazine-based frameworks prepared by ionothermal synthesis. Angew. Chem. Int. Ed. 2008, 47, 3450–3453. [Google Scholar] [CrossRef] [PubMed]
  54. Huang, W.; Byun, J.; Rörich, I.; Ramanan, C.; Blom, P.W.M.; Lu, H.; Wang, D.; da Silva, L.C.; Li, R.; Wang, L.; et al. Asymmetric covalent triazine framework for enhanced visible-light photoredox catalysis via energy transfer cascade. Angew. Chem. Int. Ed. 2018, 57, 8316–8320. [Google Scholar] [CrossRef] [PubMed]
  55. Lopez-Magano, A.; Salaverri, N.; Marzo, L.; Mas-Ballest´e, R.; Aleman´, J. Synergistic combination of triazine and phenanthroline moieties in a covalent triazine framework tailored for heterogeneous photocatalytic metal-free C-Br and C-Cl activation. Appl. Catal. B Environ. 2022, 317, 121791–121801. [Google Scholar] [CrossRef]
  56. Li, L.; Lu, F.; Xue, R.; Ma, B.; Li, Q.; Wu, N.; Liu, H.; Yao, W.; Guo, H.; Yang, W. Ultrastable triazine-based covalent organic framework with an interlayer hydrogen bonding for supercapacitor applications. ACS Appl. Mater. Interfaces 2019, 11, 26355–26363. [Google Scholar] [CrossRef] [PubMed]
  57. Liao, H.; Ding, H.; Li, B.; Ai, X.; Wang, C. Covalent organic frameworks: Potential host materials for sulfur impregnation in lithium–sulfur batteries. J. Mater. Chem. A 2014, 2, 8854–8858. [Google Scholar] [CrossRef]
  58. Yang, C.; Liu, C.; Cao, Y.; Yan, X. Facile room-temperature solution-phase synthesis of a spherical covalent organic framework for high-resolution chromatographic separation. Chem. Commun. 2015, 51, 12254–12257. [Google Scholar] [CrossRef] [PubMed]
  59. Han, X.; Huang, J.; Yuan, C.; Liu, Y.; Cui, Y. Chiral 3D covalent organic frameworks for high performance liquid chromatographic enantioseparation. J. Am. Chem. Soc. 2018, 140, 892–895. [Google Scholar] [CrossRef]
  60. DeBlase, C.R.; Dichtel, W.R. Moving beyond boron: The emergence of new linkage chemistries in covalent organic frameworks. Macromolecules 2016, 49, 5297–5305. [Google Scholar] [CrossRef]
  61. Uribe-Romo, F.J.; Hunt, J.R.; Furukawa, H.; Klöck, C.; O’Keeffe, M.; Yaghi, O.M. A Crystalline imine-linked 3-D porous covalent organic framework. J. Am. Chem. Soc. 2009, 131, 4570–4571. [Google Scholar] [CrossRef]
  62. Fang, Q.; Zhuang, Z.; Gu, S.; Kaspar, R.B.; Zheng, J.; Wang, J.; Qiu, S.; Yan, Y. Designed synthesis of large-pore crystalline polyimide covalent organic frameworks. Nat. Commun. 2014, 5, 1–8. [Google Scholar] [CrossRef]
  63. Jin, S.; Sakurai, T.; Kowalczyk, T.; Dalapati, S.; Xu, F.; Wei, H.; Chen, X.; Gao, J.; Seki, S.; Irle, S.; et al. Two-dimensional tetrathiafulvalene covalent organic frameworks: Towards latticed conductive organic salts. Chem. Eur. J. 2014, 20, 14608–14613. [Google Scholar] [CrossRef] [PubMed]
  64. Nailwal, Y.; Wonanke, A.D.D.; Addicoat, M.A.; Pal, S.K. A dual-function highly crystalline covalent organic framework for HCl sensing and visible-light heterogeneous photocatalysis. Macromolecules 2021, 54, 6595–6604. [Google Scholar] [CrossRef]
  65. Wang, Y.; Liu, Y.; Li, H.; Guan, X.; Xue, M.; Yan, Y.; Valtchev, V.; Qiu, S.; Fang, Q. Three-dimensional mesoporous covalent organic frameworks through steric hindrance engineering. J. Am. Chem. Soc. 2020, 142, 3736–3741. [Google Scholar] [CrossRef]
  66. Stegbauer, L.; Schwinghammer, K.; Lotsch, B.V. A hydrazone-based covalent organic framework for photocatalytic hydrogen production. Chem. Sci. 2014, 5, 2789–2793. [Google Scholar] [CrossRef]
  67. Zhang, K.; Cai, S.; Yan, Y.; He, Z.; Lin, H.; Huang, X.; Zheng, S.; Fan, J.; Zhang, W. Construction of a hydrazone-linked chiral covalent organic framework–silica composite as the stationary phase for high performance liquid chromatography. J. Chromatogr. A 2017, 1519, 100–109. [Google Scholar] [CrossRef]
  68. Li, Z.; Feng, X.; Zou, Y.; Zhang, Y.; Xia, H.; Liu, X.; Mu, Y. A 2D azine-linked covalent organic framework for gas storage applications. Chem. Commun. 2014, 50, 13825–13828. [Google Scholar] [CrossRef]
  69. Dalapati, S.; Jin, S.; Gao, J.; Xu, Y.; Nagai, A. An azine-linked covalent organic framework. J. Am. Chem. Soc. 2013, 135, 17310–17313. [Google Scholar] [CrossRef]
  70. Traxler, M.; Gisbertz, S.; Pachfule, P.; Schmidt, J.; Roeser, J.; Reischauer, S.; Rabeah, S.; Rabeah, J.; Pieber, B. Acridine-functionalized covalent organic frameworks (COFs) as photocatalysts for metallaphotocatalytic C-N cross-coupling. Angew. Chem. Int. Ed. 2022, 61, e202117738. [Google Scholar] [CrossRef]
  71. DeBlase, C.R.; Silberstein, K.E.; Truong, T.T.; Abruña, H.D.; Dichtel, W.R. β-Ketoenamine-linked covalent organic frameworks capable of pseudocapacitive energy storage. J. Am. Chem. Soc. 2013, 135, 16821–16824. [Google Scholar] [CrossRef]
  72. Liu, F.; Zhou, P.; Hou, Y.; Tan, H.; Liang, Y.; Liang, J.; Zhang, Q.; Guo, S.; Tong, M.; Ni, J. Covalent organic frameworks for direct photosynthesis of hydrogen peroxide from water, air and sunlight. Nat. Commun. 2013, 14, 4344–4354. [Google Scholar] [CrossRef] [PubMed]
  73. Zhang, S.; Zheng, Y.; An, H.; Aguila, B.; Yang, C.; Dong, Y.; Xie, W.; Cheng, P.; Zhang, Z.; Chen, Y.; et al. Covalent organic frameworks with chirality enriched by biomolecules for efficient chiral separation. Angew. Chem. Int. Ed. 2018, 57, 16754–16759. [Google Scholar] [CrossRef]
  74. Abuzeid, H.R.; EL-Mahdy, A.F.M.; Kuo, S.W. Covalent organic frameworks: Design principles, synthetic strategies, and diverse applications. Giant 2021, 6, 100054–100081. [Google Scholar] [CrossRef]
  75. Li, Y.; Chen, W.; Xing, G.; Jiang, D.; Chen, L. New synthetic strategies toward covalent organic frameworks. Chem. Soc. Rev. 2020, 49, 2852–2868. [Google Scholar] [CrossRef]
  76. Ding, S.Y.; Wang, W. Covalent organic frameworks (COFs): From design to applications. Chem. Soc. Rev. 2013, 42, 548–568. [Google Scholar] [CrossRef]
  77. Bisbey, R.P.; Dichtel, W.R. Covalent organic frameworks as a platform for multidimensional polymerization. ACS Cent. Sci. 2017, 3, 533–543. [Google Scholar] [CrossRef]
  78. Bhadra, M.; Kandambeth, S.; Sahoo, M.K.; Addicoat, M.; Balaraman, E.; Banerjee, R. Triazine functionalized porous covalent organic framework for photo-organocatalytic E–Z isomerization of olefins. J. Am. Chem. Soc. 2019, 141, 6152–6156. [Google Scholar] [CrossRef]
  79. Sick, T.; Rotter, J.M.; Reuter, S.; Kandambeth, S.; Bach, N.N.; Döblinger, M.; Merz, J.; Clark, T.; Marder, T.B.; Bein, T.; et al. Switching on and off interlayer correlations and porosity in 2D covalent organic frameworks. J. Am. Chem. Soc. 2019, 141, 12570–12581. [Google Scholar] [CrossRef]
  80. Maschita, J.; Banerjee, T.; Savasci, G.; Haase, F.; Ochsenfeld, C.; Lotsch, B.V. Ionothermal synthesis of imide-linked covalent organic frameworks. Angew. Chem. Int. Ed. 2020, 59, 15750–15758. [Google Scholar] [CrossRef] [PubMed]
  81. Guan, X.; Ma, Y.; Li, H.; Yusran, Y.; Xue, M.; Fang, Q.; Yan, Y.; Valtchev, V.; Qiu, S. Fast, ambient temperature and pressure ionothermal synthesis of three-dimensional covalent organic frameworks. J. Am. Chem. Soc. 2018, 140, 4494–4498. [Google Scholar] [CrossRef] [PubMed]
  82. Dong, B.; Wang, W.; Pan, W.; Kang, G. Ionic liquid as a green solvent for ionothermal synthesis of 2D keto-enamine-linked covalent organic frameworks. Mater. Chem. Phys. 2019, 226, 244–249. [Google Scholar] [CrossRef]
  83. Dąbrowska, S.; Chudoba, T.; Wojnarowicz, J.; Lojkowski, W. Current trends in the development of microwave reactors for the synthesis of nanomaterials in laboratories and industries: A review. Crystals 2018, 8, 379. [Google Scholar] [CrossRef]
  84. Kumar, S.; Kulkarni, V.V.; Jangir, R. Covalent-organic framework composites: A review report on synthesis methods. ChemistrySelect 2021, 6, 11201–11223. [Google Scholar] [CrossRef]
  85. Sharma, R.K.; Yadav, P.; Yadav, M.; Gupta, R.; Rana, P.; Srivastava, A.; Zbořil, R.; Varma, R.S.; Antonietti, M.; Gawande, M.B. Recent development of covalent organic frameworks (COFs): Synthesis and catalytic (organic-electro-photo) applications. Mater. Horiz. 2020, 7, 411–454. [Google Scholar] [CrossRef]
  86. Nain, S.; Singh, R.; Ravichandran, S. Importance of microwave heating in organic synthesis. Adv. J. Chem. A 2019, 2, 94–104. [Google Scholar] [CrossRef]
  87. Zhang, F.; Zhang, J.; Zhang, B.; Tan, X.; Shao, D.; Shi, J.; Tan, D.; Liu, L.; Feng, J.; Han, B.; et al. Room-temperature synthesis of covalent organic framework (COF-LZU1) nanobars in CO2/water solvent. ChemSusChem 2018, 11, 3576–3580. [Google Scholar] [CrossRef]
  88. Medina, D.D.; Rotter, J.M.; Hu, Y.; Dogru, M.; Werner, V.; Auras, F.; Markiewicz, J.T.; Knochel, P.; Bein, T. Room temperature synthesis of covalent–organic framework films through vapor-assisted conversion. J. Am. Chem. Soc. 2015, 137, 1016–1019. [Google Scholar] [CrossRef] [PubMed]
  89. Bagheri, A.R.; Aramesh, N. Towards the room-temperature synthesis of covalent organic frameworks: A mini-review. J. Mater. Sci. 2021, 56, 1116–1132. [Google Scholar] [CrossRef]
  90. Song, Y.; Sun, Q.; Aguila, B.; Ma, S. Opportunities of covalent organic frameworks for advanced applications. Adv. Sci. 2019, 6, 1801410. [Google Scholar] [CrossRef]
  91. Liu, C.; Yu, Y.; Zhang, W.; Zeng, Q.; Lei, S. Room-temperature synthesis of covalent organic frameworks with a boronic ester linkage at the liquid/solid interface. Chem. Eur. J. 2016, 22, 18412–18418. [Google Scholar] [CrossRef]
  92. Liu, R.; Yan, Q.; Tang, Y.; Liu, R.; Huang, L.; Shuai, Q. NaCl template-assisted synthesis of self-floating COFs foams for the efficient removal of sulfamerazine. J. Hazard. Mater. 2022, 421, 126702–126710. [Google Scholar] [CrossRef] [PubMed]
  93. Peng, Y.; Wong, W.K.; Hu, Z.; Cheng, Y.; Yuan, D.; Khan, S.A.; Zhao, D. Room temperature batch and continuous flow synthesis of water-stable covalent organic frameworks (COFs). Chem. Mater. 2016, 28, 5095–5101. [Google Scholar] [CrossRef]
  94. You, J.; Zhao, Y.; Wang, L.; Bao, W. Recent developments in the photocatalytic applications of covalent organic frameworks: A review. J. Clean. Prod. 2021, 291, 125822. [Google Scholar] [CrossRef]
  95. Li, X.; Yang, C.; Sun, B.; Cai, S.; Chen, Z.; Lv, Y.; Zhang, J.; Liu, Y. Expeditious synthesis of covalent organic frameworks: A review. J. Mater. Chem. A 2020, 8, 16045–16060. [Google Scholar] [CrossRef]
  96. Raptakis, A.; Dianat, A.; Croy, A.; Cuniberti, G. Predicting the bulk modulus of single-layer covalent organic frameworks with square-lattice topology from molecular building-block properties. Nanoscale 2021, 13, 1077–1085. [Google Scholar] [CrossRef]
  97. Chen, X.; Addicoat, M.; Jin, E.; Zhai, L.; Xu, H.; Huang, N.; Guo, Z.; Liu, L.; Irle, S.; Jiang, D. Locking covalent organic frameworks with hydrogen bonds: General and remarkable effects on crystalline structure, physical properties, and photochemical activity. J. Am. Chem. Soc. 2015, 137, 3241–3247. [Google Scholar] [CrossRef] [PubMed]
  98. Torabi, E.; Mirzaei, M.; Bazargan, M.; Amiri, A. A critical review of covalent organic frameworks-based sorbents in extraction methods. Anal. Chim. Acta 2022, 1224, 340207. [Google Scholar] [CrossRef] [PubMed]
  99. Xie, Z.; Wang, B.; Yang, Z.; Yang, X.; Yu, X.; Xing, G.; Zhang, Y.; Chen, L. Stable 2D heteroporous covalent organic frameworks for efficient ionic conduction. Angew. Chem. Int. Ed. 2019, 58, 15742–15746. [Google Scholar] [CrossRef]
  100. Chen, H.; Yuan, X.; Wang, H.; Yu, H.; Jiang, L. Nanostructured covalent organic frameworks with elevated crystallization for (electro) photocatalysis and energy storage devices. J. Mater. Sci. 2021, 56, 13875–13924. [Google Scholar] [CrossRef]
  101. He, G.; Zhang, R.; Jiang, Z. Engineering covalent organic framework membranes. Acc. Mater. Res. 2021, 2, 630–643. [Google Scholar] [CrossRef]
  102. Rodríguez-San-Miguel, D.; Zamora, F. Processing of covalent organic frameworks: An ingredient for a material to succeed. Chem. Soc. Rev. 2019, 48, 4375–4386. [Google Scholar] [CrossRef] [PubMed]
  103. Shi, Y.; Yang, J.; Gao, F.; Zhang, Q. Covalent organic frameworks: Recent progress in biomedical applications. ACS Nano 2023, 17, 1879–1905. [Google Scholar] [CrossRef] [PubMed]
  104. Shahzad, U.; Marwani, H.M.; Saeed, M.; Asiri, A.M.; Repon, M.R.; Althomali, R.H.; Rahman, M.M. Progress and Perspectives on promising covalent-organic frameworks (COFs) materials for energy storage capacity. Chem. Rec. 2024, 24, e202300285. [Google Scholar] [CrossRef]
  105. Liang, R.; Cui, F.; Qi, Q.; Zhao, X. A study on constitutional isomerism in covalent organic frameworks: Controllable synthesis, transformation, and distinct difference in properties. CCS Chem. 2020, 2, 139–145. [Google Scholar] [CrossRef]
  106. Li, Y.; Guo, X.; Li, X.; Zhang, M.; Jia, Z.; Deng, Y.; Tian, Y.; Li, S.; Ma, L. Redox-active two-dimensional covalent organic frameworks (COFs) for selective reductive separation of valence-variable, redox-sensitive and long-lived radionuclides. Angew. Chem. Int. Ed. 2020, 132, 4197–4204. [Google Scholar] [CrossRef]
  107. Sasmal, H.S.; Kumar Mahato, A.; Majumder, P.; Banerjee, R. Landscaping covalent organic framework nanomorphologies. J. Am. Chem. Soc. 2022, 144, 11482–11498. [Google Scholar] [CrossRef]
  108. Wang, G.; Li, Y.; Shi, W.; Zhang, B.; Hou, L.; Wang, Y. A robust cluster-based Eu-MOF as multi-functional fluorescence sensor for detection of antibiotics and pesticides in water. Sens. Actuators B Chem. 2021, 331, 129377–129386. [Google Scholar] [CrossRef]
  109. Xu, Q.; Tan, Z.; Liao, X.; Wang, C. Recent advances in nanoscale metal-organic frameworks biosensors for detection of biomarkers. Chin. Chem. Lett. 2021, 5, 52–63. [Google Scholar] [CrossRef]
  110. Zhang, X.; Li, G.; Wu, D.; Zhang, B.; Hu, N.; Wang, H.; Liu, J.; Wu, Y. Recent advances in the construction of functionalized covalent organic frameworks and their applications to sensing. Biosens. Bioelectron. 2019, 145, 111699–111718. [Google Scholar]
  111. Yao, D.; Li, C.; Wang, H.; Wen, G.; Liang, A.; Jiang, Z. A new dual-mode SERS and RRS aptasensor for detecting trace organic molecules based on gold nanocluster-doped covalent-organic framework catalyst. Sens. Actuators B Chem. 2020, 319, 128308–128319. [Google Scholar] [CrossRef]
  112. Wang, M.; Hu, M.; Liu, J.; Guo, C.; Peng, D.; Jia, Q.; He, L.; Zhang, Z.; Du, M. Covalent organic framework-based electrochemical aptasensors for the ultrasensitive detection of antibiotics. Biosens. Bioelectron. 2019, 132, 8–16. [Google Scholar] [CrossRef] [PubMed]
  113. Sun, Y.; Waterhouse, G.I.N.; Xu, L.; Qiao, X.; Xu, Z. Three-dimensional electrochemical sensor with covalent organic framework decorated carbon nanotubes signal amplification for the detection of furazolidone. Sens. Actuators B Chem. 2020, 321, 128501–128535. [Google Scholar] [CrossRef]
  114. Yang, Y.; Li, G.; Wang, P.; Fan, L.; Shi, Y. Highly sensitive multiplex detection of foodborne pathogens using a SERS immunosensor combined with novel covalent organic frameworks based biologic interference-free Raman tags. Talanta 2022, 243, 123369. [Google Scholar] [CrossRef] [PubMed]
  115. Zhang, X.; Chi, K.; Li, D.; Deng, Y.; Ma, Y.; Xu, Q.; Hu, R.; Yang, Y. 2D-porphrinic covalent organic framework-based aptasensor with enhanced photoelectrochemical response for the detection of C-reactive protein. Biosens. Bioelectron. 2019, 129, 64–71. [Google Scholar] [CrossRef] [PubMed]
  116. Ramos, L. Critical overview of selected contemporary sample preparation techniques. J. Chromatogr. A 2012, 1221, 84–98. [Google Scholar] [CrossRef] [PubMed]
  117. Liu, H.; Jin, P.; Zhu, F.; Nie, L.; Qiu, H. A review on the use of ionic liquids in preparation of molecularly imprinted polymers for applications in solid-phase extraction. Trends Anal. Chem. 2021, 134, 116132–116186. [Google Scholar] [CrossRef]
  118. Haniffa, M.A.C.M.; Ching, Y.C.; Illias, H.A.; Munawar, K.; Ibrahim, S.; Nguyen, D.H.; Chuah, C.H. Cellulose supported promising magnetic sorbents for magnetic solid-phase extraction: A review. Carbohyd. Polym. 2021, 253, 117245. [Google Scholar] [CrossRef]
  119. Fenton, J.L.; Burke, D.W.; Qian, D.; de la Cruz, M.O.; Dichtel, W.R. Polycrystalline covalent organic framework films act as adsorbents, not membranes. J. Am. Chem. Soc. 2021, 143, 1466–1473. [Google Scholar] [CrossRef]
  120. Li, N.; Du, J.; Wu, D.; Liu, J.; Li, N.; Sun, Z.; Li, G.; Wu, Y. Recent advances in facile synthesis and applications of covalent organic framework materials as superior adsorbents in sample pretreatment. Trends Anal. Chem. 2018, 108, 154–166. [Google Scholar] [CrossRef]
  121. Majd, M.; Yazdanpanah, M.; Bayatloo, M.R.; Nojavan, S. Recent advances and applications of cyclodextrins in magnetic solid phase extraction. Talanta 2021, 229, 122296. [Google Scholar] [CrossRef] [PubMed]
  122. Sajid, M.; Nazal, M.K.; Ihsanullah, I. Novel materials for dispersive (micro) solid-phase extraction of polycyclic aromatic hydrocarbons in environmental water samples: A review. Anal. Chim. Acta 2021, 1141, 246–262. [Google Scholar] [CrossRef] [PubMed]
  123. Han, S.; Leng, Q.; Teng, F.; Ding, Y.; Yao, A. Preparation of mesh covalent organic framework Tppa-2-based adsorption enhanced magnetic molecularly imprinted composite for selective extraction of tetracycline residues from animal-derived foods. Food Chem. 2022, 384, 132601. [Google Scholar] [CrossRef] [PubMed]
  124. Wang, G.; Zhou, T.; Lei, Y. Exploration of a novel triazine-based covalent organic framework for solid-phase extraction of antibiotics. RSC Adv. 2020, 10, 11557–11564. [Google Scholar] [CrossRef]
  125. Wang, R.; Li, C.; Li, Q.; Zhang, S.; Lv, F.; Yan, Z. Electrospinning fabrication of covalent organic framework composite nanofibers for pipette tip solid phase extraction of tetracycline antibiotics in grass carp and duck. J. Chromatogr. A 2020, 1622, 461098–461106. [Google Scholar] [CrossRef] [PubMed]
  126. Wen, L.; Liu, L.; Wang, X.; Wang, M.; Lin, J.; Zhao, R. Spherical mesoporous covalent organic framework as a solid-phase extraction adsorbent for the ultrasensitive determination of sulfonamides in food and water samples by liquid chromatography-tandem mass spectrometry. J. Chromatogr. A 2020, 1625, 461275–461305. [Google Scholar] [CrossRef]
  127. Wen, A.; Li, G.; Wu, D.; Yu, Y.; Yang, Y.; Hu, N.; Wang, H.; Chen, J.; Wu, Y. Sulphonate functionalized covalent organic framework-based magnetic sorbent for effective solid phase extraction and determination of fluoroquinolones. J. Chromatogr. A 2020, 1612, 460651–460685. [Google Scholar] [CrossRef]
  128. Belardi, R.; Pawliszyn, J. Application of chemically modified fused silica fibers in the extraction of organics from water matrix samples and their rapid transfer to capillary columns. Water Pollut. Res. J. Can. 1989, 24, 179–191. [Google Scholar] [CrossRef]
  129. Zheng, J.; Yang, Q.; Ni, C.; Xie, X.; Shi, Y.; Sun, J.; Zhu, F.; Ouyang, G. Fabrications of novel solid phase microextraction fiber coatings based on new materials for high enrichment capability. Trends Anal. Chem. 2018, 108, 135–153. [Google Scholar] [CrossRef]
  130. Feng, J.; Feng, J.; Ji, X.; Li, C.; Han, S.; Sun, H.; Sun, M. Recent advances of covalent organic frameworks for solid-phase microextraction. Trends Anal. Chem. 2021, 137, 116208–116214. [Google Scholar] [CrossRef]
  131. Zhang, Y.; Ning, Y.; Liao, W.; Wang, W.; Wang, A. Covalent organic framework-LZU1@ PEI@ Fe3O4-based magnetic dispersive micro-solid phase extraction of tetracyclines from environmental water prior to HPLC analysis. Anal. Methods 2021, 13, 4320–4327. [Google Scholar] [CrossRef]
  132. Yue, Q.; Huang, Y.; Shen, X.; Yang, C.; Pang, Y. In situ growth of covalent organic framework on titanium fiber for headspace solid-phase microextraction of 11 phthalate esters in vegetables. Food Chem. 2020, 318, 126507. [Google Scholar] [CrossRef]
  133. Guo, J.; Qian, H.; Zhao, X.; Yang, C.; Yan, X. In situ room-temperature fabrication of a covalent organic framework and its bonded fiber for solid-phase microextraction of polychlorinated biphenyls in aquatic products. J. Mater. Chem. A 2019, 7, 13249–13255. [Google Scholar] [CrossRef]
  134. Ji, W.; Guo, Y.; Xie, H.; Wang, X.; Jiang, X.; Guo, D. Rapid microwave synthesis of dioxin-linked covalent organic framework for efficient micro-extraction of perfluorinated alkyl substances from water. J. Hazard. Mater. 2020, 397, 122793–122801. [Google Scholar] [CrossRef] [PubMed]
  135. Szultka-Mlynska, M.; Pomastowski, P.; Buszewski, B. Application of solid phase microextraction followed by liquid chromatography-mass spectrometry in the determination of antibiotic drugs and their metabolites in human whole blood and tissue samples. J. Chromatogr. B 2018, 1086, 153–165. [Google Scholar] [CrossRef] [PubMed]
  136. Sun, T.; Fan, R.; Zhang, J.; Qin, M.; Chen, W.; Jiang, X.; Zhu, K.; Ji, C.; Hao, S.; Yang, Y. Stimuli-responsive metal–organic framework on a metal–organic framework heterostructure for efficient antibiotic detection and anticounterfeiting. ACS Appl. Mater. Interfaces 2021, 13, 35689–35699. [Google Scholar] [CrossRef] [PubMed]
  137. Lu, D.; Liu, C.; Qin, M.; Deng, J.; Shi, G.; Zhou, T. Functionalized ionic liquids-supported metal organic frameworks for dispersive solid phase extraction of sulfonamide antibiotics in water samples. Anal. Chim. Acta 2020, 1133, 88–98. [Google Scholar] [CrossRef] [PubMed]
  138. Ayankojo, A.G.; Reut, J.; Ciocan, V.; Öpik, A.; Syritski, V. Molecularly imprinted polymer-based sensor for electrochemical detection of erythromycin. Talanta 2020, 209, 120502–120511. [Google Scholar] [CrossRef]
  139. Zhou, T.; Halder, A.; Sun, Y. Fluorescent nanosensor based on molecularly imprinted polymers coated on graphene quantum dots for fast detection of antibiotics. Biosensors 2018, 8, 82. [Google Scholar] [CrossRef]
  140. Jamieson, O.; Soares, T.C.C.; de Faria, B.A.; Husaon, A.; Mecozzi, F.; Rowley-Neale, S.J.; Banks, C.E.; Gruber, J.; Novakovic, K.; Peeters, M.; et al. Screen printed electrode based detection systems for the antibiotic amoxicillin in aqueous samples utilising molecularly imprinted polymers as synthetic receptors. Chemosensors 2019, 8, 5. [Google Scholar] [CrossRef]
  141. Yin, J.; Meng, Z.; Du, M.; Liu, C.; Song, M.; Wang, H. Pseudo-template molecularly imprinted polymer for selective screening of trace β-lactam antibiotics in river and tap water. J. Chromatogr. A 2010, 1217, 5420–5426. [Google Scholar] [CrossRef]
  142. Tan, B.; Zhao, H.; Du, L.; Gan, X.; Quan, X. A versatile fluorescent biosensor based on target-responsive graphene oxide hydrogel for antibiotic detection. Biosens. Bioelectron. 2016, 83, 267–273. [Google Scholar] [CrossRef]
  143. Yue, X.; Li, Z.; Zhao, S. A new electrochemical sensor for simultaneous detection of sulfamethoxazole and trimethoprim antibiotics based on graphene and ZnO nanorods modified glassy carbon electrode. Microchem. J. 2020, 159, 105440. [Google Scholar] [CrossRef]
  144. Wong, A.; Santos, A.M.; Cincotto, F.H.; Moraes, F.C.; Fatibello-Filho, O.; Sotomayor, M.D.P.T. A new electrochemical platform based on low cost nanomaterials for sensitive detection of the amoxicillin antibiotic in different matrices. Talanta 2020, 206, 120252–120259. [Google Scholar] [CrossRef]
  145. Anh, N.T.; Dinh, N.X.; Tuan, H.V.; Doan, M.Q.; Anh, N.H.; Khi, N.T.; Trang, V.T.; Tri, D.Q.; Le, A.T. Eco-friendly copper nanomaterials-based dual-mode optical nanosensors for ultrasensitive trace determination of amoxicillin antibiotics residue in tap water samples. Mater. Res. Bull. 2022, 147, 111649. [Google Scholar] [CrossRef]
  146. Zhang, L.; Chen, L. Fluorescence probe based on hybrid mesoporous silica/quantum dot/molecularly imprinted polymer for detection of tetracycline. ACS Appl. Mater. Interfaces 2016, 8, 16248–16256. [Google Scholar] [CrossRef]
  147. Taranova, N.A.; Berlina, A.N.; Zherdev, A.V.; Dzantiev, B.B. ‘Traffic light’ immunochromatographic test based on multicolor quantum dots for the simultaneous detection of several antibiotics in milk. Biosens. Bioelectron. 2015, 63, 255–261. [Google Scholar] [CrossRef] [PubMed]
  148. Lu, S.; Wang, S.; Wu, P.; Wang, D.; Yi, J.; Li, L.; Ding, P.; Pan, H. A composite prepared from covalent organic framework and gold nanoparticles for the electrochemical determination of enrofloxacin. Adv. Powder Technol. 2021, 32, 2106–2115. [Google Scholar] [CrossRef]
  149. Hu, J.; Su, X.; Yuan, L.; Zheng, K.; Zou, X.; Sun, Z.; Xu, X.; Zhang, W. Competitive immunoassay using enzyme-regulated Fe3O4@COF/Fe3+ fluorescence probe for natural chloramphenicol detection. Anal. Chim. Acta 2023, 1277, 341680. [Google Scholar] [CrossRef] [PubMed]
  150. Wang, L.; Wang, D.; Tao, Y.; Wang, J.; Tian, M.; Liu, Q.; Yang, Y.; Zou, Y.; Ke, F.; Guo, X. Phenylboronic acid functionalized high-crystallinity fluorescent covalent organic framework act as a sensing platform with dual performance for rifamycin antibiotics and water and adsorbent for rifamycin antibiotics. Mater. Today Chem. 2023, 32, 10147. [Google Scholar] [CrossRef]
  151. Li, Q.; Zhu, S.; Wu, F.; Chen, F.; Guo, C. Slice-layer COFs-aerogel: A regenerative dispersive solid-phase extraction adsorbent for determination of ultra-trace quinolone antibiotics. Microchim. Acta 2023, 190, 369–381. [Google Scholar] [CrossRef]
  152. Han, X.; Yuan, C.; Hou, B.; Liu, L.; Li, H.; Liu, Y.; Cui, Y. Chiral covalent organic frameworks: Design, synthesis and property. Chem. Soc. Rev. 2020, 49, 6248–6272. [Google Scholar] [CrossRef]
  153. Gendy, E.A.; Oyekunle, D.T.; Ifthikar, J.; Jawad, A.; Chen, Z. A review on the adsorption mechanism of different organic contaminants by covalent organic framework (COF) from the aquatic environment. Environ. Sci. Pollut. R. 2022, 29, 32566–32593. [Google Scholar] [CrossRef]
  154. Jiang, W.; Cui, W.; Liang, R.; Qiu, J. Difunctional covalent organic framework hybrid material for synergistic adsorption and selective removal of fluoroquinolone antibiotics. J. Hazard. Mater. 2021, 413, 125302–125311. [Google Scholar] [CrossRef]
  155. Gendy, E.A.; Ifthikar, J.; Ali, J.; Oyekunle, D.T.; Elkhlifia, Z.; Shahib, I.I.; Khodair, A.I.; Chen, Z. Removal of heavy metals by covalent organic frameworks (COFs): A review on its mechanism and adsorption properties. J. Environ. Chem. Eng. 2021, 9, 105687–105700. [Google Scholar] [CrossRef]
  156. Haug, W.K.; Moscarello, E.M.; Wolfson, E.R.; McGrier, P.L. The luminescent and photophysical properties of covalent organic frameworks. Chem. Soc. Rev. 2020, 49, 839–864. [Google Scholar] [CrossRef]
  157. Zhang, N.; Ishag, A.; Li, Y.; Wang, H.; Guo, H.; Mei, P.; Meng, Q.; Sun, Y. Recent investigations and progress in environmental remediation by using covalent organic framework-based adsorption method: A review. J. Clean. Prod. 2020, 277, 123360. [Google Scholar] [CrossRef]
  158. Xu, G.; Dong, X.; Hou, L.; Wang, X.; Liu, L. Room-temperature synthesis of flower-shaped covalent organic frameworks for solid-phase extraction of quinolone antibiotics. Anal. Chim. Acta 2020, 1126, 82–90. [Google Scholar] [CrossRef] [PubMed]
  159. Zhou, W.; Wu, H.; Yildirim, T. Structural stability and elastic properties of prototypical covalent organic frameworks. Chem. Phys. Lett. 2010, 499, 103–107. [Google Scholar] [CrossRef]
  160. Tchinsa, A.; Hossain, M.F.; Wang, T.; Zhou, Y. Removal of organic pollutants from aqueous solution using metal organic frameworks (MOFs)-based adsorbents: A review. Chemosphere 2021, 284, 131393–131406. [Google Scholar] [CrossRef]
  161. Guan, S.; Wu, H.; Yang, L.; Wang, Z.; Wu, J. Use of a magnetic covalent organic framework material with a large specific surface area as an effective adsorbent for the extraction and determination of six fluoroquinolone antibiotics by HPLC in milk sample. J. Sep. Sci. 2020, 43, 3775–3784. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Number of papers about covalent–organic frameworks and antibiotics published since 2019. These papers were obtained from the Science Citation Index Expanded (SCIE) database of the Institute for Scientific Information (ISI), searched using “Covalent-organic frameworks, and antibiotics” as titles.
Figure 1. Number of papers about covalent–organic frameworks and antibiotics published since 2019. These papers were obtained from the Science Citation Index Expanded (SCIE) database of the Institute for Scientific Information (ISI), searched using “Covalent-organic frameworks, and antibiotics” as titles.
Polymers 16 02319 g001
Figure 2. (a) Topology diagrams of 2D COFs; (b) assembly of 2D hexagonal COFs; (c) COFs’ four modes of entanglement in frameworks. Reproduced with permissions from Ref. [37], Elsevier, copyright 2019, and Ref. [39], The American Association for the Advancement of Science, copyright 2017.
Figure 2. (a) Topology diagrams of 2D COFs; (b) assembly of 2D hexagonal COFs; (c) COFs’ four modes of entanglement in frameworks. Reproduced with permissions from Ref. [37], Elsevier, copyright 2019, and Ref. [39], The American Association for the Advancement of Science, copyright 2017.
Polymers 16 02319 g002
Figure 3. Diversity of linkages for the formation of COFs. Reproduced with permissions from Ref. [43]; Springer Nature, copyright 2016.
Figure 3. Diversity of linkages for the formation of COFs. Reproduced with permissions from Ref. [43]; Springer Nature, copyright 2016.
Polymers 16 02319 g003
Figure 4. The proposed process for synthesis of COFs. (a) Proposed process for the in situ transformation of TP–COF–BZ into TP–COF–DAB in the presence of 1,4-diaminobenzene. (b) Conversion of amorphous COP-1 to COFs using the linker-exchange strategy. (c) Assembly of 2D hexagonal COFs (COF-5 as an example) from linear and trigonal building units. (d) Assembly of 3D COFs (COF-300 as an example) from linear and tetrahedrally-structured building units. Reproduced with permissions from Ref. [75], ROYAL SOCIETY OF CHEMISTRY, copyright 1972.
Figure 4. The proposed process for synthesis of COFs. (a) Proposed process for the in situ transformation of TP–COF–BZ into TP–COF–DAB in the presence of 1,4-diaminobenzene. (b) Conversion of amorphous COP-1 to COFs using the linker-exchange strategy. (c) Assembly of 2D hexagonal COFs (COF-5 as an example) from linear and trigonal building units. (d) Assembly of 3D COFs (COF-300 as an example) from linear and tetrahedrally-structured building units. Reproduced with permissions from Ref. [75], ROYAL SOCIETY OF CHEMISTRY, copyright 1972.
Polymers 16 02319 g004
Figure 5. Synthesis of COFs using ionothermal method. Reproduced with permissions from Ref. [80]; Wiley Online Library.
Figure 5. Synthesis of COFs using ionothermal method. Reproduced with permissions from Ref. [80]; Wiley Online Library.
Polymers 16 02319 g005
Figure 6. Images of COF-5 using microwave-assisted method (gray COF-5 with purple impurities (a), removal of impurities via MW extraction process (b), and purified gray COF-5(c)). Reproduced with permission from Ref. [85], Royal Society of Chemistry, copyright 2014.
Figure 6. Images of COF-5 using microwave-assisted method (gray COF-5 with purple impurities (a), removal of impurities via MW extraction process (b), and purified gray COF-5(c)). Reproduced with permission from Ref. [85], Royal Society of Chemistry, copyright 2014.
Polymers 16 02319 g006
Figure 7. Preparation of COFs via solvothermal synthesis approach. Reproduced with permissions from Ref. [92]; Elsevier, copyright 2022.
Figure 7. Preparation of COFs via solvothermal synthesis approach. Reproduced with permissions from Ref. [92]; Elsevier, copyright 2022.
Polymers 16 02319 g007
Figure 8. Scheme of electrochemical detection of enrofloxacin and ampicillin using COF-based aptasensors. Reproduced with permissions from Ref. [112]; Elsevier, copyright 2019.
Figure 8. Scheme of electrochemical detection of enrofloxacin and ampicillin using COF-based aptasensors. Reproduced with permissions from Ref. [112]; Elsevier, copyright 2019.
Polymers 16 02319 g008
Figure 9. The fabrication process and application of magnetic COFs in MPE. Reproduced with permissions from Ref. [120]; Elsevier, copyright 2018.
Figure 9. The fabrication process and application of magnetic COFs in MPE. Reproduced with permissions from Ref. [120]; Elsevier, copyright 2018.
Polymers 16 02319 g009
Figure 10. Schematic diagrams of preparation of COFs (a) and application in SPE procedure (b) for the extraction of tetracyclines. Reproduced with permissions from Ref. [125]; Elsevier, copyright 2020.
Figure 10. Schematic diagrams of preparation of COFs (a) and application in SPE procedure (b) for the extraction of tetracyclines. Reproduced with permissions from Ref. [125]; Elsevier, copyright 2020.
Polymers 16 02319 g010
Figure 11. Schematic diagram of (A) electrochemical anodization on the surface of titanium wire, (B) preparation of TpBD-TiO2 SPME fiber, and (C) application of TpBDTiO2 fiber for HS-SPME and detection of phthalate esters. Reproduced with permissions from Ref. [132]; Elsevier, copyright 2020.
Figure 11. Schematic diagram of (A) electrochemical anodization on the surface of titanium wire, (B) preparation of TpBD-TiO2 SPME fiber, and (C) application of TpBDTiO2 fiber for HS-SPME and detection of phthalate esters. Reproduced with permissions from Ref. [132]; Elsevier, copyright 2020.
Polymers 16 02319 g011
Table 1. The textural properties of some COFs.
Table 1. The textural properties of some COFs.
Linkages
Species
COFsSBET
(m2 g–1)
Pore Volume
(cm3 g–1)
Pore Size
(nm)
Ref.
Boronate esterCOF-5--2.7[52]
ImineTpBD885-2.3[58]
ImineCCOF-56550.510.62[59]
AmideCOF-66130.420.59[59]
ImineTx-COF-21137-1.51[64]
TriazineCTF-1791 1.2[53]
TriazineCTF-Th78--[54]
TriazinePhen-CTF358 [55]
HydrazoneTFPT-COF1.185 3.8[66]
HydrazoneBtaMth COF7230.461.41[67]
AmideCOF-1714-3.7[73]
AzineACOF-11.176-1.1[68]
AzinePy-Azine COF1.210-1.9[69]
β-ketoenaminesAcridine COF654--[70]
Table 2. Comparison of the present method for antibiotics detection from environmental water with other reported methods.
Table 2. Comparison of the present method for antibiotics detection from environmental water with other reported methods.
MaterialsDetection MethodsAntibiotics TargetsLimits of DetectionRecovery (%)Relative Standard Deviation (%)Ref.
Zn(II)-MOFPhotoluminescence sensingAminoglycosides37.6 ng L−197 [136]
Zr-MOFSolid-phase extractionSulfonamides antibiotics0.20 μg·mL−191–109.41.0–8.6[137]
MIPSensorErythromycin0.10 μg·mL−191–102 [138]
Graphene Quantum DotsFluorescent emissionTetracycline1.0 μg·L−185.3–103.3 [139]
MIPSensorAmoxicillin1.89 ng mL−1 [140]
MIPSolid-phase extractionβ-lactam9.56 μg·L−160–90 [141]
Graphene oxide hydrogelFluorescent biosensorOxytetracycline25 μg·L−1 [142]
ZnO nanorodsSensorTrimethoprim0.3 μg·mL−193.2–108 [143]
NanomaterialSensorAmoxicillin50 ng L−1973.3[144]
Copper nanomaterialsSensorAmoxicillin1.71 μg·mL−1955[145]
Quantum dotFluorescence sensorTetracycline50 ng L−190.2–97.22.2–5.7[146]
Quantum dotsFluorescence sensorOfloxacin0.3 ng mL−192–1018[147]
COFs Enrofloxacin0.05 ng mL−196.7–102.20.9–6.4[148]
Fe3O4@COFsFluorescence sensorTetracycline0.092 ng mL−196.4–103.7 [149]
Ba@COFFluorescence sensorRifampicin0.03 μg·mL−175.20–123.46 [150]
COFsSolid-phase extractionQuinolone0.02–0.06 ng∙L−168.2–104<10%[151]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Hao, Y.; Xia, Y.; Huang, J.; Zhong, C.; Li, G. Covalent–Organic Frameworks for Selective and Sensitive Detection of Antibiotics from Water. Polymers 2024, 16, 2319. https://doi.org/10.3390/polym16162319

AMA Style

Hao Y, Xia Y, Huang J, Zhong C, Li G. Covalent–Organic Frameworks for Selective and Sensitive Detection of Antibiotics from Water. Polymers. 2024; 16(16):2319. https://doi.org/10.3390/polym16162319

Chicago/Turabian Style

Hao, Ying, Yanjie Xia, Jingjing Huang, Chenglin Zhong, and Guizhen Li. 2024. "Covalent–Organic Frameworks for Selective and Sensitive Detection of Antibiotics from Water" Polymers 16, no. 16: 2319. https://doi.org/10.3390/polym16162319

APA Style

Hao, Y., Xia, Y., Huang, J., Zhong, C., & Li, G. (2024). Covalent–Organic Frameworks for Selective and Sensitive Detection of Antibiotics from Water. Polymers, 16(16), 2319. https://doi.org/10.3390/polym16162319

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