Home Cart 0 Sign in  
X

[ CAS No. 100-55-0 ] {[proInfo.proName]}

,{[proInfo.pro_purity]}
Cat. No.: {[proInfo.prAm]}
3d Animation Molecule Structure of 100-55-0
Chemical Structure| 100-55-0
Chemical Structure| 100-55-0
Structure of 100-55-0 * Storage: {[proInfo.prStorage]}
Cart0 Add to My Favorites Add to My Favorites Bulk Inquiry Inquiry Add To Cart

Quality Control of [ 100-55-0 ]

Related Doc. of [ 100-55-0 ]

Alternatived Products of [ 100-55-0 ]

Product Details of [ 100-55-0 ]

CAS No. :100-55-0 MDL No. :MFCD00006407
Formula : C6H7NO Boiling Point : -
Linear Structure Formula :- InChI Key :MVQVNTPHUGQQHK-UHFFFAOYSA-N
M.W : 109.13 Pubchem ID :7510
Synonyms :
Chemical Name :3-Pyridinemethanol

Calculated chemistry of [ 100-55-0 ]

Physicochemical Properties

Num. heavy atoms : 8
Num. arom. heavy atoms : 6
Fraction Csp3 : 0.17
Num. rotatable bonds : 1
Num. H-bond acceptors : 2.0
Num. H-bond donors : 1.0
Molar Refractivity : 30.36
TPSA : 33.12 Ų

Pharmacokinetics

GI absorption : High
BBB permeant : No
P-gp substrate : No
CYP1A2 inhibitor : No
CYP2C19 inhibitor : No
CYP2C9 inhibitor : No
CYP2D6 inhibitor : No
CYP3A4 inhibitor : No
Log Kp (skin permeation) : -6.98 cm/s

Lipophilicity

Log Po/w (iLOGP) : 1.02
Log Po/w (XLOGP3) : -0.02
Log Po/w (WLOGP) : 0.42
Log Po/w (MLOGP) : -0.14
Log Po/w (SILICOS-IT) : 1.25
Consensus Log Po/w : 0.51

Druglikeness

Lipinski : 0.0
Ghose : None
Veber : 0.0
Egan : 0.0
Muegge : 1.0
Bioavailability Score : 0.55

Water Solubility

Log S (ESOL) : -0.99
Solubility : 11.1 mg/ml ; 0.102 mol/l
Class : Very soluble
Log S (Ali) : -0.23
Solubility : 64.9 mg/ml ; 0.594 mol/l
Class : Very soluble
Log S (SILICOS-IT) : -1.79
Solubility : 1.77 mg/ml ; 0.0162 mol/l
Class : Soluble

Medicinal Chemistry

PAINS : 0.0 alert
Brenk : 0.0 alert
Leadlikeness : 1.0
Synthetic accessibility : 1.0

Safety of [ 100-55-0 ]

Signal Word:Warning Class:N/A
Precautionary Statements:P261-P305+P351+P338 UN#:N/A
Hazard Statements:H315-H319-H335 Packing Group:N/A
GHS Pictogram:

Application In Synthesis of [ 100-55-0 ]

* All experimental methods are cited from the reference, please refer to the original source for details. We do not guarantee the accuracy of the content in the reference.

  • Upstream synthesis route of [ 100-55-0 ]
  • Downstream synthetic route of [ 100-55-0 ]

[ 100-55-0 ] Synthesis Path-Upstream   1~50

  • 1
  • [ 614-18-6 ]
  • [ 100-55-0 ]
YieldReaction ConditionsOperation in experiment
87.4% With lithium aluminium tetrahydride In tetrahydrofuran at 20℃; for 4 h; Tetrahydro Lithium aluminum (1.24g, 47mmol) dissolved in anhydrous tetrahydrofuran (20mL). After 0.5h at reflux, cooled to room temperature to, and then added dropwise nicotinic acid ethyl ester (2.40g, 16mmol) in anhydrous tetrahydrofuran (15mL ) solution. Was stirred at room temperature for 4h, the end of the reaction was detected by TLC. 50mL of saturated aqueous ammonium chloride solution was added under ice-cooling, remove the excess of lithium aluminum tetrahydride. Filtered off with suction, the filter cake (10mL × 3) and washed with ethyl acetate. (30mL × 3) The mother liquor was extracted with ethyl acetate, the combined ethyl acetate phases were dried over anhydrous magnesium sulfate, filtered and concentrated by column chromatography to give a colorless oil (1.52 g of, yield: 87.4percent).
Reference: [1] Patent: CN107459480, 2017, A, . Location in patent: Paragraph 0119-0120
[2] Journal of Organometallic Chemistry, 1982, vol. 224, # 1, p. 21 - 28
[3] ACS Catalysis, 2016, vol. 6, # 6, p. 3998 - 4002
[4] Journal of the Chemical Society, 1953, p. 3733,3736, 3737
[5] Journal of the American Chemical Society, 1951, vol. 73, p. 4925
[6] Recueil des Travaux Chimiques des Pays-Bas, 1952, vol. 71, p. 970,974
[7] Journal of the American Chemical Society, 1951, vol. 73, p. 107
  • 2
  • [ 93-60-7 ]
  • [ 100-55-0 ]
YieldReaction ConditionsOperation in experiment
95% With C18H28Br2N4Ru; potassium <i>tert</i>-butylate; hydrogen In 1,4-dioxane at 105℃; for 8 h; General procedure: To a mixture of catalyst (0.01 mmol), KOtBu (10 mol percent), and 1,4-dioxane (4.0 mL) in a Parr high-pressure reactor was added the ester(1.0 mmol). The dark red solution was purged with H2 and stirred under 400 psi of H2 at 105 °C for 8 h. Products isolation were performed via column chromatography using silica gel as stationary phase and n-pentane/ethylacetate or n-pentane/isopropanol mixture as eluent. The products were confirmed by 1H NMR.
60.5% With hydrogen; [2-((diphenylphospino)methyl)-2-methyl-1,3-propanediyl]bis[diphenylphosphine] In isopropyl alcohol at 150℃; for 24 h; In an autoclave, the reactant and Ru(acac)3 were dissolved with 1.4 equivalents of tris(diphenylphosphinomethyl)ethane in isopropanol, and converted at 150° C. and 150 bar of hydrogen pressure for 24 h. After the reaction, the reaction mixture was analyzed by gas chromatography.Conversion, selectivity and yield were determined by means of gas chromatography. Substrates, batch sizes and analysis are compiled in Table 1.
99 %Chromat. With [RuCl2(N-heterocyclic carbene)(bis[2-(diphenylphosphino)ethyl]amine)]; potassium <i>tert</i>-butylate; hydrogen In tetrahydrofuran at 50℃; for 6 h; Schlenk technique In a 50 mE glass Schlenk tube, 14.2 mg (0.0200 mmol/Ru) of ruthenium complex D produced in Example 5 was added, and after replacement with nitrogen gas, 0.10 mE (0.10 mmol) of 1 M K093u (solution in THF), 1.8 mE of THF, and 137 mg (1.0 mmol) of a substrate were added, and then a balloon containing hydrogen gas was attached to the Schienk tube to conduct replacement with hydrogen gas, and stirred at 50° C. for 6 hours. After cooling, the reactant was analyzed by GC, and 3-pyridinemethanol was obtained with a GC yield of 99percent.
Reference: [1] Organic Letters, 2016, vol. 18, # 15, p. 3894 - 3897
[2] Journal of Organic Chemistry, 2018, vol. 83, # 3, p. 1431 - 1440
[3] Dalton Transactions, 2012, vol. 41, # 34, p. 10136 - 10140,5
[4] Dalton Transactions, 2012, vol. 41, # 34, p. 10136 - 10140
[5] Inorganica Chimica Acta, 2019, vol. 486, p. 55 - 62
[6] Advanced Synthesis and Catalysis, 2016, vol. 358, # 5, p. 820 - 825
[7] Heterocycles, 1985, vol. 23, # 7, p. 1707 - 1722
[8] Chemical Communications, 2013, vol. 49, # 84, p. 9758 - 9760
[9] Organic Letters, 2018, vol. 20, # 9, p. 2654 - 2658
[10] Chemistry - A European Journal, 2015, vol. 21, # 7, p. 2785 - 2788
[11] Synthetic Communications, 1982, vol. 12, # 6, p. 463 - 468
[12] Bulletin of the Chemical Society of Japan, 1984, vol. 57, # 7, p. 1948 - 1953
[13] Patent: US2009/43102, 2009, A1, . Location in patent: Page/Page column 1-2
[14] Chemische Berichte, 1953, vol. 86, p. 1419,1422
[15] Patent: US2010/63294, 2010, A1, . Location in patent: Page/Page column 14
[16] European Journal of Organic Chemistry, 2013, # 11, p. 2061 - 2065
[17] Angewandte Chemie - International Edition, 2014, vol. 53, # 33, p. 8722 - 8726[18] Angew. Chem., 2014, vol. 126, # 33, p. 8867 - 8871,5
[19] Chemistry - A European Journal, 2015, vol. 21, # 30, p. 10851 - 10869
[20] Angewandte Chemie - International Edition, 2016, vol. 55, # 49, p. 15364 - 15368[21] Angew. Chem., 2016, # 128, p. 15590 - 15594,5
[22] Patent: US2017/44196, 2017, A1, . Location in patent: Paragraph 0350; 0351
[23] Chemistry - A European Journal, 2018, vol. 24, # 5, p. 1046 - 1052
[24] Organometallics, 2018, vol. 37, # 13, p. 2193 - 2201
  • 3
  • [ 500-22-1 ]
  • [ 100-55-0 ]
YieldReaction ConditionsOperation in experiment
92% With sodium dithionite; sodium hydrogencarbonate In water; isopropyl alcohol for 12 h; Inert atmosphere; Reflux General procedure: Benzaldehyde (1 g, 9.5 mmol, 1 equiv) was dissolved in 38 mL (1:1 IPA/ H2O), (0.25 M). Sodium dithionite (7.5 g, 43 mmol, 4.5 equiv) and NaHCO3 (1.6 g, 19 mmol, 2 equiv) was dissolved in water (43 mL, [1 M]) and added to the aldehyde. The mixture was refluxed for 12 hours under argon. The solution was allowed to cool to room temperature and the products were extracted using EtOAc (3 × 50 mL). This was dried using Na2SO4, filtered and dried under vacuum with a yield of 0.95 g (92percent). For entry 1.15 the compound was neutralized with 1 M HCl and extracted with EtOAc (3 × 50 mL) and washed with water (3 × 50 mL) the organic extracts were combined and dried using Na2SO4. The solvent was evaporated in vacuo and the resulting residue purified using column chromatography. Unless specified a 3:1 EtOAc/hexane eluent was used for chromatographic purification [1-3].
86.3% at 0 - 20℃; for 8 h; Example 24N-(4-(4-Fluoro-2-methoxyphenyl)pyridin-2-yl)-2-((pyridin-3-yl)methyl)-propanamideStep A: Preparation of (Pyridin-3-yl)methanol. Sodium borohydride (883 mg, 23.6 mmol) was added in three portions to a mixture of pyridine-3-carboxaldehyde (2.50 g, 23.6 mmol) in 25 ml of MeOH at 0° C. and stirred for 8 h at room temperature. The reaction mixture was quenched by ice pieces and concentrated under reduced pressure. The residue was partitioned between water and ethyl acetate. The combined ethyl acetate layer was washed with brine, dried in anhydrous sodium sulfate, filtered and concentrated in vacuo to afford 2.20 g (86.3percent) of (pyridin-3-yl)methanol as pale yellow oil.
69% With 1,1'-bis-(diphenylphosphino)ferrocene; silver(I) hexafluorophosphate; tripropylsilane; N-ethyl-N,N-diisopropylamine In water at 100℃; for 24 h; General procedure: Degassed CH2Cl2 (0.25 mL) was added to a microwave tube containing the ligand dppf (8.3 mg, 0.015 mmol) and AgPF6 (2.5 mg, 0.01 mmol) under argon. The resulting suspension was stirred at r.t., until a clear, colorless solution was obtained; then the solvent was removed under high vacuum. Benzaldehyde (1a; 20.3 μL, 0.2 mmol), tripropylsilane (2a;125 μL, 0.6 mmol), DIPEA (6.9 μL, 0.04 mmol), and H2O(0.5 mL) were subsequently added. The reaction mixture was stirred for 24 h at 100 °C, then cooled to r.t. and extracted with CH2Cl2 (3 × 10 mL). The combined organicphase was concentrated and purified by flash column chromatography on silica gel (hexane–EtOAc, 20:1) to give the desired product 3a as a colorless oil (19.5 mg, 90percent).
60% With hydrogen; triphenylphosphine; sodium hydroxide In ethanol at 50℃; for 16 h; Inert atmosphere Into a stainless steel autoclave equipped with a glass inner tube, Cu(NO3)(PPh3)2 (11.7 mg, 0.018 mmol) and triphenylphosphine (28.3 mg, 0.108 mmol) were introduced.
The inside of the autoclave was then replaced with nitrogen.
To the autoclave, an ethanolic solution of sodium hydroxide (0.03 M) (6.0 mL, 0.18 mmol) and 3-acetylpyridine (0.85 mL, 9 mmol) were added, and stirring was performed at a hydrogen pressure of 5 MPa at 50° C. for 16 hours.
The hydrogen was released with great care, and the conversion was analyzed by GC (88percent).
The contents were concentrated, and then purified by silica gel chromatography.
Thus, 588 mg of the 3-pyridylcarbinol was obtained (yield: 60percent).
60% With hydrogen; sodium hydroxide In ethanol at 50℃; for 16 h; Autoclave (Example 7)
Hydrogenation Reaction of 3-Acetylpyridine
Into a stainless steel autoclave equipped with a glass inner tube, Cu(NO3)(PPh3)2 (11.7 mg, 0.018 mmol) and triphenylphosphine (28.3 mg, 0.108 mmol) were introduced.
The inside of the autoclave was then replaced with nitrogen.
To the autoclave, an ethanolic solution of sodium hydroxide (0.03 M) (6.0mL, 0.18 mmol) and 3-acetylpyridine (0.85 mL, 9 mmol) were added, and stirring was performed at a hydrogen pressure of 5 MPa at 50°C for 16 hours.
The hydrogen was released with great care, and the conversion was analyzed by GC (88percent).
The contents were concentrated, and then purified by silica gel chromatography.
Thus, 588 mg of the 3-pyridylcarbinol was obtained (yield: 60percent).
57 %Chromat. With formaldehyd; tricarbonyl(η4-1,3-bis(trimethylsilyl)-4,5,6,7-tetrahydro-2H-inden-2-one)iron; water; sodium carbonate In dimethyl sulfoxide at 120℃; for 24 h; Inert atmosphere; Sealed tube General procedure: Knölker iron complex 2a (3 mol percent,12.6 mg), paraformaldehyde (300 mg, 10 mmol), and Na2CO3 (106 mg, 1 mmol,1.0 equiv) and a stirring bar were charged in a pressure tube and flushed withargon. DMSO (1.0 mL), degassed water (1.0 mL), and benzaldehyde (1 mmol)were added under an argon atmosphere to the pressure tube with a syringe.The pressure tube was placed in oil and heated at 120 C for 24 h, then cooledto room temperature. The reaction mixture was neutralized with HCl (1M) andstirred for 30 min. After extraction with EtOAc for 3 times, the combinedorganic layers were dried over MgSO4. The crude product was purified bycolumn chromatography (Heptane/EtOAc: 70:30). The reaction was cooled toroom temperature and hexadecane (100 lL) was added as a GC internalstandard.

Reference: [1] Journal of the American Chemical Society, 2017, vol. 139, # 47, p. 17089 - 17097
[2] Journal of Organic Chemistry, 2004, vol. 69, # 5, p. 1492 - 1496
[3] Organic Letters, 2007, vol. 9, # 15, p. 2791 - 2793
[4] European Journal of Organic Chemistry, 2011, # 27, p. 5291 - 5296
[5] Tetrahedron Letters, 1982, vol. 23, # 49, p. 5141 - 5144
[6] Organic Letters, 2007, vol. 9, # 26, p. 5429 - 5432
[7] Beilstein Journal of Organic Chemistry, 2018, vol. 14, p. 1529 - 1536
[8] Journal of Organic Chemistry, 2012, vol. 77, # 10, p. 4821 - 4825
[9] Patent: US2011/224225, 2011, A1, . Location in patent: Page/Page column 27
[10] Synlett, 2009, # 8, p. 1295 - 1298
[11] Angewandte Chemie - International Edition, 2013, vol. 52, # 45, p. 11871 - 11874[12] Angew. Chem., 2013, vol. 125, # 45, p. 12087 - 12090
[13] Bulletin of the Chemical Society of Japan, 1995, vol. 68, # 9, p. 2595 - 2602
[14] Journal of Organic Chemistry, 2018, vol. 83, # 3, p. 1204 - 1215
[15] Journal of Organic Chemistry, 2017, vol. 82, # 11, p. 5932 - 5939
[16] Tetrahedron Letters, 1986, vol. 27, # 21, p. 2357 - 2360
[17] Synlett, 2013, vol. 24, # 16, p. 2049 - 2056
[18] Patent: US2011/201820, 2011, A1, . Location in patent: Page/Page column 8
[19] Patent: EP2374783, 2011, A1, . Location in patent: Page/Page column 12
[20] Journal of Chemical Research - Part S, 2003, # 6, p. 368 - 369
[21] J. Appl. Chem. USSR (Engl. Transl.), 1980, vol. 53, # 4, p. 724 - 727[22] Zhurnal Prikladnoi Khimii (Sankt-Peterburg, Russian Federation), 1980, vol. 53, # 4, p. 917 - 920
[23] Pharmaceutical Chemistry Journal, 1984, vol. 18, # 3, p. 175 - 179[24] Khimiko-Farmatsevticheskii Zhurnal, 1984, vol. 18, # 3, p. 312 - 316
[25] Journal of the American Chemical Society, 2017, vol. 139, # 13, p. 4901 - 4915
[26] Organometallics, 2017, vol. 36, # 18, p. 3477 - 3483
[27] Synthetic Communications, 1992, vol. 22, # 4, p. 497 - 507
[28] Journal of Heterocyclic Chemistry, 1994, vol. 31, # 6, p. 1599 - 1603
[29] Bulletin of the Chemical Society of Japan, 1997, vol. 70, # 1, p. 156 - 168
[30] Chemistry - A European Journal, 2005, vol. 11, # 22, p. 6574 - 6582
[31] European Journal of Organic Chemistry, 2002, # 19, p. 3326 - 3335
[32] Chemistry - A European Journal, 2008, vol. 14, # 36, p. 11480 - 11487
[33] Archives of Biochemistry and Biophysics, 2010, vol. 503, # 2, p. 230 - 237
[34] Bioscience, Biotechnology and Biochemistry, 2011, vol. 75, # 6, p. 1055 - 1060
[35] Chemical Communications, 2012, vol. 48, # 38, p. 4567 - 4569
[36] ACS Medicinal Chemistry Letters, 2013, vol. 4, # 2, p. 278 - 282
[37] Biochemical Pharmacology, 2013, vol. 86, # 9, p. 1366 - 1375
[38] Chemistry - A European Journal, 2014, vol. 20, # 3, p. 839 - 845
[39] Journal of the American Chemical Society, 2014, vol. 136, # 6, p. 2520 - 2528
[40] Chemistry - A European Journal, 2014, vol. 20, # 46, p. 14959 - 14964
[41] Organic and Biomolecular Chemistry, 2015, vol. 13, # 6, p. 1768 - 1777
[42] Tetrahedron Letters, 2015, vol. 56, # 9, p. 1118 - 1121
[43] Journal of Biochemistry, 2015, vol. 158, # 5, p. 425 - 434
[44] Journal of the American Chemical Society, 2016, vol. 138, # 32, p. 10356 - 10364
[45] ChemSusChem, 2014, vol. 7, # 9, p. 2684 - 2689
[46] Bioorganic and Medicinal Chemistry, 2018, vol. 26, # 12, p. 3145 - 3157
[47] Advanced Synthesis and Catalysis, 2018, vol. 360, # 11, p. 2131 - 2137
[48] Catalysis Communications, 2018, vol. 115, p. 1 - 5
[49] Chemistry - A European Journal, 2018, vol. 24, # 49, p. 12920 - 12931
  • 4
  • [ 23100-12-1 ]
  • [ 100-55-0 ]
YieldReaction ConditionsOperation in experiment
88% With hydrogen; sodium acetate; palladium dichloride In methanol at 35℃; for 2 h; Typical procedures: 6-bromonicotinaldehyde (930 mg, 5.0 mmol), NaOAc (820 mg, 10.0 mmol), MeOH (30 mL), and PdCl2 (45 mg) were mixed in a glass bottle capped with a balloon filled with hydrogen. After stirred at 35 °C for 4 h, the mixture was filtered and washed with MeOH. The solvent was removed and the residue was dissolved in water, neutralized with solid NaHCO3, and extracted with ethyl acetate. The organic phase was dried over anhyd Na2SO4, and then filtered. The solvent was removed and the residue was subjected to chromatography to yield pyridin-3-ylmethanol (428 mg, 78percent).
Reference: [1] Tetrahedron Letters, 2012, vol. 53, # 29, p. 3798 - 3801
  • 5
  • [ 113118-81-3 ]
  • [ 100-55-0 ]
YieldReaction ConditionsOperation in experiment
83% With hydrogen; sodium acetate; palladium dichloride In methanol at 35℃; for 2 h; Typical procedures: 6-bromonicotinaldehyde (930 mg, 5.0 mmol), NaOAc (820 mg, 10.0 mmol), MeOH (30 mL), and PdCl2 (45 mg) were mixed in a glass bottle capped with a balloon filled with hydrogen. After stirred at 35 °C for 4 h, the mixture was filtered and washed with MeOH. The solvent was removed and the residue was dissolved in water, neutralized with solid NaHCO3, and extracted with ethyl acetate. The organic phase was dried over anhyd Na2SO4, and then filtered. The solvent was removed and the residue was subjected to chromatography to yield pyridin-3-ylmethanol (428 mg, 78percent).
Reference: [1] Tetrahedron Letters, 2012, vol. 53, # 29, p. 3798 - 3801
  • 6
  • [ 149806-06-4 ]
  • [ 100-55-0 ]
YieldReaction ConditionsOperation in experiment
78% With hydrogen; sodium acetate; palladium dichloride In methanol at 35℃; for 4 h; Typical procedures: 6-bromonicotinaldehyde (930 mg, 5.0 mmol), NaOAc (820 mg, 10.0 mmol), MeOH (30 mL), and PdCl2 (45 mg) were mixed in a glass bottle capped with a balloon filled with hydrogen. After stirred at 35 °C for 4 h, the mixture was filtered and washed with MeOH. The solvent was removed and the residue was dissolved in water, neutralized with solid NaHCO3, and extracted with ethyl acetate. The organic phase was dried over anhyd Na2SO4, and then filtered. The solvent was removed and the residue was subjected to chromatography to yield pyridin-3-ylmethanol (428 mg, 78percent).
Reference: [1] Tetrahedron Letters, 2012, vol. 53, # 29, p. 3798 - 3801
  • 7
  • [ 6968-72-5 ]
  • [ 100-55-0 ]
Reference: [1] Organic Letters, 2000, vol. 2, # 22, p. 3525 - 3526
[2] Synthesis, 1987, # 1, p. 55 - 56
[3] Synthetic Communications, 2004, vol. 34, # 17, p. 3197 - 3202
  • 8
  • [ 500-22-1 ]
  • [ 100-55-0 ]
  • [ 59-67-6 ]
Reference: [1] Organic Letters, 2005, vol. 7, # 26, p. 5893 - 5895
[2] Tetrahedron Letters, 2006, vol. 47, # 32, p. 5771 - 5774
[3] Tetrahedron Letters, 1998, vol. 39, # 46, p. 8437 - 8440
[4] Tetrahedron Letters, 1999, vol. 40, # 6, p. 1179 - 1180
  • 9
  • [ 626-60-8 ]
  • [ 859842-81-2 ]
  • [ 100-55-0 ]
YieldReaction ConditionsOperation in experiment
12% With potassium phosphate; water In 1,4-dioxane at 20 - 100℃; for 14.25 h; To a solution of 3-chloropyridine (50 mg, 0.44 mmol) in 1,4-dioxane (2 ml) and water (200 μl) were added sodium [(2,2-dimethyl)propionyloxy]methyl trifluoroborate (181 mg, 0.88 mmol), palladium (II) acetate (9.9 mg, 0.044 mmol), 2-dicyclohexylphosphino-2',6'-dimethoxybiphenyl (36mg, 0.088 mmol), and potassium phosphate (405 mg, 1.76 mmol) at room temperature. The reaction mixture was stirred at 100°C for 14 hours and 15 minutes under the nitrogen atmosphere. After the reaction mixture was cooled at room temperature, water was added to the mixture, followed by filtration using Celite. After the filtrate was extracted with ethyl acetate, the organic layer was washed with an aqueous saturated sodium chloride solution, dried over anhydrous magnesium sulfate, and filtered. The solvent was distilled off from the filtrate under reduced pressure. The residue was purified by NH silica gel column chromatography (heptane: ethyl acetate=2:1) to obtain the title compound (10 mg, 0.052 mmol, 12percent). 1H-NMR Spectrum (CDCl3) δ (ppm): 1.23(9H, s), 5.13(2H, s), 7.28-7.32(1H, m), 7.66-7.69(1H, m), 8.58(1H, dd, J=1.7 Hz, 4.8 Hz), 8.62(1H, d, J=1.7 Hz); As a side product of (Example B-13), the title compound (6.8 mg, 0.062 mmol, 14percent) was obtained. 1H-NMR Spectrum (CDCl3) δ (ppm) : 4. 73 (2H, s), 7.28-7.31(1H, m), 7.72-7.75(1H, m), 8.49(1H, dd, J=1.7 Hz, 4. 9 Hz), 8.55(1H, d, J=2.0 Hz).
Reference: [1] Patent: EP1867650, 2007, A1, . Location in patent: Page/Page column 40
  • 10
  • [ 98-92-0 ]
  • [ 100-55-0 ]
Reference: [1] ACS Catalysis, 2016, vol. 6, # 1, p. 47 - 54
[2] Chemical Science, 2017, vol. 8, # 5, p. 3576 - 3585
  • 11
  • [ 1752-96-1 ]
  • [ 100-55-0 ]
  • [ 62-53-3 ]
Reference: [1] ACS Catalysis, 2016, vol. 6, # 1, p. 47 - 54
[2] Chemical Science, 2017, vol. 8, # 5, p. 3576 - 3585
  • 12
  • [ 114-33-0 ]
  • [ 100-55-0 ]
Reference: [1] ACS Catalysis, 2016, vol. 6, # 1, p. 47 - 54
  • 13
  • [ 3731-52-0 ]
  • [ 100-55-0 ]
Reference: [1] Synlett, 2012, vol. 23, # 5, p. 706 - 710
[2] DRP/DRBP Org.Chem.,
  • 14
  • [ 3731-52-0 ]
  • [ 69966-55-8 ]
  • [ 100-55-0 ]
YieldReaction ConditionsOperation in experiment
81% With bromine; isopentyl nitrite In dichloromethane [0091] The following example explores reaction conditions and product distributions.[0092] Under reaction conditions identical to those used with Br2 in Example 1, the reaction using I2 failed to produce the corresponding iodide. Using 10 equivalents of ICl however, as the source of halogen, results in the production of the chloride in good yield without the iodide. In a further experiment, using 10 equivalents of IBr, gives again the bromide with no detectable iodide. As per the postulated mechanism, the highly reactive nitryl, and nitrosyl halides are the active intermediates, neither of which is known for the iodide.[0093] Conversion of 3-aminomethylpyridine to the bromide in THF was similar to that in CH2Cl2, but decreased in the series CH2Cl2 > CHCl3 > CCl4 with a <n="23"/>corresponding increase in the formation of the hydroxy product. Reaction in DMF or DMSO led to remaining AMP under the conditions used and also favored the formation of the 3-(hydroxymethyl)pyridine over the bromide.[0094] Using isoamyl nitrite as the nitroso source, and Br2, in water, resulted in an incomplete conversion with very little of the hydroxy product. In all solvents tried using NaNC>2 and acetic acid as the nitroso source, an increased amount of 3- (hydroxymethyl)pyridine was observed with incomplete reaction of the AMP. This latter observation is most likely due to the use of NaNO2 for the nitroso source, as in CH2CI2, increasing additions of acetic acid up to 5 equivalents with respect to AMP, favored the formation of the bromide over the hydroxy product. The stronger acid TFA, however, favored the formation of the hydroxy product over the bromide. Addition of an organic base, TEA, piperidine or DBU was well tolerated up to about 1 equivalent with respect to AMP, with higher amounts leading to less conversion of the amine to the bromide. Reaction to the bromide was concentration dependent with best conversions occurring at > 0.1 molar in AMP (variable only explored in GH2CI2).[0095] Conversion of 3-aminomethyl pyridine to 3-(bromomethyl)pyridine using 1.1 equiv. of nitrosyl agent and 5 equiv. of bromine was conducted in the solvent shown in Table 1 below. A sample of the reaction mixture was taken and quenched by high dilution into acetontrile for ESMS analysis of the products. 3- (hydroxymethyl)pyridine and 3-(isoamyloxymethyl)pyridine side products were also observed in addition to starting material and 3-(bromomethyl)pyridine product.Table 1. Product distributions based on variation in reaction conditions. <n="24"/>[0096] The gas evolved by the reaction was identified by collecting the gas in an IR cell and examining by IR Spectroscopy. The IR spectrum in Figure 2 shows the N2O gas produced by the reaction.[0097] Conversion of 3-aminomethylpyridine (AMP) and 2-(2- aminoethyl)pyrdine (AEP) is efficient, as demonstrated by the HPLC traces before and after reaction as shown in Figure 3.[0098] Titration of the addition of both isoamyl nitrite and Bτ2 to AMP shows that the amine disappears from the reaction mixture on addition of one equivalent each of isoamyl nitrite and Br2. As shown in Figure 7, the formation of the product 3- (bromomethyl)pyridine is out of phase with the disappearance of reacting amine, and in some trials was almost completely absent until greater than one equivalent of isoamyl nitrite and Br2 was added. Best yields were obtained by a total addition of about five equivalents OfBr2.in this example
79% With bromine; isopentyl nitrite In tetrahydrofuran [0091] The following example explores reaction conditions and product distributions.[0092] Under reaction conditions identical to those used with Br2 in Example 1, the reaction using I2 failed to produce the corresponding iodide. Using 10 equivalents of ICl however, as the source of halogen, results in the production of the chloride in good yield without the iodide. In a further experiment, using 10 equivalents of IBr, gives again the bromide with no detectable iodide. As per the postulated mechanism, the highly reactive nitryl, and nitrosyl halides are the active intermediates, neither of which is known for the iodide.[0093] Conversion of 3-aminomethylpyridine to the bromide in THF was similar to that in CH2Cl2, but decreased in the series CH2Cl2 > CHCl3 > CCl4 with a <n="23"/>corresponding increase in the formation of the hydroxy product. Reaction in DMF or DMSO led to remaining AMP under the conditions used and also favored the formation of the 3-(hydroxymethyl)pyridine over the bromide.[0094] Using isoamyl nitrite as the nitroso source, and Br2, in water, resulted in an incomplete conversion with very little of the hydroxy product. In all solvents tried using NaNC>2 and acetic acid as the nitroso source, an increased amount of 3- (hydroxymethyl)pyridine was observed with incomplete reaction of the AMP. This latter observation is most likely due to the use of NaNO2 for the nitroso source, as in CH2CI2, increasing additions of acetic acid up to 5 equivalents with respect to AMP, favored the formation of the bromide over the hydroxy product. The stronger acid TFA, however, favored the formation of the hydroxy product over the bromide. Addition of an organic base, TEA, piperidine or DBU was well tolerated up to about 1 equivalent with respect to AMP, with higher amounts leading to less conversion of the amine to the bromide. Reaction to the bromide was concentration dependent with best conversions occurring at > 0.1 molar in AMP (variable only explored in GH2CI2).[0095] Conversion of 3-aminomethyl pyridine to 3-(bromomethyl)pyridine using 1.1 equiv. of nitrosyl agent and 5 equiv. of bromine was conducted in the solvent shown in Table 1 below. A sample of the reaction mixture was taken and quenched by high dilution into acetontrile for ESMS analysis of the products. 3- (hydroxymethyl)pyridine and 3-(isoamyloxymethyl)pyridine side products were also observed in addition to starting material and 3-(bromomethyl)pyridine product.Table 1. Product distributions based on variation in reaction conditions. <n="24"/>[0096] The gas evolved by the reaction was identified by collecting the gas in an IR cell and examining by IR Spectroscopy. The IR spectrum in Figure 2 shows the N2O gas produced by the reaction.[0097] Conversion of 3-aminomethylpyridine (AMP) and 2-(2- aminoethyl)pyrdine (AEP) is efficient, as demonstrated by the HPLC traces before and after reaction as shown in Figure 3.[0098] Titration of the addition of both isoamyl nitrite and Bτ2 to AMP shows that the amine disappears from the reaction mixture on addition of one equivalent each of isoamyl nitrite and Br2. As shown in Figure 7, the formation of the product 3- (bromomethyl)pyridine is out of phase with the disappearance of reacting amine, and in some trials was almost completely absent until greater than one equivalent of isoamyl nitrite and Br2 was added. Best yields were obtained by a total addition of about five equivalents OfBr2.in this example
67% With bromine; isopentyl nitrite In chloroform [0091] The following example explores reaction conditions and product distributions.[0092] Under reaction conditions identical to those used with Br2 in Example 1, the reaction using I2 failed to produce the corresponding iodide. Using 10 equivalents of ICl however, as the source of halogen, results in the production of the chloride in good yield without the iodide. In a further experiment, using 10 equivalents of IBr, gives again the bromide with no detectable iodide. As per the postulated mechanism, the highly reactive nitryl, and nitrosyl halides are the active intermediates, neither of which is known for the iodide.[0093] Conversion of 3-aminomethylpyridine to the bromide in THF was similar to that in CH2Cl2, but decreased in the series CH2Cl2 > CHCl3 > CCl4 with a <n="23"/>corresponding increase in the formation of the hydroxy product. Reaction in DMF or DMSO led to remaining AMP under the conditions used and also favored the formation of the 3-(hydroxymethyl)pyridine over the bromide.[0094] Using isoamyl nitrite as the nitroso source, and Br2, in water, resulted in an incomplete conversion with very little of the hydroxy product. In all solvents tried using NaNC>2 and acetic acid as the nitroso source, an increased amount of 3- (hydroxymethyl)pyridine was observed with incomplete reaction of the AMP. This latter observation is most likely due to the use of NaNO2 for the nitroso source, as in CH2CI2, increasing additions of acetic acid up to 5 equivalents with respect to AMP, favored the formation of the bromide over the hydroxy product. The stronger acid TFA, however, favored the formation of the hydroxy product over the bromide. Addition of an organic base, TEA, piperidine or DBU was well tolerated up to about 1 equivalent with respect to AMP, with higher amounts leading to less conversion of the amine to the bromide. Reaction to the bromide was concentration dependent with best conversions occurring at > 0.1 molar in AMP (variable only explored in GH2CI2).[0095] Conversion of 3-aminomethyl pyridine to 3-(bromomethyl)pyridine using 1.1 equiv. of nitrosyl agent and 5 equiv. of bromine was conducted in the solvent shown in Table 1 below. A sample of the reaction mixture was taken and quenched by high dilution into acetontrile for ESMS analysis of the products. 3- (hydroxymethyl)pyridine and 3-(isoamyloxymethyl)pyridine side products were also observed in addition to starting material and 3-(bromomethyl)pyridine product.Table 1. Product distributions based on variation in reaction conditions. <n="24"/>[0096] The gas evolved by the reaction was identified by collecting the gas in an IR cell and examining by IR Spectroscopy. The IR spectrum in Figure 2 shows the N2O gas produced by the reaction.[0097] Conversion of 3-aminomethylpyridine (AMP) and 2-(2- aminoethyl)pyrdine (AEP) is efficient, as demonstrated by the HPLC traces before and after reaction as shown in Figure 3.[0098] Titration of the addition of both isoamyl nitrite and Bτ2 to AMP shows that the amine disappears from the reaction mixture on addition of one equivalent each of isoamyl nitrite and Br2. As shown in Figure 7, the formation of the product 3- (bromomethyl)pyridine is out of phase with the disappearance of reacting amine, and in some trials was almost completely absent until greater than one equivalent of isoamyl nitrite and Br2 was added. Best yields were obtained by a total addition of about five equivalents OfBr2.in this example
61% With bromine; isopentyl nitrite In water [0091] The following example explores reaction conditions and product distributions.[0092] Under reaction conditions identical to those used with Br2 in Example 1, the reaction using I2 failed to produce the corresponding iodide. Using 10 equivalents of ICl however, as the source of halogen, results in the production of the chloride in good yield without the iodide. In a further experiment, using 10 equivalents of IBr, gives again the bromide with no detectable iodide. As per the postulated mechanism, the highly reactive nitryl, and nitrosyl halides are the active intermediates, neither of which is known for the iodide.[0093] Conversion of 3-aminomethylpyridine to the bromide in THF was similar to that in CH2Cl2, but decreased in the series CH2Cl2 > CHCl3 > CCl4 with a <n="23"/>corresponding increase in the formation of the hydroxy product. Reaction in DMF or DMSO led to remaining AMP under the conditions used and also favored the formation of the 3-(hydroxymethyl)pyridine over the bromide.[0094] Using isoamyl nitrite as the nitroso source, and Br2, in water, resulted in an incomplete conversion with very little of the hydroxy product. In all solvents tried using NaNC>2 and acetic acid as the nitroso source, an increased amount of 3- (hydroxymethyl)pyridine was observed with incomplete reaction of the AMP. This latter observation is most likely due to the use of NaNO2 for the nitroso source, as in CH2CI2, increasing additions of acetic acid up to 5 equivalents with respect to AMP, favored the formation of the bromide over the hydroxy product. The stronger acid TFA, however, favored the formation of the hydroxy product over the bromide. Addition of an organic base, TEA, piperidine or DBU was well tolerated up to about 1 equivalent with respect to AMP, with higher amounts leading to less conversion of the amine to the bromide. Reaction to the bromide was concentration dependent with best conversions occurring at > 0.1 molar in AMP (variable only explored in GH2CI2).[0095] Conversion of 3-aminomethyl pyridine to 3-(bromomethyl)pyridine using 1.1 equiv. of nitrosyl agent and 5 equiv. of bromine was conducted in the solvent shown in Table 1 below. A sample of the reaction mixture was taken and quenched by high dilution into acetontrile for ESMS analysis of the products. 3- (hydroxymethyl)pyridine and 3-(isoamyloxymethyl)pyridine side products were also observed in addition to starting material and 3-(bromomethyl)pyridine product.Table 1. Product distributions based on variation in reaction conditions. <n="24"/>[0096] The gas evolved by the reaction was identified by collecting the gas in an IR cell and examining by IR Spectroscopy. The IR spectrum in Figure 2 shows the N2O gas produced by the reaction.[0097] Conversion of 3-aminomethylpyridine (AMP) and 2-(2- aminoethyl)pyrdine (AEP) is efficient, as demonstrated by the HPLC traces before and after reaction as shown in Figure 3.[0098] Titration of the addition of both isoamyl nitrite and Bτ2 to AMP shows that the amine disappears from the reaction mixture on addition of one equivalent each of isoamyl nitrite and Br2. As shown in Figure 7, the formation of the product 3- (bromomethyl)pyridine is out of phase with the disappearance of reacting amine, and in some trials was almost completely absent until greater than one equivalent of isoamyl nitrite and Br2 was added. Best yields were obtained by a total addition of about five equivalents OfBr2.in this example
57% With bromine; isopentyl nitrite In dimethyl sulfoxide [0091] The following example explores reaction conditions and product distributions.[0092] Under reaction conditions identical to those used with Br2 in Example 1, the reaction using I2 failed to produce the corresponding iodide. Using 10 equivalents of ICl however, as the source of halogen, results in the production of the chloride in good yield without the iodide. In a further experiment, using 10 equivalents of IBr, gives again the bromide with no detectable iodide. As per the postulated mechanism, the highly reactive nitryl, and nitrosyl halides are the active intermediates, neither of which is known for the iodide.[0093] Conversion of 3-aminomethylpyridine to the bromide in THF was similar to that in CH2Cl2, but decreased in the series CH2Cl2 > CHCl3 > CCl4 with a <n="23"/>corresponding increase in the formation of the hydroxy product. Reaction in DMF or DMSO led to remaining AMP under the conditions used and also favored the formation of the 3-(hydroxymethyl)pyridine over the bromide.[0094] Using isoamyl nitrite as the nitroso source, and Br2, in water, resulted in an incomplete conversion with very little of the hydroxy product. In all solvents tried using NaNC>2 and acetic acid as the nitroso source, an increased amount of 3- (hydroxymethyl)pyridine was observed with incomplete reaction of the AMP. This latter observation is most likely due to the use of NaNO2 for the nitroso source, as in CH2CI2, increasing additions of acetic acid up to 5 equivalents with respect to AMP, favored the formation of the bromide over the hydroxy product. The stronger acid TFA, however, favored the formation of the hydroxy product over the bromide. Addition of an organic base, TEA, piperidine or DBU was well tolerated up to about 1 equivalent with respect to AMP, with higher amounts leading to less conversion of the amine to the bromide. Reaction to the bromide was concentration dependent with best conversions occurring at > 0.1 molar in AMP (variable only explored in GH2CI2).[0095] Conversion of 3-aminomethyl pyridine to 3-(bromomethyl)pyridine using 1.1 equiv. of nitrosyl agent and 5 equiv. of bromine was conducted in the solvent shown in Table 1 below. A sample of the reaction mixture was taken and quenched by high dilution into acetontrile for ESMS analysis of the products. 3- (hydroxymethyl)pyridine and 3-(isoamyloxymethyl)pyridine side products were also observed in addition to starting material and 3-(bromomethyl)pyridine product.Table 1. Product distributions based on variation in reaction conditions. <n="24"/>[0096] The gas evolved by the reaction was identified by collecting the gas in an IR cell and examining by IR Spectroscopy. The IR spectrum in Figure 2 shows the N2O gas produced by the reaction.[0097] Conversion of 3-aminomethylpyridine (AMP) and 2-(2- aminoethyl)pyrdine (AEP) is efficient, as demonstrated by the HPLC traces before and after reaction as shown in Figure 3.[0098] Titration of the addition of both isoamyl nitrite and Bτ2 to AMP shows that the amine disappears from the reaction mixture on addition of one equivalent each of isoamyl nitrite and Br2. As shown in Figure 7, the formation of the product 3- (bromomethyl)pyridine is out of phase with the disappearance of reacting amine, and in some trials was almost completely absent until greater than one equivalent of isoamyl nitrite and Br2 was added. Best yields were obtained by a total addition of about five equivalents OfBr2.in this example
54% With bromine; isopentyl nitrite In tetrachloromethane [0091] The following example explores reaction conditions and product distributions.[0092] Under reaction conditions identical to those used with Br2 in Example 1, the reaction using I2 failed to produce the corresponding iodide. Using 10 equivalents of ICl however, as the source of halogen, results in the production of the chloride in good yield without the iodide. In a further experiment, using 10 equivalents of IBr, gives again the bromide with no detectable iodide. As per the postulated mechanism, the highly reactive nitryl, and nitrosyl halides are the active intermediates, neither of which is known for the iodide.[0093] Conversion of 3-aminomethylpyridine to the bromide in THF was similar to that in CH2Cl2, but decreased in the series CH2Cl2 > CHCl3 > CCl4 with a <n="23"/>corresponding increase in the formation of the hydroxy product. Reaction in DMF or DMSO led to remaining AMP under the conditions used and also favored the formation of the 3-(hydroxymethyl)pyridine over the bromide.[0094] Using isoamyl nitrite as the nitroso source, and Br2, in water, resulted in an incomplete conversion with very little of the hydroxy product. In all solvents tried using NaNC>2 and acetic acid as the nitroso source, an increased amount of 3- (hydroxymethyl)pyridine was observed with incomplete reaction of the AMP. This latter observation is most likely due to the use of NaNO2 for the nitroso source, as in CH2CI2, increasing additions of acetic acid up to 5 equivalents with respect to AMP, favored the formation of the bromide over the hydroxy product. The stronger acid TFA, however, favored the formation of the hydroxy product over the bromide. Addition of an organic base, TEA, piperidine or DBU was well tolerated up to about 1 equivalent with respect to AMP, with higher amounts leading to less conversion of the amine to the bromide. Reaction to the bromide was concentration dependent with best conversions occurring at > 0.1 molar in AMP (variable only explored in GH2CI2).[0095] Conversion of 3-aminomethyl pyridine to 3-(bromomethyl)pyridine using 1.1 equiv. of nitrosyl agent and 5 equiv. of bromine was conducted in the solvent shown in Table 1 below. A sample of the reaction mixture was taken and quenched by high dilution into acetontrile for ESMS analysis of the products. 3- (hydroxymethyl)pyridine and 3-(isoamyloxymethyl)pyridine side products were also observed in addition to starting material and 3-(bromomethyl)pyridine product.Table 1. Product distributions based on variation in reaction conditions. <n="24"/>[0096] The gas evolved by the reaction was identified by collecting the gas in an IR cell and examining by IR Spectroscopy. The IR spectrum in Figure 2 shows the N2O gas produced by the reaction.[0097] Conversion of 3-aminomethylpyridine (AMP) and 2-(2- aminoethyl)pyrdine (AEP) is efficient, as demonstrated by the HPLC traces before and after reaction as shown in Figure 3.[0098] Titration of the addition of both isoamyl nitrite and Bτ2 to AMP shows that the amine disappears from the reaction mixture on addition of one equivalent each of isoamyl nitrite and Br2. As shown in Figure 7, the formation of the product 3- (bromomethyl)pyridine is out of phase with the disappearance of reacting amine, and in some trials was almost completely absent until greater than one equivalent of isoamyl nitrite and Br2 was added. Best yields were obtained by a total addition of about five equivalents OfBr2.in this example
47% With bromine; acetic acid; sodium nitrite In dimethyl sulfoxide [0091] The following example explores reaction conditions and product distributions.[0092] Under reaction conditions identical to those used with Br2 in Example 1, the reaction using I2 failed to produce the corresponding iodide. Using 10 equivalents of ICl however, as the source of halogen, results in the production of the chloride in good yield without the iodide. In a further experiment, using 10 equivalents of IBr, gives again the bromide with no detectable iodide. As per the postulated mechanism, the highly reactive nitryl, and nitrosyl halides are the active intermediates, neither of which is known for the iodide.[0093] Conversion of 3-aminomethylpyridine to the bromide in THF was similar to that in CH2Cl2, but decreased in the series CH2Cl2 > CHCl3 > CCl4 with a <n="23"/>corresponding increase in the formation of the hydroxy product. Reaction in DMF or DMSO led to remaining AMP under the conditions used and also favored the formation of the 3-(hydroxymethyl)pyridine over the bromide.[0094] Using isoamyl nitrite as the nitroso source, and Br2, in water, resulted in an incomplete conversion with very little of the hydroxy product. In all solvents tried using NaNC>2 and acetic acid as the nitroso source, an increased amount of 3- (hydroxymethyl)pyridine was observed with incomplete reaction of the AMP. This latter observation is most likely due to the use of NaNO2 for the nitroso source, as in CH2CI2, increasing additions of acetic acid up to 5 equivalents with respect to AMP, favored the formation of the bromide over the hydroxy product. The stronger acid TFA, however, favored the formation of the hydroxy product over the bromide. Addition of an organic base, TEA, piperidine or DBU was well tolerated up to about 1 equivalent with respect to AMP, with higher amounts leading to less conversion of the amine to the bromide. Reaction to the bromide was concentration dependent with best conversions occurring at > 0.1 molar in AMP (variable only explored in GH2CI2).[0095] Conversion of 3-aminomethyl pyridine to 3-(bromomethyl)pyridine using 1.1 equiv. of nitrosyl agent and 5 equiv. of bromine was conducted in the solvent shown in Table 1 below. A sample of the reaction mixture was taken and quenched by high dilution into acetontrile for ESMS analysis of the products. 3- (hydroxymethyl)pyridine and 3-(isoamyloxymethyl)pyridine side products were also observed in addition to starting material and 3-(bromomethyl)pyridine product.Table 1. Product distributions based on variation in reaction conditions. <n="24"/>[0096] The gas evolved by the reaction was identified by collecting the gas in an IR cell and examining by IR Spectroscopy. The IR spectrum in Figure 2 shows the N2O gas produced by the reaction.[0097] Conversion of 3-aminomethylpyridine (AMP) and 2-(2- aminoethyl)pyrdine (AEP) is efficient, as demonstrated by the HPLC traces before and after reaction as shown in Figure 3.[0098] Titration of the addition of both isoamyl nitrite and Bτ2 to AMP shows that the amine disappears from the reaction mixture on addition of one equivalent each of isoamyl nitrite and Br2. As shown in Figure 7, the formation of the product 3- (bromomethyl)pyridine is out of phase with the disappearance of reacting amine, and in some trials was almost completely absent until greater than one equivalent of isoamyl nitrite and Br2 was added. Best yields were obtained by a total addition of about five equivalents OfBr2.in this example
22% With bromine; isopentyl nitrite In N,N-dimethyl-formamide [0091] The following example explores reaction conditions and product distributions.[0092] Under reaction conditions identical to those used with Br2 in Example 1, the reaction using I2 failed to produce the corresponding iodide. Using 10 equivalents of ICl however, as the source of halogen, results in the production of the chloride in good yield without the iodide. In a further experiment, using 10 equivalents of IBr, gives again the bromide with no detectable iodide. As per the postulated mechanism, the highly reactive nitryl, and nitrosyl halides are the active intermediates, neither of which is known for the iodide.[0093] Conversion of 3-aminomethylpyridine to the bromide in THF was similar to that in CH2Cl2, but decreased in the series CH2Cl2 > CHCl3 > CCl4 with a <n="23"/>corresponding increase in the formation of the hydroxy product. Reaction in DMF or DMSO led to remaining AMP under the conditions used and also favored the formation of the 3-(hydroxymethyl)pyridine over the bromide.[0094] Using isoamyl nitrite as the nitroso source, and Br2, in water, resulted in an incomplete conversion with very little of the hydroxy product. In all solvents tried using NaNC>2 and acetic acid as the nitroso source, an increased amount of 3- (hydroxymethyl)pyridine was observed with incomplete reaction of the AMP. This latter observation is most likely due to the use of NaNO2 for the nitroso source, as in CH2CI2, increasing additions of acetic acid up to 5 equivalents with respect to AMP, favored the formation of the bromide over the hydroxy product. The stronger acid TFA, however, favored the formation of the hydroxy product over the bromide. Addition of an organic base, TEA, piperidine or DBU was well tolerated up to about 1 equivalent with respect to AMP, with higher amounts leading to less conversion of the amine to the bromide. Reaction to the bromide was concentration dependent with best conversions occurring at > 0.1 molar in AMP (variable only explored in GH2CI2).[0095] Conversion of 3-aminomethyl pyridine to 3-(bromomethyl)pyridine using 1.1 equiv. of nitrosyl agent and 5 equiv. of bromine was conducted in the solvent shown in Table 1 below. A sample of the reaction mixture was taken and quenched by high dilution into acetontrile for ESMS analysis of the products. 3- (hydroxymethyl)pyridine and 3-(isoamyloxymethyl)pyridine side products were also observed in addition to starting material and 3-(bromomethyl)pyridine product.Table 1. Product distributions based on variation in reaction conditions. <n="24"/>[0096] The gas evolved by the reaction was identified by collecting the gas in an IR cell and examining by IR Spectroscopy. The IR spectrum in Figure 2 shows the N2O gas produced by the reaction.[0097] Conversion of 3-aminomethylpyridine (AMP) and 2-(2- aminoethyl)pyrdine (AEP) is efficient, as demonstrated by the HPLC traces before and after reaction as shown in Figure 3.[0098] Titration of the addition of both isoamyl nitrite and Bτ2 to AMP shows that the amine disappears from the reaction mixture on addition of one equivalent each of isoamyl nitrite and Br2. As shown in Figure 7, the formation of the product 3- (bromomethyl)pyridine is out of phase with the disappearance of reacting amine, and in some trials was almost completely absent until greater than one equivalent of isoamyl nitrite and Br2 was added. Best yields were obtained by a total addition of about five equivalents OfBr2.in this example
30% With bromine; acetic acid; sodium nitrite In water [0091] The following example explores reaction conditions and product distributions.[0092] Under reaction conditions identical to those used with Br2 in Example 1, the reaction using I2 failed to produce the corresponding iodide. Using 10 equivalents of ICl however, as the source of halogen, results in the production of the chloride in good yield without the iodide. In a further experiment, using 10 equivalents of IBr, gives again the bromide with no detectable iodide. As per the postulated mechanism, the highly reactive nitryl, and nitrosyl halides are the active intermediates, neither of which is known for the iodide.[0093] Conversion of 3-aminomethylpyridine to the bromide in THF was similar to that in CH2Cl2, but decreased in the series CH2Cl2 > CHCl3 > CCl4 with a <n="23"/>corresponding increase in the formation of the hydroxy product. Reaction in DMF or DMSO led to remaining AMP under the conditions used and also favored the formation of the 3-(hydroxymethyl)pyridine over the bromide.[0094] Using isoamyl nitrite as the nitroso source, and Br2, in water, resulted in an incomplete conversion with very little of the hydroxy product. In all solvents tried using NaNC>2 and acetic acid as the nitroso source, an increased amount of 3- (hydroxymethyl)pyridine was observed with incomplete reaction of the AMP. This latter observation is most likely due to the use of NaNO2 for the nitroso source, as in CH2CI2, increasing additions of acetic acid up to 5 equivalents with respect to AMP, favored the formation of the bromide over the hydroxy product. The stronger acid TFA, however, favored the formation of the hydroxy product over the bromide. Addition of an organic base, TEA, piperidine or DBU was well tolerated up to about 1 equivalent with respect to AMP, with higher amounts leading to less conversion of the amine to the bromide. Reaction to the bromide was concentration dependent with best conversions occurring at > 0.1 molar in AMP (variable only explored in GH2CI2).[0095] Conversion of 3-aminomethyl pyridine to 3-(bromomethyl)pyridine using 1.1 equiv. of nitrosyl agent and 5 equiv. of bromine was conducted in the solvent shown in Table 1 below. A sample of the reaction mixture was taken and quenched by high dilution into acetontrile for ESMS analysis of the products. 3- (hydroxymethyl)pyridine and 3-(isoamyloxymethyl)pyridine side products were also observed in addition to starting material and 3-(bromomethyl)pyridine product.Table 1. Product distributions based on variation in reaction conditions. <n="24"/>[0096] The gas evolved by the reaction was identified by collecting the gas in an IR cell and examining by IR Spectroscopy. The IR spectrum in Figure 2 shows the N2O gas produced by the reaction.[0097] Conversion of 3-aminomethylpyridine (AMP) and 2-(2- aminoethyl)pyrdine (AEP) is efficient, as demonstrated by the HPLC traces before and after reaction as shown in Figure 3.[0098] Titration of the addition of both isoamyl nitrite and Bτ2 to AMP shows that the amine disappears from the reaction mixture on addition of one equivalent each of isoamyl nitrite and Br2. As shown in Figure 7, the formation of the product 3- (bromomethyl)pyridine is out of phase with the disappearance of reacting amine, and in some trials was almost completely absent until greater than one equivalent of isoamyl nitrite and Br2 was added. Best yields were obtained by a total addition of about five equivalents OfBr2.in this example

Reference: [1] Patent: WO2007/117404, 2007, A2, . Location in patent: Page/Page column 21-23
[2] Patent: WO2007/117404, 2007, A2, . Location in patent: Page/Page column 21-23
[3] Patent: WO2007/117404, 2007, A2, . Location in patent: Page/Page column 21-23
[4] Patent: WO2007/117404, 2007, A2, . Location in patent: Page/Page column 21-23
[5] Patent: WO2007/117404, 2007, A2, . Location in patent: Page/Page column 21-23
[6] Patent: WO2007/117404, 2007, A2, . Location in patent: Page/Page column 21-23
[7] Patent: WO2007/117404, 2007, A2, . Location in patent: Page/Page column 21-23
[8] Patent: WO2007/117404, 2007, A2, . Location in patent: Page/Page column 21-23
[9] Patent: WO2007/117404, 2007, A2, . Location in patent: Page/Page column 21-23
  • 15
  • [ 122306-01-8 ]
  • [ 100-55-0 ]
YieldReaction ConditionsOperation in experiment
> 95 %Chromat. at 20℃; for 3 h; Sonication; Schlenk technique First, Pd/AlO(OH) NP catalyst (40.0 mg, 0.16 mmol percent) and 1 mL of H2O/MeOH (v/v = 1/1) were added to a Schlenk tube. Next, the halogenated compound (0.25 mmol) was added. Finally, NaBH4 (0.75 mmol) was added to the reaction mixture and the vessel was closed. The reaction then continued during vigorous stirring under ultrasonic conditions at room temperature and was monitored by GC. Most of the reactions were completed over a time period of 1.5 - 4 h. After completion of the reaction, the catalyst was removed via simple centrifugation at 6000 rpm and then washed three times with methanol and water and allowed to dry for further use. The solvent was evaporated under vacuum. The products were purified by flash column chromatography and the dehalogenated products were then determined by 1H NMR.
Reference: [1] Tetrahedron, 2016, vol. 72, # 39, p. 5898 - 5902
  • 16
  • [ 6968-72-5 ]
  • [ 500-22-1 ]
  • [ 100-55-0 ]
Reference: [1] Synlett, 2008, # 17, p. 2579 - 2582
  • 17
  • [ 59-67-6 ]
  • [ 100-55-0 ]
Reference: [1] Journal of Chemical Research - Part S, 2003, # 8, p. 522 - 525
[2] Journal of the Chemical Society - Perkin Transactions 1, 1996, # 16, p. 1993 - 1994
  • 18
  • [ 626-60-8 ]
  • [ 100-55-0 ]
Reference: [1] Organic Letters, 2012, vol. 14, # 5, p. 1278 - 1281
  • 19
  • [ 10072-09-0 ]
  • [ 100-55-0 ]
Reference: [1] Patent: CN107311917, 2017, A, . Location in patent: Paragraph 0032; 0035; 0040; 0045
  • 20
  • [ 3731-52-0 ]
  • [ 69966-55-8 ]
  • [ 100-55-0 ]
YieldReaction ConditionsOperation in experiment
44% With bromine; acetic acid; sodium nitrite In dichloromethane [0091] The following example explores reaction conditions and product distributions.[0092] Under reaction conditions identical to those used with Br2 in Example 1, the reaction using I2 failed to produce the corresponding iodide. Using 10 equivalents of ICl however, as the source of halogen, results in the production of the chloride in good yield without the iodide. In a further experiment, using 10 equivalents of IBr, gives again the bromide with no detectable iodide. As per the postulated mechanism, the highly reactive nitryl, and nitrosyl halides are the active intermediates, neither of which is known for the iodide.[0093] Conversion of 3-aminomethylpyridine to the bromide in THF was similar to that in CH2Cl2, but decreased in the series CH2Cl2 > CHCl3 > CCl4 with a <n="23"/>corresponding increase in the formation of the hydroxy product. Reaction in DMF or DMSO led to remaining AMP under the conditions used and also favored the formation of the 3-(hydroxymethyl)pyridine over the bromide.[0094] Using isoamyl nitrite as the nitroso source, and Br2, in water, resulted in an incomplete conversion with very little of the hydroxy product. In all solvents tried using NaNC>2 and acetic acid as the nitroso source, an increased amount of 3- (hydroxymethyl)pyridine was observed with incomplete reaction of the AMP. This latter observation is most likely due to the use of NaNO2 for the nitroso source, as in CH2CI2, increasing additions of acetic acid up to 5 equivalents with respect to AMP, favored the formation of the bromide over the hydroxy product. The stronger acid TFA, however, favored the formation of the hydroxy product over the bromide. Addition of an organic base, TEA, piperidine or DBU was well tolerated up to about 1 equivalent with respect to AMP, with higher amounts leading to less conversion of the amine to the bromide. Reaction to the bromide was concentration dependent with best conversions occurring at > 0.1 molar in AMP (variable only explored in GH2CI2).[0095] Conversion of 3-aminomethyl pyridine to 3-(bromomethyl)pyridine using 1.1 equiv. of nitrosyl agent and 5 equiv. of bromine was conducted in the solvent shown in Table 1 below. A sample of the reaction mixture was taken and quenched by high dilution into acetontrile for ESMS analysis of the products. 3- (hydroxymethyl)pyridine and 3-(isoamyloxymethyl)pyridine side products were also observed in addition to starting material and 3-(bromomethyl)pyridine product.Table 1. Product distributions based on variation in reaction conditions. <n="24"/>[0096] The gas evolved by the reaction was identified by collecting the gas in an IR cell and examining by IR Spectroscopy. The IR spectrum in Figure 2 shows the N2O gas produced by the reaction.[0097] Conversion of 3-aminomethylpyridine (AMP) and 2-(2- aminoethyl)pyrdine (AEP) is efficient, as demonstrated by the HPLC traces before and after reaction as shown in Figure 3.[0098] Titration of the addition of both isoamyl nitrite and Bτ2 to AMP shows that the amine disappears from the reaction mixture on addition of one equivalent each of isoamyl nitrite and Br2. As shown in Figure 7, the formation of the product 3- (bromomethyl)pyridine is out of phase with the disappearance of reacting amine, and in some trials was almost completely absent until greater than one equivalent of isoamyl nitrite and Br2 was added. Best yields were obtained by a total addition of about five equivalents OfBr2.in this example
Reference: [1] Patent: WO2007/117404, 2007, A2, . Location in patent: Page/Page column 21-23
  • 21
  • [ 98-92-0 ]
  • [ 100-55-0 ]
  • [ 3731-52-0 ]
Reference: [1] Chemical Science, 2016, vol. 7, # 5, p. 3432 - 3442
  • 22
  • [ 100-54-9 ]
  • [ 100-55-0 ]
  • [ 108-99-6 ]
Reference: [1] Bulletin of the Chemical Society of Japan, 1994, vol. 67, # 4, p. 1107 - 1112
  • 23
  • [ 500-22-1 ]
  • [ 1342295-75-3 ]
  • [ 100-55-0 ]
  • [ 59-67-6 ]
Reference: [1] Organic and Biomolecular Chemistry, 2011, vol. 9, # 20, p. 7023 - 7027
  • 24
  • [ 500-22-1 ]
  • [ 108460-23-7 ]
  • [ 100-55-0 ]
Reference: [1] Tetrahedron Letters, 1986, vol. 27, # 21, p. 2375 - 2378
[2] Tetrahedron Letters, 1986, vol. 27, # 21, p. 2375 - 2378
[3] Tetrahedron Letters, 1986, vol. 27, # 21, p. 2375 - 2378
  • 25
  • [ 500-22-1 ]
  • [ 100-55-0 ]
  • [ 41668-21-7 ]
Reference: [1] European Journal of Organic Chemistry, 2002, # 19, p. 3326 - 3335
  • 26
  • [ 64531-95-9 ]
  • [ 100-55-0 ]
Reference: [1] Journal of the Chemical Society - Perkin Transactions 1, 1996, # 16, p. 1993 - 1994
  • 27
  • [ 6968-72-5 ]
  • [ 16420-13-6 ]
  • [ 100-55-0 ]
Reference: [1] Journal of Heterocyclic Chemistry, 2007, vol. 44, # 2, p. 487 - 489
  • 28
  • [ 85719-72-8 ]
  • [ 100-55-0 ]
Reference: [1] Applied Organometallic Chemistry, 2011, vol. 25, # 2, p. 83 - 86
  • 29
  • [ 500-22-1 ]
  • [ 952-92-1 ]
  • [ 100-55-0 ]
  • [ 15519-25-2 ]
Reference: [1] Tetrahedron Letters, 1986, vol. 27, # 21, p. 2375 - 2378
[2] Tetrahedron Letters, 1986, vol. 27, # 21, p. 2375 - 2378
  • 30
  • [ 500-22-1 ]
  • [ 6290-49-9 ]
  • [ 100-55-0 ]
  • [ 93-60-7 ]
  • [ 136138-66-4 ]
Reference: [1] Tetrahedron, 1991, vol. 47, # 28, p. 5269 - 5276
  • 31
  • [ 13469-94-8 ]
  • [ 500-22-1 ]
  • [ 100-55-0 ]
Reference: [1] Journal of Heterocyclic Chemistry, 1994, vol. 31, # 6, p. 1599 - 1603
  • 32
  • [ 129803-30-1 ]
  • [ 100-55-0 ]
Reference: [1] Chemistry - A European Journal, 2007, vol. 13, # 15, p. 4334 - 4341
  • 33
  • [ 1345455-94-8 ]
  • [ 100-55-0 ]
Reference: [1] Chemical Communications, 2012, vol. 48, # 38, p. 4567 - 4569
  • 34
  • [ 553-53-7 ]
  • [ 100-55-0 ]
Reference: [1] Helvetica Chimica Acta, 1941, vol. 24, p. 24 E, 25 E
  • 35
  • [ 5433-39-6 ]
  • [ 100-55-0 ]
Reference: [1] Helvetica Chimica Acta, 1941, vol. 24, p. 24 E, 25 E
  • 36
  • [ 100-54-9 ]
  • [ 100-55-0 ]
Reference: [1] Diss.&lt;Berlin 1938&gt; S.20,
  • 37
  • [ 77934-69-1 ]
  • [ 118-90-1 ]
  • [ 100-55-0 ]
Reference: [1] Tetrahedron Letters, 1980, vol. 21, p. 4917 - 4920
  • 38
  • [ 100-55-0 ]
  • [ 3731-52-0 ]
YieldReaction ConditionsOperation in experiment
96% With (carbonyl)chloro(hydrido)tris(triphenylphosphine)ruthenium(II); ammonia; [5-(diphenylphosphanyl)-9,9-dimethyl-9H-xanthen-4-yl]diphenylphosphane In tert-Amyl alcohol at 140℃; for 20 h; Inert atmosphere; Cooling Example 6Direct Single-Stage Amination of Alcohols andHydroxy Acids by Means of Ammonia Over aHomogeneous Ruthenium Catalyst and Xantphos ata high V7J Vgas (according to the invention)Under an argon atmosphere, m g of starting material, mRU g of [carbonylchlorohydridotris(triphenylphosphane)ruthenium(II)] and mp g of 9,9-dimethyl-4,5-bis (diphenylphosphino)xanthene as catalyst and V07 ml of 2-methyl-2-butanol as solvent were introduced into a 50 mlsteel tube. The vessel was closed, pressurized three times with 20 bar of argon and depressurized each time. The vessel was then cooled by means of dry ice and m g of ammonia were condensed in. The reactor is heated to T° C. and maintained at this temperature for 20 hours. Afier cooling to room temperature, the reactor was depressurized and opened, the solvent was removed on a rotary evaporator and the residue was dissolved in methanol and then analysed by gas chromatography. Reaction parameters and conversions and selectivities to the desired reaction product are shown in Tab. 5. The results show that many different hydroxy-thnctionalized substrates can be aminated by the method described.
Reference: [1] Patent: US2013/331580, 2013, A1, . Location in patent: Paragraph 0063
[2] Angewandte Chemie - International Edition, 2011, vol. 50, # 33, p. 7599 - 7603
  • 39
  • [ 98-92-0 ]
  • [ 100-55-0 ]
  • [ 3731-52-0 ]
Reference: [1] Chemical Science, 2016, vol. 7, # 5, p. 3432 - 3442
  • 40
  • [ 100-55-0 ]
  • [ 6293-56-7 ]
Reference: [1] European Journal of Organic Chemistry, 2015, vol. 2015, # 32, p. 7009 - 7019
  • 41
  • [ 100-55-0 ]
  • [ 6443-85-2 ]
Reference: [1] ChemMedChem, 2015, vol. 10, # 11, p. 1875 - 1883
  • 42
  • [ 100-55-0 ]
  • [ 95-92-1 ]
  • [ 39931-77-6 ]
Reference: [1] Chemistry - A European Journal, 2013, vol. 19, # 23, p. 7334 - 7337
  • 43
  • [ 100-55-0 ]
  • [ 39931-77-6 ]
Reference: [1] ChemMedChem, 2015, vol. 10, # 11, p. 1875 - 1883
  • 44
  • [ 100-55-0 ]
  • [ 57933-83-2 ]
  • [ 1152-61-0 ]
  • [ 3479-47-8 ]
  • [ 4779-31-1 ]
Reference: [1] Tetrahedron Letters, 1987, vol. 28, # 15, p. 1665 - 1668
[2] Tetrahedron Letters, 1987, vol. 28, # 15, p. 1665 - 1668
[3] Tetrahedron Letters, 1987, vol. 28, # 15, p. 1665 - 1668
  • 45
  • [ 100-55-0 ]
  • [ 57933-83-2 ]
  • [ 1152-61-0 ]
  • [ 3479-47-8 ]
  • [ 4779-31-1 ]
Reference: [1] Tetrahedron Letters, 1987, vol. 28, # 15, p. 1665 - 1668
[2] Tetrahedron Letters, 1987, vol. 28, # 15, p. 1665 - 1668
[3] Tetrahedron Letters, 1987, vol. 28, # 15, p. 1665 - 1668
  • 46
  • [ 100-55-0 ]
  • [ 6959-48-4 ]
YieldReaction ConditionsOperation in experiment
63.6% With thionyl chloride In chloroform at 0℃; Reflux Step B: Preparation of 3-(Chloromethyl)pyridine hydrochloride. Thionyl chloride (9.60 g, 80.6 mmol) was added dropwise to a stirred solution of (pyridin-3-yl)methanol (2.20 g, 20.1 mmol) in 30 ml of CHCl3 at 0° C. and refluxed for 4 h. The reaction mixture was concentrated under reduced pressure to afford 2.10 g (63.6percent) of 3-(chloromethyl)pyridine hydrochloride as a brown solid.
Reference: [1] ChemMedChem, 2015, vol. 10, # 11, p. 1875 - 1883
[2] Patent: US2011/224225, 2011, A1, . Location in patent: Page/Page column 27
  • 47
  • [ 100-55-0 ]
  • [ 10035-10-6 ]
  • [ 31106-82-8 ]
Reference: [1] Dalton Transactions, 2013, vol. 42, # 42, p. 14967 - 14981
  • 48
  • [ 100-55-0 ]
  • [ 39244-80-9 ]
Reference: [1] Journal of the Chemical Society - Perkin Transactions 1, 1998, # 15, p. 2239 - 2241
  • 49
  • [ 100-55-0 ]
  • [ 209783-80-2 ]
Reference: [1] Journal of Medicinal Chemistry, 1999, vol. 42, # 15, p. 3001 - 3003
  • 50
  • [ 100-55-0 ]
  • [ 209784-85-0 ]
  • [ 209783-80-2 ]
Reference: [1] Patent: WO2017/81278, 2017, A1,
Same Skeleton Products
Historical Records

Related Functional Groups of
[ 100-55-0 ]

Alcohols

Chemical Structure| 102074-19-1

[ 102074-19-1 ]

(5-Methylpyridin-3-yl)methanol

Similarity: 0.89

Chemical Structure| 38070-80-3

[ 38070-80-3 ]

Pyridine-3,4-diyldimethanol

Similarity: 0.89

Chemical Structure| 113293-71-3

[ 113293-71-3 ]

(6-Aminopyridin-3-yl)methanol

Similarity: 0.85

Chemical Structure| 34107-46-5

[ 34107-46-5 ]

(6-Methylpyridin-3-yl)methanol

Similarity: 0.85

Chemical Structure| 120277-39-6

[ 120277-39-6 ]

(E)-3-(Pyridin-3-yl)prop-2-en-1-ol

Similarity: 0.84

Related Parent Nucleus of
[ 100-55-0 ]

Pyridines

Chemical Structure| 102074-19-1

[ 102074-19-1 ]

(5-Methylpyridin-3-yl)methanol

Similarity: 0.89

Chemical Structure| 38070-80-3

[ 38070-80-3 ]

Pyridine-3,4-diyldimethanol

Similarity: 0.89

Chemical Structure| 113293-71-3

[ 113293-71-3 ]

(6-Aminopyridin-3-yl)methanol

Similarity: 0.85

Chemical Structure| 34107-46-5

[ 34107-46-5 ]

(6-Methylpyridin-3-yl)methanol

Similarity: 0.85

Chemical Structure| 120277-39-6

[ 120277-39-6 ]

(E)-3-(Pyridin-3-yl)prop-2-en-1-ol

Similarity: 0.84