Waiting
Login processing...

Trial ends in Request Full Access Tell Your Colleague About Jove

Chemistry

Solid-phase Synthesis of [4.4] Spirocyclic Oximes

Published: February 6, 2019 doi: 10.3791/58508

Summary

Here we present a protocol to demonstrate an efficient method for the synthesis of spirocyclic heterocycles. The five-step process utilizes solid-phase synthesis and regenerating Michael linker strategies. Generally difficult to synthesize, we present a customizable method for the synthesis of spirocyclic molecules otherwise inaccessible to other modern approaches.

Abstract

A convenient synthetic route for spirocyclic heterocycles is well sought after due to the molecule's potential use in biological systems. By means of solid-phase synthesis, regenerating Michael (REM) linker strategies, and 1,3-dipolar cycloaddition, a library of structurally similar heterocycles, both with and without a spirocyclic center, can be constructed. The main advantages of the solid-support synthesis are as follows: first, each reaction step can be driven to completion using a large excess of reagents resulting in high yields; next, the use of commercially available starting materials and reagents keep the costs low; finally, the reaction steps are easy to purify via simple filtration. The REM linker strategy is attractive because of its recyclability and traceless nature. Once a reaction scheme is completed, the linker can be reused multiple times. In a typical solid-phase synthesis, the product contains either a part of or the whole linker, which can prove undesirable. The REM linker is "traceless" and the point of attachment between the product and the polymer is indistinguishable. The high diastereoselectivity of the intramolecular 1,3-dipolar cycloaddition is well documented. Limited by the insolubility of the solid support, the reaction progression can only be monitored by a change in the functional groups (if any) via infrared (IR) spectroscopy. Thus, the structural identification of intermediates cannot be characterized by conventional nuclear magnetic resonance (NMR) spectroscopy. Other limitations to this method stem from the compatibilities of the polymer/linker to the desired chemical reaction scheme. Herein we report a protocol that allows for the convenient production of spirocyclic heterocycles that, with simple modifications, can be automated with high-throughput techniques.

Introduction

Despite recent discoveries using highly-functionalized spirocyclic heterocycles in a number of biological systems1, a convenient pathway is still necessary for their easy manufacture. Such systems and uses for these heterocycles include: MDM2 inhibition and other anticancer activities2,3,4,5, enzyme inhibition6,7,8, antibiotic activity9,10, fluorescent tagging10,11,12, enantioselective binding for DNA probes13,14,15 and RNA targeting16, along with numerous potential applications to therapeutics17,18,19. With an increasing demand for these heterocycles, current literature remains divided about which synthetic pathway is best. Modern synthetic approaches to this problem use isatin and isatin derivatives as starting materials for a variety of heterocycles20,21, complicated intramolecular rearrangements22,23,24,25, Lewis acid1,26,27 or transition metal catalysis17,28,29,30, or asymmetric processes31. While these procedures have had success in producing specific spirocyclic oximes with limited functionality, a synthetic strategy for producing a library of molecules with high diastereoselectivity has been explored relatively less32.

The technique presented here shows that these molecules of interest can be generated using a number of well-understood synthetic techniques in tandem. Starting with the synthesis of the molecule on a solid support using a REM linker and intramolecular silyl nitronate-olefin cycloaddition (ISOC), the proposed pathway deploys a nonlinear route, characterized by bond severing in a tricyclic system, leaving a highly functionalized heterocycle. REM linkers, known for their convenience and recyclability, utilize a solid support to synthesize tertiary amines33. Due to the ease of purification accredited to the REM linker via simple filtration, this solid-phase synthesis technique provides scientists with a recyclable and traceless linker, which has been used here. Once the reaction is complete, the REM linker is regenerated and can be reused multiple times. The REM linker is also traceless because, unlike many solid-phase linkers, the point of attachment between the product and the polymer is indistinguishable34,35. Also well-studied and understood is the ISOC reaction, useful in the synthesis of pyrrolidine oximes36,37. Perhaps better known as a 1,3-dipolar cycloaddition, these reactions form a number of heterocycles with high diastereoselectivity38,39,40,41,42,43,44,45. Using the modified REM-coupled-ISOC technique for the synthesis of spirocyclic molecules yields a highly diastereoselective product.Herein, we report on the efficient production of spirocyclic oximes using a new synthetic approach, combining two well-understood pathways and readily available starting materials.

Subscription Required. Please recommend JoVE to your librarian.

Protocol

CAUTION: Please consult all relevant material safety data sheets (MSDS) before use. Several of the chemicals used in these syntheses are acutely toxic and carcinogenic. Please use all appropriate safety practices when performing the following reactions, including the use of engineering controls (fume hood and IR and NMR spectrometers) and personal protective equipment (safety goggles, gloves, lab coat, full-length pants, and closed-toe shoes).

1. Michael Addition of Furfurylamine to the REM Linker

NOTE: The duration of this step is 25 min for the set-up and 24 h of reaction time.

  1. Add 1 g (1 equiv.) of REM resin, 20 mL (20 equiv.) of dimethylformamide (DMF), and 2.4 mL of furfurylamine to a 25 mL solid-phase reaction vessel.
  2. Agitate the reaction vessel for 24 h at room temperature using a shaker following the reaction initiation. The vessel is capped during the reaction.
    NOTE: Ensure that the resin does not sit at the bottom of the vessel and mixes thoroughly.
  3. Drain the solution and wash the resin 1x with 5 mL of DMF after the reaction is complete.
    1. Then, wash the resin 4x, alternating between 5 mL of dichloromethane (DCM) and 5 mL of methanol.
    2. Following the washes, dry the resin thoroughly with compressed air in the reaction vessel for 30 min.
    3. Monitor the reaction progress for a change in IR stretching frequencies, as shown in Table 1.

2. Tandem Michael Addition/1,3-dipolar Cycloaddition

NOTE: The duration of this step is 25 min for the set-up and 48 h of reaction time.

  1. Take the dry resin and add 1.48 mL (5 equiv.) of triethylamine (TEA), 10 mL of dry toluene, and 0.637 g (2 equiv.) of nitro-olefin to the reaction vessel.
  2. Add 1 mL (4 equiv.) of trimethylsilyl chloride (TMSCl) to the reaction vessel in a well-ventilated fume hood.
    CAUTION: This reaction will form HCl gas. Do not cap the reaction vessel until the gas has been released under a fume hood.
  3. Securely cap the reaction vessel and agitate using a shaker for 48 h at room temperature. Ensure that the resin mixes thoroughly with the reagents.
  4. Quench the reaction with 5 mL of methanol.
    1. Drain the solution from the vessel and, then, wash the resin 4x, alternating between 5 mL of DCM and 5 mL of methanol.
    2. Following the washes, dry the resin thoroughly with compressed air in the reaction vessel for 30 min.
    3. Monitor the reaction progress by observing a change in the IR stretching frequencies, as shown in Table 1.

3. Ring Opening of Resin-bound Isoxazole by Tetra-n-butylammonium Fluoride

NOTE: The duration of this step is 10 min for the set-up and 12 h of reaction time.

  1. Place 1 mL of dry tetrahydrofuran (THF) in the reaction vessel with the dry resin. Then, add 1.24 mL (2 equiv.) of 1 M tetra-n-butylammonium fluoride (TBAF) in THF to the reaction vessel.
  2. Using a shaker, agitate the solution for 12 h at room temperature and ensure that the resin thoroughly mixes with the solution.
  3. Drain the solution and wash the resin 1x with 5 mL of THF after the reaction is complete.
    1. Then, wash the resin 4x, alternating between 5 mL of DCM and 5 mL of methanol.
    2. Following the washes, dry the resin thoroughly with compressed air in the reaction vessel for 30 min.
    3. Monitor the reaction progress by observing a change in the IR stretching frequencies, as shown in Table 1.

4. N-alkylation of the Resin-bound Heterocycle to Form Quaternary Amine

NOTE: The duration of this step is 10 min for the set-up and 24 h of reaction time.

  1. Take the dry resin in the reaction vessel and add 5 mL of DMF.
    1. Then, add 1 mL of alkyl halide (10 equiv.) to the vessel and agitate using a shaker for 24 h at room temperature. Ensure the thorough mixing of the resin with the reagents.
  2. Drain the solution and wash the resin 1x with 5 mL of DMF after the reaction is complete.
    1. Then, wash the resin 4x, alternating between 5 mL of DCM and 5 mL of methanol.
    2. Following the washes, dry the resin thoroughly with compressed air in the reaction vessel for 30 min.
    3. Monitor the reaction progress by observing a change in IR stretching frequencies as shown in Table 1.

5. β-elimination of the Quaternary Amine from the Polymer Support

NOTE: The duration of this step is 15 min for the set-up and 24 h of reaction time.

  1. Take the dry resin and add 3 mL of DCM to the reaction vessel.
    1. Then, add 1.5 mL (5 equiv.) of TEA to the reaction vessel to cleave the heterocycle from the polymer support.
    2. Agitate using a shaker for 24 h, ensuring the thorough mixing of the resin with the solution. Drain the solution from the resin.
      NOTE: Do not discard since the cleaved product is in the TEA/DCM solution.
  2. Wash the resin 4x, alternating between 5 mL of DCM and 5 mL of methanol.
    NOTE: Do not discard.
    1. Combine the elution from all washes in steps 5.1.2 and 5.2 and concentrate it via rotatory evaporation.
    2. Purify the spirocyclic oxime by trituration: add 0.5 mL of hot methanol to dissolve any impurities. The pure product will crash out of the solution and is collected via gravity filtration.
  3. Following two washes with 5 mL of DCM for reuse in future experiments, thoroughly dry the resin with compressed air in the reaction vessel for 30 min.
    1. Monitor the reaction progress by observing a change in the IR stretching frequencies, as shown in Table 1.

Subscription Required. Please recommend JoVE to your librarian.

Representative Results

As outlined in the procedure above, the synthetic route to spirocyclic oximes (see Figure 1) begins with the Michael addition of furfurylamine to compound 1, the REM linker, to afford 2. A subsequent Michael addition and 1,3-dipolar cycloaddition of the support 2 using various β-nitrostyrene derivatives yield the tricyclic compound 3, an N-silyloxy isoxazolidine with four unique stereogenic centers. Desilylation of 3 with TBAF produces the spirocyclic oxime 4, still bound to the solid-phase linker. Following the desilylation of 3, polymer-bound 4 is N-alkylated with various electrophiles of choice yielding an ammonium salt, as seen with compound 5. Finally, using β-elimination for the cleavage from the polymer support, compound 6 is generated, along with the fully intact REM linker 1. A library of spirocyclic molecules can be created and purified with ease based on the choice of R1, the β-nitrostyrene, and R2, the electrophiles used in N-alkylation.

For monitoring the progress of each reaction step shown in Figure 1, IR spectroscopy was done on the starting REM resin 1 and on each of the polymer-bound intermediates 2 - 5 to determine whether or not each step had proceeded to completion. These could be classified with a change in the functional group, including conjugated or unconjugated esters, trimethylsilyls, hydroxyls, and oximes, corresponding to a change in wavenumbers as shown in Table 1. NMR analysis was not used to monitor the progress of each step since the intermediates formed are bound to the insoluble polymer support. Corresponding diastereoselective ratios (dr) and yields of the six products 6a - 6f are depicted in Table 2. The yields between 40% and 53% are the overall yields which highlight an average, high yield of between 80% and 88% per step in this five-step route. 1H NMR analysis of the crude product mixture provided the dr values reported.

Figure 1
Figure 1: REM-coupled-ISOC technique for the synthesis of spirocyclic oximes through a tricyclic system intermediate. Customizable R1 and R2 groups using commercially available β-nitrostyrene derivatives and different alkylating reagents, respectively, allow for a library of molecules with a common, spirocyclic backbone to be created, as shown in molecule 6. Please click here to view a larger version of this figure.

Starting material and intermediate IR stretching frequency (cm-1) IR Detectable Functional Groups
1 1722 Conjugated Ester
2 1731 Unconjugated Ester
3 1731 Unconjugated Ester
1214 Trimethylsilyl
4 3600 Hydroxyl
1731 Unconjugated Ester
1655 Oxime
5 3600 Hydroxyl
1731 Unconjugated Ester
1655 Oxime

Table 1: Monitoring solid-phase reactions by infrared spectroscopy. A reaction progression determination of each step was conducted by tracking the changes in the IR stretching frequencies of the starting REM resin 1 and the intermediates 2 - 5.

Product R1 R2 dra yield (%)b
6a phenyl octyl >99:1 40%
6b phenyl methyl 95:5 50%
6c 4-bromophenyl methyl 96:4 53%
6d 4-bromophenyl allyl 96:4 45%
6e 3,4-dimethoxyphenyl benzyl 97:3 45%
6f 2,4-dichlorophenyl methyl >99:1 40%

Table 2: Solid-phase synthesis of N-octyl, -methyl, -allyl, and -benzyl, spirocyclic oximes (products 6a - 6f). (a) The diastereoselective ratio was determined by 1H NMR spectroscopy. (b) The reported yield of the five-step synthesis was determined based on the loading of the REM resin. The overall yield of 40% - 53% indicates an average of 80% - 88% yield for each step.

Subscription Required. Please recommend JoVE to your librarian.

Discussion

In a typical REM linker/solid-phase synthetic strategy, prior to the release of an amine from the solid support, it is critical to form a quaternary ammonium salt, as described in section 4 of the protocol39. Due to the steric hindrance of the tricyclic system and bulky R2 groups (benzyl and octyl halides), only small alkylating reagents (methyl and allyl halides) could be utilized in this reaction46. With a simple modification, allowing for the addition and use of larger, steric reagents, the rigidity of the tricyclic structure was decreased before the N-alkylation step by opening the isoxazoline ring first32. This is illustrated in Figure 1. The ring opening of the tricyclic intermediate 3 relieves the steric hindrance which allows for the addition of virtually any primary alkyl halide desired.

This method was successful in reporting some of the highest dr values in the synthesis of spirocyclic compounds30,47,48. Success in the diastereoselectivity is attributed to the ISOC reaction, which takes the furfurylamine moiety of 2 and creates the rigid, tricyclic system of 338,39,40. Further steps, such as the breaking of the tricyclic system, conserve the diastereoselective nature of the molecule, affording the scientist, in the end, with compounds at or above diastereoselective ratios of 95:5. Equally important is the customizability of the method: with modified β-nitrostyrene derivatives and other electrophiles for N-alkylation, a large library of molecules can be made with relative ease.

To conclude, a highly diastereoselective protocol for the construction of highly functionalized, spirocyclic molecules using a new REM-coupled-ISOC pathway has been developed. This pathway yields a rigid, tricyclic scaffold from the ISOC reaction, from which diastereoselectivity is conserved throughout the remaining reactions. The availability of β-nitrostyrene derivatives and alkylating reagents makes the route convenient and cost-effective. However, should they not be available for purchase, the synthesis of such reagents would be required. This is one such limitation of the method, another being the size of the cycles. As of now, the proposed method is suitable for the construction of a [4.4] spirocyclic framework. Limitations in the 1,3-dipolar cycloaddition method prevent the formation of other ring sizes.

We are in the process of testing the recyclability of the REM linker used in the protocol presented here and will report this shortly. In addition, future applications of the proposed method will be put to use in a number of biological assays. A high-throughput combinatorial synthesis of these spirocyclic molecules using this method can afford a large number of spirocyclic derivatives, which can be tested for anticancer activities in human cancer cells. Such tests will involve cytotoxicity assays, pull-down experiments, and cell culture viability.

Subscription Required. Please recommend JoVE to your librarian.

Disclosures

The authors have nothing to disclose.

Acknowledgments

This work was funded by a grant from the Faculty Research Council to K.S. Huang (Azusa Pacific University - United States). C.R. Drisko is a recipient of the John Stauffer Scholarship and the Gencarella Undergraduate Research Grant. S.A. Griffin received an S2S Undergraduate Research Fellowship from the Department of Biology and Chemistry.

Image 1

Authors (left to right) Cody Drisko, Dr. Kevin Huang and Silas Griffin conducted the experiments and prepared the manuscript. Cody Drisko is a John Stauffer Fellow and a recipient of the Gencarela Research Grant. Silas is a S2S Azusa Pacific University Research Fellow. Dr. Kevin Huang provided the research mentoring and is a recipient of the Azusa Pacific University Faculty Research Council Grant.

Materials

Name Company Catalog Number Comments
Chemicals
REM Resin Nova Biochem 8551010005 Solid Polymer Support; 1.1 mmol/g loading
Furfurylamine Acros Organics 119800050 Reagent
Dimethylformamide (DMF) Sigma-Aldrich 227056 Solvent
Dichloromethane (DCM) Sigma-Aldrich 270997 Solvent
Methanol Sigma-Aldrich 34860 Solvent
trans-4-bromo-β-nitrostyrene Sigma-Aldrich 400017 Nitro-olefin solid
trans-3,4-dimethoxy-β-nitrostyrene Sigma-Aldrich S752215 Nitro-olefin solid
trans-2,4-dichloro-β-nitrostyrene Sigma-Aldrich 642169 Nitro-olefin solid
trans-β-nitrostyrene Sigma-Aldrich N26806 Nitro-olefin solid
Triethylamine (TEA) Sigma-Aldrich T0886 Solvent
Trimethylsilyl chloride (TMSCl) Sigma-Aldrich 386529 Reagent; CAUTION - highly volatile; creates HCl gas
Tetra-n-butylammonium fluoride (TBAF) in Tetrahydrofuran (THF) Sigma-Aldrich 216143 Reagent
Tetrahydrofuran (THF) Sigma-Aldrich 401757 Reagent
1-Bromooctane Sigma-Aldrich 152951 Alkyl-halide
Iodomethane Sigma-Aldrich 289566 Alkyl-halide
Allylbromide Sigma-Aldrich 337528 Alkyl-halide
Benzylbromide Sigma-Aldrich B17905 Alkyl-halide
Glassware/Instrumentation
25 mL solid-phase reaction vessel Chemglass CG-1861-02 Glassware with filter
Thermo Scientific Nicole iS5 Thermo Scientific IQLAADGAAGFAHDMAZA Instrument
AVANCE III NMR Spectrometer Bruker N/A Instrument; 300 MHz; Solvents: CDCl3 and CD3OH
Wrist-Action Shaker Model 75 Burrell Scientific 757950819 Instrument

DOWNLOAD MATERIALS LIST

References

  1. Bayat, M., Amiri, Z. Chemoselective synthesis of novel spiropyrano acenaphthylene derivatives via one-pot four-component reaction. Tetrahedron Letters. 58 (45), 4260-4263 (2017).
  2. Ding, K., et al. Structure-Based Design of Potent Non-Peptide MDM2 Inhibitors. Journal of the American Chemical Society. 127 (29), 10130-10131 (2005).
  3. D'Erasmo, M. P., et al. 7,9-Diaryl-1,6,8-trioxaspiro[4.5]dec-3-en-2-ones: Readily accessible and highly potent anticancer compounds. Bioorganic & Medicinal Chemistry Letters. 24 (16), 4035-4038 (2014).
  4. Gomez, C., et al. Phosphine-Catalyzed Synthesis of 3,3-Spirocyclopenteneoxindoles from γ-Substituted Allenoates: Systematic Studies and Targeted Applications. The Journal of Organic Chemistry. 78 (4), 1488-1496 (2013).
  5. Wu, S., et al. Novel spiropyrazolone antitumor scaffold with potent activity: Design, synthesis and structure-activity relationship. European Journal of Medicinal Chemistry. 115, 141-147 (2016).
  6. Allgardsson, A., et al. Structure of a prereaction complex between the nerve agent sarin, its biological target acetylcholinesterase, and the antidote HI-6. Proceedings of the National Academy of Sciences. 113 (20), 5514-5519 (2016).
  7. Cantín, Á, et al. Novel Inhibitors of the Mitochondrial Respiratory Chain: Oximes and Pyrrolines Isolated from Penicillium brevicompactum and Synthetic Analogues. Journal of Agricultural and Food Chemistry. 53 (21), 8296-8301 (2005).
  8. Wu, E. S. C., et al. et al. In Vitro Muscarinic Activity of Spiromuscarones and Related Analogs. Journal of Medicinal Chemistry. 38 (9), 1558-1570 (1995).
  9. Gober, C. M., Carroll, P. J., Joullié, M. M. Triazaspirocycles: Occurrence, Synthesis, and Applications. Mini-reviews in organic chemistry. 13 (2), 126-142 (2016).
  10. Hong, C. Y., et al. Novel Fluoroquinolone Antibacterial Agents Containing Oxime-Substituted (Aminomethyl)pyrrolidines: Synthesis and Antibacterial Activity of 7-(4-(Aminomethyl)-3-(methoxyimino)pyrrolidin-1-yl)-1-cyclopropyl-6-fluoro- 4-oxo-1,4-dihydro[1,8]naphthyridine-3-carboxylic Acid (LB20304),1. Journal of Medicinal Chemistry. 40 (22), 3584-3593 (1997).
  11. Ryzhakov, D., Jarret, M., Guillot, R., Kouklovsky, C., Vincent, G. Radical-Mediated Dearomatization of Indoles with Sulfinate Reagents for the Synthesis of Fluorinated Spirocyclic Indolines. Organic Letters. 19 (23), 6336-6339 (2017).
  12. Wang, L., et al. A Facile Radiolabeling of [18F]FDPA via Spirocyclic Iodonium Ylides: Preliminary PET Imaging Studies in Preclinical Models of Neuroinflammation. Journal of Medicinal Chemistry. 60 (12), 5222-5227 (2017).
  13. Lin, Y., Jones, G. B., Hwang, G. -S., Kappen, L., Goldberg, I. H. Convenient Synthesis of NCS−Chromophore Metabolite Isosteres: Binding Agents for Bulged DNA Microenvironments. Organic Letters. 7 (1), 71-74 (2005).
  14. Kappen, L. S., Lin, Y., Jones, G. B., Goldberg, I. H. Probing DNA Bulges with Designed Helical Spirocyclic Molecules. Biochemistry. 46 (2), 561-567 (2007).
  15. Zhang, N., Lin, Y., Xiao, Z., Jones, G. B., Goldberg, I. H. Solution Structure of a Designed Spirocyclic Helical Ligand Binding at a Two-Base Bulge Site in DNA. Biochemistry. 46 (16), 4793-4803 (2007).
  16. Thomas, J. R., Hergenrother, P. J. Targeting RNA with Small Molecules. Chemical Reviews. 108 (4), 1171-1224 (2008).
  17. Jones, B., Proud, M., Sridharan, V. Synthesis of oxetane/azetidine containing spirocycles via the 1,3-dipolar cycloaddition reaction. Tetrahedron Letters. 57 (25), 2811-2813 (2016).
  18. Martinez, N. J., et al. A High-Throughput Screen Identifies 2,9-Diazaspiro[5.5]Undecanes as Inducers of the Endoplasmic Reticulum Stress Response with Cytotoxic Activity in 3D Glioma Cell Models. PLoS ONE. 11 (8), e0161486 (2016).
  19. Wang, Y., et al. Discovery and Optimization of Potent GPR40 Full Agonists Containing Tricyclic Spirocycles. ACS Medicinal Chemistry Letters. 4 (6), 551-555 (2013).
  20. Singh, G. S., Desta, Z. Y. Isatins As Privileged Molecules in Design and Synthesis of Spiro-Fused Cyclic Frameworks. Chemical Reviews. 112 (11), 6104-6155 (2012).
  21. Rana, S., et al. Isatin Derived Spirocyclic Analogues with α-Methylene-γ-butyrolactone as Anticancer Agents: A Structure-Activity Relationship Study. Journal of Medicinal Chemistry. 59 (10), 5121-5127 (2016).
  22. Sue, D., Kawabata, T., Sasamori, T., Tokitoh, N., Tsubaki, K. Synthesis of Spiro Compounds through Tandem Oxidative Coupling and a Framework Rearrangement Reaction. Organic Letters. 12 (2), 256-258 (2010).
  23. Perry, M. A., Hill, R. R., Rychnovsky, S. D. Trianion Synthon Approach to Spirocyclic Heterocycles. Organic Letters. 15 (9), 2226-2229 (2013).
  24. Palmer, L. I., Read de Alaniz, J. Rapid and Stereoselective Synthesis of Spirocyclic Ethers via the Intramolecular Piancatelli Rearrangement. Organic Letters. 15 (3), 476-479 (2013).
  25. Berton, J. K. E. T., Salemi, H., Pirat, J. -L., Virieux, D., Stevens, C. V. Three-Step Synthesis of Chiral Spirocyclic Oxaphospholenes. The Journal of Organic Chemistry. 82 (23), 12439-12446 (2017).
  26. Carreira, E. M., Fessard, T. C. Four-Membered Ring-Containing Spirocycles: Synthetic Strategies and Opportunities. Chemical Reviews. 114 (16), 8257-8322 (2014).
  27. Yamazaki, S., Naito, T., Niina, M., Kakiuchi, K. Lewis Acid Catalyzed Cyclization Reactions of Ethenetricarboxylates via Intramolecular Hydride Transfer. The Journal of Organic Chemistry. 82 (13), 6748-6763 (2017).
  28. Hung, A. W., et al. Route to three-dimensional fragments using diversity-oriented synthesis. Proceedings of the National Academy of Sciences. 108 (17), 6799-6804 (2011).
  29. Wright, D. L., Schulte, J. P., Page, M. A. An Imine Addition/Ring-Closing Metathesis Approach to the Spirocyclic Core of Halichlorine and Pinnaic Acid. Organic Letters. 2 (13), 1847-1850 (2000).
  30. Qiu, B., et al. Highly Enantioselective Oxidation of Spirocyclic Hydrocarbons by Bioinspired Manganese Catalysts and Hydrogen Peroxide. ACS Catalysis. 8 (3), 2479-2487 (2018).
  31. Richmond, E., Duguet, N., Slawin, A. M. Z., Lébl, T., Smith, A. D. Asymmetric Pericyclic Cascade Approach to Spirocyclic Oxindoles. Organic Letters. 14 (11), 2762-2765 (2012).
  32. Griffin, S. A., Drisko, C. R., Huang, K. S. Tricyclic heterocycles as precursors to functionalized spirocyclic oximes. Tetrahedron Letters. , (2017).
  33. Brown, A. R., Rees, D. C., Rankovic, Z., Morphy, J. R. Synthesis of Tertiary Amines Using a Polystyrene (REM) Resin. Journal of the American Chemical Society. 119 (14), 3288-3295 (1997).
  34. Blaney, P., Grigg, R., Sridharan, V. Traceless Solid-Phase Organic Synthesis. Chemical Reviews. 102 (7), 2607-2624 (2002).
  35. Morphy, J. R., Rankovic, Z., Rees, D. C. A novel linker strategy for solid-phase synthesis. Tetrahedron Letters. 37 (18), 3209-3212 (1996).
  36. Saruengkhanphasit, R., Collier, D., Coldham, I. Synthesis of Spirocyclic Amines by Using Dipolar Cycloadditions of Nitrones. The Journal of Organic Chemistry. 82 (12), 6489-6496 (2017).
  37. Li, F., et al. Assembly of Diverse Spirocyclic Pyrrolidines via Transient Directing Group Enabled Ortho-C(sp2)-H Alkylation of Benzaldehydes. Organic Letters. 20 (1), 146-149 (2018).
  38. Gottlieb, L., Hassner, A. Cycloadditions. 53. Stereoselective Synthesis of Functionalized Pyrrolidines via Intramolecular 1,3-Dipolar Silyl Nitronate Cycloaddition. The Journal of Organic Chemistry. 60 (12), 3759-3763 (1995).
  39. Namboothiri, I. N. N., Hassner, A., Gottlieb, H. E. A Highly Stereoselective One-Pot Tandem Consecutive 1,4-Addition−Intramolecular 1,3-Dipolar Cycloaddition Strategy for the Construction of Functionalized Five- and Six-Membered Carbocycles,1. The Journal of Organic Chemistry. 62 (3), 485-492 (1997).
  40. Dehaen, W., Hassner, A. Stereoselectivity in intramolecular 1,3-dipolar cycloadditions. Nitrile oxides versus silyl nitronates. Tetrahedron Letters. 31 (5), 743-746 (1990).
  41. Roger, P. -Y., Durand, A. -C., Rodriguez, J., Dulcère, J. -P. Unprecedented in Situ Oxidative Ring Cleavage of Isoxazolidines: Diastereoselective Transformation of Nitronic Acids and Derivatives into 3-Hydroxymethyl 4-Nitro Tetrahydrofurans and Pyrrolidines. Organic Letters. 6 (12), 2027-2029 (2004).
  42. Kudoh, T., Ishikawa, T., Shimizu, Y., Saito, S. Intramolecular Cycloaddition Reactions of Silyl Nitronate Tethered to Vinylsilyl Group: 2-Nitroalkanols as Precursors for Amino Polyols. Organic Letters. 5 (21), 3875-3878 (2003).
  43. Ishikawa, T., Shimizu, Y., Kudoh, T., Saito, S. Conversion of d-Glucose to Cyclitol with Hydroxymethyl Substituent via Intramolecular Silyl Nitronate Cycloaddition Reaction: Application to Total Synthesis of (+)-Cyclophellitol. Organic Letters. 5 (21), 3879-3882 (2003).
  44. Hashimoto, T., Maruoka, K. Recent Advances of Catalytic Asymmetric 1,3-Dipolar Cycloadditions. Chemical Reviews. 115 (11), 5366-5412 (2015).
  45. Li, X., et al. Highly Enantioselective One-Pot Synthesis of Spirocyclopentaneoxindoles Containing the Oxime Group by Organocatalyzed Michael Addition/ISOC/Fragmentation Sequence. Organic Letters. 13 (23), 6160-6163 (2011).
  46. Jensen, K. H., Hanson, J. E. Synthesis and Photochemistry of Tertiary Amine Photobase Generators. Chemistry of Materials. 14 (2), 918-923 (2002).
  47. Mondal, S., Mukherjee, S., Yetra, S. R., Gonnade, R. G., Biju, A. T. Organocatalytic Enantioselective Vinylogous Michael-Aldol Cascade for the Synthesis of Spirocyclic Compounds. Organic Letters. 19 (16), 4367-4370 (2017).
  48. Ni, C., et al. Phosphine-Catalyzed Asymmetric (3 + 2) Annulations of δ-Acetoxy Allenoates with β-Carbonyl Amides: Enantioselective Synthesis of Spirocyclic β-Keto γ-Lactams. Organic Letters. 19 (13), 3668-3671 (2017).

Tags

Solid-phase Synthesis [4.4] Spirocyclic Oximes Combinatorial Library Anti-cancer Activities Bioassays Spirocyclic Compounds Isoxazolines Isoxazoles REM Resin Dimethylformamide Furfurylamine DMF Dichloromethane Methanol Triethylamine Anhydrous Toluene Beta Nitrostyrene Trimethylsilyl Chloride
Solid-phase Synthesis of [4.4] Spirocyclic Oximes
Play Video
PDF DOI DOWNLOAD MATERIALS LIST

Cite this Article

Drisko, C. R., Griffin, S. A.,More

Drisko, C. R., Griffin, S. A., Huang, K. S. Solid-phase Synthesis of [4.4] Spirocyclic Oximes. J. Vis. Exp. (144), e58508, doi:10.3791/58508 (2019).

Less
Copy Citation Download Citation Reprints and Permissions
View Video

Get cutting-edge science videos from JoVE sent straight to your inbox every month.

Waiting X
Simple Hit Counter