Electrochemical (Bio)Sensors Based on Covalent Organic Frameworks: Comparison
Please note this is a comparison between Version 2 by Vivi Li and Version 1 by Emiliano Martínez-Periñán.

Covalent organic frameworks (COFs) are defined as crystalline organic polymers with programmable topological architectures using properly predesigned building blocks precursors. Since the development of the first COF in 2005, many works are emerging using this kind of material for different applications, such as the development of electrochemical sensors and biosensors. COF shows superb characteristics, such as tuneable pore size and structure, permanent porosity, high surface area, thermal stability, and low density. Apart from these special properties, COF’s electrochemical behaviour can be modulated using electroactive building blocks. Furthermore, the great variety of functional groups that can be inserted in their structures makes them interesting materials to be conjugated with biological recognition elements, such as antibodies, enzymes, DNA probe, aptamer, etc. Moreover, the possibility of linking them with other special nanomaterials opens a wide range of possibilities to develop new electrochemical sensors and biosensors.

  • COF
  • electrochemical biosensors
  • electrochemical sensors
  1. Introduction

1. Introduction

Covalent organic frameworks (COFs) are a type of polymers, which connect organic molecules in two (2D) or three dimensions (3D) via covalent bonds[1,2] [1][2]. Their most noteworthy feature is their intrinsic order, which is predetermined by the monomers, or linkers, employed in the polymerization, building crystalline structures with pre-designable architectures [3,4,5][3][4][5]. The design of the network’s topology lies in the control of the direction of covalent bond formation during the polymerization [1]. To ensure the directionality of growth, the monomers must be formed by relatively rigid bonds and present the functionalities in specific positions [6,7][6][7]. To do this, the design is approached from a simplified block model, where each monomer is represented with a specific geometric shape that symbolizes the relative positions of the reactive points [6,7,8][6][7][8]. In this way, as depicted in Figure 1 Figure 1A, a linear monomer could generate hexagonal networks by cyclotrimerization (for example boroxine formation) or produce square COFs if it is combined with a C4  linker. Following the pioneering work of Yaghi and the co-workers [9]of Yaghi [9], several structural motifs have been described, including hexagonal [10[10][11],11], kagome [12,13,14][12][13][14] or square [15,16][15][16] two-dimensional networks (2D-COFs), and diamonoid [17,18][17][18], Cubic [19][19], [19] or PtS PtS [20] [20] three-dimensional architectures (3D-COFs). It is worth pointing out that the dimensionality of the COF network is also determined by the linkers employed in the polymerization. 2D-COFs are generally built from flat molecules, while 3D-COFs are usually based on linkers endowed with sp3  centres [18] or a geometry significantly distorted from planarity via steric tuning [20].

Figure 1. A) Examples of COFs topology diagrams. B) Examples of linkages commonly used for the synthesis of COFs (from left to right: boronate ester, imine, imide and triazine. C) Examples of arrangements of COF’s networks. D) List of linkers named in this entreviewy.

The type of bond used to connect the building blocks, commonly known as linkage, is crucial during the formation of the framework as multiple covalent structures with different degrees of crystallinity can be produced during the polymerization reactions [6,7][6][7]. In fact, the main difference between COFs and Pporous Oorganic Ppolymers (POPs) is the crystalline order. The generally most widespread strategy to prevent the formation of amorphous products and favour the COF’s crystallization is the use of reversible reactions [6,7][6][7]. In this way, the balance between reactants and products allows curing and error correction of the amorphous network, usually associated to kinetic control products, until the COF crystallizes as the thermodynamically more stable product [7]. These are the characteristic features of dynamic covalent chemistry (DCC) [21,22][21][22]. In simple terms, in a system where DCC operates, it is the stability of the products that decides the final distributions [2]. Thus, several linkages have been used for the formation of COFs, such as boroxine, boronate esterification, imine, azine, imide, or triazine (Figure 1B) [6]. Other approaches to enhance the crystallinity of the systems includes the addition of monofunctional modulators [18], competitors[23] [23], and protected monomers [24] as strategies to control the nucleation and growth of the crystalline phases [25,26][25][26]. Unfortunately, the crystallization conditions to obtain the more thermodynamically more stable products are not known a priori, a priori and the COF crystallization is usually achieved via iterative screening methodology. Nevertheless, a plethora of methodologies have been reported in the last years to enhance the crystallinity of COFs, including the solvothermal synthesis [10], microwave-assisted methods [27] [27], ionothermal synthesis [28], interfacial synthesis [29], repolymerization methods [30][30], or on-surface synthesis [31]. Furthermore, screening of experimental conditions and structure-property relationships are being accelerated through computational calculations and automated machine learning [32[32][33][34],33,34], which are emerging as powerful tools to obtain tailor-made materials in record times.

Another distinctive characteristic of COFs is their inherent porosity [35,36,37][35][36][37]. On the one hand, the overlap of the 2D-COFs layers can build up a three-dimensional structure, traversed by unidirectional cavities also known as pores, which can mainly present two mainly ias two isomers. The first one is an eclipsed structure (AA), composed by adjacent layers, which are located in the same position. The second one, is a staggered structure (AB or ABC), where every COF layer is shifted with respect to the previous one. It is worth pointing out that the staggered structure in 2D-COFs also presents pores traversing the three-dimensional structure. However, in terms of porosity, the structure whichthat presents the highest surface area is the eclipsed isomer network [1,2][1][2]. On the other hand, networks entanglements can be produced in 3D-COFs, thus lowering the theoretical surface area (Figure 1C) [18].

COFs presents high thermal, mechanical, and chemical stabilities, with differences depending on the linkage and/or the monomers used in the polymerization reaction [6]. The above-mentioned characteristics, together with their inherent insolubility [1], place COFs as ideal materials to explore their applicability in heterogeneous phases. In this way, COFs have already been used in different applications, such as catalysis [38[38][39][40],39,40], solar cells [41], batteries [42,43][42][43], gas storage and separation [44[44][45],45], water remediation [39[39][46],46] and sensing [47].

Concerning with the integration in devices, COFs must face the doble edged sword of one of their main attributes, their inherent insolubility. In this way, the most common procedure to achieve tailor-made devices are two depending on the followed path: (i) Top-down protocol involves the destruction of the granular COF structure, achieving the sheet separation in a process known as exfoliation producing covalent organic nanosheets (CONs) [48]. To address this end, several methods have been developed. The most common procedure is liquid phase exfoliation, which consists in the use of different solvents to produce layer slipping and the subsequent delamination in an ultrasonic bath [49]. Another method is the mechanical delamination, which produces the exfoliation by using an external force over the bulk COF [50]. Chemical or acid spontaneous exfoliations are receiving a great deal of attention because itthey causes the COF delamination using chemical energy, reducing the costs for obtaining CONs [49]. Finally, nitrogen delamination consists in the gentlye heating of the bulk COF and subsequent addition of liquid nitrogen, which intercalates between the COF layers and produces exfoliation by the thermal expansion in the liquid to -to-gas phase transition [51]. (ii) The use of bottom-up methods involves the formation of the CONs or the tailor-made COF directly from the monomers. On the one hand, for the synthesis of nanosheets ist usually requireds the use of additives, such as tetra-n-butylammonium fluoride (TBAF) micelles or surfactants mediators producing monomodal distribution of CONs by the restriction of the growth dimensions. Thus, mi-cellar synthesis allows the growing of large few-layered COFs in the interphase between two immiscible solvents [29,52][29][52]. On the other hand,  tailor-shaped COFs are obtained by using either the 3D-printing technique, which involves the preparation of a COF “ink” and the subsequent printing following a layer-by-layer protocol [53,54][53][54] or the formation of monolithic COF aerogels by crystallization in moulds to produce tailor-made macroscopic objects [55,56][55][56].

Concerning sensing applications, between all the COF-based sensor assemblies, electrochemical sensors present important advantages, such as their high porosity, which results oin increased sensibility, and the high specificity and fantastic biocompatibility that improves the stability of the electrochemical sensors [57]. The most common procedure followed to achieve these devices is the use of COFs as supports to anchor the active species (or guests) towards the recognition [58,59] [58][59]. In this way, the COF enhances the electrochemically active surface area of the electrode by preventing the agglomeration of the active species [59,60][59][60]. There are mainly three main methods to anchor the active species into COFs. The first one involves the coordination of the metallic species in Lewis acid or basic positions of the COF [61]. The second one involves the active specie anchorage by other non-covalent interactions, such as π-interactions [59]. Finally, covalent immobilization can be also be achieved by chemical reactions involving backbone modification or pendant groups reactions, highlighting the Hüisgen’s  Hüisgen’s azide-alkyne cycloaddition as the most prominent example of these guests immobilization methods [5].

To achieve electrochemical sensing, great conductivity is desired. However, despite

To achieve electrochemical sensing, great conductivity is desired. However, despite COFs presenting an ordered and columnar π-skeleton, the electrochemical sensing is still limited due to the polygranular COF’s morphology, leading to insignificant hopping-type electrical conductivity in most of the examples [62]. In the past few years, several strategies have been developed to address this problem. To avoid this problem, conductive additives can be added during the sensor assembly. The most common example is the in situ reduction of metallic precursors to produce metallic nanoparticles (MNPs), which can be anchored to the COF pore walls by coordinative interactions preventing the NPs agglomeration [63][64]. Other common conductive additives are carbonaceous compounds, such as carbon black (CB) [65], pyrolyzed three-dimensional kenaf stem (3D-KSC) [66][67][68], or multi-wall carbon nanotubes (MWCNTs) [69].
Sensor assemblies are usually achieved by different strategies. One of them involves the mixing of the COF with the conductive additives in a hollow glass tube [70][71]. A second strategy involves the electrode modification with Covalent Organic Nanosheets (also known as CONs [72][73]), produced by delamination of COFs usually by liquid phase exfoliation (LPE) and subsequent drop casting of CONs colloids [74][75][76]. Finally, the sensor’s design can be based on simple electrodes [58][61] of dual systems, highlighting the sandwich-type electrodes [77][78].
Different examples of electrochemical sensors and biosensors are exposed in this entry, showing the great versatility of functions and capabilities of materials based on COFs, resulting in interesting approaches that can enhance the sensibility and selectivity of the final devices.

2. Electrochemical Sensors

COFs present an ordered and columnar π-skeleton, the electrochemical sensing is still limited due to the polygranular COF’s morphology, leading to insignificant hopping-type electrical conductivity in most of the examples [62]. In the past few years, several strategies have ave been developed to address this problem. To avoid this problem, conductive additives can be added during the sensor assembly. The most common example is the in situ reduction of metallic precursors to produce metallic nanoparticles (MNPs), which can be anchored to the COF pore walls by coordinative interactions preventing the NPs agglomeration [63,64]. Other common conddely uctive additives are carbonaceous compounds such as carbon black (CB) [65], pyrolyzed three-ded imensional kenaf stem (3D-KSC) [66,67,68] or multi-wall carbon nanotubes (MWCNTs) [69] .

Sensor assemblies are usually ac thieved by different strategies. One of them involves the mixing of the COF with the conductive additives in a hollow glass tube [70,71]. A second strategy involves the electrode m develodification with Covalent Organic Nanosheets (also known as CONs [72,73]), produced by delamination of COFs usually by liquid phase exfoliation (LPE) and subsequent drop casting of CONs colloids [74,75,76]. Finally, thent of e sensor’s design can be based on simple electrodes [58,61] of dual systems, highlighting the sandwich type electrodes [77,78].

Different examples of electrochemical sensors and biosensors are exposed in this review, showing the great versatility of functions and capabilities of materials based on COFs, resulting in interesting approaches that can enhance the sensibility and selectivity of the final devices.

  1. Electrochemical Sensors

COFs have been widely used in the development of elect. Researcherochemical sensor. We have classified COFs in three different groups attending to their main role in electrochemical sensor.

2.1. COFs acting as electrocatalysts

The easiest configuration for the use of COFs in the development of electrochemical sensors involves the direct deposition of the COF as modifier on top of a working electrode surface. The high porosity of COFs-based materials and their crystalline structure make them interesting materials for that purpose as shown in the work developed by Y.-H. Pang et al. [79].

2.2. COFs acting as support of Electrocatalyst and/or recognition elements.

One of the main functions of COFs in this context is their role as recognition element in the sensor. The high porosity of COFs together with the possibility of delamination of 2D-COFs allow the production of 2D-nanomaterials with enhanced surface-volume ratio which allows an increment of the sensor recognition element or the electroactive centre exposition to the measuring solution media.

2.3. COFs chelating properties for anodic stripping analysis.

The capability of COFs to include multiple functional groups in their structures allows the development of COFs endowed with chelating groups. There are several examples in the literature showing the use of COFs as chelating agents to promote pre-concentration of metal cations for differential pulse anodic stripping voltammetry.

  1. Electrochemical biosensors

3. Electrochemical biosensors

COF skeletons are superior scaffolds to allow charge migration and improve the signal amplification in electrochemical sensors. The large specific surface area and tuneable pore structure of COFs make of them good candidates for the development of functional biosensors.

3.1. Enzymatic Biosensors

The structure tunability, porosity, crystallinity, and stability of COFs are promising characteristics for their use envisaged as host arrays for the immobilization of enzymes, among them tunability [110][80]. Their customizable composition through the proper choosing of their precursor, which also determine the functional groups on their surfaces, can be tailored to favour some specific interactions between them and enzymes. Furthermore, the regular distribution of nanopores in the COFs is beneficial to provide a high surface area interface for the adsorption and desegregation of enzymes and to allow a rapid movement of reagents. Moreover, the structural robustness of COFs represents an important attribute to stabilize enzymes during the biosensor manufacture process. Some reported works have used COFs to encapsulate multiple enzymes simultaneously [112][81].

3.2. Immunosensors

COFs have also been used as support material for the development of immunosensors. COFs can be also used to develop electroactive labels linked with secondary antibodies. The high porosity together with the high amount of adsorption sites along the COFs nanostructure make them excellent candidates to guest electroactive molecules, which can be reduced or oxidized during the measuring step, obtaining an electrochemical signal proportional to the concentration of the antigen being analysed [114][82]. A different strategy to obtain a labelling agent is the use of COFs together with peroxidase enzymes, which are adsorbed on the COFs nanostructures [117][83].

3.3. Genosensors

COFs can be used as structural materials to adsorb oligonucleotides and their further use as DNA probes for the development of genosensors [119][84].

3.4. Aptasensors

COFs can be also used as supports to adsorb oligonucleotides, which are capable of recognizing target proteins with an affinity and specificity rivalling that of antibodies [122][85]. The combination of COFs with carbon nanomaterials has been also used for the development of aptasensors as recently reported by He and co-workers [126][86]. COFs have been also used in combination with metal nanoparticles for the development of aptasensors. Metal nanoparticles can supply the limited electrical conductivity of COFs [128][87]. COFs have been also combined with MOFs for the development of aptasensors [59].

 

Table 1. Examples of COFs application in different samples.

Table 1. Examples of COFs application in different samples.

COF

COF

Modifier

Modifier

Analyte

Analyte

Sample

Sample

Electrode

Electrode

LOD

LOD

Linear Range

Linear Range

Sensitivity

Sensitivity

Ref

Ref.

DQTP- COF.

β-ketoenamine

Bisphenol A, and S

Food packages

Graphite

0.15 and

0.15 µM

0.5-30 and

0.5-30 µM 

0.239 and 0.150  µA/µM 

[80]

TPA-COF

Composite Carbon Black

Dopamine

Medical injections

Glassy Carbon

0.17  µM

20-1000  µM

0.023  µA/µM

[65]

TAPB-DMTP-COF@PANI

polyanyline

Acetaminophen

tablets, human blood, serum and urine

Glassy Carbon

0.032  µM 

0.10-500  µM

0.1229 µA/µM

[84]

Fe3O4@AT-COF

Fe3O4

p-nitrophenol and o-nitrophenol

lake and tap water

Magnetic beads

0.2278 and 0.5925 μM

10-3000 and 10-3000 µM

0.7588  and 0.7799  µA/µM

[85]

FeTAPP-TA-COF

Composite graphene aerogel

NO

Complex biological system

Glassy carbon

0.030

μM 

0.18-400 μM

8.8 μA/‧μM·cm2

[86]

MA-TP-COF

β-Ketoamine

Cd2+, Cu2+, Pb2+, Hg2+ and Zn2+

Drinking water

Glassy Carbon

0.922, 0.450, 0.309, 0.208 and 0.526 nM

-

17.8, 36.6, 53.2, 79.1 and 31.3 μA/μM cm2

[106]

DTPA-TFB-COF

kenaf stem-derived macroporous carbon

Cd2+, Pb2+, Cu2+ and Hg2+

Soil and sewage

 

12.3 , 11.8 , 18.6  and 21.4 nM

0.0369-18.0,

0.0356-19.0,

0.0536-19.0, and

0.0503-18.0 µM

1337.4 , 1389.0  886.2 and 770.0  μA/μM cm2

[108]

GOD/DMFc/ PA-TFB-COF/CFMEs)

DMFc and GOD

Glucose

Rats’ brains

Carbon fibre microelectrode

0.36 µM

1.08 μM to 8.5 mM

46.55  mV/ mM cm 2

[111]

enzyme@ZIF-8@COF

GOD, HRP, AChE

Glucose, H2O2 and malathion

 

Glassy Carbon

0.85 μM, 2.81 nM, 3.0×10−13 g/L

2.83 μM–8.0 mM, 9.53 nM-7.0 μM, 10−12 g/L-10−8 g/L

-

 

[75]

COFs-AuNPs

AuNPs, Capture antiKIM-1

KIM-1

Plasma samples

Glassy Carbon

2.00 fg/ mL

0.01–50.00 pg/ mL

1.8981 µA·mL/pg

[113]

AuPt@MnO2@COF

AuNPs, PSA affinity peptide 

PSA

Human serum

 

16.7 fg/mL

0.00005–10 ng/mL

2.237 µA/log(ng/mL)

[115]

TP-PANO2-COF

AuNPs, CYFRA21-1 antibodies

CYFRA21-1

Human serun

 

0.1 pg/mL

0.5–1.0 × 104 pg/mL

6.3  µA/log(pg/mL

[116]

Cu-MOF@Cu-PcTA-COF

Cu-MOF, HIV-1 DNA probe strands

HIV-1 DNA

Human serun

Glassy carbon

0.07 fM

1 fM to 1 nM

-

[119]

Acronyms

 

1,3,5-triformylphloroglucinol (TP) and 2,6-diaminoanthraquinone (DQ)

 

TPA-COF triphenylamine-based covalent-organic framework

 

TAPB, 1,3,5-tris(4-aminophenyl)benzene; DMTA, 2,5-dimethoxyterephaldehyde

 

1,3,5-tris(4-aminophenyl) benzene (TAPB) and 1,3,5-benzenetricarboxaldehyde (TFB)

 

5,10,15,20-tetrakis (4-aminophenyl) porphinato]-iron (Fe-TAPP) and terephthalaldehyde (TA),

 

2,4,6-triformylphloroglucinol and melamine (MA-TP-COF).

 

1,4-benzenedithiol-2,5-diamino-hydrochloride (DTPA) and 1,3,5-triformylbenzene (TFB)

 

1,3,5 -Triformylbenzene (TFB) and 1,4-diaminobenzene (PA)

 

1,3,5-tris(p-formylphenyl) benzene (TFPB) and 4,4′-diaminobiphenyl-2,2′-dicarboxylic acid (DBD) in the presence of enzyme@ZIF-8

 

Triformylphloroglucinol and 2-nitrobenzene-1,4-diamine (PANO2),

 

Copper-phthalocyaninetetra-amine (Cu-PcTA) and 2,9-bis[p-(formyl)phenyl]-1,10-phenanthroline (PDB)

 

          
DQTP-COFβ-ketoenamineBisphenol A, and SFood packagesGraphite0.15 and

0.15 µM
0.5–30 and 0.5–30 µM0.239 and 0.150 µA/µM[88]
TPA-COFComposite Carbon BlackDopamineMedical injectionsGlassy Carbon0.17 µM20–1000 µM0.023 µA/µM[65]
TAPB-DMTP-COF@PANIpolyanylineAcetaminophenTablets, human blood, serum and urineGlassy Carbon0.032 µM0.10–500 µM0.1229 µA/µM[89]
Fe3O4@AT-COFFe3O4p-nitrophenol and o-nitrophenolLake and tap waterMagnetic beads0.2278 and 0.5925 μM10–3000 and 10–3000 µM0.7588 and 0.7799 µA/µM[90]
FeTAPP-TA-COFComposite graphene aerogelNOComplex biological systemGlassy carbon0.030 μM0.18–400 μM8.8 μA/μM·cm2[91]
MA-TP-COFβ-KetoamineCd2+, Cu2+, Pb2+, Hg2+ and Zn2+Drinking waterGlassy Carbon0.922, 0.450, 0.309, 0.208 and 0.526 nM-17.8, 36.6, 53.2, 79.1 and 31.3 μA/μM cm2[92]
DTPA-TFB-COFkenaf stem-derived macroporous carbonCd2+, Pb2+, Cu2+ and Hg2+Soil and sewage 12.3, 11.8, 18.6 and 21.4 nM0.0369–18.0, 0.0356–19.0, 0.0536–19.0, and 0.0503–18.0 µM1337.4, 1389.0, 886.2 and 770.0 μA/μM cm2[93]
GOD/DMFc/PA-TFB-COF/CFMEsDMFc and GODGlucoseRats’ brainsCarbon fibre microelectrode0.36 µM1.08 μM to 8.5 mM46.55 mV/mM cm2[94]
enzyme@ZIF-8@COFGOD, HRP, AChEGlucose, H2O2 and malathion Glassy Carbon0.85 μM, 2.81 nM, 3.0 × 10−13 g/L2.83 μM–8.0 mM, 9.53 nM–7.0 μM, 10−12 g/L–10−8 g/L-[75]
COFs-AuNPsAuNPs, Capture antiKIM-1KIM-1Plasma samplesGlassy Carbon2.00 fg/mL0.01–50.00 pg/mL1.8981 µA·mL/pg[95]
AuPt@MnO2@COFAuNPs, PSA affinity peptidePSAHuman serum 16.7 fg/mL0.00005–10 ng/mL2.237 µA/log (ng/mL)[96]
TP-PANO2-COFAuNPs, CYFRA21-1 antibodiesCYFRA21-1Human serum 0.1 pg/mL0.5–1.0 × 104 pg/mL6.3 µA/log (pg/mL)[97]
Cu-MOF@Cu-PcTA-COFCu-MOF, HIV-1 DNA probe strandsHIV-1 DNAHuman serumGlassy carbon0.07 fM1 fM to 1 nM-[84]
Acronyms
1,3,5-triformylphloroglucinol (TP) and 2,6-diaminoanthraquinone (DQ)
TPA-COF triphenylamine-based covalent-organic framework
TAPB, 1,3,5-tris(4-aminophenyl)benzene; DMTA, 2,5-dimethoxyterephaldehyde
1,3,5-tris(4-aminophenyl) benzene (TAPB) and 1,3,5-benzenetricarboxaldehyde (TFB)
5,10,15,20-tetrakis [(4-aminophenyl) porphinato]-iron (Fe-TAPP) and terephthalaldehyde (TA)
2,4,6-triformylphloroglucinol and melamine (MA-TP-COF)
1,4-benzenedithiol-2,5-diamino-hydrochloride (DTPA) and 1,3,5-triformylbenzene (TFB)
1,3,5-Triformylbenzene (TFB) and 1,4-diaminobenzene (PA)
1,3,5-tris(p-formylphenyl) benzene (TFPB) and 4,4′-diaminobiphenyl-2,2′-dicarboxylic acid (DBD) in the presence of enzyme@ZIF-8
Triformylphloroglucinol and 2-nitrobenzene-1,4-diamine (PANO2)
Copper-phthalocyaninetetra-amine (Cu-PcTA) and 2,9-bis[p-(formyl)phenyl]-1,10-phenanthroline (PDB)

 

4. Conclusions

4. Conclusions

COFs are new and emerging materials with multiple application fields. Among them, sensor and biosensor development is a wide area of application of these materials, as a consequence of their extraordinary properties. In particular, it is worth highlighting their high porosity, appropriate pore sizes for the immobilization of specific recognition elements and post-modification capacity. In addition, the possibility of predesigning chemical structures according to the correct selection of their precursors is opening the door to interesting sensors and biosensors developments. COF can be used in the sensor platforms in order to improve specific area, serving as supportive material for electrocatalysts, enzymes, antibodies, oligonucleotides…etc. COF can also have a specific function related with their structure, acting as electrocatalyst, recognizing element and offering the functional groups needed for specific adsorption or chelating process. Innovative strategies are using COFs as label agent developments, linking them with specific antibodies, aptamer, enzymes, oligonucleotides etc. The great variety of COFs structures and the enormous amount of new COF designs would have a starring role in the next generation of electrochemical sensors and biosensors. Finally, prospects of sensors assemblies should address the problems produced from the inherent insolubility of the COFs. Thus, the use of dynamic covalent chemistry to produce CONs with monomodal distributions through bottom-up strategies may improve the sensor quality, since the top-down strategies may be very destructive, yielding uncontrolled distribution of nanosheets sizes. Furthermore, to ensure a sustainable future, the inclusion of metallic additives on the sensor composition should be avoided. In addition, other additives could produce pore blockage limiting the analyte diffusion to the active sites. Furthermore, the characteristic electrical insulator behaviour of COFs must be ad-dressed by either developing conductive COFs or by using few-layer COFs ensuring close contact between COFs and electrodes. Finally, the new trends on tailor-made COFs by 3D-printing techniques and shapeable monolithic aerogels will pave the way for the production of tailor-made sensors based on COFs to ensure proper device integrations and scalability.

 

References

  1. Geng, K.; He, T.; Liu, R.; Dalapati, S.; Tan, K.T.; Li, Z.; Tao, S.; Gong, Y.; Jiang, Q.; Jiang, D. Covalent Organic Frameworks: Design, Synthesis, and Functions. Chem. Rev. 2020, 120, 8814–8933.
  2. Hu, J.; Gupta, S.K.; Ozdemir, J.; Beyzavi, H. Applications of Dynamic Covalent Chemistry Concept toward Tailored Covalent Organic Framework Nanomaterials: A Review. ACS Appl. Nano Mater. 2020, 3, 6239–6269.
  3. Zhang, K.; Kirlikovali, K.O.; Varma, R.S.; Jin, Z.; Jang, H.W.; Farha, O.K.; Shokouhimehr, M. Covalent Organic Frameworks: Emerging Organic Solid Materials for Energy and Electrochemical Applications. ACS Appl. Mater. Interfaces 2020, 12, 27821–27852.
  4. Lohse, M.S.; Bein, T. Covalent Organic Frameworks: Structures, Synthesis, and Applications. Adv. Funct. Mater. 2018, 28, 1705553.
  5. Segura, J.L.; Royuela, S.; Ramos, M.M. Post-synthetic modification of covalent organic frameworks. Chem. Soc. Rev. 2019, 48, 3903–3945.
  6. Zhao, F.; Liu, H.; Mathe, S.D.R.; Dong, A.; Zhang, J. Covalent Organic Frameworks: From Materials Design to Biomedical Application. Nanomaterials 2017, 8, 15.
  7. Li, X.; Yang, C.; Sun, B.; Cai, S.; Chen, Z.; Lv, Y.; Zhang, J.; Liu, Y. Expeditious synthesis of covalent organic frameworks: A review. J. Mater. Chem. A 2020, 8, 16045–16060.
  8. Bourda, L.; Krishnaraj, C.; Van Der Voort, P.; Van Hecke, K. Conquering the crystallinity conundrum: Efforts to increase quality of covalent organic frameworks. Mater. Adv. 2021, 2, 2811–2845.
  9. Diercks, C.S.; Yaghi, O.M. The atom, the molecule, and the covalent organic framework. Science 2017, 355, eaal1585.
  10. Martínez-Fernández, M.; Gavara, R.; Royuela, S.; Fernández-Ecija, L.; Martínez, J.I.; Zamora, F.; Segura, J.L. Following the light: 3D-printed (2-hydroxyethyl methacrylate) dual emissive composite with response to polarity and acidity. J. Mater. Chem. A 2022, 10, 4634–4643.
  11. Martínez-Fernández, M.; Martínez-Periñán, E.; Royuela, S.; Martínez, J.I.; Zamora, F.; Lorenzo, E.; Segura, J.L. Covalent organic frameworks based on electroactive naphthalenediimide as active electrocatalysts toward oxygen reduction reaction. Appl. Mater. Today 2022, 26, 101384.
  12. Tian, Y.; Xu, S.-Q.; Liang, R.-R.; Qian, C.; Jiang, G.-F.; Zhao, X. Construction of two heteropore covalent organic frameworks with Kagome lattices. Cryst. Eng. Comm. 2017, 19, 4877–4881.
  13. Abuzeid, H.R.; El-Mahdy, A.F.M.; Kuo, S.-W. Hydrogen bonding induces dual porous types with microporous and mesoporous covalent organic frameworks based on bicarbazole units. Micropor. Mesopor. Mat. 2020, 300, 110151.
  14. Xing, G.; Zheng, W.; Gao, L.; Zhang, T.; Wu, X.; Fu, S.; Song, X.; Zhao, Z.; Osella, S.; Martínez-Abadía, M.; et al. Nonplanar Rhombus and Kagome 2D Covalent Organic Frameworks from Distorted Aromatics for Electrical Conduction. J. Am. Chem. Soc. 2022, 144, 5042–5050.
  15. Spitler, E.L.; Colson, J.W.; Uribe-Romo, F.J.; Woll, A.R.; Giovino, M.R.; Saldivar, A.; Dichtel, W.R. Lattice Expansion of Highly Oriented 2D Phthalocyanine Covalent Organic Framework Films. Angew. Chem. Int. Ed. 2012, 51, 2623–2627.
  16. Royuela, S.; Almarza, J.; Mancheño, M.J.; Flores, J.C.P.; Michel, E.G.; Ramos, M.M.; Zamora, F.; Ocón, P.; Segura, J.L. Synergistic Effect of Covalent Bonding and Physical Encapsulation of Sulfur in the Pores of a Microporous COF to Improve Cycling Performance in Li-S Batteries. Chem. A Eur. J. 2019, 25, 12394–12404.
  17. Fang, Q.; Wang, J.; Gu, S.; Kaspar, R.B.; Zhuang, Z.; Zheng, J.; Guo, H.; Qiu, S.; Yan, Y. 3D Porous Crystalline Polyimide Covalent Organic Frameworks for Drug Delivery. J. Am. Chem. Soc. 2015, 137, 8352–8355.
  18. Zhang, Y.-B.; Su, J.; Furukawa, H.; Yun, Y.; Gándara, F.; Duong, A.; Zou, X.; Yaghi, O.M. Single-Crystal Structure of a Covalent Organic Framework. J. Am. Chem. Soc. 2013, 135, 16336–16339.
  19. Wang, X.; Bahri, M.; Fu, Z.; Little, M.A.; Liu, L.; Niu, H.; Browning, N.D.; Chong, S.Y.; Chen, L.; Ward, J.W.; et al. A Cubic 3D Covalent Organic Framework with nbo Topology. J. Am. Chem. Soc. 2021, 143, 15011–15016.
  20. Gui, B.; Liu, X.; Cheng, Y.; Zhang, Y.; Chen, P.; He, M.; Sun, J.; Wang, C. Tailoring the Pore Surface of 3D Covalent Organic Frameworks via Post-Synthetic Click Chemistry. Angew. Chem. Int. Ed. 2021, 61, e202113852.
  21. Rowan, S.J.; Cantrill, S.J.; Cousins, G.R.L.; Sanders, J.K.M.; Stoddart, J.F. Dynamic Covalent Chemistry. Angew. Chem. Int. Ed. 2002, 41, 898–952.
  22. Jin, Y.; Yu, C.; Denman, R.J.; Zhang, W. Recent advances in dynamic covalent chemistry. Chem. Soc. Rev. 2013, 42, 6634–6654.
  23. Smith, B.J.; Dichtel, W.R. Mechanistic Studies of Two-Dimensional Covalent Organic Frameworks Rapidly Polymerized from Initially Homogenous Conditions. J. Am. Chem. Soc. 2014, 136, 8783–8789.
  24. Chen, D.; Chen, W.; Xing, G.; Zhang, T.; Chen, L. An Upgraded “Two-in-One” Strategy toward Highly Crystalline Covalent Organic Frameworks Chem. Eur. J. 2020, 26, 8377.
  25. Fischbach, D.M.; Rhoades, G.; Espy, C.; Goldberg, F.; Smith, B.J. Controlling the crystalline structure of imine-linked 3D covalent organic frameworks. Chem. Commun. 2019, 55, 3594–3597.
  26. Li, R.L.; Flanders, N.C.; Evans, A.M.; Ji, W.; Castano, I.; Chen, L.X.; Gianneschi, N.C.; Dichtel, W.R. Controlled growth of imine-linked two-dimensional covalent organic framework nanoparticles. Chem. Sci. 2019, 10, 3796–3801.
  27. Ritchie, L.K.; Trewin, A.; Reguera-Galan, A.; Hasell, T.; Cooper, A.I. Synthesis of COF-5 using microwave irradiation and conventional solvothermal routes. Microporous Mesoporous Mater. 2010, 132, 132–136.
  28. Maschita, J.; Banerjee, T.; Savasci, G.; Haase, F.; Ochsenfeld, C.; Lotsch, B.V. Ionothermal Synthesis of Imide-Linked Covalent Organic Frameworks. Angew. Chem. Int. Ed. 2020, 59, 15750–15758.
  29. Liu, K.; Qi, H.; Dong, R.; Shivhare, R.; Addicoat, M.; Zhang, T.; Sahabudeen, H.; Heine, T.; Mannsfeld, S.; Kaiser, U.; et al. On-water surface synthesis of crystalline, few-layer two-dimensional polymers assisted by surfactant monolayers. Nat. Chem. 2019, 11, 994–1000.
  30. Yang, L.; Guo, Q.; Kang, H.; Chen, R.; Liu, Y.; Wei, D. Self-Controlled Growth of Covalent Organic Frameworks by Repolymerization. Chem. Mater. 2020, 32, 5634–5640.
  31. Liu, X.-H.; Guan, C.-Z.; Ding, S.-Y.; Wang, W.; Yan, H.-J.; Wang, D.; Wan, L.-J. On-Surface Synthesis of Single-Layered Two-Dimensional Covalent Organic Frameworks via Solid–Vapor Interface Reactions. J. Am. Chem. Soc. 2013, 135, 10470–10474.
  32. Zhao, Y.; Guo, L.; Gándara, F.; Ma, Y.; Liu, Z.; Zhu, C.; Lyu, H.; Trickett, C.A.; Kapustin, E.A.; Terasaki, O.; et al. A Synthetic Route for Crystals of Woven Structures, Uniform Nanocrystals, and Thin Films of Imine Covalent Organic Frameworks. J. Am. Chem. Soc. 2017, 139, 13166–13172.
  33. Priya, P.; Nguyen, T.C.; Saxena, A.; Aluru, N.R. Machine Learning Assisted Screening of Two-Dimensional Materials for Water Desalination. ACS Nano 2022, 16, 1929–1939.
  34. Yang, P.; Zhang, H.; Lai, X.; Wang, K.; Yang, Q.; Yu, D. Accelerating the Selection of Covalent Organic Frameworks with Automated Machine Learning. ACS Omega 2021, 6, 17149–17161.
  35. Maschita, J.; Banerjee, T.; Lotsch, B.V. Direct and Linker-Exchange Alcohol-Assisted Hydrothermal Synthesis of Imide-Linked Covalent Organic Frameworks. Chem. Mater. 2022, 34, 2249–2258.
  36. Sick, T.; Rotter, J.M.; Reuter, S.; Kandambeth, S.; Bach, N.N.; Döblinger, M.; Merz, J.; Clark, T.; Marder, T.B.; Bein, T.; et al. Switching on and off Interlayer Correlations and Porosity in 2D Covalent Organic Frameworks. J. Am. Chem. Soc. 2019, 141, 12570–12581.
  37. Uribe-Romo, F.J.; Hunt, J.R.; Furukawa, H.; Klöck, C.; O’Keeffe, M.; Yaghi, O.M. A Crystalline Imine-Linked 3-D Porous Covalent Organic Framework. J. Am. Chem. Soc. 2009, 131, 4570–4571.
  38. García-Arroyo, P.; Martínez-Periñán, E.; Cabrera-Trujillo, J.J.; Salagre, E.; Michel, E.G.; Martínez, J.I.; Lorenzo, E.; Segura, J.L. Pyrenetetraone-based covalent organic framework as an effective electrocatalyst for oxygen reduction reaction. Nano Res. 2022, 15, 3907–3912.
  39. Alonso-Navarro, M.J.; Barrio, J.; Royuela, S.; Karjule, N.; Ramos, M.M.; Martínez, J.I.; Shalom, M.; Segura, J.L. Photocatalytic degradation of organic pollutants through conjugated poly(azomethine) networks based on terthiophene–naphthalimide assemblies. RSC Adv. 2021, 11, 2701–2705.
  40. Xu, H.; Gao, J.; Jiang, D. Stable, crystalline, porous, covalent organic frameworks as a platform for chiral organocatalysts. Nat. Chem. 2015, 7, 905–912.
  41. Wu, C.; Liu, Y.; Liu, H.; Duan, C.; Pan, Q.; Zhu, J.; Hu, F.; Ma, X.; Jiu, T.; Li, Z.; et al. Highly Conjugated Three-Dimensional Covalent Organic Frameworks Based on Spirobifluorene for Perovskite Solar Cell Enhancement. J. Am. Chem. Soc. 2018, 140, 10016–10024.
  42. van der Jagt, R.; Vasileiadis, A.; Veldhuizen, H.; Shao, P.; Feng, X.; Ganapathy, S.; Habisreutinger, N.C.; van der Veen, M.A.; Wang, C.; Wagemaker, M.; et al. Synthesis and Structure–Property Relationships of Polyimide Covalent Organic Frameworks for Carbon Dioxide Capture and (Aqueous) Sodium-Ion Batteries. Chem. Mater. 2021, 33, 818–833.
  43. Huo, J.; Luo, B.; Chen, Y. Crystalline Covalent Organic Frameworks from Triazine Nodes as Porous Adsorbents for Dye Pollutants. ACS Omega 2019, 4, 22504–22513.
  44. Kato, M.; Ota, R.; Endo, T.; Yanase, T.; Nagahama, T.; Shimada, T. Free-Standing Nanometer-Thick Covalent Organic Framework Films for Separating CO2 and N2. ACS Appl. Nano Mater. 2022, 5, 2367–2374.
  45. Jiang, L.; Tian, Y.; Sun, T.; Zhu, Y.; Ren, H.; Zou, X.; Ma, Y.; Meihaus, K.R.; Long, J.R.; Zhu, G. A Crystalline Polyimide Porous Organic Framework for Selective Adsorption of Acetylene over Ethylene. J. Am. Chem. Soc. 2018, 140, 15724–15730.
  46. Royuela, S.; Millán, R.G.-S.; Mancheño, M.J.; Ramos, M.M.; Segura, J.L.; Navarro, J.A.R.; Zamora, F. Catalytically Active Imine-based Covalent Organic Frameworks for Detoxification of Nerve Agent Simulants in Aqueous Media. Materials 2019, 12, 1974.
  47. 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, 5266–5302.
  48. Rodríguez-San-Miguel, D.; Montoro, C.; Zamora, F. Covalent organic framework nanosheets: Preparation, properties and applications. Chem. Soc. Rev. 2020, 49, 2291–2302.
  49. Tao, Y.; Ji, W.; Ding, X.; Han, B.-H. Exfoliated covalent organic framework nanosheets. J. Mater. Chem. A 2021, 9, 7336–7365.
  50. Chandra, S.; Kandambeth, S.; Biswal, B.P.; Lukose, B.; Kunjir, S.M.; Chaudhary, M.; Babarao, R.; Heine, T.; Banerjee, R. Chemically Stable Multilayered Covalent Organic Nanosheets from Covalent Organic Frameworks via Mechanical Delamination. J. Am. Chem. Soc. 2013, 135, 17853–17861.
  51. Yan, Z.; Fang, L.; He, Z.; Xie, H.; Liu, B.; Guo, B.; Yao, Y. Surfactant-Modulated a Highly Sensitive Fluorescent Probe of Fully Conjugated Covalent Organic Nanosheets for Detecting Copper Ions in Aqueous Solution. Small 2022, 18, 2200388.
  52. Mow, R.E.; Lipton, A.S.; Shulda, S.; Gaulding, E.A.; Gennett, T.; Braunecker, W.A. Colloidal three-dimensional covalent organic frameworks and their application as porous liquids. J. Mater. Chem. A 2020, 8, 23455–23462.
  53. Liu, X.; Lim, G.J.; Wang, Y.; Zhang, L.; Mullangi, D.; Wu, Y.; Zhao, D.; Ding, J.; Cheetham, A.K.; Wang, J. Binder-free 3D printing of covalent organic framework (COF) monoliths for CO2 adsorption. Chem. Eng. J. 2020, 403, 126333.
  54. Zhang, M.; Li, L.; Lin, Q.; Tang, M.; Wu, Y.; Ke, C. Hierarchical-Coassembly-Enabled 3D-Printing of Homogeneous and Heterogeneous Covalent Organic Frameworks. J. Am. Chem. Soc. 2019, 141, 5154–5158.
  55. Karak, S.; Dey, K.; Torris, A.; Halder, A.; Bera, S.; Kanheerampockil, F.; Banerjee, R. Inducing Disorder in Order: Hierarchically Porous Covalent Organic Framework Nanostructures for Rapid Removal of Persistent Organic Pollutants. J. Am. Chem. Soc. 2019, 141, 7572–7581.
  56. Martín-Illán, J.; Rodríguez-San-Miguel, D.; Castillo, O.; Beobide, G.; Perez-Carvajal, J.; Imaz, I.; Maspoch, D.; Zamora, F. Macroscopic Ultralight Aerogel Monoliths of Imine-based Covalent Organic Frameworks. Angew. Chem. Int. Ed. 2021, 60, 13969–13977.
  57. Chen, S.; Yuan, B.; Liu, G.; Zhang, D. Electrochemical Sensors Based on Covalent Organic Frameworks: A Critical Review. Front. Chem. 2020, 8, 601044.
  58. Tan, X.; Zhang, Z.; Cao, T.; Zeng, W.; Huang, T.; Zhao, G. Control Assembly of Pillararene-Modified Ag Nanoparticles on Covalent Organic Framework Surface for Enhanced Sensing Performance toward Paraquat. ACS Sustain. Chem. Eng. 2019, 7, 20051–20059.
  59. Liu, X.; Hu, M.; Wang, M.; Song, Y.; Zhou, N.; He, L.; Zhang, Z. Novel nanoarchitecture of Co-MOF-on-TPN-COF hybrid: Ultralowly sensitive bioplatform of electrochemical aptasensor toward ampicillin. Biosens. Bioelectron. 2018, 123, 59–68.
  60. Li, H.; Kou, B.; Yuan, Y.; Chai, Y.; Yuan, R. Porous Fe3O4@COF-Immobilized gold nanoparticles with excellent catalytic performance for sensitive electrochemical detection of ATP. Biosens. Bioelectron. 2021, 197, 113758.
  61. 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.
  62. Bian, G.; Yin, J.; Zhu, J. Recent Advances on Conductive 2D Covalent Organic Frameworks. Small 2021, 17, 2006043.
  63. Sun, L.; Guo, H.; Pan, Z.; Liu, B.; Zhang, T.; Yang, M.; Wu, N.; Zhang, J.; Yang, F.; Yang, W. In-situ reducing platinum nanoparticles on covalent organic framework as a sensitive electrochemical sensor for simultaneous detection of catechol, hydroquinone and resorcinol. Colloids Surf. A Physicochem. Eng. Asp. 2021, 635, 128114.
  64. Guan, Q.; Guo, H.; Wu, N.; Cao, Y.; Wang, M.; Zhang, L.; Yang, W. Highly sensitive determination of acetaminophen and 4-aminophenol based on COF/3D NCNF-T/Au NPs composite electrochemical sensing platform. Colloids Surf. A Physicochem. Eng. Asp. 2021, 630, 127624.
  65. Geng, W.-Y.; Zhang, H.; Luo, Y.-H.; Zhu, X.-G.; Xie, A.-D.; Wang, J.; Zhang, D.-E. Facile fabrication of carbon-loaded covalent-organic framework composites with enhanced electrochemical performance for dopamine determination. Microporous Mesoporous Mater. 2021, 323, 111186.
  66. Wu, N.; Wang, L.; Xie, Y.; Du, Y.; Song, Y.; Wang, L. Double signal ratiometric electrochemical riboflavin sensor based on macroporous carbon/electroactive thionine-contained covalent organic framework. J. Colloid Interface Sci. 2021, 608, 219–226.
  67. Yang, Y.; Shen, Y.; Wang, L.; Song, Y.; Wang, L. Three-dimensional porous carbon/covalent-organic framework films integrated electrode for electrochemical sensors. J. Electroanal. Chem. 2019, 855, 113590.
  68. Liu, Y.; Zhou, M.; Jin, C.; Zeng, J.; Huang, C.; Song, Q.; Song, Y. Preparation of a Sensor Based on Biomass Porous Carbon/Covalent-Organic Frame Composites for Pesticide Residues Detection. Front. Chem. 2020, 8, 643.
  69. Wang, L.; Xie, Y.; Yang, Y.; Liang, H.; Wang, L.; Song, Y. Electroactive Covalent Organic Frameworks/Carbon Nanotubes Composites for Electrochemical Sensing. ACS Appl. Nano Mater. 2020, 3, 1412–1419.
  70. 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.
  71. Xin, Y.; Wang, N.; Wang, C.; Gao, W.; Chen, M.; Liu, N.; Duan, J.; Hou, B. Electrochemical detection of hydroquinone and catechol with covalent organic framework modified carbon paste electrode. J. Electroanal. Chem. 2020, 877, 114530.
  72. Wu, S.; Li, M.; Phan, H.; Wang, D.; Herng, T.S.; Ding, J.; Lu, Z.-G.; Wu, J. Toward Two-Dimensional π-Conjugated Covalent Organic Radical Frameworks. Angew. Chem. Int. Ed. 2018, 57, 8007–8011.
  73. Evans, A.M.; Parent, L.R.; Flanders, N.C.; Bisbey, R.P.; Vitaku, E.; Kirschner, M.S.; Schaller, R.D.; Chen, L.X.; Gianneschi, N.C.; Dichtel, W.R. Seeded growth of single-crystal two-dimensional covalent organic frameworks. Science 2018, 361, 52–57.
  74. Xie, Y.; Wang, N.; Sun, X.; Chu, H.; Wang, Y.; Hu, X. Triple-signaling amplification strategy based electrochemical sensor design: Boosting synergistic catalysis in metal–metalloporphyrin–covalent organic frameworks for sensitive bisphenol A detection. Analyst 2021, 146, 4585–4594.
  75. Liang, H.; Wang, L.; Yang, Y.; Song, Y.; Wang, L. A novel biosensor based on multienzyme microcapsules constructed from covalent-organic framework. Biosens. Bioelectron. 2021, 193, 113553.
  76. Zhou, N.; Ma, Y.; Hu, B.; He, L.; Wang, S.; Zhang, Z.; Lu, S. Construction of hybrid nanostructure: Label-free aptasensor for the ultrasensitive detection of oxytetracycline residues in aqueous solution environments. Biosens. Bioelectron. 2018, 127, 92–100.
  77. Gao, Y.; Zhang, J.; Zhang, X.; Li, J.; Zhang, R.; Song, W. Liposomal Controlled Release Ag-Activated DNAzyme Cycle Amplification on a 2D Pyrene COF-Based Photocathode for α-Synuclein Immunosensing. Anal. Chem. 2021, 93, 8647–8655.
  78. Liang, H.; Ning, G.; Wang, L.; Li, C.; Zheng, J.; Zhao, H.; Li, C.-P. Covalent Framework Particles Modified with MnO2 Nanosheets and Au Nanoparticles as Electrochemical Immunosensors for Human Chorionic Gonadotropin. ACS Appl. Nano Mater. 2021, 4, 4593–4601.
  79. Pang, Y.-H.; Huang, Y.-Y.; Wang, L.; Shen, X.-F.; Wang, Y.-Y. Determination of bisphenol A and bisphenol S by a covalent organic framework electrochemical sensor. Environ. Pollut. 2020, 263, 114616.
  80. Sun, Q.; Fu, C.-W.; Aguila, B.; Perman, J.A.; Wang, S.; Huang, H.-Y.; Xiao, F.-S.; Ma, S. Pore Environment Control and Enhanced Performance of Enzymes Infiltrated in Covalent Organic Frameworks. J. Am. Chem. Soc. 2018, 140, 984–992.
  81. Wang, L.; Liang, H.; Xu, M.; Wang, L.; Xie, Y.; Song, Y. Ratiometric electrochemical biosensing based on double-enzymes loaded on two-dimensional dual-pore COFETTA-TPAL. Sens. Actuators B Chem. 2019, 298, 126859.
  82. Liang, H.; Xu, H.; Zhaoa, Y.; Zhenga, J.; Zhaob, H.; Lic, G.; Lia, C.P. Ultrasensitive electrochemical sensor for prostate specific antigen detection with a phosphorene platform and magnetic covalent organic framework signal amplifier. Biosens. Bioelectron. 2019, 144, 111691.
  83. 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.
  84. Xu, M.; Chen, K.; Zhu, L.; Zhang, S.; Wang, M.; He, L.; Zhang, Z.; Du, M. Heterostructure Hybrid for Dual-Mode Photoelectrochemical–Electrochemical HIV-1 DNA Sensing. Langmuir 2021, 37, 13479–13492.
  85. Keefe, A.D.; Pai, S.; Ellington, A. Aptamers as therapeutics. Nat. Rev. Drug Discov. 2010, 9, 537–550.
  86. Zhu, Q.-Q.; Li, H.-K.; Sun, X.-L.; Han, Z.-Y.; Sun, J.; He, H. Rational incorporation of covalent organic framework/carbon nanotube (COF/CNT) composites for electrochemical aptasensing of ultra-trace atrazine. J. Mater. Chem. C 2021, 9, 8043–8050.
  87. Zhang, T.; Song, Y.; Xing, Y.; Gu, Y.; Yan, X.; Liu, H.; Lu, N.; Xu, H.; Xu, Z.; Zhang, Z.; et al. The synergistic effect of Au-COF nanosheets and artificial peroxidase (NiPd) rhombic dodecahedra for signal amplification for biomarker detection. Nanoscale 2019, 11, 20221–20227.
  88. Huang, Y.-Y.; Pang, Y.-H.; Shen, X.-F.; Jiang, R.; Wang, Y.-Y. Covalent organic framework DQTP modified pencil graphite electrode for simultaneous determination of bisphenol A and bisphenol S. Talanta 2021, 236, 122859.
  89. Xie, Y.; Chen, Y.; Sun, X.; Wang, Y.; Wang, Y. Conducting polymer engineered covalent organic framework as a novel electrochemical amplifier for ultrasensitive detection of acetaminophen. Chin. Chem. Lett. 2020, 32, 2061–2065.
  90. Wang, Q.; Li, R.; Zhao, Y.; Zhe, T.; Bu, T.; Liu, Y.; Sun, X.; Hu, H.; Zhang, M.; Zheng, X.; et al. Surface morphology-controllable magnetic covalent organic frameworks: A novel electrocatalyst for simultaneously high-performance detection of p-nitrophenol and o-nitrophenol. Talanta 2020, 219, 121255.
  91. Zhu, P.; Li, S.; Zhou, S.; Ren, N.; Ge, S.; Zhang, Y.; Wang, Y.; Yu, J. In situ grown COFs on 3D strutted graphene aerogel for electrochemical detection of NO released from living cells. Chem. Eng. J. 2020, 420, 127559.
  92. Han, J.; Pei, L.; Du, Y.; Zhu, Y. Tripolycyanamide-2,4,6-triformyl pyrogallol covalent organic frameworks with many coordination sites for detection and removal of heavy metal ions. J. Ind. Eng. Chem. 2021, 107, 53–60.
  93. Han, J.; Yu, J.; Guo, Y.; Wang, L.; Song, Y. COFBTLP-1/three-dimensional macroporous carbon electrode for simultaneous electrochemical detection of Cd2+, Pb2+, Cu2+ and Hg2+. Sens. Actuators B Chem. 2020, 321, 128498.
  94. Su, D.; Feng, B.; Xu, P.; Zeng, Q.; Shan, B.; Song, Y. Covalent organic frameworks and electron mediator-based open circuit potential biosensor for in vivo electrochemical measurements. Anal. Methods 2018, 10, 4320–4328.
  95. Boyacıoğlu, H.; Yola, B.B.; Karaman, C.; Karaman, O.; Atar, N.; Yola, M.L. A novel electrochemical kidney injury molecule-1 (KIM-1) immunosensor based covalent organic frameworks-gold nanoparticles composite and porous NiCo2S4@CeO2 microspheres: The monitoring of acute kidney injury. Appl. Surf. Sci. 2021, 578, 152093.
  96. Zheng, J.; Zhao, H.; Ning, G.; Sun, W.; Wang, L.; Liang, H.; Xu, H.; He, C.; Li, C.-P. A novel affinity peptide–antibody sandwich electrochemical biosensor for PSA based on the signal amplification of MnO2-functionalized covalent organic framework. Talanta 2021, 233, 122520.
  97. Cheng, J.; Hu, K.; Liu, Q.; Liu, Y.; Yang, H.; Kong, J. Electrochemical ultrasensitive detection of CYFRA21-1 using Ti3C2Tx-MXene as enhancer and covalent organic frameworks as labels. Anal. Bioanal. Chem. 2021, 413, 2543–2551.
More
ScholarVision Creations