Você está na página 1de 13

KARL HUGHES

CHEM 356, MAY 2012

Introduction
Aspirin 1 is the archetypal molecule of the class of drugs known as non-steroidal antiinflammatories (NSAIDs). Discovered in the 1700s, it is widely used to treat a variety of conditions, from fever to osteoarthritis. However, it is relatively acidic (pKa~3)1 and can have severe side effects2.

1 This is also true for many other NSAIDs which are based on the same structure. This has prompted the search for molecules that exhibit anti-inflammatory properties but arent based on this structure and consequently dont have the same side effects. The sulfur atom is about the same size as a CH2CH2 group3 and has the same electronegativity as carbon; thus, thiophene 2 and benzene 3 are roughly the same size3 and the thiophene moiety has been incorporated into many drugs.

Substituted thiophenes have found many biological uses, such as antibacterial4 and antiviral5 agents; anaesthetics6; and, recently, as a neurotoxin inhibitor7. Substituted thiophenes have also been found to have anti-inflammatory activity8. In particular 2-morpholino-3-aryl-5aroyl thiophenes 4 are particularly effective. The biological activity of this class of compounds is maximised when Ar1=Ph and Ar2 is a benzene ring substituted at the para-position by a substituent that cant be metabolically oxidised (e.g. a halogen) otherwise rapid excretion occurs9.

KARL HUGHES

CHEM 356, MAY 2012

4 The simplest of these, 5, will be prepared in this experiment.

5 Thiophenes, like all aromatic molecules can undergo Friedel-Crafts acylation or alkylation using a Lewis acid catalyst (Scheme 1).

Scheme 1

KARL HUGHES

CHEM 356, MAY 2012

However, this fails when 2-aminothiophenes are used10 because of deactivation of the catalyst by the amine and, therefore, other methods are required. Accordingly, several methods of synthesising 2,3,5-trisubstituted thiophenes have been developed. Scheme 2 depicts two such methods, the Vilsmeier-Haack reaction11 of 8 with POCl3 and that of 8 with Lawesson Reagent or P2S512. These both have disadvantages: Lawesson Reagent and P2S5 are both expensive and the Vilsmeier-Haack reaction uses explosive POCl3. Moreover, the latter may also give low yields of the intermediate hypochlorite 1113.

Scheme 2 None of these is desirable, especially when carrying out a large-scale synthesis, such as that of a pharmaceutical. Thus, if this class of products is to be a viable alternative to traditional NSAIDs, another method is needed. Liebscher and Rolfs have developed a method13, 14 which is similar to the above route but uses a Willgerodt-Kindler reaction for the first step (Scheme 3) to produce thioamide 16. The Kindler modification of the Willgerodt reaction uses elemental sulfur, a cheap alternative to the other methods. A versatile reaction, it proceeds with aromatic, alicyclic and aliphatic substrates15 and is therefore also a useful method for preparing analogues to be tested for biological activity. This will be followed by the reaction of HC(OEt)3 and morpholine to form thioacrylamide 17. Finally, an SN2 of 17 with the -bromoketone 18 and subsequent cyclisation with Et3N yields the title compound 5.
3

KARL HUGHES

CHEM 356, MAY 2012

Scheme 3

KARL HUGHES

CHEM 356, MAY 2012

Results and discussion


Phenylthioacetic acid morpholide, 16 Acetophenone, morpholine and elemental sulfur were refluxed with catalytic TsOH for 3 h. Equimolar amounts of acetophenone and sulfur were used, with morpholine in excess as both reagent and solvent. The reaction was quenched by pouring the reaction mixture into MeOH followed by cooling and recrystallisation from MeOH and water (initial 1:1 ratio). The nearcolourless 16 was obtained in 44% yield; however, it should be noted that although increasing the amount of sulfur to 2 equivalents does increase the yield of 16, it also results in a difficult-to-remove impurity, 1914.

Mechanism of the Willgerodt-Kindler reaction There is some disagreement about the mechanism and many attempts to elucidate it have been recorded16. Brown16 suggests that, given the wide range of substrates with which and conditions under which the reaction proceeds, there may be several different mechanisms. Nevertheless, those suggestions made by King and McMillan17 (Scheme 4a) and DeTar and Carmack18 (Scheme 4b) seem to be the most widely accepted.

Scheme 4a
5

KARL HUGHES

CHEM 356, MAY 2012

These differ predominantly in whether the amine adds before or after addition of sulfur. However, they both involve a so-called migrating group. In the first mechanism, it is the sulfur which migrates (after either reduction to the alcohol and then elimination to form the olefin, to which sulfur can add, or direct displacement of oxygen to yield the thioketone followed by elimination of H2S); in the second, the amine migrates after first adding onto the carbonyl and eliminating water via the imine and then tautomerising to the enamine and the intermediate in this case is an acetylene. Both of these mechanisms at some point require the formation of a primary cation; while this is obviously unfavourable, the subsequent steps are essentially irreversible and so would drive the reaction to completion.

Scheme 4b It has been reported19 that, in base, the polysulfide anion is present and from this, Kanbara19 suggested that the initial steps were simultaneous condensation of the amine and ketone, and attack by the amine on the S8 ring. In acidic media, the thiol could be present which would react in the same way; this alternative mechanism is shown here (Scheme 5).

KARL HUGHES

CHEM 356, MAY 2012

Scheme 5 Characterisation of 16 16 was characterised by elemental analysis, IR, and 1H spectroscopy (see experimental and appendices A and B). The theoretical and experimental analyses were in agreement and the IR showed a peak at 1106.94 cm-1, characteristic of the C=S stretch20. The NMR showed a characteristic singlet at 4.36 ppm, corresponding to the methylene of 16, as well as the multiplets characteristic of morpholines: one at 3.46 ppm and one centred on 3.6 ppm. 4-morpholino-2-phenylthioacrylic acid morpholide 17 The formation of a new C-C bond was effected with HC(OEt)3, and subsequent reaction with morpholine yielded the thioacrylamide. 16 was reacted with morpholine and triethyl orthoformate, both in excess to act as both reagent and solvent. EtOH was removed by distillation to drive the reaction to completion and 17 was obtained as a yellow solid in 72% yield after recrystallisation from MeOH and chloroform (1:5 ratio). The reaction took 9-10 h for completion and was monitored by TLC on silica gel, mobile phase EtOAc-Pet. ether (1:2).

KARL HUGHES

CHEM 356, MAY 2012

Mechanism The mechanism is shown in Scheme 6. The orthoester cleaves on heating to yield ethoxide and an oxonium in situ. Ethoxide attacks 16 producing the thioenolate which then reacts with the oxonium species in a conjugate type addition. This subsequently reacts with morpholine and tautomerisation leads to the enamine. Characterisation of 17 17 was characterised by elemental analysis, IR and 1H NMR spectroscopy. The theoretical and experimental analyses were in agreement and the IR showed a C= peak at 1141.65 cm -1. The 1H NMR spectrum showed a singlet at 7.4 ppm, characteristic of the methylene group, as well as a multiplet at 3.4-3.8 ppm, integrating to 16 protons and therefore characteristic of the morpholine substituents. 5-(4-bromobenzoyl)-2-(4-morpholino)-2-phenylthiophene, 5 Equimolar amounts of 17 and 18 were dissolved in MeOH and heated to reflux. The SN2 reaction was facile, due to the sulfur nucleophile, primary carbon, bromide leaving group and adjacent carbonyl. The sulfonium salt was not isolated since cyclisation could be effected

Scheme 6
8

KARL HUGHES

CHEM 356, MAY 2012

immediately with Et3N. As expected, aromaticity followed swiftly and yellow 5 was obtained in good yield (85%) and as the pure product; no further purification was necessary. Mechanism The mechanism is straight forward (Scheme 7); SN2 is followed by attack of triethylamine, forming the enolate which readily cyclises to the thiophene with loss of morpholine. Characterisation of 5 Again, elemental analysis, IR and 1H NMR spectroscopy was used. The experimental and theoretical analyses were in agreement and the only peak of note in the IR is the carbonyl C=O at 1616.06 cm-1. The NMR showed the morpholine resonances at 3.8 ppm while the characteristic singlet of the thiophene proton in the aromatic region was found at ~7.4 ppm

Scheme 7 .
13C

NMR spectra

There were no major problems in the experiment; however, as can be seen from appendices C, F and I (the
13

C spectra of 16, 17 and 5 respectively), only the solvent peaks are clearly

visible. At least 50 mg sample was used in the analysis, but evidently this was not sufficient and after other analyses and subsequent steps in the experiment were undertaken, there was
9

KARL HUGHES

CHEM 356, MAY 2012

not enough remaining sample to repeat the experiments (it was assumed that using < 50 mg would be fruitless).Therefore, these spectra could not be used for characterisation purposes.

Conclusions
In summary, the total synthesis of 5 was achieved with 27% overall yield in 3 steps. The synthetically versatile Willgerodt-Kindler reaction was employed, followed by the orthoformate method of forming C-C bonds to incorporate a methylene group; finally, an SN2 followed by cyclisation gave the final thiophene product. The synthetic sequence presented no major problems, required no non-standard laboratory equipment and has a moderate overall yield. Thus, it is a good route to this potential pharmaceutically important class of molecules, allowing for fairly efficient synthesis and its possible application to a wide range of substrates means it could also be used for the preparation of the analogous substituted thiophenes currently under investigation for use as anti-inflammatories. Further modifications of the experimental procedures (recycling of solvents, for instance) would enhance the green credentials of this experiment, something which the pharmaceutical industry is acutely aware that more focus is needed on.

10

KARL HUGHES

CHEM 356, MAY 2012

Experimental
All reagents were commercially available and used without further purification. Elemental analyses was performed by the University of Liverpool Chemistry Department; infra red spectra were run neat on a Jasco FT/IR 4200 spectrometer; 1H- and
13

C-NMR spectra were

recorded on a Bruker ARX 250 spectrometer using deuterated solvents and TMS as an internal standard; melting points were determined on a Gallenkamp melting point apparatus and TLCs were run on Sigma-Aldrich silica gel 60 F254 plates, mobile phase 1:2 ethyl acetate: pet. ether (40-60oC fraction). Step 1 Acetophenone (3.0 mL, 26 mmol), morpholine (4.6 mL, 52 mmol), sulfur (0.84 g, 26 mmol) and p-toluene sulfonic acid monohydrate (0.13 g, 0.67 mmol) was heated to reflux (T~ 130oC) for 3 h. The red-brown solution was poured into 13 mL hot (T~60oC) methanol and the precipitate isolated after cooling in an ice-water bath. The crude product was suspended in 3 mL methanol and 3mL water, heated to reflux and methanol added dropwise until complete solution occurred. The almost colourless pure product 1 was collected after cooling and dried in the open air (2.52g, 44%), mp 76-77oC (MeOH-water). (Anal. Calcd. for C12H15NOS: C, 65.12; H, 6.83; N, 6.33. Found: C, 65.10; H, 6.89; N, 6.32; max(neat)/cm-1 1106.94 (C=S); H (250 MHz; CDCl3; Me4Si, p.p.m.) 7.319 (5H, m, Ph), 4.364 (2H, m, morpholide-CH2), 4.364 (2H, s, PhCH2CS), 3.725 (2H, m, morpholide-CH2), 3.484 (2H, m, morpholide-CH2), 3.370 (2H, m, morpholide-CH2). Step 2 1 (1.43 g, 6.47 mmol), morpholine (1.2 mL, 12.9 mmol) and triethyl orthoformate (4.3 mL, 26 mmol) were added to a distillation apparatus and heated for 10 h (until TLC of the reaction mixture showed no remaining 1, stationary phase 1:2 EtOAc: pet. ether), during which time ethanol was removed by distillation. The orange mixture was evaporated in vacuo (bath T~80oC) and the orange-yellow precipitate recrystallised from chloroform and
11

KARL HUGHES

CHEM 356, MAY 2012

methanol (1:5, methanol was added until complete solution occurred). The pure yellow product 2 was isolated by filtration, washed successively with cold methanol and ether and dried in air (1.49g, 72%), mp 153-154oC (chloroform-MeOH). (Anal. Calcd. for C17H22N2O2S: C, 64.12; H, 6.96; N, 8.80. Found: C, 63.80; H, 6.77; N, 8.91; max(neat)/cm-1 1114.65 (C=S); H (250 MHz; CDCl3; Me4Si, p.p.m.) 7.4 (5H, m, Ph), 6.4 (1H, s, methyne), 3.8 (16H, m, morpholine). Step 3 2 (1.24 g, 3.88 mmo) and 1-4-dibromoacetophenone (1.08 g, 3.88 mmol) were suspended in methanol (19 mL) and the mixture was heated to reflux. Triethylamine (0.54 mL, 3.88 mmol) in methanol (3.9 mL) was added and the mixture refluxed for a further 10 minutes. Cooling and filtration yielded the target compound 3 as a yellow solid (1.41g, 85%), mp 174-175oC. (Anal. Calcd. for C21H18BrNO2S: C, 58.88; H, 4.24; N, 3.27. Found: C, 57.81; H, 4.28; N, 3.27; max(neat)/cm-1 1616.06 (C=O), 1234.22 (C-O); H (250 MHz; CDCl3; Me4Si, p.p.m.) 7.2-7.8 (10H, m, Ph), 7.5 (1H, s, thiophene), 3.75 (4H, m, morpholine), 3.1 (4H, m, morpholine).

References
1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. M. W. Whitehouse, Curr Med Chem, 2005, 12, 2931-2942. K. I. Molvi, K. K. Vasu, S. G. Yerande, V. Sudarsanam and N. Haque, Eur J Med Chem, 2007, 42, 1049-1058. S. Raju, P. R. Kumar, K. Mukkanti, P. Annamalai and M. Pal, Bioorg Med Chem Lett, 2006, 16, 6185-6189. E. Abele and E. Lukevics, Chemistry of Heterocyclic Compounds, 2001, 37, 141-169. R. L. Jarvest, I. L. Pinto, S. M. Ashman, C. E. Dabrowski, A. V. Fernandez, L. J. Jennings, P. Lavery and D. G. Tew, Bioorganic and Medicinal Chemistry Letters, 1999, 9, 443-448. K. E. Schulze, P. R. Cohen and B. R. Nelson, Dermatol Surg, 2006, 32, 407-410. I. Opsenica, V. Filipovic, J. E. Nuss, L. M. Gomba, D. Opsenica, J. C. Burnett, R. Gussio, B. A. Solaja and S. Bavari, Eur J Med Chem, 2012. A. D. Pillai, P. D. Rathod, P. X. Franklin, M. Patel, M. Nivsurkar, K. K. Vasu, H. Padh and V. Sudarsanam, Biochem Bioph Res Co, 2003, 301, 183-186. A. D. Pillai, S. Rani, P. D. Rathod, F. P. Xavier, K. K. Vasu, H. Padh and V. Sudarsanam, Bioorgan Med Chem, 2005, 13, 1275-1283. A. Noack and H. Hartmann, Tetrahedron, 2002, 58, 2137-2146. J. Liebscher and B. Abegaz, Synthesis-Stuttgart, 1982, 769-771. C. Heyde, I. Zug and H. Hartmann, Eur J Org Chem, 2000, 3273-3278. A. Rolfs and J. Liebscher, Synthesis-Stuttgart, 1994, 683-684. A. Rolfs and J. Liebscher, Org. Synth. , 1997, 74. J. A. King and F. H. Mcmillan, J Am Chem Soc, 1946, 68, 525-526. E. V. Brown, Synthesis-Stuttgart, 1975, 358-375. J. A. King and F. H. Mcmillan, J Am Chem Soc, 1946, 68, 632-636. D. F. Detar and M. Carmack, J Am Chem Soc, 1946, 68, 2025-2029. 12

KARL HUGHES 19. 20.

CHEM 356, MAY 2012

T. Kanbara, Y. Kawai, K. Hasegawa, H. Morita and T. Yamamoto, J Polym Sci Pol Chem, 2001, 39, 3739-3750. E. Spinner, J Org Chem, 1958, 23, 2037-2038.

13

Você também pode gostar