Next Article in Journal
Facile and Convenient Synthesis of New Thieno[2,3-b]-Thiophene Derivatives
Previous Article in Journal
Insecticidal Activity of Essential Oil of Carum Carvi Fruits from China and Its Main Components against Two Grain Storage Insects
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of 2-(4,6-Dimethoxy-1,3,5-triazin-2-yloxyimino) Derivatives: Application in Solution Peptide Synthesis

by
Tarfah I. Al-Warhi
1,*,
Hassan M.A. AL-Hazimi
2,
Ayman El-Faham
2,3 and
Fernando Albericio
4,5,6
1
Women Students-Medical Studies and Sciences Sections, Chemistry Department, College of Science, King Saud University, P.O. Box 22452, Riyadh 11495, Saudi Arabia
2
Chemistry Department, College of Science, King Saud University, P.O. Box 2455, Riyadh 11451, Saudi Arabia
3
Chemistry Department, Faculty of Science, Alexandria University, 426 Ibrahimia, 21321 Alexandria, Egypt
4
Institute for Research in Biomedicine, Barcelona Science Park, Baldiri Reixac 10, Barcelona 08028, Spain
5
CIBER-BBN, Networking Centre on Bioengineering, Biomaterials and Nanomedicine, Barcelona Science Park, Baldiri Reixac 10, Barcelona 08028, Spain
6
Department of Organic Chemistry, University of Barcelona, Martí i Franqués 1-11, Barcelona 08028, Spain
*
Author to whom correspondence should be addressed.
Molecules 2010, 15(12), 9403-9417; https://doi.org/10.3390/molecules15129403
Submission received: 29 November 2010 / Revised: 14 December 2010 / Accepted: 15 December 2010 / Published: 20 December 2010

Abstract

:
A new class of 1,3,5-triazinyloxyimino derivatives were prepared, characterized and tested for reactivity in solution peptide synthesis. The new triazinyloxyimino derivatives failed to activate the carboxyl group during formation of peptide bonds, but gave the corresponding N-triazinyl amino acid derivatives as a major product. The oxyma (ethyl 2-cyano-2-(hydroxyimino)acetate) uronium salt was superior to other uronium salts in terms of racemization, while 2-chloro-4,6-dimethoxy-1,3,5-triazine (CDMT, 9) gave the best results.

1. Introduction

For many years, 1-hydroxybenzotriazole (HOBt, 1, Figure 1)has been used as a coupling reagent in peptide synthesis, especially in combination with dicyclohexylcarbodiimide (DCC, 2, Figure 1) [1,2,3,4]. Later, other benzotriazole derivatives such as 6-chloro-1-hydroxybenzotriazole (6-Cl-HOBt, 3, Figure 1) and 1-hydroxy-7-azabenzotriazole (HOAt, 4, Figure 1) [5,6] have been introduced and shown more efficient that the parent (HOBt, 1).
Figure 1. Structure of additives and DCC.
Figure 1. Structure of additives and DCC.
Molecules 15 09403 g001
During the last decade, iminium/uronium and phosphonium salts of 1-hydroxybenzotriazole and 1-hydroxy-7-azabenzotriazole, 5 and 6 (Figure 2)have shown the supremacy of these kind of derivatives over carbodiimides in the presence of an additive [5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33].
Despite their widespread use, these reagents are not without ptoblems. The explosive properties of 1-hydroxybenzotriazoles are often not referenced in literature. Sometime, Material Safety Data Sheets warn that 1-hydroxybenzotriazoles may be unstable, with a relatively high sensitivity to friction and sparks, but in most cases, there is no mention of how sensitive such substances are to heating under confinement and no warning is given with respect to their ability to propagate a deflagration or a detonation.
More recently a new class of coupling reagents such as 7 and 8 (Figure 2) based on changes to the carbon skeleton structure [34,35,36] gave more marked increases in coupling efficiency, as well as reduced racemization levels during the coupling step. Particularly noteworthy are the morpholino derivatives 8, because the oxygen in the carbon skeleton increases the solubility as well as the reactivity, which reduces the racemization during the coupling step [34,35,36].
Figure 2. Structure of uronium/iminium salts.
Figure 2. Structure of uronium/iminium salts.
Molecules 15 09403 g002
Although 2-chloro-4,6-dimethoxy-1,3,5-triazine (CDMT, 9) has been successfully applied for the preparation of peptides, the mechanism of its participation in coupling reactions remains unknown [37,38]. Furthermore, in order to be successfully stored the reagent itself must be of high purity, the formation of gaseous products may generate a rapid pressure increase in the container, and accordingly, precautions should be taken to avoid the serious risk of blowout of toxic gases.
Successful activation of carboxylic acids by means of CDMT (9) confirmed the feasibility of a multistep process proceeding via triazinylammonium salts such as 10 and (DMTMM, 11) (Figure 3)formed in situ in the presence of the appropriate amine [37,38].
Figure 3. Structure of triazinylammonium salts.
Figure 3. Structure of triazinylammonium salts.
Molecules 15 09403 g003
The related triazine, 4-(4,6-dimethoxy-1,3,5-triazin-2-yl)-4-methylmorpholinium chloride (DMTMM, 11), formed by a simple reaction of CDMT (9) with N-methylmorpholine (NMM), has been found applications in amidation [38,39,40,41,42], esterification [42], glycosidation [43,44] and phosphonylation [45] methodology. The present work presents the synthesis and application of a new family of 1,3,5-triazine derivatives and their comparison with CDMT (9) as well as the uronium based type coupling reagents [46].

2. Results and Discussion

The new triazinyloxyimino derivatives 12a-c were prepared from the reaction of the oxime derivatives 13a-c with CDMT (9) in the presence of triethylamine. The potassium salts were used in case of the synthesis of malononitrile and diethylmalonate derivatives (Scheme 1).
Scheme 1. Synthesis of triazinyloxyimino derivatives.
Scheme 1. Synthesis of triazinyloxyimino derivatives.
Molecules 15 09403 g005
The pyridinyl derivative was prepared using the same conditions to afford 12d. All the products 12a-d have been identified on the basis of their spectroscopic data (see Experimental). In order to have a good comparison to the triazinyloxyimino coupling reagents, some of the related uronium-type coupling reagents were prepared according to the reported method [36] as shown in Scheme 2.
Scheme 2. Synthesis of uronium salts.
Scheme 2. Synthesis of uronium salts.
Molecules 15 09403 g006
The uronium salt coupling reagents were prepared by reaction of the chloro salts DCMH 14a or TCFH 14b with the oxime derivatives in the presence of triethylamine in DCM or with the potassium salt of the oxime derivatives in acetonitrile to give the oxyimino uronium base coupling reagents 15a-f (Scheme 2). To investigate the retention of configuration induced for the new coupling reagents, several previously studied model peptide system 16a,b and 17a-c (Figure 4) were examined [36]. These models involve stepwise coupling and (2+1) segment coupling as well.
Figure 4. Model of peptide used in racemization studies.
Figure 4. Model of peptide used in racemization studies.
Molecules 15 09403 g004
Results comparing the coupling of Z-Phe-OH to H-Ala-OMe.HCl (18a) with different coupling agents to afford the Z-Phe-Ala-OMe (16a) are indicated in Table 1. All the uronium type coupling reagents (HOTU, 15a), (HTOPC, 15e), and (HTOPT, 15f), used gave less than 1% racemization according to the NMR data but CDMT (9) gave about 15% of the racemized product (Table 1).
Table 1. Yield (%) and Racemization (%) of Z-Phe-Ala-OMe (16a) using the Uronium Salts 15a,e,f and CDMT (9).
Table 1. Yield (%) and Racemization (%) of Z-Phe-Ala-OMe (16a) using the Uronium Salts 15a,e,f and CDMT (9).
Coupling ReagentYield %DL %
(HOTU, 15a)100<1
(HTOPC, 15e)54<1
(HTOPT, 15f)66<1
(CDMT, 9)5515
CH3 DL (d) at δ = 1.20 ppm; CH3 LL (d) at δ = 1.32 ppm.
The other triazinyloxyimino derivative (DMTOC, 12a) failed to give the expected product Z-Phe-Ala-OMe (16a), due to the low activation for the carboxylic group and fast attack by the N-terminal of the amino acid which led to formation of the N-triazinyl amino acid derivative methyl 2-(4,6-dimethoxy-1,3,5-triazin-2-ylamino)propanoate (19a) in 77-85% yield, as confirmed by NMR data (Scheme 3).
Scheme 3. Coupling of Z-Phe-OH with H-Ala-OMe.HCl (18a) using DMTOC (12a) and CDMT (9).
Scheme 3. Coupling of Z-Phe-OH with H-Ala-OMe.HCl (18a) using DMTOC (12a) and CDMT (9).
Molecules 15 09403 g007
Another example, the coupling of Z-Phe-OH to H-Val-OMe.HCl (18b) to afford the dipeptide Z-Phe-Val-OMe (16b) showed the same behavior for the uronium salts and triazinyloxyimino derivatives (Scheme 4, Table 2).
Scheme 4. Coupling of Z-Phe-OH with H-Val-OMe.HCl (18b) using triazinyloxyimino derivatives and CDMT (9).
Scheme 4. Coupling of Z-Phe-OH with H-Val-OMe.HCl (18b) using triazinyloxyimino derivatives and CDMT (9).
Molecules 15 09403 g008
Table 2. Yield (%) and Racemization (%) of Z-Phe-Val-OMe (16b) using the Uronium Salts 15a,e,f and CDMT (9).
Table 2. Yield (%) and Racemization (%) of Z-Phe-Val-OMe (16b) using the Uronium Salts 15a,e,f and CDMT (9).
Coupling ReagentYield %DL %
(HOTU, 15a)83<1
(HTOPC, 15e)89<1
(HTOPT, 15f)68<1
(CDMT, 9)692.0
CH3 DL (dd) at δ = 0.66 ppm; CH3 LL (dd) at δ = 0.82 ppm.
The desired dipeptide product was not formed in case of the triazinyloxyimino derivatives (DMTOC, 12a), (DMTOPyC, 12c), and (DMTOPy, 12d) but rather the side product methyl 2-(4,6-dimethoxy-1,3,5-triazin-2-ylamino)-3-methylbutanoate (19b), due to the reaction of amino acid ester with the coupling reagents as indicated by NMR. For the rather non-sensitive case of segment coupling of Z-Gly-Phe-OH to H-Ala-OMe.HCl (18a), which leads to the tripeptide Z-Gly-Phe-Ala-OMe (17a), the Oxyma (ethyl 2-cyano-2-(hydroxyimino)acetate) derivatives gave the best results compared with others as observed from the NMR data (Scheme 5, Table 3).
Scheme 5. Coupling of Z-Gly-Phe-OH with H-Ala-OMe.HCl (18a) using triazinyloxyimino derivatives and CDMT (9).
Scheme 5. Coupling of Z-Gly-Phe-OH with H-Ala-OMe.HCl (18a) using triazinyloxyimino derivatives and CDMT (9).
Molecules 15 09403 g009
Table 3. Yield (%) and Racemization of (%)Z-Gly-Phe-Ala-OMe (17a) using the Uronium Salts 15a,e,f and CDMT (9).
Table 3. Yield (%) and Racemization of (%)Z-Gly-Phe-Ala-OMe (17a) using the Uronium Salts 15a,e,f and CDMT (9).
Coupling ReagentYield %DL %
(HOTU, 15a) 63<1
(HTOPC, 15e) 1919.0
(HTOPT, 15f) 7612.8
(CDMT, 9)gummy 30
CH3 DL (d) at δ = 1.19 ppm; CH3 LL (d) at δ = 1.32 ppm.
The rest of the triazinyloxyimino derivatives (DMTOC, 12a), (DMTOPyC, 12c), and (DMTOPy, 12d) did not afford the desired product, giving the side product methyl 2-(4,6-dimethoxy-1,3,5-triazin-2-ylamino)propanoate (19a) (Scheme 5) as a major product, as revealed from its NMR spectral data.
In the more sensitive cases of the formation of Z-Gly-Phe-Val-OMe (17b) and Z-Gly-Val-Val-OMe(17c) the same results were obtained. The best results obtained were when the Oxyma derivative (HOTU, 15a) was using as a coupling reagent, while CDMT (9) gave the highest racemization level and the triazine coupling reagents failed in the formation of the target product, giving the same side product 19b, which was obtained from the coupling of Z-Phe-OH with H-Val-OMe.HCl (18b) as observed from the HPLC and NMR data (Table 4 and Table 5).
Table 4. Yield (%) and Racemization of (%) Z-Gly-Phe-Val-OMe (17b) using the Uronium Salts 15a,e,f and CDMT (9).*
Table 4. Yield (%) and Racemization of (%) Z-Gly-Phe-Val-OMe (17b) using the Uronium Salts 15a,e,f and CDMT (9).*
Coupling ReagentYield %DL %
(HOTU, 15a)80< 0.1
(HTOPC, 15e)8126.19
(HTOPT, 15f)786.15
(CDMT, 9)6350.0
* The racemization percent was detected by reverse phase HPLC, using a SunFire C18 column (Waters) (5 μm, 4.6 × 150 mm), linear gradient 10-90 0.036% TFA in CH3CN/0.045% TFA in H2O over 8 min tR LL is 7.43 min, tR DL is 7.61 min as indicated from and authentic samples.
Table 5. Yield (%) and Racemization (%) of Z-Gly-Val-Val-OMe (17c) using the Uronium Salts 15a,e,f and CDMT (9).*
Table 5. Yield (%) and Racemization (%) of Z-Gly-Val-Val-OMe (17c) using the Uronium Salts 15a,e,f and CDMT (9).*
Coupling ReagentYield %DL %
(HOTU, 15a)81< 0.1
(HTOPC, 15e)≈10017.5
(HTOPT, 15f)956.66
(CDMT, 9)6047.3
* The percent racemization was detected by reverse phase HPLC, using a SunFire C18 column (Waters) (5 μm, 4.6 × 150 mm), linear gradient 10-90 0.036% TFA in CH3CN/0.045% TFA in H2O over 8 min tR LL is 6.90 min, tR DL is 7.07 min as indicated by authentic samples.

3. Experimental

3.1. General

All chemicals were used without further purification. Solvents are redistilled before use. N-protected amino acids and esters were purchased from Novabiochem. Melting points are uncorrected and were determined on a Gallenkamp hot stage. A Perkin Elmer Spectrum 1000 FT-IR Spectrometer was used for recording infrared (IR) spectra of the prepared compounds as KBr pellets or in spectroscopic grade dichloromethane. 1H- and 13C-NMR spectra of compounds were run on JEOL 400 MHz NMR spectrometer, in CD3COCD3, CDCl3 or DMSO-d6 at room temperature using TMS as internal standard. The instruments are located at King Saud University, College of Science, Chemistry Department. For analytical separations, characterization and determination of racemization, a reverse-phase Waters 2695 HPLC separation module was used (Barcelona Science Park, University of Barcelona, Spain) equipped with a SunFire C18 column (Waters, 5 μm, 4.6 × 150 mm), 10-90% linear gradient of 0.036% TFA in CH3CN/0.045% TFA in H2O and coupled to a Waters 2998 PDA-UV detector. Chromatograms were processed with Empower software. In the detection of the percent racemization as indicated by authentic samples. Peptide mass was detected by means of an HPLC-PDA system as described above, coupled to a Water Micromass ZQ mass detector, using the MassLynx 4.1 software. Elemental analyses were performed on a Perkin-Elmer 2400 elemental analyzer, and the values found were within ±0.3% of the theoretical values. Oxime derivatives 13a-c were prepared following the literature procedures [47,48].

3.2. General Synthesis of Triazine Coupling Reagents 12a-d from Oxime Derivatives

2-Chloro-4,6-dimethoxy-1,3,5-triazine (9, 20 mmol) was added to a solution of oxime (20 mmol) and triethylamine (20 mmol) in dichloromethane (DCM, 200 mL) at 0 ºC. The reaction mixture was stirred at the same temperature for 1 h and then at room temperature for 5-8 h. DCM (400 mL) was added, and then the reaction mixture was washed twice with saturated aqueous NaCl (200 mL each). Finally, the organic solvent was dried with anhydrous Na2SO4, filtered, and the solvent was removed under reduced pressure. The crude product was recrystallized from DCM/hexane.
Ethyl-2-cyano-2-(4,6-dimethoxy-1,3,5-triazin-2-yloxyimino)acetate (DMTOC, 12a). White powder, yield 4.87 g (86%), m.p. 110 ºC; IR (KBr): 2986.68 (aliph. CH), 2363.98, 2344.92 (C≡N), 1758.25 (C=O, ester), 1608.79 (C=N), 1286.58 (N-O), 1084.42 (C-O-C) cm-1; 1H-NMR (CDCl3): δ 1.44 (t, J = 7.2 Hz, 3H, CH3), 4.11 (s, 6H, 2 OCH3), 4.50 (q, J = 7.2 Hz, 2H, CH2) ppm; 13C-NMR (CDCl3): δ 14.19 (CH3), 56.36 (2 OCH3), 64.77 (CH2), 106.95 (C≡N), 131.58 (C=N), 156.88 (C=O), 172.96 (C-2), 174.18 (C-4,6) ppm; Elemental analysis, calculated for C10H11N5O5 (281.22): C, 42.71; H, 3.94; N, 24.90. Found: C, 42.46; H, 3.82; N, 24.66.
(4,6-Dimethoxy-1,3,5-triazin-2-yloxy)carbonimidoyl dicyanide (DMTODC, 12b). Snow white powder, yield 3.11 g (66%), m.p. 74 ºC; IR (KBr): 2948.10 (aliph. CH), 2363.80, 2344.87 (C≡N), 1558.96 (C=N), 1467.87 (Ar. C=N), 1361.80 (Ar. C-N), 1222.97 (N-O), 1042.73 (C-O-C) cm-1; 1H-NMR (CDCl3): δ 3.37 (s, 6H, 2 OCH3) ppm; 13C-NMR (CDCl3): δ 56.58 (2 OCH3), 105.11, 108.19 (2 C≡N), 114.80 (C=N), 172.34 (C-2), 174.21 (C-4,6) ppm; Elemental analysis, calculated for C8H6N6O3 (234.17): C, 41.03; H, 2.58; N, 35.89. Found: C, 41.16; H, 2.49; N, 35.68.
N-(4,6-Dimethoxy-1,3,5-triazin-2-yloxy)picolinimidoyl cyanide (DMTOPyC, 12c). Pink cotton solid, yield 6.29 g (≈100%) from (10 mmol of starting material), m.p. 198 ºC; IR (KBr): 3468.01 (Ar. C-H), 2924.86 (aliph. CH), 2364.64, 2344.88 (C≡N), 1606.27 (C=N, C=C), 1560.18 (asym. C-O-C), 1361.76 (sym. C-O-C) cm-1; 1H-NMR (CDCl3): δ 4.13 (s, 6H, 2 OCH3), 7.52 (m, 1H, H-5'), 7.86 (dt, 1H, H-4'), 8.22 (d, J = 7.32 Hz, 1H, H-3'), 8.82 (d, J = 4.4 Hz, 1H, H-6') ppm; 13C-NMR (CDCl3): δ 56.11 (2 OCH3), 108.21 (C≡N), 122.26 (C-3'), 126.74 (C-5'), 137.21 (C-4'), 139.89 (C-6'), 147.42 (C-2'), 150.43 (C=N), 173.26 (C-2), 174.04 (C-4,6) ppm; Elemental analysis, calculated for C12H10N6O3 (286.25): C, 50.35; H, 3.52; N, 29.36. Found: C, 50.59; H, 3.42; N, 29.29.
1-(4,6-Dimethoxy-1,3,5-triazin-2-yloxy)pyridin-2(1H)-one (DMTOPy, 12d). The product was obtained as a light tan powder, yield 4.28 g (85%), m.p. (104-106 ºC); IR (KBr): 3091.57 (Ar. C-H), 2953.80 (aliph. CH), 1678.35 (C=O), 1608.62 (C=N, C=C), 1558.30 (N-O), 1369.99, 1349.16 (Ar. C-N), 1280.22 (asym. C-O-C), 1084.36 (sym. C-O-C) cm-1; 1H-NMR (CDCl3): δ 3.96 (s, 6H, 2 OCH3), 6.25 (dt, 1H, H-5'), 6.75 (dd, 1H, H-3'), 7.42 (1H, H-4'), 7.53 (dd, H-6') ppm; 13C-NMR (CDCl3): δ 56.10 (2 OCH3), 105.39 (C-5'), 123.27 (C-3'), 135.45 (C-6'), 139.79 (C-4'), 157.23 (C=O), 173.96 (C-2), 174.11 (C-4,6) ppm; Elemental analysis, calculated for C10H10N4O4 (250.21): C, 48.00; H, 4.03; N, 22.39. Found: C, 48.22; H, 4.09; N, 22.45.

3.3. General Synthesis of Oximo-Uronium Type Coupling Reagents 15a-f

To a solution of the oxime potassium salts of 13a-c (20 mmol) in acetonitrile (ACN) (50 mL) was added to the chloro salts 14a,b (20 mmol) at 0 ºC. The reaction mixture was stirred at this temperature 30 min and stirred at room temperature for 6 h. Filter and wash with acetonitrile. The solvent was concentrated to small volume (1/4) under reduced pressure, and then dry ether was added to afford the product as a white solid in pure state [46].
O-[(Cyano(ethoxycarbonyl)methylidene)amino]-1,1,3,3-tetramethyluronium hexafluorophosphate (HOTU, 15a). Yield 6.32g (82%) using the potassium salt of the oxime, m.p. (135-137 ºC) (dec). The triethylamine/oxime combinations gave a lower yield (69%); IR (KBr): 3003.32 (N-CH3), 2964.16 (aliph. CH), 2345.36 (C≡N, =N+), 1771.32 (C=O), 1702.40 (C=N), 1266.93 (N-O) cm-1; 1H-NMR (acetone-d6): δ 1.37 (t, 3H, CH3), 3.37 (s, 12H, 4 CH3), 4.82 (q, 2H, CH2) ppm; 13C-NMR (acetone-d6): δ 13.46 (CH3), 40.71 (4 CH3), 64.56 (CH2), 106.78 (C≡N), 135.09 (C=N), 156.11 (C=O), 161.43 (C+) ppm; Elemental Analysis, calculated for C10H17F6N4O3P (386.23): C, 31.10; H, 4.44; N, 14.51. Found: C, 31.33; H, 4.40; N, 14.36.
1-[(1-(Cyano-2-ethoxy-2-oxoethylideneaminooxy)dimethylaminomorpholinomethylene)]methanamin- ium hexafluorophosphate (COMU, 15b). Yield 7.6g (89%), m.p. (143-144 ºC); 1H-NMR (acetone-d6): δ 1.38 (t, 3H, CH3), 3.41 (s, 6H, 2 CH3), 3.82 (t, 4H, 2 CH2), 3.89 (t, 4H, 2 CH2), 4.48 (q, 2H, CH2) ppm; 13C-NMR (acetone-d6): δ 13.48 (CH3), 40.70 (2 CH3), 49.94 (2 CH2), 64.59 (2 CH2), 66.04 (CH2), 106.76 (C≡N), 135.03 (C=N), 156.14 (C=O), 160.61 (C+) ppm; Elemental Analysis, calculated for C12H19F6N4O4P (428.27): C, 33.65; H, 4.47; N, 13.08. Found: C, 33.92; H, 4.41; N, 13.24.
O-[(Dicyanomethylidene)amino]-1,1,3,3-tetramethyluronium hexafluorophosphate (HTODC, 15c). Yield 5.0 g (74%), m.p. (180-181 ºC) (dec); 1H-NMR (acetone-d6): δ 3.27 (s, 12H, 4 CH3) ppm; 13C-NMR (acetone-d6): δ 40.80 (4 CH3), 105.10 (C≡N), 108.21 (C≡N), 119.65 (C=N), 160.67 (C+) ppm; Elemental Analysis, calculated for C8H12F6N5OP (339.18): C, 28.33; H, 3.57; N, 20.65. Found: C, 28.56; H, 3.64; N, 20.80.
1-[(1-(Dicyanomethyleneaminooxy)dimethylaminomorpholinomethylene)]methanaminium hexafluoro- phosphate (HDMODC, 15d). Yield 5.7 g (75%), m.p. (118-119 ºC); 1H-NMR (acetone-d6): δ 3.41 (s, 6H, 2 CH3), 3.87-3.88 (m, 8H, 4 CH2) ppm; 13C-NMR (acetone-d6): δ 40.97 (2 CH3), 49.93 (2 CH2), 65.92 (2 CH2), 105.13 (C≡N), 108.15 (C≡N), 119.84 (C=N), 159.77 (C+) ppm; Elemental Analysis, calculated for C10H14F6N5O2P (381.21): C, 31.51; H, 3.70; N, 18.37. Found: C, 31.77; H, 3.56; N, 18.52.
N-[(Cyano(pyridine-2-yl)methyleneaminooxy)(dimethylamino)methylene]-N-methylmethanaminium hexafluorophosphate (HTOPC, 15e). Yield 6.2 g (83%), m.p. (169-171 ºC); IR (KBr): 3069.24 (N-CH3), 2966.06 (aliph. CH), 2345.51 (C≡N, =N+), 1692.23 (C=N, C=C), 1273.86 (N-O) cm-1; 1H-NMR (DMSO-d6): δ 3.21 (s, 12H, 4 CH3), 7.74 (t, 1H, H-5), 8.07 (t, 1H, H-4), 8.14 (d, J = 8.08 Hz, 1H, H-3), 8.85(d, J = 5.12 Hz, 1H, H-6) ppm; 13C-NMR (DMSO-d6): δ 40.27 (4 CH3), 108.54 (C≡N), 122.74 (C-3), 128.23 (C-5), 138.62 (C-4), 142.07 (C-2), 146.70 (C-6), 150.88 (C=N), 161.17 (C+) ppm; Elemental Analysis, calculated for C12H16F6N5OP (391.21): C, 36.84; H, 4.12; N, 17.90. Found: C, 37.06; H, 4.22; N, 18.12.
O-(2-Pyridone)-1,1,3,3-tetramethyluronium hexafluorophosphate (HTOPT, 15f). Yield 6.4 g (82%), m.p. (166-176 ºC); IR (KBr): 3096.77 (N-CH3), 2958.98 (aliph. CH), 2345.56 (=N+), 1709.64 (C=O), 1672.42 (C=C, C=N), 1272.58 (N-O) cm-1; 1H-NMR (DMSO-d6): δ 3.11 (s, 12H, 4 CH3), 6.52 (t, J = 7.36 Hz, 1H, H-5), 6.78 (d, J = 8.8, 1H, H-3), 7.65 (t, J = 7.32 Hz, 1H, H-4), 8.42 (d, J = 7.32 Hz, 1H, H-6) ppm; 13C-NMR (DMSO-d6): δ 40.27 (4 CH3), 107.02 (C-5), 121.98 (C-3), 135.90 (C-4), 142.05 (C-6), 156.56 (C=O), 162.12 (C+) ppm; Elemental Analysis, calculated for C14H18F6N5O2P (391.21), C, 38.81; H, 4.19; N, 16.16. Found: C, 39.03; H, 4.31; N, 16.34.

3.4. Synthesis of 16a,b (1+1)

The coupling reagents 15a,e,f (0.25 mmol) were added to a mixture of Z-Phe-OH (0.25 mmol), the appropriate amino acid 18a,b (0.25 mmol) and N-methylmorpholine (0.75 mmol), in dry acetonitrile (5 mL) at 0 ºC. The reaction mixture was stirred at the same temperature for 1 h and then at room temperature for 2 h. Ethyl acetate (EtOAc, 50 mL) was added and the mixture was subsequently washed with 10% aqueous HCl (v/v), saturated aqueous Na2CO3 and saturated aqueous NaCl solution (2 × 10 mL each). Finally, the organic solvent was dried with anhydrous Na2SO4, filtered, and the solvent was removed under reduced pressure. The gummy residue obtained 16a,b (DL < 1%) were dried under vacuum. Both products were obtaind when the coupling chlorotriazine derivative 9 was used (15% racemization). Using the triazine coupling reagent 12a,c,d instead of 15a,e,f and following the different procedure whereby Z-Phe-OH was activated with 12a,c,d for 1 hour at room temperature and then added to a solution of 18a,b in the presence of NMM as a base in acetonitrile, the reaction mixture was stirred at the same temperature for 24 h gave compounds 19a,b.
Z-Phe-Ala-OMe (16a). IR (KBr): 3303.71 (NH), 2951.65, 2926.12 (aliph. CH), 1743.34 (C=O, ester), 1694.14 (C=O, urethane), 1652.91 (C=O, amide) cm-1; 1H-NMR (CDCl3): δ 1.31 (d, 3H, CH3), 3.06 (d, 2H, CH2-C), 3.72 (s, 3H, OCH3), 4.47-4.51 (m, 2H, 2 CH), 5.07 (s, 2H, CH2-O), 5.38 (1H, NH, amide), 6.42 (1H, NH, urethane), 7.25-7.32 (m, 10H, Ar. H) ppm; 13C-NMR (CDCl3): δ 18.36 (CH3), 38.64 (CH2-C), 48.23 (CH2-O), 52.57 (OCH3), 56.10 (CH, Phe), 67.16 (CH, Ala), 129.43 (C=O, amide), 170.43 (C=O, urethane), 172.88 (C=O, ester) ppm.
Z-Phe-Val-OMe (16b). IR (NaCl/DCM): 3310.29 (NH), 2963.93 (aliph. CH), 1741.07 (C=O, ester), 1704.99 (C=O, urethane), 1661.22 (C=O, amide) cm-1; 1H-NMR (CDCl3): δ 0.80-0.84 (dd, 6H, 2 CH3), 2.02-2.12 (m, 1H, CH(CH3)2), 3.05-3.07 (m, 2H, CH2-C), 3.68 (s, 3H, OCH3), 4.45-4.46 (m, 2H, 2 CH), 5.08 (s, 2H, CH2-O), 5.42 (1H, NH, amide), 6.36 (1H, NH, urethane), 7.24-7.32 (m, 10H, Ar. H) ppm.

3.5. General Method for Synthesis of 17a-c (2+1) Using Different Coupling Reagents

The coupling reagents 15a,e,f (0.25 mmol) were added to a mixture of Z-Gly-Phe-OH (0.25 mmol), 18a,b (0.25 mmol) and N-methylmorpholine (0.75 mmol), in dry ACN (5 mL) at 0 ºC. The reaction mixture was stirred at the same temperature for 1 h and then at room temperature for 24 h. Ethyl acetate (EtOAc, 50 mL) was added and the mixture was subsequently washed with 10% aqueous HCl (v/v), saturated aqueous Na2CO3 and saturated aqueous NaCl solution (2 × 10 mL each). Finally, the organic solvent was dried with anhydrous Na2SO4, filtered, and the solvent was removed under reduced pressure. The gummy residue obtained 17a,b (DL < 1% in the case of using 15a) was dried under vacuum. Both products were obtained when the coupling chlorotriazine 9 was used. Using the triazine coupling reagent 12a,c,d instead of 15a,e,f and following the different procedure whereby Z-Gly-Phe-OH was activated with 12a,c,d for 1 h at 0 ºC and then added to a solution of 18a,b in the presence of NMM as a base in acetonitrile, the reaction mixture was stirred at the same temperature for 1 h, then at room temperature for 24 h gave compounds 19a,b.
Z-Gly-Phe-Ala-OMe (17a). IR (NaCl/DCM): 3434.81, 3272.55 (NH), 2950.43 (aliph. CH), 1735.95 (C=O, ester), 1718.13 (C=O, urethane), 1685.96, 1647.73 (2 C=O, amide) cm-1; 1H-NMR (CDCl3): δ 1.30 (d, 3H, CH3), 3.03 (d, 2H, CH2-N), 3.67 (s, 3H, OCH3), 3.79-3.85 (m, 2H, CH2-C), 4.43-4.46 (m, 1H, CH, Phe), 4.68-4.70 (m, 1H, CH, Ala), 5.08 (s, 2H, CH2-O), 5.62, 6.71 (2H, 2 NH, amide), 6.91 (1H, NH, urethane), 7.14-7.32 (m, 10H, Ar. H) ppm.
Z-Gly-Phe-Val-OMe (17b). IR (NaCl/DCM): 3298.25 (br. NH), 2961.81 (aliph. CH), 1719.76 (br. C=O, ester & urethane), 1654.46 (2 C=O, amide) cm-1; 1H-NMR (CDCl3: δ 0.77-0.82 (dd, 6H, 2 CH3), 2.02-2.12 (m, 1H, CH(CH3)2), 2.90-3.10 (m, 2H, CH2-N), 3.62 (s, 3H, OCH3), 3.60-3.90 (m, 2H, CH2-C), 4.35-4.49 (m, 1H, CH, Gly), 4.75-4.85 (m, 1H, CH, Val), 5.06 (s, 2H, CH2-O), 5.85, 6.95 (2H, 2 NH, amide), 7.12-7.20 (m, 1H, NH, urethane), 7.25-7.36 (m, 10H, Ar. H) ppm.
Z-Gly-Val-Val-OMe (17c). Prepared following the same procedure above for 17a,b by replacing Z-Gly-Phe-OH with Z-Gly-Val-OH, and the amino acid 18b was used. The gummy residue of 17c obtained when the coupling reagents 15a,e,f (DL < 0.1% in the case of using 15a) was dried under vacuum. The product was obtained (47.3% racemization) when the coupling chlorotriazine 9 was used. Using the triazine coupling reagent 12a,c,d instead of 15a,e,f and following the different procedure whereby Z-Gly-Val-OH was activated with 12a,c,d for 1 hour at 0 ºC and then added to a solution of 18b in the presence of NMM as a base in acetonitrile, the reaction mixture was stirred at room temperature for 24 h to give compound 19b.
Z-Gly-Val-Val-OMe (17c). IR (NaCl/DCM): 3411.60, 3303.57 (NH), 2967.43 (aliph. CH), 1733.86 (br. C=O, ester & urethane), 1651.45 (2 C=O, amide) cm-1; 1H-NMR (CDCl3): δ 0.88-0.90 (dd, 12H, 4 CH3), 2.06-2.14 (m, 2H, 2 CH(CH3)2), 3.70 (s, 3H, OCH3), 3.90 (2H, CH2-N), 4.38 (1H, CH, Gly), 4.48-4.51 (m, 1H, CH, Val), 5.10 (s, 2H, CH2-O), 5.68, 6.74 (2H, 2 NH, amide), 6.82 (1H, NH, urethane), 7.25-7.34 (m, 5H, Ar. H) ppm.
Methyl-2-(4,6-dimethoxy-1,3,5-triazin-2-ylamino)propanoate (19a). m.p. 98-100 ºC, 1H-NMR (CDCl3): δ 1.48 (d, 3H, CH3), 3.72 (s, 3H, CH3 ester), 3.93 (s, 6H, 2 OCH3), 4.70 (q, 1H, CH), 5.93 (d, 1H, NH) ppm; Elemental analysis calculated for C9H14N4O4: C, 44.63; H, 5.83; N, 23.13. Found: C, 44.91; H, 6.00; N, 23.36.
Methyl-2-(4,6-dimethoxy-1,3,5-triazin-2-ylamino)-3-methylbutanoate (19b). m.p. 103-105 ºC, 1H-NMR (CDCl3): δ 0.96 (t, 6H, 2 CH3), 2.21-2.22 (m, 1H, CH), 3.72 (s, 3H, CH3 ester), 3.91 (s, 6H, 2 OCH3), 4.67 (q, 1H, CH), 5.84 (d, 1H, NH) ppm; Elemental analysis calculated for C11H18N4O4: C, 48.88; H, 6.71; N, 20.73. Found: C, 49.12; H, 6.89; N, 21.01.

4. Conclusions

In conclusion, herein a new family of 1,3,5-triazinyloxyimino derivatives 12a-d was compared with uronium salts. The morpholino and tetramethyluronium salts are more reactive than triazinyloxyimino derivatives for activation of the carboxylic group and formation of peptide bonds. The new triazinyloxyimino derivatives 12a-d failed to give the expected products, due to the low activation for the carboxylic group and fast attack by the N-terminal of the amino acid, which led to formation of the corresponding N-triazinyl amnio acid derivatives as confirmed by NMR and HPLC. Oxyma (ethyl 2-cyano-2-(hydroxyimino)acetate) uronium salts were confirmed to be superior to other oxime derivatives, HOAt (4) and HOBt (1), in terms of both coupling yield and retention of configuration. Finally, the new triazinyloxyimino derivatives may be useful in the preparation of biologically active N-triazinyl amino acids as well as N-triazinyl peptides which is now going to be tested in our laboratory.

Acknowledgements

We are grateful to Research Centre at College of Science, King Saud University, for supporting this work.

References

  1. Rich, D.H.; Singh, J. The Peptides: Analysis, Biology; Gross, E., Meienhofer, J., Eds.; Academic Press: New York, NY, USA, 1979; Volume 1, pp. 241–261. [Google Scholar]
  2. Knorr, R.; Trzeciak, A.; Bannwarth, W.; Gillessen, D. New coupling reagents in peptide chemistry. Tetrahedron Lett. 1989, 30, 1927–1930. [Google Scholar]
  3. König, W.; Geiger, R. A new method for the synthesis of peptides: activation of the carboxyl group with dicyclohexylcarbodiimide and 3-hydroxy-4-oxo-3,4-dihydro-1,2,3-benzotriazine. Chem. Ber. 1970, 103, 2034–2040. [Google Scholar] [CrossRef]
  4. König, W.; Geiger, R. A new method for the synthesis of peptides: activation of the carboxyl group with dicyclohexylcarbodiimide using 1-hydroxy-benzotriazoles as additives. Chem. Ber. 1970, 103, 788–789. [Google Scholar] [CrossRef]
  5. Carpino, L.A. 1-Hydroxy-7-azabenzotriazole. An efficient peptide coupling additive. J. Am. Chem. Soc. 1993, 115, 4397–4398. [Google Scholar] [CrossRef]
  6. Carpino, L.A.; Imazumi, H.; El-Faham, A.; Ferrer, F.J.; Zhang, C.; Lee, Y.; Foxman, B.M.; Henklein, P.; Hanay, C.; Gge, C.M.; Wenschuh, H.; Klose, J.; Beyermann, M.; Bienert, M. The Uronium/ Guanidinium Peptide Coupling Reagents: Finally the True Uronium Salts. Angew. Chem. Int. Ed. 2002, 41, 441–445. [Google Scholar]
  7. Albericio, F.; Bofill, J.M.; El-Faham, A.; Kates, S.A. Use of Onium Salt-Based Coupling Reagents in Peptide Synthesis. J. Org. Chem. 1998, 63, 9678–9683. [Google Scholar] [CrossRef]
  8. Wijkmans, J.C.H.M.; Kruijtzer, J.A.W.; van der Marel, G.A.; van Boom, J.H.; Bloemhoff, W. CF3-NO2-PyBOP: A new and highly efficient coupling reagent for N-methyl amino acids. Tetrahedron Lett. 1995, 36, 4643–4646. [Google Scholar]
  9. Kehler, J.; Püsch, A.; Dahl, O. Improved Syntheses of 1-Hydroxy-4-nitro-6-trifluoromethyl- benzotriazole and 1-Hydroxy-4,6-dinitrobenzotriazole. Acta Chem. Scan. 1996, 50, 1171–1173. [Google Scholar] [CrossRef]
  10. Carpino, L.A.; El-Faham, A.; Albericio, F. Racemization Studies During Solid Phase Peptide Synthesis Using Azabenzotriazole-Based Coupling Reagents. Tetrahedron Lett. 1994, 35, 2279–2282. [Google Scholar] [CrossRef]
  11. Carpino, L.A.; El-Faham, A.; Minor, C.H.; Albericio, F. Advantageous applications of azabenzotriazole (triazolopyridine)-based coupling reagents to solid-phase peptide synthesis. J. Chem. Soc., Chem. Commun 1994, 201–203. [Google Scholar]
  12. Quibell, M.; Packman, L.C.; Johnson, T. Identification of coupling conditions proceeding with low C-terminal epimerization during the assembly of fully protected backbone-substituted peptide segments. J. Chem. Soc. Perkin. Trans. 1996, 1, 1219–1225. [Google Scholar]
  13. Carpino, L.A.; El-Faham, A. The Diisopropylcarbodiimide/1-Hydroxy-7-azabenzotriazole System: Segment Coupling and Stepwise Peptide Assembly. Tetrahedron 1999, 55, 6813–6830. [Google Scholar] [CrossRef]
  14. Carpino, L.A.; Imazumi, H.; Foxman, B.M.; Vela, M.J.; Henklein, P.; El-Faham, A.; Klose, J.; Bienert, M. Comparison of the Effects of 5- and 6-HOAt on Model Peptide Coupling Reactions Relative to the Cases for the 4- and 7-Isomers. Org. Lett. 2000, 2, 2253–2256. [Google Scholar] [CrossRef]
  15. Angell, Y.M.; Gacia, C.E.; Rich, D.H. Comparative studies of the coupling of N-methylated, sterically hindered amino acids during solid-phase peptide synthesis. Tetrahedron Lett. 1994, 35, 5981–5984. [Google Scholar]
  16. Gawne, G.; Kenner, G.; Sheppard, R.C. Acyloxyphosphonium salts as acylating agents. Synthesis of peptides. J. Am. Chem. Soc. 1969, 91, 5669–5671. [Google Scholar]
  17. Castro, B.; Dormoy, J.R. Azidotris(dimethylamino)phosphonium hexafluorophosphate. Excellent reagent for peptide coupling. Bull. Soc. Chim. Fr. 1973, 3359–3361. [Google Scholar]
  18. Castro, B.; Dormoy, J.R.; Evin, G.; Selve, C. Peptide coupling reagents. IV. N-[Oxytris(dimethylamino)phosphonium]benzotriazole hexafluorophosphate. Tetrahedron Lett. 1975, 14, 1219–1222. [Google Scholar]
  19. Castro, B.; Dormoy, J.R.; Evin, G.; Selve, C. Peptide coupling reagents. Part VII. Mechanism of the formation of active esters of hydroxybenzotriazole in the reaction of carboxylate ions on the BOP reagent for peptide coupling. A comparison with Itoh's reagent. J. Chem. Res. (S) 1977, 182–185. [Google Scholar]
  20. Coste, J.; Le-Nguyen, D.; Castro, B. PyBOP: A new peptide coupling reagent devoid of toxic by-product. Tetrahedron Lett. 1990, 31, 205–208. [Google Scholar]
  21. Dykstra, R.R. Encyclopedia of Reagents for Organic Synthesis; Paquette, L.A., Ed.; John Wiley & Sons Ltd.: Chichester, UK, 1995; Volume 4, pp. 2668–2675. [Google Scholar]
  22. Carpino, L.A.; El-Faham, A. Effect of Tertiary Bases on O-Benzotriazolyluronium Salt-Induced Peptide Segment Coupling. J. Org. Chem. 1994, 59, 695–698. [Google Scholar] [CrossRef]
  23. Carpino, L.A.; Ionescu, D.; El-Faham, A. Peptide Coupling in the Presence of Highly Hindered Tertiary Amines. J. Org. Chem. 1996, 61, 2460–2465. [Google Scholar] [CrossRef]
  24. Kates, S.A.; Diekmann, E.; El-Faham, A.; Herman, L.W.; Ionescu, D.; McGuinness, B.F.; Triolo, S.A.; Albericio, F.; Carpino, L.A. Techniques in Protein Chemistry VII; Marshak, D.R., Ed.; Academic Press: New York, NY, USA, 1996; pp. 515–523. [Google Scholar]
  25. Ehrlich, A.; Heyne, H.U.; Winter, R.; Beyermann, M.; Haber, H.; Carpino, L.A.; Bienert, M. Cyclization of all-L-Pentapeptides by Means of 1-Hydroxy-7-azabenzotriazole-Derived Uronium and Phosphonium Reagents. J. Org. Chem. 1996, 61, 8831–8838. [Google Scholar] [CrossRef]
  26. Jou, G.; Gozalez, I.; Albericio, F.; Lloyd, P.W.; Giralt, E. Total Synthesis of Dehydrodidemnin B. Use of Uronium and Phosphonium Salt Coupling Reagents in Peptide Synthesis in Solution. J. Org. Chem. 1997, 62, 354–366. [Google Scholar] [CrossRef]
  27. Han, Y.; Albericio, F.; Barany, G. Occurrence and Minimization of Cysteine Racemization during Stepwise Solid-Phase Peptide Synthesis. J. Org. Chem. 1997, 62, 4307–4312. [Google Scholar] [CrossRef]
  28. Albericio, F.; Cases, M.; Alsina, J.; Triolo, S.A.; Carpino, L.A.; Kates, S.A. On the use of PyAOP, a phosphonium salt derived from HOAt, in solid-phase peptide synthesis. Tetrahedron Lett. 1997, 38, 4853–4856. [Google Scholar] [CrossRef]
  29. Kim, S.; Chang, A.; Ko, Y.K. Benzotriazol-1-yl diethyl phosphate. A new convenient coupling reagent for the synthesis of amides and peptides. Tetrahedron Lett. 1985, 26, 1341–1349. [Google Scholar] [CrossRef]
  30. Fan, C.X.; Hao, X.L.; Ye, Y.H. A novel organophosphorus compound as a coupling reagent for peptide synthesis. Synth. Commun 1996, 26, 1455–1460. [Google Scholar] [CrossRef]
  31. Carpino, L.A.; Xia, J.; El-Faham, A. 3-Hydroxy-4-oxo-3,4-dihydro-5-azabenzo-1,2,3-triazene. J. Org. Chem. 2004, 69, 54–61. [Google Scholar] [CrossRef]
  32. Carpino, L.A.; Xia, J.; Zhang, C.; El-Faham, A. Organophoshorus and Nitro-substituted Sulfone Esters of 1-Hydroxy-7-azabenzotriazole as Highly Efficient Fast-Acting Peptide Coupling Reagents. J. Org. Chem. 2004, 69, 69–71. [Google Scholar]
  33. Carpino, L.A.; El-Faham, A.; Albericio, F. Efficiency in Peptide Coupling: 1-Hydroxy-7-azabenzotriazole vs. 3,4-Dihydro-3-hydroxy-4-oxo-1,2,3-benzotriazine. J. Org. Chem. 1995, 60, 3561–3564. [Google Scholar] [CrossRef]
  34. El-Faham, A.; Khattab, Sh.N.; Abdul-Ghani, M.; Albericio, F. Design and Synthesis of New Immonium-Type Coupling Reagents. Eur. J. Org. Chem. 2006, 6, 1563–1573. [Google Scholar]
  35. El-Faham, A.; Albericio, F. Novel Proton Acceptor Immonium-Type Coupling Reagents: Application in Solution and Solid Phase Peptide Synthesis. Org. Lett. 2007, 9, 4475–4477. [Google Scholar] [CrossRef]
  36. El-Faham, A.; Albericio, F. Morpholine-Based Immonium and Halogeno amidinium Salts as Coupling Reagents in Peptide Synthesis. J. Org. Chem. 2008, 73, 2731–2737. [Google Scholar] [CrossRef]
  37. Kamiński, Z.J. Triazine-based condensing reagents. Biopolymers (Peptide Science) 2000, 55, 140–164. [Google Scholar] [CrossRef]
  38. Kamiński, Z.J.; Paneth, P.; Rudzinski, J. A Study on the Activation of Carboxylic Acids by Means of 2-Chloro-4,6-dimethoxy-1,3,5-triazine and 2-Chloro-4,6-diphenoxy-1,3,5-triazine. J. Org. Chem. 1998, 63, 4248–4255. [Google Scholar] [CrossRef]
  39. Kunishima, M.; Kawachi, C.; Iwasaki, F.; Terao, K.; Tani, S. Synthesis and characterization of 4-(4,6-dimethoxy-1,3,5-triazin-2-yl)-4-methylmorpholinium chloride. Tetrahedron Lett. 1999, 40, 5327–5330. [Google Scholar] [CrossRef]
  40. Kunishima, M.; Kawachi, C.; Morita, J.; Hioki, K.; Terao, K.; Tani, S. Formation of carboxamides by direct condensation of carboxylic acids and amines in alcohols using a new alcohol- and water-soluble condensing agent: DMT-MM. Tetrahedron 2001, 57, 1551–1558. [Google Scholar] [CrossRef]
  41. Kjell, D.P.; Hallberg, D.W.; Kalbfleisch, J.M.; McCurry, C.K.; Semo, M.J.; Sheldo, E.M.; Spitler, J.T.; Wang, M. Determination of the Source of the N-Methyl Impurity in the Synthesis of Pemetrexed Disodium Heptahydrate. Org. Proc. Res. Dev. 2005, 9, 738–742. [Google Scholar] [CrossRef]
  42. Kunishima, M.; Kawachi, C.; Morita, J.; Terao, K.; Iwasaki, F.; Tani, S. 4-(4,6-dimethoxy-1,3,5-triazin-2-yl)-4-methyl-morpholinium chloride: an efficient condensing agent leading to the formation of amides and esters. Tetrahedron 1999, 55, 13159–13170. [Google Scholar] [CrossRef]
  43. Tanaka, T.; Noguchi, M.; Kobayashi, A.; Shoda, S.I. A novel glycosyl donor for chemo-enzymatic oligosaccharide synthesis: 4,6-dimethoxy-1,3,5-triazin-2-yl glycoside. Chem. Commun 2008, 2016–2018. [Google Scholar]
  44. Paoline, I.; Nuti, F.; de la Cruz Pozo Carrero, M.; Barbetti, F.; Kolesinska, B.; Kamiński, Z.J.; Chelli, M.; Papini, A.M. A convenient microwave-assisted synthesis of N-glycosyl amino acids. Tetrahedron Lett. 2007, 48, 2901–2904. [Google Scholar]
  45. Wozniak, L.A.; Gora, M.; Stec, W.J. Chemoselective Activation of Nucleoside 3'-O-Methylphosphonothioates with 1,3,5-Triazinyl Morpholinium Salts. J. Org. Chem. 2007, 72, 8584–8587. [Google Scholar] [CrossRef]
  46. El-Faham, A.; Subirós-Funosas, R.; Prohens, R.; Albericio, F. COMU, a Safer and More Effective Replacement for Benzotriazole-based Uronium Coupling Reagents. Chem. Eur. J. 2009, 15, 9404–9416. [Google Scholar] [CrossRef]
  47. Kitamura, M.; Chiba, S.; Narasaka, K. Synthesis of Primary Amines and N-Methylamines by the Electrophilic Amination of Grignard Reagents with 2-Imidazolidinone O-Sulfonyloxime. Bull. Chem. Soc. Jpn. 2003, 76, 1063–1070. [Google Scholar] [CrossRef]
  48. Izdebski, J. New Reagents Suppressing Racemization in Peptide Synthesis by the DCC Method. Polish J. Chem. 1979, 53, 1049–1057. [Google Scholar]
  • Sample Availability: Samples of the compounds triazinyloximino derivatives are available from the authors.

Share and Cite

MDPI and ACS Style

Al-Warhi, T.I.; AL-Hazimi, H.M.A.; El-Faham, A.; Albericio, F. Synthesis of 2-(4,6-Dimethoxy-1,3,5-triazin-2-yloxyimino) Derivatives: Application in Solution Peptide Synthesis. Molecules 2010, 15, 9403-9417. https://doi.org/10.3390/molecules15129403

AMA Style

Al-Warhi TI, AL-Hazimi HMA, El-Faham A, Albericio F. Synthesis of 2-(4,6-Dimethoxy-1,3,5-triazin-2-yloxyimino) Derivatives: Application in Solution Peptide Synthesis. Molecules. 2010; 15(12):9403-9417. https://doi.org/10.3390/molecules15129403

Chicago/Turabian Style

Al-Warhi, Tarfah I., Hassan M.A. AL-Hazimi, Ayman El-Faham, and Fernando Albericio. 2010. "Synthesis of 2-(4,6-Dimethoxy-1,3,5-triazin-2-yloxyimino) Derivatives: Application in Solution Peptide Synthesis" Molecules 15, no. 12: 9403-9417. https://doi.org/10.3390/molecules15129403

Article Metrics

Back to TopTop