Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2023 Feb 2.
Published in final edited form as: J Am Chem Soc. 2022 Jan 24;144(4):1647–1662. doi: 10.1021/jacs.1c10390

Catalytic Activation of Bioorthogonal Chemistry with Light (CABL) Enables Rapid, Spatiotemporally-controlled Labeling and No-Wash, Subcellular 3D-Patterning in Live Cells using Long Wavelength Light

Andrew Jemas 1, Yixin Xie 1, Jessica E Pigga 1, Jeffrey L Caplan 2, Christopher W am Ende 3,*, Joseph M Fox 1,*
PMCID: PMC9364228  NIHMSID: NIHMS1819578  PMID: 35072462

Abstract

Described is the spatiotemporally controlled labeling and patterning of biomolecules in live cells through the catalytic activation of bioorthogonal chemistry with light, referred to as “CABL”. Here, an unreactive dihydrotetrazine (DHTz) is photocatalytically oxidized in the intracellular environment by ambient O2 to produce a tetrazine that immediately reacts with a trans-cyclooctene (TCO) dienophile. 6-(2-Pyridyl)-dihydrotetrazine-3-carboxamides were developed as stable, cell permeable DHTz reagents that upon oxidation produce the most reactive tetrazines ever used in live cells with Diels-Alder kinetics exceeding k2 106 M−1s−1. CABL photocatalysts are based on fluorescein or silarhodamine dyes with activation at 470 or 660 nm. Strategies for limiting extracellular production of singlet oxygen are described that increase the cytocompatibility of photocatalysis. The HaloTag self-labeling platform was used to introduce DHTz tags to proteins localized in the nucleus, mitochondria, actin or cytoplasm, and high-yielding subcellular activation and labeling with a TCO-fluorophore was demonstrated. CABL is light-dose dependent, and 2-photon excitation promotes CABL at the sub-organelle level to selectively pattern live cells under no-wash conditions. CABL was also applied to spatially resolved live-cell labeling of an endogenous protein target by using TIRF microscopy to selectively activate intracellular monoacylglycerol lipase tagged with DHTz-labeled small molecule covalent inhibitor. Beyond spatiotemporally controlled labeling, CABL also improves the efficiency of ‘ordinary’ tetrazine ligations by rescuing the reactivity of commonly used 3-aryl-6-methyltetrazine reporters that become partially reduced to DHTzs inside cells. The spatiotemporal control and fast rates of photoactivation and labeling of CABL should enable a range of biomolecular labeling applications in living systems.

Graphical Abstract

graphic file with name nihms-1819578-f0010.jpg

Introduction

The activation of chemical reactivity by light plays a central role in identifying and studying biological molecules in cellular context.13 Prominent among the various roles of photochemistry in chemical biology are tools for tracking the dynamics of biological molecules in living cells.4 For example, fluorescence recovery after photobleaching (FRAP) has long been used to study the movement of mobility of cellular molecules.5 FRAP and related techniques operate by photobleaching a subcellular region of interest and studying the migration and exchange of fluorescently labeled molecules into that region. As a complement to FRAP, photoactivatable proteins have emerged as tools for the direct study of subcellular molecules that become fluorescent in response to light.6 Here, recombinant proteins can be made fluorescent in real time in response to focused light. Related techniques based on photoconvertible7 and photoswitchable8 proteins offer additional options for the study of fusion proteins in living systems. As an alternative to techniques based on fluorescent proteins, small molecule probes have been advanced as alternatives to photoactivation in the cellular environment.916 Small molecule photoactivation has traditionally involved photochemical uncaging914 with more recent innovations using the Wolff rearrangement15 and photoprotonation.16 An advantage to small molecule fluorophores is that they are generally brighter and more stable than fluorescent proteins.17 Challenges for current small molecules systems include increasing photoactivation rate and/or extending to longer wavelengths to further augment cytocompatibility.

Bioorthogonal chemistry is a powerful method for selective and high-yielding covalent bond formation in living cells and organisms.18,19 Bioorthogonal chemistry is highly versatile and can be applied to essentially any biological molecule that can be adapted to incorporate small molecule chemical reporters, and as such has been used to label glycans,20 lipids,21 nucleic acids,22 and proteins23 engineered via enzymatic labeling,24 self-labeling25 or genetic code expansion.26 Chemical probes bearing bioorthogonal tags have been used to probe for targets in endogenous systems,2729 and strategies for activity-based protein profiling30 typically rely on enrichment via bioorthogonal chemistry. Because of this remarkable generality, new methodology for bioorthogonal labeling holds the potential for broad applicability in biology and medicine.

Photoactivatible bioorthogonal chemistry is an alternative method for labeling biological molecules in response to light.31,32 Initial examples included the photolysis of tetrazole and cyclopropenones to produce cycloaddition-reactive nitrile imines and cyclooctyne derivatives, respectively.3335 Further examples include photochemically inducible analogs of the Staudinger36 and CuAAC37 reactions as well as cycloadditions involving azirines,38 benzyne39, diarylsydnones,40,41 quinones,4245 o-napthaquinone methides46, o-quinodimethanes47,48 and trans-cycloheptene.49

Variations of tetrazine ligation— the bioorthogonal Diels-Alder reaction of tetrazines— have become increasingly important due to their exceptional kinetics with trans-cyclooctenes and other strained dienophiles.5053 Tetrazine ligation has become an important tool for tagging bioorthogonal reporters in live cells for fluorescent and super-resolution microscopy.5460 Photochemically inducible variants of tetrazine ligation have been described in recent years based on methods for uncaging cyclopropene61,62 and bicyclononyne63 dienophiles. Inducible versions of tetrazine ligation can also be achieved via the oxidation of dihydrotetrazine (DHTz) precursors64 with applications in electrochemically controlled bioconjugation at electrode surfaces,65 in batteries,66 and for colorimetric nitrous gas detection.67 In preprint, an o-nitrophenylphenyl protected dihydrotetrazine has been used with 405 nm light and without catalysis to uncage tetrazines that react with TCO with rates of 102 M−1s−1.68

Oxidation-induced bioorthogonal chemistry has emerged as a powerful method for on-demand bioorthogonal chemistry.69 Previously, our group described activation of tetrazine ligation using catalytic stimuli,70 where either visible light and a photocatalyst or the enzyme horseradish peroxidase was used to catalyze the oxidation of a DHTz to a tetrazine. (Fig 1). Oxygen serves as the terminal oxidant in this process, and several photocatalysts were found to be effective including methylene blue which catalyzes photooxidation upon excitation by 660 nm light. This initial system for photocatalytic oxidation using red light had found several in vitro applications,7073 but had limited use in live cell applications due to phototoxicity associated with the sensitization of singlet oxygen by methylene blue. More recently we have shown that silarhodamine (SiR) dyes, initially developed as fluorophores for biological imaging,7477 can be repurposed as photocatalysts for DHTz oxidation.78 The Janelia-SiR dyes77 were found to be especially effective even at low catalyst loadings. With SiR-photocatalysts, DHTz oxidation is more rapid than the competing sensitization of singlet oxygen, and therefore the photocatalytic activation of tetrazine ligation can be applied to protein modification while minimizing oxidative damage. In the presence of live human prostate cancer cells, SiR-red light photocatalysis was used to crosslink polymers to create hydrogels and enable their culture in 3D.78 This photoinducible hydrogel formation could also be carried out in vivo in live mice through subcutaneous injection of a solution containing SiR photocatalyst and a hydrogel precursor, followed by brief in vivo irradiation with 660 nm light to produce a stable hydrogel material.78 While these recent studies have demonstrated the ability to photocatalytically activate tetrazine ligation in the extracellular environment, the poor cellular permeability of the DHTz analogs used previously prevented us from exploring the potential for intracellular applications, and we had not explored spatiotemporal control of photoactivation at the cellular level.

Fig 1.

Fig 1.

Illustration of spatiotemporally controlled subcellular labeling via catalytic activation of bioorthogonal chemistry with light, or “CABL”. (A) Directing a DHTz-functionalized ligand to a subcellular target (illustrated here for actin) does not result in recruitment of a labeled TCO until (B) illumination in the presence of light results in the oxidation of DHTz to tetrazine, enabling rapid labeling via the fastest bioorthogonal reactions observed to date in live cells.

Described herein is the spatiotemporally controlled labeling and patterning of biomolecules inside of live cells through the catalytic activation of bioorthogonal chemistry with light, referred to as “CABL”. The new method is mechanistically different from traditional methods of photoactivation involving a chromophoric change upon irradiation. In CABL, a DHTz tag is directed to a subcellular target, where it is inactive in bioorthogonal conjugation chemistry. In the presence of a photocatalyst and light, the dormant DHTz is converted into a highly reactive tetrazine that can then undergo rapid bioorthogonal chemistry. (Fig 1). Key to the success of CABL in live cells is a new class of DHTz reagents that are cell permeable and stable in the intracellular environment. The fluorescent dyes, fluorescein (FL) and Janelia Si-Rhodamine, thienyl JF646 (SiR-tJF646)77 (Fig 2A), serve as photocatalysts in conjunction with brief irradiation at 470 nm and 660 nm, respectively, to produce the most reactive tetrazines ever used in live cells with Diels-Alder kinetics exceeding k2 106 M−1s−1. CABL can be used to photoactivate covalent attachment via bioorthogonal chemistry on a variety of subcellular protein targets to deliver bright, stable small molecule fluorophores to illuminated regions. In conjunction with 2-photon excitation it is possible to selectively pattern subcellular structures in 3D with sub-micron spatial resolution with live imaging under no-wash conditions. Photoactivation by CABL is not limited to overexpressed proteins, and was also applied to spatially resolved live-cell labeling of the protein target monoacylglycerol lipase (MAGL) at low, endogenous cellular concentration.

Fig 2.

Fig 2.

(A) Silarhodamine or fluorescein dyes catalyze the oxidation of dihydrotetrazine (DHTz) 1 by O2 to produce tetrazine 2. (B) Synthesis of amine-reactive DHTz 6 and amide analogs 1a-c. (C) Relative rate of tetrazine-TCO ligation for 2a vs other tetrazine derivatives in aqueous buffer. (D) Kinetics for reaction of 2a with TCO derivatives.

Results and Discussion

Synthesis and Evaluation of DHTz’s with Improved Permeability, Stability, and Reactivity

Essential to creating a system for photocatalysis in live cells was the development of 6-(2-pyridyl)-dihydrotetrazine-3-carboxamides 1 as a new class of compounds that display improved cell permeability and stability in the reduced DHTz state and enhanced reactivity in the oxidized tetrazine state (Fig 2A). The DHTz/Tz pair was designed to contain electron-withdrawing amide and 2-pyridyl substituents that play a dual role by stabilizing the dihydrotetrazine 1 toward background oxidation and augmenting the rate of the tetrazine 2 in inverse electron demand Diels-Alder cycloadditions.

Ethyl 6-pyridyl-DHTz-3-carboxylate 6 was synthesized through the sequence outlined in Fig 2b: hydrazide 3 and ethyl chlorooxoacetate were combined to give the unsymmetrical diacylhydrazine 4, which was reacted with PCl5 to give dichloride 5. Condensation with a stoichiometric amount of anhydrous hydrazine in ethanol gave 6 on gram scale.79,80 DHTz ester 6 is reactive enough to directly form amides upon reaction with amines due to the electron withdrawing nature of the DHTz. Butylamine can undergo direct amidation directly with 6 to give amide 1a. For the synthesis of secondary amide 1b and HaloTag-substrate 1c, activation with trimethylaluminum81 was required to promote the amidation reaction (Figure 2B).

UV-vis and NMR spectroscopies were used to monitor the stability of DHTz derivatives in metal-free PBS-buffer, prepared by simply passing PBS through chelex resin.82 DHTz 1a (40 μM) was shown by UV-vis to retain 96% and 94% of the DHTz oxidation state after 24 h and 49 h, respectively (Fig S1). Similarly, DHTz 1b (25 mM) was shown by 1H NMR to be ≥ 95% stable after 48 h in deuterated PBS (Fig S2). The stability of DHTz 1a was also studied in 10% human serum in PBS buffer; a UV-vis assay showed a 150 μM solution of 1a retained 92%, 81% and 71% DHTz oxidation state after 20 h, 2 d and 3.5 d respectively (Fig S3). For cellular experiments, FBS was replaced with serum-free Opti-MEM minimal media prior to DHTz introduction.

Photooxidation of DHTz 1a generates tetrazine 2a which participates in the fastest bioorthogonal reactions measured to date. Tetrazine 2a was also prepared independently via chemical oxidation of 1a using phenyliodonium diacetate.64 Using stopped-flow kinetics with UV-vis monitoring, the rate of reactivity of eq-5-hydroxy-trans-cyclooctene (5-OH-TCO) toward 2a was compared to known tetrazines as shown in Fig 2C. Chosen for comparison were monoaryltetrazine 783 and dipyridyltetrazine 8,64,84 which are commonly employed for their exceptional reactivity, and 3-methyl-6-aryltetrazine 9,83 which is commonly employed for its high stability. In PBS, the rate of 2a with 5-OH-TCO was measured to be 173,000 M−1s−1, which is 7-times and 10-times faster than analogous reactions of 5-OH-TCO toward 7 and 8, respectively. In MeOH, tetrazine 2a is more than 300 times more reactive than 9. The rate constants for the reaction of tetrazine 2a toward other trans-cyclooctene dienophiles are displayed in Figure 2D. The high reactivity of 2a is expected due to the activating nature of the pyridyl and amide substituents, which may increase reactivity through a combination of electron withdrawing and ground-state destabilizing effects.85 As expected, faster reactivity was observed with soluble a-TCO86 and o-TCO87 derivatives, which reacted with 2a in PBS with rates of 1.46 × 106 M−1s−1 and 1.13 × 106 M−1s−1. Under aqueous conditions, the reaction with the conformationally strained alkene s-TCO50,88 was too rapid for us to measure using stopped flow spectroscopy (>107 M−1s−1).

The hydrolytic stability of tetrazine 2a was studied by UV-vis spectroscopy by monitoring the disappearance of the signature tetrazine absorbance at 520 nm. Tetrazine 2a has a hydrolysis half-life of 45 min in PBS at r.t. (Figure S4). In 5 mM glutathione, the tetrazine absorbance at 520 nm decreases with a half-life of 70 seconds, but 65% of the absorbance of 2a is recovered upon photocatalytic oxidation, indicating 1a was formed as a reduction product of glutathione (Fig S7A). In HeLa cell lysate, diluted to 0.5 mg protein/mL, 2a decays with a half-life of 33 min with minor reduction to 1a, presumably due to glutathione in the lysate (Fig S7B). Any concerns about background reactivity are ameliorated by the ‘on demand’ nature of CABL as bioorthogonal reactions of 2a are much more rapid than background reactions. Thus, by activating tetrazine ligation in the presence of TCO dienophiles, CABL should not only provide spatiotemporal control but would also enable bioorthogonal reactions with exceptional rates.

Photocatalytic tetrazine ligation of 1a was initially demonstrated using BCN as the dienophile for in situ Diels-Alder reaction (Fig 3A). While BCN is less reactive than TCO in Diels-Alder chemistry, for small molecule characterization BCN conjugation has the advantage of producing a single aromatic product. As shown in Fig 3A, treatment of 1a with BCN under photocatalytic conditions using FL (2 μM) or SiR-tJF646 (2 μM) produced conjugate 10 in 95–98% yield after irradiation at 470 nm and 660 nm, respectively (Figs S89). UV-vis spectroscopy was used to monitor the photocatalyzed transformation of 1a (20 μM) to tetrazine 2a in the absence of BCN as a trapping agent.

Fig 3.

Fig 3

(A) DHTz oxidation/Diels-Alder reaction of 1a with BCN is photocatalyzed by either SiR or fluorescein gives conjugate 10. (B) Photocatalytic DHTz oxidation/Diels-Alder reaction of Halo-DHTz in the presence of methionine, a singlet oxygen scavenger, produces conjugate 11.

Fluorescein (FL) is a particularly efficient photocatalyst for DHTz oxidation.70 Fluorescein has been used as an in vitro photocatalyst89 and in cellular experiments to promote arylboronate ester oxidation via singlet oxygen sensitization.90 With 1 μM FL and 470 nm light, the formation of 2a is light dependent and yield of 2a reached a maximum of 79% yield after 40 s (Fig S5), and then decreased upon prolonged irradiation by a degradation process that is also photocatalyzed (Fig S6). Similarly, an 80% yield of 2a was observed after 400 s with photocatalysis by 1 μM SiR-tJF646 and 660 nm light (Fig S7).

Photocatalytic tetrazine ligation was also demonstrated with a protein-DHTz conjugate with reaction monitoring by mass spectrometry as shown in Figure 3B. The self-labeling HaloTag protein was labeled with the alkylchloride 1c to produce Halo-DHTz (33,928 Da). As shown in Figure 3B, conjugation with a-TCO-TAMRA took place efficiently upon irradiation at 470 nm in the presence of fluorescein as a photocatalyst to give conjugate 11 (34,578 Da) while exposure to photocatalyst and a-TCO-TAMRA in the absence of light showed no change in mass. Under photocatalytic conditions, both FL and SiR can also sensitize the formation of singlet oxygen at a rate competitive with the photooxidation of DHTz 1. Methionine (5 mM) was added as a singlet oxygen scavenger to limit protein oxidation and mimic the intracellular mechanisms for regulating singlet oxygen. As will be discussed below, strategies for limiting extracellular singlet oxygen production were developed and shown to maximize the efficacy of CABL in experiments in live cells.

Labeling Live Cells with CABL

We next sought to demonstrate that CABL could be carried out in a variety of intracellular targets in live cells. To this end, the HaloTag self-labeling platform was used to introduce DHTz tags to a variety of mammalian subcellular targets as well as to live bacteria. As displayed in Fig 4, HeLa cells were transfected with a green fluorescent protein HaloTag-GFP construct fused to a ‘protein of interest’ (POI) that controls subcellular localization, and then labeled by the small molecule DHTz-HaloTag. Subsequently, cells are treated with a photocatalyst and either a-TCO-TAMRA or o-TCO-TAMRA (Fig 4B)—fluorescent conjugates of TCO derivatives that were chosen for their hydrophilicity and improved permeability relative to traditional TCO derivatives.86 Upon irradiation, conjugation is expected only in those cells that express the HaloTag fusion protein, and analysis by SDS-PAGE with fluorescent imaging was used to provide a measure of conjugation efficiency. Unlike silarhodamine and fluorescein derivatives, TAMRA does not act as a photocatalyst for DHTz oxidation, and therefore functions only as a fluorescent reporter.

Fig 4.

Fig 4

(A) Workflow for introducing DHTz labels to Halotag fusion proteins in HeLa cells or E. coli with subsequent fluorescent labeling via CABL. (B) Structure of TAMRA-TCO conjugates. (C) Upon hydrolysis by esterases, fluoresein diacetate becomes fluorescent and an active photocatalyst in live cells. (D, E) In-gel fluorescence was used to monitor the progress of subcellular photocatalysis using (D) Fluorescein/470 nm light/60 mW/cm2 and (E) SiR-tJF646/660 nm/450 mW/cm2 light.

Fluorescein diacetate (FDA) was found to be superior to the direct use of fluorescein (FL). FDA is a cell permeable derivative of the ‘closed-form’ of FL. In live mammalian cells, esterases rapidly hydrolyze FDA to produce FL, which due to limited permeability becomes ‘trapped’ intracellularly, manifesting in intracellular green fluorescence (Fig 4C). Therefore, FDA has the advantage of delivering and concentrating FL within the intracellular environment and thereby increases the efficacy of localized photocatalysis. As shown in Fig 4D, HeLa cells expressing HaloTag-H2B-GFP (nucleus) tagged with DHTz-Halo (10 μM), washed and then treated with o-TCO-TAMRA (2.5 μM) and FDA for 30 min. Cells were then washed to remove excess unbound FDA and irradiated at 470 nm at 60 mW/cm2 using an LED light source. Because the nascent FL tends to leak out of the cell after hydrolysis, irradiation was performed within 30 minutes of the washing step. All labeling experiments were carried out for a total of 10 min: after irradiation for the desired amount of time (0–10 min), the cells were kept in the dark for the balance of the 10 min, at which point the TCO reagent was chased by a non-fluorescent tetrazine, and the cells were lysed and analyzed by SDS-PAGE. As shown in Fig 4D, fluorescent labeling increased with irradiation time, with 95% maximum labeling achieved after 5 minutes of illumination. As a qualitative measure of the degree of labeling efficiency, we directly labeled HeLa cells expressing HaloTag-H2B-GFP for 30 min with TAMRA-Halo (structure in Fig 5C): the degree of labeling by in-gel fluorescence was similar to that observed using CABL (Figs 4D and S19). Efficient light-dependent labeling was also observed in similar experiments where DHTz-Halo was targeted to mitochondria or cytosol and labeled with o-TCO-TAMRA under photocatalytic conditions using FDA and 470 nm light (60 mW/cm2). In all cases maximum labeling was achieved within 10 min. CABL was also efficient in E. coli expressing HaloTag that was tagged with DHTz-Halo and labeled with s-TCO-TAMRA using SiR-tJF646 as the photocatalyst, with maximum labeling achieved within 10 min upon irradiation at 660 nm (60 mW/cm2) (Fig. S17).

Fig 5.

Fig 5.

CABL improves the efficiency of a ‘regular’ bioorthogonal tetrazine ligation. (A) Intracellular labeling of a MeTzHalo-tagged protein proceeds with low labeling efficiency relative to direct labeling by a (C) TAMRA-halo control. (B) The low efficiency of the tetrazine ligation was hypothesized to be a result of inactivation of the tetrazine in the cellular environment. (D) UV-vis spectroscopic monitoring of Tz 9 in PBS containing 5 mM glutathione in the presence of SiR shows that tetrazine absorption slowly decreases in the dark, rapidly recovers upon illumination, and again decreases in the dark. (E) The efficiency of tetrazine ligation with MeTzHalo in live cells is improved 3.6-fold through photocatalysis and is more comparable to that observed when HeLa cells expressing Halo-Tag-H2B-GFP were directly tagged with TAMRA-Halo (Fig 5E, right).

As shown in Fig 4E, SiR-tJF646 can be used as photocatalyst in conjunction with 660 nm light for CABL in live cells. Unlike FL, SiR-tJF646 is membrane permeable and is not selectively delivered to the intracellular environment. Therefore, the dye can sensitize singlet oxygen in the extracellular environment where reactive oxygen species are not actively regulated by cellular mechanisms. Under the conditions used with FDA/470 nm, photocatalysis with SiR-tJF646/660 nm led to cellular lysis, which we hypothesized was a result of oxidative stress due to the sensitization of 1O2 in the extracellular environment. While a range of enzymes and reductants including glutathione can mitigate intracellular oxidative stress,91 the extracellular environment lacks comparable mechanisms for resolvingoxidative damage due to 1O2. We found that cellular SiR-photocatalysis was successful when carried out in the presence of 2 mM sodium ascorbate– a non-toxic reductant with low membrane permeability.92 When irradiated by 660 nm light, SiR-tJF646 is deactivated by ascorbate and is therefore incapable of sensitizing extracellular 1O2. However, due to the impermeability of ascorbate, intracellular SiR-tJF646 remains active and capable of catalyzing DHTz oxidation. As shown in Fig 4E, HeLa cells expressing HaloTag-Mito-GFP (mitochondria) were tagged with DHTz-Halo (10 μM), washed and then treated with o-TCO-TAMRA (2.5 μM), sodium ascorbate (2.0 mM) and SiR (2.5 μM) for 20 min. In time course experiments, cells were then irradiated at 660 nm and 450 mW/cm2 using an LED light source. For photocatalytic oxidation by SiR dyes, the rate of photooxidation increases proportionally with increasing light power density from 18 mW/cm2 to 450 mW cm2. 78 All labeling experiments were carried out for a total of 30 min: after irradiation for the desired amount of time (0–30 min), the cells were kept in the dark for the balance of the 10 min, at which point the TCO reagent was chased by a non-fluorescent tetrazine, and the cells were lysed and analyzed by SDS-PAGE. (Fig 4E). Paralleling the observations in Figure 3, CABL is less rapid with SiR/far red catalysis, with maximum labeling achieved after 30 minutes of illumination. CABL using SiR/660 nm light could also be used to activate labeling in the nucleus, actin or cytosol (Fig S20). As discussed above, the inclusion of ascorbate had a dramatic protective effect on cellular viability for SiR-photocatalysis. While high toxicity is observed in the absence of ascorbate, HeLa cells treated with 3.2 μM SiR-tJF646 and 660 nm light (450 mW/cm2) for 30 min show no decrease in cell viability (MTT) after 24 hours when sodium ascorbate (2 mM) is included (Fig S26). By contrast, cells treated with FDA/470 nm light were initially viable but showed ~30% decrease in viability after 2 hours (Fig S26). Thus, SiR-tJF646 photocatalysis is the preferred method where long-term cell viability is desired, whereas FDA is preferred catalyst when shorter reaction times or 2-photon excitation is desired.

Overall, CABL was successful in the majority of subcellular environments that were tested, but we did observe significant background oxidation/Diels-Alder reaction in the absence of light in experiments with a ceramide-DHTz derivative. The especially high oxidizing environment of the Golgi,93 where ceramide is localized, may be responsible for the premature activation of the DHTz in this experiment. We also observed ~10% background oxidation in the activation of a DHTz-Halotag conjugate in HeLa cells expressing HalomCherry-PDGFR on the extracellular surface (Fig S23A). Here, to localize the catalyst to the extracellular environment, impermeant fluorescein (FL) was used as the catalyst instead of fluorescein diacetate.

CABL improves the efficiency of ‘regular’ tetrazine ligations

As discussed above, the efficiency of CABL in live HeLa cells with DHTz-tagged HaloTag-H2B-GFP is comparable to direct labeling by TAMRA-Halo (Fig 5C). By contrast, tagging HaloTag-H2B-GFP with a conventional tetrazine, MeTzHalo, followed by labeling with o-TCO-TAMRA is considerably less efficient (Fig 5A, 5E.). We reasoned that the low labeling efficiency may be partly a consequence of tetrazine reduction in the cellular environment,94 as illustrated in Fig 5B. Tetrazines are reduced to DHTzs by thiols.94 In the presence of SiR (4 μM) but in the dark, incubating 25 μM tetrazine 9 with GSH (5 mM) leads to a reduction in absorption at 262 nm, with ~12% of tetrazine 9 consumed after 110 min (Fig 5D). Irradiation at 660 nm light led to rapid and nearly complete recovery of the absorbance at 262 nm. When the light was turned off, reduction of tetrazine 9 resumed. The observation is consistent with a partial reduction of tetrazine 9 to DHTz 13 by GSH, with the recovery of the tetrazine oxidation state upon catalytic photooxidation.

The efficiency of tetrazine ligation with MeTzHalo in live cells was also improved through photocatalysis. As shown in left of Fig 5E, relatively weak fluorescence was observed in HeLa cells expressing HaloTag-H2B-GFP that were sequentially tagged with MeTzHalo, incubated for 2 h, and then labeled with o-TCO-TAMRA. However, when SiR (2.5 μM) and 660 nm light was applied during the conjugation of o-TCO-TAMRA, a 3.6-fold increase in fluorescence was observed (Fig 5E, center). The efficiency of photocatalytic conjugation is more comparable to that observed when HeLa cells expressing HaloTag-H2B-GFP were directly tagged with TAMRA-Halo (Fig 5E, right).

Fluorescence imaging of subcellular targets in live cells labeled by CABL

The ability to photocatalyze the activation of tetrazine ligation was next applied to selective fluorescent labeling and imaging of subcellular targets. Following the same workflow outlined in Fig 4, HeLa cells were transfected with a GFP-HaloTag construct fused to a POI that controls subcellular localization, and then labeled by the small molecule DHTz-Halo. The cells are treated with SiR photocatalyst (2.5 μM) and o-TCO-TAMRA (1.0 μM), irradiated with 660 nm light for 20 min, and then fixed (4% paraformaldehyde), washed, and imaged by confocal microscopy (Fig 6A).

Fig 6.

Fig 6.

Fluorescence imaging of subcellular targets in live cells labeled by CABL. (A) DHTz-HaloTag is introduced to HeLa cells were transfected with a HaloTag-POI-GFP fusion where the protein of interest (POI) controls subcellular localization. Successful photoactivation and Diels-Alder reaction with TAMRA-TCO results in the colocalization of green fluorescence from the target and red fluorescence from TAMRA. Cells were labeled while live and fixed prior to imaging, and DAPI was added to stain cell nuclei. (B) Confocal fluorescence microscopy images of photoactivate cells with DHTz targeted to LifeAct (actin), GAP43 (cytoplasm), ActA (mitochondria) and H2B (nucleus). Scale bar = 10 μm

Because DHTz-Halo is tagged to a GFP-fusion protein, a successful experiment is expected to result in colocalization of green (GFP) and red (TAMRA) fluorescence at the subcellular target in transfected cells. As shown in Fig 6B, fluorescent colocalization was observed in CABL experiments with various subcellular targets including LifeAct (actin), GAP43 (cytoplasm), ActA (mitochondria) and H2B (nucleus). TAMRA conjugation was not observed in control experiments where light was omitted, neither by confocal microscopy nor by gel electrophoresis (see irradiation t=0 in Fig 4E and Fig S19S20). As expected for transient transfection protocol, only a subset of cells express the fluorescent fusion protein. Fluorophore colocalization was only observed in those cells that expressed the HaloTag-GFP protein complex, and not in neighboring cells that did not display GFP fluorescence (Fig S21 A.)

Live Cell, no-wash Sub-organelle Photopatterning

Two-photon excitation can provide focused light to biological samples with spatiotemporal control. Fig 7A schematically illustrates how CABL was used in combination with two-photon excitation and confocal microscopy in order to pattern well-defined 3D structures at the subcellular level. Using CABL, two photon excitation microscopy (880 nm) is used to excite fluorescein and photocatalyze the generation of reactive tetrazines in the subcellular environment, and fluorescent reporters are concentrated in the region of illumination through Diels-Alder chemistry. The rapid kinetics of the tetrazine-TCO conjugation enables stoichiometric labeling95,96 at very low concentrations (<1 μM) of the fluorescent reporter which is covalently attached in the illuminated area of the live cells under no-wash conditions.

Fig 7.

Fig 7.

Live cell, no-wash photopatterning on the cell nucleus. (A) In the presence of a-TCO-SiR, two-photon excitation microscopy (880 nm) is used to activate fluorescein and photocatalyze the generation of reactive tetrazines in nuclei of HaloTag-H2B-RFP transfected cells. The whole nucleus is red-fluorescent, but only the photopatterned regions are labeled by the far-red SiR-dye. Cells are imaged live immediately after 2-photon excitation. (B-D) Illuminating for 3.3 seconds with focused, 2-photon light is used to label (B) square, (C) an ‘X’ and (D) letters in the cell nucleus. A high laser power (details) was used in (C) to demonstrate that labeling is effective even under conditions that photobleach the RFP fluorophore. (E) For the experiment in Fig 7B, the intensity of fluorescence at 633 nm due to the SiR fluorophore was monitored in the illuminated region of the nucleus as well as in a non-illuminated square region of equivalent area found directly below. (F). Fluorescence intensity timecourse for a cell nucleus of a single cell that was periodically pulsed with light from the 2-photon source at low laser power. The fluorescence threshold was set to the background fluorescence of the SiR dye, and detection above the threshold is displayed. After irradiation, the fluorescence intensity is approximately an order of magnitude higher in the irradiated nucleus relative to the nucleus of a neighboring nucleus. Scale bar = 10 μm

As depicted in Figure 7A, HeLa cells were transfected with HaloTag-H2B-RFP, where RFP is the red fluorescent protein mCherry. The nuclei of the transfected cells were then tagged with DHTz-Halo (10 μM), washed and then treated with a-TCO-SiR (100 nM) and FDA (10 μM). For this experiment, the SiR serves as fluorescent reporter and a two-photon laser provided focused light at the optimal wavelength (880 nm) for 2-photon excitation of the fluorescein photocatalyst. Because of the high intensity of the light, only brief illumination was required to promote efficient photoactivation and fluorescent labeling. As shown in Figure 7B, illuminating a square region within the cell nucleus resulted in effective labeling after irradiating for 3.3 seconds. As shown in Fig 7E and Fig S22 the intensity of fluorescence at 633 nm due to the SiR fluorophore was monitored in the illuminated region of the nucleus as well as in a non-illuminated square region of equivalent area found directly below. The fluorescence intensity was normalized to a region outside of the cell. In the irradiated area, the fluorescence intensity increased by order of magnitude over background, whereas the non-illuminated region did not change (Fig 7E). To demonstrate that CABL is light-dose dependent, a separate experiment was carried out where the entire nucleus of a HaloTag-H2B-RFP transfected cell was periodically pulsed with light from the 2-photon source at 10% laser power. A movie (supplementary movie 1) and Fig7F show that fluorescence intensity rapidly increases (t1/2~7 sec) and then stabilizes after each of three light pulses at which point the maximum fluorescence intensity for the region is reached. During CABL, the signal-to-noise ratio increases by concentrating the fluorescent reporter in the irradiated region through Diels-Alder chemistry. Additionally, the fluorescent signal may also be enhanced by the moderate fluorogenicity of the silarhodamine reporter upon covalent attachment to a protein target.97 As depicted in Fig 7F, in the illuminated cell nucleus the fluorescence intensity is an order of magnitude higher than in a neighboring cell nucleus that was not irradiated. The apparent bimolecular rate constant k2(app) 1 × 106 M−1s−1 measured for this reaction in a live cell nucleus is similar to the in vitro rates observed for a-TCO with Tz 2a (k2 1.46 × 106 M−1s−1, Fig 2c). Interestingly, the rate is ~40% slower for the experiment in Figs 7B, 7E, and supplementary movie 2 where only the interior of the cell nucleus was activated k2(app) 6 × 105 M−1s−1, and may reflect a limitation of diffusion within the nuclear environment. As shown in Fig 7C and 7D, 2-photon irradiation could be used to introduce additional patterns in the live cell nucleus. For the “X” (Fig 7C, Supplementary movie 3), a higher laser power was intentionally used to photobleach the RFP fluorophore, and to show that SiR-fluorescence and RFP photobleaching are colocalized in the experiment. As shown in Fig 7D, CABL can also be used to create letters in the cell nucleus. Supplementary movie 4 of a rotating 3D projection generated from the z-stack of the cell demonstrates that the photopatterning exhibits a high degree of both lateral and axial resolution.

Spatiotemporally resolved labeling of endogenous MAGL

CABL was also used for the spatiotemporally resolved labeling of an endogenous protein target in live cells. We constructed a DHTz probe for the covalent modification of monoacylglycerol lipase (MAGL), a serine hydrolase from the endocannabinoid signaling pathway with broad therapeutic potential (Figure 8A).98 DHTz probe 14 was prepared by conjugating 6 (Figure 2) to a 1-oxa-8-azaspiro[4.5]decane scaffold99 with an electrophilic hexafluoroisopropyl (HFIP) carbamate group for covalently labeling the active site serine100 (Fig 8A). Probe 14 inhibited MAGL activity with 13 nM IC50 in an in vitro assay (Fig S24).101 To test the labeling of endogenous MAGL in live cells, human prostate cancer PC3 cells were treated with probe 14 in Opti-Mem buffer for 1 h, washed and then treated with FDA (10 μM) for 30 min. The cells were again washed and then incubated with 100 nM of a-TCO-SiR in Opti-Mem for 5 min in live cells (Fig 8B), and then irradiated with a 470 nm LED for 1 min. 3-(p-(aminomethyl)phenyl-6-methyltetrazine 9 (125 μM) was then added to quench any unreacted a-TCO-SiR. Cells were lysed and in-gel fluorescence was used to assess MAGL-labeling (Fig 8C,D).102 Fluorescence was observed for two isoforms of MAGL with minimal non-specific labeling from a-TCO-SiR. Less prominent labeling of an additional protein at ~35 kDa was also observed, which is consistent with the reactivity of HFIP-carbamate probes toward toward α/β-hydrolase domain 6 (ABHD6) targets.100,102 The labeling by probe 14 was dose responsive with a cellular IC50 of 2.5 nM, and was competed by a MAGL inhibitor, KML29.100 Consistent with the need for photoactivation, labeling was not observed in controls (1) with FDA catalyst but without irradiation, (2) with irradiation but without FDA, or (3) with neither FDA nor irradiation (Fig 8E).

Fig 8.

Fig 8.

(A) Activity-based labeling of endogenous MAGL in live cells followed by photocatalytic oxidation (B) Structure of MAGL reactive probe 14 and competitive inhibitor KML-29. (C) Live cells were treated with probe 14 for 1 h, washed and then treated with FDA (10 μM) for 30 min, followed by 2 μM a-TCO-SiR for 30 min, and irradiation for 1 min with 470 nm light. A non-fluorescent tetrazine was added to quench unreacted a-TCO-SiR, and cells were lysed and analyzed by in-gel fluorescence. (C) In-gel fluorescence signals for a dose response of probe 14. (D) Dose response fitting of the fluorescence signals of MAGL normalized by the total protein amount indicated by Coomassie staining. Data are reported as mean ± SEM (n = 2). (E) Probe 14 (3.2 nM, 1 h) was competed by pre-treatment with MAGL inhibitor KML29 (300 nM, 1 h). Additional controls include the exclusion of FDA photocatalyst, light or both light and photocatalyst. See Fig S25 for Coomassie staining. (F-I) Confocal microscopy data for live PC3 cells that were labeled by 14 and incubated with a-TCO-SiR before and after wide field illumination with 470 nm light. The fluorescence threshold was set to the background fluorescence of the SiR dye, and detection above the threshold is displayed. (F) SiR fluorescence + DIC prior to illumination. (G) SiR fluorescence + DIC after illumination. (H) SiR fluorescence after illumination. (I) Fluorescence intensity timecourse monitoring of a region of a cell that was periodically pulsed with 488 nm light (Scale bar=10 μm).

In order to determine the kinetics of the activation of probe 14 on endogenous targets, we labeled PC3 cells, we labeled MAGL using the previously described strategy, treating them with FDA (10 μM), washing, and incubating with 50 nm a-TCO-SiR. Without washing, e fluorescence intensity was monitored via confocal microscopy while periodically irradiating short pulses of light using the 488 nm laser line (Figs 8FI). We once again found that SiR signal only increased during photoactivation, and that the fluorescence intensity reached saturation after six pulses of light. After CABL, the cellular distribution of the fluorophore is consistent with that observed in with direct labeling with a fluorescent probe for MAGL.102

Total internal reflection fluorescence (TIRF)103 microscopy was used to study the ability of CABL to activate spatially resolved live-cell MAGL (Fig 9A). In TIRF, incident light is internally reflected in a glass substrate, resulting in the generation labeling of endogenous of an evanescent wave at the glass-liquid interface. As the intensity of the evanescent wave drops exponentially with distance, only chromophores near the coverslip can become efficiently excited.103 We reasoned that TIRF would provide method to study the photocatalytic activation and labeling of DHTz-tagged MAGL at the liquid-glass surface. As DHTz-tagged protein beyond the region evanescent illumination would not be oxidized, labeling is only expected in the thin region near the glass surface (Fig 9A). A monolayer of PC3 cells were plated on piranha-cleaned, poly-L-lysine-coated glass, and sequentially treated with probe 14 in Opti-Mem buffer for 1 h, washed and then treated with FDA (10 μM) for 30 min. The cells were washed and then incubated with 100 nM of a-TCO-SiR in Opti-Mem for 5 min in live cells (Fig 7B), and then promptly illuminated with a 488 nm laser at the critical angle for TIRF irradiation. a-TCO-SiR fluorescence in the cells first was detected with TIRF using the 637 nm laser. Fig 9B displays two adjacent 200 μm2 areas of cells, of which only one was irradiated. a-TCO-SiR fluorescence was higher in the cells irradiated (Fig 9B). To examine the depth of labeling, a z-stack of images through the sample was acquired with spinning disk confocal microscopy (Fig 9CE). In the 3D projection image of an illuminated area in Fig 9C, it is evident that cells were prominently labeled only near the glass interface. An axial view orthogonal stack in Fig 9D was plotted in Fig 9E, and shows that fluorescent intensity was highest within 1 μm from the glass surface and dropped off rapidly thereafter. Together, the results illustrate spatiotemporally controlled activation of the DHTz-labeled MAGL protein in response to photocatalysis.

Fig 9.

Fig 9.

(A) Total internal reflection fluorescence (TIRF) microscopy was used to activate spatially resolved live-cell labeling of endogenous MAGL proteins that were covalently labeled by probe 14. Only protein-DHTz conjugates in the thin region of evanescent illumination become activated and labeled by the a-TCO-SiR fluorophore. (B) Top view of illuminated and non-illuminated cells visualized by fluorescence microscopy. (C,D) Perspective and orthogonal view of illuminated cells visualized by microscopy. In the orthogonal view, the arrow points to the thin layer that becomes fluorescently labeled near the glass surface. (E) Plot of fluorescent intensity vs distance from surface for the orthogonal projection.

Conclusions

Catalytic Activation of Bioorthogonal Chemistry with Light (CABL) is a spatiotemporally controlled method for the labeling and patterning of biomolecules in live cells. Unreactive dihydrotetrazines are photocatalytically oxidized inside live cells to produce reactive tetrazines that are immediately captured by a trans-cyclooctene (TCO) dienophile with kinetics exceeding k2 106 M−1s−1. Fluorescein or silarhodamine dyes are used with activation at 470 or 660 nm, and strategies for limiting the extracellular production of singlet oxygen were developed to increase the cytocompatibility.

Fusions of the HaloTag self-labeling protein were used to localize DHTz-tags and demonstrate high-yielding subcellular activation and labeling of proteins in the nucleus, mitochondria, actin or cytoplasm. 2-Photon excitation microscopy was used to demonstrate that CABL is light-dose dependent and to selectively pattern live cells at the sub-organelle level under no-wash conditions. Spatially resolved live-cell labeling of an endogenous protein target was accomplished by using TIRF microscopy to selectively activate intracellular monoacylglycerol lipase tagged with DHTz-labeled small molecule covalent inhibitor. Beyond spatiotemporally controlled labeling, photocatalysis also improves the efficiency of an ‘ordinary’ tetrazine ligations by rescuing the reactivity of commonly used 3-aryl-6-methyltetrazine reporters that become partially reduced to DHTzs inside cells. We anticipate that the spatiotemporal control and fast rates of photoactivation and labeling of CABL will enable a range of biomolecular labeling applications in living systems.

Supplementary Material

Supp Info
SI Movie 1
Download video file (7.3MB, mov)
SI Movie 2
Download video file (5.9MB, mov)
SI Movie 3
Download video file (6.4MB, mov)
SI Movie 4
Download video file (1.7MB, mov)
SI Movie 5
Download video file (19MB, mov)

ACKNOWLEDGMENT

We thank Luke Lavis and Jonathan Grimm of HHMI for a gift of SiR-tJF646. We thank Lucy Stevens (Pfizer) for determining biochemical MAGL potency. We thank Colin Thorpe (UD), Sam Scinto (UD) and Julia Rosenberger (UD) for insightful discussions. This work was supported by NIH (R01GM132460) and Pfizer. Instrumentation was supported by NIH awards P20GM104316, P20GM103446, S10OD025185, S10OD026951, S10OD016267, S10 OD016361, and S10 OD30321. Facilities and instrumentation were also supported by NSF through the University of Delaware Materials Research Science and Engineering Center, DMR-2011824.

Footnotes

SUPPORTING INFORMATION

Synthetic procedures and compound characterization data; methods for determining dihydrotetrazine stability and photooxidation kinetics; Diels-Alder kinetics; description of plasmids; protocols for HaloTag conjugation and photoactivation; protocols for photoactivation of DHTz-HaloTag in e. coli; protocols for photoactivation of DHTz-HaloTag in HeLa Cells; protocols for photoactivation of DHTz-MAGL in PC3 Cells.

The Supporting Information is available free of charge on the ACS Publications website. The file type is PDF.

The authors declare the following competing financial interest(s): C.W.A. is an employee of Pfizer Inc.

REFERENCES

  • (1).Klán P; Šolomek T; Bochet CG; Blanc A; Givens R; Rubina M; Popik V; Kostikov A; Wirz J. Photoremovable Protecting Groups in Chemistry and Biology: Reaction Mechanisms and Efficacy. Chem. Rev 2013, 113, 119–191. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (2).Kumar GS; Lin Q. Light-Triggered Click Chemistry. Chem. Rev 2021, 121, 6991–7031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (3).Hüll K; Morstein J; Trauner D. In Vivo Photopharmacology. Chem. Rev 2018, 118, 10710–10747. [DOI] [PubMed] [Google Scholar]
  • (4).Lippincott-Schwartz J; Snapp E; Kenworthy A. Studying protein dynamics in living cells. Nat. Rev. Mol. Cell Biol 2001, 2, 444–456. [DOI] [PubMed] [Google Scholar]
  • (5).Lorén N; Hagman J; Jonasson JK; Deschout H; Bernin D; Cella-Zanacchi F; Diaspro A; McNally JG; Ameloot M; Smisdom N; Nydén M; Hermansson A-M; Rudemo M; Braeckmans K. Fluorescence recovery after photobleaching in material and life sciences: putting theory into practice. Q. Rev. Biophys 2015, 48, 323–387. [DOI] [PubMed] [Google Scholar]
  • (6).Lippincott-Schwartz J; Patterson GH Photoactivatable fluorescent proteins for diffraction-limited and super-resolution imaging. Trends Cell Biol. 2009, 19, 555–565. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (7).Wachter RM Photoconvertible Fluorescent Proteins and the Role of Dynamics in Protein Evolution. Int. J. Mol. Sci 2017, 18, 1792. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (8).Zhou XX; Lin MZ Photoswitchable fluorescent proteins: ten years of colorful chemistry and exciting applications. Curr. Opin. Chem. Biol 2013, 17, 682–690. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (9).Banala S; Maurel D; Manley S; Johnsson K. A caged, localizable rhodamine derivative for superresolution microscopy. ACS Chem. Biol 2012, 7, 289–293. [DOI] [PubMed] [Google Scholar]
  • (10).Belov VN; Wurm CA; Boyarskiy VP; Jakobs S; Hell SW Rhodamines NN: a novel class of caged fluorescent dyes. Angew. Chem. Int. Ed. Engl 2010, 49, 3520–3523. [DOI] [PubMed] [Google Scholar]
  • (11).Grimm JB; Klein T; Kopek BG; Shtengel G; Hess HF; Sauer M; Lavis LD Synthesis of a Far-Red Photoactivatable Silicon-Containing Rhodamine for Super-Resolution Microscopy. Angew. Chem. Int. Ed. Engl 2016, 55, 1723–1727. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (12).Kashima H; Kamiya M; Obata F; Kojima R; Nakano S; Miura M; Urano Y. Photoactivatable fluorophores for durable labelling of individual cells. Chem. Commun 2021, 57, 5802–5805. [DOI] [PubMed] [Google Scholar]
  • (13).Puliti D; Warther D; Orange C; Specht A; Goeldner M. Small photoactivatable molecules for controlled fluorescence activation in living cells. Bioorg. Med. Chem 2011, 19, 1023–1029. [DOI] [PubMed] [Google Scholar]
  • (14).Wysocki LM; Grimm JB; Tkachuk AN; Brown TA; Betzig E; Lavis LD Facile and general synthesis of photoactivatable xanthene dyes. Angew. Chem. Int. Ed. Engl 2011, 50, 11206–11209. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (15).Grimm JB; English BP; Choi H; Muthusamy AK; Mehl BP; Dong P; Brown TA; Lippincott-Schwartz J; Liu Z; Lionnet T; Lavis LD Bright photoactivatable fluorophores for single-molecule imaging. Nat. Meth 2016, 13, 985–988. [DOI] [PubMed] [Google Scholar]
  • (16).Frei MS; Hoess P; Lampe M; Nijmeijer B; Kueblbeck M; Ellenberg J; Wadepohl H; Ries J; Pitsch S; Reymond L; Johnsson K. Photoactivation of silicon rhodamines via a light-induced protonation. Nat. Commun 2019, 10, 4580. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (17).Patterson G; Davidson M; Manley S; Lippincott-Schwartz J. Superresolution imaging using single-molecule localization. Annu. Rev. Phys. Chem 2010, 61, 345–367. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (18).Scinto SL; Bilodeau DA; Hincapie R; Lee W; Nguyen SS; Xu M; am Ende CW; Finn MG; Lang K; Lin Q; Pezacki JP; Prescher JA; Robillard MS; Fox JM Bioorthogonal chemistry. Nat. Rev. Methods Primers 2021, 1, 30. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (19).Sletten EM; Bertozzi CR Bioorthogonal Chemistry: Fishing for Selectivity in a Sea of Functionality. Angew. Chem. Int. Ed. Engl 2009, 48, 6974–6998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (20).Lopez Aguilar A; Briard JG; Yang L; Ovryn B; Macauley MS; Wu P. Tools for Studying Glycans: Recent Advances in Chemoenzymatic Glycan Labeling. ACS Chem. Biol 2017, 12, 611–621. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (21).Liang D; Wu K; Tei R; Bumpus TW; Ye J; Baskin JM A real-time, click chemistry imaging approach reveals stimulus-specific subcellular locations of phospholipase D activity. Proc. Nat. Acad. Sci 2019, 116, 15453–15462. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (22).Ganz D; Harijan D; Wagenknecht H-A Labelling of DNA and RNA in the cellular environment by means of bioorthogonal cycloaddition chemistry. RSC Chem. Biol 2020, 1, 86–97. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (23).Lang K; Chin JW Bioorthogonal Reactions for Labeling Proteins. ACS Chem. Biol 2014, 9, 16–20. [DOI] [PubMed] [Google Scholar]
  • (24).Zhang Y; Park K-Y; Suazo KF; Distefano MD Recent progress in enzymatic protein labelling techniques and their applications. Chem. Soc. Rev 2018, 47, 9106–9136. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (25).Murrey HE; Judkins JC; am Ende CW; Ballard TE; Fang Y; Riccardi K; Di L; Guilmette ER; Schwartz JW; Fox JM; Johnson DS Systematic Evaluation of Bioorthogonal Reactions in Live Cells with Clickable HaloTag Ligands: Implications for Intracellular Imaging. J. Am. Chem. Soc 2015, 137, 11461–11475. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (26).Lang K; Chin JW Cellular incorporation of unnatural amino acids and bioorthogonal labeling of proteins. Chem. Rev 2014, 114, 4764–4806. [DOI] [PubMed] [Google Scholar]
  • (27).Cañeque T; Müller S; Rodriguez R. Visualizing biologically active small molecules in cells using click chemistry. Nat. Rev. Chem 2018, 2, 202–215. [Google Scholar]
  • (28).Lambert WD; Fang Y; Mahapatra S; Huang Z; am Ende CW; Fox JM Installation of Minimal Tetrazines through Silver-Mediated Liebeskind-Srogl Coupling with Arylboronic Acids. J. Am. Chem. Soc 2019, 141, 17068–17074. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (29).Xie Y; Fang Y; Huang Z; Tallon A; am Ende CW; Fox J. Divergent Synthesis of Monosubstituted and Unsymmetrical 3,6-Disubstituted Tetrazines from Carboxylic Ester Precursors. Angew. Chem. Int. Ed. Engl 2020, 59, 16967–16973. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (30).Cravatt BF; Hsu K-L; Weerapana E: Activity-Based Protein Profiling; Springer, 2019; Vol. 420. [Google Scholar]
  • (31).Herner A; Lin Q. Photo-Triggered Click Chemistry for Biological Applications. Top. Curr. Chem 2015, 374, 1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (32).Tasdelen MA; Yagci Y. Light-Induced Click Reactions. Angew. Chem. Int. Ed 2013, 52, 5930–5938. [DOI] [PubMed] [Google Scholar]
  • (33).Song W; Wang Y; Qu J; Madden MM; Lin Q. A photoinducible 1,3-dipolar cycloaddition reaction for rapid, selective modification of tetrazole-containing proteins. Angew. Chem. Int. Ed 2008, 47, 2832–2835. [DOI] [PubMed] [Google Scholar]
  • (34).An P; Lewandowski TM; Erbay TG; Liu P; Lin Q. Sterically Shielded, Stabilized Nitrile Imine for Rapid Bioorthogonal Protein Labeling in Live Cells. J. Am. Chem. Soc 2018, 140, 4860–4868. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (35).Yu Z; Lin Q. Design of Spiro[2.3]hex-1-ene, a Genetically Encodable Double-Strained Alkene for Superfast Photoclick Chemistry. J. Am. Chem. Soc 2014, 136, 4153–4156. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (36).Shah L; Laughlin ST; Carrico IS Light-Activated Staudinger–Bertozzi Ligation within Living Animals. J. Am. Chem. Soc 2016, 138, 5186–5189. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (37).Shete AU; El-Zaatari BM; French JM; Kloxin CJ Blue-light activated rapid polymerization for defect-free bulk Cu(i)-catalyzed azide–alkyne cycloaddition (CuAAC) crosslinked networks. Chem. Commun 2016, 52, 10574–10577. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (38).Lim RKV; Lin Q. Azirine ligation: fast and selective protein conjugation via photoinduced azirine–alkene cycloaddition. Chem. Commun 2010, 46, 7993–7995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (39).Gann AW; Amoroso JW; Einck VJ; Rice WP; Chambers JJ; Schnarr NA A photoinduced, benzyne click reaction. Org. Lett 2014, 16, 2003–2005. [DOI] [PubMed] [Google Scholar]
  • (40).Gao J; Xiong Q; Wu X; Deng J; Zhang X; Zhao X; Deng P; Yu Z. Direct ring-strain loading for visible-light accelerated bioorthogonal ligation via diarylsydnone-dibenzo[b,f ][1,4,5]thiadiazepine photo-click reactions. Commun. Chem 2020, 3, 29. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (41).Zhang L; Zhang X; Yao Z; Jiang S; Deng J; Li B; Yu Z. Discovery of Fluorogenic Diarylsydnone-Alkene Photoligation: Conversion of ortho-Dual-Twisted Diarylsydnones into Planar Pyrazolines. J. Am. Chem. Soc 2018, 140, 7390–7394. [DOI] [PubMed] [Google Scholar]
  • (42).Li J; Kong H; Huang L; Cheng B; Qin K; Zheng M; Yan Z; Zhang Y. Visible Light-Initiated Bioorthogonal Photoclick Cycloaddition. J Am Chem Soc 2018, 140, 14542–14546. [DOI] [PubMed] [Google Scholar]
  • (43).Li J; Kong H; Zhu C; Zhang Y. Photo-controllable bioorthogonal chemistry for spatiotemporal control of bio-targets in living systems. Chem. Sci 2020, 11, 3390–3396. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (44).Bruins JJ; Albada B; van Delft F. ortho-Quinones and Analogues Thereof: Highly Reactive Intermediates for Fast and Selective Biofunctionalization. Chem. Eur. J 2018, 24, 4749–4756. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (45).Bruins JJ; Blanco-Ania D; Van Der Doef V; Van Delft FL; Albada B. Orthogonal, dual protein labelling by tandem cycloaddition of strained alkenes and alkynes to ortho-quinones and azides. Chem. Commun 2018, 54, 7338–7341. [DOI] [PubMed] [Google Scholar]
  • (46).Arumugam S; Popik VV Photochemical Generation and the Reactivity ofo-Naphthoquinone Methides in Aqueous Solutions. J. Am. Chem. Soc 2009, 131, 11892–11899. [DOI] [PubMed] [Google Scholar]
  • (47).Feist F; Rodrigues LL; Walden SL; Krappitz TW; Dargaville TR; Weil T; Goldmann AS; Blinco JP; Barner-Kowollik C. Light-induced Ligation of o-Quinodimethanes with Gated Fluorescence Self-reporting. J. Am. Chem. Soc 2020, 142, 7744–7748. [DOI] [PubMed] [Google Scholar]
  • (48).Bruins JJ; Westphal AH; Albada B; Wagner K; Bartels L; Spits H; Van Berkel WJH; Van Delft FL Inducible, Site-Specific Protein Labeling by Tyrosine Oxidation–Strain-Promoted (4 + 2) Cycloaddition. Bioconj. Chem 2017, 28, 1189–1193. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (49).Singh K; Fennell CJ; Coutsias EA; Latifi R; Hartson S; Weaver JD Light Harvesting for Rapid and Selective Reactions: Click Chemistry with Strain-Loadable Alkenes. Chem 2018, 4, 124–137. [Google Scholar]
  • (50).Darko A; Wallace S; Dmitrenko O; Machovina MM; Mehl RA; Chin JW; Fox JM Conformationally Strained trans-Cyclooctene with Improved Stability and Excellent Reactivity in Tetrazine Ligation. Chem. Sci 2014, 5, 3770–3776. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (51).Selvaraj R; Fox JM trans-Cyclooctene--a stable, voracious dienophile for bioorthogonal labeling. Curr. Opin. Chem. Biol 2013, 17, 753–760. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (52).Wu H; Devaraj NK Inverse Electron-Demand Diels-Alder Bioorthogonal Reactions. Top. Curr. Chem 2016, 374, 3. [DOI] [PubMed] [Google Scholar]
  • (53).Wu H; Devaraj NK Advances in Tetrazine Bioorthogonal Chemistry Driven by the Synthesis of Novel Tetrazines and Dienophiles. Acc Chem Res 2018, 51, 1249–1259. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (54).Devaraj NK; Hilderbrand S; Upadhyay R; Mazitschek R; Weissleder R. Bioorthogonal Turn-On Probes for Imaging Small Molecules inside Living Cells. Angew. Chem. Int. Ed. Engl 2010, 49, 2869–2872. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (55).Thompson AD; Bewersdorf J; Toomre D; Schepartz A. HIDE Probes: A New Toolkit for Visualizing Organelle Dynamics, Longer and at Super-Resolution. Biochemistry 2017, 56, 5194–5201. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (56).Werther P; Yserentant K; Braun F; Grußmayer K; Navikas V; Yu M; Zhang Z; Ziegler MJ; Mayer C; Gralak AJ; Busch M; Chi W; Rominger F; Radenovic A; Liu X; Lemke EA; Buckup T; Herten D-P; Wombacher R. Bio-orthogonal Red and Far-Red Fluorogenic Probes for Wash-Free Live-Cell and Super-resolution Microscopy. ACS Cent. Sci 2021, 7, 1561–1571. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (57).Beliu G; Kurz AJ; Kuhlemann AC; Behringer-Pliess L; Meub M; Wolf N; Seibel J; Shi Z-D; Schnermann M; Grimm JB; Lavis LD; Doose S; Sauer M. Bioorthogonal labeling with tetrazine-dyes for super-resolution microscopy. Commun. Biol 2019, 2, 261. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (58).Nikić I; Estrada Girona G; Kang JH; Paci G; Mikhaleva S; Koehler C; Shymanska NV; Ventura Santos C; Spitz D; Lemke EA Debugging Eukaryotic Genetic Code Expansion for Site-Specific Click-PAINT Super-Resolution Microscopy. Angew. Chem. Int. Ed. Engl 2016, 55, 16172–16176. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (59).Uttamapinant C; Howe JD; Lang K; Beránek V; Davis L; Mahesh M; Barry NP; Chin JW Genetic Code Expansion Enables Live-Cell and Super-Resolution Imaging of Site-Specifically Labeled Cellular Proteins. J. Am. Chem. Soc 2015, 137, 4602–4605. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (60).Peng T; Hang HC Site-Specific Bioorthogonal Labeling for Fluorescence Imaging of Intracellular Proteins in Living Cells. J. Am. Chem. Soc 2016, 138, 14423–14433. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (61).Kumar P; Zainul O; Camarda FM; Jiang T; Mannone JA; Huang W; Laughlin ST Caged Cyclopropenes with Improved Tetrazine Ligation Kinetics. Org. Lett 2019, 21, 3721–3725. [DOI] [PubMed] [Google Scholar]
  • (62).Jiang T; Kumar P; Huang W; Kao WS; Thompson AO; Camarda FM; Laughlin ST Modular Enzyme- and Light-Based Activation of Cyclopropene–Tetrazine Ligation. ChemBioChem 2019, 20, 2222–2226. [DOI] [PubMed] [Google Scholar]
  • (63).Mayer SV; Murnauer A; Wrisberg MK; Jokisch ML; Lang K. Photo-induced and Rapid Labeling of Tetrazine-Bearing Proteins via Cyclopropenone-Caged Bicyclononynes. Angew. Chem. Int. Ed 2019, 58, 15876–15882. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (64).Selvaraj R; Fox JM An efficient and mild oxidant for the synthesis of s-tetrazines. Tetrahedron Lett. 2014, 55, 4795–4797. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (65).Ehret F; Wu H; Alexander SC; Devaraj NK Electrochemical Control of Rapid Bioorthogonal Tetrazine Ligations for Selective Functionalization of Microelectrodes. J. Am. Chem. Soc 2015, 137, 8876–8879. [DOI] [PubMed] [Google Scholar]
  • (66).Min DJ; Miomandre F; Audebert P; Kwon JE; Park SY s-Tetrazines as a New Electrode-Active Material for Secondary Batteries. ChemSusChem 2019, 12, 503–510. [DOI] [PubMed] [Google Scholar]
  • (67).Nickerl G; Senkovska I; Kaskel S. Tetrazine functionalized zirconium MOF as an optical sensor for oxidizing gases. Chem. Commun 2015, 51, 2280–2282. [DOI] [PubMed] [Google Scholar]
  • (68).Liu L; Zhang D; Johnson M; Devaraj NK Light-activated tetrazines enable live-cell spatiotemporal control of bioorthogonal reactions. 2020–12-02. bioRxiv 2020, 2020.2012.2001.405423 (accessed 2021-12-29). [Google Scholar]
  • (69).Albada B; Keijzer JF; Zuilhof H; van Delft F. Oxidation-Induced “One-Pot” Click Chemistry. Chem. Rev 2021, 121, 7032–7058. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (70).Zhang H; Trout WS; Liu S; Andrade GA; Hudson DA; Scinto SL; Dicker KT; Li Y; Lazouski N; Rosenthal J; Thorpe C; Jia X; Fox JM Rapid Bioorthogonal Chemistry Turn-on through Enzymatic or Long Wavelength Photocatalytic Activation of Tetrazine Ligation. J. Am. Chem. Soc 2016, 138, 5978–5983. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (71).Carthew J; Frith JE; Forsythe JS; Truong VX Polyethylene glycol–gelatin hydrogels with tuneable stiffness prepared by horseradish peroxidase-activated tetrazine–norbornene ligation. J. Mater. Chem. B 2018, 6, 1394–1401. [DOI] [PubMed] [Google Scholar]
  • (72).Truong VX; Tsang KM; Ercole F; Forsythe JS Red Light Activation of Tetrazine–Norbornene Conjugation for Bioorthogonal Polymer Cross-Linking across Tissue. Chem. Mat 2017, 29, 3678–3685. [Google Scholar]
  • (73).Liu Z; Lv Y; Zhu A; An Z. One-Enzyme Triple Catalysis: Employing the Promiscuity of Horseradish Peroxidase for Synthesis and Functionalization of Well-Defined Polymers. ACS Macro Lett. 2018, 7, 1–6. [DOI] [PubMed] [Google Scholar]
  • (74).Fu M; Xiao Y; Qian X; Zhao D; Xu Y. A design concept of long-wavelength fluorescent analogs of rhodamine dyes: replacement of oxygen with silicon atom. Chem. Commun 2008, 1780–1782. [DOI] [PubMed] [Google Scholar]
  • (75).Koide Y; Urano Y; Hanaoka K; Terai T; Nagano T. Evolution of Group 14 Rhodamines as Platforms for Near-Infrared Fluorescence Probes Utilizing Photoinduced Electron Transfer. ACS Chem. Biol 2011, 6, 600–608. [DOI] [PubMed] [Google Scholar]
  • (76).Koide Y; Urano Y; Hanaoka K; Piao W; Kusakabe M; Saito N; Terai T; Okabe T; Nagano T. Development of NIR Fluorescent Dyes Based on Si–rhodamine for in Vivo Imaging. J. Am. Chem. Soc 2012, 134, 5029–5031. [DOI] [PubMed] [Google Scholar]
  • (77).Grimm JB; Brown TA; Tkachuk AN; Lavis LD General Synthetic Method for Si-Fluoresceins and Si-Rhodamines. ACS Cent. Sci 2017, 3, 975–985. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (78).Wang C; Zhang H; Zhang T; Zou X; Wang H; Rosenberger JE; Vannam R; Trout WS; Grimm JB; Lavis LD; Thorpe C; Jia X; Li Z; Fox JM Enabling In Vivo Photocatalytic Activation of Rapid Bioorthogonal Chemistry by Repurposing Silicon-Rhodamine Fluorophores as Cytocompatible Far-Red Photocatalysts. J. Am. Chem. Soc 2021, 143, 10793–10803. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (79).Wang D; Chen W; Zheng Y; Dai C; Wang L; Wang B: A general and efficient entry to asymmetric tetrazines for click chemistry applications. Heterocycl. Commun, 2013, 19, 171–177. [Google Scholar]
  • (80).Pilgram K; Skiles RD Unsymmetrically 3,6-disubstituted s-tetrazines. Synthesis of 3-aryl-6-(perfluoroalkyl)-1,2,4,5-tetrazines and 1,2-dihydro derivatives. J. Org. Chem 1976, 41, 3392–3395. [Google Scholar]
  • (81).Chung SW; Uccello DP; Choi HW; Montgomery JI; Chen JS Trimethylaluminium-Facilitated Direct Amidation of Carboxylic Acids. Synlett 2011, 2072–2074. [Google Scholar]
  • (82).To remove adventitous metal impurities, PBS buffer was purified by filtration through chelex resin prior to use.
  • (83).Karver MR; Weissleder R; Hilderbrand SA Synthesis and Evaluation of a Series of 1,2,4,5-Tetrazines for Bioorthogonal Conjugation. Bioconj. Chem 2011, 22, 2263–2270. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (84).Blackman ML; Royzen M; Fox JM Tetrazine ligation: Fast bioconjugation based on inverse-electron-demand Diels-Alder reactivity. J. Am. Chem. Soc 2008, 130, 13518–13519. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (85).Svatunek D; Wilkovitsch M; Hartmann L; Houk K; Mikula H. Uncovering the key role of distortion in tetrazine ligations guides the design of bioorthogonal tools with high reactivity and superior stability. 2021–08-17. ChemRxiv 2021, 10.33774/chemrxiv-32021-vnwzk. (accessed 2021-12-29). [DOI] [Google Scholar]
  • (86).Pigga JE; Rosenberger JE; Jemas A; Boyd SJ; Dmitrenko O; Xie Y; Fox JM General, Divergent Platform for Diastereoselective Synthesis of trans-Cyclooctenes with High Reactivity and Favorable Physiochemical Properties. Angew. Chem. Int. Ed. Engl 2021, 60, 14975–14980. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (87).Lambert WD; Scinto SL; Dmitrenko O; Boyd SJ; Magboo R; Mehl RA; Chin JW; Fox JM; Wallace S. Computationally guided discovery of a reactive, hydrophilic trans-5-oxocene dienophile for bioorthogonal labeling. Org. Biomol. Chem 2017, 15, 6640–6644. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (88).Taylor MT; Blackman ML; Dmitrenko O; Fox JM Design and synthesis of highly reactive dienophiles for the tetrazine-trans-cyclooctene ligation. J. Am. Chem. Soc 2011, 133, 9646–9649. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (89).Romero NA; Nicewicz DA Organic Photoredox Catalysis. Chem. Rev 2016, 116, 10075–10166. [DOI] [PubMed] [Google Scholar]
  • (90).Wang H; Li W-G; Zeng K; Wu Y-J; Zhang Y; Xu T-L; Chen Y. Photocatalysis Enables Visible-Light Uncaging of Bioactive Molecules in Live Cells. Angew. Chem. Int. Ed. Engl 2019, 58, 561–565. [DOI] [PubMed] [Google Scholar]
  • (91).Pompella A; Visvikis A; Paolicchi A; Tata VD; Casini AF The changing faces of glutathione, a cellular protagonist. Biochem. Pharmacology 2003, 66, 1499–1503. [DOI] [PubMed] [Google Scholar]
  • (92).Hannesschlaeger C; Pohl P. Membrane Permeabilities of Ascorbic Acid and Ascorbate. Biomolecules 2018, 8, 73. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (93).Hatori Y; Kubo T; Sato Y; Inouye S; Akagi R; Seyama T. Visualization of the Redox Status of Cytosolic Glutathione Using the Organelle- and Cytoskeleton-Targeted Redox Sensors. Antioxidants 2020, 9, 129. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (94).Umlauf BJ; Mix KA; Grosskopf VA; Raines RT; Shusta EV Site-Specific Antibody Functionalization Using Tetrazine-Styrene Cycloaddition. Bioconj. Chem 2018, 29, 1605–1613. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (95).Jang HS; Jana S; Blizzard RJ; Meeuwsen JC; Mehl RA Access to Faster Eukaryotic Cell Labeling with Encoded Tetrazine Amino Acids. J. Am. Chem. Soc 2020, 142, 7245–7249. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (96).Blizzard RJ; Backus DR; Brown W; Bazewicz CG; Li Y; Mehl RA Ideal Bioorthogonal Reactions Using A Site-Specifically Encoded Tetrazine Amino Acid. J. Am. Chem. Soc 2015, 137, 10044–10047. [DOI] [PubMed] [Google Scholar]
  • (97).Lukinavičius G; Umezawa K; Olivier N; Honigmann A; Yang G; Plass T; Mueller V; Reymond L; Corrêa IR Jr; Luo Z-G; Schultz C; Lemke EA; Heppenstall P; Eggeling C; Manley S; Johnsson K. A near-infrared fluorophore for live-cell super-resolution microscopy of cellular proteins. Nature Chemistry 2013, 5, 132–139. [DOI] [PubMed] [Google Scholar]
  • (98).Blankman JL; Cravatt BF Chemical probes of endocannabinoid metabolism. Pharmacological Rev. 2013, 65, 849–871. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (99).Kohrt JT; Dorff PH; Burns M; Lee C; O’Neil SV; Maguire RJ; Kumar R; Wagenaar M; Price L; Lall MS Application of Flow and Biocatalytic Transaminase Technology for the Synthesis of a 1-Oxa-8-azaspiro[4.5]decan-3-amine. Org. Proc. Res. Dev 2021, 10.1021/acs.oprd.1c00075. [DOI] [Google Scholar]
  • (100).Chang JW; Niphakis MJ; Lum KM; Cognetta AB 3rd; Wang C; Matthews ML; Niessen S; Buczynski MW; Parsons LH; Cravatt BF Highly selective inhibitors of monoacylglycerol lipase bearing a reactive group that is bioisosteric with endocannabinoid substrates. Chem. Biol 2012, 19, 579–588. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (101).Butler CR; Beck EM; Harris A; Huang Z; McAllister LA; am Ende CW; Fennell K; Foley TL; Fonseca K; Hawrylik SJ; Johnson DS; Knafels JD; Mente S; Noell GS; Pandit J; Phillips TB; Piro JR; Rogers BN; Samad TA; Wang J; Wan S; Brodney MA Azetidine and Piperidine Carbamates as Efficient, Covalent Inhibitors of Monoacylglycerol Lipase. J. Med. Chem 2017, 60, 9860–9873. [DOI] [PubMed] [Google Scholar]
  • (102).Chang JW; Cognetta AB 3rd; Niphakis MJ; Cravatt BF Proteome-wide reactivity profiling identifies diverse carbamate chemotypes tuned for serine hydrolase inhibition. ACS Chem. Biol 2013, 8, 1590–1599. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (103).Fish KN Total internal reflection fluorescence (TIRF) microscopy. Curr. Protocols in Cytometry 2009, Chapter 12. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supp Info
SI Movie 1
Download video file (7.3MB, mov)
SI Movie 2
Download video file (5.9MB, mov)
SI Movie 3
Download video file (6.4MB, mov)
SI Movie 4
Download video file (1.7MB, mov)
SI Movie 5
Download video file (19MB, mov)

RESOURCES