Next Article in Journal
Cytotoxic Xanthones from Hypericum stellatum, an Ethnomedicine in Southwest China
Next Article in Special Issue
Late-Stage Copper-Catalyzed Radiofluorination of an Arylboronic Ester Derivative of Atorvastatin
Previous Article in Journal
Comparative Study of Aryl O-, C-, and S-Mannopyranosides as Potential Adhesion Inhibitors toward Uropathogenic E. coli FimH
Previous Article in Special Issue
Fluorine-18 Labeled Fluorofuranylnorprogesterone ([18F]FFNP) and Dihydrotestosterone ([18F]FDHT) Prepared by “Fluorination on Sep-Pak” Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Bioorthogonal Click Chemistry for Efficient Synthesis of Radiotracers and Radiopharmaceuticals

1
Department of Nuclear Engineering, Pakistan Institute of Engineering and Applied Sciences, Islamabad 45650, Pakistan
2
IT Convergence Materials Group, Korea Institute of Industrial Technology, Cheonan 31056, Korea
3
Department of Applied Chemistry, School of Applied Chemical Engineering, Kyungpook National University, Daegu 41566, Korea
*
Author to whom correspondence should be addressed.
These authors contributed equally to this paper.
Molecules 2019, 24(19), 3567; https://doi.org/10.3390/molecules24193567
Submission received: 4 September 2019 / Revised: 23 September 2019 / Accepted: 27 September 2019 / Published: 2 October 2019
(This article belongs to the Special Issue Past, Present, and Future of Radiochemical Synthesis)

Abstract

:
In recent years, several catalyst-free site-specific reactions have been investigated for the efficient conjugation of biomolecules, nanomaterials, and living cells. Representative functional group pairs for these reactions include the following: (1) azide and cyclooctyne for strain-promoted cycloaddition reaction, (2) tetrazine and trans-alkene for inverse-electron-demand-Diels–Alder reaction, and (3) electrophilic heterocycles and cysteine for rapid condensation/addition reaction. Due to their excellent specificities and high reaction rates, these conjugation methods have been utilized for the labeling of radioisotopes (e.g., radiohalogens, radiometals) to various target molecules. The radiolabeled products prepared by these methods have been applied to preclinical research, such as in vivo molecular imaging, pharmacokinetic studies, and radiation therapy of cancer cells. In this review, we explain the basics of these chemical reactions and introduce their recent applications in the field of radiopharmacy and chemical biology. In addition, we discuss the significance, current challenges, and prospects of using bioorthogonal conjugation reactions.

1. Introduction

The term ‘click chemistry’ has been introduced to describe specific chemical reactions, which are fast, reliable and can be selectively applied to the synthesis of functional materials and biomolecule conjugates [1,2,3,4,5,6]. Click chemistry can be broadly defined as a ligation reaction in which two reactants are joined under ambient conditions to provide the desired product in high chemical yield and short time [7,8,9,10]. Over the last two decades, tremendous development and progress has been achieved in these conjugation reactions to encompass wide substrate scopes in the click reaction. Additionally, in several cases, these ligations proceed in aqueous media without significant decrease of the selectivity and reaction rate. Furthermore, click chemistries enable the facile isolation of the desired products from the reaction mixtures and facilitate the removal of the non-reacted substrates and byproducts, without the need for sophisticated separation methods [11,12,13,14,15,16]. Therefore, click chemistry-based conjugation methods have been applied to several avenues of research, including biochemical sciences, material sciences [17,18,19,20,21,22,23,24], drug discovery [25,26,27,28], pharmaceutical sciences [29,30,31,32,33,34], and synthesis of radiolabeled products [35,36,37,38,39,40,41]. Several typically used ligation reactions which are closely related to click chemistry include the thiol-Michael addition reaction [42], ring-opening reactions of aziridinium ions and epoxides [43], hydrazone and oxime formation from an aldehyde group [44] and so on. However, these reactions showed certain disadvantages such as poor specificity and stability under aqueous conditions, because of the reactivity of these functional groups with biomolecule residues and water. In 2003, K. B. Sharpless and M. G. Finn et al. reported that copper(I)-catalyzed azide-alkyne [3+2] cycloaddition reaction (CuAAC) can be employed as a new class of click reactions for rapid and reliable bioconjugation [45]. As both azide and alkyne groups are unreactive toward protein residues or other biomolecules, this ligation brought about a great impact and has been utilized as an efficient site-specific ligation methodology. Later, some researchers reported that the exogenous metals used to catalyze the click reaction (e.g., copper) could cause mild to severe cytotoxic effects and thus the use of metal catalyst-free chemical reaction has been recommended for several applications [46]. Therefore, catalyst-free, rapid, biocompatible, and bioorthogonal reactions such as strain-promoted azide-alkyne cycloaddition reaction (SPAAC) [47] and inverse-electron-demand Diels–Alder reaction (IEDDA) [48] have been developed as useful alternatives, and have been extensively used in various research fields (Figure 1).
In recent years, these conjugation reactions have also been applied to the synthesis of radioisotope-labeled molecules, which have been used for nuclear imaging using positron emission tomography (PET) and single-photon emission computed tomography (SPECT) as well as for therapeutic applications. Particularly, several important diagnostic radioisotopes including 11C (t1/2 = 20 min), 18F (t1/2 = 110 min), 99mTc (t1/2 = 360 min), and 68Ga (t1/2 = 68 min) have short half-lives, and thus their radiolabeling procedures require rapid and efficient reactions which can provide reliable radiochemical results, such as high radiochemical yield (RCY) and purity, and minimal undesired by-product formation [49]. In this regard, the catalyst-free click reactions can be highly useful tools for radiolabeling complex small molecules and biomacromolecules, which are sensitive to harsh reaction conditions such as elevated temperatures, extreme pH, and the presence of metal catalysts [50]. In addition to in vitro radiolabeling applications, these ligation methods have also been investigated for in vivo pre-targeted strategies for specific imaging and cancer therapy in animal xenograft models [51].
This review aims to highlight the recent and noteworthy results for the synthesis of radiolabeled molecules using site-specific click reactions. In detail, this review will mainly focus on the following bioconjugation reactions: (1) strain-promoted azide-alkyne cycloaddition (SPAAC); (2) inverse-electron-demand Diels–Alder cycloaddition reaction (IEDDA); (3) rapid condensation/cycloaddition reactions based on electrophilic heterocycles. The review will also showcase the advantages of these reactions, which have empowered radiochemists in the production of radiolabeled products and radiopharmaceuticals for imaging and therapeutic purposes. Finally, future directions and emerging trends of these ligation methods will be discussed.

2. Strained Promoted Copper-Free Click Reaction for Synthesis of Radiolabeled Molecules

In Vitro Radiolabeling of Biomolecules

In SPAAC, the ring strain of cyclic alkynes such as dibenzocyclooctyne (DBCO) is used to drive the reaction with azide groups in the absence of copper(I) catalysis [52,53]. Generally, two strategies have been employed for SPAAC-based radiolabeling. The first is the synthesis of radiolabeled cyclooctyne precursors, which can be used for the labeling of azide containing biomolecules, and the other is the preparation of radioisotope-tagged azide tracers which are reacted with cyclooctyne modified biomolecules. In 2011, Feringa group investigated SPAAC reaction for the efficient 18F-labeling of biomolecules [54]. In this study, three 18F-labeled azides were synthesized, and the prepared tracers were conjugated with DBCO modified bombesin peptide derivatives. Notably, the reaction proceeded with high efficiency to provide 18F-labeled cancer-targeting peptides in 15 min with good radiochemical yields (RCYs) (Figure 2). Particularly, radiolabeling studies using these reactions were also explored in human plasma to determine their reactivity and specificity in biological media.
Wuest et al. reported the synthesis of an 18F-labeled DBCO analog for efficient preparation of a diagnostic probe. (Table 1, entry 1) This tracer was reacted with several azide group-bearing geldanamycine moieties and carbohydrates to furnish the corresponding 18F-labeled products. Importantly, these radiolabeling reactions were performed in various media, including in methanol, DMSO/water (1:1), and bovine serum albumin, wherein the observed RCYs did not decrease significantly [55]. Along similar lines, Carpenter et al. used a modified 18F-labeled DBCO analog for the radiolabeling of azide conjugated substrates (Table 1, entry 2). The radiolabeling was performed at room temperature in N,N-dimethylformamide (DMF) to afford the desired radiolabeled products [56]. The peptide A20FMDV2 (Table 1, entry 3), which has a strong binding affinity with integrin αvβ6-receptor, was successfully labeled with 18F, using a SPAAC-based ligation. The radiolabeling of the azide group bearing A20FMDV2 was performed at ambient temperature to give the product in 11% of isolated RCY. The radiolabeled peptide was highly stable in rat serum, and its binding affinity towards the target receptor was not affected. However, in vivo studies revealed its decreased targeting ability due to the structural differences and increased lipophilicity compared to the parent structure [57]. Several other 18F-DBCO analogs have shown good RCY for the preparation of radiolabeled peptides for targeting cancer [58,59]. To improve the efficiency of 18F radiolabeling, Roche et al. explored a new 18F-labeled azide prosthetic group, 18F-FPyZIDE (Table 1, entry 6). In their study, the radiolabeled tracers were evaluated in both CuAAC- and SPAAC-based ligations and the labeling results showed that both radiolabeling methods provided high RCYs under mild reaction conditions (room temperature or 40 °C) [60]. Evans et al. investigated the radiosynthesis of 68Ga-labeled peptide using azide group-bearing 1,4,7,10-tetraazacyclododecane-tetraacetic acid (DOTA) chelator and DBCO group conjugated cRGD peptide (Table 1, entry 7) [61]. The developed novel bioorthogonal click reaction has been used in the design and preparation of multimodal imaging tracers. Ghosh et al. studied a dual-modal scaffold in which the precursor was first labeled with 68Ga using a DOTA chelator, and then, a near-infrared (NIR)-absorbing fluorescent dye, IR Dye 800CW, was incorporated into the tracer using SPAAC ligation. The dual-labeled tracer was then applied to the targeted imaging of a somatostatin receptor and the quantification of its biological uptake in vivo (Table 1, entry 8) [62].
Generally, most SPAAC ligations based on DBCO derivatives display second-order rate constants in the 1–2 M−1 s−1 range with azide groups [63] due to which, the observed RCY is not satisfactory when using low substrate concentrations. To improve reaction kinetics, 18F-labeled oxa-dibenzocyclooctyne (ODIBO), which has a k2 value of 45 M−1 s−1, was synthesized to label azide containing biomolecules with high efficiency (Figure 3) [64,65]. This new prosthetic group enabled site-specific radiolabeling using much smaller amounts (about one-tenth) of azide bearing molecules than those of DBCO-based reactions.
The study reported by Kim et al. used SPAAC ligation in both radiolabeling reaction and purification steps [66]. For this application, 18F-labeled azide tracer was first reacted with DBCO- modified cancer-targeting peptide (cRGD) and then the desired product was separated from unreacted peptide substrates using an azide modified resin as a scavenger for the DBCO group. The remarkable two-steps process provided the radiolabeled peptide in high decay-corrected RCY (92%) and radiochemical purity (98%) (Figure 4). Notably, PET imaging and biodistribution data confirmed the high tumor uptake value of the 18F-labeled peptides in U87MG xenograft along with significant enhancement of the tumor to background ratio [67].
SPAAC reaction has also been applied to the labeling of radioisotopes with longer half-lives, such as radioactive metals and radioactive iodine. We reported the use of 125I-labeled azide prosthetic groups for synthesizing radiolabeled biomolecules and nanomaterials. In this process, DBCO group modified cRGD peptides were efficiently conjugated with 125I-labeled azides in high RCY and radiochemical purity after HPLC purification (Figure 5) [68,69]. It was reported that 64Cu could be labeled with cross-bridged cyclam chelator CB-TE1K1P under mild conditions. To employ this chelator for radiolabeling biomolecules, Anderson et al. synthesized a DBCO modified chelator (DBCO-PEG4-CB-TE1K1P) and reacted it with an azide-bearing Cetuximab by SPAAC ligation (Figure 6). The 64Cu labeling proceeded in high RCY (>95%) at 37 °C, and the radiolabeled antibody showed enhanced serum stability when compared with those of previously reported 64Cu chelators [70].
Yuan et al. explored the synthesis of 89Zr-labeled PET imaging agents using SPAAC ligation on the surface of the superparamagnetic feraheme (FH). For this study, azide-functionalized FH nanoparticles were prepared and were mixed with 89Zr under elevated temperature to deliver the 89Zr-labeled azide-FH. In the next step, DBCO-conjugated RGD peptide, or Cy5.5 tagged protamine was reacted with 89Zr-azide-FH to give the desired radiolabeled products with good radiochemical results and specific radioactivity (Figure 7) [71]. This strategy provided an efficient approach for the preparation of multimodal/multifunctional nanoprobes, which are suitable for a wide range of diagnostic and therapeutic applications.
Several kinds of liposomes are known to be useful vehicles for targeted delivery in biomedical research as well as clinically approved platforms [72]. Hood and co-workers used SPAAC ligation for efficient conjugation between 111In-labeled liposomes and single-chain variable fragments (scFv) or monoclonal antibodies. The radiolabeled tracer, 111In-liposomes-mAb/scFv, was used in the targeted imaging of the platelet-endothelial cell adhesion molecule (PECAM-1) and intracellular adhesion molecule (ICAM-1). The uptake value of 111In-liposomes/scFv into the target cells was much higher than that of 111In-liposomes/mAb [73]. Recently, thermosensitive hydrogels comprising polyisocyanopeptide (PIC) were labeled with 111In via a SPAAC method. In this research, azide-modified PIC hydrogel was first conjugated with DBCO-modified diethylenetriaminepentaacetic acid (DTPA) chelator to afford the PIC-DTPA conjugate. Next, PIC-DTPA was reacted with 111InCl3 to give the 111In-labeled PIC in high RCY. The radiolabeled PIC was applied in a SPECT imaging study for evaluating the efficacy of PIC gels in wound mouse models [74]. Figure 8 shows the 99mTc labeling of human serum albumin (HSA) via a SPAAC reaction. After labeling 99mTc(CO)3 with an azide group-modified dipyridine chelator, it was then reacted with ADIBO bearing HSA under mild condition to give the radiolabeled protein in high RCY (76–99%). The 99mTc-labeled HSA prepared by this procedure showed better stability in vivo, as compared with those previously reported 99mTc-labeled HSA, which were obtained by direct 99mTc labeling [75]. The radiolabeled HSA thus prepared, was used in blood pool imaging using SPECT.

3. Inverse-Electron-Demand Diels–Alder Reaction for Synthesis of Radiolabeled Molecules

3.1. In Vitro Radiolabeling of Biomolecules

The inverse electron demand Diels–Alder (IEDDA) between 1,2,4,5-tetrazine and strained alkene (such as trans-cyclooctene, TCO) is a well-established bioorthogonal reaction, which is typically regarded as the fastest click reaction with first-order rate constants ranging up to 105 M−1 S−1 [76,77,78,79] Since the first report on IEDDA reaction, several kinds of strained alkenes/alkynes and tetrazine analogs have been synthesized, and these functional group pairs have been applied to the radiolabeling of various small molecules, biomolecules, and nanomaterials [80,81]. Due to the extremely rapid reaction rate of IEDDA under mild conditions such as room temperature, neutral pH, and in aqueous media, this reaction has been a highly useful ligation approach for labeling radioisotopes with short half-lives. In 2010, Fox et al. reported the IEDDA-mediated 18F-labeling of small molecules. The radiolabeled TCO (Table 2, entry 1) could be synthesized by a nucleophilic substitution reaction of the tosylated precursor in 71% RCY. Remarkably, the IEDDA reaction between a model tetrazine substrate and 18F-labeled TCO provided the desired product in more than 98% RCY in 10 seconds [82]. Conti et al. applied IEEDA to the synthesis of an 18F-labeled cancer-targeting peptide [83]. The labeling reaction of a tetrazine conjugated cRGD peptide was carried out using an 18F-labeled TCO analog, which was prepared using a similar protocol, and delivered the radiolabeled peptide in excellent RCY (Figure 9). The 18F-labeled cRGD thus prepared, was evaluated in the U87MG xenograft model and exhibited clear visualization of tumor cells by PET imaging.
Later, the same group reported a maleimide-conjugated tetrazine analog, which was used to incorporate the tetrazine group onto biomolecules comprising free cysteine moieties. The tetrazine bearing biomolecules (cRGD peptide and VEGF protein) prepared by the above method was then reacted with 18F-labeled TCO to give PET-imaging tracers for diagnosis of cancer cells in vivo (Table 2, entry 2) [84]. Weissleder and coworkers synthesized 18F-AZD2281, a poly-ADP-ribose-polymerase 1, as a PET imaging tracer (Table 2, entry 3). In this report, 18F-labeled TCO and a tetrazine group-bearing AZD2281 were incubated for 3 minutes, and the crude product was purified using a magnetic TCO-scavenger resin for removing the unreacted substrate, without the need for carrying out the traditional HPLC purification. The process delivered the 18F-labeled AZD2281 in 92% RCY using the scavenger-assisted method [85]. The prepared radiolabeled tracer was then evaluated in xenograft models to visualize MDA-MB-436 tumors. Wu and coworkers extended the application of IEDDA ligation to the radiolabeling of the exendin-4 peptide and applied the 18F-labeled exendin-4 to the targeted imaging of GLP-1R receptor in an animal model [86]. In 2015, the Schirrmacher group reported the novel silicon-fluoride acceptor (SiFA) labeling method, which is based on an isotopic exchange reaction (Table 2, entry 5). This simple labeling step (19F →18F), which is based on a silicon-fluorine scaffold, provided the 18F-labeled tetrazine in much higher RCY (78%) than those realized with other 18F chemistries [87].
Norbornene analogs are known to be reactive toward tetrazines. Although the reaction rate was much slower than those of TCO analogs, the preparation of a norbornene substrate is straightforward. Importantly, norbornene analogs are known to be more stable than TCO analogs, which are prone to isomerization to their cis-isomers under physiological conditions. Knight and coworkers reported the reaction of the tetrazine group-conjugated bombesin peptide with an 18F-labeled norbornene prosthetic group, to provide the radiolabeled product with high efficiency and radiochemical purity (Table 2, entry 6) [88]. In addition to the radioactive fluoride, 11C is another important cyclotron-produced radioisotope for preclinical and clinical PET imaging. Particularly, the 11C-labeled methyl triflate and methyl iodide are the most prominent synthons for nucleophilic methylation of alcohols, amines, and thiols, which are commonly used for the production of various radiotracers and radiopharmaceuticals. Herth and coworkers reported the first synthesis of an 11C-labeled tetrazine and its reaction with a strained cyclooctene (Table 2, entry 7) [89]. The radioactive precursor [11C]CH3I was reacted with a tetrazine-conjugated phenol group to give the desired radiolabeled tetrazine in 33% RCY, which underwent a click reaction with a trans-cyclooctenol in 20 seconds, suggesting the suitability of this conjugation method for preparation of radiolabeled molecules with short-lived isotopes such as 11C. Devaraj et al. reported the development of a 68Ga-labeled tetrazine modified dextran polymer for increasing the half-life and in vivo stability of the tracer in blood (Table 2, entry 8), and evaluated its use in human colon cancer cells (LS174T) and xenograft models [90].
Radioactive iodines have been used for the preparation of various radiotracers for in vivo imaging and biodistribution studies. The traditional radioiodination method via an electrophilic substitution reaction typically provides high RCY in a short time. However, the radiolabeled tracer synthesized using the above reaction generally exhibited considerable deiodination in the living subjects, and the liberated radioactive iodines rapidly accumulated in the thyroid and stomach which and resulted in high background signals in the images. Moreover, the use of a strong oxidant that requires radioiodination often resulted in decreased biological activity of the molecules. To address these problems, the radioactive iodine-labeled tetrazine can be used as an alternative method for the efficient radiolabeling of biomolecules. Valliant et al. reported the rapid radiolabeling of antibody based on IEDDA. In this study, the 125I-labeled tetrazine analog was incubated with the TCO-modified anti-VEGFR2 for 5 minutes to afford the desired product in 69% RCY. Interestingly, the radiolabeled antibody, which was prepared using this procedure, displayed a 10-fold increase in stability to in vivo deiodination, then the same antibody prepared by direct radioiodination using iodogen (Table 2, entry 9) [91]. Along similar lines, we investigated a modified 125I-labeled tetrazine tracer via oxidative halo-destannylation of the corresponding precursor (Table 2, entry 10) [92]. The prepared radiolabeled tetrazine was then applied to the labeling of TCO derived cRGD peptide and human serum albumin (HSA) and delivered excellent RCYs (>99%). The biodistribution study of the 125I-labeled HSA in normal ICR mice demonstrated enhanced in vivo stability toward deiodination than the radiolabeled HSA obtained using the conventional iodination method. Valliant group also synthesized 123/125I-labeled carborane-tetrazine and employed it for the radiolabeling of TCO-bound H520 cells [93].
Several radioactive metal-labeled tracers have also been prepared by IEDDA ligation for diagnostic purposes. Lewis et al. reported tetrazine conjugated metal-chelating agents such as DOTA and deferoxamine (DFO) for the radiolabeling of norbornene bearing trastuzumab using 64Cu or 89Zr (Figure 10) [94]. By using this procedure, radiolabeled trastuzumab was obtained in high RCY (>80%) and high specific radioactivity (>2.9 mCi/mg). Furthermore, PET imaging studies demonstrated that radiolabeled antibodies were quite stable in vivo conditions and showed specific uptake in HER2-positive BT-474 tumor cells. In 2018, IEDDA ligation was employed for the preparation of therapeutic radioisotope-labeled human antibodies 5B1 and huA33 (Figure 11) [95]. In this study, a tetrazine conjugated DOTA chelator was synthesized, and labeled with 225Ac, a useful therapeutic radioisotope. The radiolabeled tetrazine tracer was then reacted with TCO-modified antibodies to give the desired products within 5 min. This two-step method provided superior RCYs compared to the conventional approaches used in clinical applications. In addition, the biodistribution results demonstrated that the 225Ac-labeled antibody showed high tumor uptake values and relatively low non-specific accumulation in normal organs.
Recently, Long et al. reported the radiolabeling process for microbubble, which is a contrast agent used in ultrasound imaging and relies on an IEDDA reaction for its operation. First, a tetrazine-bearing metal chelator (HBED-CC) was labeled with 68Ga. The TCO-modified phospholipids were then treated with 68Ga-HBED-CC-tetrazine under mild conditions to give the 68Ga-labeled lipid molecule (68Ga-PE). Next, the prepared 68Ga-PE was combined with other types of lipids, and the resulting formulation was activated to form gas-filled microbubbles (Figure 12). This strategy enabled the PET-based real-time monitoring and pharmacokinetic study of newly developed contrast agents for ultrasound analysis [96].

3.2. In Vivo Pre-Targeted Imaging and Therapy

The tetrazine and TCO groups are not reactive towards amine or thiol nucleophiles and show high reaction specificity in biological media. In addition, IEDDA can proceed with fast kinetics even at very low reactant concentrations. Due to these reasons, IEDDA-based ligation is one of the most potent tools for pre-targeting applications among the existing click reaction approaches. In the pre-targeted approach, a cancer-targeting ligand and a radiolabeled small molecule are administered separately into a living subject. Generally, the TCO (or tetrazine) conjugated tumor-targeting antibody is injected first into the tumor xenograft model and is allowed to accumulate in the tumor cells for a certain period (Figure 13). Next, the radiolabeled tetrazine (or TCO) group is administered after the excess amount of antibody in healthy tissues is excreted from the body. The in vivo click reaction through the above procedure decreases the circulation time of the radioligand and results in reduced non-specific uptake of radioactivity in healthy tissues. Furthermore, this approach also facilitates the delivery of radioisotopes with short half-lives, which would not be feasible with antibody-based imaging studies [97]. Table 3 shows in vivo pre-targeted studies that were conducted using IEDDA-based ligation in animal models.
In 2010, the Robillard group reported their pioneering work on in vivo pre-targeted imaging of cancers using IEDDA ligation [98]. In the first step, TCO group bearing CC49 antibody was injected to target colon cancer cells in a mouse model. Post administration of the antibody (24 h), only a small excess (3.4 equivalent) amount of 111In-labeled tetrazine tracer was injected into the same mouse model. The obtained SPECT images showed the efficient delivery of the radioisotope into the tumor and indicated a high tumor-to-normal tissue ratio (Table 3, entry 1). In the next study, the same research group revealed that TCO could be converted to its (Z)-isomer, which is unreactive to tetrazine in the presence of copper-containing proteins [99]. Thus, a shorter linker was introduced in the tetrazine tracer to impede interactions with the copper-containing proteins in albumin. By this structural modification, the reactivity and isomerization half-life of TCO was increased compared to that of the previously used TCO analog. Later, Robillard and coworkers reported the use of tetrazine-functionalized clearing agents as a modified pre-targeting system (Table 3, entry 3) [100]. While a portion of the administered antibody accumulated in the tumor tissue in this approach, a significant portion of it still remained in the blood. This accumulated antibody could cause a reduced target-to-background ratio because IEDDA reaction is also feasible at non-specific areas in the body. To address this problem, the group added one more step in the animal study (Figure 14). After the TCO-modified antibody was injected into the xenograft model to target tumor cells in vivo, the tetrazine bearing galactose-albumin conjugate was injected as a TCO clearance agent to mask the unbound TCO-modified antibody in the blood. The radiolabeled tetrazine was then injected to enable the IEDDA reaction at the surface of tumor site. This approach demonstrated that the use of a clearing agent could lead to the doubling of the tetrazine tumor uptake and a greater than 100-fold improvement of the tumor-to-blood ratio at 3 h could be realized after injection of the radiolabeled tetrazine.
The same group reported a pre-targeted radioimmunotherapy study using a similar strategy. To achieve high tumor uptake and improved tumor-to-blood ratio, the group employed a linker with higher hydrophilicity to prepare the TCO-tagged CC49 antibody [101]. In 2015, TCO-functionalized diabody, AVP04-07 was evaluated in the pre-targeted strategy [102]. In this study, the TAG72-targeting dimers of single-chain Fv fragments and 177Lu-labeled tetrazine tracers were evaluated in the LS174T tumor xenograft. As the diabody showed rapid renal clearance kinetics, this strategy could provide high tumor-to-blood ratio and low non-specific retention in the kidneys. In a related study, the authors successfully performed an IEDDA-based pre-targeted study by employing HER2 affibody molecules and 111In/177Lu-labeled tetrazine tracers (Table 3, entry 6) [103].
In addition to these results, several research groups have investigated a variety of pre-targeted approaches using short half-life radioisotope-labeled TCO or tetrazine derivatives. Denk et al. developed a novel 18F-labeled tetrazine by the direct 18F-fluorination of the tosylated precursor, which proceeded in an RCY up to 18% (Table 3, entry 7) [104]. The PET imaging study exhibited fast homogeneous biodistribution of the 18F-labeled tetrazine, which can also cross the blood–brain barrier. The high reactivity of this tracer towards TCO-bearing molecules and favorable pharmacokinetic properties indicated that 18F-labeled tetrazine can be a useful tracer for bioorthogonal PET imaging. Lewis et al., reported 18F-based pre-targeted PET imaging studies using TCO-modified anti-CA19.9 antibody 5B1 and a 1,4,7-triazacyclononane-1,4,7-triacetic acid (NOTA)-conjugated tetrazine analog. The complexation reaction using AlCl3 and [18F]F provided the desired radioligand in 54–65% decay-corrected RCY [105]. The in vivo pre-targeted images displayed its effective targeting ability with radioactivity up to 6.4% ID/g in the tumors at 4 h post administration. Sarparanta and co-workers investigated in vivo IEDDA reaction between TCO conjugated monoclonal antibodies and 18F-labeled tetrazine molecule [106]. For this study, TCO conjugated antibody (trastuzumab and cetuximab) was injected into tumor-bearing (BT-474 cells and A431 cells) mice and the 18F-labeled tetrazine-containing hydrophilic linker was injected into the same xenograft models after given time points (1, 2, or 3 days). The highest tumor-to-background ratio was observed when the radioisotope was injected after 3 days post the administration of the TCO-modified antibody. In addition, the 18F-labeled tetrazine was applied to the pre-targeted in vivo imaging of TCO-modified porous silicon nanoparticles (Table 3, entry 10) [107]. Bormans et al. developed a new 18F-labeled TCO tracer for in vivo IEDDA. To prepare radiolabeled TCO, the authors synthesized a dioxolane-fused TCO analog from its cis isomer by using a micro-flow photochemistry process (Figure 15) [108]. The nucleophilic substitution of mesylated precursor using dry K[18F]F, K222 complex provided the desired 18F-labeled tracer in 12% RCY and >99% radiochemical purity. This product showed excellent reactivity and stability toward a tetrazine and thus it was applied to pre-targeted PET Imaging. In this approach, a tetrazine-modified trastuzumab monoclonal antibody was injected initially into SKOV-3 xenograft models (Figure 16). After 2 or 3 days, the 18F-labeled TCO was injected, and then the PET images were obtained after 2 h post the administration of the radioligand. The obtained results showed that the pre-targeted imaging strategy provided better tumor-to-muscle ratio when compared to that of control groups which did not use the pre-targeting approach (Table 3, entry 12) [109].
The radiolabeled small molecule tracers often underwent rapid renal or hepatobiliary clearance, and therefore, the efficiency of in vivo click reactions is reduced. To increase the blood circulation time of the functional group, the Weissleder group designed the tetrazine group-bearing polymers comprising dextran scaffolds [110]. An 18F-labeled polymer-modified tetrazine and TCO-bearing CD45 monoclonal antibodies were investigated in a living mouse, and the PET imaging study revealed excellent conversion of reactants and high tumor uptake in the tumor xenograft, which suggested that the radiolabeled polymer will be a promising candidate for pre-targeted imaging. The use of IEDDA for pre-targeted PET imaging has also been investigated with 11C. In 2016, Mikula et al. reported the use of 11C-labeled tetrazine for in vivo click reaction. An amino tetrazine analog was reacted with [11C]CH3OTf to provide 11C-labeled tetrazine in 52% of RCY (Figure 17) [111], and the resulting product exhibited high reaction rate with TCO derivatives. Furthermore, the product also showed good stability under physiological conditions and demonstrated rapid clearance kinetics in mice. This 11C-labeled tracer was then applied to animal imaging studies with TCO-modified mesoporous silica nanoparticles in normal mice. Herth et al. reported the improved radiosynthesis of 11C-labeled tetrazine for pre-targeted PET imaging (Table 3, entry 15) [112]. In this study, the radioligand was evaluated with TCO-functionalized polyglutamic acid and indicated potential use for brain imaging.
The use of the TCO-tetrazine ligation in living subjects was also extended to several metal radioisotopes. Lewis et al. reported a pre-targeted strategy using a TCO-bearing huA33 antibody and 64Cu-labeled NOTA-tetrazine conjugate (Table 3, entry 16). In this study, a tumor-targeted antibody was administered to SW1222 colorectal cancer xenografts, and the tetrazine tracer was then injected one-day post administration of the antibody. This approach exhibited enhanced tumor-to-background ratio and reduced non-specific radiation dose in normal tissues [113]. In the following study, the authors reported a site-specific conjugation method to construct the huA33-TCO immunoconjugate by using enzymatic transformations and a bifunctional linker [114]. A similar bioconjugation strategy was also applied to the preparation of a TCO and fluorescent dye-bearing antibody (huA33-Dye800-TCO) for bimodal PET/optical pre-targeted imaging of colorectal cancer cells (Table 3, entry 19) [116] using a 64Cu-sarcophagine-based tetrazine tracer. This strategy demonstrated the non-invasive visualization of tumors and the image-guided excision of malignant tumor tissue. Aboagye et al., synthesized 68Ga-labeled tetrazine to study its use in pre-targeted PET imaging of EGFR-expressing A431 tumor. After administration of the TCO-bearing cetuximab, the 68Ga-labeled tracer was injected to the mouse model, and PET imaging showed a significant improvement in the tumor-to-background ratio compared to that with the traditional direct radiolabeling method [117]. Recently, Lewis et al. used a modified pre-targeted PET imaging strategy for obtaining a better tumor-to-blood ratio. The authors employed a tetrazine-modified dextran polymer to reduce injected TCO-bearing antibody, which remained in blood circulation. After the TCO-modified antibody was injected into the xenograft model to target tumor cells, the TCO scavenger was administrated to mask unbounded TCO modified antibody in the blood. Next, 68Ga-labeled tetrazine radioligand was injected to allow the IEDDA reaction at the surface of tumor cells. Further, the use of the TCO masking agent in this study showed a significant improvement in the PET image quality and tumor-to-background ratio (Table 3, entry 21) [118].
The Valliant group demonstrated a pre-targeted strategy for bone imaging and radiotherapy based on the IEDDA between the TCO-conjugated bisphosphonate and radiolabeled tetrazines (Figure 18) [119]. In this experiment, TCO-bisphosphonate conjugate was first injected into an animal model for accumulation of the dienophile in the skeleton. After 12 h post administration, 99mTc-labeled tetrazine was administered intravenously, and the acquired SPECT/CT imaging revealed high radioactivity in the knees and shoulder, which suggested that the TCO-bisphosphonate can be a useful probe for targeting functionalized tetrazine in the bone tissue. A therapeutic radioisotope (177Lu)-labeled radioligand was also investigated in the same study.
In 2018, Garcia et al. investigated an antibody pre-targeting approach using TCO-bearing bevacizumab and 99mTc-labeled tetrazine tracer. To increase renal clearance kinetics of the radioisotope, a hydrophilic peptide linker was introduced between tetrazine and the 6-hydrazinonicotinyl group, which is a well-known chelator of 99mTc. The pre-targeted bevacizumab SPECT imaging was then investigated in B16-F10 melanoma cell’s xenograft [120]. In addition to various diagnostic research, the alpha-particle emitting radioisotope (212Pb) was applied to the pre-targeted radioimmunotherapy by Quinn et al. (Table 4, entry 24). In this study, the LS174T tumor-bearing mice were injected with CC49-TCO monoclonal antibody. Two doses of the tetrazine bearing Galactose-albumin as a TCO clearing agent were injected after 30 and 48 h to remove the unbound antibodies in blood and normal organs. Then, 212Pb-labeled tetrazine was injected for targeted tumor therapy. This pre-targeted alpha-particle therapy successfully reduced the tumor growth and improved the survival of model mice [121].

4. Other Click Reactions Based on Aromatic Prosthetic Groups

4.1. Condensation/Addition Reactions Using Aromatic Compounds

As shown in previous sections, SPAAC and IEDDA have been two of the most frequently used radiolabeling methods for several years. To apply these reactions in the labeling procedure, the target molecule (e.g., peptide, antibody) needs to be modified to incorporate an artificial functional group, which is then reacted with the radiolabeled prosthetic group. For example, a TCO analog needs to be conjugated with the target molecule, to facilitate its reaction with a radioisotope-containing tetrazine. Such modification of biomolecule requires additional synthetic, and purification steps. Furthermore, the presence of excess amounts of randomly conjugated functional groups can cause decreased biological activities of the molecules. Therefore, several labeling procedures, which do not involve a modification of the biomolecules, have been developed. In many cases, these methods utilized electrophilic aromatic prosthetic groups that displayed rapid reaction rates and high selectivities toward a specific nucleophile such as thiol or 1,2-amino thiol. Table 4 summarizes recent studies on the applications of aromatic prosthetic groups for radiolabeling reactions.
In 2012, Jeon et al. investigated the rapid condensation reaction between 18F-labeled 2-cyanobenzothiazole (18F-CBT) and N-terminal cysteine-bearing biomolecules (Table 4, entry 1) [122]. The 18F-CBT was synthesized from the corresponding tosylated precursor using K[18F]F and 18-crown-6 as the phase transfer catalyst. This radiolabeled CBT (18F-CBT) can be reacted with N-terminal cysteine with a second-order reaction rate of ca. 9 M−1 s−1. The rapid condensation reaction between the N-terminal cysteine-bearing dimeric cRGD peptide and the 18F-CBT provided the 18F-labeled peptide (18F-CBT-RGD2) in a high (>80%) RCY under mild conditions, and the prepared 18F-CBT-RGD2 was investigated for its use in specific tumor imaging in U87MG xenograft models. Later, 18F-CBT was also applied for the efficient radiolabeling of EGFR-targeting affibody molecules (ZEFGR:1907), and the radiolabeled affibody provided clear visualization of the A431 tumors in animal models [123]. As the heterocyclic adducts, which result from the condensation reaction between CBT and N-terminal cysteine are hydrophobic, the injected tracers prepared by the above method showed high non-specific uptake in normal organs. Therefore, the Seimbille group synthesized a more hydrophilic 18F-labeled CBT tracer containing a diethylene glycol linker and 2-fluoropyridine moiety (Table 4, entry 2) The optimized radiolabeling condition provided 18F-labeled cancer-targeting peptide, which is more hydrophilic than the ones reported in the previous studies [124]. The same research group also reported the synthesis of the metal-chelating agent-conjugated CBT prosthetic groups for 68Ga-labeled tracers for PET imaging of tumor hypoxia [125]. The rapid condensation for radiolabeling procedure provided the desired radiotracers in high RCY under mild conditions (Table 4, entry 3). In another study, the same group synthesized the two bifunctional chelators, the desferrioxamine B-bearing CBT (DFO-CBT) and the cysteine-bearing CBT (DFO-Cys) for efficient radiolabeling. These chelators were employed in the labeling with the [89Zr]Zr-oxalate and rapid conjugation with cRGD peptide. The two-step radiochemical process exhibited high RCY under mild reaction conditions [126]. As CBT structure contained a hydroxy group, it can be a good substrate for facile labeling of radioactive iodines [127]. Thus, we synthesized a 125I-labeled CBT (125I-CBT) via electrophilic iodination reaction under mild reaction conditions. The 125I-CBT was then applied to the rapid radiolabeling of N-terminal cysteine-bearing cRGD peptide in high RCY (Table 4, entry 4).
In 2013, Barbas III group reported the chemoselective ligation of thiol-containing proteins using methylsulfonyl derivatives [128]. They showed that phenyloxadiazole methylsulfone and phenyltetrazole methylsulfone react rapidly and selectively with the sulfhydryl group of cysteine residues in aqueous media under mild conditions (at neutral pH and room temperature) In addition, the structures resulting from these ligation reactions were more stable under physiological conditions in comparison to the corresponding products obtained by maleimide-thiol chemistry. These advantages lead to the development of new prosthetic groups for site-specific radiolabeling reactions. Mindt et al. reported a 18F-labeled phenyloxadiazole methylsulfone analog([18F]FPOS) for the rapid and chemoselective radiolabeling of thiol-bearing biomolecules under mild conditions (Table 4, entry 5) [129]. In this study, [18F]FPOS was applied to efficient radiolabeling of free thiol group-bearing biomolecules. The radiolabeled affibody (ZHER2:2395) could be successfully applied to the PET imaging of HER2-positive tumor cells in animal models. Recently, we have reported a radioiodinated phenyltetrazole methylsulfone derivative as a new thiol-reactive prosthetic group (Table 4, entry 6) [130]. The 125I-labeled (4-(5-methane-sulfonyl-[1,2,3,4]tetrazole-1-yl)-phenol) (125I-MSTP) can be prepared by using a simple iodination reaction from the phenolic precursor in high RCY (73% isolated yield) and purity (>99%). The 125I-MSTP was used for site-specific radiolabeling of a single free-thiol-bearing peptide and protein by using radioiodinated labeling of thiol-containing biomolecules. The radiolabeled HSA prepared by this method exhibited enhanced in vivo stability upon deiodination compared with radioiodinated HSA prepared by a direct iodination reaction. In 2018, Park et al. reported a novel condensation reaction using an aryl diamine linker and 125I-labeled aldehyde prosthetic group (Table 4, entry 7) [131]. This method was applied to rapid and efficient radiolabeling of bioactive molecules and the labeled products showed high in vitro and in vivo stability. Samnick et al. proposed a new phenol-reactive prosthetic group for site-specific radiolabeling reaction of tyrosine-containing biomolecules (Table 4, entry 8) [132]. The 18F-labeled 1,2,4-triazoline-3,5-dione([18F]FS-PTAD) was reacted with the model compounds such as phenol, l-tyrosine and N-acetyl-l-tyrosine methyl amide to evaluate the efficacy of the labeling reaction, which proceeded rapidly under mild aqueous conditions to furnish the corresponding radiolabeled compounds in good RCY (45–58%) within 5 min.

4.2. Miscellaneous

Neumaier and coworkers demonstrated the 18F-radiolabeling of biomolecules using [3+2] cycloaddition reactions between 18F-labeled nitrone and maleimide-bearing molecules [133]. This reaction can provide high efficiency for the synthesis of radiolabeled small molecules. However, the cycloaddition reaction required elevated temperatures in organic solvents, and thus, this method was not suitable for radiolabeling of proteins or antibodies. Continuing this theme, the same group explored more efficient [3+2] cycloaddition reactions using 18F-labeled nitriloxides and N-hydroxyimidoyl chloride (Figure 19). Interestingly, these radiolabeled tracers showed high reactivity with a strained alkene and norbornene analogs under ambient temperature, suggesting that this method can be a useful alternative to the copper-free azide–alkyne click reactions for the radiolabeling of biomolecules [134].
Recently, Wuest et al. demonstrated the first application of the sulfo-click chemistry in the 18F-labeling reaction (Figure 20). In this study, 18F-labeled thiol acids were synthesized and treated with sulfonyl azide-modified small molecules and peptide substrates to afford the corresponding radiolabeled products in moderate to good RCYs [135]. Furthermore, this labeling reaction can be selectively performed in aqueous solvents with a high degree of functional group compatibility.
Recently, a novel photochemical conjugation reaction was developed for one-pot radiolabeling of antibodies by the Holland group [136,137]. The group synthesized the 68Ga-labeled photoactivatable ligand, which contained an aryl azide group ([68Ga]GaNODAGA-PEG3-ArN3) (Figure 21). The prepared radiotracer underwent a facile reaction with an amino group of the antibodies, including GMP-grade HerceptinTM upon light irradiation (λmax ~ 365 nm) within 5 min. The radiolabeled product was also utilized for the specific tumor imaging in SK-OV-3 tumor xenograft. A similar method was also applied to the radiosynthesis of 89Zr-labeled antibody by using a desferrioxamine B conjugated aryl azide group [138]. As the radiolabeling of trastuzumab has been carried out over a short time with high efficiency and purity, this approach will be applicable for the efficient radiolabeling of various biologically active molecules.

5. Conclusions and Future Perspectives

In this review, we focused on recent examples that highlight the application of bioorthogonal click chemistries for the preparation of radiolabeled products. For many years, rapid and selective conjugation reactions including SPAAC and IEDDA have been successfully employed for the straightforward, site-specific, and efficient labeling of various radioisotopes to the small molecules, biomacromolecules, functional nanomaterials, and living cells. In addition, electrophilic aromatic prosthetic groups which display fast reaction kinetics and high selectivity towards the amine or thiol groups could also be the preferred methods for the radiolabeling procedure, because these reactions do not need the introduction of an artificial functional group to the target the biomolecule. Regarding future perspectives, it is anticipated that the relevance of bioorthogonal strategies will continue to be applicable beyond the rapid labeling of a radioisotope to a target molecule of interest. For example, the development of in vivo ligation based on IEDDA enabled the investigation of various approaches for specific tumor imaging with decreased non-specific accumulation of radioligand in normal tissues. Particularly, the introduction of clearing agents before administration of radiotracers demonstrated improved tumor-to-background ratio with enhanced uptake values in the target sites. Although some recent advancements can provide potent tools in nuclear medicine, several key challenges need to be addressed for their further development. The functional groups and resulting adducts obtained by bioorthogonal ligations are normally hydrophobic, which may result in non-specific uptake and retarded excretion kinetics in a living subject. Moreover, conjugation of a relatively large functional group to the small molecule probes or short peptides will affect their pharmacokinetic profiles and induce undesired accumulation or retention of radioactive signals in healthy tissues. For instance, we have synthesized 18F-labeled dimeric cRGD peptide by using the condensation reaction between CBT and N-terminal cysteine (Table 4, entry 1). This method provided an efficient radiochemical result. However, the hydrophobic adduct produced by the radiolabeling reaction afforded high uptake values in normal organs, including in liver and kidneys compared with [18F]FPPRGD2, which is a clinically approved radiopharmaceutical [122]. Such undesired biodistribution results would hamper further investigation of new radiotracers. Therefore, development of fine-tuned functional group pairs, which are smaller and less lipophilic, and at the same time possess high reactivity and selectivity must be investigated to maximize specific targeting ability of the radioligand with minimal background signal. Consequently, bioorthogonal click reactions have exhibited enormous potential for development of radiopharmaceuticals and applications in the field of nuclear medicine. The optimization of these ligation methods will enable the exploration of advanced theranostic strategies as well as the investigation of sophisticated biological phenomena. We expect that these tools will continue to be used as a key technology for the development of various radiolabeled molecules and radiopharmaceuticals, which can offer benefits across preclinical studies and ultimately in clinical applications in the future.

Author Contributions

S.M., and S.-J.Y. wrote the review paper. J.J. conceived the main concept and edited the review paper.

Funding

This work was supported by the research grant from the National Research Foundation of Korea (grant number: 2017M2A2A6A01070858).

Acknowledgments

We would like to thank Lubna Ghani for formatting the reference list.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

CBT2-cyanobenzothiazole
CuAACcopper(I)-catalyzed azide-alkyne [3+2] cycloaddition reaction
DBCOdibenzocyclooctyne
DFOdeferoxamine
DMFN,N-dimethylformamide
DMSOdimethyl sulfoxide
DOTA1,4,7,10-tetraazacyclododecane-tetraacetic acid
DTPAdiethylenetriaminepentaacetic acid
FHferaheme
HPLChigh-performance liquid chromatography
HAShuman serum albumin
ICAM-1intercellular adhesion molecule
IEDDAinverse-electron-demand Diels–Alder reaction
MSTP(4-(5-methane-sulfonyl-[1,2,3,4]tetrazole-1-yl)-phenol)
NOTA1,4,7-triazacyclononane-1,4,7-triacetic acid
ODIBOoxa-dibenzocyclooctyne
PBSphosphate-buffered saline
PECAM-1platelet-endothelial cell adhesion molecule
PETpositron emission tomography
PICpolyisocyanopeptide
RCYradiochemical yield
SPAACstrain-promoted azide-alkyne cycloaddition reaction
SPECTsingle-photon emission computed tomography
TCOtrans-cyclooctene

References

  1. Schulze, B.; Schubert, U.S. Beyond click chemistry–supramolecular interactions of 1,2,3-triazoles. Chem. Soc. Rev. 2014, 43, 2522–2571. [Google Scholar] [CrossRef] [PubMed]
  2. Jewett, J.C.; Bertozzi, C.R. Synthesis of a fluorogenic cyclooctyne activated by Cu-free click chemistry. Org. Lett. 2011, 13, 5937–5939. [Google Scholar] [CrossRef] [PubMed]
  3. Xu, L.; Li, Y.; Li, Y. Application of “click” chemistry to the construction of supramolecular functional systems. Chem. Asian J. 2014, 3, 582–602. [Google Scholar] [CrossRef]
  4. McKay, C.S.; Finn, M.G. Click chemistry in complex mixtures: Bioorthogonal bioconjugation. Chem. Biol. 2014, 21, 1075–1101. [Google Scholar] [CrossRef] [PubMed]
  5. Hua, Y.; Flood, A.H. Click chemistry generates privileged CH hydrogen-bonding triazoles: The latest addition to anion supramolecular chemistry. Chem. Soc. Rev. 2010, 39, 1262–1271. [Google Scholar] [CrossRef] [PubMed]
  6. Aucagne, V.; Leigh, D.A. Chemoselective formation of successive triazole linkages in One Pot: “Click− Click” chemistry. Org. Lett. 2006, 8, 4505–4507. [Google Scholar] [CrossRef]
  7. Ganesh, V.; Sudhir, V.S.; Kundu, T.; Chandrasekaran, S. 10 Years of Click Chemistry: Synthesis and Applications of Ferrocene-Derived Triazoles. Chem. Asian J. 2011, 6, 2670–2694. [Google Scholar] [CrossRef] [Green Version]
  8. Becer, C.R.; Hoogenboom, R.; Schubert, U.S. Click chemistry beyond metal-catalyzed cycloaddition. Angew. Chem. Int. Ed. 2009, 48, 4900–4908. [Google Scholar] [CrossRef]
  9. Agalave, S.G.; Maujan, S.R.; Pore, V.S. Click chemistry: 1, 2, 3-triazoles as pharmacophores. Chem. Asian J. 2011, 6, 2696–2718. [Google Scholar] [CrossRef]
  10. Li, L.; Zhang, Z. Development and Applications of the Copper-Catalyzed Azide-Alkyne Cycloaddition (CuAAC) as a Bioorthogonal Reaction. Molecules 2016, 21, 1393. [Google Scholar] [CrossRef]
  11. Kappe, C.O.; Van der Eycken, E. Click chemistry under non-classical reaction conditions. Chem. Soc. Rev. 2010, 39, 1280–1290. [Google Scholar] [CrossRef] [PubMed]
  12. Mamidyala, S.K.; Finn, M.G. In situ click chemistry: Probing the binding landscapes of biological molecules. Chem. Soc. Rev. 2010, 39, 1252–1261. [Google Scholar] [CrossRef] [PubMed]
  13. Van Steenis, D.J.V.; David, O.R.; van Strijdonck, G.P.; van Maarseveen, J.H.; Reek, J.N. Click-chemistry as an efficient synthetic tool for the preparation of novel conjugated polymers. Chem. Commun. 2005, 34, 4333–4335. [Google Scholar] [CrossRef] [PubMed]
  14. Wong, C.H.; Zimmerman, S.C. Orthogonality in organic, polymer, and supramolecular chemistry: From Merrifield to click chemistry. Chem. Commun. 2013, 49, 1679–1695. [Google Scholar] [CrossRef]
  15. Van Dijk, M.; Rijkers, D.T.; Liskamp, R.M.; van Nostrum, C.F.; Hennink, W.E. Synthesis and applications of biomedical and pharmaceutical polymers via click chemistry methodologies. Bioconjugate Chem. 2009, 20, 2001–2016. [Google Scholar] [CrossRef] [PubMed]
  16. Moses, J.E.; Moorhouse, A.D. The growing applications of click chemistry. Chem. Soc. Rev. 2007, 36, 1249–1262. [Google Scholar] [CrossRef]
  17. Sanchez-Sanchez, A.; Pérez-Baena, I.; Pomposo, J. Advances in click chemistry for single-chain nanoparticle construction. Molecules 2013, 18, 3339–3355. [Google Scholar] [CrossRef] [PubMed]
  18. Lummerstorfer, T.; Hoffmann, H. Click chemistry on surfaces: 1, 3-dipolar cycloaddition reactions of azide-terminated monolayers on silica. J. Phys. Chem. B 2004, 108, 3963–3966. [Google Scholar] [CrossRef]
  19. Devadoss, A.; Chidsey, C.E. Azide-modified graphitic surfaces for covalent attachment of alkyne-terminated molecules by “click” chemistry. J. Am. Chem. Soc. 2007, 129, 5370–5371. [Google Scholar] [CrossRef]
  20. Schlossbauer, A.; Schaffert, D.; Kecht, J.; Wagner, E.; Bein, T. Click chemistry for high-density biofunctionalization of mesoporous silica. J. Am. Chem. Soc. 2008, 130, 12558–12559. [Google Scholar] [CrossRef]
  21. Li, N.; Binder, W.H. Click-chemistry for nanoparticle-modification. J. Mater. Chem. 2011, 21, 16717–16734. [Google Scholar] [CrossRef]
  22. Xi, W.; Scott, T.F.; Kloxin, C.J.; Bowman, C.N. Click chemistry in materials science. Adv. Funct. Mater. 2014, 24, 2572–2590. [Google Scholar] [CrossRef]
  23. Tasdelen, M.A. Diels–Alder “click” reactions: Recent applications in polymer and material science. Polym. Chem. 2011, 2, 2133–2145. [Google Scholar] [CrossRef]
  24. Nandivada, H.; Jiang, X.; Lahann, J. Click chemistry: Versatility and control in the hands of materials scientists. Adv. Mater. 2007, 19, 2197–2208. [Google Scholar] [CrossRef]
  25. Thirumurugan, P.; Matosiuk, D.; Jozwiak, K. Click chemistry for drug development and diverse chemical–biology applications. Chem. Rev. 2013, 113, 4905–4979. [Google Scholar] [CrossRef]
  26. Kolb, H.C.; Sharpless, K.B. The growing impact of click chemistry on drug discovery. Drug Discov. Today 2003, 8, 1128–1137. [Google Scholar] [CrossRef]
  27. Whiting, M.; Muldoon, J.; Lin, Y.C.; Silverman, S.M.; Lindstrom, W.; Olson, A.J.; Kolb, H.C.; Finn, M.G.; Sharpless, K.B.; Elder, J.H.; et al. Inhibitors of HIV-1 protease by using in situ click chemistry. Angew. Chem. Int. Ed. 2006, 45, 1435–1439. [Google Scholar] [CrossRef]
  28. Pagliai, F.; Pirali, T.; Del Grosso, E.; Di Brisco, R.; Tron, G.C.; Sorba, G.; Genazzani, A.A. Rapid synthesis of triazole-modified resveratrol analogues via click chemistry. J. Med. Chem. 2006, 49, 467–470. [Google Scholar] [CrossRef]
  29. Kamal, A.; Shankaraiah, N.; Reddy, C.R.; Prabhakar, S.; Markandeya, N.; Srivastava, H.K.; Sastry, G.N. Synthesis of bis-1,2,3-triazolo-bridged unsymmetrical pyrrolobenzodiazepine trimers via ‘click’ chemistry and their DNA-binding studies. Tetrahedron 2010, 66, 5498–5506. [Google Scholar] [CrossRef]
  30. De Geest, B.G.; Van Camp, W.; Du Prez, F.E.; De Smedt, S.C.; Demeester, J.; Hennink, W.E. Degradable multilayer films and hollow capsules via a ‘click’ strategy. Macromol. Rapid Commun. 2008, 29, 1111–1118. [Google Scholar] [CrossRef]
  31. Xu, X.D.; Chen, C.S.; Lu, B.; Wang, Z.C.; Cheng, S.X.; Zhang, X.Z.; Zhuo, R.X. Modular synthesis of thermosensitive P (NIPAAm-co-HEMA)/β-CD based hydrogels via click chemistry. Macromol. Rapid Commun. 2009, 30, 157–164. [Google Scholar] [CrossRef] [PubMed]
  32. Moorhouse, A.D.; Santos, A.M.; Gunaratnam, M.; Moore, M.; Neidle, S.; Moses, J.E. Stabilization of G-quadruplex DNA by highly selective ligands via click chemistry. J. Am. Chem. Soc. 2006, 128, 15972–15973. [Google Scholar] [CrossRef] [PubMed]
  33. De Geest, B.G.; Van Camp, W.; Du Prez, F.E.; De Smedt, S.C.; Demeester, J.; Hennink, W.E. Biodegradable microcapsules designed via ‘click’ chemistry. Chem. Commun. 2008, 190–192. [Google Scholar] [CrossRef] [PubMed]
  34. Lallana, E.; Fernandez-Megia, E.; Riguera, R. Surpassing the use of copper in the click functionalization of polymeric nanostructures: A strain-promoted approach. J. Am. Chem. Soc. 2009, 131, 5748–5750. [Google Scholar] [CrossRef]
  35. Schieferstein, H.; Betzel, T.; Fischer, C.R.; Ross, T.L. 18F-click labeling and preclinical evaluation of a new 18 F-folate for PET imaging. EJNMMI Res. 2013, 3, 68. [Google Scholar] [CrossRef] [PubMed]
  36. Notni, J.; Šimeček, J.; Hermann, P.; Wester, H.J. TRAP, a Powerful and Versatile Framework for Gallium-68 Radiopharmaceuticals. Chem. Asian J. 2011, 17, 14718–14722. [Google Scholar] [CrossRef]
  37. Pretze, M.; Mamat, C. Automated preparation of [18F]AFP and [18F]BFP: Two novel bifunctional 18F-labeling building blocks for Huisgen-click. J. Fluorine Chem. 2013, 150, 25–35. [Google Scholar] [CrossRef]
  38. Roberts, M.P.; Pham, T.Q.; Doan, J.; Jiang, C.D.; Hambley, T.W.; Greguric, I.; Fraser, B.H. Radiosynthesis and ‘click’ conjugation of ethynyl-4-[18F]fluorobenzene - an improved [18F]synthon for indirect radiolabeling. J. Label. Compd. Radiopharm. 2015, 58, 473–478. [Google Scholar] [CrossRef]
  39. Mindt, T.L.; Müller, C.; Melis, M.; de Jong, M.; Schibli, R. “Click-to-chelate”: In vitro and in vivo comparison of a 99mTc(CO)3-labeled N(τ)-histidine folate derivative with its isostructural, clicked 1,2,3-triazole analogue. Bioconjugate Chem. 2008, 19, 1689–1695. [Google Scholar] [CrossRef]
  40. Zeng, D.; Lee, N.S.; Liu, Y.; Zhou, D.; Dence, C.S.; Wooley, K.L.; Katzenellenbogen, J.A.; Welch, M.J. 64Cu Core-labeled nanoparticles with high specific activity via metal-free click chemistry. ACS Nano 2012, 6, 5209–5219. [Google Scholar] [CrossRef]
  41. Liu, Z.; Li, Y.; Lozada, J.; Schaffer, P.; Adam, M.J.; Ruth, T.J.; Perrin, D.M. Stoichiometric Leverage: Rapid 18F-Aryltrifluoroborate Radiosynthesis at High Specific Activity for Click Conjugation. Angew. Chem. Int. Ed. 2013, 52, 2303–2307. [Google Scholar] [CrossRef] [PubMed]
  42. Nair, D.P.; Podgorski, M.; Chatani, S.; Gong, T.; Xi, W.; Fenoli, C.R.; Bowman, C.N. The thiol-Michael addition click reaction: A powerful and widely used tool in materials chemistry. Chem. Mater. 2013, 26, 724–744. [Google Scholar] [CrossRef]
  43. Sweeney, J.B. Aziridines: Epoxides’ ugly cousins? Chem. Soc. Rev. 2002, 31, 247–258. [Google Scholar] [CrossRef] [PubMed]
  44. Crisalli, P.; Kool, E.T. Water-soluble organocatalysts for hydrazone and oxime formation. J. Org. Chem. 2013, 78, 1184–1189. [Google Scholar] [CrossRef]
  45. Wang, Q.; Chan, T.R.; Hilgraf, R.; Fokin, V.V.; Sharpless, K.B.; Finn, M.G. Bioconjugation by copper (I)-catalyzed azide-alkyne [3+2] cycloaddition. J. Am. Chem. Soc. 2003, 125, 3192–3193. [Google Scholar] [CrossRef] [PubMed]
  46. Dai, L.; Xue, Y.; Qu, L.; Choi, H.J.; Baek, J.B. Metal-free catalysts for oxygen reduction reaction. Chem. Rev. 2015, 115, 4823–4892. [Google Scholar] [CrossRef]
  47. Jewett, J.C.; Bertozzi, C.R. Cu-free click cycloaddition reactions in chemical biology. Chem. Soc. Rev. 2010, 39, 1272–1279. [Google Scholar] [CrossRef]
  48. Juhl, K.; Jørgensen, K.A. The First Organocatalytic Enantioselective Inverse-Electron-Demand Hetero-Diels–Alder Reaction. Angew. Chem. Int. Ed. 2003, 42, 1498–1501. [Google Scholar] [CrossRef]
  49. Li, Z.; Conti, P.S. Radiopharmaceutical chemistry for positron emission tomography. Adv. Drug Deliv. Rev. 2010, 62, 1031–1051. [Google Scholar] [CrossRef]
  50. Walsh, J.C.; Kolb, H.C. Applications of click chemistry in radiopharmaceutical development. CHIMIA Int. J. Chem. 2010, 64, 29–33. [Google Scholar] [CrossRef]
  51. Šečkutė, J.; Devaraj, N.K. Expanding room for tetrazine ligations in the in vivo chemistry toolbox. Curr. Opin. Chem. Biol. 2013, 17, 761–767. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Agard, N.J.; Prescher, J.A.; Bertozzi, C.R. A strain-promoted [3+2] azide−alkyne cycloaddition for covalent modification of biomolecules in living systems. J. Am. Chem. Soc. 2004, 126, 15046–15047. [Google Scholar] [CrossRef] [PubMed]
  53. Mbua, N.E.; Guo, J.; Wolfert, M.A.; Steet, R.; Boons, G.J. Strain-promoted alkyne–azide cycloadditions (SPAAC) reveal new features of glycoconjugate biosynthesis. ChemBioChem 2011, 12, 1912–1921. [Google Scholar] [CrossRef] [PubMed]
  54. Campbell-Verduyn, L.S.; Mirfeizi, L.; Schoonen, A.K.; Dierckx, R.A.; Elsinga, P.H.; Feringa, B.L. Strain-promoted copper-free “click” chemistry for 18F radiolabeling of bombesin. Angew. Chem. Int. Ed. 2011, 50, 11117–11120. [Google Scholar] [CrossRef] [PubMed]
  55. Bouvet, V.; Wuest, M.; Wuest, F. Copper-free click chemistry with the short-lived positron emitter fluorine-18. Org. Biomol. Chem. 2011, 9, 7393–7399. [Google Scholar] [CrossRef]
  56. Carpenter, R.D.; Hausner, S.H.; Sutcliffe, J.L. Copper-free click for PET: Rapid 1,3-dipolar cycloadditions with a fluorine-18 cyclooctyne. ACS Med. Chem. Lett. 2011, 2, 885–889. [Google Scholar] [CrossRef] [PubMed]
  57. Hausner, S.H.; Carpenter, R.D.; Bauer, N.; Sutcliffe, J.L. Evaluation of an integrin αvβ6-specific peptide labeled with [18F] fluorine by copper-free, strain-promoted click chemistry. Nucl. Med. Biol. 2013, 40, 233–239. [Google Scholar] [CrossRef]
  58. Arumugam, S.; Chin, J.; Schirrmacher, R.; Popik, V.V.; Kostikov, A.P. [18F]Azadibenzocyclooctyne ([18F]ADIBO): A biocompatible radioactive labeling synthon for peptides using catalyst free [3+2] cycloaddition. Bioorg. Med. Chem. Lett. 2011, 21, 6987–6991. [Google Scholar] [CrossRef]
  59. Kettenbach, K.; Ross, T.L. A 18F-labeled dibenzocyclooctyne (DBCO) derivative for copper-free click labeling of biomolecules. Med. Chem. Comm. 2016, 7, 654–657. [Google Scholar] [CrossRef]
  60. Roche, M.; Specklin, S.; Richard, M.; Hinnen, F.; Génermont, K.; Kuhnast, B. [18F]FPyZIDE: A versatile prosthetic reagent for the fluorine-18 radiolabeling of biologics via copper-catalyzed or strain-promoted alkyne-azide cycloadditions. J. Label. Compd. Radiopharm. 2019, 62, 95–108. [Google Scholar] [CrossRef]
  61. Evans, H.L.; Carroll, L.; Aboagye, E.O.; Spivey, A.C. Bioorthogonal chemistry for 68Ga radiolabelling of DOTA-containing compounds. J. Label. Compd. Radiopharm. 2014, 57, 291–297. [Google Scholar] [CrossRef] [PubMed]
  62. Ghosh, S.C.; Hernandez-Vargas, S.; Rodriguez, M.; Kossatz, S.; Voss, J.; Carmon, K.S.; Reiner, T.; Schonbrunn, A.; Azhdarinia, A. Synthesis of a fluorescently labeled 68Ga-DOTA-TOC analog for somatostatin receptor targeting. ACS Med. Chem. Let. 2017, 8, 720–725. [Google Scholar] [CrossRef] [PubMed]
  63. Gordon, C.G.; Mackey, J.L.; Jewett, J.C.; Sletten, E.M.; Houk, K.N.; Bertozzi, C.R. Reactivity of Biarylazacyclooctynones in Copper-Free Click Chemistry. J. Am. Chem. Soc. 2012, 134, 9199–9208. [Google Scholar] [CrossRef]
  64. McNitt, C.D.; Popik, V.V. Photochemical generation of oxa-dibenzocyclooctyne (ODIBO) for metal-free click ligations. Org. Biomol. Chem. 2012, 10, 8200–8202. [Google Scholar] [CrossRef] [PubMed]
  65. Boudjemeline, M.; McNitt, C.D.; Singleton, T.A.; Popik, V.V.; Kostikov, A.P. [18F]ODIBO: A prosthetic group for bioorthogonal radiolabeling of macromolecules via strain-promoted alkyne–azide cycloaddition. Org. Biomol. Chem. 2018, 16, 363–366. [Google Scholar] [CrossRef]
  66. Sachin, K.; Jadhav, V.H.; Kim, E.M.; Kim, H.L.; Lee, S.B.; Jeong, H.J.; Lim, S.T.; Sohn, M.H.; Kim, D.W. F-18 Labeling protocol of peptides based on chemically orthogonal strain-promoted cycloaddition under physiologically friendly reaction condition. Bioconjugate Chem. 2012, 23, 1680–1686. [Google Scholar] [CrossRef] [PubMed]
  67. Kim, H.L.; Sachin, K.; Jeong, H.J.; Choi, W.; Lee, H.S.; Kim, D.W. F-18 labeled RGD probes based on bioorthogonal strain-promoted click reaction for PET imaging. ACS Med. Chem. Lett. 2015, 6, 402–407. [Google Scholar] [CrossRef]
  68. Jeon, J.; Kang, J.A.; Shim, H.E.; Nam, Y.R.; Yoon, S.; Kim, H.R.; Lee, D.E.; Park, S.H. Efficient method for iodine radioisotope labeling of cyclooctyne-containing molecules using strain-promoted copper-free click reaction. Bioorg. Med. Chem. 2015, 23, 3303–3308. [Google Scholar] [CrossRef]
  69. Choi, M.H.; Shim, H.E.; Nam, Y.R.; Kim, H.R.; Kang, J.A.; Lee, D.E.; Park, S.H.; Choi, D.S.; Jang, B.S.; Jeon, J. Synthesis and evaluation of an 125I-labeled azide prosthetic group for efficient and bioorthogonal radiolabeling of cyclooctyne-group containing molecules using copper-free click reaction. Bioorg. Med. Chem. Lett. 2016, 26, 875–878. [Google Scholar] [CrossRef]
  70. Zeng, D.; Ouyang, Q.; Cai, Z.; Xie, X.Q.; Anderson, C.J. New cross-bridged cyclam derivative CB-TE1K1P, an improved bifunctional chelator for copper radionuclides. Chem. Commun. 2014, 50, 43–45. [Google Scholar] [CrossRef]
  71. Yuan, H.; Wilks, M.Q.; El Fakhri, G.; Normandin, M.D.; Kaittanis, C.; Josephson, L. Heat-induced-radiolabeling and click chemistry: A powerful combination for generating multifunctional nanomaterials. PloS ONE 2017, 12, 0172722. [Google Scholar] [CrossRef] [PubMed]
  72. Öztürk Öncel, M.Ö.; Garipcan, B.; Inci, F. Biomedical Applications: Liposomes and Supported Lipid Bilayers for Diagnostics, Theranostics, Imaging, Vaccine Formulation, and Tissue Engineering. In Biomimetic Lipid Membranes: Fundamentals, Applications, and Commercialization; Kök, F.N., Arslan Yildiz, A., Inci, F., Eds.; Springer: Cham, Switzerland, 2019; pp. 193–212. [Google Scholar]
  73. Hood, E.D.; Greineder, C.F.; Shuvaeva, T.; Walsh, L.; Villa, C.H.; Muzykantov, V.R. Vascular targeting of radiolabeled liposomes with bio-orthogonally conjugated ligands: Single chain fragments provide higher specificity than antibodies. Bioconjugate Chem. 2018, 29, 3626–3637. [Google Scholar] [CrossRef] [PubMed]
  74. Op’t Veld, R.C.; Joosten, L.; van den Boomen, O.; Boerman, O.; Kouwer, P.H.; Middelkoop, E.; Rowan, A.; Jansen, J.A.; Walboomers, F.; Wagener, F. Monitoring 111In-labelled polyisocyanopeptide (PIC) hydrogel wound dressings in full-thickness wounds. Biomater. Sci. 2019, 7, 3041–3050. [Google Scholar] [CrossRef] [PubMed]
  75. Lodhi, N.A.; Park, J.Y.; Kim, K.; Kim, Y.J.; Shin, J.H.; Lee, Y.S.; Im, H.J.; Jeong, J.M.; Khalid, M.; Cheon, G.J.; et al. Development of 99mTc-Labeled Human Serum Albumin with Prolonged Circulation by Chelate-then-Click Approach: A Potential Blood Pool Imaging Agent. Mol. Pharm. 2019, 16, 1586–1595. [Google Scholar] [CrossRef] [PubMed]
  76. Knall, A.C.; Slugovc, C. Inverse electron demand Diels–Alder (iEDDA)-initiated conjugation: A (high) potential click chemistry scheme. Chem. Soc. Rev. 2013, 42, 5131–5142. [Google Scholar] [CrossRef]
  77. Liu, F.; Liang, Y.; Houk, K.N. Theoretical elucidation of the origins of substituent and strain effects on the rates of Diels–Alder reactions of 1,2,4,5-tetrazines. J. Am. Chem. Soc. 2014, 136, 11483–11493. [Google Scholar] [CrossRef] [PubMed]
  78. Selvaraj, R.; Fox, J.M. Trans-Cyclooctene - a stable, voracious dienophile for bioorthogonal labeling. Curr. Opin. Chem. Biol. 2013, 17, 753–760. [Google Scholar] [CrossRef]
  79. Versteegen, R.M.; Rossin, R.; Ten Hoeve, W.; Janssen, H.M.; Robillard, M.S. Click to release: Instantaneous doxorubicin elimination upon tetrazine ligation. Angew. Chem. Int. Ed. 2013, 52, 14112–14116. [Google Scholar] [CrossRef]
  80. Mushtaq, S.; Jeon, J. Synthesis of PET and SPECT radiotracers using inverse electron-demand Diels–Alder reaction. Appl. Chem. Eng. 2017, 28, 141–152. [Google Scholar]
  81. Oliveira, B.L.; Guo, Z.; Bernardes, G.J.L. Inverse electron demand Diels–Alder reactions in chemical biology. Chem. Soc. Rev. 2017, 46, 4895–4950. [Google Scholar] [CrossRef]
  82. Li, Z.; Cai, H.; Hassink, M.; Blackman, M.L.; Brown, R.C.; Conti, P.S.; Fox, J.M. Tetrazine-trans-cyclooctene ligation for the rapid construction of 18F labeled probes. Chem. Commun. 2010, 46, 8043–8045. [Google Scholar] [CrossRef] [PubMed]
  83. Selvaraj, R.; Liu, S.; Hassink, M.; Huang, C.W.; Yap, L.P.; Park, R.; Fox, J.M.; Li, Z.; Conti, P.S. Tetrazine-trans-cyclooctene ligation for the rapid construction of integrin αvβ3 targeted PET tracer based on a cyclic RGD peptide. Bioorg. Med. Chem. Lett. 2011, 21, 5011–5014. [Google Scholar] [CrossRef] [PubMed]
  84. Liu, S.; Hassink, M.; Selvaraj, R.; Yap, L.P.; Park, R.; Wang, H.; Chen, X.; Fox, J.M.; Li, Z.; Conti, P.S. Efficient 18F labeling of cysteine-containing peptides and proteins using tetrazine–trans-cyclooctene ligation. Mol. Imaging 2013, 12, 121–128. [Google Scholar] [CrossRef]
  85. Reiner, T.; Keliher, E.J.; Earley, S.; Marinelli, B.; Weissleder, R. Synthesis and in vivo imaging of a 18F-labeled PARP1 inhibitor using a chemically orthogonal scavenger-assisted high-performance method. Angew. Chem. Int. Ed. 2011, 50, 1922–1925. [Google Scholar] [CrossRef] [PubMed]
  86. Wu, Z.; Liu, S.; Hassink, M.; Nair, I.; Park, R.; Li, L.; Todorov, I.; Fox, J.M.; Li, Z.; Shively, J.E.; et al. Development and evaluation of 18F-TTCO-Cys40-Exendin-4: A PET Probe for Imaging Transplanted Islets. J. Nucl. Med. 2013, 54, 244–251. [Google Scholar] [CrossRef] [PubMed]
  87. Zhu, J.; Li, S.; Wängler, C.; Wängler, B.; Lennox, R.B.; Schirrmacher, R. Synthesis of 3-chloro-6-((4-(di-tert-butyl[18F]-fluorosilyl)-benzyl)oxy)-1,2,4,5-tetrazine ([18F]SiFA-OTz) for rapid tetrazine-based 18F-radiolabeling. Chem. Commun. 2015, 51, 12415–12418. [Google Scholar] [CrossRef] [PubMed]
  88. Knight, J.C.; Richter, S.; Wuest, M.; Way, J.D.; Wuest, F. Synthesis and evaluation of an 18F-labelled norbornene derivative for copper-free click chemistry reactions. Org. Biomol. Chem. 2013, 11, 3817–3825. [Google Scholar] [CrossRef]
  89. Herth, M.M.; Andersen, V.L.; Lehel, S.; Madsen, J.; Knudsen, G.M.; Kristensen, J.L. Development of a 11C-labeled tetrazine for rapid tetrazine–trans-cyclooctene ligation. Chem. Commun. 2013, 49, 3805–3807. [Google Scholar] [CrossRef]
  90. Nichols, B.; Qin, Z.; Yang, J.; Vera, D.R.; Devaraj, N.K. 68Ga chelating bioorthogonal tetrazine polymers for the multistep labeling of cancer biomarkers. Chem. Commun. 2014, 50, 5215–5217. [Google Scholar] [CrossRef]
  91. Albu, S.A.; Al-Karmi, S.A.; Vito, A.; Dzandzi, J.P.; Zlitni, A.; Beckford-Vera, D.; Blacker, M.; Janzen, N.; Patel, R.M.; Capretta, A.; et al. 125I-Tetrazines and inverse-electron-demand Diels–Alder chemistry: A convenient radioiodination strategy for biomolecule labeling, screening, and biodistribution studies. Bioconjugate Chem. 2016, 27, 207–216. [Google Scholar] [CrossRef]
  92. Choi, M.H.; Shim, H.E.; Yun, S.J.; Kim, H.R.; Mushtaq, S.; Lee, C.H.; Park, S.H.; Choi, D.S.; Lee, D.E.; Byun, E.B.; et al. Highly efficient method for 125I-radiolabeling of biomolecules using inverse-electron-demand Diels–Alder reaction. Bioorg. Med. Chem. 2016, 24, 2589–2594. [Google Scholar] [CrossRef] [PubMed]
  93. Genady, A.R.; Tan, J.; Mohamed, E.; Zlitni, A.; Janzen, N.; Valliant, J.F. Synthesis, characterization and radiolabeling of carborane-functionalized tetrazines for use in inverse electron demand Diels–Alder ligation reactions. J. Organomet. Chem. 2015, 798, 278–288. [Google Scholar] [CrossRef]
  94. Zeglis, B.M.; Mohindra, P.; Weissmann, G.I.; Divilov, V.; Hilderbrand, S.A.; Weissleder, R.; Lewis, J.S. Modular strategy for the construction of radiometalated antibodies for positron emission tomography based on inverse electron demand Diels–Alder click chemistry. Bioconjugate Chem. 2011, 22, 2048–2059. [Google Scholar] [CrossRef] [PubMed]
  95. Poty, S.; Membreno, R.; Glaser, J.M.; Ragupathi, A.; Scholz, W.W.; Zeglis, B.M.; Lewis, J.S. The inverse electron-demand Diels–Alder reaction as a new methodology for the synthesis of 225Ac-labelled radioimmunoconjugates. Chem. Commun. 2018, 54, 2599–2602. [Google Scholar] [CrossRef]
  96. Hernández-Gil, J.; Braga, M.; Harriss, B.I.; Carroll, L.S.; Leow, C.H.; Tang, M.X.; Aboagye, E.O.; Long, N.J. Development of 68Ga-labelled ultrasound microbubbles for whole-body PET imaging. Chem. Sci. 2019, 10, 5603–5615. [Google Scholar] [CrossRef] [PubMed]
  97. Altai, M.; Membreno, R.; Cook, B.; Tolmachev, V.; Zeglis, B.M. Pretargeted imaging and therapy. J. Nucl. Med. 2017, 58, 1553–1559. [Google Scholar] [CrossRef]
  98. Rossin, R.; Renart Verkerk, P.; Van Den Bosch, S.M.; Vulders, R.C.; Verel, I.; Lub, J.; Robillard, M.S. In vivo chemistry for pretargeted tumor imaging in live mice. Angew. Chem. Int. Ed. 2010, 49, 3375–3378. [Google Scholar] [CrossRef]
  99. Rossin, R.; van den Bosch, S.M.; ten Hoeve, W.; Carvelli, M.; Versteegen, R.M.; Lub, J.; Robillard, M.S. Highly Reactive trans-cyclooctene tags with improved stability for Diels–Alder chemistry in living systems. Bioconjugate Chem. 2013, 24, 1210–1217. [Google Scholar] [CrossRef]
  100. Rossin, R.; Läppchen, T.; van den Bosch, S.M.; Laforest, R.; Robillard, M.S. Diels–Alder reaction for tumor pretargeting: In vivo chemistry can boost tumor radiation dose compared with directly antibody. J. Nucl. Med. 2013, 54, 1989–1995. [Google Scholar] [CrossRef]
  101. Rossin, R.; van Duijnhoven, S.M.; Läppchen, T.; van den Bosch, S.M.; Robillard, M.S. Trans-Cyclooctene tag with improved properties for tumor pretargeting with the Diels–Alder reaction. Mol. Pharm. 2014, 11, 3090–3096. [Google Scholar] [CrossRef]
  102. Van Duijnhoven, S.M.; Rossin, R.; van den Bosch, S.M.; Wheatcroft, M.P.; Hudson, P.J.; Robillard, M.S. Diabody pretargeted with click chemistry in vivo. J. Nucl. Med. 2015, 56, 1422–1428. [Google Scholar] [CrossRef] [PubMed]
  103. Altai, M.; Perols, A.; Tsourma, M.; Mitran, B.; Honarvar, H.; Robillard, M.; Rossin, R.; ten Hoeve, W.; Lubberink, M.; Orlova, A.; et al. Feasibility of Affibody-Based Bioorthogonal Chemistry–Mediated Radionuclide Pretargeting. J. Nucl. Med. 2016, 57, 431–436. [Google Scholar] [CrossRef] [PubMed]
  104. Denk, C.; Svatunek, D.; Filip, T.; Wanek, T.; Lumpi, D.; Fröhlich, J.; Kuntner, C.; Mikula, H. Development of a 18F-Labeled Tetrazine with Favorable Pharmacokinetics for Bioorthogonal PET Imaging. Angew. Chem. Int. Ed. 2014, 53, 9655–9659. [Google Scholar] [CrossRef] [PubMed]
  105. Meyer, J.P.; Houghton, J.L.; Kozlowski, P.; Abdel-Atti, D.; Reiner, T.; Pillarsetty, N.V.K.; Scholz, W.W.; Zeglis, B.M.; Lewis, J.S. 18F-Based Pretargeted PET Imaging Based on Bioorthogonal Diels− Alder Click Chemistry. Bioconjugate Chem. 2016, 27, 298–301. [Google Scholar] [CrossRef]
  106. Keinänen, O.; Fung, K.; Pourat, J.; Jallinoja, V.; Vivier, D.; Pillarsetty, N.K.; Airaksinen, A.J.; Lewis, J.S.; Zeglis, B.M.; Sarparanta, M. Pretargeting of internalizing trastuzumab and cetuximab with a 18F-tetrazine tracer in xenograft models. EJNMMI Res. 2017, 7, 95. [Google Scholar] [CrossRef]
  107. Keinänen, O.; Mäkilä, E.M.; Lindgren, R.; Virtanen, H.; Liljenbäck, H.; Oikonen, V.; Sarparanta, M.; Molthoff, C.; Windhorst, A.D.; Roivainen, A.; et al. Pretargeted PET Imaging of trans-Cyclooctene-Modified Porous Silicon Nanoparticles. ACS Omega 2017, 2, 62–69. [Google Scholar]
  108. Billaud, E.M.F.; Shahbazali, E.; Ahamed, M.; Cleeren, F.; Noël, T.; Koole, M.; Verbruggen, A.; Hessel, V.; Bormans, G. Micro-flow photosynthesis of new dienophiles for inverse-electron-demand Diels–Alder reactions. Potential applications for pretargeted in vivo PET imaging. Chem. Sci. 2017, 8, 1251–1258. [Google Scholar]
  109. Billaud, E.M.; Belderbos, S.; Cleeren, F.; Maes, W.; Van de Wouwer, M.; Koole, M.; Verbruggen, A.; Himmelreich, U.; Geukens, N.; Bormans, G. Pretargeted PET imaging using a bioorthogonal 18F-Labeled trans-cyclooctene in an ovarian carcinoma model. Bioconjugate Chem. 2017, 28, 2915–2920. [Google Scholar] [CrossRef]
  110. Devaraj, N.K.; Thurber, G.M.; Keliher, E.J.; Marinelli, B.; Weissleder, R. Reactive polymer enables efficient in vivo bioorthogonal chemistry. Proc. Natl. Acad. Sci. USA 2012, 109, 4762–4767. [Google Scholar] [CrossRef] [Green Version]
  111. Denk, C.; Svatunek, D.; Mairinger, S.; Stanek, J.; Filip, T.; Matscheko, D.; Kuntner, C.; Wanek, T.; Mikula, H. Design, Synthesis, and Evaluation of a Low-Molecular-Weight 11C-Labeled Tetrazine for Pretargeted PET Imaging Applying Bioorthogonal in vivo Click Chemistry. Bioconjugate Chem. 2016, 27, 1707–1712. [Google Scholar] [CrossRef]
  112. Stéen, E.J.L.; Jørgensen, J.T.; Petersen, I.N.; Nørregaard, K.; Lehel, S.; Shalgunov, V.; Birke, A.; Edem, P.E.; L’Estrade, E.T.; Hansen, H.D.; et al. Improved radiosynthesis and preliminary in vivo evaluation of the 11C-labeled tetrazine [11C]AE-1 for pretargeted PET imaging. Bioorg. Med. Chem. Lett. 2019, 20, 986–990. [Google Scholar] [CrossRef] [PubMed]
  113. Zeglis, B.M.; Sevak, K.K.; Reiner, T.; Mohindra, P.; Carlin, S.D.; Zanzonico, P.; Weissleder, R.; Lewis, J.S. A pretargeted PET imaging strategy based on bioorthogonal Diels–Alder click chemistry. J. Nucl. Med. 2013, 54, 1389–1396. [Google Scholar] [CrossRef] [PubMed]
  114. Cook, B.E.; Adumeau, P.; Membreno, R.; Carnazza, K.E.; Brand, C.; Reiner, T.; Agnew, B.J.; Lewis, J.S.; Zeglis, B.M. Pretargeted PET imaging Using a Site-Specifically Labeled Immunoconjugate. Bioconjugate Chem. 2016, 27, 1789–1795. [Google Scholar] [CrossRef] [PubMed]
  115. Houghton, J.L.; Zeglis, B.M.; Abdel-Atti, D.; Sawada, R.; Scholz, W.W.; Lewis, J.S. Pretargeted Immuno-PET of Pancreatic Cancer: Overcoming Circulating Antigen and Internalized Antibody to Reduce Radiation Doses. J. Nucl. Med. 2016, 57, 453–459. [Google Scholar] [CrossRef] [PubMed]
  116. Adumeau, P.; Carnazza, K.E.; Brand, C.; Carlin, S.D.; Reiner, T.; Agnew, B.J.; Lewis, J.S.; Zeglis, B.M. A Pretargeted Approach for the Multimodal PET/NIRF Imaging of Colorectal Cancer. Theranostics 2016, 6, 2267–2277. [Google Scholar] [CrossRef] [PubMed]
  117. Evans, H.L.; Nguyen, Q.D.; Carroll, L.S.; Kaliszczak, M.; Twyman, F.J.; Spivey, A.C.; Aboagye, E.O. A bioorthogonal 68Ga-labelling strategy for rapid in vivo imaging. Chem. Commun. 2014, 50, 9557–9560. [Google Scholar] [CrossRef]
  118. Meyer, J.P.; Tully, K.M.; Jackson, J.; Dilling, T.R.; Reiner, T.; Lewis, J.S. Bioorthogonal Masking of Circulating Antibody–TCO Groups Using Tetrazine-Functionalized Dextran Polymers. Bioconjugate Chem. 2018, 29, 538–545. [Google Scholar] [CrossRef] [PubMed]
  119. Yazdani, A.; Bilton, H.; Vito, A.; Genady, A.R.; Rathmann, S.M.; Ahmad, Z.; Janzen, N.; Czorny, S.; Zeglis, B.M.; Francesconi, L.C.; et al. A bone-seeking trans-cyclooctene for pretargeting and bioorthogonal chemistry: A proof of concept study using 99mTc- and 177Lu-labeled tetrazines. J. Med. Chem. 2016, 59, 9381–9389. [Google Scholar] [CrossRef]
  120. García, M.F.; Gallazzi, F.; de Souza Junqueira, M.; Fernández, M.; Camacho, X.; da Silva Mororó, J.; Faria, D.; de Godoi Carneiro, C.; Couto, M.; Carrión, F.; et al. Synthesis of hydrophilic HYNIC-[1,2,4,5] tetrazine conjugates and their use in antibody pretargeting with 99mTc. Org. Biomol. Chem. 2018, 16, 5275–5285. [Google Scholar] [CrossRef] [PubMed]
  121. Shah, M.A.; Zhang, X.; Rossin, R.; Robillard, M.S.; Fisher, D.R.; Bueltmann, T.; Hoeben, F.J.; Quinn, T.P. Metal-free cycloaddition chemistry driven pretargeted radioimmunotherapy using α-particle radiation. Bioconjugate chem. 2017, 28, 3007–3015. [Google Scholar] [CrossRef]
  122. Jeon, J.; Shen, B.; Xiong, L.; Miao, Z.; Lee, K.H.; Rao, J.; Chin, F.T. Efficient method for site-specific 18F-labeling of biomolecules using the rapid condensation reaction between 2-cyanobenzothiazole and cysteine. Bioconjugate Chem. 2012, 23, 1902–1908. [Google Scholar] [CrossRef] [PubMed]
  123. Su, X.; Cheng, K.; Jeon, J.; Shen, B.; Venturin, G.T.; Hu, X.; Rao, J.; Chin, F.T.; Wu, H.; Cheng, Z. Comparison of Two Site-Specifically 18F-Labeled Affibodies for PET Imaging of EGFR Positive Tumors. Mol. Pharm. 2014, 11, 3947–3956. [Google Scholar] [CrossRef] [PubMed]
  124. Inkster, J.A.; Colin, D.J.; Seimbille, Y. A novel 2-cyanobenzothiazole-based 18F prosthetic group for conjugation to 1,2-aminothiol-bearing targeting vectors. Org. Biomol. Chem. 2015, 13, 3667–3676. [Google Scholar] [CrossRef] [PubMed]
  125. Chen, K.T.; Nguyen, K.; Ieritano, C.; Gao, F.; Seimbille, Y. A Flexible Synthesis of 68Ga-Labeled Carbonic Anhydrase IX (CAIX)-Targeted Molecules via CBT/1,2-Aminothiol Click Reaction. Molecules 2019, 24, 23. [Google Scholar] [CrossRef] [PubMed]
  126. Gao, F.; Ieritano, C.; Chen, K.-T.; Dias, G.M.; Rousseau, J.; Bénard, F.; Seimbille, Y. Two bifunctional desferrioxamine chelators for bioorthogonal labeling of biovectors with zirconium-89. Org. Biomol. Chem. 2018, 16, 5102–5106. [Google Scholar] [CrossRef]
  127. Mushtaq, S.; Choi, D.S.; Jeon, J. Radiosynthesis of 125I-labeled 2-cyanobenzothiazole: A new prosthetic group for efficient radioiodination reaction. J. Radiopharm. Mol. Probes 2017, 3, 44–51. [Google Scholar]
  128. Toda, N.; Asano, S.; Barbas, C.F., III. Rapid, Stable, Chemoselective Labeling of Thiols with Julia Kocieński like Reagents: A Serum-Stable Alternative to Maleimide-Based Protein Conjugation. Angew. Chem. Int. Ed. 2013, 52, 12592–12596. [Google Scholar] [CrossRef]
  129. Chiotellis, A.; Sladojevich, F.; Mu, L.; Herde, A.M.; Valverde, I.E.; Tolmachev, V.; Schibli, R.; Ametamey, S.M.; Mindt, T.L. Novel chemoselective 18F-radiolabeling of thiol-containing biomolecules under mild aqueous conditions. Chem. Commun. 2016, 52, 6083–6086. [Google Scholar] [CrossRef]
  130. Shim, H.E.; Mushtaq, S.; Song, L.; Lee, C.H.; Lee, H.; Jeon, J. Development of a new thiol-reactive prosthetic group for site-specific labeling of biomolecules with radioactive iodine. Bioorg. Med. Chem. Lett. 2018, 28, 2875–2878. [Google Scholar] [CrossRef]
  131. Mushtaq, S.; Nam, Y.R.; Kang, J.A.; Choi, D.S.; Park, S.H. Efficient and Site-Specific 125I Radioiodination of Bioactive Molecules Using Oxidative Condensation Reaction. ACS Omega 2018, 3, 6903–6911. [Google Scholar] [CrossRef]
  132. Al-Momani, E.; Israel, I.; Buck, A.K.; Samnick, S. Improved synthesis of [18F]FS-PTAD as a new tyrosine-specific prosthetic group for radiofluorination of biomolecules. Appl. Radiat. Isot. 2015, 104, 136–142. [Google Scholar] [CrossRef] [PubMed]
  133. Zlatopolskiy, B.D.; Kandler, R.; Mottaghy, F.M.; Neumaier, B. C-(4-[18F]fluorophenyl)-N-phenyl nitrone: A novel 18F-labeled building block for metal free [3+2] cycloaddition. Appl. Radiat. Isot. 2012, 70, 184–192. [Google Scholar] [CrossRef]
  134. Zlatopolskiy, B.D.; Kandler, R.; Kobus, D.; Mottaghy, F.M.; Neumaier, B. Beyond azide–alkyne click reaction: Easy access to 18F-labelled compounds via nitrile oxide cycloadditions. Chem. Commun. 2012, 48, 7134–7136. [Google Scholar] [CrossRef] [PubMed]
  135. Urkow, J.; Bergman, C.; Wuest, F. Sulfo-click chemistry with 18F-labeled thio acids. Chem. Commun. 2019, 55, 1310–1313. [Google Scholar] [CrossRef] [PubMed]
  136. Patra, M.; Eichenberger, L.S.; Fischer, G.; Holland, J.P. Photochemical Conjugation and One-Pot Radiolabelling of Antibodies for Immune-PET. Angew. Chem. Int. Ed. 2019, 58, 1928–1933. [Google Scholar] [CrossRef] [PubMed]
  137. Eichenberger, L.S.; Patra, M.; Holland, J.P. Photoactive chelates for radiolabelling proteins. Chem. Commun. 2019, 55, 2257–2260. [Google Scholar] [CrossRef] [PubMed]
  138. Patra, M.; Klingler, S.; Eichenberger, L.S.; Holland, J.P. Simultaneous Photoradiochemical Labeling of Antibodies for Immuno-Positron Emission Tomography. IScience 2019, 13, 416–431. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Selected bioorthogonal conjugation reactions. (1) Copper-catalyzed azide-alkyne cycloaddition reaction (CuAAC); (2) strain-promoted azide-alkyne cycloaddition reaction (SPAAC); (3) tetrazine and trans-alkene substrates for inverse electron-demand-Diels–Alder reaction (IEDDA); (4) condensation reaction between 2-cyanobenzothiazole (CBT) and 1,2-aminothiol (N-terminal cysteine).
Figure 1. Selected bioorthogonal conjugation reactions. (1) Copper-catalyzed azide-alkyne cycloaddition reaction (CuAAC); (2) strain-promoted azide-alkyne cycloaddition reaction (SPAAC); (3) tetrazine and trans-alkene substrates for inverse electron-demand-Diels–Alder reaction (IEDDA); (4) condensation reaction between 2-cyanobenzothiazole (CBT) and 1,2-aminothiol (N-terminal cysteine).
Molecules 24 03567 g001
Figure 2. 18F-radiolabeling of bombesin derivative using SPAAC: a) human plasma or dimethyl sulfoxide (DMSO), room temperature, 15 min. R = Pyr-Gln, Pyr = pyroglutamic acid, R1 = Leu-Gly-Asn-Gln-Trp-Ala-Val-Gly-His-Leu-Met-NH2, RCY = radiochemical yield.
Figure 2. 18F-radiolabeling of bombesin derivative using SPAAC: a) human plasma or dimethyl sulfoxide (DMSO), room temperature, 15 min. R = Pyr-Gln, Pyr = pyroglutamic acid, R1 = Leu-Gly-Asn-Gln-Trp-Ala-Val-Gly-His-Leu-Met-NH2, RCY = radiochemical yield.
Molecules 24 03567 g002
Figure 3. Radiolabeling of peptides or proteins using 18F-labeled ODIBO.
Figure 3. Radiolabeling of peptides or proteins using 18F-labeled ODIBO.
Molecules 24 03567 g003
Figure 4. 18F-radiolabeling of DBCO-modified cRGD dimer using 18F-labeled azide precursor and polystyrene-supported azide-modified resin for purification of unreacted substrate.
Figure 4. 18F-radiolabeling of DBCO-modified cRGD dimer using 18F-labeled azide precursor and polystyrene-supported azide-modified resin for purification of unreacted substrate.
Molecules 24 03567 g004
Figure 5. Radiolabeling of DBCO-conjugated cRGD peptide using 125I-labeled azide tracers.
Figure 5. Radiolabeling of DBCO-conjugated cRGD peptide using 125I-labeled azide tracers.
Molecules 24 03567 g005
Figure 6. Reaction of azide-conjugated Cetuximab antibody with DBCO-conjugated crossed bridged macrocyclic CB-TE1K1P chelator for 64Cu radiolabeling.
Figure 6. Reaction of azide-conjugated Cetuximab antibody with DBCO-conjugated crossed bridged macrocyclic CB-TE1K1P chelator for 64Cu radiolabeling.
Molecules 24 03567 g006
Figure 7. Preparation of 89Zr-labeled multifunctional nanoprobes using SPAAC ligation.
Figure 7. Preparation of 89Zr-labeled multifunctional nanoprobes using SPAAC ligation.
Molecules 24 03567 g007
Figure 8. SPAAC for 99mTc-based radiolabeling of human serum protein.
Figure 8. SPAAC for 99mTc-based radiolabeling of human serum protein.
Molecules 24 03567 g008
Figure 9. Radiolabeling of tetrazine conjugated cRGD peptide using 18F-labeled TCO; a) DMSO, 10 s, room temperature.
Figure 9. Radiolabeling of tetrazine conjugated cRGD peptide using 18F-labeled TCO; a) DMSO, 10 s, room temperature.
Molecules 24 03567 g009
Figure 10. IEDDA-mediated radiolabeling of trastuzumab.
Figure 10. IEDDA-mediated radiolabeling of trastuzumab.
Molecules 24 03567 g010
Figure 11. IEDDA-mediated synthesis of 225Ac-labeled monoclonal antibody.
Figure 11. IEDDA-mediated synthesis of 225Ac-labeled monoclonal antibody.
Molecules 24 03567 g011
Figure 12. Synthesis of 68Ga-labeled microbubble using IEDDA.
Figure 12. Synthesis of 68Ga-labeled microbubble using IEDDA.
Molecules 24 03567 g012
Figure 13. General strategy for pre-targeted imaging and therapy using IEDDA.
Figure 13. General strategy for pre-targeted imaging and therapy using IEDDA.
Molecules 24 03567 g013
Figure 14. A modified strategy using tetrazine-bearing clearing agent.
Figure 14. A modified strategy using tetrazine-bearing clearing agent.
Molecules 24 03567 g014
Figure 15. Radiosynthesis of 18F-labeled TCO.
Figure 15. Radiosynthesis of 18F-labeled TCO.
Molecules 24 03567 g015
Figure 16. Two-step pre-targeting strategy using 18F-labeled TCO.
Figure 16. Two-step pre-targeting strategy using 18F-labeled TCO.
Molecules 24 03567 g016
Figure 17. 11C-labeled tetrazine tracers for in vivo IEDDA reaction, (a) from ref [111], (b) from ref [112].
Figure 17. 11C-labeled tetrazine tracers for in vivo IEDDA reaction, (a) from ref [111], (b) from ref [112].
Molecules 24 03567 g017
Figure 18. Pre-targeted IEDDA ligation between TCO-bisphosphonate and radiolabeled tetrazine in bone tissue.
Figure 18. Pre-targeted IEDDA ligation between TCO-bisphosphonate and radiolabeled tetrazine in bone tissue.
Molecules 24 03567 g018
Figure 19. [3+2] cycloaddition reaction using 18F-labeled nitriloxides or N-hydroxyimidoyl chloride.
Figure 19. [3+2] cycloaddition reaction using 18F-labeled nitriloxides or N-hydroxyimidoyl chloride.
Molecules 24 03567 g019
Figure 20. Radiolabeling of biomolecules using sulfo-click chemistry.
Figure 20. Radiolabeling of biomolecules using sulfo-click chemistry.
Molecules 24 03567 g020
Figure 21. Radiolabeling of antibodies using photochemical conjugation reaction.
Figure 21. Radiolabeling of antibodies using photochemical conjugation reaction.
Molecules 24 03567 g021
Table 1. Examples of SPAAC in labeling reactions using short half-life radioisotopes.
Table 1. Examples of SPAAC in labeling reactions using short half-life radioisotopes.
Entry DBCO PrecursorAzide PrecursorProduct aRCY(%)Ref
1 Molecules 24 03567 i001 Molecules 24 03567 i002
R = 4-azidoaniline, 11-azido-3,6,9-trioxaundane-1-amine, 6-azido-6-deoxyglucose, 2-azido-deoxyglucose, azido-geldanamycin
Molecules 24 03567 i00369–98[55]
2 Molecules 24 03567 i004 Molecules 24 03567 i005
R = Ph, PEGylated acid
Molecules 24 03567 i00664–75[56]
3 Molecules 24 03567 i007 Molecules 24 03567 i008
R = A20FMDV2 peptide
Molecules 24 03567 i00912[57]
4 Molecules 24 03567 i010 Molecules 24 03567 i011
R = Tyr3-octreotate peptide
Molecules 24 03567 i01295[58]
5 Molecules 24 03567 i013 Molecules 24 03567 i014
R = cRGD peptide
Molecules 24 03567 i01593[59]
6 Molecules 24 03567 i016 Molecules 24 03567 i017 Molecules 24 03567 i018>95[60]
7 Molecules 24 03567 i019 Molecules 24 03567 i020
R = (PEG)3-DOTA-68Ga
Molecules 24 03567 i02194–100[61]
8 Molecules 24 03567 i022 Molecules 24 03567 i023
R = Tyr3-octreotate peptide
Molecules 24 03567 i02480[62]
a Products were obtained as isomeric mixtures.
Table 2. IEDDA-mediated in vitro radiolabeling.
Table 2. IEDDA-mediated in vitro radiolabeling.
EntryTetrazineDienophileProduct aRCY (%)Ref
1 Molecules 24 03567 i025 Molecules 24 03567 i026 Molecules 24 03567 i027>98[82]
2 Molecules 24 03567 i028
R = c(RGDyC) or VEGF protein
Molecules 24 03567 i029 Molecules 24 03567 i03095 c(RGDyC), 75 (VEGF)[84]
3 Molecules 24 03567 i031
R= AZD2281
Molecules 24 03567 i032 Molecules 24 03567 i03392[85]
4 Molecules 24 03567 i034
R= Cys40-exendin-4
Molecules 24 03567 i035 Molecules 24 03567 i036>80[86]
5 Molecules 24 03567 i037 Molecules 24 03567 i038 Molecules 24 03567 i039>99[87]
6 Molecules 24 03567 i040
R = Bombesin
Molecules 24 03567 i041 Molecules 24 03567 i04246[88]
7 Molecules 24 03567 i043 Molecules 24 03567 i044 Molecules 24 03567 i045>98[89]
8 Molecules 24 03567 i046
Tz-polymer
Molecules 24 03567 i047
anti-A33 antibody
Molecules 24 03567 i048-[90]
9 Molecules 24 03567 i049 Molecules 24 03567 i050
anti-VEGFR2 antibody
Molecules 24 03567 i05169[91]
10 Molecules 24 03567 i052 Molecules 24 03567 i053
R = cRGD peptide, HSA protein
Molecules 24 03567 i054>99 (cRGD), 93 (HSA)[92]
a Products were obtained as isomeric mixtures.
Table 3. IEDDA-based in vivo pre-targeted approach.
Table 3. IEDDA-based in vivo pre-targeted approach.
EntryBiomoleculeRadiotracerAnimal ModelRef
1CC49-TCO antibody111In-labeled tetrazineLS174T cells (Balb/C mouse)[98]
2CC49-TCO antibody111In-labeled tetrazineLS174T cells (Balb/C mouse)[99]
3CC49-TCO antibody177Lu-labeled tetrazineLS174T cells (Balb/C mouse)[100]
4CC49-TCO antibody177Lu-labeled tetrazineLS174T cells (Balb/C mouse)[101]
5AVP04-07-TCO diabody177Lu-labeled tetrazineLS174T cells (Balb/C mouse)[102]
6Z2395-TCO affibody111In-labeled tetrazine
177Lu-labeled tetrazine
SKOV-3 cells (Balb/C mouse)[103]
7PEGylated-TCO18F-labeled tetrazineHealthy Balb/C mouse[104]
85B1-TCO antibody18F-labeled tetrazineBxPC3 cells (athymic nude mice)[105]
9Cetuximab-TCO antibody
Trastuzumab-TCO antibody
18F-labeled tetrazineA431 cells (nu/nu mouse)
BT-474 cells (nu/nu mouse)
[106]
10Porous silicon-TCO nanoparticle18F-labeled tetrazineHealthy (Balb/C mouse)[107]
11PSMA antagonist-tetrazine conjugate18F-labeled TCOLNCaP cells (Balb/C mouse)[108]
12Trastuzumab-tetrazine antibody18F-labeled TCOSKOV-3 cells (Balb/C mouse)[109]
13A33-TCO antibody18F-labeled tetrazineLS174T cells (Balb/C mouse)
A431 cells (Balb/C mouse)
[110]
14Mesoporous silica-TCO nanoparticle11C-labeled tetrazineHealthy Balb/C mouse[111]
15Polyglutamic acid-TCO11C-labeled tetrazineCT26 cell (Balb/C mouse)[112]
16A33-TCO antibody64Cu-labeled tetrazineSW1222 cell (mouse)[113]
17HuA33-TCO antibody64Cu-labeled tetrazineSW1222 cell (mouse)[114]
185B1-TCO antibody64Cu-labeled tetrazineBxPC3 and Capan-2 cells (athymic nude mice)[115]
19HuA33-dye-800-TCO64Cu-labeled tetrazineSW1222 cell (mouse)[116]
20C225-TCO antibody68Ga-labeled tetrazineA431 cells (Balb/C mouse)[117]
21HuA33-TCO antibody68Ga-labeled tetrazineSW1222 cell (CrTac:NCr- Foxn1nu mouse)[118]
22Bisphosphonate-TCO conjugate177Lu-labeled tetrazine
99mTc-labeled tetrazine
Healthy Balb/C mouse[119]
23Bevacizumab-TCO antibody99mTc-labeled tetrazineB16-F10 cell (C57 Bl/6J mouse)[120]
24CC49-TCO antibody212Pb-labeled tetrazineLS174T cells (Balb/C mouse)[121]
Table 4. Aromatic prosthetic groups for radiolabeling reactions.
Table 4. Aromatic prosthetic groups for radiolabeling reactions.
EntryRadiotracerTarget MoleculeProductRCY (%)Ref
1 Molecules 24 03567 i055 Molecules 24 03567 i056
R = cRGD2 peptide, RLuc8 protein, ZEGFR:1907 affibody
Molecules 24 03567 i05780 (cRGD), 12 (RLuc8), 41 (ZEGFR:1907)[122,123]
2 Molecules 24 03567 i058 Molecules 24 03567 i059
R = cRGD peptide
Molecules 24 03567 i0607[124]
3 Molecules 24 03567 i061 Molecules 24 03567 i062 Molecules 24 03567 i06399[125]
4 Molecules 24 03567 i064 Molecules 24 03567 i065
R = cRGD peptide
Molecules 24 03567 i066>99[127]
5 Molecules 24 03567 i067 Molecules 24 03567 i068
R= BBN peptide, ZHER2:2395 affibody
Molecules 24 03567 i069>99 (BBN), 40 (ZHER2:2395)[129]
6 Molecules 24 03567 i070HSA protein, GCQRPPR peptide Molecules 24 03567 i071
R = HSA protein, GCQRPPR peptide
65 (HSA), 99 (GCQRPPR)[130]
7 Molecules 24 03567 i072 Molecules 24 03567 i073
R = cRGD peptide, HSA protein
Molecules 24 03567 i07499 (cRGD), 94 (HSA)[131]
8 Molecules 24 03567 i075 Molecules 24 03567 i076 Molecules 24 03567 i07745[132]

Share and Cite

MDPI and ACS Style

Mushtaq, S.; Yun, S.-J.; Jeon, J. Recent Advances in Bioorthogonal Click Chemistry for Efficient Synthesis of Radiotracers and Radiopharmaceuticals. Molecules 2019, 24, 3567. https://doi.org/10.3390/molecules24193567

AMA Style

Mushtaq S, Yun S-J, Jeon J. Recent Advances in Bioorthogonal Click Chemistry for Efficient Synthesis of Radiotracers and Radiopharmaceuticals. Molecules. 2019; 24(19):3567. https://doi.org/10.3390/molecules24193567

Chicago/Turabian Style

Mushtaq, Sajid, Seong-Jae Yun, and Jongho Jeon. 2019. "Recent Advances in Bioorthogonal Click Chemistry for Efficient Synthesis of Radiotracers and Radiopharmaceuticals" Molecules 24, no. 19: 3567. https://doi.org/10.3390/molecules24193567

Article Metrics

Back to TopTop