fbpx
Wikipedia

Bioorthogonal chemistry

The term bioorthogonal chemistry refers to any chemical reaction that can occur inside of living systems without interfering with native biochemical processes.[1][2][3] The term was coined by Carolyn R. Bertozzi in 2003.[4][5] Since its introduction, the concept of the bioorthogonal reaction has enabled the study of biomolecules such as glycans, proteins,[6] and lipids[7] in real time in living systems without cellular toxicity. A number of chemical ligation strategies have been developed that fulfill the requirements of bioorthogonality, including the 1,3-dipolar cycloaddition between azides and cyclooctynes (also termed copper-free click chemistry),[8] between nitrones and cyclooctynes,[9] oxime/hydrazone formation from aldehydes and ketones,[10] the tetrazine ligation,[11] the isocyanide-based click reaction,[12] and most recently, the quadricyclane ligation.[13]

Shown here is a bioorthogonal ligation between biomolecule X and reactive partner Y. To be considered bioorthogonal, these reactive partners cannot perturb other chemical functionality naturally found within the cell.

The use of bioorthogonal chemistry typically proceeds in two steps. First, a cellular substrate is modified with a bioorthogonal functional group (chemical reporter) and introduced to the cell; substrates include metabolites, enzyme inhibitors, etc. The chemical reporter must not alter the structure of the substrate dramatically to avoid affecting its bioactivity. Secondly, a probe containing the complementary functional group is introduced to react and label the substrate.

Although effective bioorthogonal reactions such as copper-free click chemistry have been developed, development of new reactions continues to generate orthogonal methods for labeling to allow multiple methods of labeling to be used in the same biosystems. Carolyn R. Bertozzi was awarded the Nobel Prize in Chemistry in 2022 for her development of click chemistry and bioorthogonal chemistry.[14]

Etymology edit

The word bioorthogonal comes from Greek bio- "living" and orthogōnios "right-angled". Thus literally a reaction that goes perpendicular to a living system, thus not disturbing it.

Requirements for bioorthogonality edit

To be considered bioorthogonal, a reaction must fulfill a number of requirements:

  • Selectivity: The reaction must be selective between endogenous functional groups to avoid side reactions with biological compounds
  • Biological inertness: Reactive partners and resulting linkage should not possess any mode of reactivity capable of disrupting the native chemical functionality of the organism under study.
  • Chemical inertness: The covalent link should be strong and inert to biological reactions.
  • Kinetics: The reaction must be rapid so that covalent ligation is achieved prior to probe metabolism and clearance. The reaction must be fast, on the time scale of cellular processes (minutes) to prevent competition in reactions which may diminish the small signals of less abundant species. Rapid reactions also offer a fast response, necessary in order to accurately track dynamic processes.
  • Reaction biocompatibility: Reactions have to be non-toxic and must function in biological conditions taking into account pH, aqueous environments, and temperature. Pharmacokinetics are a growing concern as bioorthogonal chemistry expands to live animal models.
  • Accessible engineering: The chemical reporter must be capable of incorporation into biomolecules via some form of metabolic or protein engineering. Optimally, one of the functional groups is also very small so that it does not disturb native behavior.

Staudinger ligation edit

The Staudinger ligation is a reaction developed by the Bertozzi group in 2000 that is based on the classic Staudinger reaction of azides with triarylphosphines.[15] It launched the field of bioorthogonal chemistry as the first reaction with completely abiotic functional groups although it is no longer as widely used. The Staudinger ligation has been used in both live cells and live mice.[5]

Bioorthogonality edit

The azide can act as a soft electrophile that prefers soft nucleophiles such as phosphines. This is in contrast to most biological nucleophiles which are typically hard nucleophiles. The reaction proceeds selectively under water-tolerant conditions to produce a stable product.

Phosphines are completely absent from living systems and do not reduce disulfide bonds despite mild reduction potential. Azides had been shown to be biocompatible in FDA-approved drugs such as azidothymidine and through other uses as cross linkers. Additionally, their small size allows them to be easily incorporated into biomolecules through cellular metabolic pathways.

Mechanism edit

Classic Staudinger reaction edit

 

The nucleophilic phosphine attacks the azide at the electrophilic terminal nitrogen. Through a four-membered transition state, N2 is lost to form an aza-ylide. The unstable ylide is hydrolyzed to form phosphine oxide and a primary amine. However, this reaction is not immediately bioorthogonal because hydrolysis breaks the covalent bond in the aza-ylide.

Staudinger ligation edit

 

The reaction was modified to include an ester group ortho to the phosphorus atom on one of the aryl rings to direct the aza-ylide through a new path of reactivity in order to outcompete immediate hydrolysis by positioning the ester to increase local concentration. The initial nucleophilic attack on the azide is the rate-limiting step. The ylide reacts with the electrophilic ester trap through intramolecular cyclization to form a five-membered ring. This ring undergoes hydrolysis to form a stable amide bond.

Limitations edit

The phosphine reagents slowly undergo air oxidation in living systems. Additionally, it is likely that they are metabolized in vitro by cytochrome P450 enzymes.

The kinetics of the reactions are slow with second order rate constants around 0.0020 M−1•s−1. Attempts to increase nucleophilic attack rates by adding electron-donating groups to the phosphines improved kinetics, but also increased the rate of air oxidation.

The poor kinetics require that high concentrations of the phosphine be used which leads to problems with high background signal in imaging applications. Attempts have been made to combat the problem of high background through the development of a fluorogenic phosphine reagents based on fluorescein and luciferin, but the intrinsic kinetics remain a limitation.[16]

Copper-free click chemistry edit

Copper-free click chemistry is a bioorthogonal reaction first developed by Carolyn Bertozzi as an activated variant of an azide alkyne Huisgen cycloaddition, based on the work by Karl Barry Sharpless et al. Unlike CuAAC, Cu-free click chemistry has been modified to be bioorthogonal by eliminating a cytotoxic copper catalyst, allowing reaction to proceed quickly and without live cell toxicity. Instead of copper, the reaction is a strain-promoted alkyne-azide cycloaddition (SPAAC). It was developed as a faster alternative to the Staudinger ligation, with the first generations reacting over sixty times faster. The bioorthogonality of the reaction has allowed the Cu-free click reaction to be applied within cultured cells, live zebrafish, and mice.

 
click chemistry labeling

Copper toxicity edit

The classic copper-catalyzed azide-alkyne cycloaddition has been an extremely fast and effective click reaction for bioconjugation, but it is not suitable for use in live cells due to the toxicity of Cu(I) ions. Toxicity is due to oxidative damage from reactive oxygen species formed by the copper catalysts. Copper complexes have also been found to induce changes in cellular metabolism and are taken up by cells.

There has been some development of ligands to prevent biomolecule damage and facilitate removal in in vitro applications. However, it has been found that different ligand environments of complexes can still affect metabolism and uptake, introducing an unwelcome perturbation in cellular function.[17]

Bioorthogonality edit

The azide group is particularly bioorthogonal because it is extremely small (favorable for cell permeability and avoids perturbations), metabolically stable, and does not naturally exist in cells and thus has no competing biological side reactions. Although azides are not the most reactive 1,3-dipole available for reaction, they are preferred for their relative lack of side reactions and stability in typical synthetic conditions.[18] The alkyne is not as small, but it still has the stability and orthogonality necessary for in vivo labeling. Cyclooctynes are traditionally the most common cycloalkyne for labeling studies, as they are the smallest stable alkyne ring.

Mechanism edit

 

The reaction proceeds as a standard 1,3-dipolar cycloaddition, a type of asynchronous, concerted pericyclic shift. The ambivalent nature of the 1,3-dipole should make the identification of an electrophilic or nucleophilic center on the azide impossible such that the direction of the cyclic electron flow is meaningless. [p] However, computation has shown that the electron distribution amongst nitrogens causes the innermost nitrogen atom to bear the greatest negative charge.[19]

Regioselectivity edit

Although the reaction produces a regioisomeric mixture of triazoles, the lack of regioselectivity in the reaction is not a major concern for most current applications. More regiospecific and less bioorthogonal requirements are best served by copper-catalyzed Huisgen cycloaddition, especially given the synthetic difficulty (compared to the addition of a terminal alkyne) of synthesizing a strained cyclooctyne.

Development of cyclooctynes edit

Cyclooctyne Second order rate constant (M−1s−1)
OCT 0.0024
ALO 0.0013
MOFO 0.0043
DIFO 0.076
DIBO 0.057
BARAC 0.96
DIBAC (ADIBO) 0.31
DIMAC 0.0030
 
Strained cyclooctynes developed for copper-free click chemistry

OCT was the first cyclooctyne developed for Cu-free click chemistry. While linear alkynes are unreactive at physiological temperatures, OCT was able readily react with azides in biological conditions while showing no toxicity. However, it was poorly water-soluble, and the kinetics were barely improved over the Staudinger ligation. ALO (aryl-less octyne) was developed to improve water solubility, but it still had poor kinetics.

Monofluorinated (MOFO) and difluorinated (DIFO) cyclooctynes were created to increase the rate through the addition of electron-withdrawing fluorine substituents at the propargylic position. Fluorine is a good electron-withdrawing group in terms of synthetic accessibility and biological inertness. In particular, it cannot form an electrophilic Michael acceptor that may side-react with biological nucleophiles.[8] DIBO (dibenzocyclooctyne) was developed as a fusion to two aryl rings, resulting in very high strain and a decrease in distortion energies. It was proposed that biaryl substitution increases ring strain and provides conjugation with the alkyne to improve reactivity. Although calculations have predicted that mono-aryl substitution would provide an optimal balance between steric clash (with azide molecule) and strain,[20] monoarylated products have been shown to be unstable.

BARAC (biarylazacyclooctynone) followed with the addition of an amide bond which adds an sp2-like center to increase rate by distortion. Amide resonance contributes additional strain without creating additional unsaturation which would lead to an unstable molecule. Additionally, the addition of a heteroatom into the cyclooctyne ring improves both solubility and pharmacokinetics of the molecule. BARAC has sufficient rate (and sensitivity) to the extent that washing away excess probe is unnecessary to reduce background. This makes it extremely useful in situations where washing is impossible as in real-time imaging or whole animal imaging. Although BARAC is extremely useful, its low stability requires that it must be stored at 0 °C, protected from light and oxygen.[21]

 
The synthesis was designed by the Bertozzi group as a modular route to facilitate future modifications in SAR analysis. The first step is Fischer indole synthesis. The product is alkylated with allyl bromide as a handle for future probe attachment; TMS is then added. Oxidation opens the central rings to form a cyclic amide. The ketone is treated as an enolate to add a triflate group. Reaction of the terminal alkene generates a linker for conjugation to a molecule. The final reaction with CsF introduces the strained alkyne at the last step.

Further adjustments variations on BARAC to produce DIBAC/ADIBO were performed to add distal ring strain and reduce sterics around the alkyne to further increase reactivity. Keto-DIBO, in which the hydroxyl group has been converted to a ketone, has a three-fold increase in rate due to a change in ring conformation. Attempts to make a difluorobenzocyclooctyne (DIFBO) were unsuccessful due to the instability.

Problems with DIFO with in vivo mouse studies illustrate the difficulty of producing bioorthogonal reactions. Although DIFO was extremely reactive in the labeling of cells, it performed poorly in mouse studies due to binding with serum albumin. Hydrophobicity of the cyclooctyne promotes sequestration by membranes and serum proteins, reducing bioavailable concentrations. In response, DIMAC (dimethoxyazacyclooctyne) was developed to increase water solubility, polarity, and pharmacokinetics,[22] although efforts in bioorthogonal labeling of mouse models is still in development.

Reactivity edit

Computational efforts have been vital in explaining the thermodynamics and kinetics of these cycloaddition reactions which has played a vital role in continuing to improve the reaction. There are two methods for activating alkynes without sacrificing stability: decrease transition state energy or decrease reactant stability.

 
The red arrow shows the direction of energy change. Black arrows show the difference in activation energy before and after the effects.

Decreasing reactant stability: Houk[23] has proposed that differences in the energy (Ed) required to distort the azide and alkyne into the transition state geometries control the barrier heights for the reaction. The activation energy (E) is the sum of destabilizing distortions and stabilizing interactions (Ei). The most significant distortion is in the azide functional group with lesser contribution of alkyne distortion. However, it is only the cyclooctyne that can be easily modified for higher reactivity. Calculated barriers of reaction for phenyl azide and acetylene (16.2 kcal/mol) versus cyclooctyne (8.0 kcal/mol) results in a predicted rate increase of 106. The cyclooctyne requires less distortion energy (1.4 kcal/mol versus 4.6 kcal/mol) resulting in a lower activation energy despite smaller interaction energy.

 
Relationship between activation energy, distortion energy, and interaction energy

Decreasing transition state energy: Electron withdrawing groups such as fluorine increase rate by decreasing LUMO energy and the HOMO-LUMO gap. This leads to a greater charge transfer from the azide to the fluorinated cyclooctyne in the transition state, increasing interaction energy (lower negative value) and overall activation energy.[24] The lowering of the LUMO is the result of hyperconjugation between alkyne π donor orbitals and CF σ* acceptors. These interactions provide stabilization primarily in the transition state as a result of increased donor/acceptor abilities of the bonds as they distort. NBO calculations have shown that transition state distortion increases the interaction energy by 2.8 kcal/mol.

The hyperconjugation between out-of-plane π bonds is greater because the in-plane π bonds are poorly aligned. However, transition state bending allows the in-plane π bonds to have a more antiperiplanar arrangement that facilitates interaction. Additional hyperconjugative interaction energy stabilization is achieved through an increase in the electronic population of the σ* due to the forming CN bond. Negative hyperconjugation with the σ* CF bonds enhances this stabilizing interaction.[19]

Regioselectivity edit

Although regioselectivity is not a great issue in the current imaging applications of copper-free click chemistry, it is an issue that prevents future applications in fields such as drug design or peptidomimetics.[25]

Currently most cyclooctynes react to form regioisomeric mixtures. [m] Computation analysis has found that while gas phase regioselectivity is calculated to favor 1,5 addition over 1,4 addition by up to 2.9 kcal/mol in activation energy, solvation corrections result in the same energy barriers for both regioisomers. While the 1,4 isomer in the cycloaddition of DIFO is disfavored by its larger dipole moment, solvation stabilizes it more strongly than the 1,5 isomer, eroding regioselectivity.[24]

 

Symmetrical cyclooctynes such as BCN (bicyclo[6.1.0]nonyne) form a single regioisomer upon cycloaddition[26] and may serve to address this problem in the future.

Applications edit

The most widespread application of copper-free click chemistry is in biological imaging in live cells or animals using an azide-tagged biomolecule and a cyclooctyne bearing an imaging agent.

Fluorescent keto and oxime variants of DIBO are used in fluoro-switch click reactions in which the fluorescence of the cyclooctyne is quenched by the triazole that forms in the reaction.[27] On the other hand, coumarin-conjugated cyclooctynes such as coumBARAC have been developed such that the alkyne suppresses fluorescence while triazole formation increases the fluorescence quantum yield by ten-fold.[28]

 
coumBARAC fluorescence increases with reaction

Spatial and temporal control of substrate labeling has been investigated using photoactivatable cyclooctynes. This allows equilibration of the alkyne prior to reaction in order to reduce artifacts as a result of concentration gradients. Masked cyclooctynes are unable to react with azides in the dark but become reactive alkynes upon irradiation with light.[29]

 

Copper-free click chemistry is being explored for use in synthesizing PET imaging agents which must be made quickly with high purity and yield in order to minimize isotopic decay before the compounds can be administered. Both the high rate constants and the bioorthogonality of SPAAC are amenable to PET chemistry.[30]

Other bioorthogonal reactions edit

Nitrone dipole cycloaddition edit

Copper-free click chemistry has been adapted to use nitrones as the 1,3-dipole rather than azides and has been used in the modification of peptides.[9]

 

This cycloaddition between a nitrone and a cyclooctyne forms N-alkylated isoxazolines. The reaction rate is enhanced by water and is extremely fast with second order rate constants ranging from 12 to 32 M−1•s−1, depending on the substitution of the nitrone. Although the reaction is extremely fast, it faces problems in incorporating the nitrone into biomolecules through metabolic labeling. Labeling has only been achieved through post-translational peptide modification.

Norbornene cycloaddition edit

1,3 dipolar cycloadditions have been developed as a bioorthogonal reaction using a nitrile oxide as a 1,3-dipole and a norbornene as a dipolarophile. Its primary use has been in labeling DNA and RNA in automated oligonucleotide synthesizers,[31] and polymer crosslinking in the presence of living cells.[32]

 

Norbornenes were selected as dipolarophiles due to their balance between strain-promoted reactivity and stability. The drawbacks of this reaction include the cross-reactivity of the nitrile oxide due to strong electrophilicity and slow reaction kinetics.

Oxanorbornadiene cycloaddition edit

The oxanorbornadiene cycloaddition is a 1,3-dipolar cycloaddition followed by a retro-Diels Alder reaction to generate a triazole-linked conjugate with the elimination of a furan molecule.[33] Preliminary work has established its usefulness in peptide labeling experiments, and it has also been used in the generation of SPECT imaging compounds.[34] More recently, the use of an oxanorbornadiene was described in a catalyst-free room temperature "iClick" reaction, in which a model amino acid is linked to the metal moiety, in a novel approach to bioorthogonal reactions.[35]

 

Ring strain and electron deficiency in the oxanorbornadiene increase reactivity towards the cycloaddition rate-limiting step. The retro-Diels Alder reaction occurs quickly afterwards to form the stable 1,2,3 triazole. Problems include poor tolerance for substituents which may change electronics of the oxanorbornadiene and low rates (second order rate constants on the order of 10−4).

Tetrazine ligation edit

The tetrazine ligation is the reaction of a trans-cyclooctene and an s-tetrazine in an inverse-demand Diels Alder reaction followed by a retro-Diels Alder reaction to eliminate nitrogen gas.[36] The reaction is extremely rapid with a second order rate constant of 2000 M−1–s−1 (in 9:1 methanol/water) allowing modifications of biomolecules at extremely low concentrations.

 

Based on computational work by Bach, the strain energy for Z-cyclooctenes is 7.0 kcal/mol compared to 12.4 kcal/mol for cyclooctane due to a loss of two transannular interactions. E-cyclooctene has a highly twisted double bond resulting in a strain energy of 17.9 kcal/mol.[37] As such, the highly strained trans-cyclooctene is used as a reactive dienophile. The diene is a 3,6-diaryl-s-tetrazine which has been substituted in order to resist immediate reaction with water. The reaction proceeds through an initial cycloaddition followed by a reverse Diels Alder to eliminate N2 and prevent reversibility of the reaction.[11]

Not only is the reaction tolerant of water, but it has been found that the rate increases in aqueous media. Reactions have also been performed using norbornenes as dienophiles at second order rates on the order of 1 M−1•s−1 in aqueous media. The reaction has been applied in labeling live cells[38] and polymer coupling.[39]

[4+1] Cycloaddition edit

This isocyanide click reaction is a [4+1] cycloaddition followed by a retro-Diels Alder elimination of N2.[12]

 

The reaction proceeds with an initial [4+1] cycloaddition followed by a reversion to eliminate a thermodynamic sink and prevent reversibility. This product is stable if a tertiary amine or isocyanopropanoate is used. If a secondary or primary isocyanide is used, the produce will form an imine which is quickly hydrolyzed.

Isocyanide is a favored chemical reporter due to its small size, stability, non-toxicity, and absence in mammalian systems. However, the reaction is slow, with second order rate constants on the order of 10−2 M−1•s−1.

Tetrazole photoclick chemistry edit

Photoclick chemistry utilizes a photoinduced cycloelimination to release N2. This generates a short-lived 1,3 nitrile imine intermediate via the loss of nitrogen gas, which undergoes a 1,3-dipolar cycloaddition with an alkene to generate pyrazoline cycloadducts.[12]

 

Photoinduction takes place with a brief exposure to light (wavelength is tetrazole-dependent) to minimize photodamage to cells. The reaction is enhanced in aqueous conditions and generates a single regioisomer.

The transient nitrile imine is highly reactive for 1,3-dipolar cycloaddition due to a bent structure which reduces distortion energy. Substitution with electron-donating groups on phenyl rings increases the HOMO energy, when placed on the 1,3 nitrile imine and increases the rate of reaction.

Advantages of this approach include the ability to spatially or temporally control reaction and the ability to incorporate both alkenes and tetrazoles into biomolecules using simple biological methods such as genetic encoding.[40] Additionally, the tetrazole can be designed to be fluorogenic in order to monitor progress of the reaction.[41]

Quadricyclane ligation edit

The quadricyclane ligation utilizes a highly strained quadricyclane to undergo [2+2+2] cycloaddition with π systems.[13]

 

Quadricyclane is abiotic, unreactive with biomolecules (due to complete saturation), relatively small, and highly strained (~80 kcal/mol). However, it is highly stable at room temperature and in aqueous conditions at physiological pH. It is selectively able to react with electron-poor π systems but not simple alkenes, alkynes, or cyclooctynes.

Bis(dithiobenzil)nickel(II) was chosen as a reaction partner out of a candidate screen based on reactivity. To prevent light-induced reversion to norbornadiene, diethyldithiocarbamate is added to chelate the nickel in the product.

 

These reactions are enhanced by aqueous conditions with a second order rate constant of 0.25 M−1•s−1. Of particular interest is that it has been proven to be bioorthogonal to both oxime formation and copper-free click chemistry.

Uses edit

Bioorthogonal chemistry is an attractive tool for pretargeting experiments in nuclear imaging and radiotherapy.[42]

References edit

  1. ^ Sletten, Ellen M.; Bertozzi, Carolyn R. (2009). "Bioorthogonal Chemistry: Fishing for Selectivity in a Sea of Functionality". Angewandte Chemie International Edition. 48 (38): 6974–98. doi:10.1002/anie.200900942. PMC 2864149. PMID 19714693.
  2. ^ Prescher, Jennifer A.; Dube, Danielle H.; Bertozzi, Carolyn R. (2004). "Chemical remodelling of cell surfaces in living animals". Nature. 430 (7002): 873–7. Bibcode:2004Natur.430..873P. doi:10.1038/nature02791. PMID 15318217. S2CID 4371934.
  3. ^ Prescher, Jennifer A; Bertozzi, Carolyn R (2005). "Chemistry in living systems". Nature Chemical Biology. 1 (1): 13–21. doi:10.1038/nchembio0605-13. PMID 16407987. S2CID 40548615.
  4. ^ Hang, Howard C.; Yu, Chong; Kato, Darryl L.; Bertozzi, Carolyn R. (2003-12-09). "A metabolic labeling approach toward proteomic analysis of mucin-type O-linked glycosylation". Proceedings of the National Academy of Sciences. 100 (25): 14846–14851. Bibcode:2003PNAS..10014846H. doi:10.1073/pnas.2335201100. ISSN 0027-8424. PMC 299823. PMID 14657396.
  5. ^ a b Sletten, Ellen M.; Bertozzi, Carolyn R. (2011). "From Mechanism to Mouse: A Tale of Two Bioorthogonal Reactions". Accounts of Chemical Research. 44 (9): 666–676. doi:10.1021/ar200148z. PMC 3184615. PMID 21838330.
  6. ^ Plass, Tilman; Milles, Sigrid; Koehler, Christine; Schultz, Carsten; Lemke, Edward A. (2011). "Genetically Encoded Copper-Free Click Chemistry". Angewandte Chemie International Edition. 50 (17): 3878–3881. doi:10.1002/anie.201008178. PMC 3210829. PMID 21433234.
  7. ^ Neef, Anne B.; Schultz, Carsten (2009). "Selective Fluorescence Labeling of Lipids in Living Cells". Angewandte Chemie International Edition. 48 (8): 1498–500. doi:10.1002/anie.200805507. PMID 19145623.
  8. ^ a b Baskin, J. M.; Prescher, J. A.; Laughlin, S. T.; Agard, N. J.; Chang, P. V.; Miller, I. A.; Lo, A.; Codelli, J. A.; Bertozzi, C. R. (2007). "Copper-free click chemistry for dynamic in vivo imaging". Proceedings of the National Academy of Sciences. 104 (43): 16793–7. Bibcode:2007PNAS..10416793B. doi:10.1073/pnas.0707090104. PMC 2040404. PMID 17942682.
  9. ^ a b Ning, Xinghai; Temming, Rinske P.; Dommerholt, Jan; Guo, Jun; Blanco-Ania, Daniel; Debets, Marjoke F.; Wolfert, Margreet A.; Boons, Geert-Jan; Van Delft, Floris L. (2010). "Protein Modification by Strain-Promoted Alkyne-Nitrone Cycloaddition". Angewandte Chemie International Edition. 49 (17): 3065–8. doi:10.1002/anie.201000408. PMC 2871956. PMID 20333639.
  10. ^ Yarema, K. J.; Mahal, LK; Bruehl, RE; Rodriguez, EC; Bertozzi, CR (1998). "Metabolic Delivery of Ketone Groups to Sialic Acid Residues. Application to Cell Surface Glycoform Engineering". Journal of Biological Chemistry. 273 (47): 31168–79. doi:10.1074/jbc.273.47.31168. PMID 9813021.
  11. ^ a b Blackman, Melissa L.; Royzen, Maksim; Fox, Joseph M. (2008). "The Tetrazine Ligation: Fast Bioconjugation based on Inverse-electron-demand Diels-Alder Reactivity". Journal of the American Chemical Society. 130 (41): 13518–9. doi:10.1021/ja8053805. PMC 2653060. PMID 18798613.
  12. ^ a b c Stöckmann, Henning; Neves, André A.; Stairs, Shaun; Brindle, Kevin M.; Leeper, Finian J. (2011). "Exploring isonitrile-based click chemistry for ligation with biomolecules". Organic & Biomolecular Chemistry. 9 (21): 7303–5. doi:10.1039/C1OB06424J. PMID 21915395.
  13. ^ a b Sletten, Ellen M.; Bertozzi, Carolyn R. (2011). "A Bioorthogonal Quadricyclane Ligation". Journal of the American Chemical Society. 133 (44): 17570–3. doi:10.1021/ja2072934. PMC 3206493. PMID 21962173.
  14. ^ "The Nobel Prize in Chemistry". The Nobel Prize. Retrieved 6 October 2022.
  15. ^ Saxon, E.; Bertozzi, CR (2000). "Cell Surface Engineering by a Modified Staudinger Reaction". Science. 287 (5460): 2007–10. Bibcode:2000Sci...287.2007S. doi:10.1126/science.287.5460.2007. PMID 10720325. S2CID 19720277.
  16. ^ Pamela, Chang.; Prescher, Jennifer A.; Hangauer, Matthew J.; Bertozzi, Carolyn R. (2008). "Imaging Cell Surface Glycans with Bioorthogonal Chemical Reporters". J Am Chem Soc. 129 (27): 8400–8401. doi:10.1021/ja070238o. PMC 2535820. PMID 17579403.
  17. ^ Kennedy, David C.; McKay, Craig S.; Legault, Marc C. B.; Danielson, Dana C.; Blake, Jessie A.; Pegoraro, Adrian F.; Stolow, Albert; Mester, Zoltan; Pezacki, John Paul (2011). "Cellular Consequences of Copper Complexes Used to Catalyze Bioorthogonal Click Reactions". Journal of the American Chemical Society. 133 (44): 17993–8001. doi:10.1021/ja2083027. PMID 21970470.
  18. ^ Huisgen, Rolf. (1976). "1,3-Dipolar cycloadditions. 76. Concerted nature of 1,3-dipolar cycloadditions and the question of diradical intermediates". The Journal of Organic Chemistry. 41 (3): 403–419. doi:10.1021/jo00865a001.
  19. ^ a b Gold, Brian; Shevchenko, Nikolay E.; Bonus, Natalie; Dudley, Gregory B.; Alabugin, Igor V. (2011). "Selective Transition State Stabilization via Hyperconjugative and Conjugative Assistance: Stereoelectronic Concept for Copper-Free Click Chemistry". The Journal of Organic Chemistry. 77 (1): 75–89. doi:10.1021/jo201434w. PMID 22077877.
  20. ^ Chenoweth, Kimberly; Chenoweth, David; Goddard Iii, William A. (2009). "Cyclooctyne-based reagents for uncatalyzed click chemistry: A computational survey" (PDF). Organic & Biomolecular Chemistry. 7 (24): 5255–8. doi:10.1039/B911482C. PMID 20024122.
  21. ^ Jewett, John C.; Sletten, Ellen M.; Bertozzi, Carolyn R. (2010). "Rapid Cu-Free Click Chemistry with Readily Synthesized Biarylazacyclooctynones". Journal of the American Chemical Society. 132 (11): 3688–90. doi:10.1021/ja100014q. PMC 2840677. PMID 20187640.
  22. ^ Sletten, Ellen M.; Bertozzi, Carolyn R. (2008). "A Hydrophilic Azacyclooctyne for Cu-Free Click Chemistry". Organic Letters. 10 (14): 3097–9. doi:10.1021/ol801141k. PMC 2664610. PMID 18549231.
  23. ^ Ess, Daniel H.; Jones, Gavin O.; Houk, K. N. (2008). "Transition States of Strain-Promoted Metal-Free Click Chemistry: 1,3-Dipolar Cycloadditions of Phenyl Azide and Cyclooctynes". Organic Letters. 10 (8): 1633–6. doi:10.1021/ol8003657. PMID 18363405.
  24. ^ a b Schoenebeck, Franziska; Ess, Daniel H.; Jones, Gavin O.; Houk, K. N. (2009). "Reactivity and Regioselectivity in 1,3-Dipolar Cycloadditions of Azides to Strained Alkynes and Alkenes: A Computational Study". Journal of the American Chemical Society. 131 (23): 8121–33. doi:10.1021/ja9003624. PMID 19459632.
  25. ^ Lutz, Jean-François (2008). "Copper-Free Azide Alkyne Cycloadditions: New Insights and Perspectives". Angewandte Chemie International Edition. 47 (12): 2182–4. doi:10.1002/anie.200705365. PMID 18264961.
  26. ^ Dommerholt, Jan; Schmidt, Samuel; Temming, Rinske; Hendriks, Linda J. A.; Rutjes, Floris P. J. T.; Van Hest, Jan C. M.; Lefeber, Dirk J.; Friedl, Peter; Van Delft, Floris L. (2010). "Readily Accessible Bicyclononynes for Bioorthogonal Labeling and Three-Dimensional Imaging of Living Cells". Angewandte Chemie International Edition. 49 (49): 9422–5. doi:10.1002/anie.201003761. PMC 3021724. PMID 20857472.
  27. ^ Mbua, Ngalle Eric; Guo, Jun; Wolfert, Margreet A.; Steet, Richard; Boons, Geert-Jan (2011). "Strain-Promoted Alkyne-Azide Cycloadditions (SPAAC) Reveal New Features of Glycoconjugate Biosynthesis". ChemBioChem. 12 (12): 1912–21. doi:10.1002/cbic.201100117. PMC 3151320. PMID 21661087.
  28. ^ Jewett, John C.; Bertozzi, Carolyn R. (2011). "Synthesis of a fluorogenic cyclooctyne activated by Cu-free click chemistry". Organic Letters. 13 (22): 5937–9. doi:10.1021/ol2025026. PMC 3219546. PMID 22029411.
  29. ^ Poloukhtine, Andrei A.; Mbua, Ngalle Eric; Wolfert, Margreet A.; Boons, Geert-Jan; Popik, Vladimir V. (2009). "Selective Labeling of Living Cells by a Photo-Triggered Click Reaction". Journal of the American Chemical Society. 131 (43): 15769–76. doi:10.1021/ja9054096. PMC 2776736. PMID 19860481.
  30. ^ Carpenter, Richard D.; Hausner, Sven H.; Sutcliffe, Julie L. (2011). "Copper-Free Click for PET: Rapid 1,3-Dipolar Cycloadditions with a Fluorine-18 Cyclooctyne". ACS Medicinal Chemistry Letters. 2 (12): 885–889. doi:10.1021/ml200187j. PMC 4018166. PMID 24900276.
  31. ^ Gutsmiedl, Katrin; Wirges, Christian T.; Ehmke, Veronika; Carell, Thomas (2009). "Copper-Free "Click" Modification of DNA via Nitrile Oxide Norbornene 1,3-Dipolar Cycloaddition". Organic Letters. 11 (11): 2405–8. doi:10.1021/ol9005322. PMID 19405510.
  32. ^ Truong, Vinh X.; Zhou, Kun; Simon P., George; Forsythe, John S. (2015). "Nitrile Oxide-Norbornene Cycloaddition as a Bioorthogonal Crosslinking Reaction for the Preparation of Hydrogels". Macromolecular Rapid Communications. 36 (19): 1729–34. doi:10.1002/marc.201500314. PMID 26250120.
  33. ^ Van Berkel, Sander S.; Dirks, A. (Ton) J.; Debets, Marjoke F.; Van Delft, Floris L.; Cornelissen, Jeroen J. L. M.; Nolte, Roeland J. M.; Rutjes, Floris P. J. T. (2007). "Metal-Free Triazole Formation as a Tool for Bioconjugation". ChemBioChem. 8 (13): 1504–8. doi:10.1002/cbic.200700278. hdl:2066/34475. PMID 17631666. S2CID 45813826.
  34. ^ Van Berkel, Sander S.; Dirks, A. (Ton) J.; Meeuwissen, Silvie A.; Pingen, Dennis L. L.; Boerman, Otto C.; Laverman, Peter; Van Delft, Floris L.; Cornelissen, Jeroen J. L. M.; Rutjes, Floris P. J. T. (2008). "Application of Metal Free Triazole Formation in the Synthesis of Cyclic RGD DTPA Conjugates". ChemBioChem. 9 (11): 1805–15. doi:10.1002/cbic.200800074. hdl:2066/69881. PMID 18623291. S2CID 205552916.
  35. ^ Henry, Lucas; Schneider, Christopher; Mützel, Benedict; Simpson, Peter V.; Nagel, Christoph; Fucke, Katharina; Schatzschneider, Ulrich (2014). "Amino acid bioconjugation via iClick reaction of an oxanorbornadiene-masked alkyne with a MnI(bpy)(CO)3-coordinated azide" (PDF). ChemComm. 50 (99): 15692–95. doi:10.1039/C4CC07892F. PMID 25370120. S2CID 24060126.
  36. ^ Row, R. David; Prescher, Jennifer A. (2016). "Tetrazine Marks the Spot". ACS Central Science. 2 (8): 493–494. doi:10.1021/acscentsci.6b00204. PMC 4999966. PMID 27610408.
  37. ^ Bach, Robert D. (2009). "Ring Strain Energy in the Cyclooctyl System. The Effect of Strain Energy on [3 + 2] Cycloaddition Reactions with Azides". Journal of the American Chemical Society. 131 (14): 5233–43. doi:10.1021/ja8094137. PMID 19301865.
  38. ^ Devaraj, Neal K.; Weissleder, Ralph; Hilderbrand, Scott A. (2008). "Tetrazine-Based Cycloadditions: Application to Pretargeted Live Cell Imaging". Bioconjugate Chemistry. 19 (12): 2297–9. doi:10.1021/bc8004446. PMC 2677645. PMID 19053305.
  39. ^ Hansell, Claire F.; Espeel, Pieter; Stamenovic, Milan M.; Barker, Ian A.; Dove, Andrew P.; Du Prez, Filip E.; o Reilly, Rachel K. (2011). "Additive-Free Clicking for Polymer Functionalization and Coupling by Tetrazine Norbornene Chemistry". Journal of the American Chemical Society. 133 (35): 13828–31. doi:10.1021/ja203957h. PMID 21819063.
  40. ^ Lim, Reyna K. V.; Lin, Qing (2011). "Photoinducible Bioorthogonal Chemistry: A Spatiotemporally Controllable Tool to Visualize and Perturb Proteins in Live Cells". Accounts of Chemical Research. 44 (9): 828–839. doi:10.1021/ar200021p. PMC 3175026. PMID 21609129.
  41. ^ Song, Wenjiao; Wang, Yizhong; Qu, Jun; Lin, Qing (2008). "Selective Functionalization of a Genetically Encoded Alkene-Containing Protein via "Photoclick Chemistry" in Bacterial Cells". Journal of the American Chemical Society. 130 (30): 9654–5. doi:10.1021/ja803598e. PMID 18593155.
  42. ^ Knight, James C.; Cornelissen, Bart (2014). "Bioorthogonal chemistry: implications for pretargeted nuclear (PET/SPECT) imaging and therapy". American Journal of Nuclear Medicine and Molecular Imaging. 4 (2): 96–113. ISSN 2160-8407. PMC 3992206. PMID 24753979.

bioorthogonal, chemistry, term, bioorthogonal, chemistry, refers, chemical, reaction, that, occur, inside, living, systems, without, interfering, with, native, biochemical, processes, term, coined, carolyn, bertozzi, 2003, since, introduction, concept, bioorth. The term bioorthogonal chemistry refers to any chemical reaction that can occur inside of living systems without interfering with native biochemical processes 1 2 3 The term was coined by Carolyn R Bertozzi in 2003 4 5 Since its introduction the concept of the bioorthogonal reaction has enabled the study of biomolecules such as glycans proteins 6 and lipids 7 in real time in living systems without cellular toxicity A number of chemical ligation strategies have been developed that fulfill the requirements of bioorthogonality including the 1 3 dipolar cycloaddition between azides and cyclooctynes also termed copper free click chemistry 8 between nitrones and cyclooctynes 9 oxime hydrazone formation from aldehydes and ketones 10 the tetrazine ligation 11 the isocyanide based click reaction 12 and most recently the quadricyclane ligation 13 Shown here is a bioorthogonal ligation between biomolecule X and reactive partner Y To be considered bioorthogonal these reactive partners cannot perturb other chemical functionality naturally found within the cell The use of bioorthogonal chemistry typically proceeds in two steps First a cellular substrate is modified with a bioorthogonal functional group chemical reporter and introduced to the cell substrates include metabolites enzyme inhibitors etc The chemical reporter must not alter the structure of the substrate dramatically to avoid affecting its bioactivity Secondly a probe containing the complementary functional group is introduced to react and label the substrate Although effective bioorthogonal reactions such as copper free click chemistry have been developed development of new reactions continues to generate orthogonal methods for labeling to allow multiple methods of labeling to be used in the same biosystems Carolyn R Bertozzi was awarded the Nobel Prize in Chemistry in 2022 for her development of click chemistry and bioorthogonal chemistry 14 Contents 1 Etymology 2 Requirements for bioorthogonality 3 Staudinger ligation 3 1 Bioorthogonality 3 2 Mechanism 3 2 1 Classic Staudinger reaction 3 2 2 Staudinger ligation 3 3 Limitations 4 Copper free click chemistry 4 1 Copper toxicity 4 2 Bioorthogonality 4 3 Mechanism 4 4 Regioselectivity 4 5 Development of cyclooctynes 4 6 Reactivity 4 7 Regioselectivity 4 8 Applications 5 Other bioorthogonal reactions 5 1 Nitrone dipole cycloaddition 5 2 Norbornene cycloaddition 5 3 Oxanorbornadiene cycloaddition 5 4 Tetrazine ligation 5 5 4 1 Cycloaddition 5 6 Tetrazole photoclick chemistry 5 7 Quadricyclane ligation 6 Uses 7 ReferencesEtymology editThe word bioorthogonal comes from Greek bio living and orthogōnios right angled Thus literally a reaction that goes perpendicular to a living system thus not disturbing it Requirements for bioorthogonality editTo be considered bioorthogonal a reaction must fulfill a number of requirements Selectivity The reaction must be selective between endogenous functional groups to avoid side reactions with biological compounds Biological inertness Reactive partners and resulting linkage should not possess any mode of reactivity capable of disrupting the native chemical functionality of the organism under study Chemical inertness The covalent link should be strong and inert to biological reactions Kinetics The reaction must be rapid so that covalent ligation is achieved prior to probe metabolism and clearance The reaction must be fast on the time scale of cellular processes minutes to prevent competition in reactions which may diminish the small signals of less abundant species Rapid reactions also offer a fast response necessary in order to accurately track dynamic processes Reaction biocompatibility Reactions have to be non toxic and must function in biological conditions taking into account pH aqueous environments and temperature Pharmacokinetics are a growing concern as bioorthogonal chemistry expands to live animal models Accessible engineering The chemical reporter must be capable of incorporation into biomolecules via some form of metabolic or protein engineering Optimally one of the functional groups is also very small so that it does not disturb native behavior Staudinger ligation editThe Staudinger ligation is a reaction developed by the Bertozzi group in 2000 that is based on the classic Staudinger reaction of azides with triarylphosphines 15 It launched the field of bioorthogonal chemistry as the first reaction with completely abiotic functional groups although it is no longer as widely used The Staudinger ligation has been used in both live cells and live mice 5 Bioorthogonality edit The azide can act as a soft electrophile that prefers soft nucleophiles such as phosphines This is in contrast to most biological nucleophiles which are typically hard nucleophiles The reaction proceeds selectively under water tolerant conditions to produce a stable product Phosphines are completely absent from living systems and do not reduce disulfide bonds despite mild reduction potential Azides had been shown to be biocompatible in FDA approved drugs such as azidothymidine and through other uses as cross linkers Additionally their small size allows them to be easily incorporated into biomolecules through cellular metabolic pathways Mechanism edit This section may contain material not related to the topic of the article Please help improve this section or discuss this issue on the talk page September 2014 Learn how and when to remove this message Classic Staudinger reaction edit nbsp The nucleophilic phosphine attacks the azide at the electrophilic terminal nitrogen Through a four membered transition state N2 is lost to form an aza ylide The unstable ylide is hydrolyzed to form phosphine oxide and a primary amine However this reaction is not immediately bioorthogonal because hydrolysis breaks the covalent bond in the aza ylide Staudinger ligation edit nbsp The reaction was modified to include an ester group ortho to the phosphorus atom on one of the aryl rings to direct the aza ylide through a new path of reactivity in order to outcompete immediate hydrolysis by positioning the ester to increase local concentration The initial nucleophilic attack on the azide is the rate limiting step The ylide reacts with the electrophilic ester trap through intramolecular cyclization to form a five membered ring This ring undergoes hydrolysis to form a stable amide bond Limitations edit The phosphine reagents slowly undergo air oxidation in living systems Additionally it is likely that they are metabolized in vitro by cytochrome P450 enzymes The kinetics of the reactions are slow with second order rate constants around 0 0020 M 1 s 1 Attempts to increase nucleophilic attack rates by adding electron donating groups to the phosphines improved kinetics but also increased the rate of air oxidation The poor kinetics require that high concentrations of the phosphine be used which leads to problems with high background signal in imaging applications Attempts have been made to combat the problem of high background through the development of a fluorogenic phosphine reagents based on fluorescein and luciferin but the intrinsic kinetics remain a limitation 16 Copper free click chemistry editMain article Copper free click chemistry Copper free click chemistry is a bioorthogonal reaction first developed by Carolyn Bertozzi as an activated variant of an azide alkyne Huisgen cycloaddition based on the work by Karl Barry Sharpless et al Unlike CuAAC Cu free click chemistry has been modified to be bioorthogonal by eliminating a cytotoxic copper catalyst allowing reaction to proceed quickly and without live cell toxicity Instead of copper the reaction is a strain promoted alkyne azide cycloaddition SPAAC It was developed as a faster alternative to the Staudinger ligation with the first generations reacting over sixty times faster The bioorthogonality of the reaction has allowed the Cu free click reaction to be applied within cultured cells live zebrafish and mice nbsp click chemistry labeling Copper toxicity edit The classic copper catalyzed azide alkyne cycloaddition has been an extremely fast and effective click reaction for bioconjugation but it is not suitable for use in live cells due to the toxicity of Cu I ions Toxicity is due to oxidative damage from reactive oxygen species formed by the copper catalysts Copper complexes have also been found to induce changes in cellular metabolism and are taken up by cells There has been some development of ligands to prevent biomolecule damage and facilitate removal in in vitro applications However it has been found that different ligand environments of complexes can still affect metabolism and uptake introducing an unwelcome perturbation in cellular function 17 Bioorthogonality edit The azide group is particularly bioorthogonal because it is extremely small favorable for cell permeability and avoids perturbations metabolically stable and does not naturally exist in cells and thus has no competing biological side reactions Although azides are not the most reactive 1 3 dipole available for reaction they are preferred for their relative lack of side reactions and stability in typical synthetic conditions 18 The alkyne is not as small but it still has the stability and orthogonality necessary for in vivo labeling Cyclooctynes are traditionally the most common cycloalkyne for labeling studies as they are the smallest stable alkyne ring Mechanism edit nbsp The reaction proceeds as a standard 1 3 dipolar cycloaddition a type of asynchronous concerted pericyclic shift The ambivalent nature of the 1 3 dipole should make the identification of an electrophilic or nucleophilic center on the azide impossible such that the direction of the cyclic electron flow is meaningless p However computation has shown that the electron distribution amongst nitrogens causes the innermost nitrogen atom to bear the greatest negative charge 19 Regioselectivity edit Although the reaction produces a regioisomeric mixture of triazoles the lack of regioselectivity in the reaction is not a major concern for most current applications More regiospecific and less bioorthogonal requirements are best served by copper catalyzed Huisgen cycloaddition especially given the synthetic difficulty compared to the addition of a terminal alkyne of synthesizing a strained cyclooctyne Development of cyclooctynes edit Cyclooctyne Second order rate constant M 1s 1 OCT 0 0024 ALO 0 0013 MOFO 0 0043 DIFO 0 076 DIBO 0 057 BARAC 0 96 DIBAC ADIBO 0 31 DIMAC 0 0030 nbsp Strained cyclooctynes developed for copper free click chemistry OCT was the first cyclooctyne developed for Cu free click chemistry While linear alkynes are unreactive at physiological temperatures OCT was able readily react with azides in biological conditions while showing no toxicity However it was poorly water soluble and the kinetics were barely improved over the Staudinger ligation ALO aryl less octyne was developed to improve water solubility but it still had poor kinetics Monofluorinated MOFO and difluorinated DIFO cyclooctynes were created to increase the rate through the addition of electron withdrawing fluorine substituents at the propargylic position Fluorine is a good electron withdrawing group in terms of synthetic accessibility and biological inertness In particular it cannot form an electrophilic Michael acceptor that may side react with biological nucleophiles 8 DIBO dibenzocyclooctyne was developed as a fusion to two aryl rings resulting in very high strain and a decrease in distortion energies It was proposed that biaryl substitution increases ring strain and provides conjugation with the alkyne to improve reactivity Although calculations have predicted that mono aryl substitution would provide an optimal balance between steric clash with azide molecule and strain 20 monoarylated products have been shown to be unstable BARAC biarylazacyclooctynone followed with the addition of an amide bond which adds an sp2 like center to increase rate by distortion Amide resonance contributes additional strain without creating additional unsaturation which would lead to an unstable molecule Additionally the addition of a heteroatom into the cyclooctyne ring improves both solubility and pharmacokinetics of the molecule BARAC has sufficient rate and sensitivity to the extent that washing away excess probe is unnecessary to reduce background This makes it extremely useful in situations where washing is impossible as in real time imaging or whole animal imaging Although BARAC is extremely useful its low stability requires that it must be stored at 0 C protected from light and oxygen 21 nbsp The synthesis was designed by the Bertozzi group as a modular route to facilitate future modifications in SAR analysis The first step is Fischer indole synthesis The product is alkylated with allyl bromide as a handle for future probe attachment TMS is then added Oxidation opens the central rings to form a cyclic amide The ketone is treated as an enolate to add a triflate group Reaction of the terminal alkene generates a linker for conjugation to a molecule The final reaction with CsF introduces the strained alkyne at the last step Further adjustments variations on BARAC to produce DIBAC ADIBO were performed to add distal ring strain and reduce sterics around the alkyne to further increase reactivity Keto DIBO in which the hydroxyl group has been converted to a ketone has a three fold increase in rate due to a change in ring conformation Attempts to make a difluorobenzocyclooctyne DIFBO were unsuccessful due to the instability Problems with DIFO with in vivo mouse studies illustrate the difficulty of producing bioorthogonal reactions Although DIFO was extremely reactive in the labeling of cells it performed poorly in mouse studies due to binding with serum albumin Hydrophobicity of the cyclooctyne promotes sequestration by membranes and serum proteins reducing bioavailable concentrations In response DIMAC dimethoxyazacyclooctyne was developed to increase water solubility polarity and pharmacokinetics 22 although efforts in bioorthogonal labeling of mouse models is still in development Reactivity edit Computational efforts have been vital in explaining the thermodynamics and kinetics of these cycloaddition reactions which has played a vital role in continuing to improve the reaction There are two methods for activating alkynes without sacrificing stability decrease transition state energy or decrease reactant stability nbsp The red arrow shows the direction of energy change Black arrows show the difference in activation energy before and after the effects Decreasing reactant stability Houk 23 has proposed that differences in the energy Ed required to distort the azide and alkyne into the transition state geometries control the barrier heights for the reaction The activation energy E is the sum of destabilizing distortions and stabilizing interactions Ei The most significant distortion is in the azide functional group with lesser contribution of alkyne distortion However it is only the cyclooctyne that can be easily modified for higher reactivity Calculated barriers of reaction for phenyl azide and acetylene 16 2 kcal mol versus cyclooctyne 8 0 kcal mol results in a predicted rate increase of 106 The cyclooctyne requires less distortion energy 1 4 kcal mol versus 4 6 kcal mol resulting in a lower activation energy despite smaller interaction energy nbsp Relationship between activation energy distortion energy and interaction energy Decreasing transition state energy Electron withdrawing groups such as fluorine increase rate by decreasing LUMO energy and the HOMO LUMO gap This leads to a greater charge transfer from the azide to the fluorinated cyclooctyne in the transition state increasing interaction energy lower negative value and overall activation energy 24 The lowering of the LUMO is the result of hyperconjugation between alkyne p donor orbitals and CF s acceptors These interactions provide stabilization primarily in the transition state as a result of increased donor acceptor abilities of the bonds as they distort NBO calculations have shown that transition state distortion increases the interaction energy by 2 8 kcal mol The hyperconjugation between out of plane p bonds is greater because the in plane p bonds are poorly aligned However transition state bending allows the in plane p bonds to have a more antiperiplanar arrangement that facilitates interaction Additional hyperconjugative interaction energy stabilization is achieved through an increase in the electronic population of the s due to the forming CN bond Negative hyperconjugation with the s CF bonds enhances this stabilizing interaction 19 Regioselectivity edit Although regioselectivity is not a great issue in the current imaging applications of copper free click chemistry it is an issue that prevents future applications in fields such as drug design or peptidomimetics 25 Currently most cyclooctynes react to form regioisomeric mixtures m Computation analysis has found that while gas phase regioselectivity is calculated to favor 1 5 addition over 1 4 addition by up to 2 9 kcal mol in activation energy solvation corrections result in the same energy barriers for both regioisomers While the 1 4 isomer in the cycloaddition of DIFO is disfavored by its larger dipole moment solvation stabilizes it more strongly than the 1 5 isomer eroding regioselectivity 24 nbsp Symmetrical cyclooctynes such as BCN bicyclo 6 1 0 nonyne form a single regioisomer upon cycloaddition 26 and may serve to address this problem in the future Applications edit The most widespread application of copper free click chemistry is in biological imaging in live cells or animals using an azide tagged biomolecule and a cyclooctyne bearing an imaging agent Fluorescent keto and oxime variants of DIBO are used in fluoro switch click reactions in which the fluorescence of the cyclooctyne is quenched by the triazole that forms in the reaction 27 On the other hand coumarin conjugated cyclooctynes such as coumBARAC have been developed such that the alkyne suppresses fluorescence while triazole formation increases the fluorescence quantum yield by ten fold 28 nbsp coumBARAC fluorescence increases with reaction Spatial and temporal control of substrate labeling has been investigated using photoactivatable cyclooctynes This allows equilibration of the alkyne prior to reaction in order to reduce artifacts as a result of concentration gradients Masked cyclooctynes are unable to react with azides in the dark but become reactive alkynes upon irradiation with light 29 nbsp Copper free click chemistry is being explored for use in synthesizing PET imaging agents which must be made quickly with high purity and yield in order to minimize isotopic decay before the compounds can be administered Both the high rate constants and the bioorthogonality of SPAAC are amenable to PET chemistry 30 Other bioorthogonal reactions editNitrone dipole cycloaddition edit Copper free click chemistry has been adapted to use nitrones as the 1 3 dipole rather than azides and has been used in the modification of peptides 9 nbsp This cycloaddition between a nitrone and a cyclooctyne forms N alkylated isoxazolines The reaction rate is enhanced by water and is extremely fast with second order rate constants ranging from 12 to 32 M 1 s 1 depending on the substitution of the nitrone Although the reaction is extremely fast it faces problems in incorporating the nitrone into biomolecules through metabolic labeling Labeling has only been achieved through post translational peptide modification Norbornene cycloaddition edit 1 3 dipolar cycloadditions have been developed as a bioorthogonal reaction using a nitrile oxide as a 1 3 dipole and a norbornene as a dipolarophile Its primary use has been in labeling DNA and RNA in automated oligonucleotide synthesizers 31 and polymer crosslinking in the presence of living cells 32 nbsp Norbornenes were selected as dipolarophiles due to their balance between strain promoted reactivity and stability The drawbacks of this reaction include the cross reactivity of the nitrile oxide due to strong electrophilicity and slow reaction kinetics Oxanorbornadiene cycloaddition edit The oxanorbornadiene cycloaddition is a 1 3 dipolar cycloaddition followed by a retro Diels Alder reaction to generate a triazole linked conjugate with the elimination of a furan molecule 33 Preliminary work has established its usefulness in peptide labeling experiments and it has also been used in the generation of SPECT imaging compounds 34 More recently the use of an oxanorbornadiene was described in a catalyst free room temperature iClick reaction in which a model amino acid is linked to the metal moiety in a novel approach to bioorthogonal reactions 35 nbsp Ring strain and electron deficiency in the oxanorbornadiene increase reactivity towards the cycloaddition rate limiting step The retro Diels Alder reaction occurs quickly afterwards to form the stable 1 2 3 triazole Problems include poor tolerance for substituents which may change electronics of the oxanorbornadiene and low rates second order rate constants on the order of 10 4 Tetrazine ligation edit The tetrazine ligation is the reaction of a trans cyclooctene and an s tetrazine in an inverse demand Diels Alder reaction followed by a retro Diels Alder reaction to eliminate nitrogen gas 36 The reaction is extremely rapid with a second order rate constant of 2000 M 1 s 1 in 9 1 methanol water allowing modifications of biomolecules at extremely low concentrations nbsp Based on computational work by Bach the strain energy for Z cyclooctenes is 7 0 kcal mol compared to 12 4 kcal mol for cyclooctane due to a loss of two transannular interactions E cyclooctene has a highly twisted double bond resulting in a strain energy of 17 9 kcal mol 37 As such the highly strained trans cyclooctene is used as a reactive dienophile The diene is a 3 6 diaryl s tetrazine which has been substituted in order to resist immediate reaction with water The reaction proceeds through an initial cycloaddition followed by a reverse Diels Alder to eliminate N2 and prevent reversibility of the reaction 11 Not only is the reaction tolerant of water but it has been found that the rate increases in aqueous media Reactions have also been performed using norbornenes as dienophiles at second order rates on the order of 1 M 1 s 1 in aqueous media The reaction has been applied in labeling live cells 38 and polymer coupling 39 4 1 Cycloaddition edit This isocyanide click reaction is a 4 1 cycloaddition followed by a retro Diels Alder elimination of N2 12 nbsp The reaction proceeds with an initial 4 1 cycloaddition followed by a reversion to eliminate a thermodynamic sink and prevent reversibility This product is stable if a tertiary amine or isocyanopropanoate is used If a secondary or primary isocyanide is used the produce will form an imine which is quickly hydrolyzed Isocyanide is a favored chemical reporter due to its small size stability non toxicity and absence in mammalian systems However the reaction is slow with second order rate constants on the order of 10 2 M 1 s 1 Tetrazole photoclick chemistry edit Photoclick chemistry utilizes a photoinduced cycloelimination to release N2 This generates a short lived 1 3 nitrile imine intermediate via the loss of nitrogen gas which undergoes a 1 3 dipolar cycloaddition with an alkene to generate pyrazoline cycloadducts 12 nbsp Photoinduction takes place with a brief exposure to light wavelength is tetrazole dependent to minimize photodamage to cells The reaction is enhanced in aqueous conditions and generates a single regioisomer The transient nitrile imine is highly reactive for 1 3 dipolar cycloaddition due to a bent structure which reduces distortion energy Substitution with electron donating groups on phenyl rings increases the HOMO energy when placed on the 1 3 nitrile imine and increases the rate of reaction Advantages of this approach include the ability to spatially or temporally control reaction and the ability to incorporate both alkenes and tetrazoles into biomolecules using simple biological methods such as genetic encoding 40 Additionally the tetrazole can be designed to be fluorogenic in order to monitor progress of the reaction 41 Quadricyclane ligation edit The quadricyclane ligation utilizes a highly strained quadricyclane to undergo 2 2 2 cycloaddition with p systems 13 nbsp Quadricyclane is abiotic unreactive with biomolecules due to complete saturation relatively small and highly strained 80 kcal mol However it is highly stable at room temperature and in aqueous conditions at physiological pH It is selectively able to react with electron poor p systems but not simple alkenes alkynes or cyclooctynes Bis dithiobenzil nickel II was chosen as a reaction partner out of a candidate screen based on reactivity To prevent light induced reversion to norbornadiene diethyldithiocarbamate is added to chelate the nickel in the product nbsp These reactions are enhanced by aqueous conditions with a second order rate constant of 0 25 M 1 s 1 Of particular interest is that it has been proven to be bioorthogonal to both oxime formation and copper free click chemistry Uses editBioorthogonal chemistry is an attractive tool for pretargeting experiments in nuclear imaging and radiotherapy 42 References edit Sletten Ellen M Bertozzi Carolyn R 2009 Bioorthogonal Chemistry Fishing for Selectivity in a Sea of Functionality Angewandte Chemie International Edition 48 38 6974 98 doi 10 1002 anie 200900942 PMC 2864149 PMID 19714693 Prescher Jennifer A Dube Danielle H Bertozzi Carolyn R 2004 Chemical remodelling of cell surfaces in living animals Nature 430 7002 873 7 Bibcode 2004Natur 430 873P doi 10 1038 nature02791 PMID 15318217 S2CID 4371934 Prescher Jennifer A Bertozzi Carolyn R 2005 Chemistry in living systems Nature Chemical Biology 1 1 13 21 doi 10 1038 nchembio0605 13 PMID 16407987 S2CID 40548615 Hang Howard C Yu Chong Kato Darryl L Bertozzi Carolyn R 2003 12 09 A metabolic labeling approach toward proteomic analysis of mucin type O linked glycosylation Proceedings of the National Academy of Sciences 100 25 14846 14851 Bibcode 2003PNAS 10014846H doi 10 1073 pnas 2335201100 ISSN 0027 8424 PMC 299823 PMID 14657396 a b Sletten Ellen M Bertozzi Carolyn R 2011 From Mechanism to Mouse A Tale of Two Bioorthogonal Reactions Accounts of Chemical Research 44 9 666 676 doi 10 1021 ar200148z PMC 3184615 PMID 21838330 Plass Tilman Milles Sigrid Koehler Christine Schultz Carsten Lemke Edward A 2011 Genetically Encoded Copper Free Click Chemistry Angewandte Chemie International Edition 50 17 3878 3881 doi 10 1002 anie 201008178 PMC 3210829 PMID 21433234 Neef Anne B Schultz Carsten 2009 Selective Fluorescence Labeling of Lipids in Living Cells Angewandte Chemie International Edition 48 8 1498 500 doi 10 1002 anie 200805507 PMID 19145623 a b Baskin J M Prescher J A Laughlin S T Agard N J Chang P V Miller I A Lo A Codelli J A Bertozzi C R 2007 Copper free click chemistry for dynamic in vivo imaging Proceedings of the National Academy of Sciences 104 43 16793 7 Bibcode 2007PNAS 10416793B doi 10 1073 pnas 0707090104 PMC 2040404 PMID 17942682 a b Ning Xinghai Temming Rinske P Dommerholt Jan Guo Jun Blanco Ania Daniel Debets Marjoke F Wolfert Margreet A Boons Geert Jan Van Delft Floris L 2010 Protein Modification by Strain Promoted Alkyne Nitrone Cycloaddition Angewandte Chemie International Edition 49 17 3065 8 doi 10 1002 anie 201000408 PMC 2871956 PMID 20333639 Yarema K J Mahal LK Bruehl RE Rodriguez EC Bertozzi CR 1998 Metabolic Delivery of Ketone Groups to Sialic Acid Residues Application to Cell Surface Glycoform Engineering Journal of Biological Chemistry 273 47 31168 79 doi 10 1074 jbc 273 47 31168 PMID 9813021 a b Blackman Melissa L Royzen Maksim Fox Joseph M 2008 The Tetrazine Ligation Fast Bioconjugation based on Inverse electron demand Diels Alder Reactivity Journal of the American Chemical Society 130 41 13518 9 doi 10 1021 ja8053805 PMC 2653060 PMID 18798613 a b c Stockmann Henning Neves Andre A Stairs Shaun Brindle Kevin M Leeper Finian J 2011 Exploring isonitrile based click chemistry for ligation with biomolecules Organic amp Biomolecular Chemistry 9 21 7303 5 doi 10 1039 C1OB06424J PMID 21915395 a b Sletten Ellen M Bertozzi Carolyn R 2011 A Bioorthogonal Quadricyclane Ligation Journal of the American Chemical Society 133 44 17570 3 doi 10 1021 ja2072934 PMC 3206493 PMID 21962173 The Nobel Prize in Chemistry The Nobel Prize Retrieved 6 October 2022 Saxon E Bertozzi CR 2000 Cell Surface Engineering by a Modified Staudinger Reaction Science 287 5460 2007 10 Bibcode 2000Sci 287 2007S doi 10 1126 science 287 5460 2007 PMID 10720325 S2CID 19720277 Pamela Chang Prescher Jennifer A Hangauer Matthew J Bertozzi Carolyn R 2008 Imaging Cell Surface Glycans with Bioorthogonal Chemical Reporters J Am Chem Soc 129 27 8400 8401 doi 10 1021 ja070238o PMC 2535820 PMID 17579403 Kennedy David C McKay Craig S Legault Marc C B Danielson Dana C Blake Jessie A Pegoraro Adrian F Stolow Albert Mester Zoltan Pezacki John Paul 2011 Cellular Consequences of Copper Complexes Used to Catalyze Bioorthogonal Click Reactions Journal of the American Chemical Society 133 44 17993 8001 doi 10 1021 ja2083027 PMID 21970470 Huisgen Rolf 1976 1 3 Dipolar cycloadditions 76 Concerted nature of 1 3 dipolar cycloadditions and the question of diradical intermediates The Journal of Organic Chemistry 41 3 403 419 doi 10 1021 jo00865a001 a b Gold Brian Shevchenko Nikolay E Bonus Natalie Dudley Gregory B Alabugin Igor V 2011 Selective Transition State Stabilization via Hyperconjugative and Conjugative Assistance Stereoelectronic Concept for Copper Free Click Chemistry The Journal of Organic Chemistry 77 1 75 89 doi 10 1021 jo201434w PMID 22077877 Chenoweth Kimberly Chenoweth David Goddard Iii William A 2009 Cyclooctyne based reagents for uncatalyzed click chemistry A computational survey PDF Organic amp Biomolecular Chemistry 7 24 5255 8 doi 10 1039 B911482C PMID 20024122 Jewett John C Sletten Ellen M Bertozzi Carolyn R 2010 Rapid Cu Free Click Chemistry with Readily Synthesized Biarylazacyclooctynones Journal of the American Chemical Society 132 11 3688 90 doi 10 1021 ja100014q PMC 2840677 PMID 20187640 Sletten Ellen M Bertozzi Carolyn R 2008 A Hydrophilic Azacyclooctyne for Cu Free Click Chemistry Organic Letters 10 14 3097 9 doi 10 1021 ol801141k PMC 2664610 PMID 18549231 Ess Daniel H Jones Gavin O Houk K N 2008 Transition States of Strain Promoted Metal Free Click Chemistry 1 3 Dipolar Cycloadditions of Phenyl Azide and Cyclooctynes Organic Letters 10 8 1633 6 doi 10 1021 ol8003657 PMID 18363405 a b Schoenebeck Franziska Ess Daniel H Jones Gavin O Houk K N 2009 Reactivity and Regioselectivity in 1 3 Dipolar Cycloadditions of Azides to Strained Alkynes and Alkenes A Computational Study Journal of the American Chemical Society 131 23 8121 33 doi 10 1021 ja9003624 PMID 19459632 Lutz Jean Francois 2008 Copper Free Azide Alkyne Cycloadditions New Insights and Perspectives Angewandte Chemie International Edition 47 12 2182 4 doi 10 1002 anie 200705365 PMID 18264961 Dommerholt Jan Schmidt Samuel Temming Rinske Hendriks Linda J A Rutjes Floris P J T Van Hest Jan C M Lefeber Dirk J Friedl Peter Van Delft Floris L 2010 Readily Accessible Bicyclononynes for Bioorthogonal Labeling and Three Dimensional Imaging of Living Cells Angewandte Chemie International Edition 49 49 9422 5 doi 10 1002 anie 201003761 PMC 3021724 PMID 20857472 Mbua Ngalle Eric Guo Jun Wolfert Margreet A Steet Richard Boons Geert Jan 2011 Strain Promoted Alkyne Azide Cycloadditions SPAAC Reveal New Features of Glycoconjugate Biosynthesis ChemBioChem 12 12 1912 21 doi 10 1002 cbic 201100117 PMC 3151320 PMID 21661087 Jewett John C Bertozzi Carolyn R 2011 Synthesis of a fluorogenic cyclooctyne activated by Cu free click chemistry Organic Letters 13 22 5937 9 doi 10 1021 ol2025026 PMC 3219546 PMID 22029411 Poloukhtine Andrei A Mbua Ngalle Eric Wolfert Margreet A Boons Geert Jan Popik Vladimir V 2009 Selective Labeling of Living Cells by a Photo Triggered Click Reaction Journal of the American Chemical Society 131 43 15769 76 doi 10 1021 ja9054096 PMC 2776736 PMID 19860481 Carpenter Richard D Hausner Sven H Sutcliffe Julie L 2011 Copper Free Click for PET Rapid 1 3 Dipolar Cycloadditions with a Fluorine 18 Cyclooctyne ACS Medicinal Chemistry Letters 2 12 885 889 doi 10 1021 ml200187j PMC 4018166 PMID 24900276 Gutsmiedl Katrin Wirges Christian T Ehmke Veronika Carell Thomas 2009 Copper Free Click Modification of DNA via Nitrile Oxide Norbornene 1 3 Dipolar Cycloaddition Organic Letters 11 11 2405 8 doi 10 1021 ol9005322 PMID 19405510 Truong Vinh X Zhou Kun Simon P George Forsythe John S 2015 Nitrile Oxide Norbornene Cycloaddition as a Bioorthogonal Crosslinking Reaction for the Preparation of Hydrogels Macromolecular Rapid Communications 36 19 1729 34 doi 10 1002 marc 201500314 PMID 26250120 Van Berkel Sander S Dirks A Ton J Debets Marjoke F Van Delft Floris L Cornelissen Jeroen J L M Nolte Roeland J M Rutjes Floris P J T 2007 Metal Free Triazole Formation as a Tool for Bioconjugation ChemBioChem 8 13 1504 8 doi 10 1002 cbic 200700278 hdl 2066 34475 PMID 17631666 S2CID 45813826 Van Berkel Sander S Dirks A Ton J Meeuwissen Silvie A Pingen Dennis L L Boerman Otto C Laverman Peter Van Delft Floris L Cornelissen Jeroen J L M Rutjes Floris P J T 2008 Application of Metal Free Triazole Formation in the Synthesis of Cyclic RGD DTPA Conjugates ChemBioChem 9 11 1805 15 doi 10 1002 cbic 200800074 hdl 2066 69881 PMID 18623291 S2CID 205552916 Henry Lucas Schneider Christopher Mutzel Benedict Simpson Peter V Nagel Christoph Fucke Katharina Schatzschneider Ulrich 2014 Amino acid bioconjugation via iClick reaction of an oxanorbornadiene masked alkyne with a MnI bpy CO 3 coordinated azide PDF ChemComm 50 99 15692 95 doi 10 1039 C4CC07892F PMID 25370120 S2CID 24060126 Row R David Prescher Jennifer A 2016 Tetrazine Marks the Spot ACS Central Science 2 8 493 494 doi 10 1021 acscentsci 6b00204 PMC 4999966 PMID 27610408 Bach Robert D 2009 Ring Strain Energy in the Cyclooctyl System The Effect of Strain Energy on 3 2 Cycloaddition Reactions with Azides Journal of the American Chemical Society 131 14 5233 43 doi 10 1021 ja8094137 PMID 19301865 Devaraj Neal K Weissleder Ralph Hilderbrand Scott A 2008 Tetrazine Based Cycloadditions Application to Pretargeted Live Cell Imaging Bioconjugate Chemistry 19 12 2297 9 doi 10 1021 bc8004446 PMC 2677645 PMID 19053305 Hansell Claire F Espeel Pieter Stamenovic Milan M Barker Ian A Dove Andrew P Du Prez Filip E o Reilly Rachel K 2011 Additive Free Clicking for Polymer Functionalization and Coupling by Tetrazine Norbornene Chemistry Journal of the American Chemical Society 133 35 13828 31 doi 10 1021 ja203957h PMID 21819063 Lim Reyna K V Lin Qing 2011 Photoinducible Bioorthogonal Chemistry A Spatiotemporally Controllable Tool to Visualize and Perturb Proteins in Live Cells Accounts of Chemical Research 44 9 828 839 doi 10 1021 ar200021p PMC 3175026 PMID 21609129 Song Wenjiao Wang Yizhong Qu Jun Lin Qing 2008 Selective Functionalization of a Genetically Encoded Alkene Containing Protein via Photoclick Chemistry in Bacterial Cells Journal of the American Chemical Society 130 30 9654 5 doi 10 1021 ja803598e PMID 18593155 Knight James C Cornelissen Bart 2014 Bioorthogonal chemistry implications for pretargeted nuclear PET SPECT imaging and therapy American Journal of Nuclear Medicine and Molecular Imaging 4 2 96 113 ISSN 2160 8407 PMC 3992206 PMID 24753979 Retrieved from https en wikipedia org w index php title Bioorthogonal chemistry amp oldid 1208848124, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.