Search results

Search for "GC/MS" in Full Text gives 251 result(s) in Beilstein Journal of Organic Chemistry. Showing first 200.

Crystal design using multipolar electrostatic interactions: A concept study for organic electronics

  • Peer Kirsch,
  • Qiong Tong and
  • Harald Untenecker

Beilstein J. Org. Chem. 2013, 9, 2367–2373, doi:10.3762/bjoc.9.272

Graphical Abstract
  • processability. Experimental General remarks: Reagents and solvents were obtained commercially and used as supplied. 1H and 19F NMR spectra were collected using a Bruker Avance 400 spectrometer. 1H NMR chemical shifts were referenced to the solvent signal. GCMS experiments were performed on an Agilent 7890A
PDF
Album
Full Research Paper
Published 05 Nov 2013

Elucidation of the regio- and chemoselectivity of enzymatic allylic oxidations with Pleurotus sapidus – conversion of selected spirocyclic terpenoids and computational analysis

  • Verena Weidmann,
  • Mathias Schaffrath,
  • Holger Zorn,
  • Julia Rehbein and
  • Wolfgang Maison

Beilstein J. Org. Chem. 2013, 9, 2233–2241, doi:10.3762/bjoc.9.262

Graphical Abstract
  • hydrogen abstraction in position 2, spiroether 7 was selectively converted to the spirolactone 14. The alternative oxidation product 15 was not found by GCMS. The regioisomeric spiroether 8 offers only one plausible path for allylic oxidation in position 3. However, this spiroether was not converted at
PDF
Album
Supp Info
Full Research Paper
Published 29 Oct 2013

A concise enantioselective synthesis of the guaiane sesquiterpene (−)-oxyphyllol

  • Martin Zahel and
  • Peter Metz

Beilstein J. Org. Chem. 2013, 9, 2028–2032, doi:10.3762/bjoc.9.239

Graphical Abstract
  • (triplet), q quartet), br (broad). Coupling constants (J) are quoted to the nearest 0.1 Hz. Mass spectra were recorded with an Agilent 5973N detector coupled with an Agilent 6890N GC (GCMS, 70 eV). HRMS spectra were recorded on a Finnigan MAT 95 (EI, 70 eV). Optical rotations were measured on a Perkin
  • , 977 cm−1; GCMS: m/z = 296 [M]+; anal. calcd for C15H20O4S: C 60.79, H 6.80; found: C 60.88, H 6.92. Alcohol 3: Epoxide 4 (100.0 mg, 454 μmol) and cobalt(II) acetylacetonate (23.3 mg, 91 μmol) were dissolved in THF (5 mL), and the solution was cooled to 0 °C. Oxygen was bubbled through the solution
  • : 3446, 2957, 2930, 2874, 1708, 1130, 852 cm−1; GCMS: m/z = 238 [M]+; HRMS (EI) m/z: [M]+ calcd for C14H22O3, 238.1569; found, 238.1564. Allyl ether 11: Methyltriphenylphosphonium bromide (131.5 mg, 368 μmol) was suspended in THF (1.5 mL), and a solution of sodium hexamethyldisilazide in THF (2.0 M, 172
PDF
Album
Supp Info
Full Research Paper
Published 08 Oct 2013

Damage of polyesters by the atmospheric free radical oxidant NO3: a product study involving model systems

  • Catrin Goeschen and
  • Uta Wille

Beilstein J. Org. Chem. 2013, 9, 1907–1916, doi:10.3762/bjoc.9.225

Graphical Abstract
  • )] in deuterated DMSO. If necessary the assignment of the chemical shifts was confirmed by utilising 2D NMR techniques. GCMS (EI, 70 eV) analysis was run on an Agilent 7890A GC/5975C MSD, column from SUPELCO 30 m, 0.32 mm ID, 0.25 μm film thickness fused silica capillary column, using the temperature
PDF
Album
Supp Info
Full Research Paper
Published 20 Sep 2013

Integrating reaction and analysis: investigation of higher-order reactions by cryogenic trapping

  • Skrollan Stockinger and
  • Oliver Trapp

Beilstein J. Org. Chem. 2013, 9, 1837–1842, doi:10.3762/bjoc.9.214

Graphical Abstract
  • evaporation the reaction mixture was dissolved in pentane and washed with water. The diluted pentane solution was analyzed by GCMS. For the ocRGC analysis separate DCM solutions of anthracene or anthracene derivative and benzenediazonium-2-carboxylate were prepared, both containing 1 mg per mL. The injection
  • -order reactions. Experimental Benzenediazonium-2-carboxylate was synthesized according to a procedure described in [33][34]. Anthracene, 9-bromoanthracene, 9-anthracenecarboxaldehyde and 9-anthracenemethanol were purchased from Aldrich Chemical, Acros or ABCR and used without further purification. As GC
  • MS a Trace GC Ultra combined with an ISQ Single Quadrupole mass spectrometer, both from Thermo Scientific, were used. For cryofocusing a Cryogenic CO2 Cold Trap System by SGE Analytical Science was installed, according to the experimental setup outlined in Figure 2. For cryofocusing CO2 was used as a
PDF
Album
Letter
Published 10 Sep 2013

Activation of cryptic metabolite production through gene disruption: Dimethyl furan-2,4-dicarboxylate produced by Streptomyces sahachiroi

  • Dinesh Simkhada,
  • Huitu Zhang,
  • Shogo Mori,
  • Howard Williams and
  • Coran M. H. Watanabe

Beilstein J. Org. Chem. 2013, 9, 1768–1773, doi:10.3762/bjoc.9.205

Graphical Abstract
  • characterization of the major metabolite was carried out using LC–MS analysis and 1D and 2D cryoprobe NMR spectroscopy (data provided in Supporting Information File 1). The molecular formula was established as C8H8O5 based upon a molecular ion peak observed by LC–APCIMS at m/z 185.0453 [M + H]+ and GCMS. The GC
  • MS fragmentation of the compound matched that provided in the Ultra GC–DSQ (ThermoElectron, Waltham, MA) GCMS database. The structure of the compound was determined by NMR analysis in CDCl3 (Table 1). The presence of two O-CH3 groups at C-2’ (δc 51.99, δH 3.94) and C-2” (δc 52.31, δH 3.88) and
  • flow rate of 0.75 mL/min in 254 nm UV. The same program was used for LC–APCI analysis. For GCMS analysis, the sample was diluted 1:10 in DCM and was performed on an Ultra GC/DSQ (ThermoElectron, Waltham, MA). A Rxi-5ms gas chromatographic column was utilized with dimensions of 60 m length, 0.25 mm i.d
PDF
Album
Supp Info
Letter
Published 29 Aug 2013

Gold-catalyzed intermolecular coupling of sulfonylacetylene with allyl ethers: [3,3]- and [1,3]-rearrangements

  • Jungho Jun,
  • Hyu-Suk Yeom,
  • Jun-Hyun An and
  • Seunghoon Shin

Beilstein J. Org. Chem. 2013, 9, 1724–1729, doi:10.3762/bjoc.9.198

Graphical Abstract
  • the presence of the Au-catalyst, no cross-over product was observed by GCMS and NMR spectrometry, indicating the [3,3]- and [1,3]-rearrangement occurred intramolecularly (Scheme 2, reaction 2). This was further confirmed by the cross-over experiment employing 2i and 2m, two α-substituted allyl ethers
PDF
Album
Supp Info
Full Research Paper
Published 22 Aug 2013

The preparation of several 1,2,3,4,5-functionalized cyclopentane derivatives

  • André S. Kelch,
  • Peter G. Jones,
  • Ina Dix and
  • Henning Hopf

Beilstein J. Org. Chem. 2013, 9, 1705–1712, doi:10.3762/bjoc.9.195

Graphical Abstract
  • derivatives of 16, it should also be mentioned that attempts to prepare the corresponding pentachloride (by treatment with either SOCl2 or PCl3) and pentaiodide (red P/iodine) also failed. In all these experiments complex product mixtures resulted, the GCMS analysis of which demonstrated that only partially
  •  5). Finally, the hydrogenation of 28 over Pd/C in methanol furnished a mixture of the saturated pentaesters 26 and 27 (ratio 4:1, capillary GCMS-analysis) in practically quantitative yield (98%). Although we were unable to separate the two components quantitatively by column chromatography on
PDF
Album
Supp Info
Full Research Paper
Published 19 Aug 2013

Bromination of hydrocarbons with CBr4, initiated by light-emitting diode irradiation

  • Yuta Nishina,
  • Bunsho Ohtani and
  • Kotaro Kikushima

Beilstein J. Org. Chem. 2013, 9, 1663–1667, doi:10.3762/bjoc.9.190

Graphical Abstract
  • species for the bromination. The generation of CHBr3 was confirmed by NMR spectroscopy and GCMS spectrometry analysis, indicating that the present bromination involves the homolytic cleavage of a C–Br bond in CBr4 followed by radical abstraction of a hydrogen atom from a hydrocarbon. Keywords
  • (0.50 mmol) dissolved in cyclohexane (0.10 mL) and CDCl3 (0.40 mL) (Figure 2b). Additionally, the generation of CHBr3 in the present bromination was confirmed by 1H NMR spectroscopy and GCMS spectrometry. These results support the reaction pathway described above, although the chemical species after
PDF
Album
Letter
Published 14 Aug 2013

Mechanistic studies on the CAN-mediated intramolecular cyclization of δ-aryl-β-dicarbonyl compounds

  • Brian M. Casey,
  • Dhandapani V. Sadasivam and
  • Robert A. Flowers II

Beilstein J. Org. Chem. 2013, 9, 1472–1479, doi:10.3762/bjoc.9.167

Graphical Abstract
  • system from Teledyne Isco, Inc. All new compounds were characterized by 1H NMR, 13C NMR, GCMS, IR, and LC–HRMS. Known compounds were characterized by 1H NMR, 13C NMR and GCMS. 1H NMR and 13C NMR spectra were recorded on a Bruker 500 MHz spectrometer. Mass spectra were obtained by using a HP 5890 series
  • GCMS instrument. A Satellite FTIR from Thermo-Mattson was used to obtain IR spectra. LC–HRMS data were recorded at the Mass Spectrometry Facility at Notre Dame University. General procedure for the synthesis of δ-aryl-β-dicarbonyl compounds 1a–j. Sodium hydride (11 mmol) was suspended in 25 mL of
PDF
Album
Supp Info
Full Research Paper
Published 23 Jul 2013

Isolation and X-ray characterization of palladium–N complexes in the guanylation of aromatic amines. Mechanistic implications

  • Abdessamad Grirrane,
  • Hermenegildo Garcia and
  • Eleuterio Álvarez

Beilstein J. Org. Chem. 2013, 9, 1455–1462, doi:10.3762/bjoc.9.165

Graphical Abstract
  • S21 and S22 for 5a, Figures S24 and S25 for 5b, our recent published work for 5c [20] and experimental section in Supporting Information File 1). ESIMS and GCMS of solutions obtained respectively after dissolving guanidines 5a and 5b in CH2Cl2/MeOH (1:1) and CH2Cl2 shows a single positive MS peak at
  • and GCMS spectra for new compounds. Supporting Information File 380: X-Ray structure analysis data for 4a (CCDC-931786, 4b (CCDC-931787), 4c (CCDC-931788) and 5a (CCDC-931789) are given. These data can also be obtained free of charge from The Cambridge Crystallographic Data Centre via http
PDF
Album
Supp Info
Full Research Paper
Published 22 Jul 2013

Palladium-catalyzed synthesis of N-arylated carbazoles using anilines and cyclic diaryliodonium salts

  • Stefan Riedmüller and
  • Boris J. Nachtsheim

Beilstein J. Org. Chem. 2013, 9, 1202–1209, doi:10.3762/bjoc.9.136

Graphical Abstract
  • equal or better in efficiency and yield. Changing Pd2(dba)3 to Pd(OAc)2 had no significant impact on product yields (Table 1, entry 12). Further increase of the catalyst ratio from 5 to 10 mol % had little effect (Table 1, entry 15). Next, we decided to analyze the byproducts of this reaction by GCMS
  • GCMS analysis, we could detect 2-iodobiphenyl, 2,2'-diiodobiphenyl, 2-(2,5-dimethylphenyl)-2'-iodobiphenyl, and 2'-iodo-N-phenylbiphenyl-2-amine in the absence of any produced 3a. Contrary to that observation, when we conducted an analogous experiment without aniline, we observed no decomposition or
PDF
Album
Supp Info
Full Research Paper
Published 21 Jun 2013

Interplay of ortho- with spiro-cyclisation during iminyl radical closures onto arenes and heteroarenes

  • Roy T. McBurney and
  • John C. Walton

Beilstein J. Org. Chem. 2013, 9, 1083–1092, doi:10.3762/bjoc.9.120

Graphical Abstract
  • communicated previously, 3-substituted phenanthridines 10a–f were isolated in good to quantitative yields (52–99%, Scheme 1) irrespective of the nature of the 4-substituent [26]. No spiro-products were detected even by GCMS analyses of reaction mixtures. Byproducts included traces of imines ArC(R2)=NH (ImH
PDF
Album
Supp Info
Full Research Paper
Published 04 Jun 2013

Isotopically labeled sulfur compounds and synthetic selenium and tellurium analogues to study sulfur metabolism in marine bacteria

  • Nelson L. Brock,
  • Christian A. Citron,
  • Claudia Zell,
  • Martine Berger,
  • Irene Wagner-Döbler,
  • Jörn Petersen,
  • Thorsten Brinkhoff,
  • Meinhard Simon and
  • Jeroen S. Dickschat

Beilstein J. Org. Chem. 2013, 9, 942–950, doi:10.3762/bjoc.9.108

Graphical Abstract
  • ) were used in feeding experiments with the Roseobacter clade members Phaeobacter gallaeciensis DSM 17395 and Ruegeria pomeroyi DSS-3, and their volatile metabolites were analyzed by closed-loop stripping and solid-phase microextraction coupled to GCMS. Feeding experiments with [2H6]DMSP resulted in the
  • by solid-phase microextraction (SPME) and by CLSA. The obtained headspace extracts were subsequently analyzed by GCMS, leading to the identification of the volatiles dimethyl disulfide (1), dimethyl trisulfide (2), S-methyl methanethiosulfonate (3), dimethyl telluride (4), dimethyl telluryl sulfide
  • GCMS. No incorporation of the 34S-labeling from sulfate or thiosulfate into the MeSH-derived volatiles was observed, pointing strongly away from a significant involvement of sulfate reduction in the biosynthesis of MeSH. Instead, feeding of [methyl-2H3]methionine led to an effective incorporation of
PDF
Album
Supp Info
Full Research Paper
Published 15 May 2013

4-Pyridylnitrene and 2-pyrazinylcarbene

  • Curt Wentrup,
  • Ales Reisinger and
  • David Kvaskoff

Beilstein J. Org. Chem. 2013, 9, 754–760, doi:10.3762/bjoc.9.85

Graphical Abstract
  • ’-Azopyridine (25) (54% yield; red solid, mp 107–108 °C) was identified by comparison with a sample prepared according to den Hertog et al. (mp 107.5 °C) [22]. Traces of unreacted azide 18 and 2- and 3-cyanopyrroles 16 and 17were detected by GCMS and IR spectroscopy. “Violent pyrolysis” was achieved by holding
PDF
Album
Supp Info
Full Research Paper
Published 17 Apr 2013
Graphical Abstract
  • g) was distilled into the FVT apparatus from a sample flask held at −30 °C and thermolysed at 370–500 °C/10−3–10−4 mbar in the course of 2 h. The pyrolysate was examined by GC/MS and 1H NMR spectroscopy. The 2- and 3-cyanopyrroles 7 and 8 were separated by flash chromatography on silica gel 60
  • remained unreacted); 400 °C: 1.63:1 (36% 3-azidopyridine remained unreacted); 500 °C: 1.48:1 (6% 3-azidopyridine remained unreacted). GC/MS retention times and molecular masses: 2-cyanopyrrole 3.60 min (m/z 92); 3-cyanpyrrole 5.45 min (m/z 92); 3-azidopyridine 2.19 min (m/z 120); 1H NMR (CDCl3) 2
PDF
Album
Supp Info
Full Research Paper
Published 17 Apr 2013

Quantification of N-acetylcysteamine activated methylmalonate incorporation into polyketide biosynthesis

  • Stephan Klopries,
  • Uschi Sundermann and
  • Frank Schulz

Beilstein J. Org. Chem. 2013, 9, 664–674, doi:10.3762/bjoc.9.75

Graphical Abstract
  • EtOAc. The combined organic layers were dried over Na2SO4 and purified by column chromatography (DCM/MeOH 99:1) to obtain 7.56 g (61%) of the desired product as a colorless oil. Rf 0.42 (DCM/MeOH 9:1); HRMS (GCMS): [M + H]+ calcd for C4H10ONS, 120.04776; found, 120.04730; 1H NMR (400 MHz, CDCl3-d1
  • :5) to obtain 200 mg (37%) of 2 as a slightly yellow oil. Rf 0.5 (DCM/MeOH 9:1); HRMS (GCMS): [M + H]+ calcd for C8H122H3O4, 178.11532; found, 178.11512; 1H NMR (400 MHz, CDCl3-d1) 1.46 (s, 9H), 3.36 (s, 1H); 13C NMR (101 MHz, CDCl3-d1) 21.1, 28.2, 47.0 ,82.7, 169.6, 176.6. Synthesis of tert-butyl 3
PDF
Album
Supp Info
Full Research Paper
Published 05 Apr 2013

Kinetics and mechanism of the anilinolysis of aryl phenyl isothiocyanophosphates in acetonitrile

  • Hasi Rani Barai and
  • Hai Whang Lee

Beilstein J. Org. Chem. 2013, 9, 615–620, doi:10.3762/bjoc.9.68

Graphical Abstract
  • , 1H, aliphatic), 7.01–7.43 (m, 15H, aromatic); 13C NMR (100 MHz, MeCN-d3) δ 118.59–131.20 (m, 18C; aromatic); 31P NMR (162 MHz, MeCN-d3) δ 3.62 (d, J = 8.6 Hz, 1P, P=O); GCMS (EI, m/z): 325 (M+). Brönsted plots with X [log kH versus pKa(X)] of the reactions of Y-aryl phenyl isothiocyanophosphates
PDF
Album
Supp Info
Full Research Paper
Published 26 Mar 2013
Graphical Abstract
  • % isolated yield. Quenching with MeOD afforded an authentic sample of 4 deuterated at the C(2) position (2H NMR: δ 5.84–5.75 (m, 1D)); the lack of a molecular ion in the GCMS of triene 4 precluded accurate determination of the deuterium content. Addition at –78 °C of dry, oxygen-free N,N,N’,N
  • GC and by GCMS affording baseline separation of the three products (4–6), illustrated in Scheme 4, which accounted for essentially the total material balance. The structures of the bicyclic products, 5 and 6, were established as detailed in the Experimental Section (see Supporting Information File 1
  • for details) by NMR and GCMS: an authentic sample of 5 was prepared as illustrated below and 6 is a known compound [20]. It is noteworthy that no 1-methylstellane was detected as a product from any of the reactions. Stirring a reaction mixture for 1 h at room temperature demonstrated that the first
PDF
Album
Supp Info
Full Research Paper
Published 14 Mar 2013

Easy and direct conversion of tosylates and mesylates into nitroalkanes

  • Alessandro Palmieri,
  • Serena Gabrielli and
  • Roberto Ballini

Beilstein J. Org. Chem. 2013, 9, 533–536, doi:10.3762/bjoc.9.58

Graphical Abstract
  • methodology [42]. 1H NMR was recorded at 400 MHz on a Varian Mercury Plus 400. 13C NMR was recorded at 100 MHz. Microanalyses were performed with a CHNS-O analyser Model EA 1108 from Fisons Instruments. Mass spectra were performed with a GCMS system Agilent Technologies 6850 II/5973 Inert by means of the EI
  • –n. Supporting Information All synthesized products were characterized by IR, NMR, GCMS and elementary analysis. The data is reported in Supporting Information File 1, while copies of 1H and 13C NMR spectra are attached as Supporting Information File 2. Supporting Information File 365
PDF
Album
Supp Info
Full Research Paper
Published 14 Mar 2013

Efficient Cu-catalyzed base-free C–S coupling under conventional and microwave heating. A simple access to S-heterocycles and sulfides

  • Silvia M. Soria-Castro and
  • Alicia B. Peñéñory

Beilstein J. Org. Chem. 2013, 9, 467–475, doi:10.3762/bjoc.9.50

Graphical Abstract
  • amounts of the substituted thioacetate derivative S-(2-acetoamidophenyl)thioacetate (15) and N-(2-mercaptophenyl)acetamide (16) or S-(2-aminophenyl)thioacetate (17) without cyclization, as detected by GCMS spectrometry (Scheme 6). In order to accelerate the ring-closure reaction, a mixture of 13 and 1 in
PDF
Album
Supp Info
Full Research Paper
Published 04 Mar 2013

Synthesis of 5-(ethylsulfonyl)-2-methoxyaniline: An important pharmacological fragment of VEGFR2 and other inhibitors

  • Miroslav Murár,
  • Gabriela Addová and
  • Andrej Boháč

Beilstein J. Org. Chem. 2013, 9, 173–179, doi:10.3762/bjoc.9.20

Graphical Abstract
  • acquired on FT-IR-ATR REACT IR 1000 (ASI Applied Systems) with diamond probe and MTS detector. Mass spectra were performed on LC–MS (Agilent Technologies 1200 Series equipped with Mass spectrometer Agilent Technologies 6100 Quadrupole LC–MS) and GCMS (Agilent Technologies 6890N gas chromatograph with a
  • one occurring in the range of 134.0–126.3 and, therefore, cannot be exactly assigned; GCMS m/z: 245.1 (m, M+), 216.0 (s, M+ − Et), 214.0 (m, M+ − OMe), 200.0 (m, M+ − NO2 + H), 198.0 (m, M+ − NO2 − H), 168.1 (m, M+ − NO2 − OMe), 124.0 (m), 122.0 (m), 106.1 (m), 76.1 (m); Anal. calcd for C9H11NO5S
PDF
Album
Full Research Paper
Published 25 Jan 2013

Polar reactions of acyclic conjugated bisallenes

  • Reiner Stamm and
  • Henning Hopf

Beilstein J. Org. Chem. 2013, 9, 36–48, doi:10.3762/bjoc.9.5

Graphical Abstract
  • 17 and 19 only GCMS data were available, and the given structures are hence speculative and rest only on their mass spectra. Hydrocarbon 14 was obtained in 80% purity by preparative gas chromatography; its complete separation from 17 failed. The bisallyl derivative 20 was isolated in analytically
  • trimethylsilyl substituent. As it turned out, 3 is considerably less reactive than 2: after comparable reactions times (see Supporting Information File 1) only 37% of the substrate had been converted into products. As shown by GCMS analysis the reaction mixture contained four mono silylated products (m/z = 206
  • Scheme 6 all of these experiments failed. Neither could we prepare the “coupling product” 27 by treatment of 4 with the biselectrophile dimethylysilyl dichloride nor the dimer 28 by the direct action of iodine on the organolithium compound 4. In this latter case we noted after work-up and GCMS analysis
PDF
Album
Supp Info
Full Research Paper
Published 08 Jan 2013

Chemical–biological characterization of a cruzain inhibitor reveals a second target and a mammalian off-target

  • Jonathan W. Choy,
  • Clifford Bryant,
  • Claudia M. Calvet,
  • Patricia S. Doyle,
  • Shamila S. Gunatilleke,
  • Siegfried S. F. Leung,
  • Kenny K. H. Ang,
  • Steven Chen,
  • Jiri Gut,
  • Juan A. Oses-Prieto,
  • Jonathan B. Johnston,
  • Michelle R. Arkin,
  • Alma L. Burlingame,
  • Jack Taunton,
  • Matthew P. Jacobson,
  • James M. McKerrow,
  • Larissa M. Podust and
  • Adam R. Renslo

Beilstein J. Org. Chem. 2013, 9, 15–25, doi:10.3762/bjoc.9.3

Graphical Abstract
  • employing GC/MS as reported previously for compound 5 [33]. The GC/MS trace for uninfected host cells establishes that the additional peaks observed in infected cells are of T. cruzi origin (peaks labeled a-i, Figure 4). Treatment with the known TcCYP51 inhibitor posaconazole produces an increase in the
  • . Synthesis of the additional control compound 6–8, the reduced forms of analogues 1, 3, and 4 respectively. GC/MS analysis of lipid extracts from T. cruzi parasites treated with test compounds. DMSO and K777 (1) were used as negative controls; posaconazole served as a positive control. The analysis of 4 was
  • experimentally determined and computationally predicted binding affinities, additional GC/MS spectra from lipid-analysis studies, time courses for reaction of compounds 1 and 6 with glutathione in vitro, and synthetic schemes for analogues 4, 9, 11, 12, and 13, as well as experimental procedures. Supporting
PDF
Album
Supp Info
Full Research Paper
Published 04 Jan 2013

A novel asymmetric synthesis of cinacalcet hydrochloride

  • Veera R. Arava,
  • Laxminarasimhulu Gorentla and
  • Pramod K. Dubey

Beilstein J. Org. Chem. 2012, 8, 1366–1373, doi:10.3762/bjoc.8.158

Graphical Abstract
  • was removed by recrystallizing the crude product (Table 1, entry 5) in n-hexane at 0 °C twice and the pure product was obtained in 90% yield. The obtained product purity was >99% (GCMS) which was sufficient for further conversions. We determined that 14 was not a carryover impurity from 8 by spiking
  • recrystallized twice from n-hexane at 0 °C. Decanting the hexane layer from the crystalline solid affords the pure 10 (clear liquid at 25 °C). Yield = 47.3 g (90%); purity (GCMS) = 99.45%; IR (neat): 3300–3400 (–COOH), 1712 (–CO–) cm−1; 1H NMR (CDCl3/TMS) δ 2.72 (t, 2H, –CH2), 3.03 (t, 2H, –CH2), 7.41 (d, 2H
PDF
Album
Supp Info
Letter
Published 24 Aug 2012
Other Beilstein-Institut Open Science Activities