Hostname: page-component-848d4c4894-sjtt6 Total loading time: 0 Render date: 2024-07-07T10:38:52.827Z Has data issue: false hasContentIssue false

Protein click chemistry and its potential for medical applications

Published online by Cambridge University Press:  15 April 2024

Ahmad Amiri
Affiliation:
Institute of Biochemistry and Biophysics (IBB), University of Tehran, Tehran, Iran
Sedigheh Abedanzadeh
Affiliation:
Faculty of Chemistry, Kharazmi University, Tehran, Iran
Bagher Davaeil
Affiliation:
Institute of Biochemistry and Biophysics (IBB), University of Tehran, Tehran, Iran
Ahmad Shaabani
Affiliation:
Department of Chemistry, Shahid Beheshti University, Tehran, Iran
Ali A. Moosavi-Movahedi*
Affiliation:
Institute of Biochemistry and Biophysics (IBB), University of Tehran, Tehran, Iran
*
Corresponding author: Ali A. Moosavi-Movahedi; Email: [email protected]
Rights & Permissions [Opens in a new window]

Abstract

A revolution in chemical biology occurred with the introduction of click chemistry. Click chemistry plays an important role in protein chemistry modifications, providing specific, sensitive, rapid, and easy-to-handle methods. Under physiological conditions, click chemistry often overlaps with bioorthogonal chemistry, defined as reactions that occur rapidly and selectively without interfering with biological processes. Click chemistry is used for the posttranslational modification of proteins based on covalent bond formations. With the contribution of click reactions, selective modification of proteins would be developed, representing an alternative to other technologies in preparing new proteins or enzymes for studying specific protein functions in different biological processes. Click-modified proteins have potential in diverse applications such as imaging, labeling, sensing, drug design, and enzyme technology. Due to the promising role of proteins in disease diagnosis and therapy, this review aims to highlight the growing applications of click strategies in protein chemistry over the last two decades, with a special emphasis on medicinal applications.

Type
Review
Copyright
© The Author(s), 2024. Published by Cambridge University Press

Introduction

Chemical modification of proteins has become a valuable tool for developing modified proteins. Playing complementary roles to genetic techniques, we have a broad toolkit that allows us to create an almost unlimited number of protein constructs with natural or synthetically modified residues using chemical modifications (Stephanopoulos and Francis, Reference Stephanopoulos and Francis2011). The ideal requirements for such reactions include functional group tolerance/compatibility, water as a reaction medium, selectivity, high reaction rates, neutral pH and room temperature (or up to 40 °C), nontoxic reagents, and low reactant concentrations. Reactions must be designed and implemented to achieve high modification efficiencies without the need for tedious and inefficient purification/characterization protocols. For in vivo studies, chemical modification methods involving those listed above are appropriate since they do not interfere with normal cell function (Boutureira and Bernardes, Reference Boutureira and Bernardes2015). Protein modifications have a significant impact on signaling, migration, differentiation, and trafficking as important cellular processes through sulfation, phosphorylation, methylation, acylation, ubiquitination, glycosylation, farnesylation, and so on (Walsh et al., Reference Walsh, Garneau-Tsodikova and Gatto2005).

Posttranslational protein modifications (PTMs) are commonly thought to be responsible for the vast biodiversity found in nature today (Boutureira and Bernardes, Reference Boutureira and Bernardes2015). These modifications usually occur after protein translation. In this regard, considering the characteristics of the natural modification of proteins, it can be concluded that the efficient and controlled reproducing of PTM provides a valuable tool in the study and precise function of proteins. In addition to the facilities provided by (bio)orthogonal methods; it allows precise and site-selective modification of proteins, making it a valuable tool for in vivo and in vitro studies (Bernardes et al., Reference Bernardes, Chalker, Davis and Pignataro2010; Kee and Muir, Reference Kee and Muir2012). Considering the various methods used for chemical modification, it will now be feasible to choose the target residue and modification type to provide the required property/function such as chemical affinity probes, fluorophores, reactive tags, and so forth. Despite the enormous progress made in the bioconjugation of proteins, there are still many serious challenges, not only from a synthetic point of view but also in manufacturing, processing, stability, and safety. Some types of proteins can be modified using methods that are not appropriate for all kinds of proteins (Boutureira and Bernardes, Reference Boutureira and Bernardes2015). Therefore, there is still a need for the development of complementary reactions for site-selective modifications of proteins that are efficient, robust, and mild. The various aspects of protein synthesis have been discussed in detail (Bernardes et al., Reference Bernardes, Chalker, Davis and Pignataro2010), from general chemical ligation strategies (Hackenberger and Schwarzer, Reference Hackenberger and Schwarzer2008; Kent, Reference Kent2009; Siman and Brik, Reference Siman and Brik2012), endogenous amino acid modification (Baslé et al., Reference Baslé, Joubert and Pucheault2010), to click modification protocols (Lallana et al., Reference Lallana, Riguera and Fernandez-MEGIA2011; Van Berkel et al., Reference Van Berkel, Van Eldijk and Van Hest2011; Palomo, Reference Palomo2012; Tasdelen and Yagci, Reference Tasdelen and Yagci2013), which are more specialized on specific PTMs, including glycosylation (Gamblin et al., Reference Gamblin, Scanlan and Davis2009; Schmaltz et al., Reference Schmaltz, Hanson and Wong2011; Villalonga et al., Reference Villalonga, Diez, Sanchez, Gamella, Pingarron and Villalonga2014; Wang and Amin, Reference Wang and Amin2014), PEGylation (Nischan and Hackenberger, Reference Nischan and Hackenberger2014), and polymerization of protein-based initiator (Matyjaszewski and Tsarevsky, Reference Matyjaszewski and Tsarevsky2014; Wallat et al., Reference Wallat, Rose and Pokorski2014). However, most coupling methods are inspired by naturally occurring reactions or engineered protein-catalyzed reactions (Lange and Polizzi, Reference Lange and Polizzi2021). Based on biomimetic research studies, there is a general attraction toward carbon–heteroatom bond formation over carbon–carbon bonds; for instance, proteins, polysaccharides, and nucleic acids are condensed polymeric assemblies of subunits bound by carbon-heteroatom (Kolb and Sharpless, Reference Kolb and Sharpless2003).

With the emergence of click chemistry, a new promising line of chemical protein modification has been developed. Click reactions were first introduced by Kolb et al. (Reference Kolb, Finn and Sharpless2001). The basis of this method is the simple assembly of molecules. These reactions are based on covalent bonds that can be used in various applications (Kolb et al., Reference Kolb, Finn and Sharpless2001). Caroline Bertozzi published a new click reaction in 2004, called strain-promoted azide-alkyne cycloaddition (SPAAC). This achievement opened a new field in click chemistry (Agard et al., Reference Agard, Prescher and Bertozzi2004). This method has simplified the use of the click method in biological studies by eliminating the need for copper. In 2022, the Nobel Prize in Chemistry was awarded to three influential scientists, Barry Sharpless, Morten Meldahl, and Caroline Bertozzi, who were pioneered the field of click reactions. The click strategy is a powerful tool to produce new molecules and can be summarized by the following statement: ˝All searches must be restricted to molecules that are easy to make˝ (Kolb et al., Reference Kolb, Finn and Sharpless2001). Click reactions are modular, applicable reactions that give very high yields, produce no unpleasant by-products, and require simple reaction conditions (Shaabani et al., Reference Shaabani, Maleki and Mofakham2008; Shaabani et al., Reference Shaabani, Afshari and Hooshmand2017a; Shaabani et al., Reference Shaabani, Tavousi Tabatabaei and Shaabani2017b; Shaabani et al., Reference Shaabani, Shadi, Mohammadian, Javanbakht, Nazeri and Bahri2019; Khodkari et al., Reference Khodkari, Nazeri, Javanbakht and Shaabani2023). When there is a need for new molecular properties, small molecular building blocks can be joined together to produce such properties (Suazo et al., Reference Suazo, Park and Distefano2021). As shown in Figure 1, there has been an increase in the number of publications on the topic of click protein chemistry with biological applications in recent years. Click protein chemistry has given rise to a myriad of highly interesting developments in the field of medicine and has significantly impacted the synthesis of structurally diverse molecules through shorter, stereoselective, and efficient synthetic routes. In continuation of our previous review articles on modified/functionalized materials via multicomponent reactions strategy as “click” reactions (Afshari and Shaabani, Reference Afshari and Shaabani2018; Javanbakht and Shaabani, Reference Javanbakht and Shaabani2019; Javanbakht et al., Reference Javanbakht, Nasiriani, Farhid, Nazeri and Shaabani2022), herein, an attempt has been made to review the medicinal applications of protein click chemistry.

Figure 1. The number of published papers in the field of protein click chemistry with biological applications. (The method of extraction is fully described in the Supplementary Material.)

Bioorthogonal click chemistry

The widespread use of straightforward click reactions in biological environments led to the advent of new definition “bio-click chemistry” which can be understood toward reactions of functional groups commonly found in biological molecules like proteins or live cells and by exploiting the potential of click reactions (Rodríguez et al., Reference Rodríguez, Moglie, Ramírez-Sarmiento, Singh, Dua and Zacconi2022). There is a close overlap between click and bioorthogonal chemistry. Consequently, click chemistry tools will enable the development of bioorthogonal reactions, while technological advancements in chemical biology will stimulate the possibility of developing novel click chemistry methods (Suazo et al., Reference Suazo, Park and Distefano2021). Bioorthogonal reactions and protein labeling in complex cellular mixtures have also been studied in several types of research over the past decade (Lang and Chin, Reference Lang and Chin2014; Patterson et al., Reference Patterson, Nazarova and prescher2014). In another definition, bioorthogonal chemistry represents high-yielding rapid and selective chemical reactions that proceed in biological environments without any by-products (Scinto et al., Reference Scinto, Bilodeau, Hincapie, Lee, Nguyen, Xu, Am Ende, Finn, Lang and Lin2021). The significant advantage of bioorthogonal approaches is related to the fact that they usually do not affect other normal biochemical processes (Sletten and Bertozzi, Reference Sletten and Bertozzi2009). For example, bioorthogonal click chemistry has enabled selective labeling of enzymatic processes in vitro and in vivo, allowing real-time analysis of enzymatic processes in both environments. In this way, we have better understood about many biomedical challenges and biological questions. These include the causes of Alzheimer’s disease, Coronavirus, and Cancer, among others (Suazo et al., Reference Suazo, Park and Distefano2021).

Different amino acids are used in protein click reactions, including histidine, lysine, cysteine, and so forth. Besides the direct use of amino acids in the protein click reactions, many reports suggest that the amino acids can be modified before the click reaction. For instance, the use of iodoacetamide can be used to modify nucleophilic amino acids such as cysteine, lysine, and histidine nonspecifically. It has also been noted that activated carboxylic acid derivatives, such as N-hydroxysuccinimide esters, have been extensively used to modify the ε-amine of lysine residues in peptides and proteins (Fisher et al., Reference Fisher, Baker and Shoichet2017). With the aim of modifying proteins, it should be considered that the bioactivity of the protein may be affected if the modification disrupts the active site. Currently, modification of site-specific proteins with unnatural amino acids while maintaining protein structure and function is very popular (Raliski et al., Reference Raliski, Howard and Young2014; Maza et al., Reference Maza, McKenna, Raliski, Freedman and Young2015). Click chemistry involves a range of important reactions in synthetic organic chemistry, including strained ring opening, conjugate addition, aldehyde capture by α-effect nucleophiles, cycloaddition, and acylation/sulfonylation (Devaraj and Finn, Reference Devaraj and Finn2021). Figure 2 presents various types of click reactions. Therefore, click methods are among the most efficient strategies for protein modification, binding the protein to a new substrate or another protein.

Figure 2. The most recently used click reactions in protein conjugations.

Thiol-alkene (thiol-ene) and thiol-alkyne (thiol-yne) couplings have attracted great attention in recent years because they utilize readily available chemical functions found in the structure of proteins, peptides, polymers, and materials (Campos et al., Reference Campos, Killops, Sakai, Paulusse, Damiron, Drockenmuller, Messmore and Hawker2008; Hensarling et al., Reference Hensarling, Doughty, Chan and Patton2009; Hoogenboom, Reference Hoogenboom2010). An advantage of this method is that it requires only low concentrations of initiators, typically proceeds rapidly, and can be easily isolated from the products (Colak et al., Reference Colak, Da Silva, Soares and Gautrot2016). In general, the kinetics of a reaction is strongly influenced by the structure of the alkene moiety. A strained alkene moiety, as well as an electron-rich alkene moiety, will have the greatest reactivity (Reddy et al., Reference Reddy, Cramer and Bowman2006; Hoyle and Bowman, Reference Hoyle and Bowman2010). The thiol-Michael addition reaction has been widely applied for protein conjugation because it can work under mild conditions in an aqueous environment (Jones et al., Reference Jones, Mantovani, Ryan, Wang, Brayden and Haddleton2009). While UV irradiation or photoinitiators may be harmful to cells or materials, it is often preferable to use thiol-Michael addition instead of thiol-ene reactions (Fisher et al., Reference Fisher, Baker and Shoichet2017). The Cu-catalyzed azide-alkyne cycloaddition (CuAAC) reaction is used in pharmacy and chemical biology as an efficient method for the covalent modification of active and proactive biological molecules (McKay and Finn, Reference McKay and Finn2014). CuAAC offers a unique advantage over other non-Cu click technologies. The small triazole linkage is identical in size and polarity to the peptide linkage, minimizing disruption of the biological function of the conjugate (Valverde et al., Reference Valverde, Lecaille, Lalmanach, Aucagne and Delmas2012; Birts et al., Reference Birts, Sanzone, El-Sagheer, Blaydes, Brown and Tavassoli2014). However, there is a significant drawback in using CuAAC reactions for bioconjugations related to the presence of Cu, which generates reactive oxygen species (Thirumurugan et al., Reference Thirumurugan, Matosiuk and Jozwiak2013). Because of the oxidation of Cu(I) to Cu(II), it is often necessary to carry out the reaction under inert gas or to use reducing agents such as ascorbic acid or sodium ascorbate (Rostovtsev et al., Reference Rostovtsev, Green, Fokin and Sharpless2002). In contrast to CuAAC, SPAAC reaction does not require a metal catalyst. A cycloalkyne’s electronic structure strongly influences the reaction rate and efficiency (Fisher et al., Reference Fisher, Baker and Shoichet2017). Despite SPAAC is effectively applied for conjugating proteins and peptides, strained cyclooctyin exhibits high reactivity, resulting in poor stability and difficult synthetic processes (Jewett et al., Reference Jewett, Sletten and Bertozzi2010). In a Diels–Alder reaction, electron-rich dienophiles react with electron-poor dienophiles to form a [4 + 2] cyclization reaction (Fisher et al., Reference Fisher, Baker and Shoichet2017). It is important to remember that the Diels–Alder reaction is a highly selective transformation and can proceed more rapidly and selectively in water than in organic solvents (Li and Chan, Reference Li and Chan1997). In general, the Diels–Alder reaction is quite slow; however, when heated to higher temperatures, the reaction rate can be increased. However, temperatures above 37 °C should be avoided because of the possibility of protein denaturation, while at higher temperatures the Diels–Alder reaction is reversible (Koehler et al., Reference Koehler, Anseth and Bowman2013). Binding of the C-terminal part of a thioester to the N-terminal part of a cysteine residue provides the formation of amide moiety in the thioester-amine method (native chemical ligation (NCL)) (Dawson et al., Reference Dawson, Muir, Clark-Lewis and Kent1994). NCL allows peptides larger than 50 amino acids to be constructed by synthesizing the peptide fragments and connecting them together. This method overcomes the limitations of the solid-phase peptide synthetic method (Dawson and Kent, Reference Dawson and Kent2000). The NCL procedure is performed in aqueous solutions, at neutral pH levels, in the presence of denaturing agents such as guanidine hydrochloride to prevent protein aggregation. This reaction is very sensitive to the pH value of the system, whereas high pH values can hydrolyze thioesters, and low pH values can reduce the reactivity of cysteine thiol amine, decreasing the rate of the reaction (Fisher et al., Reference Fisher, Baker and Shoichet2017). The Bertozzi group developed Staudinger ligation for cell surface modification to form amide bonds between azides and triarylphosphine derivatives, in which cells with azide moieties reacted with phosphines to generate amide bonds (Saxon and Bertozzi, Reference Saxon and Bertozzi2000). Proteins can be conjugated to polymers or immobilized on the glass or gold surfaces through this reaction (Fisher et al., Reference Fisher, Baker and Shoichet2017). Nucleophilic oxyamines attack the electron-deficient aldehydes or ketones in the oxime ligation reaction, resulting in the formation of an oxime bond, with water molecules being formed as a by-product of the reaction. It is a highly selective bioorthogonal process that has almost quantitative conversions, proceeds under mildly acidic aqueous conditions, and does not require metal catalysts (Fisher et al., Reference Fisher, Baker and Shoichet2017).

In summary, click reactions play notable roles in the design of chemical motifs and help to construct bioorthogonal covalent conjugations under physiological conditions. Considering the extent of click chemistry, new strategies have emerged to improve target specificity and increase efficacy in diagnosing and treating disease. With this background, we highlight a range of medicinal applications that involve the bioorthogonal chemistry toolbox.

Diagnosis

Protein labeling

There are tens of thousands of proteins in the human proteome, many of them occurring in minute concentrations below the limits of detection (LODs) of current technologies such as ELISA, mass spectrometry (MS), and protein microarrays (Wilson, Reference Wilson2013). It is imperative to develop a molecular instrument that can detect disease-related protein biomarkers at low levels in the body without the need for any further manipulation (Senapati et al., Reference Senapati, Manna, Lindsay and Zhang2013).

Imaging is a method of distinguishing the target biomolecule from a living system. This is possible through the use of a spectroscopic probe (Baskin et al., Reference Baskin, Prescher, Laughlin, Agard, Chang, Miller, Lo, Codelli and Bertozzi2007). If less structural disruption is desired, the target protein can be chemically (Carrico et al., Reference Carrico, Carlson and Bertozzi2007) or enzymatically (Chen and Ting, Reference Chen and Ting2005) labeled with small molecules. There seems to be a growing interest in imaging biomolecules that are not easily modified by genetic modification, even though proteins remain the primary targets of cellular imaging (Carrico et al., Reference Carrico, Carlson and Bertozzi2007). For example, because lipids and glycans have independent functions and these molecules are the result of PTM, they cannot be imaging using protein-specific methods. Although these biomolecules have been studied in vitro in static systems, the dynamic behavior of these biomolecules in living cells is poorly understood (Carrico et al., Reference Carrico, Carlson and Bertozzi2007). Therefore, in vivo biomolecular labeling can be achieved without genetic manipulation using a bioorthogonal click strategy (Carrico et al., Reference Carrico, Carlson and Bertozzi2007). In this regard, bioorthogonal functional groups are installed in target biomolecules by a cell’s metabolic machinery. The next step involves the covalent labeling of a probe to the functional group (Prescher and Bertozzi, Reference Prescher and Bertozzi2005).

There are various conventional methods for click chemistry, depending on the type of chemical protein modification. Combining photoaffinity labeling with liquid chromatography (LC)–MS/MS method can effectively identify cognate proteins in biological systems (Preston and Wilson, Reference Preston and Wilson2013). Because these analyses are based on comparing the signal strength of unlabeled peptides, small MS signals at the LOD are still difficult to distinguish from false positives. However, the multifunctional cross-linkers with highly sensitive functions can be applied as alternatives that can effectively distinguish labeled peptides from various contaminants (He et al., Reference He, Xie, Yang, Zhang, Su, Ge, Song and Chen2017; Horne et al., Reference Horne, Walko, Calabrese, Levenstein, Brockwell, Kapur, Wilson and Radford2018). Hence, the sulfo-click reaction can simplify the synthesis of photoprobes of biomolecules with different functional groups. In addition, highly controlled cleavage of the N-acylsulfonamide linkage significantly improves the handling of the labeled proteins. A synthetic advantage makes it possible to determine the specific direction in which ligands bind to the surface of the interacting proteins. Another advantage of this method is that it can be combined with a cross-linking photocyclization reaction to produce a special MS tag whose mass changes under the influence of irradiation, leading to an understandable distinction between labeled peptides and false positive signals. Thanks to these improvements, target proteins and interacting sites of biomolecules can be identified more quickly in practical applications (Hayashi et al., Reference Hayashi, Morimoto and Tomohiro2019).

Through a metal- and oxidant-free synthetic approach, a monosubstituted 3 bromo-1,2,4,5-tetrazine fragment (Tz) was synthesized. Since it has superb reactivity toward different nucleophiles, its application as a biomolecule labeling agent was investigated, especially as a simple and straightforward way to incorporate a bioorthogonal handle into proteins. Based on the results of LC–MS/MS analyses, this compound appears to be chemoselective for lysines and forms monosubstituted amino-Tz on the surface of the protein with preserved folding. The chemical properties of the labeled lysines were investigated for click-to-release (CtR) reactions. They were additionally can be used in cell culture as a therapeutically relevant context (Figure 3) (Ros et al., Reference Ros, Bellido, Verdaguer, Ribas De Pouplana and Riera2020).

Figure 3. Targeted drug release via the click-to-release reaction using 3-bromo-1,2,4,5-tetrazine (Tz) for protein labeling. A: Activation of circulating inactive trans-cyclooct-2-en-1-yl (TCO)-drug conjugate by monosubstituted amino Tz reaction. B: Labeling of Trastuzumab using 3-bromo-1,2,4,5-tetrazine. C: The reaction of labeled Trastuzumab and TCO-Dox in BT474 (HER2+) cell culture to release drug (Ros et al., Reference Ros, Bellido, Verdaguer, Ribas De Pouplana and Riera2020).

It has been reported that nitroso species can be selectively incorporated into a number of proteins, including lysozyme, BSA, KRAS, MiaA, and HRAS, using lysine residues as targets. Under physiological conditions, it was observed that the corresponding azo functionalities were formed in a highly selective manner with excellent yields, displaying rather good stability. It was found that a fluorescent and/or dual fluorescent labeling protocol could be efficiently and selectively applied for the preparation of different labeled proteins, including HRAS, KRAS, and lysozyme. Consequently, the interactions between prenylated proteins and enzymes were evaluated through fluorescence resonance energy transfer (FRET) assays. By incorporating pyrenyl functionality into specific proteins at lysine residues, established via the so-called NEL process, along with the Selectfluor and click reaction approach targeting prenyl functionality, chemically modified proteins with a 1-pyrenyl-fluorophore by 254 nm UV irradiation were interestingly synthesized. The sequential azidation and click reaction of the protein prenyl functionality allow the incorporation of naphthene to increase the transmitted fluorescence energy. Moreover, significantly increased absorbance at 218 nm in lysed HEK293T cells and greatly enhanced greenish fluorescence in live HEK293T cells were observed (Gan et al., Reference Gan, Li, Zhou and Wang2022). This was the first report of using a chemical approach to figure out protein–protein interactions through the FRET assay.

A bioorthogonal reaction for dynamic cellular imaging was developed by combining the biocompatibility of Staudinger ligation with the rapid reaction kinetics of the click strategy. The development of an alternative method to activate alkynes for [3 + 2]-cycloaddition with azides would be one approach to achieve this goal (Carrico et al., Reference Carrico, Carlson and Bertozzi2007). However, cyclooctynes as the smallest of the stable cycloalkynes, have achieved bioorthogonal azide labeling by using ring strain (Agard et al., Reference Agard, Prescher and Bertozzi2004). Even though the SPAAC reaction showed no better sensitivity than that of the Staudinger ligation, the cyclooctyne probes demonstrated no cellular cytotoxicity (Agard et al., Reference Agard, Baskin, Prescher, Lo and Bertozzi2006). By means of ring strain and electron-withdrawing groups as two important rate-accelerating factors, the sensitivity of cyclooctynes for azide detection can be increased in Cu-free click reagent design (Baskin et al., Reference Baskin, Prescher, Laughlin, Agard, Chang, Miller, Lo, Codelli and Bertozzi2007). It is expected that azide labeling and Cu-free click chemistry will find numerous applications in glycobiology. Cu-free clicks using difluorinated cyclooctyne have been demonstrated to proceed selectively inside living mice (Baskin et al., Reference Baskin, Prescher, Laughlin, Agard, Chang, Miller, Lo, Codelli and Bertozzi2007). In addition, Cu-free click chemistry can be used to characterize other metabolites and enzyme activities (Speers and Cravatt, Reference Speers and Cravatt2004), PTMs (Prescher and Bertozzi, Reference Prescher and Bertozzi2005), and site-specifically labeled proteins (Chin et al., Reference Chin, Cropp, Anderson, Mukherji, Zhang and Schultz2003; Chen and Ting, Reference Chen and Ting2005) in living systems.

Protein biomarkers can be detected using affinity molecules attached to atomic force microscopy (AFM) tips using catalyst-free click reactions. Molecular recognition force spectroscopy (MRFS), a technique based on the use of AFM measurements, is being developed to determine and characterize the interactions between antibodies and antigens, ligands and receptors, DNA probes and targets, and so forth, at the single-molecule level (Florin et al., Reference Florin, Moy and Gaub1994; Dammer et al., Reference Dammer, Hegner, Anselmetti, Wagner, Dreier, Huber and Güntherodt1996; Avci et al., Reference Avci, Schweitzer, Boyd, Wittmeyer, Steele, Toporski, Beech, Arce, Spangler and Cole2004; Neuert et al., Reference Neuert, Albrecht, Pamir and Gaub2006; Carvalho et al., Reference Carvalho, Connell, Miltenberger-Miltenyi, Pereira, Tavares, Ariëns and Santos2010; Meng et al., Reference Meng, Paetzell, Bogorad and Soboyejo2010; Zapotoczny et al., Reference Zapotoczny, Biedroń, Marcinkiewicz and Nowakowska2012). AFM, with its single-molecule sensitivity, is a candidate for nanodiagnostics (Archakov and Ivanov, Reference Archakov and Ivanov2007). In combination with irreversible binding, it has been shown that AFM can reach a concentration sensitivity limit of 10−17 M (Archakov et al., Reference Archakov, Ivanov, Lisitsa and Zgoda2007). AFM is being used in DNA, protein, and cell analysis, and its chemical sensibility has also been greatly enhanced. When the tip of the AFM is coated with an affinity molecule, it can see and count the target molecules (Senapati et al., Reference Senapati, Manna, Lindsay and Zhang2013). With an affinity molecule tethered to the tip, AFM can scan individual proteins on a surface, called recognition imaging (RI) (Stroh et al., Reference Stroh, Wang, Bash, Ashcroft, Nelson, Gruber, Lohr, Lindsay and Hinterdorfer2004; Lin et al., Reference Lin, Wang, Liu, Yan and Lindsay2006; Wang et al., Reference Wang, Dalal, Henikoff and Lindsay2008; Chtcheglova and Hinterdorfer, Reference Chtcheglova and Hinterdorfer2011; Wang et al., Reference Wang, Guo, Zhang, Park and Xu2012). Several protein biomarkers can be identified and detected by MRFS and RI in the clinical setting. A new method combining two orthogonal click chemistries, thiol-vinyl sulfone-Michael addition and catalyst-free azide-alkyne cycloaddition, was developed for affinity molecule attachment to AFM tips for force spectroscopy and RI. All reactions were carried out in aqueous solutions. It is noteworthy that this method is used for both AFM-based force measurement and detection imaging. In addition, no specific reaction conditions are required for the attachment process (Senapati et al., Reference Senapati, Manna, Lindsay and Zhang2013).

By using positron emission tomography (PET) and single photon emission computed tomography techniques, scientists have the ability to take imaging of blood pools (Kumar and Boddeti, Reference Kumar, Boddeti, Baum and Rösch2013). Imaging plays an important role in the diagnosis of cardiovascular function evaluation (Millar et al., Reference Millar, Hannan, Sapru and Muir1979; Nishimura et al., Reference Nishimura, Hamada, Hayashida, Uehara, Katabuchi and Hayashi1989), a bleeding gastrointestinal tract (Grady, Reference Grady2016), and testing cancerous tissues for vascular permeability (Niu et al., Reference Niu, Lang, Kiesewetter, Ma, Sun, Guo, Guo, Wu and Chen2014). Radiolabeling of human serum albumin (HSA) with 99mTc using the chelate-then-click SPAAC method was developed and optimized to make an agent that can be used to image blood pools. Under mild reaction conditions, 99mTc and HSA are paired in excellent radiochemical yields. When compared with commercially available conventional 99mTc-HSA, 99mTc-DPA-HSA, 2,2′-dipicolylamine (DPA) as chelator agent, showed a high degree of stability in vivo, indicating higher blood retention and improved visualization of vasculatures in healthy mice for up to 3 hours following injection. Accordingly, the present radiolabeled method for the detection of biomolecules with similar properties can also be applied to other biomolecules with similar sensitivity (Figure 4) (Lodhi et al., Reference Lodhi, Park, Kim, Kim, Shin, Lee, Im, Jeong, Khalid and Cheon2019).

Figure 4. Radiolabeling of human serum albumin (HSA) with 99mTc via chelate-then-click strain-promoted azide-alkyne cycloaddition (SPAAC) approach (Lodhi et al., Reference Lodhi, Park, Kim, Kim, Shin, Lee, Im, Jeong, Khalid and Cheon2019).

Specific functional group labeling in living cells is one of the most potent applications of bioorthogonal couplings. Tetrazines, for example, irreversibly form dihydropyrazine products and dinitrogen by reacting with a strained dienophile norbornene, which is a rapid, selective, and high-yield reaction in aqueous media. A norbornene-modified monoclonal antibody was used in human breast cancer cells to target Her2/neu receptors. In the presence of serum, tetrazines conjugated to a near-infrared fluorochrome label the pre-targeted antibody selectively and rapidly in the presence of the pre-targeted antibody. Based on these findings, this chemistry may be useful for pre-targeted imaging in vivo under numerous modalities under in vitro labeling experiments (Figure 5) (Devaraj et al., Reference Devaraj, Weissleder and Hilderbrand2008).

Figure 5. Pretargeting of SKBR3 cells with norbornene and tetramethylrhodamine co-labeled trastuzumab. Tagging the live cells with tetrazine-VT680 via an inverse electron demand using Diels–Alder coupling technique (Devaraj et al., Reference Devaraj, Weissleder and Hilderbrand2008).

A series of novel heterobifunctional linkers have been developed based on substituted dimethylmaleic anhydride. This hybrid linker combines the advantages of click chemistry with pH-sensitive binding between conjugated biomolecules that can be used in various applications. Due to the acidic environment of tumors or early endosomes (Murphy et al., Reference Murphy, Powers and Cantor1984; Engin et al., Reference Engin, Leeper, Cater, Thistlethwaite, Tupchong and McFarlane1995), the pH labiality of linkers to the acidic environment is helpful. A further advantage of this method is that the linker is cleaved off traceless, yielding an unmodified molecule of interest. The versatility of the technique in protein modification is illustrated in examples such as the reversible dye labeling of proteins, pH-sensitive modification with polyethylene glycol (PEG), and the intracellular transduction of proteins in bioactive form (Maier and Wagner, Reference Maier and Wagner2012).

A biaryl-linker probe produced by the click strategy has a unique T-shaped conformation that significantly enhances its labeling performance. A new generation of biaryl-linker probes made by click reactions makes activity-based protein profiling analysis of trace membrane targets a powerful tool for identifying bioactive compounds (Nakamura et al., Reference Nakamura, Inomata, Ebine, Manabe, Iwakura and Ueda2010).

PET is widely acknowledged as a robust and noninvasive molecular imaging technique due to its unique properties. By utilizing PET, valuable functional insights into various physiological, biochemical, and pharmacological processes can be obtained (Phelps, Reference Phelps2000). The expanding array of novel targets utilized for PET imaging has led to the advent of new advanced radiolabeling techniques, particularly for high-weight molecules like peptides, proteins, antibodies, and antibody fragments. The short-lived positron emitter fluorine-18 (18F) has proven to be beneficial in designing and synthesizing PET radiotracers, thanks to its advantageous nuclear and chemical properties. Nevertheless, incorporating 18F into high-molecular-weight compounds, such as proteins, presents a significant challenge (Ramenda et al., Reference Ramenda, Kniess, Bergmann, Steinbach and Wuest2009). In this regard, a Cu(I)-mediated 1,3-dipolar [3 + 2] cycloaddition reaction was employed to label azide-functionalized HSA with the short-lived positron emitter 18F. This groundbreaking achievement expands the potential of click chemistry as a versatile tool for a wide range of radiolabeling reactions in future applications. The study showcases the promising application of click chemistry in the field of molecular imaging (Ramenda et al., Reference Ramenda, Kniess, Bergmann, Steinbach and Wuest2009). It highlights its importance in developing innovative radiotracers for noninvasive PET imaging of high-molecular-weight compounds like proteins.

In another study, a new pyrene-based excimer fluorescence labeling technique was developed. Through a thiol-ene click strategy, biological samples were linked to an excimer fluorescent precursor (EFP). Thiol groups of proteins in reduced cells and tissues could be successfully labeled with radicals generated by UV light or electron beams. Correlative light/electron microscopy (CLEM) was used to observe cells or tissues with atmospheric scanning electron microscope (SEM). The labeling reaction was induced by the electron beam from an inverted SEM, while the subsequent excimer fluorescence of the samples was observed using fluorescence microscopy. Because the labeling process is limited to the specific area scanned by the electron beam, there is potential for space-limited labeling within a narrow electron beam trajectory, which results in high resolution in the imaging process. The presented technique can be helpful in various biological analyses such as oxidative stress sensing and protein folding. EFP-based CLEM techniques together with protein dynamics studies hold immense potential to improve our understanding of diseases and contribute to the development of clinical applications, including X-ray radiodynamic therapy for cancer (Naya and Sato, Reference Naya and Sato2020).

In general, bioorthogonal click chemistry has dramatically affected the field of radiochemistry and molecular imaging via making radiotracers with specific functions more feasible, accessible, and efficient (Zhong et al., Reference Zhong, Yan, Ding, Su, Xu and Yang2023).

Sensor technology

Absolutely, the assembly of proteins at the electrode surfaces presents an exciting opportunity to harness the diverse capabilities of biology and electronics in a synergistic manner. By combining the molecular recognition properties of proteins with the signal transduction capabilities of electronics, bioelectronic devices can be developed with a wide range of potential applications. For example, the successful development of a device to monitor blood glucose levels shows promise for the broader application of multiplex biosensors in wearable devices. This achievement suggests that this technology can be extended to other essential functions, including on-site contamination detection and point-of-care disease diagnosis (Shi et al., Reference Shi, Qiu, Nie, Xiao, Payne and Du2013). The surface-immobilized biological material acts as a connector that interacts with the target analyte to produce a measurable signal. For practical applications, the biosensor design focuses on miniaturization and portability. By optimizing the size of the biosensor, it becomes more practical and can be easily integrated into various systems for different applications.

In recent years, there has been a significant interest in the field of glycobiology, which involves the study of carbohydrates and glycoconjugates (glycans) in biological systems and their intricate interactions with proteins, cells, and other biomolecules (Norberg et al., Reference Norberg, Deng, Aastrup, Yan and Ramstrom2011). The exploration of the complex functions of glycans is restricted as a result of a number of challenges, including the synthesis and/or purification of glycans, as well as analyzing the biomolecular interactions (Gabius, Reference Gabius2011). To achieve robust and efficient high-throughput analysis, several powerful glycan array formats have been developed using more or less specific chemical ligation techniques to selective immobilization of glycans (Norberg et al., Reference Norberg, Deng, Aastrup, Yan and Ramstrom2011). In this regard, the photoclick surface functionalization method has been suggested as a general polymeric system for protein–carbohydrate interactions (Norberg et al., Reference Norberg, Deng, Yan and Ramstrom2009; Norberg et al., Reference Norberg, Deng, Aastrup, Yan and Ramstrom2011). A panel of carbohydrate structures were applied to fabricate lectin-binding evaluating sensors using a quartz crystal microbalances setup that was possible to efficiently demonstrate protein binding (Norberg et al., Reference Norberg, Deng, Aastrup, Yan and Ramstrom2011).

The availability of simple and powerful detection technologies for protein biomarkers is essential for disease diagnosis. ELISAs, Western blots, and antibody-based enzyme-linked immunosorbent assays are commonly used methods to detect proteins. However, these traditional techniques are time-consuming and labor-intensive. In recent years, a variety of new biosensors have been introduced that feature high sensitivity and specificity, inexpensive instruments or devices that do not require washing, and rapid response times. With this background, a fast affinity-induced reaction sensor (FAIRS) technique based on differential kinetics was developed for rapid, one-step antigen detection. The FAIRS detection is based on the rapid affinity of antibodies and antigens and the slow reaction of fluorogenic click chemistry. Complete characterization of the sensor revealed a response time of 6.5 ± 1.0 minutes. The significant increase in the local concentration of click chemicals results from the binding of tagged antibodies to antigens quantitatively matches the difference in the intrinsic second-order rate constants. A kinetic discrepancy provides a guideline for further designing such sensors. Considering the important properties of the FAIRS method, including its high specificity, sensitivity, and simple detection procedure, this method has found wide application in drug and biomarker discovery, inflammatory disease diagnosis, and similar cases (Liu et al., Reference Liu, Abdullah, Yang and Wang2019).

Impedimetric biosensors allow the detection of biological targets with no need to use any prior labeling steps (fluorophores, redox enzymes, etc.) (Daniels and Pourmand, Reference Daniels and Pourmand2007). Impedimetric technique has also been used for monitoring the catalyzed reactions of enzymes or the biomolecular recognition events of lectins, specific binding proteins, nucleic acids, receptors, antibodies, antibody-related substances, and whole cells (Prodromidis, Reference Prodromidis2010). Physical adsorption, intermolecular cross-linking, covalent bonding, and entrapment are techniques commonly used to immobilize biomolecules to develop specific biosensors (Turner et al., Reference Turner, Karube and Wilson1987). However, electrochemical immobilization and especially electro-addressing are new approaches to immobilize active biomolecules on conductive surfaces. Biomolecules can be immobilized on an electrode (chip) by applying an electrochemical potential that must be compatible with biochip microarray fabrication for multiple detections. With this background, an electro-addressing strategy compatible with the fabrication of multidetector microarrays was used to monitor the label-free immobilization of proteins on a gold surface using electrochemical impedance spectroscopy (EIS). Azide-alkyne cycloaddition click strategy is used to achieve the functionalization process. The main advantage of this method lies in the fact that proteins can be spatially addressed on a single gold chip in a microarray while avoiding any unspecific addressing by maintaining the other chips at a positive potential (+300 mV) (Meini et al., Reference Meini, Ripert, Chaix, Farre, De Crozals, Kherrat and Jaffrezic-Renault2014).

Nucleic acid modifications

Indeed, functionalized oligonucleotides hold significant potential for a wide range of applications in various fields, including nucleic acid diagnostics, therapy, and nanobiotechnology (Nåbo et al., Reference Nåbo, Madsen, Jensen, Kongsted and Astakhova2015). There has been a growing interest in the conjugation of proteins and nucleic acids for diverse applications in both scientific research and industrial settings (Baranda Pellejero et al., Reference Baranda Pellejero, Nijenhuis, Ricci and Gothelf2023; Freitag et al., Reference Freitag, Möser, Belay, Altattan, Grasse, Pothineni, Schnauß and Smith2023; Rück et al., Reference Rück, Mishra, Sørensen, Liisberg, Sloth, Cerretani, Mollerup, Kjaer, Lou and Jensen2023). This emerging field offers exciting possibilities as it combines the multiple functionalities of proteins with the precise recognition and encoding properties of nucleic acid (Manderville and Wetmore, Reference Manderville and Wetmore2016; Trads et al., Reference Trads, Tørring and Gothelf2017). Absolutely, proteins offer a wide range of functionalities that include catalyzing chemical reactions (Köhler and Turner, Reference Köhler and Turner2015), generating forces (Derr et al., Reference Derr, Goodman, Jungmann, Leschziner, Shih and Reck-Peterson2012), and participating in receptor–ligand interactions (Li et al., Reference Li, Wong and Liu2014). On the other hand, nucleic acid provides a powerful tool for positioning a molecule of interest with nanometer-level accuracy (Chandrasekaran, Reference Chandrasekaran2016; Ramakrishnan et al., Reference Ramakrishnan, Krainer, Grundmeier, Schlierf and Keller2016; Zhang et al., Reference Zhang, Yang, Jiang, Liu and Yan2016). Although protein-nucleic acid conjugation has been achieved by a variety of chemical (Baranda Pellejero et al., Reference Baranda Pellejero, Nijenhuis, Ricci and Gothelf2023) and genetic methods (Siggers and Gordân, Reference Siggers and Gordân2014; Manderville and Wetmore, Reference Manderville and Wetmore2016; Praetorius and Dietz, Reference Praetorius and Dietz2017; Trads et al., Reference Trads, Tørring and Gothelf2017; Chen et al., Reference Chen, Cheng, Huang and Chen2018), there are still few strategies that have high reaction rates, specificity, and biocompatibility (Manderville and Wetmore, Reference Manderville and Wetmore2016; Trads et al., Reference Trads, Tørring and Gothelf2017). By using click chemistry-based methods, researchers can overcome the challenges of conventional protein-nucleic acid conjugation approaches. These strategies provide a versatile and efficient platform for developing bioconjugates with high reaction speed, specificity, and biocompatibility.

RNA–protein interactions (RPIs) are critical in modulating many aspects of coding and noncoding RNA biology and are potential future drug targets (Gerstberger et al., Reference Gerstberger, Hafner and Tuschl2014). In this regard, the researchers developed a homogeneous complementation assay using click chemistry for studying RPIs. The newly developed assay shares many of the same advantages as catalytic enzyme-linked click chemistry assay. One of these benefits is catalytic signal amplification, which allows for the creation of a robust and highly sensitive assay system. This approach should be adaptable to other RPI systems for high-throughput screening design and development, as chemical RNA synthesis allows the synthesis of a variety of labeled RNAs and HT fusion products are available at both the N-terminal and C-terminal (Sherman et al., Reference Sherman, Lorenz and Garner2019).

In DNA sequencing and diagnostics, electrochemical detection of redox-labeled DNA is an alternative to fluorescence techniques (Palecek et al., Reference Palecek, Scheller and Wang2005; Palecek and Bartosik, Reference Palecek and Bartosik2012). Even though, there has been extensive research and numerous oxidizable or reducible labels available, redox labeling of DNA is often encounters challenges related to stability, sensitivity, and cross-reactivity due to the labels (Brázdilová et al., Reference Brázdilová, Vrábel, Pohl, Pivoňková, Havran, Hocek and Fojta2007; Hocek and Fojta, Reference Hocek and Fojta2011). Therefore, redox labeling and electrochemistry have been applied to study DNA–protein interactions that are relatively scarce. This application is limited to methods based on changes in DNA-mediated charge transfer (CT) upon protein binding. Accordingly, developing a new redox labeling method for DNA involving an azido group represents a significant advancement in the field. This novel approach allows for the chemical transformation of the azido group to either nitrophenyltriazole or silence phenyltriazole, while this technique provides electrochemical detection of DNA–protein interactions. The preparation of 5-(4-azidophenyl)-20-deoxycytidine and 7-(4-azidophenyl)-7-deaza-20-deoxyadenosine nucleosides involved an aqueous-phase Suzuki cross-coupling reaction. These azido-labeled nucleosides were then converted into nucleoside triphosphates to serve as substrates for incorporation into DNA. The incorporation process was facilitated by DNA polymerase. Due to a reduction in the azido function, the azidoophenyl-modified nucleotides and azidophenyl-modified DNA have shown a strong signal in voltammetry studies at 0.9 V. It has been shown that the Cu-catalyzed click reactions of azidophenyl-modified nucleosides or azidophenyl-modified DNA with 4-nitrophenylacetylene lead to nitrophenyl-substituted triazoles that exhibit a reduction peak of −0.4 V under voltammetry; however, the click reaction with phenylacetylene leads to electrochemically silent phenyltriazoles. In this study, converting the azidophenyl label to nitrophenyltriazole was used for the electrochemical detection of DNA–protein interactions, in which the p53 protein is explicitly involved. This conversion served as a selective indicator, as only the azidophenyl groups in the regions of DNA that were not protected by the bound p53 protein were converted to nitrophenyltriazoles (Figure 6) (Balintová et al., Reference Balintová, Špaček, Pohl, Brázdová, Havran, Fojta and Hocek2015).

Figure 6. The electrochemical detection of protein–DNA interactions via azidophenyl as a click-transformable redox label of DNA (Balintová et al., Reference Balintová, Špaček, Pohl, Brázdová, Havran, Fojta and Hocek2015).

There has been considerable interest in developing methods for measuring Mn2+ in biological and environmental samples due to the important role and wide distribution of Mn2+ in biological systems, as well as its extensive utility as an excellent contrast agent for MRI scanning, which has attracted a lot of attention (Carter et al., Reference Carter, Young and Palmer2014). For example, overexposure to Mn2+ poses a significant risk to cells and can potentially lead to serious diseases, including neurodegenerative disorders and developmental disorders in children (Guilarte, Reference Guilarte2010; Horning et al., Reference Horning, Caito, Tipps, Bowman and Aschner2015). However, the development of fluorescent Mn2+ sensors has been challenged, due to the chemical similarity between Mn2+ and Mg2+/Ca2+, and quenching of fluorescence, which originates from the paramagnetic properties of Mn2+. Consequently, researchers employed the click-type MnDDC (Mn2+-activated DCV-DNA conjugate) reaction as a fluorescent sensor for Mn2+ detection. In this approach, a molecular beacon was used to facilitate the ligation of DCV-DNA (rep protein of duck circovirus-DNA) and signal readout. The click properties of the MnDDC reaction make this assay extremely versatile for Mn2+ detection and extend its applicability beyond serum and food samples. Importantly, it enables Mn2+ detection directly in live cells without the need for additional washing steps. It must also be noted that using DCV’s protein nature, it could be used on the cell surface as a subcellular targeting sensor for monitoring Mn2+ in the extracellular microenvironment as well. This work demonstrated that click-type MnDDC suits site-specific covalent protein–DNA linkages in complex biological environments (Hu et al., Reference Hu, Zhang, Tang, Fan, Liu, Kang, Lei, Nie, Huang and Yao2019).

With the help of the CuAAC reaction, a new bifunctional bioconjugation reagent, N-(3-azidopropyl) vinylsulfonamide, has been attached to an alkyne-modified DNA or protein. Although it has been shown that partial hydrolysis (ca. 12%) of the sulfonamide occurred in the click modification of DNA, it could still be used for efficient biorthogonal modification of biomolecules to attach vinylsulfonamide (VS) as a Michael acceptor. The VS-linked DNA or protein is capable of specific reactivity with cysteine -containing peptides and/or proteins, leading to the formation of stable covalent cross-links. Indeed, the approach involving VS-linked DNA or protein can be extended to work with any combination of two biomolecules, where one contains an alkyne and the other contains a thiol group. Evidently, pull-down experiments have great potential to identify and isolate DNA-binding proteins containing cysteine near to recognition sequences to identify and isolate them (Dadová et al., Reference Dadová, Vrábel, Adámik, Brázdová, Pohl, Fojta and Hocek2015).

A pioneering method of DNA assembly has been developed, enabling the creation of DNA-modified surfaces for the electrochemical detection of biomolecules. The Cu-free click strategy allows the formation of monolayers with lower density and more uniform spacing while maintaining surface passivation against the redox indicator. These monolayers have been characterized by both electrochemical and imaging techniques. The developed platform facilitates DNA-mediated CT, making it highly sensitive to changes or perturbations in the DNA structure. As a result, this system can achieve exquisite electrochemical discrimination between DNA duplexes that are well-matched and those with mismatches. Furthermore, because this platform has a more significant number of surface-exposed binding sites than conventional high-density films, it can be used to detect protein binding events with higher sensitivity than conventional high-density films. It is possible to detect TATA-binding proteins on low-density membranes at a concentration as low as 4 nM (Furst et al., Reference Furst, Hill and Barton2013).

Researchers have successfully developed an innovative protein–DNA assembly strategy utilizing Cu-free click chemistry. This breakthrough approach provides site specificity and bioorthogonality under mild reaction conditions, enabling efficient and rapid protein–DNA assembly. With the wide availability of oligonucleotides featuring azide modifications at the 5′- and 3′-ends and at internal positions, the presented protein–DNA assembly method offers a versatile and powerful approach. Through this method, researchers could attach proteins to any position on DNA oligos. Even more, the developed protein–DNA assembly method enables site specificity and high coupling efficiency while maintaining the biological activity of the proteins involved. The demonstrated protein–DNA assembly technique is a promising tool for single-molecule studies and DNA-based nanotechnology applications involving functional protein–DNA hybrids (Mukhortava and Schlierf, Reference Mukhortava and Schlierf2016).

Treatment

Click chemistry has attracted significant attention as an ideal approach for drug design and discovery for the following reasons: the azole linkages have small structures that can easily form larger molecules and do not cause severe structural alterations to the whole molecular environment (Wang et al., Reference Wang, Huang, Liu and Zhan2016; Jiang et al., Reference Jiang, Hao, Jing, Wu, Kang, Liu and Zhan2019). Protein-/peptide-derived drugs, due to their structural and functional diversity, high stability, biocompatibility, and low immunogenicity, will become an important part of the pharmaceutical market in the future (Li et al., Reference Li, Aneja and Chaiken2013; Lin et al., Reference Lin, Jiang, Ren and Liu2023).

It should be noted that, whether proteins and peptides are broken down under physiological conditions, the breakdown products are amino acids, which are not toxic and are readily absorbed and can be easily absorbed or excreted (Tang and Becker, Reference Tang and Becker2014). Recently, the development of new peptidomedicines has been dramatically accelerated by the use of modern synthesis techniques such as click chemistry (Zhang et al., Reference Zhang, Chen, Chen, Du, Ding, He, Ding, Hu, Qin and Tang2023). Click chemistry provides a range of peptide/protein modifications and could be combined with other methods to easily create complex structures, drugs, and multicomponent functionalized systems. Furthermore, click chemistry is a very attractive approach for developing practical drug delivery tools and targeted drug delivery (van Dijk et al., Reference Van Dijk, Van Nostrum, Hennink, Rijkers and Liskamp2010). For example, nucleolin is a protein that is located between the cell nucleus and the cell surface. This protein is found to be overexpressed in certain types of cancer cells (Porkka et al., Reference Porkka, Laakkonen, Hoffman, Bernasconi and Ruoslahti2002; Christian et al., Reference Christian, Pilch, Akerman, Porkka, Laakkonen and Ruoslahti2003). Nucleolin can specifically bind to a short peptide known as the F3 peptide. The F3 peptide is conjugated to the surface of nanoparticles and hydrogel substrates via a click reaction, enabling targeted drug delivery (Qin et al., Reference Qin, Zong and Kopelman2014). One of the disadvantages of using peptides as drugs is their low stability compared to protein drugs due to the fragility of peptide and disulfide bonds under physiological conditions. Click chemistry improves peptide stability by replacing these unstable natural bonds with stable structures without compromising peptide function (Li et al., Reference Li, Aneja and Chaiken2013). In general, there are various strategies in click chemistry to produce peptides and proteins with medicinal applications, some of these strategies include cyclization, cross-linking, unstable bond surrogates, introduction of contact groups to improve affinity, and conjugation of multiple functions (Figure 7).

Figure 7. Peptide click chemistry strategies. Each strategy improves the stability and performance of peptides incorporating click-derived triazoles: Conjugation of multiple functional groups refers to the linking of multiple functional groups within a molecule. This can lead to the creation of complex structures with diverse properties. Cyclization in click chemistry refers to the formation of a cyclic compound through a chemical reaction. Protein cross-linking by click chemistry involves the covalent linking of proteins using highly efficient and selective reactions. By using stable and well-characterized building blocks, researchers can minimize the risk of encountering unstable bound surrogates and improve the efficiency and reliability of click chemistry reactions. The introduction of contact groups to improve affinity by click chemistry involves the strategic attachment of specific functional groups to a molecule to enhance its binding or interaction with a target.

Cyclization

Although peptides are good options for medicinal applications due to their high functional diversity, their low bioavailability has limited their use. One of the methods that help to solve this problem is the head-to-tail cyclization method, which reduces the biological vulnerability and increases the permeability of peptides by removing the free C- and N-terminus of the peptides. Therefore, cyclopeptides are relatively more stable for proteolysis than linear peptides (Besser et al., Reference Besser, Müller, Kleinwächter, Greiner, Seyfarth, Steinmetzer, Arad and Reissmann2000a; Shibata et al., Reference Shibata, Suzawa, Soga, Mizukami, Yamada, Hanai and Yamasaki2003; Tugyi et al., Reference Tugyi, Mezö, Fellinger, Andreu and Hudecz2005). For peptide cyclization, the L and D configurations of residues and the presence or absence of amino acids such as proline and glycine are important. This is particularly important for synthesizing tetra-, penta-, and hexacyclic peptides (Klose et al., Reference Klose, Ehrlich and Bienert1998; El Haddadi et al., Reference El Haddadi, Cavelier, Vives, Azmani, Verducci and Martinez2000; Besser et al., Reference Besser, Reissmann, Olender, Rosenfeld and Arad2000b). Additionally, the cyclization of peptides with smaller sequences is also challenging because these small peptides are highly susceptible to oligomerization. Click chemistry macrocyclization is an efficient tool that helps to solve these challenges.

Natural macrocyclization of peptides and proteins often occurs via thioether bridges or disulfide bonds as a constraining element in some conformations, such as turn structures (Musiol et al., Reference Musiol, Siedler, Quarzago and Moroder1994; Siedler et al., Reference Siedler, Quarzago, Rudolph-Böhner and Moroder1994; Turner et al., Reference Turner, Oliver and Lokey2007). However, cyclic peptides in synthetic chemistry and drug discovery are mainly created by forming a disulfide bond between two cysteine residues or an isopeptide bond (Hill et al., Reference Hill, Shepherd, Diness and Fairlie2014). The microwave-assisted click reaction was also used for the cyclization of a peptide derived from the Alzheimer Aβ sequence, and recent work showed that microwave-assisted click chemistry could be used to prepare peptide triazole-based polymers from bifunctional peptide monomers (Elgersma et al., Reference Elgersma, Van Dijk, Dechesne, Van Nostrum, Hennink, Rijkers and Liskamp2009). Generally, there are many methods and reagents for synthesizing cyclic peptides, but the production efficiency of cyclic peptides is the main limitation of these methods (Turner et al., Reference Turner, Oliver and Lokey2007).

Conjugation

Conjugation is the covalent linking of two or more molecular components to create multifunctional molecules. In drug design, conjugation is among the most favorable and widely used parts of click chemistry reactions. Typically, peptide conjugation is less complicated than protein conjugation since protein conjugation suffers from denaturation or loss of biological activity due to its complex structures. On the other hand, sometimes peptides can perform the biological role of a whole protein, which is important in the design of peptide drugs. Therefore, in drug discovery, it is advantageous to conjugate bioactive peptides with synthetic chemical compounds rather than whole proteins (Tang and Becker, Reference Tang and Becker2014). Many click-conjugated proteins and peptides are coupled to biologically relevant tiny molecules whit CuAAC, SPAACs, thiol-ene reaction, thiol-Michael addition, and so forth (Li et al., Reference Li, Aneja and Chaiken2013; Tang and Becker, Reference Tang and Becker2014). Although CuAAC is widely used in biopolymer science, its use in medicinal products is restricted based on its toxicity. This problem, identified by researchers working in the field of bio-click, led to the discovery and use of Cu-free SPAACs in scientific and pharmaceutical research (Agard et al., Reference Agard, Prescher and Bertozzi2004; Takahashi et al., Reference Takahashi, Suzuki, Suhara, Omichi, Shimizu, Hasegawa, Kokudo, Ohta and Ito2013).

The conjugation of peptides with other compounds has opened up many possibilities and applications for their use in the pharmaceutical field. For example, Zhou et al. design click-conjugated protein-drug micelles with anti-ferroptotic and anti-inflammatory properties that stimulate regeneration in spinal cord injuries. In this study, protein-drug micelles are created by conjugating irresolvable ferrostatin-1 and dibenzocyclooctyne modules to amphiphilic polymers using an azido linker-modified acidic fibroblast growth factor (FGF) linker (Zhou et al., Reference Zhou, Zhang, Xin, Yang, Pan, Liu, Liu, Yu, Li and Jiao2022). Moreover, in 2014, conjugated compounds were designed that triterpene sapogenins were linked to helix zone-binding domain (HBD)-bearing peptides of T20 (gp41-specific human immunodeficiency virus type 1 (HIV-1) fusion inhibitor) using the CuAAC reaction, resulting in increased efficacy of each of these compounds compared to their individual states. Triterpenes and peptides individually displayed low potency toward HIV-1 Env-mediated cell–cell fusion (Jiang et al., Reference Jiang, Hao, Jing, Wu, Kang, Liu and Zhan2019). The conjugation of peptides with carbohydrates, hydrogels, and radiolabeling reagents is used to produce vaccines and antibiotics, to properly deliver drugs, and to study the biodistribution of drugs as well as biologically related interactions (such as ligand-receptor binding, protein structures, and enzyme activities) (De Groot et al., Reference De Groot, Cadee, Koten, Hennink and Den Otter2002; Wan et al., Reference Wan, Chen, Chen and Danishefsky2006; Hausner et al., Reference Hausner, Marik, Gagnon and Sutcliffe2008; Li et al., Reference Li, Aneja and Chaiken2013). According to the studies of Skwarczynski et al., the lysine residue in proteins and peptides is a suitable site for conjugation because it has an alpha and epsilon group (Skwarczynski et al., Reference Skwarczynski, Ah Ahmad Fuaad, Rustanti, M Ziora, Aqil, R Batzloff, F Good and Toth2011; Gupta et al., Reference Gupta, Singh, Gupta, Khan, Sehgal, Haridas and Roy2012).

Cross-linking and surrogates for unstable bonds

Peptides are highly flexible due to their small size, which impacts their pharmaceutical performance. On the other hand, proteolysis of peptide bonds and the sensitivity of disulfide bonds to environmental pH are the main and dominant reasons for the low stability of peptides, which is an obstacle to using peptides as drugs (Li et al., Reference Li, Aneja and Chaiken2013). Today, in addition to glycosylation of the N- and C-termini of peptides and other methods to increase their stability, click chemistry has provided solutions to increase the stability of peptides (Powell et al., Reference Powell, Stewart, Jr Otvos, Urge, Gaeta, Sette, Arrhenius, Thomson, Soda and Colon1993). Click chemistry can improve structural rigidity and potentially constrain the structure within the bioactive conformation through chemical alteration such as secondary structure imitation, cross-linking, cyclization, and replacement of degradable peptide bonds. 1,2,3-Triazoles are among the most widely used compounds in click chemistry to replace unstable peptide bonds, cyclization, and cross-linking of peptides (Li et al., Reference Li, Aneja and Chaiken2013). Triazole rings were chosen because of their similarity to the amide bond in terms of molecular dimensions and the compatibility of their linkage geometry for some β-turns (Oh and Guan, Reference Oh and Guan2006).

Introduction of contact groups to improve affinity

Proteins and peptides, with their structural diversity, perform a large number of diverse biological functions compared to other compounds and molecules. Click chemistry significantly expands the range of functions of peptides and proteins. This is particularly important when dealing with pharmaceutical peptides and proteins interacting with other molecules. Li et al. identified a series of triazole-incorporated peptide-based receptors that can bind to the gp120 protein of the HIV-1 envelope and disrupt the virus function. In this study, the proline residue of the peptide (which was located near a binding hotspot residue) was replaced with cis-4-azidoproline. This allowed linkage to a panel of substituted alkynes via click reaction, and finally, high binding affinity was achieved by substituted alkynes with different characteristics (Gopi et al., Reference Gopi, Tirupula, Baxter, Ajith and Chaiken2006; Li et al., Reference Li, Aneja and Chaiken2013).

Enzyme

Click chemistry is a chemical approach based on cycloaddition reactions with high chemoselectivity, which has recently starred in a boom in the field of protein chemistry. Herein, a number of studies in the field of enzyme chemistry will be reviewed, highlighting the role of click reactions (Palomo, Reference Palomo2012).

Enzyme-mediated protein modifications

PTMs of proteins through synthetic methods provide opportunities to precisely attach a wide range of moieties on demand to create new proteins. Site-specific conjugation empowers proteins with new characteristics, further expanding their applications, specifically in diagnosis and therapeutics (Ataie et al., Reference Ataie, Moosavi-Movahedi, Saboury, Hakimelahi, Hwu and Tsay2000; Khajeh et al., Reference Khajeh, Naderi-Manesh, Ranjbar, Akbar Moosavi-Movahedi and Nemat-Gorgani2001; Safarian et al., Reference Safarian, Moosavi-Movahedi, Hosseinkhani, Xia, Habibi-Rezaei, Hosseini, Sorenson and Sheibani2003; Hashemnia et al., Reference Hashemnia, Moosavi-Movahedi, Ghourchian, Ahmad, Hakimelahi and Saboury2006; Tavakoli et al., Reference Tavakoli, Ghourchian, Moosavi-Movahedi and Saboury2006). A better understanding of the protein structure–function relationship is another invaluable achievement of the synthetic modification of proteins. To modify biomolecules, click chemistry offers promising strategies to develop reactions as orthogonal, selective, and reactive as those of natural systems. In 2013, a new approach was designed by Spokoyny et al. for mild functionalization of cysteine thiolate moieties in unprotected peptides based on a nucleophilic aromatic substitution reaction (SNAr) between perfluoroarenes and cysteine residues (Spokoyny et al., Reference Spokoyny, Zou, Ling, Yu, Lin and Pentelute2013). It should be mentioned that perfluoroaromatic reagents are insoluble and, therefore, have low reactivity in aqueous media, and this was a main drawback for its general use. In the same year, a bioconjugation strategy for site-specific cysteine modification was reported by the same research group in which the developed perfluoroarene-cysteine SNAr click reaction was accompanied by glutathione S-transferase enzyme catalysis (Zhang et al., Reference Zhang, Spokoyny, Zou, Simon and Pentelute2013). Utilizing click synthetic transformation, an enzyme-mediated reaction for the chemoselective modification of biomolecules has been demonstrated. Enzyme-catalyzed conjugation is a promising pathway with high biological target specificity under mild reaction conditions. In 2015, CuAAC click reactions were applied by Rachel et al. for enzymatic transamidation to yield covalently conjugated peptides and proteins (Figure 8) (Rachel and Pelletier, Reference Rachel and Pelletier2016).

Figure 8. Combination of enzymatic transamidation and click chemistry for one-pot peptide and protein covalent conjugation (Rachel and Pelletier, Reference Rachel and Pelletier2016).

In 2015, a two-step modular process was designed by Alt et al. for site-specific modification of recombinant antibodies using an enzyme-mediated bioconjugation combined with click reactions (Alt et al., Reference Alt, Paterson, Westein, Rudd, Poniger, Jagdale, Ardipradja, Connell, Krippner and Nair2015). The first step occurred in the presence of transpeptidase Sortase A to incorporate strained cyclooctyne functional groups, and the second step involved the azide-alkyne cycloaddition click reaction. The combination of enzymatic bioconjugation with click chemistry in the current study demonstrated a convenient approach that can be readily utilized for a wide variety of functional groups in all biological macromolecules. In 2016, Nienberg et al. used SPAAC click reaction to modify the protein kinase α-subunit with unnatural amino acid para azidophenylalanine as a fluorophore (Nienberg et al., Reference Nienberg, Retterath, Becher, Saenger, Mootz and Jose2016). In 2010, Peters et al. had developed an alternative cofactor for protein methyltransferases that transfer the activated methyl group from the cofactor mainly to lysine and arginine side chains in the protein substrates. They replaced the methyl group of S-adenosyl-l-methionine (AdoMet) with a pent-2-en-4-ynyl side chain to construct a new cofactor based on AdoMet. The steric effects within the SN2-like transition state is compensated by the double bond in the vicinity of sulfonium center through conjugative stabilization, and the terminal alkyne serves as a valuable tool for chemical protein modifications (Peters et al., Reference Peters, Willnow, Duisken, Kleine, Macherey, Duncan, Litchfield, Lüscher and Weinhold2010). S-acylation of cysteine residues with predominantly C16:0 fatty acids can be mentioned as one of the most common forms of PTMs of proteins. Different in vitro assays have been employed via radiolabeled fatty acids to quantify protein lipidation. However, the cost and safety are mentioned as the main drawbacks. In 2015, a click-based ELISA format was designed to measure enzyme-catalyzed acylation of the protein sonic hedgehog (Lanyon-Hogg et al., Reference Lanyon-Hogg, Masumoto, Bodakh, Konitsiotis, Thinon, Rodgers, Owens, Magee and Tate2015). Alkynylated palmitoyl-coenzyme A substrate was clicked with azido-functionalized peptides to allow colorimetric readout of protein palmitoylation.

Enzymatic protein labeling

In 2011, Willnow et al. reported the synthesis of a new selenium-based S-adenosyl-l-methionine analogue for enzymatic transfer of a small propargyl group. The modified proteins were amenable to be labeled with biotin or fluorophores via CuAAC click reaction (Willnow et al., Reference Willnow, Martin, Lüscher and Weinhold2012). In the same year, Kamaruddin et al. described the development of a series of new fluorescent chemosensor peptide substrates of Src-family protein tyrosine kinases using the Cu(I)-assisted Huisgen cycloaddition click reaction (Kamaruddin et al., Reference Kamaruddin, Ung, Hossain, Jarasrassamee, O’Malley, Thompson, Scanlon, Cheng and Graham2011). In 2011, Heal et al. reported a protocol for the selective and site-specific enzymatic labeling of proteins. The method suggested a click-tagging approach for the enzymatic transfer of myristic acid to an N-terminal glycine (Heal et al., Reference Heal, Wright, Thinon and Tate2012).

Biologically smart carriers

Numerous studies have been performed on smart nanoengineered carriers as effective therapeutic delivery systems. Peptide sequences are promising candidates to be applied as carriers that would be degraded with a specific enzyme. The use of peptide sequence as an enzyme-specific trigger for engineered carriers is an emerging area in targeted drug delivery. Considering the pH-responsive behavior of poly(2-diisopropylaminoethyl methacrylate in the presence of Cathepsin B, which plays as an enzyme-specific degradable cross-linker, Gunawan et al. reported the design of hybrid click capsules (Gunawan et al., Reference Gunawan, Liang, Such, Johnston, Leung, Cui and Caruso2014).

In another study, Skrinjar et al. used click chemistry to develop a building block strategy for enzyme substrate assembly. A sugar moiety as enzyme responsive unit, a linker that can easily be labeled, and a tunable modifier compound combined to construct a click substrate successfully to assay enzyme activity in the newborn screening of lysosomal storage disorders (Skrinjar et al., Reference Skrinjar, Schwarz, Lexmüller, Mechtler, Zeyda, Greber-Platzer, Trometer, Kasper and Mikula2018).

Enzyme immobilization

There is a vast array of research on enzyme immobilization techniques due to its significance in industrial and analytical applications, including cross-linking, adsorption, and entrapment. With respect to the promising role of click chemistry (Debelouchina and Muir, Reference Debelouchina and Muir2017), a wide range of natural and synthetic supports have been applied for this purpose (Moghaddam et al., Reference Moghaddam, Ganjali, Dinarvand, Saboury, Razavi, Moosavi-Movahedi and Norouzi2007; Hashemnia et al., Reference Hashemnia, Ghourchian, Moosavi-Movahedi and Faridnouri2009; Nabati et al., Reference Nabati, Habibi-Rezaei, Amanlou and Moosavi-Movahedi2011; Hong et al., Reference Hong, Yang, Zhao, Xiao, Gao, Yang, Ghourchian, Moosavi-Movahedi, Sheibani and Li2013; Karimi et al., Reference Karimi, Habibi-Rezaei, Safari, Moosavi-Movahedi, Sayyah, Sadeghi and Kokini2014). In 2011, Durmaz et al. reported the synthesis of several types of core microspheres with polydivinylbenzene cross-linkers that carry hydrophilic and/or hydrophobic chains employing click methods (Figure 9) (Durmaz et al., Reference Durmaz, Karagoz, Bicak, Demirkol, Yalcinkaya, Timur and Yagci2011). The modified-polydivinylbenzene microspheres were studied as support for the reversible immobilization of Agaricus bisporus laccase.

Figure 9. Modification of polydivinylbenzene (PDVB) microspheres containing hydrophilic and/or hydrophobic polymer chains by a hydrobromination/click chemistry protocol (N,N,N′,N″,N″-pentamethyldiethylenetriamine (PMDETA), poly(ethylene glycol) (PEG), poly(methylmethacrylate) (PMMA), poly(tert-butylacrylate) (PtBA)) (Durmaz et al., Reference Durmaz, Karagoz, Bicak, Demirkol, Yalcinkaya, Timur and Yagci2011).

In 2012, Çelebi et al. developed an immobilized enzyme reactor as a capillary monolith for a microliquid chromatography system (Çelebi et al., Reference Çelebi, Bayraktar and Tuncel2012). Then, α-chymotrypsin was covalently attached to the monolith via the click approach. In 2016, Strzemińska et al. designed a sensor with electrografted quinone and azido moieties on an electrode surface (Strzemińska et al., Reference Strzemińska, Fanchine, Anquetin, Reisberg, Noël, Pham and Piro2016). Then, using CuAAC click reaction, the peptide probe was coupled to the azido group.

In 2018, Matsumoto et al. reported a synthetic method for a tetrameric streptavidin-based hydrogel by click ligation as a new platform for immobilizing enzymes (Matsumoto et al., Reference Matsumoto, Isogawa, Tanaka and Kondo2018). The hydrogel-coated electrodes were consequently used for the biocatalytic oxidation of glucose. Next year, Wang et al. proposed a new microfluidic fabrication method for enzyme immobilization through “electro click chemistry method” (Wang et al., Reference Wang, Jinlong, An, Kimura and Ono2019).

In another study in 2019, Oktay et al. reported the synthesis of a PEG-based hydrogel for the immobilization of amylase in which PEG was initially functionalized using a thiol-ene click reaction (Oktay et al., Reference Oktay, Demir and Kayaman-Apohan2019). In 2020, a green strategy was reported by Zhao et al. for the construction of a monolithic enzyme reactor in a capillary in which trypsin was immobilized through thiol-ene click reaction (Zhao et al., Reference Zhao, Fan, Mo, Huang and Liu2020). Through thiol-ene click reaction, Fan et al. fabricated an immobilized enzyme reactor based on trimethylolpropane trimethacrylate monolith being applied as a matrix (Fan et al., Reference Fan, Zhao, Wei, Huang and Liu2020). In 2022, Shi et al. reported a stable, efficient, and site-specific method based on a combination of SPAAC click reaction and enzymatic ligation to immobilize green fluorescent protein (Shi et al., Reference Shi, Wang, Deng, Tian, Wu and Zheng2022).

Enzyme inhibitors

Protein inhibitors have great potential for therapeutic applications since many human diseases are associated with protein dysfunction (Angeli and Supuran, Reference Angeli and Supuran2023; Kugler et al., Reference Kugler, Hadzima, Dzijak, Rampmaier, Srb, Vrzal, Voburka, Majer, Řezáčová and Vrabel2023). Extensive efforts have been made to provide potent inhibitors with improved selectivity (Saboury and Moosavi-Movahedi, Reference Saboury and Moosavi-Movahedi1997; Hakimelahi et al., Reference Hakimelahi, Moosavi-Movahedi, Sambaiah, Zhu, Ethiraj, Pasdar and Hakimelahi2002; Mahinpour et al., Reference Mahinpour, Ghasemi, Mosavi Nejad and Zahrayi2019; Hajizadeh et al., Reference Hajizadeh, Moosavi-Movahedi, Sheibani and Moosavi-Movahedi2021). The in situ click chemistry referred to the enzyme-mediated azide-alkyne cycloaddition reactions provides a robust approach for alkyne cycloaddition reactions provides a robust approach for identifying enzyme inhibitors (Mamidyala and Finn, Reference Mamidyala and Finn2010; Linkuvienė et al., Reference Linkuvienė, Zubrienė, Manakova, Petrauskas, Baranauskienė, Zakšauskas, Smirnov, Gražulis, Ladbury and Matulis2018). In this regard, Manetsch et al. published a study in 2004, on the optimization of target-guided strategy using acetylcholinesterase as a test system (Manetsch et al., Reference Manetsch, Krasiński, Radić, Raushel, Taylor, Sharpless and Kolb2004). Carbonic anhydrase inhibitors were identified using in situ click chemistry by Mocharla et al. (Reference Mocharla, Colasson, Lee, Röper, Sharpless, Wong and Kolb2005). In the study carried out by Xie et al. in 2007, CuAAC click reactions were employed to generate two sequential libraries of protein tyrosine phosphatase inhibitors (Xie and Seto, Reference Xie and Seto2007). In 2011, Anand et al. described the synthesis of 1,2,3-1H-triazolyl glycohybrids with various sugar units or a chromenone moiety via CuAAC click reactions, which were consequently screened for inhibitory activities (Anand et al., Reference Anand, Jaiswal, Pandey, Srivastava and Tripathi2011). In the next year, Gu et al. reported the preparation of bisaryl maleimide derivatives to mimic natural kinase inhibitors through a click approach (Gu et al., Reference Gu, Wang, Liu, Fu, Li, Cao, Li, Fang, Xu and Shen2012). In another study published in 2013 by Tieu et al. reported a method to improve the general utility of a multicomponent in situ click approach to ligand optimization (Tieu et al., Reference Tieu, Da Costa, Yap, Keeling, Wilce, Wallace, Booker, Polyak and Abell2013). In this regard, a leaky mutant of Staphylococcus aureus biotin protein ligase was applied to enhance the turnover rate for the reaction of biotin alkyne with an azide to give a triazole.

Summary

Chemical protein modification is a powerful tool for generating new protein constructs. Click chemistry contributes effectively to chemical protein modifications, representing a new way of bioorthogonal chemistry with a promising green and productive future. Small-molecule probes can be chemically or enzymatically attached to the target protein when less structurally disruptive labels are required. Bioorthogonal functional groups can be installed in target biomolecules by a cell’s metabolic machinery, which would consequently be covalently labeled by a probe. Modular click units manipulate different molecules in living cells and image them with super-resolution. Biology and electronics cooperate to fabricate new electrochemical biosensors for molecular recognition and signal transmission using click approaches through attaching proteins to electrode surfaces. Functionalized oligonucleotides can be used in nucleic acid diagnostics, therapy, and nanobiotechnology. Protein–DNA conjugation is becoming increasingly popular in research and industry as it combines proteins’ diverse functionalities with DNA’s precise recognition and encoding abilities. A site-specific covalent protein–DNA linkage with high reaction rates, specificity, and biocompatibility has been achieved using click chemistry. In addition, click chemistry has attracted great attention as an ideal approach for drug design and discovery. However, click strategies are one of the main future lines in protein chemistry to design enzymes with improved catalytic efficiency or a broad substrate activity scope. This perspective focused on the most recent advances in designing and creating new modified proteins using click reactions, specifically those with medicinal applications. However, research and development in this field are exponentially increasing.

Supplementary material

The supplementary material for this article can be found at http://doi.org/10.1017/S0033583524000027.

Acknowledgements

The support of the University of Tehran and Iran National Science Foundation (INSF) is greatly acknowledged.

Competing interest

There are no conflicts to declare.

References

Afshari, R and Shaabani, A (2018) Materials functionalization with multicomponent reactions: State of the art. ACS Combinatorial Science 20(9), 499528.CrossRefGoogle ScholarPubMed
Agard, NJ, Baskin, JM, Prescher, JA, Lo, A and Bertozzi, CR (2006) A comparative study of bioorthogonal reactions with azides. ACS Chemical Biology 1(10), 644648.CrossRefGoogle ScholarPubMed
Agard, NJ, Prescher, JA and Bertozzi, CR (2004) A strain-promoted [3+2] azide−alkyne cycloaddition for covalent modification of biomolecules in living systems. Journal of the American Chemical Society 126(46), 1504615047.CrossRefGoogle ScholarPubMed
Alt, K, Paterson, BM, Westein, E, Rudd, SE, Poniger, SS, Jagdale, S, Ardipradja, K, Connell, TU, Krippner, GY and Nair, AK (2015) A versatile approach for the site-specific modification of recombinant antibodies using a combination of enzyme-mediated bioconjugation and click chemistry. Angewandte Chemie International Edition 54(26), 75157519.CrossRefGoogle ScholarPubMed
Anand, N, Jaiswal, N, Pandey, SK, Srivastava, A and Tripathi, RP (2011) Application of click chemistry towards an efficient synthesis of 1, 2, 3-1H-triazolyl glycohybrids as enzyme inhibitors. Carbohydrate Research 346(1), 1625.CrossRefGoogle ScholarPubMed
Angeli, A and Supuran, CT (2023) Click chemistry approaches for developing carbonic anhydrase inhibitors and their applications. Journal of Enzyme Inhibition and Medicinal Chemistry 38(1), 2166503.CrossRefGoogle ScholarPubMed
Archakov, A and Ivanov, YD (2007) Analytical nanobiotechnology for medicine diagnostics. Molecular BioSystems 3(5), 336342.CrossRefGoogle ScholarPubMed
Archakov, AI, Ivanov, YD, Lisitsa, AV and Zgoda, VG (2007) AFM fishing nanotechnology is the way to reverse the Avogadro number in proteomics. Proteomics 7(1), 49.CrossRefGoogle ScholarPubMed
Ataie, G, Moosavi-Movahedi, A, Saboury, A, Hakimelahi, G, Hwu, JR and Tsay, S (2000) Enthalpy and enzyme activity of modified histidine residues of adenosine deaminase and diethyl pyrocarbonate complexes. International Journal of Biological Macromolecules 27(1), 2933.CrossRefGoogle ScholarPubMed
Avci, R, Schweitzer, M, Boyd, RD, Wittmeyer, J, Steele, A, Toporski, J, Beech, I, Arce, FT, Spangler, B and Cole, KM (2004) Comparison of antibody− antigen interactions on collagen measured by conventional immunological techniques and atomic force microscopy. Langmuir 20(25), 1105311063.CrossRefGoogle ScholarPubMed
Balintová, J, Špaček, J, Pohl, R, Brázdová, M, Havran, L, Fojta, M and Hocek, M (2015) Azidophenyl as a click-transformable redox label of DNA suitable for electrochemical detection of DNA–protein interactions. Chemical Science 6(1), 575587.CrossRefGoogle ScholarPubMed
Baranda Pellejero, L, Nijenhuis, MA, Ricci, F and Gothelf, KV (2023) Protein-Templated reactions using DNA-antibody conjugates. Small 19(13), 2200971.CrossRefGoogle ScholarPubMed
Baskin, JM, Prescher, JA, Laughlin, ST, Agard, NJ, Chang, PV, Miller, IA, Lo, A, Codelli, JA and Bertozzi, CR (2007) Copper-free click chemistry for dynamic in vivo imaging. Proceedings of the National Academy of Sciences 104(43), 1679316797.CrossRefGoogle ScholarPubMed
Baslé, E, Joubert, N and Pucheault, M (2010) Protein chemical modification on endogenous amino acids. Chemistry & Biology 17(3), 213227.CrossRefGoogle ScholarPubMed
Bernardes, GJL, Chalker, JM and Davis, BG (2010) Chemical protein modification. In Pignataro, B (ed.). Ideas in Chemistry and Molecular Sciences: Where Chemistry Meets Life. Wiley-VCH Verlag GmbH & Co. KGaA, pp. 5991. https://doi.org/10.1002/9783527630516.ch3.CrossRefGoogle Scholar
Besser, D, Müller, B, Kleinwächter, P, Greiner, G, Seyfarth, L, Steinmetzer, T, Arad, O and Reissmann, S (2000a) Synthesis and characterization of octapeptide somatostatin analogues with backbone cyclization: Comparison of different strategies, biological activities and enzymatic stabilities. Journal für Praktische Chemie 342(6), 537545.3.0.CO;2-2>CrossRefGoogle Scholar
Besser, D, Reissmann, S, Olender, R, Rosenfeld, R and Arad, O (2000b) Study on the cyclization tendency of backbone cyclic tetrapeptides. Journal of Peptide Research 56(6), 337345.CrossRefGoogle Scholar
Birts, CN, Sanzone, AP, El-Sagheer, AH, Blaydes, JP, Brown, T and Tavassoli, A (2014) Transcription of click-linked DNA in human cells. Angewandte Chemie International Edition 53(9), 23622365.CrossRefGoogle ScholarPubMed
Boutureira, O and Bernardes, GAJ (2015) Advances in chemical protein modification. Chemical Reviews 115(5), 21742195.CrossRefGoogle ScholarPubMed
Brázdilová, P, Vrábel, M, Pohl, R, Pivoňková, H, Havran, L, Hocek, M and Fojta, M (2007) Ferrocenylethynyl derivatives of nucleoside triphosphates: Synthesis, incorporation, electrochemistry, and bioanalytical applications. Chemistry–A European Journal 13(34), 95279533.CrossRefGoogle ScholarPubMed
Campos, LM, Killops, KL, Sakai, R, Paulusse, JM, Damiron, D, Drockenmuller, E, Messmore, BW and Hawker, CJ (2008) Development of thermal and photochemical strategies for thiol− ene click polymer functionalization. Macromolecules 41(19), 70637070.CrossRefGoogle Scholar
Carrico, IS, Carlson, BL and Bertozzi, CR (2007) Introducing genetically encoded aldehydes into proteins. Nature Chemical Biology 3(6), 321322.CrossRefGoogle ScholarPubMed
Carter, KP, Young, AM and Palmer, AE (2014) Fluorescent sensors for measuring metal ions in living systems. Chemical Reviews 114(8), 45644601.CrossRefGoogle ScholarPubMed
Carvalho, FA, Connell, S, Miltenberger-Miltenyi, G, Pereira, SV, Tavares, A, Ariëns, RA and Santos, NC (2010) Atomic force microscopy-based molecular recognition of a fibrinogen receptor on human erythrocytes. ACS Nano 4(8), 46094620.CrossRefGoogle ScholarPubMed
Çelebi, B, Bayraktar, A and Tuncel, A (2012) Synthesis of a monolithic, micro-immobilised enzyme reactor via click-chemistry. Analytical and Bioanalytical Chemistry 403, 26552663.CrossRefGoogle ScholarPubMed
Chandrasekaran, AR (2016) Programmable DNA scaffolds for spatially-ordered protein assembly. Nanoscale 8(8), 44364446.CrossRefGoogle ScholarPubMed
Chen, I and Ting, AY (2005) Site-specific labeling of proteins with small molecules in live cells. Current Opinion in Biotechnology 16(1), 3540.CrossRefGoogle ScholarPubMed
Chen, T-Y, Cheng, Y-S, Huang, P-S and Chen, P (2018) Facilitated unbinding via multivalency-enabled ternary complexes: New paradigm for protein–DNA interactions. Accounts of Chemical Research 51(4), 860868.CrossRefGoogle ScholarPubMed
Chin, JW, Cropp, TA, Anderson, JC, Mukherji, M, Zhang, Z and Schultz, PG (2003) An expanded eukaryotic genetic code. Science 301(5635), 964967.CrossRefGoogle Scholar
Christian, S, Pilch, J, Akerman, ME, Porkka, K, Laakkonen, P and Ruoslahti, E (2003) Nucleolin expressed at the cell surface is a marker of endothelial cells in angiogenic blood vessels. Journal of Cell Biology 163(4), 871878.CrossRefGoogle ScholarPubMed
Chtcheglova, L and Hinterdorfer, P (2011) Simultaneous topography and recognition imaging on endothelial cells. Journal of Molecular Recognition 24(5), 788794.CrossRefGoogle ScholarPubMed
Colak, B, Da Silva, JC, Soares, TA and Gautrot, JE (2016) Impact of the molecular environment on thiol–ene coupling for biofunctionalization and conjugation. Bioconjugate Chemistry 27(9), 21112123.CrossRefGoogle ScholarPubMed
Dadová, J, Vrábel, M, Adámik, M, Brázdová, M, Pohl, R, Fojta, M and Hocek, M (2015) Azidopropylvinylsulfonamide as a new bifunctional click reagent for bioorthogonal conjugations: Application for DNA–protein cross-linking. Chemistry–A European Journal 21(45), 1609116102.CrossRefGoogle ScholarPubMed
Dammer, U, Hegner, M, Anselmetti, D, Wagner, P, Dreier, M, Huber, W and Güntherodt, H-J (1996) Specific antigen/antibody interactions measured by force microscopy. Biophysical Journal 70(5), 24372441.CrossRefGoogle ScholarPubMed
Daniels, JS and Pourmand, N (2007) Label-free impedance biosensors: Opportunities and challenges. Electroanalysis: An International Journal Devoted to Fundamental and Practical Aspects of Electroanalysis 19(12), 12391257.CrossRefGoogle ScholarPubMed
Dawson, PE and Kent, SB (2000) Synthesis of native proteins by chemical ligation. Annual Review of Biochemistry 69(1), 923960.CrossRefGoogle ScholarPubMed
Dawson, PE, Muir, TW, Clark-Lewis, I and Kent, SB (1994) Synthesis of proteins by native chemical ligation. Science 266(5186), 776779.CrossRefGoogle ScholarPubMed
De Groot, CJ, Cadee, JA, Koten, JW, Hennink, WE and Den Otter, W (2002) Therapeutic efficacy of IL-2-loaded hydrogels in a mouse tumor model. International Journal of Cancer 98(1), 134140.CrossRefGoogle Scholar
Debelouchina, GT and Muir, TW (2017) A molecular engineering toolbox for the structural biologist. Quarterly Reviews of Biophysics 50, e7.CrossRefGoogle ScholarPubMed
Derr, ND, Goodman, BS, Jungmann, R, Leschziner, AE, Shih, WM and Reck-Peterson, SL (2012) Tug-of-war in motor protein ensembles revealed with a programmable DNA origami scaffold. Science 338(6107), 662665.CrossRefGoogle ScholarPubMed
Devaraj, NK and Finn, M (2021) Introduction: Click Chemistry. Chemical Reviews 121(12), 66976698.CrossRefGoogle ScholarPubMed
Devaraj, NK, Weissleder, R and Hilderbrand, SA (2008) Tetrazine-based cycloadditions: Application to pretargeted live cell imaging. Bioconjugate Chemistry 19(12), 22972299.CrossRefGoogle ScholarPubMed
Durmaz, YY, Karagoz, B, Bicak, N, Demirkol, DO, Yalcinkaya, EE, Timur, S and Yagci, Y (2011) Modification of polydivinylbenzene microspheres by a hydrobromination/click-chemistry protocol and their protein-adsorption properties. Macromolecular Bioscience 11(1), 141150.CrossRefGoogle ScholarPubMed
El Haddadi, M, Cavelier, F, Vives, E, Azmani, A, Verducci, J and Martinez, J (2000) All-l-Leu-pro-Leu-pro: A challenging cyclization. Journal of Peptide Science 6(11), 560570.3.0.CO;2-I>CrossRefGoogle Scholar
Elgersma, RC, Van Dijk, M, Dechesne, AC, Van Nostrum, CF, Hennink, WE, Rijkers, DT and Liskamp, RM (2009) Microwave-assisted click polymerization for the synthesis of Aβ (16–22) cyclic oligomers and their self-assembly into polymorphous aggregates. Organic & Biomolecular Chemistry 7(21), 45174525.CrossRefGoogle ScholarPubMed
Engin, K, Leeper, D, Cater, J, Thistlethwaite, A, Tupchong, L and McFarlane, J (1995) Extracellular pH distribution in human tumours. International Journal of Hyperthermia 11(2), 211216.CrossRefGoogle ScholarPubMed
Fan, P-R, Zhao, X, Wei, Z-H, Huang, Y-P and Liu, Z-S (2020) Robust immobilized enzyme reactor based on trimethylolpropane trimethacrylate organic monolithic matrix through “thiol-ene” click reaction. European Polymer Journal 124, 109456.CrossRefGoogle Scholar
Fisher, SA, Baker, AE and Shoichet, MS (2017) Designing peptide and protein modified hydrogels: Selecting the optimal conjugation strategy. Journal of the American Chemical Society 139(22), 74167427.CrossRefGoogle ScholarPubMed
Florin, E-L, Moy, VT and Gaub, HE (1994) Adhesion forces between individual ligand-receptor pairs. Science 264(5157), 415417.CrossRefGoogle ScholarPubMed
Freitag, JS, Möser, C, Belay, R, Altattan, B, Grasse, N, Pothineni, BK, Schnauß, J and Smith, DM (2023) Integration of functional peptides into nucleic acid-based nanostructures. Nanoscale 15(17), 76087624.CrossRefGoogle ScholarPubMed
Furst, AL, Hill, MG and Barton, JK (2013) DNA-modified electrodes fabricated using copper-free click chemistry for enhanced protein detection. Langmuir 29(52), 1614116149.CrossRefGoogle ScholarPubMed
Gabius, H-J (Ed.) (2011) The Sugar Code: Fundamentals of Glycosciences. Germany: John Wiley & Sons.Google Scholar
Gamblin, DP, Scanlan, EM and Davis, BG (2009) Glycoprotein synthesis: An update. Chemical Reviews 109(1), 131163.CrossRefGoogle ScholarPubMed
Gan, Y, Li, Y, Zhou, H and Wang, R (2022) Deciphering regulatory proteins of prenylated protein via the FRET technique using nitroso-based ene-ligation and sequential azidation and click reaction. Organic Letters 24(36), 66256630.CrossRefGoogle ScholarPubMed
Gerstberger, S, Hafner, M and Tuschl, T (2014) A census of human RNA-binding proteins. Nature Reviews Genetics 15(12), 829845.CrossRefGoogle ScholarPubMed
Gopi, HN, Tirupula, KC, Baxter, S, Ajith, S and Chaiken, IM (2006) Click chemistry on azidoproline: High-affinity dual antagonist for HIV-1 envelope glycoprotein gp120. ChemMedChem: Chemistry Enabling Drug Discovery 1(1), 5457.CrossRefGoogle ScholarPubMed
Grady, E (2016) Gastrointestinal bleeding scintigraphy in the early 21st century. Journal of Nuclear Medicine 57(2), 252259.CrossRefGoogle ScholarPubMed
Gu, G, Wang, H, Liu, P, Fu, C, Li, Z, Cao, X, Li, Y, Fang, Q, Xu, F and Shen, J (2012) Discovery and structural insight of a highly selective protein kinase inhibitor hit through click chemistry. Chemical Communications 48(22), 27882790.CrossRefGoogle ScholarPubMed
Guilarte, TR (2010) Manganese and Parkinson’s disease: A critical review and new findings. Environmental Health Perspectives 118(8), 10711080.CrossRefGoogle ScholarPubMed
Gunawan, ST, Liang, K, Such, GK, Johnston, AP, Leung, MK, Cui, J and Caruso, F (2014) Engineering enzyme-cleavable hybrid click capsules with a pH-Sheddable coating for intracellular degradation. Small 10(20), 40804086.CrossRefGoogle ScholarPubMed
Gupta, K, Singh, S, Gupta, K, Khan, N, Sehgal, D, Haridas, V and Roy, RP (2012) A bioorthogonal chemoenzymatic strategy for defined protein dendrimer assembly. Chembiochem 13(17), 24892494.CrossRefGoogle ScholarPubMed
Hackenberger, CP and Schwarzer, D (2008) Chemoselective ligation and modification strategies for peptides and proteins. Angewandte Chemie International Edition 47(52), 1003010074.CrossRefGoogle ScholarPubMed
Hajizadeh, M, Moosavi-Movahedi, Z, Sheibani, N and Moosavi-Movahedi, AA (2021) An outlook on suicide enzyme inhibition and drug design. Journal of the Iranian Chemical Society 19, 15751592.CrossRefGoogle Scholar
Hakimelahi, GH, Moosavi-Movahedi, AA, Sambaiah, T, Zhu, J-L, Ethiraj, KS, Pasdar, M and Hakimelahi, S (2002) Reactions of purines-containing butenolides with L-cysteine or N-acetyl-L-cysteine as model biological nucleophiles: A potent mechanism-based inhibitor of ribonucleotide reductase caused apoptosis in breast carcinoma MCF7 cells. European Journal of Medicinal Chemistry 37(3), 207217.CrossRefGoogle ScholarPubMed
Hashemnia, S, Ghourchian, H, Moosavi-Movahedi, AA and Faridnouri, H (2009) Direct electrochemistry of chemically modified catalase immobilized on an oxidatively activated glassy carbon electrode. Journal of Applied Electrochemistry 39, 714.CrossRefGoogle Scholar
Hashemnia, S, Moosavi-Movahedi, A, Ghourchian, H, Ahmad, F, Hakimelahi, G and Saboury, A (2006) Diminishing of aggregation for bovine liver catalase through acidic residues modification. International Journal of Biological Macromolecules 40(1), 4753.CrossRefGoogle ScholarPubMed
Hausner, SH, Marik, J, Gagnon, MKJ and Sutcliffe, JL (2008) In vivo positron emission tomography (PET) imaging with an αvβ6 specific peptide radiolabeled using 18F-“click” chemistry: Evaluation and comparison with the corresponding 4-[18F] fluorobenzoyl-and 2-[18F] fluoropropionyl-peptides. Journal of Medicinal Chemistry 51(19), 59015904.CrossRefGoogle ScholarPubMed
Hayashi, R, Morimoto, S and Tomohiro, T (2019) Tag-convertible photocrosslinker with click-on/off N-acylsulfonamide linkage for protein identification. Chemistry 14(18), 31453148.Google ScholarPubMed
He, D, Xie, X, Yang, F, Zhang, H, Su, H, Ge, Y, Song, H and Chen, PR (2017) Quantitative and comparative profiling of protease substrates through a genetically encoded multifunctional photocrosslinker. Angewandte Chemie International Edition 56(46), 1452114525.CrossRefGoogle ScholarPubMed
Heal, WP, Wright, MH, Thinon, E and Tate, EW (2012) Multifunctional protein labeling via enzymatic N-terminal tagging and elaboration by click chemistry. Nature Protocols 7(1), 105117.CrossRefGoogle Scholar
Hensarling, RM, Doughty, VA, Chan, JW and Patton, DL (2009) “Clicking” polymer brushes with thiol-yne chemistry: Indoors and out. Journal of the American Chemical Society 131(41), 1467314675.CrossRefGoogle ScholarPubMed
Hill, TA, Shepherd, NE, Diness, F and Fairlie, DP (2014) Constraining cyclic peptides to mimic protein structure motifs. Angewandte Chemie International Edition 53(48), 1302013041.CrossRefGoogle ScholarPubMed
Hocek, M and Fojta, M (2011) Nucleobase modification as redox DNA labelling for electrochemical detection. Chemical Society Reviews 40(12), 58025814.CrossRefGoogle ScholarPubMed
Hong, J, Yang, W-Y, Zhao, Y-X, Xiao, B-L, Gao, Y-F, Yang, T, Ghourchian, H, Moosavi-Movahedi, Z, Sheibani, N and Li, J-G (2013) Catalase immobilized on a functionalized multi-walled carbon nanotubes–gold nanocomposite as a highly sensitive bio-sensing system for detection of hydrogen peroxide. Electrochimica Acta 89, 317325.CrossRefGoogle Scholar
Hoogenboom, R (2010) Thiol–yne chemistry: A powerful tool for creating highly functional materials. Angewandte Chemie International Edition 49(20), 34153417.CrossRefGoogle ScholarPubMed
Horne, JE, Walko, M, Calabrese, AN, Levenstein, MA, Brockwell, DJ, Kapur, N, Wilson, AJ and Radford, SE (2018) Rapid mapping of protein interactions using tag-transfer photocrosslinkers. Angewandte Chemie International Edition 57(51), 1668816692.CrossRefGoogle ScholarPubMed
Horning, KJ, Caito, SW, Tipps, KG, Bowman, AB and Aschner, M (2015) Manganese is essential for neuronal health. Annual Review of Nutrition 35, 71108.CrossRefGoogle ScholarPubMed
Hoyle, CE and Bowman, CN (2010) Thiol–ene click chemistry. Angewandte Chemie International Edition 49(9), 15401573.CrossRefGoogle ScholarPubMed
Hu, S, Zhang, J, Tang, R, Fan, J, Liu, H, Kang, W, Lei, C, Nie, Z, Huang, Y and Yao, S (2019) Click-type protein–DNA conjugation for Mn2+ imaging in living cells. Analytical Chemistry 91(15), 1018010187.CrossRefGoogle ScholarPubMed
Javanbakht, S, Nasiriani, T, Farhid, H, Nazeri, MT and Shaabani, A (2022) Sustainable functionalization and modification of materials via multicomponent reactions in water. Frontiers of Chemical Science and Engineering 16(9), 13181344.CrossRefGoogle Scholar
Javanbakht, S and Shaabani, A (2019) Multicomponent reactions-based modified/functionalized materials in the biomedical platforms. ACS Applied Bio Materials 3(1), 156174.CrossRefGoogle ScholarPubMed
Jewett, JC, Sletten, EM and Bertozzi, CR (2010) Rapid cu-free click chemistry with readily synthesized biarylazacyclooctynones. Journal of the American Chemical Society 132(11), 36883690.CrossRefGoogle ScholarPubMed
Jiang, X, Hao, X, Jing, L, Wu, G, Kang, D, Liu, X and Zhan, P (2019) Recent applications of click chemistry in drug discovery. Expert Opinion on Drug Discovery 14(8), 779789.CrossRefGoogle ScholarPubMed
Jones, MW, Mantovani, G, Ryan, SM, Wang, X, Brayden, DJ and Haddleton, DM (2009) Phosphine-mediated one-pot thiol–ene “click” approach to polymer–protein conjugates. Chemical Communications (35), 52725274.CrossRefGoogle ScholarPubMed
Kamaruddin, MA, Ung, P, Hossain, MI, Jarasrassamee, B, O’Malley, W, Thompson, P, Scanlon, D, Cheng, H-C and Graham, B (2011) A facile, click chemistry-based approach to assembling fluorescent chemosensors for protein tyrosine kinases. Bioorganic & Medicinal Chemistry Letters 21(1), 329331.CrossRefGoogle ScholarPubMed
Karimi, M, Habibi-Rezaei, M, Safari, M, Moosavi-Movahedi, AA, Sayyah, M, Sadeghi, R and Kokini, J (2014) Immobilization of endo-inulinase on poly-D-lysine coated CaCO3 micro-particles. Food Research International 66, 485492.CrossRefGoogle Scholar
Kee, J-M and Muir, TW (2012) Chasing phosphohistidine, an elusive sibling in the phosphoamino acid family. ACS Chemical Biology 7(1), 4451.CrossRefGoogle ScholarPubMed
Kent, SB (2009) Total chemical synthesis of proteins. Chemical Society Reviews 38(2), 338351.CrossRefGoogle ScholarPubMed
Khajeh, K, Naderi-Manesh, H, Ranjbar, B, Akbar Moosavi-Movahedi, A and Nemat-Gorgani, M (2001) Chemical modification of lysine residues in bacillus α-amylases: Effect on activity and stability. Enzyme and Microbial Technology 28(6), 543549.CrossRefGoogle ScholarPubMed
Khodkari, V, Nazeri, MT, Javanbakht, S and Shaabani, A (2023) In situ copper nanoparticle immobilization on the indigo carmine-functionalized chitosan: A versatile biocatalyst towards CO 2 fixation and click reactions in water. Reaction Chemistry & Engineering 8(1), 152163.CrossRefGoogle Scholar
Klose, J, Ehrlich, A and Bienert, M (1998) Influence of proline and β-turn mimetics on the cyclization of penta-and hexapeptides. Letters in Peptide Science 5, 129131.CrossRefGoogle Scholar
Koehler, KC, Anseth, KS and Bowman, CN (2013) Diels–Alder mediated controlled release from a poly (ethylene glycol) based hydrogel. Biomacromolecules 14(2), 538547.CrossRefGoogle ScholarPubMed
Köhler, V and Turner, NJ (2015) Artificial concurrent catalytic processes involving enzymes. Chemical Communications 51(3), 450464.CrossRefGoogle ScholarPubMed
Kolb, HC, Finn, M and Sharpless, KB (2001) Click chemistry: Diverse chemical function from a few good reactions. Angewandte Chemie International Edition 40(11), 20042021.3.0.CO;2-5>CrossRefGoogle ScholarPubMed
Kolb, HC and Sharpless, KB (2003) The growing impact of click chemistry on drug discovery. Drug Discovery Today 8(24), 11281137.CrossRefGoogle ScholarPubMed
Kugler, M, Hadzima, M, Dzijak, R, Rampmaier, R, Srb, P, Vrzal, L, Voburka, Z, Majer, P, Řezáčová, P and Vrabel, M (2023) Identification of specific carbonic anhydrase inhibitors via in situ click chemistry, phage-display and synthetic peptide libraries: Comparison of the methods and structural study. RSC Medicinal Chemistry 14(1), 144153.CrossRefGoogle ScholarPubMed
Kumar, V and Boddeti, DK (2013) 68Ga-radiopharmaceuticals for PET imaging of infection and inflammation. In Baum, R and Rösch, F (eds.). Theranostics, Gallium-68, and Other Radionuclides: A Pathway to Personalized Diagnosis and Treatment, vol 194, Berlin, Heidelberg: Springer, pp. 189–219. https://doi.org/10.1007/978-3-642-27994-2_11CrossRefGoogle Scholar
Lallana, E, Riguera, R and Fernandez-MEGIA, E (2011) Reliable and efficient procedures for the conjugation of biomolecules through Huisgen azide–alkyne cycloadditions. Angewandte Chemie International Edition 50(38), 87948804.CrossRefGoogle ScholarPubMed
Lang, K and Chin, JW (2014) Cellular incorporation of unnatural amino acids and bioorthogonal labeling of proteins. Chemical Reviews 114(9), 47644806.CrossRefGoogle ScholarPubMed
Lange, OJ and Polizzi, KM (2021) Click it or stick it: Covalent and non-covalent methods for protein-self assembly. Current Opinion in Systems Biology 28, 100374.CrossRefGoogle Scholar
Lanyon-Hogg, T, Masumoto, N, Bodakh, G, Konitsiotis, AD, Thinon, E, Rodgers, UR, Owens, RJ, Magee, AI and Tate, EW (2015) Click chemistry armed enzyme-linked immunosorbent assay to measure palmitoylation by hedgehog acyltransferase. Analytical Biochemistry 490, 6672.CrossRefGoogle ScholarPubMed
Li, C-J and Chan, T-H (1997) Organic reactions in aqueous media. New York: John Wiley & Sons.Google Scholar
Li, H, Aneja, R and Chaiken, I (2013) Click chemistry in peptide-based drug design. Molecules 18(8), 97979817.CrossRefGoogle ScholarPubMed
Li, S, Wong, AH and Liu, F (2014) Ligand-gated ion channel interacting proteins and their role in neuroprotection. Frontiers in Cellular Neuroscience 8, 125.CrossRefGoogle ScholarPubMed
Lin, L, Jiang, L, Ren, E and Liu, G (2023) Bioorthogonal chemistry based on-demand drug delivery system in cancer therapy. Frontiers of Chemical Science and Engineering 17(4), 483489.CrossRefGoogle Scholar
Lin, L, Wang, H, Liu, Y, Yan, H and Lindsay, S (2006) Recognition imaging with a DNA aptamer. Biophysical Journal 90(11), 42364238.CrossRefGoogle ScholarPubMed
Linkuvienė, V, Zubrienė, A, Manakova, E, Petrauskas, V, Baranauskienė, L, Zakšauskas, A, Smirnov, A, Gražulis, S, Ladbury, JE and Matulis, D (2018) Thermodynamic, kinetic, and structural parameterization of human carbonic anhydrase interactions toward enhanced inhibitor design. Quarterly Reviews of Biophysics 51, e10.CrossRefGoogle ScholarPubMed
Liu, J, Abdullah, MA, Yang, L and Wang, J (2019) Fast affinity induced reaction sensor based on a fluorogenic click reaction for quick detection of protein biomarkers. Analytical Chemistry 92(1), 647653.CrossRefGoogle ScholarPubMed
Lodhi, NA, Park, JY, Kim, K, Kim, YJ, Shin, JH, Lee, Y-S, Im, H-J, Jeong, JM, Khalid, M and Cheon, GJ (2019) Development of 99mTc-labeled human serum albumin with prolonged circulation by chelate-then-click approach: A potential blood pool imaging agent. Molecular Pharmaceutics 16(4), 15861595.CrossRefGoogle Scholar
Mahinpour, R, Ghasemi, M, Mosavi Nejad, Z and Zahrayi, Z (2019) The mechanism of inhibitory effect of caffeine from Iranian tea on sigmoidal kinetics of adenosine deaminase enzyme activity. Applied Biology 32(1), 138148.Google Scholar
Maier, K and Wagner, E (2012) Acid-labile traceless click linker for protein transduction. Journal of the American Chemical Society 134(24), 1016910173.CrossRefGoogle ScholarPubMed
Mamidyala, SK and Finn, M (2010) In situ click chemistry: Probing the binding landscapes of biological molecules. Chemical Society Reviews 39(4), 12521261.CrossRefGoogle ScholarPubMed
Manderville, RA and Wetmore, SD (2016) C-linked 8-aryl guanine nucleobase adducts: Biological outcomes and utility as fluorescent probes. Chemical Science 7(6), 34823493.CrossRefGoogle ScholarPubMed
Manetsch, R, Krasiński, A, Radić, Z, Raushel, J, Taylor, P, Sharpless, KB and Kolb, HC (2004) In situ click chemistry: Enzyme inhibitors made to their own specifications. Journal of the American Chemical Society 126(40), 1280912818.CrossRefGoogle ScholarPubMed
Matsumoto, T, Isogawa, Y, Tanaka, T and Kondo, A (2018) Streptavidin-hydrogel prepared by sortase A-assisted click chemistry for enzyme immobilization on an electrode. Biosensors and Bioelectronics 99, 5661.CrossRefGoogle ScholarPubMed
Matyjaszewski, K and Tsarevsky, NV (2014) Macromolecular engineering by atom transfer radical polymerization. Journal of the American Chemical Society 136(18), 65136533.CrossRefGoogle ScholarPubMed
Maza, JC, McKenna, JR, Raliski, BK, Freedman, MT and Young, DD (2015) Synthesis and incorporation of unnatural amino acids to probe and optimize protein bioconjugations. Bioconjugate Chemistry 26(9), 18841889.CrossRefGoogle ScholarPubMed
McKay, CS and Finn, M (2014) Click chemistry in complex mixtures: Bioorthogonal bioconjugation. Chemistry & Biology 21(9), 10751101.CrossRefGoogle ScholarPubMed
Meini, N, Ripert, M, Chaix, C, Farre, C, De Crozals, G, Kherrat, R and Jaffrezic-Renault, N (2014) Label-free electrochemical monitoring of protein addressing through electroactivated “click” chemistry on gold electrodes. Materials Science and Engineering: C 38, 286291.CrossRefGoogle ScholarPubMed
Meng, J, Paetzell, E, Bogorad, A and Soboyejo, W (2010) Adhesion between peptides/antibodies and breast cancer cells. Journal of Applied Physics 107(11), 114301.CrossRefGoogle Scholar
Millar, A, Hannan, W, Sapru, R and Muir, A (1979) An evaluation of six kits of technetium 99m human serum albumin injection for cardiac blood pool imaging. European Journal of Nuclear Medicine 4, 9194.CrossRefGoogle ScholarPubMed
Mocharla, VP, Colasson, B, Lee, LV, Röper, S, Sharpless, KB, Wong, CH and Kolb, HC (2005) In situ click chemistry: Enzyme-generated inhibitors of carbonic anhydrase II. Angewandte Chemie 117(1), 118122.CrossRefGoogle Scholar
Moghaddam, AB, Ganjali, MR, Dinarvand, R, Saboury, AA, Razavi, T, Moosavi-Movahedi, AA and Norouzi, P (2007) Fundamental studies of the cytochrome c immobilization by the potential cycling method on nanometer-scale nickel oxide surfaces. Biophysical Chemistry 129(2–3), 259268.CrossRefGoogle ScholarPubMed
Mukhortava, A and Schlierf, M (2016) Efficient formation of site-specific protein–DNA hybrids using copper-free click chemistry. Bioconjugate Chemistry 27(7), 15591563.CrossRefGoogle ScholarPubMed
Murphy, RF, Powers, S and Cantor, CR (1984) Endosome pH measured in single cells by dual fluorescence flow cytometry: Rapid acidification of insulin to pH 6. The Journal of Cell Biology 98(5), 17571762.CrossRefGoogle ScholarPubMed
Musiol, HJ, Siedler, F, Quarzago, D and Moroder, L (1994) Redox-active bis-cysteinyl peptides. I. Synthesis of cyclic cystinyl peptides by conventional methods in solution and on solid supports. Biopolymers: Original Research on Biomolecules 34(11), 15531562.CrossRefGoogle ScholarPubMed
Nabati, F, Habibi-Rezaei, M, Amanlou, M and Moosavi-Movahedi, A (2011) Dioxane enhanced immobilization of urease on alkyl modified nano-porous silica using reversible denaturation approach. Journal of Molecular Catalysis B: Enzymatic 70(1–2), 1722.CrossRefGoogle Scholar
Nåbo, LJ, Madsen, CS, Jensen, KJ, Kongsted, J and Astakhova, K (2015) Ultramild protein-mediated click chemistry creates efficient oligonucleotide probes for targeting and detecting nucleic acids. Chembiochem 16(8), 11631167.CrossRefGoogle ScholarPubMed
Nakamura, Y, Inomata, S, Ebine, M, Manabe, Y, Iwakura, I and Ueda, M (2010) Click-made” biaryl-linker improving efficiency in protein labelling for the membrane target protein of a bioactive compound. Organic & Biomolecular Chemistry 9(1), 8385.CrossRefGoogle ScholarPubMed
Naya, M and Sato, C (2020) Pyrene excimer-based fluorescent labeling of cysteines brought into close proximity by protein dynamics: ASEM-induced thiol-ene click reaction for high spatial resolution CLEM. International Journal of Molecular Sciences 21(20), 7550.CrossRefGoogle ScholarPubMed
Neuert, G, Albrecht, C, Pamir, E and Gaub, H (2006) Dynamic force spectroscopy of the digoxigenin–antibody complex. FEBS Letters 580(2), 505509.CrossRefGoogle ScholarPubMed
Nienberg, C, Retterath, A, Becher, K-S, Saenger, T, Mootz, HD and Jose, J (2016) Site-specific labeling of protein kinase CK2: Combining surface display and click chemistry for drug discovery applications. Pharmaceuticals 9(3), 36.CrossRefGoogle ScholarPubMed
Nischan, N and Hackenberger, CP (2014) Site-specific PEGylation of proteins: Recent developments. Journal of Organic Chemistry 79(22), 1072710733.CrossRefGoogle ScholarPubMed
Nishimura, T, Hamada, S, Hayashida, K, Uehara, T, Katabuchi, T and Hayashi, M (1989) Cardiac blood-pool scintigraphy using technetium-99m DTPA-HSA: Comparison with in vivo technetium-99m RBC labeling. Journal of Nuclear Medicine 30(10), 17131717.Google ScholarPubMed
Niu, G, Lang, L, Kiesewetter, DO, Ma, Y, Sun, Z, Guo, N, Guo, J, Wu, C and Chen, X (2014) In vivo labeling of serum albumin for PET. Journal of Nuclear Medicine 55(7), 11501156.CrossRefGoogle ScholarPubMed
Norberg, O, Deng, L, Aastrup, T, Yan, M and Ramstrom, O (2011) Photo-click immobilization on quartz crystal microbalance sensors for selective carbohydrate− protein interaction analyses. Analytical Chemistry 83(3), 10001007.CrossRefGoogle ScholarPubMed
Norberg, O, Deng, L, Yan, M and Ramstrom, O (2009) Photo-click immobilization of carbohydrates on polymeric surfaces: A quick method to functionalize surfaces for biomolecular recognition studies. Bioconjugate Chemistry 20(12), 23642370.CrossRefGoogle ScholarPubMed
Oh, K and Guan, Z (2006) A convergent synthesis of new β-turn mimics by click chemistry. Chemical Communications (29), 30693071.CrossRefGoogle ScholarPubMed
Oktay, B, Demir, S and Kayaman-Apohan, N (2019) Preparation of a poly (ethylene glycol)-based cross-linked network from a click reaction for enzyme immobilization. ChemistrySelect 4(20), 60556059.CrossRefGoogle Scholar
Palecek, E and Bartosik, M (2012) Electrochemistry of nucleic acids. Chemical Reviews 112(6), 34273481.CrossRefGoogle ScholarPubMed
Palecek, E, Scheller, F and Wang, J (2005) Electrochemistry of nucleic acids and proteins: Towards electrochemical sensors for genomics and proteomics. Vol. 1, Book series: Perspectives in bioanalysis. Amsterdam, Tokyo: Elsevier.Google Scholar
Palomo, JM (2012) Click reactions in protein chemistry: From the preparation of semisynthetic enzymes to new click enzymes. Organic & Biomolecular Chemistry 10(47), 93099318.CrossRefGoogle ScholarPubMed
Patterson, DM, Nazarova, LA and prescher, JA (2014) Finding the right (bioorthogonal) chemistry. ACS Chemical Biology 9(3), 592605.CrossRefGoogle ScholarPubMed
Peters, W, Willnow, S, Duisken, M, Kleine, H, Macherey, T, Duncan, KE, Litchfield, DW, Lüscher, B and Weinhold, E (2010) Enzymatic site-specific functionalization of protein methyltransferase substrates with alkynes for click labeling. Angewandte Chemie International Edition 49(30), 51705173.CrossRefGoogle ScholarPubMed
Phelps, ME (2000) PET: The merging of biology and imaging into molecular imaging. Journal of Nuclear Medicine 41(4), 661681.Google ScholarPubMed
Porkka, K, Laakkonen, P, Hoffman, JA, Bernasconi, M and Ruoslahti, E (2002) A fragment of the HMGN2 protein homes to the nuclei of tumor cells and tumor endothelial cells in vivo. Proceedings of the National Academy of Sciences 99(11), 74447449.CrossRefGoogle Scholar
Powell, MF, Stewart, T, Jr Otvos, L, Urge, L, Gaeta, FC, Sette, A, Arrhenius, T, Thomson, D, Soda, K and Colon, SM (1993) Peptide stability in drug development II. Effect of single amino acid substitution and glycosylation on peptide reactivity in human serum. Pharmaceutical Research 10, 12681273.CrossRefGoogle ScholarPubMed
Praetorius, F and Dietz, H (2017) Self-assembly of genetically encoded DNA-protein hybrid nanoscale shapes. Science 355(6331), eaam5488.CrossRefGoogle ScholarPubMed
Prescher, JA and Bertozzi, CR (2005) Chemistry in living systems. Nature Chemical Biology 1(1), 1321.CrossRefGoogle ScholarPubMed
Preston, GW and Wilson, AJ (2013) Photo-induced covalent cross-linking for the analysis of biomolecular interactions. Chemical Society Reviews 42(8), 32893301.CrossRefGoogle ScholarPubMed
Prodromidis, MI (2010) Impedimetric immunosensors—A review. Electrochimica Acta 55(14), 42274233.CrossRefGoogle Scholar
Qin, M, Zong, H and Kopelman, R (2014) Click conjugation of peptide to hydrogel nanoparticles for tumor-targeted drug delivery. Biomacromolecules 15(10), 37283734.CrossRefGoogle ScholarPubMed
Rachel, N and Pelletier, J (2016) One-pot peptide and protein conjugation: A combination of enzymatic transamidation and click chemistry. Chemical Communications 52(12), 25412544.CrossRefGoogle ScholarPubMed
Raliski, BK, Howard, CA and Young, DD (2014) Site-specific protein immobilization using unnatural amino acids. Bioconjugate Chemistry 25(11), 19161920.CrossRefGoogle ScholarPubMed
Ramakrishnan, S, Krainer, G, Grundmeier, G, Schlierf, M and Keller, A (2016) Structural stability of DNA origami nanostructures in the presence of chaotropic agents. Nanoscale 8(19), 1039810405.CrossRefGoogle ScholarPubMed
Ramenda, T, Kniess, T, Bergmann, R, Steinbach, J and Wuest, F (2009) Radiolabelling of proteins with fluorine-18 via click chemistry. Chemical Communications (48), 75217523.CrossRefGoogle ScholarPubMed
Reddy, SK, Cramer, NB and Bowman, CN (2006) Thiol−vinyl mechanisms. 1. Termination and propagation kinetics in thiol−ene photopolymerizations. Macromolecules 39(10), 36733680.CrossRefGoogle Scholar
Rodríguez, DF, Moglie, Y, Ramírez-Sarmiento, CA, Singh, SK, Dua, K and Zacconi, FC (2022) Bio-click chemistry: A bridge between biocatalysis and click chemistry. RSC Advances 12(4), 19321949.CrossRefGoogle ScholarPubMed
Ros, E, Bellido, M, Verdaguer, X, Ribas De Pouplana, L and Riera, A (2020) Synthesis and application of 3-bromo-1, 2, 4, 5-tetrazine for protein labeling to trigger click-to-release biorthogonal reactions. Bioconjugate Chemistry 31(3), 933938.CrossRefGoogle ScholarPubMed
Rostovtsev, VV, Green, LG, Fokin, VV and Sharpless, KB (2002) A stepwise huisgen cycloaddition process: Copper (I)-catalyzed regioselective “ligation” of azides and terminal alkynes. Angewandte Chemie 114(14), 27082711.3.0.CO;2-0>CrossRefGoogle Scholar
Rück, V, Mishra, NK, Sørensen, KK, Liisberg, MB, Sloth, AB, Cerretani, C, Mollerup, CB, Kjaer, A, Lou, C and Jensen, KJ (2023) Bioconjugation of a near-infrared DNA-stabilized silver nanocluster to peptides and human insulin by copper-free click chemistry. Journal of the American Chemical Society 145(30), 1677116777.CrossRefGoogle ScholarPubMed
Saboury, A and Moosavi-Movahedi, A (1997) A simple novel method for studying the combined inhibitory effects of ethylurea and N, N-dimethylurea on jack bean urease. Journal of Enzyme Inhibition 11(3), 217222.CrossRefGoogle Scholar
Safarian, S, Moosavi-Movahedi, AA, Hosseinkhani, S, Xia, Z, Habibi-Rezaei, M, Hosseini, G, Sorenson, C and Sheibani, N (2003) The structural and functional studies of His119 and His12 in RNase a via chemical modification. Journal of Protein Chemistry 22, 643654.CrossRefGoogle ScholarPubMed
Saxon, E and Bertozzi, CR (2000) Cell surface engineering by a modified Staudinger reaction. Science 287(5460), 20072010.CrossRefGoogle ScholarPubMed
Schmaltz, RM, Hanson, SR and Wong, C-H (2011) Enzymes in the synthesis of glycoconjugates. Chemical Reviews 111(7), 42594307.CrossRefGoogle ScholarPubMed
Scinto, SL, Bilodeau, DA, Hincapie, R, Lee, W, Nguyen, SS, Xu, M, Am Ende, CW, Finn, M, Lang, K and Lin, Q (2021) Bioorthogonal chemistry. Nature Reviews Methods Primers 1(1), 30.CrossRefGoogle ScholarPubMed
Senapati, S, Manna, S, Lindsay, S and Zhang, P (2013) Application of catalyst-free click reactions in attaching affinity molecules to tips of atomic force microscopy for detection of protein biomarkers. Langmuir 29(47), 1462214630.CrossRefGoogle ScholarPubMed
Shaabani, A, Afshari, R and Hooshmand, SE (2017a) Crosslinked chitosan nanoparticle-anchored magnetic multi-wall carbon nanotubes: A bio-nanoreactor with extremely high activity toward click-multi-component reactions. New Journal of Chemistry 41(16), 84698481.CrossRefGoogle Scholar
Shaabani, A, Maleki, A and Mofakham, H (2008) Click reaction: Highly efficient synthesis of 2, 3-dihydroquinazolin-4 (1 H)-ones. Synthetic Communications 38(21), 37513759.CrossRefGoogle Scholar
Shaabani, A, Shadi, M, Mohammadian, R, Javanbakht, S, Nazeri, MT and Bahri, F (2019) Multi-component reaction-functionalized chitosan complexed with copper nanoparticles: An efficient catalyst toward A3 coupling and click reactions in water. Applied Organometallic Chemistry 33(9), e5074.CrossRefGoogle Scholar
Shaabani, S, Tavousi Tabatabaei, A and Shaabani, A (2017b) Copper (I) oxide nanoparticles supported on magnetic casein as a bio-supported and magnetically recoverable catalyst for aqueous click chemistry synthesis of 1, 4-disubstituted 1, 2, 3-triazoles. Applied Organometallic Chemistry 31(2), e3559.CrossRefGoogle Scholar
Sherman, EJ, Lorenz, DA and Garner, AL (2019) Click chemistry-mediated complementation assay for RNA–protein interactions. ACS Combinatorial Science 21(7), 522527.CrossRefGoogle ScholarPubMed
Shi, S, Wang, Z, Deng, Y, Tian, F, Wu, Q and Zheng, P (2022) Combination of click chemistry and enzymatic ligation for stable and efficient protein immobilization for single-molecule force spectroscopy. CCS Chemistry 4(2), 598604.CrossRefGoogle Scholar
Shi, X-W, Qiu, L, Nie, Z, Xiao, L, Payne, GF and Du, Y (2013) Protein addressing on patterned microchip by coupling chitosan electrodeposition and ‘electro-click’chemistry. Biofabrication 5(4), 041001.CrossRefGoogle ScholarPubMed
Shibata, K, Suzawa, T, Soga, S, Mizukami, T, Yamada, K, Hanai, N and Yamasaki, M (2003) Improvement of biological activity and proteolytic stability of peptides by coupling with a cyclic peptide. Bioorganic & Medicinal Chemistry Letters 13(15), 25832586.CrossRefGoogle ScholarPubMed
Siedler, F, Quarzago, D, Rudolph-Böhner, S and Moroder, L (1994) Redox-active bis-cysteinyl peptides. II. Comparative study on the sequence-dependent tendency for disulfide loop formation. Biopolymers: Original Research on Biomolecules 34(11), 15631572.CrossRefGoogle ScholarPubMed
Siggers, T and Gordân, R (2014) Protein–DNA binding: Complexities and multi-protein codes. Nucleic Acids Research 42(4), 20992111.CrossRefGoogle ScholarPubMed
Siman, P and Brik, A (2012) Chemical and semisynthesis of posttranslationally modified proteins. Organic & Biomolecular Chemistry 10(30), 56845697.CrossRefGoogle ScholarPubMed
Skrinjar, P, Schwarz, M, Lexmüller, S, Mechtler, TP, Zeyda, M, Greber-Platzer, S, Trometer, J, Kasper, DC and Mikula, H (2018) Rapid and modular assembly of click substrates to assay enzyme activity in the newborn screening of lysosomal storage disorders. ACS Central Science 4(12), 16881696.CrossRefGoogle ScholarPubMed
Skwarczynski, M, Ah Ahmad Fuaad, A, Rustanti, L, M Ziora, Z, Aqil, M, R Batzloff, M, F Good, M and Toth, I (2011) Group a streptococcal vaccine candidates based on the conserved conformational epitope from M protein. Drug Delivery Letters 1(1), 28.Google Scholar
Sletten, EM and Bertozzi, CR (2009) Bioorthogonal chemistry: Fishing for selectivity in a sea of functionality. Angewandte Chemie International Edition 48(38), 69746998.CrossRefGoogle Scholar
Speers, AE and Cravatt, BF (2004) Profiling enzyme activities in vivo using click chemistry methods. Chemistry & Biology 11(4), 535546.CrossRefGoogle ScholarPubMed
Spokoyny, AM, Zou, Y, Ling, JJ, Yu, H, Lin, Y-S and Pentelute, BL (2013) A perfluoroaryl-cysteine SNAr chemistry approach to unprotected peptide stapling. Journal of the American Chemical Society 135(16), 59465949.CrossRefGoogle ScholarPubMed
Stephanopoulos, N and Francis, MB (2011) Choosing an effective protein bioconjugation strategy. Nature Chemical Biology 7(12), 876884.CrossRefGoogle ScholarPubMed
Stroh, C, Wang, H, Bash, R, Ashcroft, B, Nelson, J, Gruber, H, Lohr, D, Lindsay, S and Hinterdorfer, P (2004) Single-molecule recognition imaging microscopy. Proceedings of the National Academy of Sciences 101(34), 1250312507.CrossRefGoogle ScholarPubMed
Strzemińska, I, Fanchine, SSR, Anquetin, G, Reisberg, S, Noël, V, Pham, M and Piro, B (2016) Grafting of a peptide probe for prostate-specific antigen detection using diazonium electroreduction and click chemistry. Biosensors and Bioelectronics 81, 131137.CrossRefGoogle ScholarPubMed
Suazo, KF, Park, K-Y and Distefano, MD (2021) A not-so-ancient grease history: Click chemistry and protein lipid modifications. Chemical Reviews 121(12), 71787248.CrossRefGoogle ScholarPubMed
Takahashi, A, Suzuki, Y, Suhara, T, Omichi, K, Shimizu, A, Hasegawa, K, Kokudo, N, Ohta, S and Ito, T (2013) In situ cross-linkable hydrogel of hyaluronan produced via copper-free click chemistry. Biomacromolecules 14(10), 35813588.CrossRefGoogle ScholarPubMed
Tang, W and Becker, ML (2014) “Click” reactions: A versatile toolbox for the synthesis of peptide-conjugates. Chemical Society Reviews 43(20), 70137039.CrossRefGoogle ScholarPubMed
Tasdelen, MA and Yagci, Y (2013) Light-induced click reactions. Angewandte Chemie International Edition 52(23), 59305938.CrossRefGoogle ScholarPubMed
Tavakoli, H, Ghourchian, H, Moosavi-Movahedi, A and Saboury, A (2006) Histidine and serine roles in catalytic activity of choline oxidase from Alcaligenes species studied by chemical modifications. Process Biochemistry 41(2), 477482.CrossRefGoogle Scholar
Thirumurugan, P, Matosiuk, D and Jozwiak, K (2013) Click chemistry for drug development and diverse chemical–biology applications. Chemical Reviews 113(7), 49054979.CrossRefGoogle ScholarPubMed
Tieu, W, Da Costa, TPS, Yap, MY, Keeling, KL, Wilce, MC, Wallace, JC, Booker, GW, Polyak, SW and Abell, AD (2013) Optimising in situ click chemistry: The screening and identification of biotin protein ligase inhibitors. Chemical Science 4(9), 35333537.CrossRefGoogle Scholar
Trads, JB, Tørring, T and Gothelf, KV (2017) Site-selective conjugation of native proteins with DNA. Accounts of Chemical Research 50(6), 13671374.CrossRefGoogle ScholarPubMed
Tugyi, R, Mezö, G, Fellinger, E, Andreu, D and Hudecz, F (2005) The effect of cyclization on the enzymatic degradation of herpes simplex virus glycoprotein D derived epitope peptide. Journal of Peptide Science 11(10), 642649.CrossRefGoogle ScholarPubMed
Turner, A, Karube, I and Wilson, GS (1987) Biosensors: Fundamentals and Applications. New York: Oxford University Press.Google Scholar
Turner, RA, Oliver, AG and Lokey, RS (2007) Click chemistry as a macrocyclization tool in the solid-phase synthesis of small cyclic peptides. Organic Letters 9(24), 50115014.CrossRefGoogle ScholarPubMed
Valverde, IE, Lecaille, F, Lalmanach, G, Aucagne, V and Delmas, AF (2012) Synthesis of a biologically active triazole-containing analogue of cystatin A through successive peptidomimetic alkyne–azide ligations. Angewandte Chemie International Edition 51(3), 718722.CrossRefGoogle ScholarPubMed
Van Berkel, SS, Van Eldijk, MB and Van Hest, JC (2011) Staudinger ligation as a method for bioconjugation. Angewandte Chemie International Edition 50(38), 88068827.CrossRefGoogle ScholarPubMed
Van Dijk, M, Van Nostrum, CF, Hennink, WE, Rijkers, DT and Liskamp, RM (2010) Synthesis and characterization of enzymatically biodegradable PEG and peptide-based hydrogels prepared by click chemistry. Biomacromolecules 11(6), 16081614.CrossRefGoogle ScholarPubMed
Villalonga, ML, Diez, P, Sanchez, A, Gamella, M, Pingarron, JM and Villalonga, R (2014) Neoglycoenzymes. Chemical Reviews 114(9), 48684917.CrossRefGoogle ScholarPubMed
Wallat, JD, Rose, KA and Pokorski, JK (2014) Proteins as substrates for controlled radical polymerization. Polymer Chemistry 5(5), 15451558.CrossRefGoogle Scholar
Walsh, CT, Garneau-Tsodikova, S and Gatto, GJ (2005) Protein posttranslational modifications: The chemistry of proteome diversifications. Angewandte Chemie International Edition 44(45), 73427372.CrossRefGoogle ScholarPubMed
Wan, Q, Chen, J, Chen, G and Danishefsky, SJ (2006) A potentially valuable advance in the synthesis of carbohydrate-based anticancer vaccines through extended cycloaddition chemistry. Journal of Organic Chemistry 71(21), 82448249.CrossRefGoogle ScholarPubMed
Wang, B, Guo, C, Zhang, M, Park, B and Xu, B (2012) High-resolution single-molecule recognition imaging of the molecular details of ricin–aptamer interaction. Journal of Physical Chemistry B 116(17), 53165322.CrossRefGoogle ScholarPubMed
Wang, H, Dalal, Y, Henikoff, S and Lindsay, S (2008) Single-epitope recognition imaging of native chromatin. Epigenetics & Chromatin 1(1), 19.CrossRefGoogle ScholarPubMed
Wang, L-X and Amin, MN (2014) Chemical and chemoenzymatic synthesis of glycoproteins for deciphering functions. Chemistry & Biology 21(1), 5166.CrossRefGoogle ScholarPubMed
Wang, X, Huang, B, Liu, X and Zhan, P (2016) Discovery of bioactive molecules from CuAAC click-chemistry-based combinatorial libraries. Drug Discovery Today 21(1), 118132.CrossRefGoogle ScholarPubMed
Wang, Z, Jinlong, L, An, Z, Kimura, M and Ono, T (2019) Enzyme immobilization in completely packaged freestanding SU-8 microfluidic channel by electro click chemistry for compact thermal biosensor. Process Biochemistry 79, 5764.CrossRefGoogle Scholar
Willnow, S, Martin, M, Lüscher, B and Weinhold, E (2012) A selenium-based click AdoMet analogue for versatile substrate labeling with wild-type protein methyltransferases. Chembiochem 13(8), 11671173.CrossRefGoogle ScholarPubMed
Wilson, R (2013) Sensitivity and specificity: Twin goals of proteomics assays. Can they be combined?. Expert Review of Proteomics 10(2), 135149.CrossRefGoogle ScholarPubMed
Xie, J and Seto, CT (2007) A two stage click-based library of protein tyrosine phosphatase inhibitors. Bioorganic & Medicinal Chemistry 15(1), 458473.CrossRefGoogle ScholarPubMed
Zapotoczny, S, Biedroń, R, Marcinkiewicz, J and Nowakowska, M (2012) Atomic force microscopy–based molecular studies on the recognition of immunogenic chlorinated ovalbumin by macrophage receptors. Journal of Molecular Recognition 25(2), 8288.CrossRefGoogle ScholarPubMed
Zhang, C, Spokoyny, AM, Zou, Y, Simon, MD and Pentelute, BL (2013) Enzymatic “click” ligation: Selective cysteine modification in polypeptides enabled by promiscuous glutathione S-transferase. Angewandte Chemie 125(52), 1425114255.CrossRefGoogle Scholar
Zhang, C, Yang, J, Jiang, S, Liu, Y and Yan, H (2016) DNAzyme-based logic gate-mediated DNA self-assembly. Nano Letters 16(1), 736741.CrossRefGoogle ScholarPubMed
Zhang, G, Chen, X, Chen, X, Du, K, Ding, K, He, D, Ding, D, Hu, R, Qin, A and Tang, BZ (2023) Click-reaction-mediated chemotherapy and photothermal therapy synergistically inhibit breast cancer in mice. ACS Nano 17(15), 1480014813.CrossRefGoogle ScholarPubMed
Zhao, X, Fan, P-R, Mo, C-E, Huang, Y-P and Liu, Z-S (2020) Green synthesis of monolithic enzyme microreactor based on thiol-ene click reaction for enzymatic hydrolysis of protein. Journal of Chromatography A 1611, 460618.CrossRefGoogle ScholarPubMed
Zhong, X, Yan, J, Ding, X, Su, C, Xu, Y and Yang, M (2023) Recent advances in bioorthogonal click chemistry for enhanced PET and SPECT radiochemistry. Bioconjugate Chemistry 34(3), 457476.CrossRefGoogle ScholarPubMed
Zhou, P, Zhang, X, Xin, X, Yang, J, Pan, Q, Liu, C, Liu, Y, Yu, X, Li, Z and Jiao, G (2022) Click chemistry-conjugated protein-drug micelles with anti-ferroptotic and anti-inflammatory properties promote regeneration in spinal cord injury. Chemical Engineering Journal 428, 132118s.CrossRefGoogle Scholar
Figure 0

Figure 1. The number of published papers in the field of protein click chemistry with biological applications. (The method of extraction is fully described in the Supplementary Material.)

Figure 1

Figure 2. The most recently used click reactions in protein conjugations.

Figure 2

Figure 3. Targeted drug release via the click-to-release reaction using 3-bromo-1,2,4,5-tetrazine (Tz) for protein labeling. A: Activation of circulating inactive trans-cyclooct-2-en-1-yl (TCO)-drug conjugate by monosubstituted amino Tz reaction. B: Labeling of Trastuzumab using 3-bromo-1,2,4,5-tetrazine. C: The reaction of labeled Trastuzumab and TCO-Dox in BT474 (HER2+) cell culture to release drug (Ros et al., 2020).

Figure 3

Figure 4. Radiolabeling of human serum albumin (HSA) with 99mTc via chelate-then-click strain-promoted azide-alkyne cycloaddition (SPAAC) approach (Lodhi et al., 2019).

Figure 4

Figure 5. Pretargeting of SKBR3 cells with norbornene and tetramethylrhodamine co-labeled trastuzumab. Tagging the live cells with tetrazine-VT680 via an inverse electron demand using Diels–Alder coupling technique (Devaraj et al., 2008).

Figure 5

Figure 6. The electrochemical detection of protein–DNA interactions via azidophenyl as a click-transformable redox label of DNA (Balintová et al., 2015).

Figure 6

Figure 7. Peptide click chemistry strategies. Each strategy improves the stability and performance of peptides incorporating click-derived triazoles: Conjugation of multiple functional groups refers to the linking of multiple functional groups within a molecule. This can lead to the creation of complex structures with diverse properties. Cyclization in click chemistry refers to the formation of a cyclic compound through a chemical reaction. Protein cross-linking by click chemistry involves the covalent linking of proteins using highly efficient and selective reactions. By using stable and well-characterized building blocks, researchers can minimize the risk of encountering unstable bound surrogates and improve the efficiency and reliability of click chemistry reactions. The introduction of contact groups to improve affinity by click chemistry involves the strategic attachment of specific functional groups to a molecule to enhance its binding or interaction with a target.

Figure 7

Figure 8. Combination of enzymatic transamidation and click chemistry for one-pot peptide and protein covalent conjugation (Rachel and Pelletier, 2016).

Figure 8

Figure 9. Modification of polydivinylbenzene (PDVB) microspheres containing hydrophilic and/or hydrophobic polymer chains by a hydrobromination/click chemistry protocol (N,N,N′,N″,N″-pentamethyldiethylenetriamine (PMDETA), poly(ethylene glycol) (PEG), poly(methylmethacrylate) (PMMA), poly(tert-butylacrylate) (PtBA)) (Durmaz et al., 2011).

Supplementary material: File

Amiri et al. supplementary material

Amiri et al. supplementary material
Download Amiri et al. supplementary material(File)
File 52.3 KB