Next Article in Journal
Determination of Brain Tissue Samples Storage Conditions for Reproducible Intraoperative Lipid Profiling
Next Article in Special Issue
Preparation, Characterization, and In Vitro Release of Curcumin-Loaded IRMOF-10 Nanoparticles and Investigation of Their Pro-Apoptotic Effects on Human Hepatoma HepG2 Cells
Previous Article in Journal
Recent Advances in the Antiproliferative and Proapoptotic Activity of Various Plant Extracts and Constituents against Murine Malignant Melanoma
Previous Article in Special Issue
A Facile and Highly Efficient Approach to Obtain a Fluorescent Chromogenic Porous Organic Polymer for Lymphatic Targeting Imaging
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Covalent Organic Frameworks for Chemical and Biological Sensing

1
School of Material Science and Chemical Engineering, Harbin University of Science and Technology, Harbin 150080, China
2
Key Laboratory of Micro-Systems and Micro-Structures Manufacturing (Ministry of Education), Harbin Institute of Technology, Harbin 150080, China
*
Authors to whom correspondence should be addressed.
Submission received: 22 February 2022 / Revised: 29 March 2022 / Accepted: 1 April 2022 / Published: 18 April 2022
(This article belongs to the Special Issue Potential Applications of Functional Porous Organic Frameworks)

Abstract

:
Covalent organic frameworks (COFs) are a class of crystalline porous organic polymers with polygonal porosity and highly ordered structures. The most prominent feature of the COFs is their excellent crystallinity and highly ordered modifiable one-dimensional pores. Since the first report of them in 2005, COFs with various structures were successfully synthesized and their applications in a wide range of fields including gas storage, pollution removal, catalysis, and optoelectronics explored. In the meantime, COFs also exhibited good performance in chemical and biological sensing, because their highly ordered modifiable pores allowed the selective adsorption of the analytes, and the interaction between the analytes and the COFs’ skeletons may lead to a detectable change in the optical or electrical properties of the COFs. In this review, we firstly demonstrate the basic principles of COFs-based chemical and biological sensing, then briefly summarize the applications of COFs in sensing some substances of practical value, including some gases, ions, organic compounds, and biomolecules. Finally, we discuss the trends and the challenges of COFs-based chemical and biological sensing.

Graphical Abstract

1. Introduction

Detecting chemicals or biomolecules is one of the essential goals of analytical chemistry and plays an important role in a wide range of fields including environmental protection, clinic diagnosis, health monitoring, public security, food industry, et al. Thanks to the fast development of instrument science and material science in recent decades, instruments for detecting diverse analytes have been developed. Nowadays, various spectrophotometers, chromatographs, and electrochemical analyzers can be found as conventional analytical equipment in laboratories, factories, and even hospitals. A large number of chemicals and biomolecules can be detected with satisfactory accuracy. However, due to the growing concern for environmental security and healthcare, the demand for fast, real-time, and portable analytical instruments has been increasing. In recent years, plenty of efforts have been devoted to developing portable, wearable, or even implantable sensors for detecting impurities, pollutants, drugs, and physiological indexes [1,2,3,4,5,6,7].
In principle, two key concerns should be carefully considered for detecting a particular analyte. One is to selectively recognize and separate the analyte from the samples. Although some analytes with special physical properties can be separated by simple methods such as filtration, centrifugation, evaporation, et al., most analytes’ separation so far relies on molecular recognition and selective adsorption based on physical, chemical, or biological interaction. Therefore, it is crucial to design and prepare a sort of material that can selectively interact and capture the analyte from the samples during the development of a chemical or biological sensor. Until now, 2D materials such as graphene [8,9], BN [10], MoS2 [11], WS2 [12], and MXene [13], and porous materials including zeolites [14], metal organic frameworks (MOFs) [15,16,17], and covalent organic frameworks (COFs) [18,19,20] were usually chosen as the candidates for selective analytes’ capture. A number of chemical or biological sensors based on the above materials have been constructed. The other key concern for chemical and biological sensing is converting the uptake of the analyte to a detectable signal. The signal can be optical, electrical, thermal, or acoustic. In practice, electrical and optical signals were usually applied since they can be real-time and quantitatively monitored.
COFs are a class of crystalline porous organic polymers with polygonal porosity and highly ordered structures [21,22,23]. As shown in Figure 1, the porosity of the COFs originated from their unique structural design of monomers. In general, the monomers for synthesizing COFs can be roughly divided into two categories. One is the star-like multi-functional monomer that acts as the “knot” in the framework and the other is the linear double functional monomer that can be used as the “edge”. The knots and edges are usually linked by dynamic covalent bonds such as imine, borate ester, hydrazone, carbon-carbon double bond, et al. (Table 1). The dynamic nature of these linkages allows the formation of highly regular, repeated structures with ordered polygonal pores of which the sizes and shapes are determined by the molecular structures of the “knot” and “edge” molecules [19,21,24,25]. Furthermore, functional substituents can be easily introduced to the pore of the COFs by pre-linking it onto the “knots” or “edges” before the synthesis of the COFs or by immobilizing it onto the wall of the pore after the formation of the COFs. In addition, the pores and functional groups on the COFs’ skeletons allows further modification of the COFs by loading nanoparticles or immobilizing nucleic acid or enzymes into their pores or onto their 2D sheets [26,27,28,29]. Thus, it seems that there are infinite possibilities for creating new functional COFs or COF-based hybrid materials. Since the first report by Yaghi et al., in 2005 [30], numerous COFs with various structures and functions have been presented. The application of COFs has been extended to the fields of energy storage and conversion [31,32], optoelectronics [33], gas adsorption and separation [31,34,35,36], sensing [18,19,20], catalysis [37,38,39,40,41], pollutant removal [42,43], battery [44], drug release [45], and so on.
COFs nowadays have been proved to be an ideal platform for chemical and biological sensing. The ordered pore structure of COFs benefits the selective adsorption of the analytes. More importantly, by the pre-design of the monomers, binding sites that can specifically interact with the analyte can be introduced into the “knots” or “edges” of the COFs to further enhance the selectivity of the analytes’ adsorption. Therefore, COFs have exhibited a great advantage in the selective capture and enrichment of the analyte. Furthermore, the analyte adsorption may also affect the optical, electrical, or catalytic properties of the COFs, resulting in a detectable change of color, fluorescence, conductivity, capacitance, or (photo) electrochemical currents, et al. [18]. It can be found that the two key problems of fabricating a chemical or biological sensor are possibly simultaneously solved by one well-designed COF or COF-based hybrid material. Therefore, plenty of chemical or biological sensors on the basis of COFs or COFs-based hybrids have been reported in recent years [18,19,20,21,22,46].
In this review, we firstly discuss the basic principles to design a COF that can selectively adsorb the analyte, then demonstrate the detectable signals that are probably generated by COFs resulting from the analyte adsorption. Successively, we briefly summarize the application of COFs in detecting some substances of practical value, including some gases, ions, organic compounds, and biomolecules. Finally, we discuss the main challenges and future perspectives of COFs-based chemical or biological sensors.

2. Basic Principles of COF-Based Sensing

In recent years, COF-based chemical or biological sensors on the basis of various mechanisms were presented. Although there are still a few exceptions, some principles, strategies, and experiences for fabricating a COF-based sensor can be summarized from the numerous reports. In this section, we will firstly discuss the basic principles for designing a COF that can enrich the analyte. Then we demonstrate some detectable signals that can be produced by the adsorption of the analyte.

2.1. Designing a COF or COF-Based Hybird Material for Selective Adsorption

The porous structure of the COF results in its extraordinary adsorption capacity. However, compared with other types of porous materials, the predominant characteristic of COFs is the designability of the pores. Inspired by the recognition between the enzyme and its substrate, the key points to design a COF that can enrich the analyte are creating a suitable space allowing the load of the analyte and introducing binding sites to selectively interact with the analyte. Hence, the shape and size of the COF pores should be firstly considered. In general, the shape of the pores is mainly determined by the topology of the “knot” monomer (Figure 1). For example, to obtain a hexagonal pore, three-armed planar rigid “knot” monomers, such as the derivatives of triazine, triphenyl benzene, triphenylene derivatives, et al. were usually applied. In addition, four-armed planar rigid “knot” monomers were often used for fabricating a tetragonal pore, while the size of the COF pores depends on the length of both the “edge” monomer and the arms of the “knot” monomers. As shown in Figure 2, it can be easily enlarged by choosing a longer “edge” monomer or a “knot” molecule with longer arms. Thus, it can be found that both the shape and size of the COFs pores can be pre-designed by choosing the “knot” and “edge” monomers [21,24]. To effectively adsorb the analyte, the pore shape and size should be carefully designed to at least allow the analyte molecule to enter. Furthermore, for discriminating the interferents of a similar molecular shape and size, in general, a binding site is needed that can selectively interact with the analyte. The binding site can be the innate groups on the COF skeleton. For example, as shown in Figure 3a, the imine linkage is pH sensitive. It is a good binding site for capturing H+ in the solution or the acid gases like HCl and Trifluoroacetic acid (TFA) [47,48]. However, in other cases, as shown in Figure 3b, the binding site is pre-linked onto one of the monomers before the synthesis of the COFs. For example, thiol or thioether groups are able to selectively adsorb the heavy metal ions like Hg2+. Thus, thioether was covalently pre-linked onto the “edge” monomer, yielding a thioether-containing COF (COF−LZU8) [42]. It was shown that COF−LZU8 performed well in selectively capturing and detecting Hg2+. In addition, the binding site can be also introduced after the formation of COFs, which is also called “post-functionalization” of COFs. For instance, Au nanoparticles (AuNPs) can form an amalgam layer with Hg2+ under the reduction of citric acid (Figure 3c). Hence, to efficiently capture and detect the trace amount of Hg2+, AuNPs were doped on the nanosheet of a bipyridine-containing COF by the in situ growth method to form a COF-based hybrid material. It was found that the huge specific surface area of the COF-AuNPs hybrid material can significantly enhance the sensitivity of the Hg2+ detection [29]. Moreover, except for in situ growth, the binding site can be also externally introduced onto the 2D sheet of the COFs. For example, as shown in Figure 3d, to selectively detect enrofloxacin, a kind of antibiotic, an enrofloxacin-targeted aptamer was immobilized onto a 2D sheet of COF via π-π stacking and electrostatic interaction to yield an aptamer-functionalized COF hybrid [49]. Then the hybrid was coated onto a gold electrode. It can selectively capture the enrofloxacin in the sample solution and result in a detectable change in the electrode conductivity.

2.2. Signals Produced by the Adsorption of the Analyte

Apart from the selective uptake and enrichment of the analyte, the other key problem in constructing a COF-based sensor is to make the uptake of the analyte produce a detectable signal. By ingenious structural design, a large number of COFs of which the color, fluorescence, conductivity or capacitance can quantitatively change with the uptake of the analyte have been synthesized [18]. Besides that, in some reports the adsorption of the analyte can also produce some detectable signals which do not originate from COFs [29,50]. In this section, we will briefly discuss the generation, application, and advantages of the above signals in COF-based sensors. In addition, other analytical techniques such as QCM, Raman spectra, MS, et al. were also used occasionally. These techniques will be discussed with the particular examples demonstrated in the next section.

2.2.1. Fluorescence

Among the detectable signals caused by the uptake of the analyte, fluorescence change of COFs including fluorescence enhancement (“turn on”) and fluorescence quenching (“turn off”) is the most frequently used to respond to the concentration of analyte. Due to the rigid structure, the large π-conjugated fluorescent chromophores, such as phenyl, naphthalenyl, pyrenyl, perylenyl, triazine, and triphyenyl-benzene are usually applied for building COFs [51,52]. Moreover, the conjugated linkages like imine and olefins can further enlarge the π-conjugated structure in the COFs. Therefore, it is not unusual to find a COF with fluorescent emission. However, it was found that the adsorption of some analytes may influence the fluorescent emission, including the change of the intensity and movement of the emission peak. The fluorescence change may be induced by various mechanisms, such as the charge and energy transfer between the COF and analyte, aggregation-induced emission, and exciplex formation, et al. The most common one among the above mechanisms is the charge transfer from an electron-rich COF to an electron-deficient analyte. It can lead to a significant fluorescent quenching which is relative to the concentration of the analyte. In the following demonstration, it can be found that many COF-based nitroaromatic explosive sensors, iodine sensors, and transition metal ions sensors are constructed by the charge transfer mechanism.

2.2.2. Chromism

Chromism is a signal that can be directly found by naked eyes. The advantage of the sensing approach based on chromism is easy-operate. It is very convenient to be further developed into rapid testing equipment like test paper, test kit, and wearable monitoring devices. Therefore, the COF-based sensors based on chromism are frequently reported. The color change of the COFs can be induced in different manners. Firstly, the non-covalent interaction between COFs and the analyte disturbs the electronic transition responsible for the coloration. For example, the adsorption of molecules with different polarities can influence the intramolecular charge transfer from the electron donor to the electron receptor on the COF skeleton, causing a red or blue shift of the UV-Vis absorbance [53]. Secondly, in some COF-based chromism sensors, the interaction with the analyte can isomerize some groups on the COF skeleton, resulting in the color change [54,55]. In addition, it should be noted that there is also a type of COF-based chromism sensor of which the color change is not caused by the COF itself. For example, in some sensors, COFs are used as a catalyst or a catalyst carrier, and the adsorption of the analyte can influence the catalytic capacity which can be detected by the formation of the colored product. Thus, the concentration of the analyte can be determined by the color change [29,50].

2.2.3. Capacitance and Conductivity

The adsorption of the analyte can lead to a change in the capacitance and conductivity of the COFs [56,57,58]. Hence the concentration of the analyte can be also determined by the change of such two electrical signals. In general, the quantitative measurement of electrical signals is much more convenient compared with photo signals. The measuring devices involved are much simpler. Therefore, COF-containing sensors based on the change of capacitance and conductivity are more likely to be realized on a chip-like device. Many presented works have demonstrated that a fingertip-size interdigital electrode (IDE) coated with a specific COF can be applied for real-time detection of humidity, corrosive gases, or toxic gases since their adsorption results in the change of the capacitance or conductivity of the COFs. It can be expected that such a type of device can be further developed into portable, wearable, or even implantable sensors for real-time sensing.

2.2.4. Electrochemical or Photoelectrochemical Signals

Electrochemical or photoelectrochemical signals are generated by the redox reaction of the active substance. Plenty of analytical techniques based on electrochemistry or photoelectrochemistry have been built. The signal produced by electrochemical or photoelectrochemical reactions can be influenced by many factors such as the concentration of the active substances, the conductivity of the electrode/electrolyte interface, the existence of catalysts, and so on. In recent years, a number of examples of electrochemical or photoelectrochemical analysis with the participation of COFs were reported. The COFs coated on the electrode in these examples can play various roles including: (1) enriching the analyte that can be directly detected by the electrochemical or photoelectrochemical method; (2) concentrating the analytes to inhibit or facilitate the transportation of charges or substances; (3) a carrier for immobilizing the catalyst which adsorption can be affected by the analytes. All of the above allow a significant influence of the analyte on the electrochemical or photoelectrochemical signal. Compared with the above type of COF-based sensors the advantage of the electrochemical or photoelectrochemical signal is the high sensitivity. It is not rare to realize trace amounts of the analyte at or lower than ppb or ng/mL level with this type of COF-based sensor.

3. Application of COFs for Sensing Various Analytes

3.1. Gas Sensing

Generally, the key challenge of gas sensing is to capture the target molecule from the sample. Therefore, porous materials like zeolite, MOFs, and COFs show a great advantage in gas sensing. Driven by the increasing demand for healthcare and environmental protection, some COFs that can be used for detecting irritating, harmful, or toxic gases have been developed (Table 2).

3.1.1. Acidic and Alkaline Gases

Ammonia is widely applied in agriculture, the chemical industry, and refrigerated storage. Ammonia leakage can lead to discomfort and do harm to health. Although the pungent odor of ammonia can put us on high alert immediately, slow leakage of ammonia will decrease our sense of smell to find it. Therefore, it is necessary to detect trace amounts of ammonia. Early in 2010, O. M. Yaghi reported that the boronate linkage for synthesizing COFs can be used for ammonia uptake due to the Lewis acid–base interaction [59]. In 2016, D. Jiang and coworkers reported a highly emissive covalent organic framework (TPE−Ph−COF) based on dynamic boronate linkage [60] (Figure 4a). The boronate linkages can serve as a Lewis acid when it interacts with ammonia, which is a Lewis base. The interaction with ammonia significantly decreased the luminescence of TPE−Ph−COF. Hence, the dispersion of TPE−Ph−COF in organic solvent, such as toluene or cyclohexane, can be used for ammonia sensing. It can be found that the fluorescence of TPE−Ph−COF linearly decreased with the concentration of NH3 and 1 ppm NH3 could lead to a 30% reduction in fluorescence intensity. This result indicated that TPE−Ph−COF is a good candidate for detecting trace amounts of ammonia.
Table 2. COF-based gas sensing.
Table 2. COF-based gas sensing.
AnalyteYearCOF NamesSpecific Binding SiteType of Detectable SignalDetection RangeLODReference
NH32016TPE−Ph−COFBoronateFluorescence (turn off)-sub ppm level[60]
2018HMP−TAPB−1HMP−TAPB−1Heptazineconductivity1–200 ppm1 ppm[61]
2018COP−COP−1TriazineFluorescence (turn on)-5.8925 × 10−4 mL/mL[62]
2019Ph−An−COFBoronateFluorescence (turn off)--[63]
2019COF−DC−8-Conductivity2–80 ppm56.8–70 ppb[64]
2021TAPB−BPDA COFImine Conductivity5–100 ppm10 ppb[65]
TFA2019Per−N COFImineChromism0.035–110 mg L−135 μg L−1[47]
HCl2018COP−COP−1TriazineFluorescence (turn off) 1.0967 × 10−4 mL/mL[62]
2019PBHP−TAPT COFTriazineChromism, conductivity20–3000 ppm20 ppm[57]
2019COF−ETBA−DABImineFluorescence 4.7 ppm[66]
2020BCTB−BCTA COFImineFluorescence (turn off)1–25 mM10 nM[48]
H2O2013TAPP−DHNDA−COFIminolChromism20–100% RH [54]
2017COF−TXDBABoronateConductivity 11–98% RH [56]
2018Py−TT COF Chromism0.64–0.98 p/p0 [53]
2020TAPB−PDA−OH COFIminolChromism [55]
2021DUT−175ImineChromism33–94% RH [67]
Benzene2020BTA−TAPT-COFAromatic groupCapacitance500 ppb–100 ppm340 ppb[58]
NO22019COF−DC−8 Conductivity2–40 ppm1–16 ppb[64]
2020CON−10 Conductivity 2.242 ppb[68]
2020T−2DP Conductivity0.15–5 ppm2.2 ppb[69]
2021NiPc−CoTAA Conductivity 1–40 ppm [70]
NO2019COF−DC−8 Conductivity0.02–40 ppm1–5 ppb[64]
H2S2017PNT−1Triazine, pyridineFluorescence (turn off) 53 ppb[71]
2019COF−DC−8 Conductivity2–80 ppm121 ppb[64]
O32021P−COFTPB−DMTP−COFimineChromism 0.1 ppm[72]
It can be found that the acid–base interaction between NH3 and the skeleton of the COFs played an essential role in NH3 detection since NH3 is a typical alkaline gas [60,61,62,63,64,65]. Correspondingly, it can be expected that acidic gas, such as HCl and TFA, can also be detected in a similar vein [47,48,57,62,66]. As with the boronate bond, the imine bond (Schiff base) is another frequently used dynamic linkage for COF synthesis. However, the imine linkage in COFs is usually a proton receptor that can easily react with acidic gases, allowing imine-based COFs to be sensitive to acids [47,66]. For example, F. Auras and co-workers developed a series of star-shaped COFs (Per-1P, Per−N, and Per-Py COFs) via the condensation of perylene tetraaniline with the phenylene-, naphthalene-, or pyrene-based dicarbaldehydes (Figure 3b) [47]. They found that the imine linkages of the star-shaped COFs can be reversibly protonated, resulting in a marked color change of the COFs. For example, the color of the Per−N-COFs thin film quickly changed from yellow to dark brown upon exposure to the vapor of trifluoroacetic acid (TFA). Its adsorption band around 385 nm decreased with the concentration of TFA vapor, simultaneously a new increasing absorption band around 450 nm could be observed. As a colorimetric acid vapor sensor, the detection limit of Per−N-COFs to trifluoroacetic acid can be as low as 35 μg L−1 and the response range was at least four orders of magnitude.
Apart from the imine bond, triazine is another acid-sensitive group that has been frequently used in the fabrication of covalent organic porous materials. The long pair of three N atoms in triazine is a good H+ acceptor. In 2018, N. Xu and coworkers prepared a high-performance fluorescent triazine-based covalent organic polymer (COP−COP−1) sensor for detecting both HCl and NH3 gases [62]. It was found that when the COP−COP−1 was exposed to gaseous HCl, a redshift of the fluorescence happened and the intensity of the fluorescence decreased linearly with increasing HCl concentration. Conversely, NH3 can lead to recovery of both color and fluorescence of the COP−COP−1 saturated by HCl, indicating that COP−COP−1 saturated by HCl can be applied as NH3 detection. Later, R. Kulkarni and coworkers reported a triazine-based COF (PBHP−TAPT COF) with similar property [57] (Figure 3c). Gaseous HCl can protonate the triazine groups in PBHP−TAPT COF and quickly turn its color from yellow to red. A very low concentration of HCl (20–50 ppm) can result in a detectable color change. The reverse color switch can be realized by the deprotonation by NH3. Furthermore, R. Kulkarni and coworkers found that the protonation of the triazine groups enhanced the conductivity of PBHP−TAPT COF by 170-fold, indicating that apart from the colorimetric method the acidic gas detection by PBHP−TAPT COF can be conducted by also monitoring electrical conductivity.
Notably, the COF-based detection of NH3 without acid–base interaction was also reported [65]. In 2021, F. Niu and coworkers synthesized a COF (TAPB−BPDA) via the condensation of 4,4′-biphenyldicarboxaldehyde with 1,3,5-tris(4-aminophenyl)benzene catalyzed by a proper amount of Sc(OTf)3 and coated it onto an IDE. It can be measured that the resistance of TAPB−BPDA decreased when it was exposed to NH3. The limit of detection (LOD) of this resistance-based approach can be as low as 10 ppb. Meanwhile, owing to the hydrogen bond between the COF and analyte, the NH3 sensing of TAPB−BPDA exhibited a good selectivity. Other gases, such as NO2, NO, SO2, H2S, H2, CO, CO2, et al., cannot result in the resistance decrease. The work by F. Niu and coworkers provide an example of NH3 sensing with noncovalent interaction.

3.1.2. Water Vapor (Humidity) Sensing

Humidity is one of the essential indexes of our working and living environment. The equipment for real-time humidity monitoring can be widely found in laboratories and factories. In recent years, developing wearable respiratory monitoring devices has drawn much attention in the fields of healthcare and sports, because they can be used to record the breathing rate and depth of patients and athletes. For fabricating a wearable respiratory monitor, a sensitive humidity responsive material is vital. Due to the designable pore structure, COFs are good candidates for selectively capturing and detecting water vapor in the air [53,54,55,56,67].
The mechanism firstly used for humidity detection is the keto-iminol tautomerism. Early in 2013, X. H. Liu and coworkers solvothermally prepared crystalline COF (TAPP−DHNDA−COF) nanofibers by the imine condensation between 2,4,6-tris(4-aminophenyl) pyridine (TAPP) and 2,6-Dihydroxynaphthalene-1,5-dicarbaldehyde (DHNDA) [54]. SEM images showed that the nanofibers can form when the time of solvothermal synthesis was longer than 24 h. By using aramid fabric as a nucleating agent, X. H. Liu and coworkers successfully prepared a TAPP−DHNDA−COF/aramid hybrid material, where TAPP−DHNDA−COF nanofiber was coated on the surface of the aramid fibers. The hybrid material exhibited a reversible colorimetric humidity-sensitive property because the addition of H2O can disturb the equilibrium of keto-iminol tautomerism (Figure 5a). The color of the hybrid can switch from yellow to red gradually with the increasing relative humidity (R.H.) from 20% to 100%. Similarly, L. Ascherl et al. reported orienting a thin film formed by a tetrakis(4-aminophenyl)pyrene-based COF (Py–TT COF) (Figure 5b) [53]. The thin film exhibited a high humidity sensitivity. UV–Vis spectra revealed that a new absorption band between the 520–640 nm appeared after the exposure to enough H2O(g) and the original absorption across the 440–500 nm and 280–380 nm spectral regions simultaneously decreased. The highest sensitivity towards humidity appeared between H2O relative pressures of 0.64 and 0.79. Further investigation suggested that the mechanism of the color change is solvatochromism. Notably, the color change from a dry to humid atmosphere cost only 0.21s, and the converse procedure was even faster, when only 0.15 s was needed. It indicated that Py–TT COF thin film was potential material for real-time chromic humidity monitoring. Similarly, S. Jhulki and coworkers also reported a COF (TAPB-−PDA−OH COF) containing 2,5-di(imine)-substituted 1,4-dihydroxybenzene (diiminol) moieties [55]. H2O molecules can be adsorbed in the 1D channel of TAPB−PDA−OH COF via hydrogen bond. The adsorption of H2O molecules resulted in the color of TAPB−PDA−OH COF changing from orange to dark red (Figure 5c). The response to the humid air cost only 9 s and the recovery procedure needed less than 1 s, indicating that TAPB−PDA−-OH COF is also a good candidate for real-time humidity sensing.
In addition to color change, the adsorption of water vapor can also induce a decrease in impedance. H. Singh and coworkers synthesized truxene-based COFs with boron ester linkages (COF−TXDBA) [56]. The empty orbitals of the boron atom in the COF skeleton allowed the adsorption of H2O via the Lewis acid–base interaction with the electron lone pair of the oxygen atom in the H2O molecule. When R. H. was low, the amount of H2O molecules was too low to form a continuous layer, hindering the charge transportation process among the water molecules by the Grotthuss chain reaction (H2O + H3O+ → H3O+ + H2O). The conductivity of the COF−TXDBA was very low. However, if the amount of H2O molecules allowed the formation of the continuous water layer and occurrence of the Grotthuss chain reaction, the impedance would sharply decrease. A linear relationship between R. H. and the logarithm of the impedance can be found in the range of R. H. 11 to 98%. Compared with the above-mentioned chromic method, the humidity sensing based on electrical impedance reported in this study was more quantitative and accurate. However, this approach also displayed a shortage. Its response time was measured as about 40 s. which is much longer than that of the chromic humidity sensor mentioned above.

3.1.3. Harmful Gases Sensing

With the development of the industry, the number of pollution incidences resulting from harmful gases has been increasing in recent decades. For example, the huge coal or oil consumption releases massive amounts of sulfur oxides and nitrogen oxides into the atmosphere, resulting in the high frequency of acid rain. In the indoor environment, the harmful gases released by some shoddy decoration materials such as benzene and formaldehyde are endangering human health seriously. The huge demand for alerting about harmful gases is promoting the development of the COFs for sensing levels of harmful gases.
As shown in Figure 6a, D. Zhao developed a COF-based chip-like device for benzene vapor sensing [58]. The device was prepared by in situ growing the film of a COF (BTA−TAPT−COF) which was synthesized by the imine condensation between benzene-1,3,5-tricarboxaldehyde (BTA) and 2,4,6-tris(4-aminophenyl)-1,3,5-triazine on an IDE. The exposure to benzene vapor significantly enhanced the capacitance of the device, which can be detected by an LCR bridge. The limit of benzene detection can reach 340 ppb. Notably, due to the interaction between electron-rich benzene and the electron-deficient triazine group, the COF-based benzene sensors showed good selectivity toward benzene over CO2, CH4, and C3H8 at room temperature. It suggested that the performance of the capacitive benzene sensor cannot be interfered by common indoor gases.
Among various nitrogen oxides, the detection of NO2 drew the most attention, because it is the most likely to appear in the atmosphere. For NO2 sensing, the electrical method based on IDE was also applied [64,68,69]. K. A. Mirica and coworkers prepared a COF-coated IDE of which the resistance is sensitive to several harmful gases, including NH3, H2S, NO, and NO2 [64]. The quantitative detection of these gases was realized. The LOD to these gases can reach the level of ppb. However, selective sensing of NO2 had not yet been realized. Subsequently, D. Xia and coworkers fabricated a selective NO2 sensor by drop-casting a triazine-based 2D organic polymer (T−2DP) on a gold IDE [69]. The T−2DP was prepared by the exfoliation of a covalent triazine framework synthesized through a triflic acid catalytic cyclotrimerization of 1,4-dicyanobenzene. The T−2DP sensor displayed a good performance in sensing trace amounts of NO2. The linear range was < 1 ppm and the theoretical LOD is estimated to be 2.2 ppb. The sensor also showed a fast response to the analyte (response time: 35–47 s, recovery time: 56–140 s) and excellent selectivity. This work provided an example of selective high-performance NO2 sensing.
In 2017, L. Guo and coworkers reported luminescent organic polymer nanotubes (PNT−1) for S2− sensing [71]. The nanotubes were formed by a 2D structure synthesized by the condensation of 2,4,6-tris-(4-bromo-phenyl)-1,3,5-triazine with 2,6-dibromopyridine. As shown in Figure 6b, the fluorescence of the PNT−1 suspended in CH3OH can be quenched by adding Na2S. The limit of the “turn off” fluorescent S2− detection of PNT−1 was measured as 53 ppb. Other anions including halide ions, SO42−, NO3, HSO3, PO43−, CO32−, et al. cannot result in the same phenomenon, indicating the good selectivity of the S2− sensing. However, more importantly, Guo and coworkers subsequently further prepared a test paper for gaseous H2S detection by dipping a piece of filter paper into the nanotube suspension for a few minutes (Figure 6b). It was found that the fluorescence of the newly prepared moist test paper can be immediately quenched upon exposure to the H2S atmosphere. It indicated that fast detection of gaseous H2S can be realized by PNT−1.
Very recently, a COF-based fast detection of ozone, which is an air pollutant brought by excessive emission of nitrogen oxides, was reported by Z. J. Zhang et al. [72]. The COF used for O3 sensing is the frequently-reported TPB−DMTP−COF which was synthesized by 2,5-dimethoxyterephthalaldehyde and 1,3,5-tris-(4-aminophenyl) benzene. In the absence of moisture, ozone can nucleophilically attack imine groups to form nitro, aldehyde, and carboxylate moieties, while in the presence of moisture (R. H. = 40%) imine groups prefer to bind with water molecules to facilitate its protonation, leading to the visible color change from light yellow to red. Furthermore, the author found that TPB−DMTP−COF exhibited a better capacity of ozone removal than other porous materials such as activated carbon, amorphous polymer, monomer, and MOFs.

3.2. Inorganic Ions Sensing

3.2.1. Metal Ions Sensing

The concentrations of some metal ions, such as Fe2+, Zn2+, Mg2+, et al., are important indicators of the health of human beings. For example, the analysis of Fe2+/Fe3+ has great significance in the diagnosis of anemia. Beyond that, heavy metal ions, such as Cu2+, Hg2+, Pb2+, et al., are important industrial pollutants. Excessive heavy metal ions in water sources, food, or drink can cause serious poisoning and irreversible damage to human bodies. Therefore, the simple, rapid, accurate, and sensitive detection of all kinds of metal ions is widely expected in the fields of healthcare, environmental protection, and the food industry. Owing to the unique features of structural designability and selective absorbability, a variety of COFs have been designed for sensing a wide range of metal ions (Table 3).
Hg2+ is highly toxic and its existence in natural water or food must be particularly monitored. In 1956, the bioaccumulation of Hg2+ from industrial discharge caused the emergence of Minamata disease and finally resulted in a well-known social pollution nuisance. Therefore, many efforts have been devoted to the sensitive detection of Hg2+ in water. In general, for designing a COF that can effectively capture Hg2+, special groups that can particularly bind with Hg2+, such as thiol or thioether, were usually introduced [29,42,73,74]. The high affinity between S and Hg can significantly enhance the capacity of mercury uptake. In 2016, W. Wang et al. reported a fluorescent COF (COF−LZU8) with a hexagonal pore containing thioether group as a selective Hg2+ receptor [42]. The electron transformation from COF−LZU8 to Hg2+ can result in a markable fluorescence quenching, and the intensity of fluorescence linearly decreased with the concentration of Hg2+ (Figure 7a). The LOD was measured as 25.0 ppb. More importantly, the fluorescence quenching exhibited a significant selectivity, except for Hg2+, fluorescence quenching caused by a wide range of metal ions is insignificant. Meanwhile, due to the excellent Hg2+ uptake capacity, COF−LZU8 can also be easily used for mercury (II) removal. Moreover, upon the simple treatment with concentrated Na2S solution, COF−LZU8 can be easily recovered for the next Hg2+ detection and removal.
In addition to the S(II) containing groups, it was reported that the carbohydrazide group can be also applied for capturing Hg2+ in the aqueous solution. J.-D. Qiu et al. designed a pyrene-based luminescent COF (TFPPy−CHYD) with a flexible carbohydrazide (CHYD) linkage [73] (Figure 6b). The CHYD group can effectively bond Hg2+ in the tetragonal pore of TFPPy−CHYD by coordination interaction. The maximum adsorption capacity for Hg2+ can be as high as 758 mg g−1. The binding between Hg2+ and CHYD also exhibited good reversibility. The Hg2+ captured by TFPPy−CHYD can be effectively removed by 10 equiv. of Na2S solution. In the meantime, the binding of Hg2+ led to a quick, efficient quenching of the fluorescence which cannot be found in presence of other metal ions. The intensity of the fluorescence decreased linearly with [Hg2+] when 0.05 μM < [Hg2+] < 4 μM. The LOD was measured as 17 nM.
Due to the high toxicity, the threshold level of mercury in drinking water is as low as 10 nM. Therefore, it is still necessary to develop Hg2+ sensors with lower LOD. As well as capturing Hg2+ into the pore of COFs, other strategies to detect trace amounts of mercury have been also attempted. For example, J.-D. Qiu and coworkers reported a COF-based nanocomposite for Hg2+ sensing for which the LOD can be lower than 1 nM [29]. As shown in Figure 6c, they synthesized bipyridine-containing COF nanosheets, Tp−Bpy NSs, with a regular hexagonal pore structure. Due to the nitrogen-containing bipyridine groups, a large number of AuNPs can be doped onto the Tp−BpyTp−Bpy NSs by in situ growth method to form a AuNPs@Tp−Bpy nanocomposite. In citric acid-phosphate buffer solution, AuNPs@Tp−Bpy can adsorb trace amounts of Hg2+ from the sample solution to form a gold amalgam layer on the surface of the AuNPs. The amalgam layer allows H2O2 to be decomposed efficiently to form highly reactive hydroxyl radicals which can rapidly oxidize colorless 3,3′,5,5′-tetramethylbenzidine (TMB) to its bright blue oxidation state (oxTMB). Hence, the absorbance at 652 nm caused by oxTMB can be used to quantitatively determine the concentration of Hg2+ in the sample solution. Notably, this approach for Hg2+ sensing exhibited an ultrahigh sensitivity. The color changes caused by even 1 nM Hg2+ were easily detected.
Apart from Hg2+, the detection of Cu2+, Fe3+, and Pb2+ has also drawn much attention recently. Although the toxicity of Cu2+ is far less than Hg2+, the widespread existence of Cu2+ may still risk the health of citizens. Therefore, COFs that can sensitively and selectively detect Cu2+ have been developed. The strategy to develop COFs for Cu2+ detection had much in common with COF-based Hg2+ sensing. In recent years, several luminescent COFs, of which fluorescence can be quantitatively quenched by the addition of Cu2+, were presented. In 2016, X. Liu and co-workers synthesized a hydrogen bond-assisted azine-linked COF (COF−JLU3) with robust chemical stability and thermal stability [75]. It displayed a high luminescence efficiency due to the bulky tert-butyl groups on the skeleton, which can adjust the π–π interaction between the layers. The numerous N atoms and hydroxyl groups in the pore of COF−JLU3 acted as receptors to benefit the uptake of metal cations. It was found that transition-metal ions with different electron configurations such as Fe3+, Co2+, especially Cu2+, can be captured resulting in significant quenching the fluorescence of COF−JLU3. The decrease of fluorescence emission intensity was almost proportional to [Cu2+] in the range from 0 to 0.4 mM, suggesting that COF−JLU3 is a suitable candidate for Cu2+ detection. The LOD of Cu2+ sensing was measured as 0.31 mM. In 2018, D. Jiang and coworkers designed and successfully synthesized a series of stable light-emitting two-dimensional sp2 carbon-conjugated COFs (sp2c-COFs) [77]. They exhibited excellent luminescence activity and high stability simultaneously. Interestingly, the -CN side groups in sp2c-COFs were able to interact with Cu2+ resulting in the quenching of the luminescence. Only 88 ppb (about 1.4 μM) Cu2+ can result in an observable quenching. It indicated that sp2c-COFs can serve as a highly selective and effective sensor for detecting Cu2+ ions. Later, H. Wang et al. reported a fluorescent QG−scaffolded COF synthesized by the one-step covalent reactions of melamine-aldehyde and phenol-aldehyde poly-condensations using paraformaldehyde, initially with Q-Graphene (QG) scaffolds [78]. The N atoms and hydroxyl groups on the COF skeleton can effectively chelate Cu2+. Different from other metal ions, Cu2+ can significantly quench the fluorescence of QG−scaffolded COF, indicating that the QG−scaffolded COF is a good candidate for Cu2+ sensing. The linear range of the Cu2+ sensing was measured as 0.0010~10.0 μM and the LOD can reach 0.50 nM. Moreover, as with the Hg2+ detection, this strategy relies on the enzyme-mimicking property of nanomaterials can be also applied in Cu2+ detection. In 2017, Y. Xiong et al. developed a covalent triazine framework (CTF) that can exhibit peroxidase-like catalytic activity once it is coordinated with Cu2+ [50]. The copper ions acted as the active sites in the complex formed by CTF and Cu2+ (CTF/Cu2+). It can be found that the catalytic activity of CTF/Cu2+ can be tuned by the concentration of Cu2+. Upon the addition of H2SO4, H2O2, and TMB, the concentration of Cu2+ can be easily determined by UV-Vis absorbance. This colorimetric method showed good selectivity. Its linear range was measured as 1.0 μg/L to 80.0 μg/L. and the LOD was determined as 0.05 μg/L.
Iron is one of the essential trace elements of the human body. It plays an important role in oxygen uptake and transfers into red blood cells. Iron is also an important part of some enzymes, for example, catalase. Hence the detection of Fe3+ has important significance in the food industry and health examination. In recent years, a number of COFs that can be used for Fe3+ detection were reported [79,80,81,82,83,84,85]. Most of them were based on Ion-induced fluorescence quenching. In 2017, W. Yang and coworkers reported two polyimide-based porous COFs (PI−COF 201 and PI−COF 202) [79]. Owing to the p*-n transition caused by the high electro-delocalization and the inherently rigid structure, the two COFs can emit strong fluorescence. However, the adsorption of Fe3+ remarkably quenches their fluorescence due to the energy transfer from the emission level of the COFs to the unoccupied d-orbital of Fe3+. A linear relationship between the fluorescence intensity and [Fe3+] was found in the range of 5.0–40 μM. The LODs were determined as 0.13 μM (PI−COF 201) and 0.22 μM (PI−COF 202), respectively. Later, several luminescent COFs of which fluorescence can be quenched by Fe3+ were reported [80,81,82,83,84,85]. The LOD of the COF-based Fe3+ sensing has been decreased to 64 nM. Notably, some COFs for Fe3+ sensing can be also used to detect other ions, such as Pb2+ [85], CrO42−, Cr2O72−, and MnO4 [80]. Meanwhile, besides fluorescence quenching, COF-based Fe3+ sensing in another manner has been also reported. In 2021, W.-G. Zhang and coworkers prepared a robust hydrazone-linked COF (Tfpa–Mth COF), of which its fluorescence can be quenched by Fe3+ [84]. In the meantime, W.-G. Zhang and coworkers prepared a uniform Tfpa–Mth COF thin film on an amino-modified quartz crystal microbalance (QCM) chip by in situ growth method. The Tfpa–Mth COF thin film can capture Fe3+, resulting in a frequency shift in QCM measurement. Compared with the fluorescence quenching, the Tfpa–Mth COF-coated QCM chip displayed a significant advantage in real-time Fe3+ monitoring.
Pb2+ is another metal ion with which COF-based sensing has been widely concerned. Similar to mercury, lead can be also accumulated in plants and animals and further threaten human beings through the food chain, resulting in irreversible damage to the liver, nervous or circulatory system. Even nowadays, lead is still widely used in batteries, gasoline additives, and paint pigments. Therefore, fast and sensitive lead sensing is still needed. Unlike Hg2+ sensing, the reported COF-based Pb2+ sensors were more likely to depend on electrochemical signals [85,86,87,88]. In 2018, Y. Wang and J. Wang [86] and co-workers mixed TAPB−DMTTAPB−DMTP−COF [85] with graphite powder to prepare a TAPB−DMTTAPB−DMTP−COF-based carbon paste electrode (TAPB−DMTTAPB−DMTP−COF/CPE) for sensitive and selective Pb2+ sensing. The huge surface area and a large number of amino groups of TAPB−DMTTAPB−DMTP−COF greatly benefited the capture of Pb2+ from the sample solution. By using TAPB−DMTTAPB−DMTP−COF/CPE, the concentration of Pb2+ can be determined by differential pulse stripping voltammetry. The change of current intensity showed excellent linearity to the concentration of lead in the range of 0.0050 to 2.0 μM and the LOD of this method was measured as 1.9 nM. Similarly, X. Wang and coworkers prepared an Au electrode coated with sulfhydryl-modified TAPB−DMTTAPB−DMTP−COF for Pb2+ detection [87]. The [Pb2+] was determined by square wave anodic stripping voltammetry with the electrode. Benefiting from the affinity between S(II) and Pb, the LOD of the Pb2+ detection can be as low as 0.015 ng/mL (0.072 nM). Recently, Y. Zhang et al. further enhanced the sensitivity of COF-based Pb2+ detection by introducing the photoelectrochemical (PEC) sensing technique [88]. They firstly prepared a photocathode by fabricating a porphyrin-based COF (TAPP−COF) on a PET-ITO electrode. The photocathode presented an outstanding photovoltaic property owing to the unique charge channels of COFs and the good photoelectric properties of porphyrin. Then the CdSe quantum dots coated with a SiO2 shell (CdSe@SiO2QDs), which herein acted as a quenching agent, were immobilized onto the TAPP−COF thin film through a hybridization chain reaction, significantly decreasing the photocurrent. However, the presence of Pb2+ can detach the CdSe@SiO2QDs from the TAPP−COF thin film, leading to the quantitative recovery of the photocurrent. Hence, accurate sensing of Pb2+ can be achieved in a wide detection range of [Pb2+] (0.05~1000 nM) with an extremely low LOD of 0.012 nM.
In recent years, the interest in developing COF-based metal ion sensing has been gradually extended. In addition to the above-mentioned metal ions, the COFs for sensing other ions such as Au+, UO22+, Ni2+, Cr3+, Pd2+, et al. have been developed one after another. In 2018, Y. Yu designed and synthesized a thioether-functionalized fluorescent COF (TTB−COF) for selective sensing Au ions [89]. Observable fluorescence quenching of TTB−COF can be induced by adding Au ions. The fluorescence intensity and the concentrations of Au ions showed a linear relationship in the range of 1.0–10.0 mM and the LOD is calculated to be 0.87mM. In addition, TTB−COF exhibited a large capture capacity to Au ions (560 mg g−1); thus, it is also a promising material for the recovery of ultra-low concentration Au ions in the aqueous solution. Uranium is one of the essential raw materials in the nuclear industry. However, its content in the earth’s crust is quite low, therefore, porous materials for uranium extraction from seawater and detection of uranium have been recently focused on [90,96,97,98,99]. J.-D. Qiu and coworkers prepared a fluorescent COF (TFPT−BTAN−AO) containing a large amount of selective uranium-binding groups as shown in Figure 8a [90,91]. The TFPT−BTAN−AO can selectively capture UO22+ ions. The uptake capacity was measured as 427 mg g−1, suggesting that it is possible to be applied for the purposes of uranium extraction. The fluorescence emission of TFPT−BTAN−AO linearly decreased upon the gradual addition of UO22+. The LOD was determined as 6.7 nM. Similarly, Q. L. Deng and coworkers reported that a COF containing the bipyridine group in its skeleton of fluorescence can be selectively quenched by Ni2+ when pH = 10 [92]. As a platform for Ni2+ detection, an extremely low LOD of (68 pM) can be achieved. Meanwhile, the COFs have been also used for chromium sensing. As mentioned above, the COF for Fe3+ sensing reported by Z. Shi et al. [80] can also be applied for the detection of Cr(VI), which is a highly toxic carcinogen. Meanwhile, M. Du and coworkers prepared a COF-containing nanocomposite as a platform for the high-efficiency determination of Cr(III) [93]. As shown in Figure 8b, the nanocomposite (DNA/CoPc−PT−COF@Cu−MOF) was prepared by firstly growing two-dimensional phthalocyanine-based COF on Cu-MOF and then immobilizing DNA strands that can recognize Cr3+. The DNA/CoPc−PT−COF@Cu−MOF was coated onto a glass carbon electrode (GCE), yielding a Cr3+ sensor. It was found that the binding with the Cr3+ enhanced the charge transfer resistance, Rct, of the DNA strands, which can be detected by electrochemical impedance spectroscopy (EIS). In the meantime, DNA/CoPc−PT−COF@Cu−MOF also displayed a photoelectrochemical property. Cr3+ can result in a decrease in the photocurrent. Hence the Cr3+ can be detected by monitoring both Rct and photocurrent. The calculated LOD of the Cr3+ biosensor was as low as 14.5 fM. Very recently, two fluorescent COFs that can selectively detect Pd2+ were reported by J.-Y. Yue et al. and Y. Lu. et al., respectively [94,95]. Their fluorescence can be selectively quenched by Pd2+. Interestingly, the Pd2+-sensitive COF by Y. Lu. et al. was prepared in situ on a piece of paper, yielding a COF-based test paper for Pd2+ detection. Considering the wide application of palladium in the industry, it can be expected that the Pd2+ test paper will show good practicability in the fast detection of the catalyst residual.

3.2.2. pH Sensing (H+ Sensing)

Similar to the detection of the acidic/alkaline gases, introducing an acid/alkaline responsive group is the key to creating a COF for pH sensing (H+ sensing in water). [100,101,102,103,104] Early in 2016, Y. Mu and X. M. Liu et al. reported a COF (COF−JLU4) with protonatable β-ketoenamine in its skeleton [104]. The fluorescence spectra of COF−JLU4 can gradually change with the pH value in aqueous solution, indicating that pH-responsive luminescent COFs can be a candidate for pH sensing. Subsequently, several COFs of which fluorescence is responsive to pH value were developed (Table 4) [101,103,104]. Notably, C. L. Zhang et al. synthesized an imine-linked fluorescent COF containing triazine groups (COF2) [100]. The protonation of both imine linkage and triazine affected the intensity of fluorescence. Interestingly, the plot of fluorescence titration vs. pH showed a significant jump when pH = 6.5. It is suggested that COF2 can be used in the detection of cancer as a fluorescent probe. Thus, C. L. Zhang et al. modified the 2D nanosheet of COF2 with poly-D-lysine to improve its dispersity in aqueous solution, biocompatibility, and endocytosis efficiency. The experiment result exhibited that COF2 modified by poly-D-lysine can be successfully used in the in vitro cancer cell imaging and in vivo pH imaging in zebrafish.
In addition to fluorescent quenching, COF-based pH sensing can be conducted in other ways. For example, L. Wang and coworkers synthesized an imine-linked COF (COFDHTA-TTA) with multiple redox-active states [102]. Owing to the redox activity, the COFDHTA-TTA, showed a good performance in catalyzing the decomposition of H2O2. Therefore, it can be expected that a COFDHTA-TTA coated glassy carbon electrode (GCE) can be applied in the electrochemical analysis of H2O2. Additionally, the electron transport process of COFDHTA-TTA/GCE was accompanied by H+ transfer. Thus, pH could be also determined by using COFDHTA-TTA/GCE based on the peak potential as signal outputs. The experiment result showed that the peak potential exhibited a good linear relationship with the pH value of the solution in the experiment. In addition, the authors loaded glucose oxidase into the pores of COFDHTA-TTA. It made the COFDHTA-TTA coated GCE give a good performance for glucose detection.

3.2.3. Inorganic Anions Sensing

The inorganic anions in aqueous solution mainly include various acid radical ions. However, the COF for sensing inorganic anions was not frequently investigated as cations. Only a few reports can be found (Table 5). As mentioned above, the Fe3+ sensitive COF reported by Z. Shi et al. [80] can also be used for detecting CrO42−, Cr2O72−, and MnO4 via fluorescence quenching. The detection of Cr(VI) is of practical value because it is a water pollutant with significant carcinogenicity. So far, the inorganic ion that attracted the most attention is F [103,104,105], because excessive F in drinking water can damage human teeth and bones.
Early in 2015, X. M. Liu et al. reported on a boron-containing COF (BCMP−3) [105]. Unlike other anions, the Lewis acid–base interaction between the B atom and F can result in the change of fluorescence of BCMP−3 (Figure 9a). A remarkable blueshift and intensity decrease can be found upon the addition of F. The BCMP−3 can be also used for F removal. The adsorption capacity was 24 mg/g. Subsequently, D. L. Jiang and his coworkers synthesized a luminescent COF (TFPPy−DETHz−COF) that can be used for F sensing [106]. They found that F can react with the N−H unit in the hydrazone linkage of TFPPy−DETHz−COF based on an acid−base reaction mechanism, while other halogen anions like Cl, Br, I were inert (Figure 8b). The reaction with F deprotonated the N−H bond to form an anionic nitrogen species which can eliminate the nitrogen-related fluorescence quenching pathway, leading to a remarkable enhancement of the fluorescence intensity. Therefore, TFPPy−DETHz−COF was a rare fluorescence switch-on sensor because its intensity increases as [F] is increased. The detection limit of F detection was down to 50.5 ppb. Additionally, the F sensing based on the analogues of COFs, covalent triazine frameworks (CTF), was also reported. The F- sensing can be conducted by the CTF in a pathway without fluorescence. In the same year, Y. H. Xiong demonstrated an iron-modified two-dimensional CTF (2D−Fe−CTF) which exhibited remarkable peroxidase-like activity. F can coordinate with the iron ion in the CTF and significantly decreased the enzyme-mimicking activity [107]. Therefore, the concentration of F can be detected by monitoring the rate of H2O2 decomposition. The LOD of this approach can be as low as 5 nM, and other anions such as Cl, Br, SO42−, NO3, HCO3, HCOO, CO32−, PO43−, and CH3COO did not show significant interference.
In addition to F, the COFs for sensing S2− in solution have also been focused on [71,108], because H2S can act as a signal transmission molecule in the human body, the same as NO and CO [109,110], and participates in some important pathophysiologic processes such as neuromodulation and vasodilation. As demonstrated above, the organic polymer nanotubes PNT−1 can be used for detecting Na2S in CH3OH solution. Its fluorescent emission decreased with the addition of Na2S. However, for sensing the H2S in living bodies, it is still necessary to develop the COF that can respond to the trace amount of H2S (HS, S2−) in aqueous solution. In 2018, P. Wang et al. reported a COF-based hybrid probe, TpASH-NPHS, for targeting H2S around the living cells [108]. The hybrid probe was prepared by covalently anchoring a two-photon fluorescent probe 4-amino-1,8-naphthalimide derivative (NPHS) onto an imine-linked COF, TpASH. To further enhance the dispersibility, the micron-sized, bulky TpASH-NPHS was treated by solvent-assisted liquid sonication. The resulting nano-sized TpASH-NPHS can be well dispersed into the aqueous solution. The azide units on NPHS can be reduced to NH2 in the presence of H2S (HS, S2−), which will increase the fluorescence emission of NPHS. Thus, TpASH-NPHS showed a good performance in two-photon fluorescence imaging of H2S (HS, S2−) in live tumor cells and deep tumor tissues. The hybrid fluorescent probe TpASH–NPHS was able to detect the concentration of H2S (HS, S2−) in the range of 0–20 mM with a LOD of 0.11 mM. Furthermore, TpASH–NPHS was inert to a wide range of interferents, including common anions, biothiols, cysteine, homocysteine, and reactive oxygen species. Notably, since the NPHS was immobilized in the 1D channel of which the diameter is about 1.5 nm, the H2S (HS, S2−) imaging by TpASH–NPHS was not interfered by some intracellular enzymes (diameter > 3 nm), which are also able to reduce the azide group under hypoxic conditions. Hence, this COF-based hybrid probe displayed a great promise in clinical applications.
Figure 9. (a) BCMP−3 for F sensing [109]. Copyright 2015 John Wiley and Sons; (b) TFPPy−DETHz−COF for F sensing [110]. Copyright 2018 American Chemical Society.
Figure 9. (a) BCMP−3 for F sensing [109]. Copyright 2015 John Wiley and Sons; (b) TFPPy−DETHz−COF for F sensing [110]. Copyright 2018 American Chemical Society.
Molecules 27 02586 g009

3.3. Molecular Sensing

3.3.1. Explosive Sensing

Explosive detection is widely required in the security of airports, railway stations, and border checks. Therefore, plenty of effort has been devoted to the fast, portable detection of explosives. Various COFs that can be used for explosive sensing have been developed (Table 6) [76,111,112,113,114,115,116,117,118,119,120,121,122,123,124,125,126,127,128129,130].
Nitroaromatics, the most widely used explosive, have an electron-deficient aromatic ring which is an effective fluorescence quencher. It can be expected that the fluorescence of some COFs can be easily quenched by adsorbed nitroaromatics. Therefore, the key point to realizing COF-based nitroaromatics detection is designing a fluorescent COF that can effectively adsorb nitroaromatics. Early in 2012, L.-G. Qiu and coworkers demonstrated a melamine-based COF (SNW−1) for trace-level detection of nitroaromatic explosives [111]. The melamine-based fluorescent COF was synthesized by the microwave-assisted reaction between terephthalaldehyde and melamine. It was found that the fluorescent emission of SNW−1 can be quenched after adsorbing nitroaromatics due to the transfer of photoexcited electrons from the excited luminescence network donor to the electron-deficient nitroaromatic acceptors. In the THF–water (1:9, v/v) mixture, the intensity of the fluorescence linearly decreased with the concentration of nitroaromatics including nitrobenzene, 4-nitrotoluene, 4-nitrophenol, 2,4-dinitrotoluene (DNT), picric acid, 2,4,6-trinitrotoluene (TNT), and 2,4,6-trinitrophenylmethylnitramine. The LODs for TNT and DNT were 1.51 × 10−6 mol L−1 and 2.8 × 10−6 mol L−1, respectively. Remarkably, the LOD for picric acid can be as low as 5.0 × 10−8 mol L−1 (11.5 ppb). Furthermore, L.-G. Qiu and coworkers also applied SNW−1 to detect the vapor of DNT, which is a volatile by-product of commercial TNT. It can be found that the fluorescence of SNW−1 decreased by 50% within 10 s, suggesting an ultrafast response to DNT. The LOD of SNW−1 for DNT vapor detection was calculated to be 9.8 ppb, much lower than most DNT detectors that had been reported.
In the following years, many COFs of which the fluorescence can be quenched by nitroaromatics were found. For example, D. Jiang and coworkers reported an azine-linked COF, of which the fluorescence can be quenched by some nitroaromatics like picric acid, 2,4-dinitrophenol (DNP), DNT, 2-nitrophenol, and 2-nitrotoluene (Figure 10a) [109]. It can be also found that the quenching result from picric acid was much more significant than other nitroaromatics. The authors demonstrated the strong quenching effect can be attributed to the hydrogen bond formed by the -OH group of the analytes and the open N atoms on the COF skeleton. In addition, the three nitro groups of picric acid resulted in the most deficient π system for driving the fluorescence quenching and the strongest hydrogen bond with the azine for promoting the quenching process. After the reveal of the mechanism, many COFs for picric acid detections were successfully designed and synthesized [112,113,114,115,116,117,118,119,120,121,122,123,124,125,126,129,130]. The LOD has been decreased to ~10 nM (Table 6). Furthermore, the interference from other nitroaromatics especially dinitrophenol and nitrophenol has been reduced to the largest extent [112,113,114]. Moreover, considering that other nitroaromatics including dinitrophenol and nitrophenol are also explosive, it can still be concluded that the fluorescence quenching strategy based on COFs is an effective method for the nitroaromatic explosive initial inspection.
In addition, the COF-based detection of non-nitroaromatic explosives has been also presented. In 2018, D. F. Perepichka synthesized two kinds of fluorescent COFs (3BD and 3′PD) via Michael Addition−Elimination [115]. The fluorescence of the two COFs cannot only be quenched by nitroaromatics like picric acid but is also sensitive to triacetone triperoxide (TATP), which is another well-known explosive. The quenching caused by TATP was possibly attributed to the oxidation of the enamine moieties in 3BD and 3′PD (Figure 10b). A total of 1 μM TATP solved in CH2Cl2 can be detected. The authors suggested that the quench should not result from the oxidation of the enamine moiety in the COF skeleton rather than the electron transfer from excited COFs to the analyte, like nitroaromatics.

3.3.2. Iodine Sensing

Radioiodine (129I and 131I), which is generated by nuclear fission, is an important index for detecting nuclear fallout. It is pernicious to the human body because of its long radioactive half-life (129I, 1.57 × 107 years) [131] and high volatility. Therefore, it is widely concerned to detect radioiodine fast and sensitively. COF and its analogs have been proved to be excellent adsorbents for iodine removal. For example, TPB−DMTP−COF synthesized by D. L. Jiang and coworkers can effectively adsorb I2 vapor with an adsorption capacity of 6.2 g/g [132]. As with nitroaromatics, the molecule of iodine is an electron-deficient species. The charge transfer from the electron-rich COF skeleton to the iodine can significantly quench the fluorescence of the COF. Thus, it is very likely to achieve “turn off” fluorescence sensing to iodine by COF or its analogs. Recently, a number of iodine sensors based on COPs was presented [133,134,135,136,137,138,139]. In 2020, T. M. Geng et al. prepared a fluorescent covalent triazine-based framework (TDPDB) by Friedel–Crafts polymerization reaction of N, N−-diphenyl-N, N−-di(m-tolyl)benzidine (DPDB) with 2,4,6-trichloro-1,3,5-triazine (TCT) [140]. It exhibited an excellent capacity for capturing iodine (3.93 g/g). In the meantime, the adsorption of iodine quenched the fluorescence of TDPDB remarkably. The quenching coefficient (Ksv) was measured as 5.83 × 104 L mol−1 and the LOD reached 2.57 pM. It indicated that TDPDB can be used for highly sensitive I2 detection. Furthermore, the concentration of PA can be also measured by TDPDB based on the same mechanism. The Ksv and LOD for PA detection were determined as 1.55 × 104 L mol−1 and 19.3 pM, respectively. The next year, Y. X. Zhao et al. reported a COF (COF-PA) with acetylene groups in its 1D channels [131]. The acetylene groups greatly accelerated the adsorption rate of I2 vapor. The charge transfer between COF-PA and I2 effectively decreased the intensity of the fluorescence of COF-PA. I2 at ppm level in THF solution can be well detected by the fluorescence quenching of COF-PA.

3.3.3. Drug Sensing

Drug sensing has a strong practical value and significance because it is a key essential step in investigating pharmacokinetics, which is quite instructive in clinical treatment. Moreover, the drug residue, especially the antibiotics’ residue, in food and drinkable water because of the abuse of drugs was frequently reported in recent decades. It is likely to result in the appearance of the “superbug” by which the infection caused is incurable because it is resistant to the existing antibiotics. Hence, the fast detection of drugs, especially antibiotics, in bodily fluids, food, drink, and environment attracted increasing attention. COF-based detections of drugs, especially various antibiotics [49,107,140,141,142,143,144,145,146,147,148,149,150,151,152,153] were frequently reported in recent years, especially in the past five years. Notably, as shown in Table 7, it can be found that, unlike gases or ions, most of the COF-based drug detection is accomplished by a hybrid material containing COFs, rather than by a separate COF, because most of the binding sites for the selective capture of the drug molecules, including aptamers, metal ions and so on, were difficult to be robustly pre-immobilized onto the monomers. They are easy to detach from the monomers or be damaged during the solvothermal synthesis of COFs. Hence, they can only be introduced by the “post-functionalization” of the COFs. In addition, in some COF-based sensors for drug detection, the selective capture of drug molecules was not always achieved by the COFs, but by other materials like molecular imprinting membranes.
Tetracycline, a broad-spectrum antibiotic, is widely used in the treatment of bacterial infections in humans and animals. Therefore, tetracycline residues were frequently found in the products of the breeding industry or aquaculture. Thus, as shown in Figure 11a, X. H. Ma synthesized a Zr-coordinated amide porphyrin-based COF by a liquid-liquid interface method [144]. The COF exhibited strong catalytic effects and high electrical conductivity. Hence it was immobilized onto a GCE via the polymerization of the o-phenylenediamine to yield an electrochemiluminescence (ECL) sensor. In the meantime, molecular imprinting technology was used to enhance the selectivity of tetracycline. It was found that the intensity of the ECL signal obtained using a luminol-H2O2 system linearly increased with the concentration of tetracycline. The linear range was 5–60 pM and the LOD was calculated as 2.3 pM. In 2021, Y. S. Zhang reported a COF-containing porous hybrid, Mg@Fe−MIL−101/TpPa−1−COF, grown on stainless steel mesh (Figure 11b) [152]. The porous property of Mg@Fe-MIL-101 exposed more catalytical active sites, and the hybrid structure facilitated the electrons’ migration. Therefore, the Mg@Fe−MIL−101/TpPa−1−COF exhibited excellent photocatalytic tetracycline degradation capacity. Meanwhile, the Mg2+ in the hybrid can coordinate with tetracycline and form the fluorescent Mg-tetracycline complex. Hence, fluorescent detection can be realized simultaneously with photocatalytic tetracycline removal (Figure 11b).
Due to the side effects and increasing resistance, the use of antibiotics in the clinic treatment is being updated, therefore, the COF-based antibiotic detection needs to be constantly extended. B. Yan and coworkers prepared a Eu3+ functionalized COF hybrid material (Eu@TpPa−1) by simply soaking TpPa-1 COF [158] in a EuCl3 solution [145]. It was found that due to the coordination between Eu3+ and levofloxacin, Eu@TpPa−1 can acted as a fluorescence sensor toward levofloxacin with a turn-on response. The linear range of the levofloxacin sensing was 1−104 μM, and the LOD was calculated as 0.2 μM. The sensor also displayed a fast response to levofloxacin which was less than 1 min. In the meantime, it also exhibited a good selectivity refraining from the interference of other components in serum and urine. In 2021, S. Wang reported a polyimide COF of which fluorescence can be quenched by levofloxacin, tetracycline, and some metal ions [153]. It exhibited a potential application in the fast identification of multiple harmful substances in drinkable water.
In addition to antibiotics, COF-based sensing of other drugs was also reported (Table 7). For example, 5-fluorouracil (5FU), an anticancer drug, can be detected by a COF hybrid material (EB-TFP:Eu(BTA)4) prepared by B. Yan et al. (Figure 11c) [154]. The hybrid material was prepared by soaking a previously reported COF [159] in an Eu complex. The adsorption of 5FU can disturb the π−π stacking interaction between the 2D sheets of EB-TFP:Eu(BTA)4), leading to the change in the fluorescent spectra. It resulted in an increase in the fluorescent peak at 615 nm and an decrease in the peak at 500 nm. The intensity ratio I615/I500 linearly decreased with the logarithm of the concentration of 5FU. Therefore, a 5FU sensing based on fluorescence can be achieved. The linear detection range of the 5FU sensing was 10−7−10−3 M. The calculated LOD was 5.13 × 10−8 M. In addition, B. Yan et al. demonstrated a dye-functionalized COF for ultrasensitive monitoring of sialic acid, which is a widely-used cosmetic additive and biomarker of ovarian cancer. Similar to the above-mentioned Eu@TpPa−1, B. Yan et al. loaded fluorescein sodium (TpPa−1@Dye) into the porous TpPa-1 COF [154]. The TpPa−1@Dye was mixed in sodium alginate solution and finally formed a hydrogel with the assistance of Cr3+. The interaction between TpPa−1@Dye and Cr3+ caused fluorescence quenching of fluorescein, while the sialic acid can take away Cr3+ from the TpPa−1@Dye leading to the recovery of the fluorescence. The fluorescence recovery is fast and very sensitive to the addition of sialic acid, the LOD of the sialic acid detection in serum by this approach can be as low as 8.71 × 10−9 M.

3.3.4. Small Biomolecules Sensing

The level of some small biomolecules, including glucose, uric acid, Glutathione (GSH), et al., are key marks of some serious diseases. Hence the detection of them is very important in the healthcare and clinic diagnosis. Recently, with the demands for portable and wearable healthcare devices, the reports on new materials for facile detection of the above small biomolecules have been increasing. Recently, some COFs or COF-based hybrid materials were designed and synthesized for detecting glucose, uric acids, GSH, and other small biomolecules (Table 8) [26,27,102,160,161,162,163].
The concentration of glucose in blood is an essential index for the diagnosis of diabetes. Plenty of methods to detect the concentration of glucose have been developed. The basic strategy to sense glucose is by detecting H2O2 yielded by the enzymic oxidation of glucose. In 2018, J. N. Wang et al. and co-workers prepared a Fe-porphyrin-based peroxidase-mimicking COF (Fe−COF) [160]. Like the COF demonstrated by J. D. Qiu and coworkers [148], Fe−COF can be used to quantitatively detect the [H2O2] by the generation of the oxTMP with blue color. Furthermore, upon the mixing of glucose oxidase (GOx), the combination of the enzymic glucose oxidation and Fe−COF-catalyzed oxTMP production resulted in the realization of a glucose sensor. the detection range of the sensor was from 5 to 350 μM. The detection limit was 1.0 μM. On the basis of this principle, several COF-based glucose sensors with improved properties were developed [102,161,162]. For example, the function of the pH sensor mentioned above by L. Wang and coworker can be extended for glucose sensing [102]. Recently, J.-Yu Yue et al. prepared a glucose sensor by coupling the GOx on a carboxyl-containing COF via a covalent amide bond [162]. The sensor showed good recoverability and storage stability. Even after 100 days of storage, the relative activity of GOx was still more than 85%. In the meantime, this sensor displayed a wider linear range from 0.005 to 2 mM and a lower LOD which was calculated as 0.54 mM.
The level of uric acid in serum is the most important index for the diagnosis of gout, which is a common ailment in developed countries. In 2021, Y. S. He et al. developed a COF-based electrochemical sensor for detecting uric acid [163]. For fabricating the sensor, a poly-fuchsin basic film was fabricated on the surface of a bare glass GCE. Then Au nanoparticles were doped on a carboxylated COF (ACOF-TaTp), yielding a hybrid, AuNPs@ACOF-TaTp. Finally, the sensor was achieved by dropping the hybrid on the poly-fuchsin basic coated GCE. The COF in the sensor provided a high surface area and numerous binding sites (H-bond binding) for the analyte. The existence of AuNPs enhanced the conductivity, while the poly-fuchsin basic film can accelerate the electron transfer process. Therefore, the sensor exhibited high affinity to the analyte and high sensitivity and selectivity in uric acid sensing. It was reported that there were two linear response ranges of the AuNPs@ACOF-TaTp-based uric acid sensor, 5.0–25 μM, and 25–250 μM. The LOD was calculated as 0.77 μM, (normal uric acid level of human: 89–416 μM) indicating the sensor has the potential to be used in the detection of uric acid in serum. In addition, the sensor can be applied to detect the concentrations of ascorbic acid and dopamine, which are also important biomolecules.
GSH is a biomolecule that widely exists in all kinds of cells in the human body. The excessive glutathione in serum usually suggests hepatocellular injury or tumor. In 2020, H. L. Chen and coworkers reported an amine-linked COF (COF−300−AR) obtained by reducing the corresponding imine-linked COF (COF-300) (Figure 12a) [26]. It exhibited excellent oxidase-mimicking activity under light irradiation of which λ = 400 nm, catalyzing the oxidation of TMB by the formation of reactive oxygen species like •OH and O2•− free radicals in the presence of dissolved oxygen. However, the reducibility of GSH can prevent the oxidation of TMB and lighten the blue color resulting from oxTMB. Therefore, the concentration of glutathione can be measured by colorimetric analysis. The authors successfully applied COF−300−AR to detect GSH levels in HL60 cells. It displayed good selectivity and high sensitivity. Similarly, Z. Lin et al. synthesized a COF (Py−TT COF) using the condensation between tetrakis(4-aminophenyl)pyrene (Py), which can act as an electron donor, and thieno[3,2-b]thiophene-2,5-dicarbaldehyde (TT), which can act as an electron acceptor (Figure 12b) [27]. Like COF−300−AR, Py−TT COF exhibited superior enzymatic catalytic activity, which could catalyze the oxidation of TMB by the formation of ROS under visible light irradiation. Therefore, based on the similar mechanism, the concentration of GSH can be determined by the decrease of the blue color generated by oxTMB. The linear range of this approach was 0.4 − 60 μM, and the LOD was 0.225 μM, satisfying the detection of the GSH in human serum.

3.3.5. Other Small Molecules Sensing

Besides the residue of antibiotics, the detection of insecticides residual is also an important problem in agriculture [155,156]. For insecticide sensing, Y. H. Song et al. synthesized a COF by the imine condensation (DAAQ−TFP) between 2,6-diaminoanthraquinone (DAAQ) and 1,3,5-triformylphloroglucinol (TFP) [155]. DAAQ−TFP exhibited a very good capacity for adsorbing benzoylurea insecticides due to the hydrogen bond and π-π interaction between the benzoylurea and COF skeleton. The benzoylurea insecticides adsorbed by DAAQ−TFP can be eluted by organic solvent for quantitative detection by HPLC. This approach can be applied for detecting insecticide residuals in fruit samples and provided very low detection limits (~0.1 ng mL−1).
The COF-based detection of spice, of which concentration is quite important in the food industry, was also investigated. H. L. Liu can co-workers reported a fluorescent nanoprobe (CN-grafted COF@MIPs) for the detection of 4-ethylguaiacol, which is a crucial index to evaluate the quality of some kinds of wine [157]. The probe was a 4-ethylguaiacol imprinted polymer doped with a COF grafted onto carbon nanodots. The molecular imprinting and -NH2 in the hybrid material enhanced the selective uptake of the analyte. The adsorption of 4-ethylguaiacol quenched the fluorescence of carbon nanodots through charge transfer interaction. Under optimized conditions, the fluorescence of CN-grafted COF@MIPs drops linearly as the concentrations of 4-ethylguaiacol increase from 0.025 to 1 μg mL−1, with a LOD of 17 ng mL−1.
Finally, it is worth mentioning that Y. Cui and co-workers reported a chiral 1,1′-bi-2-naphthol-based (BINOL-based) COF for sensing chiral organic compounds [164]. The COF was synthesized by imine condensations of tetrakis(4-aminophenyl)ethene (TPE-TAM) and the chiral linker (R or S) 6,6′-dichloro-2,2′-diethoxy-1,1′-binaphthyl-4,4′-dialdehyde (BINOL-DA). The chiral linker created a 1D channel with high enantioselectivity for the discriminating uptake of chiral analytes. For example, the nanosheets of the COF synthesized by (R)-BINOL-DA can adsorb more (−)-α-pinene than (+)-α-pinene. Thus, the (−)-α-pinene can result in a faster and greater fluorescence quench than (+)-α-pinene at the same concentration. The authors found that the BINOL-based COF nanosheets exhibited much better selectivity and enantio-sensitivity to other BINOL-based sensing systems, probably due to the steric confinement of the COF channels and conformational rigidity of the immobilized BINOL groups.

3.3.6. Biomacromolecule Sensing

Biomacromolecules, including polysaccharides, proteins, nucleic acids, et al., are essential components of living organisms and play an important role in various life activities. The detection of biomacromolecules, especially proteins and nucleic acids, is extensively needed in clinical treatment, public health, and environmental protection. Recently, a number of COFs or COF-containing hybrid materials have been developed for sensing proteins and nucleic acids (Table 9).
Although protein is a kind of biomacromolecule with huge diversity, it is not difficult to specifically enrich the target protein from the sample solution, because most proteins can be recognized and captured by the corresponding antibody. Due to the large specific surface area, COFs are good carriers for immobilizing the antibodies. Moreover, other functional species can be simultaneously introduced onto the COF sheets to produce a detectable signal after the target protein was bound via antigen-antibody recognition. Based on this strategy, some COF-containing sensors for detecting protein with diagnostic significance were developed recently [165,166,167,168,169].
In 2018, T. Zhang et al. reported a COF-based immunosensor for detecting cardiac troponin I (cTnI), which is an important marker for angina pectoris and myocardial infarction (Figure 13a) [165]. Firstly, AuNPs were loaded into the pore of a previously-reported COF [170], then cTnI secondary antibody (Ab2-cTnI) was immobilized onto the AuNPs. Finally, toluidine blue (TB) was loaded into the COF, yielding a COF-based label, TB−Au−COFs−Ab2. Meanwhile, a GCE was coated by TiO2 NPs which is successively modified by polypyrrole, AuNPs, and cTnI primary antibody (Ab1-cTnI). By the specific recognition antigen-antibody, the TB−Au−COFs−Ab2, analyte, and modified GCE can form a sandwich-like structure, resulting in a redox signal that can be detected by square-wave voltammetry. The concentration of CTnI can be determined by the intensity of the square-wave voltammetric signal. The authors found that the AuNP-doped COF can amplify the electrochemical signal and enhance the performance of the immunosensor. The immunosensor showed a linear range from 0.5 pg mL−1 to 10.0 ng mL−1 and a low detection limit of 0.17 pg mL−1. In 2021, S. N. Feng et al. reported a similar approach to detect cTnI [166] (Figure 13b). TB in the above work was replaced by horseradish peroxidase (HRP) to fabricate the COF-based label. Meanwhile, Ab1-cTnI was modified directly on a gold electrode instead of GCE, avoiding the complex modification procedure. After the formation of the antibody-antigen-antibody sandwich structure, H2O2 and hydroquinone was added, and the intensity of the redox signal of hydroquinone can be detected by cyclic voltammetry and used to determine the concentration of cTnI.
X. Fang and coworkers prepared a COF-coated Fe3O4 NPs as an immunoaffinity probe for the detection of heat shock protein 90α (Hsp90α), a biomarker for liver cancer [167]. The probe was prepared by in situ growth of COFs on an -NH2 coated Fe3O4 NPs, and successively modified by Hsp90α antibody and PEG. The probe can selectively capture Hsp90α in the sample and be collected by the magic field. The Hsp90α on the probe was digested with free trypsin and the resulting peptides’ mixture can be collected for further MALDI-TOF-MS analysis to determine the existence of Hsp90α.
Nucleic acid is an essential biomacromolecule in living bodies and the carrier of genetic information. Nucleic acid detection is widely applied in the identification of bacteria and viruses. During the pandemic caused by COVID-19, nucleic acid detection played a huge part in tracking the virus and preventing the spread of the disease. The key feature of nucleic acid is the base pairing. It allows the detection of a particular nucleic acid segment by using its complementary sequence. Based on this principle, some COF-based detection of DNA was developed [28,171,172,173].
In 2017, X.-P. Yan demonstrated a versatile COF-based platform for sensing DNA. The sensor is on the basis of the recognition of the complementary DNA sequences and fluorescent quenching (Figure 14a). For detecting a particular DNA sequence (target DNA), its complementary sequence modified with a carboxyfluorescein (FAM) label was adsorbed onto a COF (TpTta) [171] containing triazine, amino- and carbonyl-group via hydrogen bonding, and π-π interaction. The adsorption of the FAM-labeled complementary sequence led to the quench of the fluorescence. However, the hybridization between the target DNA and the FAM-labeled complementary sequence led to the formation of double-stranded DNA that can detach from TpTta, resulting in the fluorescent recovery of the carboxyfluorescein. Moreover, the author found that quantitively ATP detection can be also achieved by replacing the above FAM-labeled complementary sequence with the FAM-labeled adenosine 5′-triphosphate (ATP) aptamer probe. The enhanced fluorescence intensity showed a good linear relationship with the ATP concentration in the concentration range of 25–200 μM and good selectivity.
M. R. Xu et al. developed a novel MOF@ COF hybrid with excellent electrochemical activity and high photoactivity (Figure 14b) [28]. By immobilizing the complementary sequence, the MOF@ COF hybrid can be applied for HIV-1 DNA detection, because the complementary hybridization enhanced the impedance of the hybrid and decreased both the photoelectrochemical and electrochemical currents. Hence, it is possible to achieve a dual-mode detection of HIV-1 DNA in human serum. It was found that both the currents decreased linearly with the increasing lg [HIV-1 DNA], in the range from 1 fM to 1 nM, and the LODs were calculated as 0.07 and 0.18 fM for PEC and DPV techniques, respectively, indicating that the ultra-sensitive sensing of HIV-1 DNA was realized.
Table 9. COF-based sensing of some typical biomacromolecules.
Table 9. COF-based sensing of some typical biomacromolecules.
AnalyteYearCOF/COF Hybrid NamesSpecific Binding SiteType of Detectable SignalDetection RangeLODReference
Cardiac Troponin I2018TB−Au−COFs−Ab2AntibodyChromism 0.5 pg/mL–10.0 ng/mL 0.17 pg/mL[165]
2021HRP−Ab2−Au−COFAntibodyChromism5 pg/mL–10 ng/mL1.7 pg/mL[166]
Heat shock protein 90α2019Fe3O4@TpBD−DSS−Ab−MEGAntibodyMS 50 pg/mL[167]
C-reactive protein2018AuNPs @COF−TPPa-1AntibodyElectrochemical signal (EIS, CV)0.017 ng/mL0.05–80 ng/mL[168]
2018COF−LZU8AntibodyElectrochemical signal (DPV)0.016 ng/mL0.05–150 ng/mL[169]
2019p−COFAptamerPhotoelectrochemical signal0.1 ng/mL0.5–100 ng/mL[174]
DNA2017TpTtaDNA hybridizationFluorescence, “turn on”10–100 nM3.7 nM[171]
2017TPA−COFDNA hybridizationFluorescence, “turn on”0.02–5 nM20 pM[172]
2018EB-TFP iCOFDNA hybridizationFluorescence, “turn on”0–32 mM-[173]
2021Cu−MOF@CuPc−TA−COFDNA hybridizationElectrochemical signal 1 fM–1 nM 0.18 fM[28]
Photoelectrochemical signal0.07 fM

4. Challenges and Future Perspectives

In this review, we discussed the basic principles of COF-based chemical and biological sensing, then demonstrated the application of COFs in the detection of a large number of analytes, including corrosive and harmful gases, water vapor (humidity), ions, explosives, drugs, biomolecules, et al. It can be found that COFs exhibited many advantages in chemical and biological sensing. Firstly, due to the designable structure and easily modified properties, the molecular structure and function of the COFs can be easily extended. There is always a good possibility of designing and synthesizing a COF with a specific function. Hence, a wide range of chemicals and biomolecules can be selectively detected by COFs or COF-based hybrid materials. Secondly, the porous structure of the COFs greatly benefits the enrichment of the analyte, significantly enhancing the sensitivity of the detection. Furthermore, the binding site that can selectively interact with the analyte can be easily linked to the COFs. It will remarkably enhance the selectivity of the sensing. Thus, it can be seen that COFs are good candidates for realizing highly sensitive and selective sensing. Thirdly, diverse aromatic structures endow the COFs with rich optical and electrical properties. Moreover, the unique nature of the COFs allows the COFs to play multiple roles in the chemical and biological sensors, including an analyte concentrator, signal generator, and functional particles’ carrier. It will significantly decrease the complexity of the sensing devices.
The above advantages exhibited that COFs are a class of promising materials for chemical and biological sensing. Although great progress has been made, COF-based sensors are still under investigation in the laboratory. However, there are still some remaining disadvantages that inhibit the COF-based sensors from practical application. They will be the important challenges to promoting the utilization of the COF-based sensors. The first challenge is the mass production of the COFs. The monomers for preparing functional COFs need to be obtained by organic synthesis. The complex synthetic route may significantly decrease the yield of the final products, and enhance the cost of COFs production. Furthermore, the method to control the polymerization of the monomers and the crystallization process during the COFs formation is still far from perfect. Obtaining COFs with uniform quality is still difficult to guarantee. Secondly, the most widely used technique for COFs synthesis is the solvothermal method. Therefore, most COFs are obtained as a powder. However, for sensor fabrication, the thin film can exhibit much better performance than the powder due to its uniformity and compactness. Therefore, effective approaches to preparing COFs thin films on various substrates are still eagerly expected. Thirdly, with the development of wearable and implantable sensing devices, the demand for materials with good biocompatibility is increasing. However, the data on the biocompatibility of reported COFs are still deficient. It may limit the further application of COFs in biosensing. COF materials with biocompatibility and biosafety should be developed for promoting its application in domestic, wearable, and implantable sensing devices.
In recent years, substantial efforts have been devoted to creating new COFs with various functions. The increasing number of new COFs will continue to promote its application in the development of chemical and biological sensors. Although the COF-based sensors have been utilized for detecting various analytes, it can still be expected that the emergence of new COFs can further extend the range of the analytes. The selectivity and sensitivity will be further improved. Moreover, with the growing demand for fast and real-time detection of pollutants, dangerous articles, and physiological indexes. Portable, wearable, or even implantable sensors is drawing increasing attention. As mentioned above, COFs are usually able to play multiple roles in a sensor, for example, being an analyte concentrator and signal generator simultaneously. It can remarkably reduce the size and complexity of the sensor. Therefore, COFs also display great potential in constructing miniaturized sensing devices. Additionally, it is also promising to develop COF-based sensors which are suitable to work in special environments, such as the polar region, space environment, and body fluids. Thanks to the increasing eagerness for environmental protection, security, food safety, and healthcare, developing new COF-based chemical or biological sensors with better performance and new applications is still far from the end.

Author Contributions

Writing—original draft preparation, G.W. and S.Z.; writing—review and editing, S.Z. and D.L.; project administration, D.L.; funding acquisition, D.L. and G.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the National Natural Science Foundation of China, grant number 52172085, 21704023.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

We thank Liangcan He from Harbin Institute of Technology for helpful discussion.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Diamond, D.; Coyle, S.; Scarmagnani, S.; Hayes, J. Wireless Sensor Networks and Chemo-/Biosensing. Chem. Rev. 2008, 108, 652–679. [Google Scholar] [CrossRef] [PubMed]
  2. Joo, S.; Brown, R.B. Chemical Sensors with Integrated Electronics. Chem. Rev. 2008, 108, 638–651. [Google Scholar] [CrossRef] [PubMed]
  3. Aksyuk, V.A. Sensing without Power. Nat. Nanotechnol. 2017, 12, 940–941. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Lou, Z.; Wang, L.; Shen, G. Recent Advances in Smart Wearable Sensing Systems. Adv. Mater. Technol. 2018, 3, 1800444. [Google Scholar] [CrossRef] [Green Version]
  5. Mondal, S.; Zehra, N.; Choudhury, A.; Iyer, P.K. Wearable Sensing Devices for Point of Care Diagnostics. ACS Appl. Bio Mater. 2021, 4, 47–70. [Google Scholar] [CrossRef]
  6. Yang, B.; Jiang, X.; Fang, X.; Kong, J. Wearable Chem-biosensing Devices: From Basic Research to Commercial Market. Lab Chip 2021, 21, 4285–4310. [Google Scholar] [CrossRef]
  7. Liu, H.; Wang, L.; Lin, G.; Feng, Y. Recent Progress in the Fabrication of Flexible Materials for Wearable Sensors. Biomater. Sci. 2022, 10, 614–632. [Google Scholar] [CrossRef]
  8. Liu, Y.; Dong, X.; Chen, P. Biological and Chemical Sensors Based on Graphene Materials. Chem. Soc. Rev. 2012, 41, 2283–2307. [Google Scholar] [CrossRef]
  9. Zhao, Y.; Li, X.; Zhou, X.; Zhang, Y. Review on the Graphene Based Optical Fiber Chemical and Biological Sensors. Sensors Actuat. B-Chem. 2016, 231, 324–340. [Google Scholar] [CrossRef]
  10. Angizi, S.; Khalaj, M.; Alem, S.A.A.; Pakdel, A.; Willander, M.; Hatamie, A.; Simchi, A. Towards the Two-dimensional Hexagonal Boron Nitride (2D h-BN) Electrochemical Sensing Platforms. J. Electrochem. Soc. 2020, 167, 126513. [Google Scholar] [CrossRef]
  11. Gan, X.; Zhao, H.; Quan, X. Two-dimensional MoS2: A Promising Building Block for Biosensors. Biosens. Bioelectron. 2017, 89, 56–71. [Google Scholar] [CrossRef]
  12. Pumera, M.; Loo, A.H. Layered Transition-metal Dichalcogenides (MoS2 and WS2) for Sensing and Biosensing. TrAC-Trend. Anal. Chem. 2014, 61, 49–53. [Google Scholar] [CrossRef]
  13. Sinha, A.; Dhanjai; Zhao, H.M.; Huang, Y.J.; Lu, X.B.; Chen, J.P.; Jain, R. MXene: An Emerging Material for Sensing and Biosensing. TrAC-Trend. Anal. Chem. 2018, 105, 424–435. [Google Scholar] [CrossRef]
  14. Wales, D.J.; Grand, J.; Ting, V.P.; Burke, R.D.; Edler, K.J.; Bowen, C.R.; Mintova, S.; Burrows, A.D. Gas Sensing Using Porous Materials for Automotive Applications. Chem. Soc. Rev. 2015, 44, 4290–4321. [Google Scholar] [CrossRef] [Green Version]
  15. Lin, R.-B.; Liu, S.-Y.; Ye, J.-W.; Li, X.-Y.; Zhang, J.-P. Photoluminescent Metal-Organic Frameworks for Gas Sensing. Adv. Sci. 2016, 3, 1500434. [Google Scholar] [CrossRef]
  16. Koo, W.T.; Jang, J.S.; Kim, I.D. Metal-Organic Frameworks for Chemiresistive Sensors. Chem 2019, 5, 1938–1963. [Google Scholar] [CrossRef]
  17. Teo, W.L.; Liu, J.; Zhou, W.; Zhao, Y. Facile Preparation of Antibacterial MOF-fabric Systems for Functional Protective Wearables. SmartMat 2021, 2, 567–578. [Google Scholar] [CrossRef]
  18. Meng, Z.; Mirica, K.A. Covalent Organic Frameworks as Multifunctional Materials for Chemical Detection. Chem. Soc. Rev. 2021, 50, 13498–13558. [Google Scholar] [CrossRef]
  19. Liu, X.; Huang, D.; Lai, C.; Zeng, G.; Qin, L.; Wang, H.; Yi, H.; Li, B.; Liu, S.; Zhang, M.; et al. Recent Advances in Covalent Organic Frameworks (COFs) as a Smart Sensing Material. Chem. Soc. Rev. 2019, 48, 5266. [Google Scholar] [CrossRef]
  20. Allendorf, M.D.; Dong, R.; Feng, X.; Kaskel, S.; Matoga, D.; Stavila, V. Electronic Devices Using Open Framework Materials. Chem. Rev. 2020, 120, 8581–8640. [Google Scholar] [CrossRef]
  21. Geng, K.; He, T.; Liu, R.; Dalapati, S.; Tan, K.; Li, Z.; Tao, S.; Gong, Y.; Jiang, Q.; Jiang, D. Covalent Organic Frameworks: Design, Synthesis, and Functions. Chem. Rev. 2020, 120, 8814–8933. [Google Scholar] [CrossRef]
  22. Liu, R.; Tan, K.; Gong, Y.; Chen, Y.; Li, Z.; Xie, S.; He, T.; Lu, Z.; Yang, H.; Jiang, D. Covalent Organic Frameworks: An Ideal Platform for Designing Ordered Materials and Advanced Applications. Chem. Soc. Rev. 2021, 50, 120–242. [Google Scholar] [CrossRef]
  23. Jiang, D. Covalent Organic Frameworks: An Amazing Chemistry Platform for Designing Polymers. Chem 2020, 6, 2461–2483. [Google Scholar] [CrossRef]
  24. Li, Z.; He, T.; Gong, Y.; Jiang, D. Covalent Organic Frameworks: Pore Design and Interface Engineering. Acc. Chem. Res. 2020, 53, 1672–1685. [Google Scholar] [CrossRef]
  25. Li, Y.; Chen, W.; Xing, G.; Jiang, D.; Chen, L. New Synthetic Strategies Towards Covalent Organic Frameworks. Chem. Soc. Rev. 2020, 49, 2852–2868. [Google Scholar] [CrossRef]
  26. Jin, P.; Niu, X.; Zhang, F.; Dong, K.; Dai, H.; Zhang, H.; Wang, W.; Chen, H.; Chen, X. Stable and Reusable Light-Responsive Reduced Covalent Organic Framework (COF−300−AR) as an Oxidase-Mimicking Catalyst for GSH Detection in Cell Lysate. ACS Appl. Mater. Interfaces 2020, 12, 20414–20422. [Google Scholar] [CrossRef]
  27. Li, G.; Ma, W.; Yang, Y.; Zhong, C.; Huang, H.; Ouyang, D.; He, Y.; Tian, W.; Lin, J.; Lin, Z. Nanoscale Covalent Organic Frameworks with Donor−Acceptor Structures as Highly Efficient Light-Responsive Oxidase-like Mimics for Colorimetric Detection of Glutathione. ACS Appl. Mater. Interfaces 2021, 13, 49482–49489. [Google Scholar] [CrossRef]
  28. Xu, M.; Chen, K.; Zhu, L.; Zhang, S.; Wang, M.; He, L.; Zhang, Z.; Du, M. MOF@COF Heterostructure Hybrid for Dual-Mode Photoelectrochemical−Electrochemical HIV-1 DNA Sensing. Langmuir 2021, 37, 13479–13492. [Google Scholar] [CrossRef]
  29. Cui, W.-R.; Zhang, C.-R.; Jiang, W.; Liang, R.-P.; Wen, S.-H.; Peng, D.; Qiu, J.-D. Covalent Organic Framework Nanosheet-Based Ultrasensitive and Selective Colorimetric Sensor for Trace Hg2+ Detection. ACS Sustain. Chem. Eng. 2019, 7, 9408–9415. [Google Scholar] [CrossRef]
  30. Côtè, 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] [Green Version]
  31. Furukawa, H.; Yaghi, O.M. Storage of Hydrogen, Methane, and Carbon Dioxide in Highly Porous Covalent Organic Frameworks for Clean Energy Applications. J. Am. Chem. Soc. 2009, 131, 8875–8883. [Google Scholar] [CrossRef] [PubMed]
  32. Tao, S.; Jiang, D. Covalent Organic Frameworks for Energy Conversions: Current Status, Challenges, and Perspectives. CCS Chem. 2021, 3, 2003–2024. [Google Scholar] [CrossRef]
  33. Wang, S.; Wang, Q.; Shao, P.; Han, Y.; Gao, X.; Ma, L.; Yuan, S.; Ma, X.; Zhou, J.; Feng, X.; et al. Exfoliation of Covalent Organic Frameworks into Few-Layer Redox-Active Nanosheets as Cathode Materials for Lithium-Ion Batteries. J. Am. Chem. Soc. 2017, 139, 4258–4261. [Google Scholar] [CrossRef] [PubMed]
  34. Han, S.S.; Furukawa, H.; Yaghi, O.M.; Goddard, W.A., III. Covalent Organic Frameworks as Exceptional Hydrogen Storage Materials. J. Am. Chem. Soc. 2008, 130, 11580–11581. [Google Scholar] [CrossRef] [Green Version]
  35. Wang, Z.; Zhang, S.; Chen, Y.; Zhang, Z.; Ma, S. Covalent Organic Frameworks for Separation Applications. Chem. Soc. Rev. 2020, 49, 708–735. [Google Scholar] [CrossRef] [PubMed]
  36. Yuan, S.; Li, X.; Zhu, J.; Zhang, G.; Van Puyvelde, P.; Van der Bruggen, B. Covalent Organic Frameworks for Membrane Separation. Chem. Soc. Rev. 2019, 48, 2665–2681. [Google Scholar] [CrossRef] [PubMed]
  37. Liang, Z.; Wang, H.-Y.; Zheng, H.; Zhang, W.; Cao, R. Porphyrin-based Frameworks for Oxygen Electrocatalysis and Catalytic Reduction of Carbon Dioxide. Chem. Soc. Rev. 2021, 50, 2540–2581. [Google Scholar] [CrossRef] [PubMed]
  38. 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]
  39. Liu, J.G.; Wang, N.; Ma, L.L. Recent Advances in Covalent Organic Frameworks for Catalysis. Chem-Asian J. 2020, 15, 338–351. [Google Scholar] [CrossRef]
  40. Yusran, Y.; Li, H.; Guan, X.; Fang, Q.; Qiu, S. Covalent Organic Frameworks for Catalysis. EnergyChem 2020, 2, 100035. [Google Scholar] [CrossRef]
  41. Ren, X.; Li, C.; Kang, W.; Li, H.; Ta, N.; Ye, S.; Hu, L.; Wang, X.; Li, C.; Yang, Q. Enormous Promotion of Photocatalytic Activity through the Use of Near-Single Layer Covalent Organic Frameworks. CCS Chem. 2021, 3, 2453–2463. [Google Scholar] [CrossRef]
  42. Ding, S.-Y.; Dong, M.; Wang, Y.-W.; Chen, Y.-T.; Wang, H.-Z.; Su, C.-Y.; Wang, W. Thioether-Based Fluorescent Covalent Organic Framework for Selective Detection and Facile Removal of Mercury(II). J. Am. Chem. Soc. 2016, 138, 3031–3037. [Google Scholar] [CrossRef]
  43. Wang, J.; Zhuang, S. Covalent Organic Frameworks (COFs) for Environmental Applications. Coord. Chem. Rev. 2019, 400, 213046. [Google Scholar] [CrossRef]
  44. Li, C.; Liu, L.; Kang, J.; Xiao, Y.; Feng, Y.; Cao, F.-F.; Zhang, H. Pristine MOF and COF Materials for Advanced Batteries. Energy Stor. Mater. 2020, 31, 115–134. [Google Scholar] [CrossRef]
  45. Feng, L.; Qian, C.; Zhao, Y. Recent Advances in Covalent Organic Framework-Based Nanosystems for Bioimaging and Therapeutic Applications. ACS Mater. Lett. 2020, 2, 1074–1092. [Google Scholar] [CrossRef]
  46. Bhambri, H.; Khullar, S.; Sakshi; Mandal, S.K. Nitrogen-rich Covalent Organic Frameworks: A Promising Class of Sensory Materials. Mater. Adv. 2022, 3, 19–124. [Google Scholar] [CrossRef]
  47. Ascherl, L.; Evans, E.W.; Gorman, J.; Orsborne, S.; Bessinger, D.; Bein, T.; Friend, R.H.; Auras, F. Perylene-Based Covalent Organic Frameworks for Acid Vapor Sensing. J. Am. Chem. Soc. 2019, 141, 15693–15699. [Google Scholar] [CrossRef]
  48. EL-Mahdy, A.F.M.; Laia, M.-Y.; Kuo, S.-W. A Highly Fluorescent Covalent Organic Framework as A Hydrogen Chloride Sensor: Roles of Schiff Base Bonding and π-Stacking. J. Mater. Chem. C 2020, 8, 9520–9528. [Google Scholar] [CrossRef]
  49. 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]
  50. Xiong, Y.; Su, L.; He, X.; Duan, Z.; Zhang, Z.; Chen, Z.; Xie, W.; Zhu, D.; Luo, Y. Colorimetric Determination of Copper Ions Based on Regulation of the Enzyme-Mimicking Activity of Covalent Triazine Frameworks. Sens. Actuators B 2017, 253, 384–391. [Google Scholar] [CrossRef]
  51. Skorjanc, T.; Shetty, D.; Valant, M. Covalent Organic Polymers and Frameworks for Fluorescence-Based Sensors. ACS Sens. 2021, 6, 1461–1481. [Google Scholar] [CrossRef]
  52. 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]
  53. Ascherl, L.; Evans, E.W.; Hennemann, M.; Nuzzo, D.D.; Hufnagel, A.G.; Beetz, M.; Friend, R.H.; Clark, T.; Bein, T.; Auras, F. Solvatochromic Covalent Organic Frameworks. Nat. Commun. 2018, 9, 3802. [Google Scholar] [CrossRef] [Green Version]
  54. Huang, W.; Jiang, Y.; Li, X.; Li, X.; Wang, J.; Wu, Q.; Liu, X. Solvothermal Synthesis of Microporous, Crystalline Covalent Organic Framework Nanofibers and Their Colorimetric Nanohybrid Structures. ACS Appl. Mater. Interfaces 2013, 5, 8845–8849. [Google Scholar] [CrossRef]
  55. Jhulki, S.; Evans, A.M.; Hao, X.L.; Cooper, M.W.; Feriante, C.H.; Leisen, J.; Li, H.; Lam, D.; Hersam, M.C.; Barlow, S.; et al. Humidity Sensing through Reversible Isomerization of a Covalent Organic Framework. J. Am. Chem. Soc. 2020, 142, 783–791. [Google Scholar] [CrossRef]
  56. Singh, H.; Tomer, V.K.; Jena, N.; Bala, I.; Sharma, N.; Nepak, D.; Sarkar, A.D.; Kailasam, K.; Pal, S.K. A Porous, Crystalline Truxene-Based Covalent Organic Framework and Its Application in Humidity Sensing. J. Mater. Chem. A 2017, 5, 21820–21827. [Google Scholar] [CrossRef]
  57. Kulkarni, R.; Noda, Y.; Barange, D.K.; Kochergin, Y.S.; Lyu, P.; Balcarova, B.; Nachtigall, P.; Bojdys, M.J. Real-Time Optical and Electronic Sensing with a β-Amino Enone Linked, Triazine-Containing 2D Covalent Organic Framework. Nat. Commun. 2019, 10, 3228. [Google Scholar] [CrossRef] [Green Version]
  58. Yuan, H.; Li, N.; Linghu, J.; Dong, J.; Wang, Y.; Karmakar, A.; Yuan, J.; Li, M.; Buenconsejo, P.J.S.; Liu, G.; et al. Chip-Level Integration of Covalent Organic Frameworks for Trace Benzene Sensing. ACS Sens. 2020, 5, 1474–1481. [Google Scholar] [CrossRef]
  59. Doonan, C.J.; Tranchemontagne, D.J.; Glover, T.G.; Hunt, J.R.; Yaghi, O.M. Exceptional Ammonia Uptake by a Covalent Organic Framework. Nat. Chem. 2010, 2, 235–238. [Google Scholar] [CrossRef] [PubMed]
  60. Dalapati, S.; Jin, E.; Addicoat, M.; Heine, T.; Jiang, D. Highly Emissive Covalent Organic Frameworks. J. Am. Chem. Soc. 2016, 138, 5797–5800. [Google Scholar] [CrossRef] [PubMed]
  61. Sharma, N.; Sharma, N.; Srinivasan, P.; Kumar, S.; Rayappan, J.B.B.; Kailasam, K. Heptazine Based Organic Framework as a Chemiresistive Sensor for Ammonia Detection at Room Temperature. J. Mater. Chem. A 2018, 6, 18389–18395. [Google Scholar] [CrossRef]
  62. Xu, N.; Wang, R.-L.; Li, D.-P.; Zhou, Z.-Y.; Zhang, T.; Xie, Y.-Z.; Su, Z.-M. Continuous Detection of HCl and NH3 Gases with a High-Performance Fluorescent Polymer Sensor. New J. Chem. 2018, 42, 13367–13374. [Google Scholar] [CrossRef]
  63. Yang, Y.; Zhao, Z.; Yan, Y.; Li, G.; Hao, C. A Mechanism of the Luminescent Covalent Organic Framework for the Detection of NH3. New J. Chem. 2019, 43, 9274–9279. [Google Scholar] [CrossRef]
  64. Meng, Z.; Stolz, R.M.; Mirica, K.A. Two-Dimensional Chemiresistive Covalent Organic Framework with High Intrinsic Conductivity. J. Am. Chem. Soc. 2019, 141, 11929–11937. [Google Scholar] [CrossRef]
  65. Niu, F.; Shao, Z.-W.; Zhu, J.-L.; Tao, L.-M.; Ding, Y. Structural Evolution of Imine-linked Covalent Organic Frameworks and Their NH3 Sensing Performance. J. Mater. Chem. C 2021, 9, 8562–8569. [Google Scholar] [CrossRef]
  66. Cui, F.-Z.; Xie, J.-J.; Jiang, S.-Y.; Gan, S.-X.; Ma, D.-L.; Liang, R.-R.; Jiang, G.-F.; Zhao, X. A Gaseous Hydrogen Chloride Chemosensor Based on a 2D Covalent Organic Framework. Chem. Commun. 2019, 55, 4550–4553. [Google Scholar] [CrossRef]
  67. Gilmanova, L.; Bon, V.; Shupletsov, L.; Pohl, D.; Rauche, M.; Brunner, E.; Kaskel, S. Chemically Stable Carbazole-Based Imine Covalent Organic Frameworks with Acidochromic Response for Humidity Control Applications. J. Am. Chem. Soc. 2021, 143, 18368–18373. [Google Scholar] [CrossRef]
  68. Ko, W.C.; Kim, M.-S.; Kwon, Y.J.; Jeong, J.; Kim, W.R.; Choi, H.; Park, J.K.; Jeong, Y.K. Two-Dimensional Semiconducting Covalent Organic Nanosheets for Highly Sensitive and Stable NO2 Sensing Under Humid Conditions. J. Mater. Chem. A 2020, 8, 19246–19253. [Google Scholar] [CrossRef]
  69. Yang, K.; Yuan, W.; Hua, Z.; Tang, Y.; Yin, F.; Xia, D. Triazine-Based Two-Dimensional Organic Polymer for Selective NO2 Sensing with Excellent Performance. ACS Appl. Mater. Interfaces 2020, 12, 3919–3927. [Google Scholar] [CrossRef]
  70. Yue, Y.; Cai, P.; Xu, X.; Li, H.; Chen, H.; Zhou, H.-C.; Huang, N. Conductive Metallophthalocyanine Framework Films with High Carrier Mobility as Efficient Chemiresistors. Angew. Chem. Int. Ed. 2021, 60, 10806–10813. [Google Scholar] [CrossRef]
  71. Guo, L.; Wang, M.; Zeng, X.; Cao, D. Luminescent Porous Organic Polymer Nanotubes for Highly Selective Sensing of H2S. Mater. Chem. Front. 2017, 1, 2643–2650. [Google Scholar]
  72. Yan, D.; Wang, Z.; Cheng, P.; Chen, Y.; Zhang, Z. Rational Fabrication of Crystalline Smart Materials for Rapid Detection and Efficient Removal of Ozone. Angew. Chem. Int. Ed. 2021, 60, 6055–6060. [Google Scholar] [CrossRef] [PubMed]
  73. Cui, W.-R.; Jiang, W.; Zhang, C.-R.; Liang, R.-P.; Liu, J.; Qiu, J.-D. Regenerable Carbohydrazide-Linked Fluorescent Covalent Organic Frameworks for Ultrasensitive Detection and Removal of Mercury. ACS Sustain. Chem. Eng. 2020, 8, 445–451. [Google Scholar] [CrossRef]
  74. Yin, Y.; Liu, G. Application of Synthesized 2,5-bis(allyloxy)terephthalohydrazide Functionalized Covalent Organic Framework Material as a Fluorescence Probe for Selective Detection of Mercury (II). Mater. Today Commun. 2021, 27, 102440. [Google Scholar] [CrossRef]
  75. Li, Z.; Zhang, Y.; Xi, H.; Mu, Y.; Liu, X. A Robust and Luminescent Covalent Organic Framework as A Highly Sensitive and Selective Sensor for the Detection of Cu2+ Ions. Chem. Commun. 2016, 52, 6613–6616. [Google Scholar] [CrossRef]
  76. Cui, Y.; Liu, Y.; Liu, J.; Du, J.; Yu, Y.; Wang, S.; Liang, Z.; Yu, J. Multifunctional Porous Tröger’s Base Polymers with TetraPhenylethene Units: CO2 Adsorption, Luminescence and Sensing Properties. Polym. Chem. 2017, 8, 4842–4848. [Google Scholar] [CrossRef]
  77. Jin, E.; Li, J.; Geng, K.; Jiang, Q.; Xu, H.; Xu, Q.; Jiang, D. Designed Synthesis of Stable Light-Emitting Two-Dimensional sp2 Carbon-Conjugated Covalent Organic Frameworks. Nat. Commun. 2018, 9, 4143. [Google Scholar] [CrossRef] [Green Version]
  78. Cai, Y.; Jiang, Y.; Feng, L.; Hua, Y.; Liu, H.; Fan, C.; Yin, M.; Li, S.; Lv, X.; Wang, H. Q-graphene-scaffolded Covalent Organic Frameworks as Fluorescent Probes and Sorbents for the Fluorimetry and Removal of Copper Ions. Anal. Chim. Acta 2019, 1057, 88–97. [Google Scholar] [CrossRef]
  79. Wang, T.; Xue, R.; Chen, H.; Shi, P.; Lei, X.; Wei, Y.; Guo, H.; Yang, W. Preparation of Two New Polyimide Bond Linked Porous Covalent Organic Frameworks and Their Fluorescence Sensing Application for Sensitive and Selective Determination of Fe3+. New J. Chem. 2017, 41, 14272–14278. [Google Scholar] [CrossRef]
  80. Li, M.; Cui, Z.; Pang, S.; Meng, L.; Ma, D.; Li, Y.; Shi, Z.; Feng, S. Luminescent Covalent Organic Framework as a Recyclable Turn-off Fluorescent Sensor for Cations and Anions in Aqueous Solution. J. Mater. Chem. C 2019, 7, 11919–11925. [Google Scholar] [CrossRef]
  81. Wu, X.; Wang, B.; Yang, Z.; Chen, L. Novel Imine-linked Covalent Organic Frameworks: Preparation, Characterization and Application. J. Mater. Chem. A 2019, 7, 5650–5655. [Google Scholar] [CrossRef]
  82. Chen, G.; Lan, H.-H.; Cai, S.-L.; Sun, B.; Li, X.-L.; He, Z.-H.; Zheng, S.-R.; Fan, J.; Liu, Y.; Zhang, W.-G. Stable Hydrazone-Linked Covalent Organic Frameworks Containing O,N,O′-Chelating Sites for Fe(III) Detection in Water. ACS Appl. Mater. Interfaces 2019, 11, 12830–12837. [Google Scholar] [CrossRef] [Green Version]
  83. Cui, D.; Ding, X.; Xie, W.; Xu, G.; Su, Z.; Xu, Y.; Xie, Y. A Tetraphenylethylene-based Covalent Organic Framework for Waste Gas Adsorption and Highly Selective Detection of Fe3+. CrystEngComm 2021, 23, 5569–5574. [Google Scholar] [CrossRef]
  84. Li, D.-M.; Zhang, S.-Y.; Wan, J.-Y.; Zhang, W.-Q.; Yan, Y.-L.; Tang, X.-H.; Zheng, S.-R.; Cai, S.-L.; Zhang, W.-G. A New Hydrazone-linked Covalent Organic Framework for Fe(III) Detection by Fluorescence and QCM Technologies. CrystEngComm 2021, 23, 3594–3601. [Google Scholar] [CrossRef]
  85. Liang, X.; Ni, Z.; Zhao, L.; Ge, B.; Zhao, H.; Li, W. Multifunctional Triphenylbenzene-based Polyimide Covalent Organic Framework with Absolute Eclipsed Stacking Models for Fluorescence Detecting of Fe3+ and Electrochemical Detecting of Pb2+. Microchem. J. 2021, 170, 106663. [Google Scholar] [CrossRef]
  86. Zhang, T.; Gao, C.; Huang, W.; Chen, Y.; Wang, Y.; Wang, J. Covalent Organic Framework as A Novel Electrochemical Platform for Highly Sensitive and Stable Detection of Lead. Talanta 2018, 188, 578–583. [Google Scholar] [CrossRef]
  87. Wang, R.; Ji, W.; Huang, L.; Guo, L.; Wang, X. Electrochemical Determination of Lead(II) in Environmental Waters Using a Sulfydryl Modified Covalent Organic Framework by Square Wave Anodic Stripping Voltammetry (SWASV). Anal. Lett. 2019, 52, 1757–1770. [Google Scholar] [CrossRef]
  88. Zhao, C.; Zhang, L.; Wang, Q.; Zhang, L.; Zhu, P.; Yu, J.; Zhang, Y. Porphyrin-Based Covalent Organic Framework Thin Films as Cathodic Materials for “On−Off−On” Photoelectrochemical Sensing of Lead Ions. ACS Appl. Mater. Interfaces 2021, 13, 20397–20404. [Google Scholar] [CrossRef]
  89. Zhou, Z.; Zhong, W.; Cui, K.; Zhuang, Z.; Li, L.; Li, L.; Bi, J.; Yu, Y. A Covalent Organic Framework Bearing Thioether Pendant Arms for Selective Detection and Recovery of Au from Ultra-low Concentration Aqueous Solution. Chem. Commun. 2018, 54, 9977–9980. [Google Scholar] [CrossRef]
  90. Cui, W.-R.; Zhang, C.-R.; Jiang, W.; Li, F.-F.; Liang, R.-P.; Liu, J.; Qiu, J.-D. Regenerable and Stable sp2 Carbon-conjugated Covalent Organic Frameworks for Selective Detection and Extraction of Uranium. Nat. Commun. 2020, 11, 436. [Google Scholar] [CrossRef] [Green Version]
  91. Zhang, L.; Yang, G.-P.; Xiao, S.-J.; Tan, Q.-G.; Zheng, Q.-Q.; Liang, R.-P.; Qiu, J.-D. Facile Construction of Covalent Organic Framework Nanozyme for Colorimetric Detection of Uranium. Small 2021, 17, 2102944. [Google Scholar] [CrossRef]
  92. Chen, Y.; Sun, R.; Zhu, W.; Zhang, Z.; Chen, Y.; Wang, S.; Deng, Q. Desirability of Position 2, 2′-bipyridine Group into COFs for the Fluorescence Sensing of Ni(II). Sens. Actuators B Chem. 2021, 344, 130216. [Google Scholar] [CrossRef]
  93. Zhang, S.; Chen, K.; Zhu, L.; Xu, M.; Song, Y.; Zhang, Z.; Du, M. Direct Growth of Two-dimensional Phthalocyanine-based COF on Cu-MOF to Construct a Photoelectrochemical-Electrochemical Dual-mode Biosensing Platform for High-efficiency Determination of Cr(III). Dalton Trans. 2021, 50, 14285–14295. [Google Scholar] [CrossRef]
  94. Lu, Y.; Liang, Y.; Zhao, Y.; Xia, M.; Liu, X.; Shen, T.; Feng, L.; Yuan, N.; Chen, Q. Fluorescent Test Paper via the In Situ Growth of COFs for Rapid and Convenient Detection of Pd(II) Ions. ACS Appl. Mater. Interfaces 2021, 13, 1644–1650. [Google Scholar] [CrossRef]
  95. Yue, J.-Y.; Ding, X.-L.; Wang, Y.-T.; Wen, Y.-X.; Yang, P.; Ma, Y.; Tang, B. Dual Functional sp2 Carbon-conjugated Covalent Organic Frameworks for Fluorescence Sensing and Effective Removal and Recovery of Pd2+ Ions. J. Mater. Chem. A 2021, 9, 26861–26866. [Google Scholar] [CrossRef]
  96. Wang, Z.; Meng, Q.; Ma, R.; Wang, Z.; Yang, Y.; Sha, H.; Ma, X.; Ruan, X.; Zou, X.; Yuan, Y.; et al. Constructing an Ion Pathway for Uranium Extraction from Seawater. Chem 2020, 6, 1683–1691. [Google Scholar] [CrossRef]
  97. Yuan, Y.; Yang, Y.; Ma, X.; Meng, Q.; Wang, L.; Zhao, S.; Zhu, G. Molecularly Imprinted Porous Aromatic Frameworks and Their Composite Components for Selective Extraction of Uranium Ions. Adv. Mater. 2018, 30, 1706507. [Google Scholar] [CrossRef]
  98. Li, Z.; Meng, Q.; Yang, Y.; Zou, X.; Yuan, Y.; Zhu, G. Constructing Amidoxime-modified Porous Adsorbents with Open Architecture for Cost-effective and Efficient Uranium Extraction. Chem. Sci. 2020, 11, 4747–4752. [Google Scholar] [CrossRef] [Green Version]
  99. Cui, W.-R.; Li, F.; Xu, R.-H.; Zhang, C.-R.; Chen, X.-R.; Yan, R.-H.; Liang, R.-P.; Qiu, J.-D. Regenerable Covalent Organic Frameworks for Photo-enhanced Uranium Adsorption from Seawater. Angew. Chem. Int. Ed. 2020, 59, 17684–17690. [Google Scholar] [CrossRef] [PubMed]
  100. Zhang, Y.; Shen, X.; Feng, X.; Xia, H.; Mu, Y.; Liu, X. Covalent Organic Frameworks as pH Responsive Signaling Scaffolds. Chem. Commun. 2016, 52, 11088–11091. [Google Scholar] [CrossRef]
  101. Chen, L.; He, L.; Ma, F.; Liu, W.; Wang, Y.; Silver, M.A.; Chen, L.; Zhu, L.; Gui, D.; Diwu, J.; et al. Covalent Organic Framework Functionalized with 8-Hydroxyquinoline as a Dual-Mode Fluorescent and Colorimetric pH Sensor. ACS Appl. Mater. Interfaces 2018, 10, 15364–15368. [Google Scholar] [CrossRef] [PubMed]
  102. Xu, M.; Wang, L.; Xie, Y.; Song, Y.; Wang, L. Ratiometric Electrochemical Sensing and Biosensing Based on Multiple Redox-Active State COFDHTA-TTA. Sens. Actuators B Chem. 2019, 281, 1009–1015. [Google Scholar] [CrossRef]
  103. Yuan, F.; Kong, Y.; You, J.; Zhang, C.; Xian, Y. Rational Synthesis of Imine-Linked Fluorescent Covalent Organic Frameworks with Different pKa for pH Sensing In Vitro and In Vivo. ACS Appl. Mater. Interfaces 2021, 13, 51351–51361. [Google Scholar] [CrossRef] [PubMed]
  104. Chen, Z.; Wang, K.; Hu, X.; Shi, P.; Guo, Z.; Zhan, H. Novel One-Dimensional Covalent Organic Framework as a H+ Fluorescent Sensor in Acidic Aqueous Solution. ACS Appl. Mater. Interfaces 2021, 13, 1145–1151. [Google Scholar] [CrossRef]
  105. Li, Z.; Li, H.; Xia, H.; Ding, X.; Luo, X.; Liu, X.; Mu, Y. Triarylboron-Linked Conjugated Microporous Polymers: Sensing and Removal of Fluoride Ions. Chem. Eur. J. 2015, 21, 17355–17362. [Google Scholar] [CrossRef]
  106. Li, Z.; Huang, N.; Lee, K.H.; Feng, Y.; Tao, S.; Jiang, Q.; Nagao, Y.; Irle, S.; Jiang, D. Light-Emitting Covalent Organic Frameworks: Fluorescence Improving via Pinpoint Surgery and Selective Switch-On Sensing of Anions. J. Am. Chem. Soc. 2018, 140, 12374–12377. [Google Scholar] [CrossRef]
  107. Su, L.; Zhang, Z.; Xiong, Y. Water Dispersed Two-dimensional Ultrathin Fe(III)-Modified Covalent Triazine Framework Nanosheets: Peroxidase Like Activity and Colorimetric Biosensing Applications. Nanoscale 2018, 10, 20120–20125. [Google Scholar] [CrossRef]
  108. Wang, P.; Zhou, F.; Zhang, C.; Yin, S.-Y.; Teng, L.; Chen, L.; Hu, X.-X.; Liu, H.-W.; Yin, X.; Zhang, X.-B. Ultrathin Two-dimensional Covalent Organic Framework Nanoprobe for Interference-Resistant Two-Photon Fluorescence Bioimaging. Chem. Sci. 2018, 9, 8402–8408. [Google Scholar] [CrossRef] [Green Version]
  109. Paul, B.D.; Snyder, S.H. H2S Signalling Through Protein Sulfhydration and Beyond. Nat. Rev. Mol. Cell Biol. 2012, 13, 499–507. [Google Scholar] [CrossRef]
  110. Gadalla, M.M.; Snyder, S.H. Hydrogen sulfide as a gasotransmitter. J. Neurochem. 2010, 113, 14–26. [Google Scholar] [CrossRef] [Green Version]
  111. Zhang, W.; Qiu, L.-G.; Yuan, Y.-P.; Xie, A.-J.; Shen, Y.-H.; Zhu, J.-F. Microwave-assisted Synthesis of Highly Fluorescent Nanoparticles of a Melamine-based Porous Covalent Organic Framework for Trace-level Detection of Nitroaromatic Explosives. J. Hazard. Matter. 2012, 221–222, 147–154. [Google Scholar] [CrossRef]
  112. Sang, N.; Zhan, C.; Cao, D. Highly Sensitive and Selective Detection of 2,4,6-trinitrophenol Using Covalent-Organic Polymer Luminescent Probes. J. Mater. Chem. A 2015, 3, 92–96. [Google Scholar] [CrossRef]
  113. Guo, L.; Cao, D.; Yun, J.; Zeng, X. Highly Selective Detection of Picric Acid from Multicomponent Mixtures of Nitro Explosives by Using COP Luminescent Probe. Sens. Actuators B Chem. 2017, 243, 753–760. [Google Scholar] [CrossRef]
  114. Jiang, S.; Meng, L.; Ma, W.; Qia, Q.; Zhang, W.; Xua, B.; Liu, L.; Tian, W. Morphology Controllable Conjugated Network Polymers Based on AIE-active Building Block for TNP Detection. Chin. Chem. Lett. 2021, 32, 1037–1040. [Google Scholar] [CrossRef]
  115. Rao, M.R.; Fang, Y.; Feyter, S.D.; Perepichka, D.F. Conjugated Covalent Organic Frameworks via Michael Addition-Elimination. J. Am. Chem. Soc. 2017, 139, 2421–2427. [Google Scholar] [CrossRef]
  116. Dalapati, S.; Jin, S.; Gao, J.; Xu, Y.; Nagai, A.; Jiang, D. An Azine-Linked Covalent Organic Framework. J. Am. Chem. Soc. 2013, 135, 17310–17313. [Google Scholar] [CrossRef]
  117. Xiang, Z.; Cao, D. Synthesis of Luminescent Covalent–Organic Polymers for Detecting Nitroaromatic Explosives and Small Organic Molecules. Macromol. Rapid Commun. 2012, 33, 1184–1190. [Google Scholar] [CrossRef]
  118. Kaleeswaran, D.; Vishnoi, P.; Murugavel, R. [3+3] Imine and β-ketoenamine Tethered Fluorescent Covalent-Organic FrameWorks for CO2 Uptake and Nitroaromatic Sensing. J. Mater. Chem. C 2015, 3, 7159–7171. [Google Scholar] [CrossRef]
  119. Das, G.; Biswal, B.P.; Kandambeth, S.; Venkatesh, V.; Kaur, G.; Addicoat, M.; Heine, T.; Verma, S.; Banerjee, R. Chemical Sensing in Two Dimensional Porous Covalent Organic Nanosheets. Chem. Sci. 2015, 6, 3931–3939. [Google Scholar] [CrossRef]
  120. Gomes, R.; Bhaumik, A. A New Triazine Functionalized Luminescent Covalent Organic Framework for Nitroaromatic Sensing and CO2 Storage. RSC Adv. 2016, 6, 28047–28054. [Google Scholar] [CrossRef]
  121. Lin, G.; Ding, H.; Yuan, D.; Wang, B.; Wang, C. A Pyrene-Based, Fluorescent Three-Dimensional Covalent Organic Framework. J. Am. Chem. Soc. 2016, 138, 3302–3305. [Google Scholar] [CrossRef] [PubMed]
  122. Zhang, C.; Zhang, S.; Yan, Y.; Xia, F.; Huang, A.; Xian, Y. Highly Fluorescent Polyimide Covalent Organic Nanosheets as Sensing Probes for the Detection of 2,4,6-Trinitrophenol. ACS Appl. Mater. Interfaces 2017, 9, 13415–13421. [Google Scholar] [CrossRef] [PubMed]
  123. Kaleeswaran, D.; Murugavel, R. Picric Acid Sensing and CO2 Capture by a Sterically Encumbered Azo-linked Fluorescent Triphenylbenzene Based Covalent Organic Polymer. J. Chem. Sci. 2018, 130, 1. [Google Scholar] [CrossRef] [Green Version]
  124. Das, P.; Mandal, S.K. A Dual-functionalized, Luminescent and Highly Crystalline Covalent Organic Framework: Molecular Decoding Strategies for VOCs and Ultrafast TNP Sensing. J. Mater. Chem. A 2018, 6, 16246–16256. [Google Scholar] [CrossRef]
  125. Wang, S.; Liu, Y.; Yu, Y.; Du, J.; Cui, Y.; Song, X.; Liang, Z. Conjugated Microporous Polymers Based on Biphenylene for CO2 Adsorption and Luminescence Detection of Nitroaromatic Compounds. New J. Chem. 2018, 42, 9482–9487. [Google Scholar] [CrossRef]
  126. Gao, Q.; Li, X.; Ning, G.-H.; Leng, K.; Tian, B.; Liu, C.; Tang, W.; Xu, H.-S.; Loh, K.P. Highly Photoluminescent Two-dimensional Iminebased Covalent Organic Frameworks for Chemical Sensing. Chem. Commun. 2018, 54, 2349–2352. [Google Scholar] [CrossRef]
  127. Faheem, M.; Aziz, S.; Jing, X.; Ma, T.; Du, J.; Sun, F.; Tian, Y.; Zhu, G. Dual Luminescent Covalent Organic Frameworks for Nitro-explosive Detection. J. Mater. Chem. A 2019, 7, 27148–27155. [Google Scholar] [CrossRef]
  128. Pan, L.; Liu, Z.; Tian, M.; Schroeder, B.C.; Aliev, A.E.; Faul, C.F.J. Luminescent and Swellable Conjugated Microporous Polymers for Detecting Nitroaromatic Explosives and Removing Harmful Organic Vapors. ACS Appl. Mater. Interfaces 2019, 11, 48352–48362. [Google Scholar] [CrossRef]
  129. Das, P.; Chakraborty, G.; Mandal, S.K. Comprehensive Structural and Microscopic Characterization of an Azine−Triazine-Functionalized Highly Crystalline Covalent Organic Framework and Its Selective Detection of Dichloran and 4-Nitroaniline. ACS Appl. Mater. Interfaces 2020, 12, 10224–10232. [Google Scholar] [CrossRef]
  130. Chen, J.-Q.; Zheng, Q.-Q.; Xiao, S.-J.; Zhang, L.; Liang, R.-P.; Ouyang, G.; Qiu, J.-D. Construction of Two-Dimensional Fluorescent Covalent Organic Framework Nanosheets for the Detection and Removal of Nitrophenols. Anal. Chem. 2022, 94, 2517–2526. [Google Scholar] [CrossRef]
  131. Zhao, Y.; Liu, X.; Li, Y.; Xia, M.; Xia, T.; Sun, H.; Sui, Z.; Hu, X.-M.; Chen, Q. Ultra-Stable Fluorescent 2D Covalent Organic Framework for Rapid Adsorption and Selective Detection of Radioiodine. Microporous Mesoporous Mater. 2021, 319, 111046. [Google Scholar] [CrossRef]
  132. Wang, P.; Xu, Q.; Li, Z.; Jiang, W.; Jiang, Q.; Jiang, D. Exceptional Iodine Capture in 2D Covalent Organic Frameworks. Adv. Mater. 2018, 30, 1801991. [Google Scholar] [CrossRef]
  133. Geng, T.; Zhang, W.; Zhu, Z.; Kai, X. Triazine-based conjugated microporous polymers constructing triphenylamine and its derivatives with nitrogen as core for iodine adsorption and flfluorescence sensing I2. Microporous Mesoporous Mater. 2019, 273, 163–170. [Google Scholar] [CrossRef]
  134. Geng, T.; Ma, L.; Chen, G.; Zhang, C.; Zhang, W.; Xia, H.; Zhu, H. Poly[1,3,6,8-tetra(2-thiophenyl)pyrene] and poly[1,3,6,8-tetra(3-thiophenyl)pyrene] conjugated microporous polymers for reversible adsorbing and fluorescent sensing iodine. J. Polym. Res. 2019, 26, 113. [Google Scholar] [CrossRef]
  135. Geng, T.; Liu, M.; Zhang, C.; Hu, C.; Xu, H. Synthesis of secondary amine-based fluorescent porous organic polymers via Friedel–Crafts polymerization reaction for adsorbing and fluorescent sensing iodine. J. Appl. Polym. Sci. 2020, 137, e49255. [Google Scholar] [CrossRef]
  136. Geng, T.; Zhang, C.; Chen, G.; Ma, L.; Zhang, W.; Xia, H. Synthesis of tetraphenylethylene-based fluorescent conjugated microporous polymers for fluorescent sensing and adsorbing iodine. Microporous Mesoporous Mater. 2019, 284, 468–475. [Google Scholar] [CrossRef]
  137. Geng, T.; Chen, G.; Ma, L.; Zhang, C.; Zhang, W.; Xu, H. The spirobifluorene-based fluorescent conjugated microporous polymers for reversible adsorbing iodine, fluorescent sensing iodine and nitroaromatic compounds. Eur. Polym. J. 2019, 115, 37–44. [Google Scholar] [CrossRef]
  138. Geng, T.; Ma, L.; Chen, G.; Zhang, C.; Zhang, W.; Niu, Q. Fluorescent conjugated microporous polymers containing pyrazine moieties for adsorbing and fluorescent sensing of iodine. Environ. Sci. Pollut. Res. 2020, 27, 20235–20245. [Google Scholar] [CrossRef]
  139. Geng, T.; Zhang, C.; Hu, C.; Liu, M.; Fei, Y.; Xia, H. Synthesis of 1,6-disubstituted pyrene-based conjugated microporous polymers for reversible adsorption and fluorescence sensing of iodine. New J. Chem. 2020, 44, 2312–2320. [Google Scholar] [CrossRef]
  140. Zhang, T.; Chen, Y.; Huang, W.; Wang, Y.; Hu, X. A Novel AuNPs−doped COFs Composite as Electrochemical Probe for Chlorogenic Acid Detection with Enhanced Sensitivity and Stability. Sens. Actuators B Chem. 2018, 276, 362–369. [Google Scholar] [CrossRef]
  141. Wang, Y.; Wang, Y.; Liu, H. A Novel Fluorescence and SPE Adsorption Nanomaterials of Molecularly Imprinted Polymers Based on Quantum Dot-Grafted Covalent Organic Frameworks for the High Selectivity and Sensitivity Detection of Ferulic Acid. Nanomaterials 2019, 9, 305. [Google Scholar] [CrossRef] [Green Version]
  142. Sun, Y.; Xua, L.; Waterhouse, G.I.N.; Wang, M.; Qiao, X.; Xu, Z. Novel Three-dimensional Electrochemical Sensor with Dual Signal Amplification Based on MoS2 Nanosheets and High-conductive NH2MWCNT@COF for Sulfamerazine Determination. Sens. Actuators B Chem. 2019, 281, 107–114. [Google Scholar] [CrossRef]
  143. Zhou, N.; Ma, Y.; Hu, B.; He, L.; Wang, S.; Zhang, Z.; Lu, S. Construction of Ce−MOF@COF Hybrid Nanostructure: Label-free Aptasensor for the Ultrasensitive Detection of Oxytetracycline Residues in Aqueous Solution Environments. Biosens. Bioelectron. 2019, 127, 92–100. [Google Scholar] [CrossRef]
  144. Ma, X.; Pang, C.; Li, S.; Xiong, Y.; Li, J.; Luo, J.; Yang, Y. Synthesis of Zr-coordinated Amide Porphyrin-based Two-dimensional Covalent Organic Framework at Liquid-liquid Interface for Electrochemical Sensing of Tetracycline. Biosens. Bioelectron. 2019, 146, 111734. [Google Scholar] [CrossRef]
  145. Wang, J.-M.; Lian, X.; Yan, B. Eu3+-Functionalized Covalent Organic Framework Hybrid Material as a Sensitive Turn-On Fluorescent Switch for Levofloxacin Monitoring in Serum and Urine. Inorg. Chem. 2019, 58, 9956–9963. [Google Scholar] [CrossRef]
  146. Zhang, D.; Liu, H.; Geng, W.; Wang, Y. A Dual-function Molecularly Imprinted Optopolymer Based on Quantum Dotsgrafted Covalent-Organic Frameworks for the Sensitive Detection of Tyramine in Fermented Meat Products. Food Chem. 2019, 277, 639–645. [Google Scholar] [CrossRef]
  147. Dong, J.; Li, X.; Peh, S.B.; Yuan, Y.D.; Wang, Y.; Ji, D.; Peng, S.; Liu, G.; Ying, S.; Yuan, D.; et al. Restriction of Molecular Rotors in Ultrathin Two-Dimensional Covalent Organic Framework Nanosheets for Sensing Signal Amplification. Chem. Mater. 2019, 31, 146–160. [Google Scholar] [CrossRef]
  148. Gua, Y.; Wang, J.; Shia, H.; Pana, M.; Liu, B.; Fang, G.; Wang, S. Electrochemiluminescence sensor based on upconversion nanoparticles and oligoaniline-crosslinked gold nanoparticles imprinting recognition sites for the determination of dopamine. Biosens. Bioelectron. 2019, 128, 129–136. [Google Scholar] [CrossRef]
  149. Yao, X.; Shen, J.; Liu, Q.; Fa, H.; Yang, M.; Hou, C. A Novel Electrochemical Aptasensor for the Sensitive Detection of Kanamycin Based on UiO-66NH2/MCA/MWCNT@rGONR Nanocomposites. Anal. Methods 2020, 12, 4967–4976. [Google Scholar] [CrossRef]
  150. Wang, J.; Zhao, L.; Yan, B. Indicator Displacement Assay Inside Dye-Functionalized Covalent Organic Frameworks for Ultrasensitive Monitoring of Sialic Acid, an Ovarian Cancer Biomarker. ACS Appl. Mater. Interfaces 2020, 12, 12990–12997. [Google Scholar] [CrossRef]
  151. Han, Z.-Y.; Zhang, H.; Li, H.-K.; Zhu, Q.-Q.; He, H. Ingenious Construction of an Electrochemical Aptasensor based on a Au@COF/GO−NH2 Composite with Excellent Detection Performance. J. Mater. Chem. C 2021, 9, 4576. [Google Scholar] [CrossRef]
  152. Yang, M.; Mao, Y.; Wang, B.; Lin, L.; Wang, Y.; Zhang, L.; Jiang, Y.; Zhao, M.; Chen, H.; Zhang, Y. Heterometallic Mg@Fe−MIL−101/TpPa−1−COF Grown on Stainless Steel Mesh: Enhancing Photo-degradation, Fluorescent Detection and Toxicity Assessment for Tetracycline Hydrochloride. Colloid Surf. A 2021, 631, 127725. [Google Scholar] [CrossRef]
  153. Zhao, N.; Liu, J.; Yang, F.; Lv, S.; Wang, J.; Wang, S. Easy Green Construction of a Universal Sensing Platform Based on Crystalline Polyimide Covalent Organic Frameworks with Sensitive Fluorescence Response to Metal Ions and Antibiotics. ACS Appl. Bio Mater. 2021, 4, 995–1002. [Google Scholar] [CrossRef]
  154. Wang, J.; Qu, X.; Zhao, L.; Yan, B. Fabricating Nanosheets and Ratiometric Detection of 5-Fluorouracil by Covalent Organic Framework Hybrid Material. Anal. Chem. 2021, 93, 4308–4316. [Google Scholar] [CrossRef]
  155. Song, Y.; Ma, R.; Hao, L.; Yang, X.; Wang, C.; Wu, Q.; Wang, Z. Application of Covalent Organic Framework as the Adsorbent for Solid-phase Extraction of Trace Levels of Pesticide Residues Prior to High-performance Liquid Chromatography-Ultraviolet Detection. J. Chromatogr. A 2018, 1572, 20–26. [Google Scholar] [CrossRef]
  156. Ji, W.-H.; Guo, Y.-S.; Wang, X.; Lu, X.-F.; Guo, D.-S. Amino-modified Covalent Organic Framework as Solid Phase Extraction Absorbent for Determination of Carboxylic Acid Pesticides in Environmental Water Samples. J. Chromatogr. A 2019, 1595, 11–18. [Google Scholar] [CrossRef]
  157. Liu, H.; Zhang, Y.; Zhang, D.; Zheng, F.; Huang, M.; Sun, J.; Sun, X.; Li, H.; Wang, J.; Sun, B. A Fluorescent Nanoprobe for 4-ethylguaiacol Based on the Use of A Molecularly Imprinted Polymer Doped with A Covalent Organic Framework Grafted onto Carbon Nanodots. Microchim. Acta 2019, 186, 182. [Google Scholar] [CrossRef]
  158. Kandambeth, S.; Mallick, A.; Lukose, B.; Mane, M.V.; Heine, T.; Banerjee, R. Construction of Crystalline 2D Covalent Organic Frameworks with Remarkable Chemical (Acid/Base) Stability via a Combined Reversible and Irreversible Route. J. Am. Chem. Soc. 2012, 134, 19524–19527. [Google Scholar] [CrossRef]
  159. Liu, X.; Chen, H.; Wang, R.; Zang, S.-Q.; Mak, T.C.W. Cationic Covalent-Organic Framework as Efficient Redox Motor for High-Performance Lithium–Sulfur Batteries. Small 2020, 16, 2002932. [Google Scholar] [CrossRef]
  160. Wang, J.; Yang, X.; Wei, T.; Bao, J.; Zhu, Q.; Dai, Z. Fe-Porphyrin-Based Covalent Organic Framework As a Novel Peroxidase Mimic for a One-Pot Glucose Colorimetric Assay. ACS Appl. Bio Mater. 2018, 1, 382–388. [Google Scholar] [CrossRef]
  161. Zhao, Y.; Xu, X.; Ma, Y.; Tan, H.; Li, Y. A novel Peroxidase/oxidase Mimetic Fe-porphyrin Covalent Organic Framework Enhanced the Luminol Chemiluminescence Reaction and Its Application in Glucose Sensing. Luminescence 2020, 35, 1366–1372. [Google Scholar] [CrossRef]
  162. Yue, J.-Y.; Ding, X.-L.; Wang, L.; Yang, R.; Bi, J.-S.; Song, Y.-W.; Yang, P.; Ma, Y.; Tang, B. Novel Enzyme-Functionalized Covalent Organic Frameworks for the Colorimetric Sensing of Glucose in Body Fluids and Drinks. Mater. Chem. Front. 2021, 5, 3859–3866. [Google Scholar] [CrossRef]
  163. He, Y.; Lin, X.; Tang, Y.; Ye, L. A Selective Sensing Platform for the Simultaneous Detection of Ascorbic acid, Dopamine, and Uric Acid Based on AuNPs/carboxylated COFs/Poly(fuchsin basic) Film. Anal. Methods 2021, 13, 4503–4514. [Google Scholar] [CrossRef]
  164. Wu, X.; Han, X.; Xu, Q.; Liu, Y.; Yuan, C.; Yang, S.; Liu, Y.; Jiang, J.; Cui, Y. Chiral BINOL-Based Covalent Organic Frameworks for Enantioselective Sensing. J. Am. Chem. Soc. 2019, 141, 7081–7089. [Google Scholar] [CrossRef]
  165. Zhang, T.; Ma, N.; Ali, A.; Wei, Q.; Wu, D.; Ren, X. Electrochemical Ultrasensitive Detection of Cardiac Troponin I Using Covalent Organic Frameworks for Signal Amplification. Biosens. Bioelectron. 2018, 119, 176–181. [Google Scholar] [CrossRef]
  166. Feng, S.; Yan, M.; Xue, Y.; Huang, J.; Yang, X. Electrochemical Immunosensor for Cardiac Troponin I Detection Based on Covalent Organic Framework and Enzyme-Catalyzed Signal Amplification. Anal. Chem. 2021, 93, 13572–13579. [Google Scholar] [CrossRef]
  167. Zhai, R.; Gong, X.Y.; Xie, J.; Yuan, Y.; Xu, F.; Jiang, Y.; Huang, Z.; Dai, X.; Zhang, Y.; Qian, X.; et al. Ultrasensitive Analysis of Heat Shock Protein 90α with Antibodies Orderly Arrayed on a Novel Type of Immunoprobe Based on Magnetic COFs. Talanta 2019, 191, 553–560. [Google Scholar] [CrossRef]
  168. Ma, Y.; Lu, M.; Deng, Y.; Bai, R.; Zhang, X.; Li, D.; Zhang, K.; Hu, R.; Yang, Y. The Preparation of C-Reactive Protein Immunosensor Based on Nano-Mimetic Enzyme Co3O4. J. Biomed. Nanotechnol. 2018, 14, 1169–1177. [Google Scholar] [CrossRef]
  169. Deng, Y.; Du, X.; Ma, Y.; Zhang, K.; Zhang, X.; Li, D.-L.; Bai, R.; Hu, R.; Yang, Y.-H. Development of C-Reactive Protein Immunosensor Using Thionine/Au Nanoparticles-Covalent Organic Framework-LZU8 as Label. Nanosci. Nanotechnol. Lett. 2018, 10, 520–527. [Google Scholar] [CrossRef]
  170. Liu, T.-Z.; Hu, R.; Zhang, X.; Zhang, K.-L.; Liu, Y.; Zhang, X.-B.; Bai, R.-Y.; Li, D.; Yang, Y.-H. Metal–Organic Framework Nanomaterials as Novel Signal Probes for Electron Transfer Mediated Ultrasensitive Electrochemical Immunoassay. Anal. Chem. 2016, 88, 12516–12523. [Google Scholar] [CrossRef]
  171. Li, W.; Yang, C.-X.; Yan, X.-P. A Versatile Covalent Organic Framework-based Platform for Sensing Biomolecules. Chem. Commun. 2017, 53, 11469–11471. [Google Scholar] [CrossRef]
  172. Peng, Y.; Huang, Y.; Zhu, Y.; Chen, B.; Wang, L.; Lai, Z.; Zhang, Z.; Zhao, M.; Tan, C.; Yang, N.; et al. Ultrathin Two-Dimensional Covalent Organic Framework Nanosheets: Preparation and Application in Highly Sensitive and Selective DNA Detection. J. Am. Chem. Soc. 2017, 139, 8698–8704. [Google Scholar] [CrossRef] [Green Version]
  173. Mal, A.; Mishra, R.K.; Praveen, V.K.; Khayum, M.A.; Banerjee, R.; Ajayaghosh, A. Supramolecular Reassembly of Self-Exfoliated Ionic Covalent Organic Nanosheets for Label-Free Detection of Double-Stranded DNA. Angew. Chem. Int. Ed. 2018, 57, 8443–8447. [Google Scholar] [CrossRef] [PubMed]
  174. Zhang, X.; Chi, K.-N.; Li, D.-L.; Deng, Y.; Ma, Y.-C.; Xu, Q.-Q.; Hu, R.; Yang, Y.-H. 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]
Figure 1. The basic topology of the COFs.
Figure 1. The basic topology of the COFs.
Molecules 27 02586 g001
Figure 2. Modulating the pore size of the COFs.
Figure 2. Modulating the pore size of the COFs.
Molecules 27 02586 g002
Figure 3. Introducing the binding site that can selectively capture the analyte. (a) capturing H+ by imine linkage; (b) Intruding thioether binding site for capturing Hg2+; (c) Introducing AuNPs by in situ growth for mercury capture; (d) Externally immobilizing enrofloxacin-targeted aptamer onto the 2D sheet of the COF for capturing enrofloxacin.
Figure 3. Introducing the binding site that can selectively capture the analyte. (a) capturing H+ by imine linkage; (b) Intruding thioether binding site for capturing Hg2+; (c) Introducing AuNPs by in situ growth for mercury capture; (d) Externally immobilizing enrofloxacin-targeted aptamer onto the 2D sheet of the COF for capturing enrofloxacin.
Molecules 27 02586 g003
Figure 4. (a) COF for NH3 sensing based on the protonation and deprotonation of boronate linkage [60]. Copyright 2016 American Chemical Society; (b) COF for HCl sensing is based on the protonation and deprotonation of imine linkage [47]. Copyright 2019 American Chemical Society; (c) COF for HCl, NH3 sensing based on the protonation and deprotonation of imine linkage [57]. Copyright 2019 Nature Publishing Group.
Figure 4. (a) COF for NH3 sensing based on the protonation and deprotonation of boronate linkage [60]. Copyright 2016 American Chemical Society; (b) COF for HCl sensing is based on the protonation and deprotonation of imine linkage [47]. Copyright 2019 American Chemical Society; (c) COF for HCl, NH3 sensing based on the protonation and deprotonation of imine linkage [57]. Copyright 2019 Nature Publishing Group.
Molecules 27 02586 g004
Figure 5. (a) TAPP−DHNDA−COF/aramid hybrid material for humidity detection [54]. Copyright 2013 American Chemical Society; (b) Py–TT COF thin film for humidity detection [53]. Copyright 2018 Nature Publishing Group; (c) Py−TT COF for humidity detection [55]. Copyright 2020 American Chemical Society.
Figure 5. (a) TAPP−DHNDA−COF/aramid hybrid material for humidity detection [54]. Copyright 2013 American Chemical Society; (b) Py–TT COF thin film for humidity detection [53]. Copyright 2018 Nature Publishing Group; (c) Py−TT COF for humidity detection [55]. Copyright 2020 American Chemical Society.
Molecules 27 02586 g005
Figure 6. (a) COF-based IDE sensor for benzene vapor sensing [58]. Copyright 2020 American Chemical Society; (b) COF-based H2S detection [71]. Copyright 2017 Royal Society of Chemistry.
Figure 6. (a) COF-based IDE sensor for benzene vapor sensing [58]. Copyright 2020 American Chemical Society; (b) COF-based H2S detection [71]. Copyright 2017 Royal Society of Chemistry.
Molecules 27 02586 g006
Figure 7. (a) COF−LZU8 for Hg2+ detection via fluorescence quenching [42]. Copyright 2016 American Chemical Society; (b) TFPPy−CHYD for Hg2+ detection via fluorescence quenching [73]. Copyright 2020 American Chemical Society; (c) Tp−Bpy NSs for Hg2+ detection via catalytic oxidation of TMB [29]. Copyright 2019 American Chemical Society.
Figure 7. (a) COF−LZU8 for Hg2+ detection via fluorescence quenching [42]. Copyright 2016 American Chemical Society; (b) TFPPy−CHYD for Hg2+ detection via fluorescence quenching [73]. Copyright 2020 American Chemical Society; (c) Tp−Bpy NSs for Hg2+ detection via catalytic oxidation of TMB [29]. Copyright 2019 American Chemical Society.
Molecules 27 02586 g007
Figure 8. (a) TFPT−BTAN−AO for UO22+ detection via fluorescence quenching [90]. Copyright 2020 Nature Publishing Group; (b) DNA/CoPc−PT−COF@Cu−MOF nanocomposite for Cr3+ detection [93]. Copyright 2021 Royal Society of Chemistry.
Figure 8. (a) TFPT−BTAN−AO for UO22+ detection via fluorescence quenching [90]. Copyright 2020 Nature Publishing Group; (b) DNA/CoPc−PT−COF@Cu−MOF nanocomposite for Cr3+ detection [93]. Copyright 2021 Royal Society of Chemistry.
Molecules 27 02586 g008
Figure 10. (a) The azine-linked COF for picric acid (2,4,6-trinitrophenol) detection. 2,4,6-trinitrophenol (TNP), 2,4-dinitrophenol (DNP), 2,4-dinitrotoluene (DNT), 2-nitrophenol (NP), and 2-nitrotoluene (NT) [116]. Copyright 2013 American Chemical Society; (b) The COF for TATP sensing [115]. Copyright 2017 American Chemical Society.
Figure 10. (a) The azine-linked COF for picric acid (2,4,6-trinitrophenol) detection. 2,4,6-trinitrophenol (TNP), 2,4-dinitrophenol (DNP), 2,4-dinitrotoluene (DNT), 2-nitrophenol (NP), and 2-nitrotoluene (NT) [116]. Copyright 2013 American Chemical Society; (b) The COF for TATP sensing [115]. Copyright 2017 American Chemical Society.
Molecules 27 02586 g010
Figure 11. (a) Zr-coordinated amide porphyrin-based COF for tetracycline detection with the aid of molecular imprinting technology [144]. Copyright 2019 Elsevier B.V.; (b) The preparation of Mg@Fe−MIL−101/TpPa−1−COF, and its fluorescent detection of tetracycline realized simultaneously with photocatalytic tetracycline removal [152]. Copyright 2021 Elsevier B.V.; (c) EB-TFP:Eu(BTA)4 for the detection of 5-Fluorouracil [154]. Copyright 2021 American Chemical Society.
Figure 11. (a) Zr-coordinated amide porphyrin-based COF for tetracycline detection with the aid of molecular imprinting technology [144]. Copyright 2019 Elsevier B.V.; (b) The preparation of Mg@Fe−MIL−101/TpPa−1−COF, and its fluorescent detection of tetracycline realized simultaneously with photocatalytic tetracycline removal [152]. Copyright 2021 Elsevier B.V.; (c) EB-TFP:Eu(BTA)4 for the detection of 5-Fluorouracil [154]. Copyright 2021 American Chemical Society.
Molecules 27 02586 g011
Figure 12. (a) COF−300−AR for GSH detection [26]. Copyright 2020 American Chemical Society; (b) Py−TT COF for GSH sensing [27]. Copyright 2021 American Chemical Society.
Figure 12. (a) COF−300−AR for GSH detection [26]. Copyright 2020 American Chemical Society; (b) Py−TT COF for GSH sensing [27]. Copyright 2021 American Chemical Society.
Molecules 27 02586 g014
Figure 13. (a) Immunosensor for cTnI Detection based on TB−Au−COFs−Ab2 [165]. Copyright 2018 Elsevier B.V; (b) Immunosensor for cTnI Detection based on HRP−Ab2−Au−COF [166]. Copyright 2021 American Chemical Society.
Figure 13. (a) Immunosensor for cTnI Detection based on TB−Au−COFs−Ab2 [165]. Copyright 2018 Elsevier B.V; (b) Immunosensor for cTnI Detection based on HRP−Ab2−Au−COF [166]. Copyright 2021 American Chemical Society.
Molecules 27 02586 g012
Figure 14. (a) A versatile COF-based platform for sensing DNA [171]. Copyright 2017 Royal Society of Chemistry; (b) MOF@ COF Heterostructure Hybrid for Dual-Mode Photoelectrochemical−Electrochemical HIV-1 DNA Sensing [173]. Copyright 2021 American Chemical Society.
Figure 14. (a) A versatile COF-based platform for sensing DNA [171]. Copyright 2017 Royal Society of Chemistry; (b) MOF@ COF Heterostructure Hybrid for Dual-Mode Photoelectrochemical−Electrochemical HIV-1 DNA Sensing [173]. Copyright 2021 American Chemical Society.
Molecules 27 02586 g013
Table 1. The dynamic linkages for COF synthesis.
Table 1. The dynamic linkages for COF synthesis.
NameMonomer 1Monomer 2Linker
Boronate Molecules 27 02586 i001 Molecules 27 02586 i002 Molecules 27 02586 i003
Imine Molecules 27 02586 i004 Molecules 27 02586 i005 Molecules 27 02586 i006
Hydrazone Molecules 27 02586 i007 Molecules 27 02586 i008 Molecules 27 02586 i009
Maleimide Molecules 27 02586 i010 Molecules 27 02586 i011 Molecules 27 02586 i012
Phenazine Molecules 27 02586 i013 Molecules 27 02586 i014 Molecules 27 02586 i015
Acrylonitrile Molecules 27 02586 i016 Molecules 27 02586 i017 Molecules 27 02586 i018
Triazine Molecules 27 02586 i019 Molecules 27 02586 i020
Borazine Molecules 27 02586 i021 Molecules 27 02586 i022
Benzimidazole Molecules 27 02586 i023 Molecules 27 02586 i024 Molecules 27 02586 i025
Benzobisoxazole Molecules 27 02586 i026 Molecules 27 02586 i027 Molecules 27 02586 i028
Table 3. COF-based ion sensing.
Table 3. COF-based ion sensing.
AnalyteYearCOF NamesSpecific Binding SiteType of Detectable SignalDetection RangeLODReference
Hg2+2016COF−LZU8Thioether Fluorescence, “turn off”-25.0 ppb[42]
2019Tp−Bpy NSsAuNPsChromism-0.33 nM[29]
2020TFPPy−CHYDCarbohydrazide Fluorescence, “turn off”0.05 μM–4 μM17 nM.[73]
2021BATHz−BtCarbon-carbon double bonds Fluorescence, “turn off”0–27.5 mM26 nM[74]
Cu2+2016COF−JLU3Hydroxyl and azineFluorescence, “turn off”0–0.4 mM0.31 mM[75]
2017CTFTriazineChromism1.0 g/L-80.0 g/L0.05 g/L[50]
2017LMOP−15LMOP−15Tertiary amino groupFluorescence, “turn off” 5.1 × 10−8 M[76]
2018sp2c-COFsCyano groupsFluorescence, “turn off” 88 ppb[77]
2019QG−scaffolded COFN atoms and hydroxyl groupsFluorescence, “turn off”0.0010~10.0 μM0.50 nM[78]
Fe3+2017PI−COF 201Amino groupsFluorescence, “turn off”5.0–400 μM0.13 μM[79]
PI−COF 2025.0–300 μM0.22 μM
2019COF−TTAmino groupsFluorescence, “turn off”0–1.2 mM0.369 mM[80]
2019TaDAP TaDAImineFluorescence, “turn off”0.02–0.2 mM18 μM[81]
2019Bth−Dha, Bth−DmaO,N,O′-chelating sitesFluorescence, “turn off”-0.17 μM[82]
2021TTPE−COF Fluorescence “turn off”10−8–10−2 M3.07 μM[83]
2021Tfpa−Mth COFHydrazide and phenol etherFluorescence, “turn off” and QCM-64 nM[84]
2021PMDA−TAPBCarbonyl groupFluorescence, “turn off”--[85]
Pb2+2018TAPB-DMTP−COFAmino groups Electrochemical signals0.0050–2.0 μM0.0019 μmol/L[86]
2019Sulfhydryl modified TAPB−DMTTAPB−DMTTAPB−DMTP−COFSulfhydryl Electrochemical signals0.05–20 ng·mL−10.015 ng·mL−1[87]
2021TAPP−COF Photoelectrochemical signal0.05–1000 nM0.012 nM[88]
2021PMDA−TAPBAmino groupsElectrochemical signals5–9000 nM1.22 nM[85]
Au+2018TTB−COFThioetherFluorescence, “turn on”1.0–10.0 mM1.39 mM[89]
UO22+2020TFPT−BTAN−AOCarbon-carbon double bondsFluorescence, “turn off”-6.7 nM[90]
2021Tph−BDPImines of the CT complexChromism0.18–75 μM0.05 μM[91]
Ni2+2021BPD−COFsN atomsFluorescence, “turn off”0.420–1.26 × 103 pM68.0 pM[92]
Cr3+2021CoPc−PT−COF@Cu−MOFBipyridineElectrochemical signals10−1–105 pM0.0229 pM[93]
Pd2+2021XB−COFsCarbon-carbon double bondsFluorescence, “turn off”-0.29 μM[94]
2021PY−SE−COFSelenodiazoleFluorescence, “turn off”20–450 mM0.45 mM[95]
Table 4. COF-based pH sensing.
Table 4. COF-based pH sensing.
AnalyteYearCOF/COF Hybrid NamesSpecific Binding SiteType of Detectable SignalDetection RangeReference
H+2016COF−JLU4AmineFluorescence “turn off”pH 0.9–13.0[100]
2018COF−HQQuinolineFluorescence “turn off”pH 1.0–5.0[101]
2019COFDHTA-TTA-Electrochemical signalspH 3.0–11.0[102]
2021COF2Imine or triazine Fluorescence “turn on”pH 5.0–8.0[103]
2021COF−TPAmineFluorescence “turn off”pH 0–6.0[104]
Table 5. COF-based anion sensing.
Table 5. COF-based anion sensing.
AnalyteYearCOF NamesSpecific Binding SiteType of Detectable SignalDetection RangeLODReference
CrO42−(Cr2O72−)2019COF−TT Fluorescence, “turn off” 0.343 mM[80]
MnO42019COF−TT Fluorescence, “turn off” 0.320 mM[80]
F2015BCMP−3Boron sitesFluorescence, “turn off” [105]
2018TFPPy−DETHz−COFAmineFluorescence, “turn off”-50.5 ppb[106]
20182D−Fe−CTFTriazineChromism10–100 μM0.56 μM[107]
S22018TpASHAzideFluorescence, “turn on”1 μM–5 mM0.12 μM[108]
Table 6. COF-based explosive sensing.
Table 6. COF-based explosive sensing.
COF NamesYearAnalyteType of Detectable SignalDetection RangeLODReference
SNW−12012picric acidFluorescence, “turn off”0.2–52.4 μM0.05 μM[111]
COP−2
COP−3
COP−4
2012picric acidFluorescence, “turn off”-~1 ppm[117]
TNT~1 ppm
Py−Azine COF2013picric acidFluorescence, “turn off”0–70 ppm-[116]
COF−301
COF−401
2015picric acidFluorescence, “turn off”-1 ppm[112]
iPrTAPB−TFP2015picric acidFluorescence, “turn off”-1 ppm[118]
TfpBDH−CONs2015picric acidFluorescence, “turn on/off”-1 × 10−3 M[119]
TRIPTA2016picric acidFluorescence, “turn off”-51.96 nM[120]
3D−Py−COF2016picric acidFluorescence, “turn off”0−20 ppm-[121]
3′PD2017picric acidFluorescence, “turn off” [115]
triacetone triperoxide
PI−CONs2017picric acidFluorescence, “turn off”0.5–10 μM0.25 μM[122]
COP−612
COP−616
2017picric acidFluorescence, “turn off”-15 ppm[113]
LMOP−152017picric acidFluorescence, “turn off”-0.33 μM[76]
iPrTAPB−Azo−COP2018picric acidFluorescence, “turn off”-13 ppm[123]
12018picric acidFluorescence, “turn off”-68 ppb[124]
CMP−LS1
CMP−LS2
2018picric acidFluorescence, “turn off”--[125]
Py−TPE−COF2018picric acidFluorescence, “turn off”-10 ppm[126]
DL−COF2019picric acid Fluorescence, “turn off”-13.10 ppb[127]
2,4-dinitrophenol8.56 ppb
2,4-dinitrotoluene10.40 ppb
4-nitrophenol5.15 ppb
4-nitrotoluene6.92 ppb
LPCMP1−42019TNTFluorescence, “turn off”0–100 ppm-[128]
ANCOF2020DichloranFluorescence, “turn off”-142 ppb[129]
4-nitroaniline-89 ppb
A−COF2021picric acidFluorescence, “turn off”-0.09 μM[114]
TFPB−TTA COF2022DNPFluorescence, “turn off”50 nM–10 μM18 nM[130]
picric acid50 nM–12.5 μM16 nM
Table 7. COF-based detection of drugs, insecticides and spices.
Table 7. COF-based detection of drugs, insecticides and spices.
YearCOF/COF Hybrid NamesAnalyteSpecific Binding SiteType of Detectable SignalDetection RangeLODReference
2018TAPB−DMTTAPB−DMTP−COFs/AuNPsChlorogenic acid-Electrochemical signals (CV)1.0 × 10−8–4.0 × 10−5 mol L−19.5 × 10−9 mol L−1[140]
20182D Fe−CTFsSarcosineSarcosine oxidaseChromism10–100 μM0.56μM[107]
ochratoxin AAptamer0.2–0.8 μM-
2019Py−M−COFEnrofloxacinAptamerElectrochemical signals (EIS)0.01 pg mL−1–2 ng mL−16.07 fg mL−1[49]
Ampicillin0.001–1000 pg mL−10.04 fg mL−1
2019QD−grafted COFsFerulic AcidMolecular imprinting, amino groupsFluorescence, “turn on”0.03–60 mg kg−15 μg kg−1[141]
2019MIP/MoS2/NH2−MWCNT@COFSulfamerazineMolecular imprintingElectrochemical signals0.30–2.0 × 102 μM0.11 μM[142]
2019Ce−MOF@MCAOxytetracyclineAptamerElectrochemical signal (EIS)0.1–0.5 ng mL−117.4 fg/mL[143]
2019Zr−amide−Por-based 2D COFTetracyclineMolecular imprintingElectrochemical signals5–60 pM2.3 pM[144]
2019Eu@TpPa−1LevofloxacinEuropium ionsFluorescence, “turn off”10−6–10−2 M0.2 μM[145]
2019MIOP based on QDs−grafted COFsTyramineH-bond, shape selectivityFluorescence, “turn on”35–35,000 µg/kg7.0 µg/kg[146]
SPE–HPLC20–2000 µg/kg5.0 µg/kg
2019NUS−30L-dopaAzineFluorescence, “turn off” [147]
2019PATP@AuNPs−crosslinked MIPDopamineMolecular imprintingelectrochemiluminescence10–14−10−6 M2 × 10−15 M[148]
2020UiO−66−NH2/MCA/MWCNT@rGONRKanamycinAptamerElectrochemical signals25–900 nM13 nM[149]
2020TpPa−1@DyeSialic acidCr3+Fluorescence, “turn on”10−8–10−2 M7.08 × 10−9 M[150]
2021Au@COF/GO−NH2ChloramphenicolAptamerElectrochemical signal (EIS)0.0001–1 ng mL-116.13 fg mL−1[151]
2021Mg@Fe−MIL−101/TpPa−1−COFTetracyclineMg2+Fluorescence, “turn off”--[152]
2021COF−1 or COF−2Tetracycline -Fluorescence, “turn off”0.005−0.0625 mM0.002 mM[153]
ofloxacin 0.025−0.25 mM0.0065 mM
2021Eu@TpPa−15-Fluorouracilπ−π stacking interactionsFluorescence10−7–10−3 M6.45 × 10−8 M[154]
2018DAAQ−TFPDiflubenzuronAmino and carbonyl group (H bonding) π−π stacking interactionsHPLC0.2–160.0 ng mL−10.02 ng mL−1[155]
Triflumuron0.2–160.0 ng mL−10.02 ng mL−1
Hexaflumuron0.2–160.0 ng mL−10.05 ng mL−1
Teflubenzuron0.2–160.0 ng mL−10.04 ng mL−1
2019NH2@COFCarboxylic acid pesticidesAmino groupHPLC-DAD0.2–100 ng mL−10.04–0.20 ng mL−1[156]
2019CNs−grafted COFs@MIP4-ethylguaiacolThe surface of the silica matrix by acid–base pairing interactionsFluorescence, “turn off”0.025–1μg ml−117 ng mL−1[157]
Table 8. COFs or COF-based hybrid materials for detecting glucose, uric acids, GSH, and other small biomolecules.
Table 8. COFs or COF-based hybrid materials for detecting glucose, uric acids, GSH, and other small biomolecules.
AnalyteYearCOF/COF Hybrid NamesSpecific Binding SiteType of Detectable SignalDetection RangeLODReference
Glucose2018Fe−COFGlucose oxidaseChromism 5–350 μM1.1 μM[160]
2019COFDHTA-TTAGlucose oxidaseElectrochemical signal0.60 μM–6.0 mM0.38 μM[102]
2020Fe−PorCOFGlucose oxidaseChemiluminescence0.01–10 μM5.3 nM[161]
2021COFHD–GOXGlucose oxidaseChromism 5–2000 μM0.54 μM[162]
Uric acid2021COF−DC−8Hydroxyl,
triazine
Electrochemical signals5.0–25 μM,
25–250 μM
0.77 μM[163]
Ascorbic acid2021COF−DC−8Hydroxyl,
triazine
Electrochemical signals30–180 μM,
0.18–1.5 μM
12.0 μM[163]
Dopamine2021COF−DC−8Hydroxyl,
triazine
Electrochemical signals1.0–6.0 μM,
8.0–50 μM
0.25 μM[163]
GSH2020COF−300−AR Chromism1–15 μM1.0 μM[26]
2021Py−TT COF Chromism0.4 − 60 μM0.225 μM[27]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, S.; Liu, D.; Wang, G. Covalent Organic Frameworks for Chemical and Biological Sensing. Molecules 2022, 27, 2586. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27082586

AMA Style

Zhang S, Liu D, Wang G. Covalent Organic Frameworks for Chemical and Biological Sensing. Molecules. 2022; 27(8):2586. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27082586

Chicago/Turabian Style

Zhang, Shiji, Danqing Liu, and Guangtong Wang. 2022. "Covalent Organic Frameworks for Chemical and Biological Sensing" Molecules 27, no. 8: 2586. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27082586

Article Metrics

Back to TopTop