Properties of Pyrazine (C4H4N2):
Alternative Names1,3-Diazine ''m''-Diazine 1,3-Diazabenzene Elemental composition of C4H4N2
Related compounds
Pyrimidine (C4H4N2): Chemical CompoundScientific Review Article | Chemistry Reference Series
AbstractPyrimidine, a six-membered heterocyclic aromatic compound with the molecular formula C4H4N2, represents one of the three diazine structural isomers, characterized by nitrogen atoms at the 1 and 3 ring positions. This π-deficient aromatic system exhibits a melting point of 20.0–22.0 °C and a boiling point of 123.0–124.0 °C. With a density of 1.016 g·cm−3 at 25 °C, pyrimidine is miscible with water and most common organic solvents. The compound demonstrates significantly reduced basicity compared to pyridine, with a pKa of 1.10 for its conjugate acid. Pyrimidine serves as the fundamental structural motif for numerous biologically significant derivatives and finds extensive applications in pharmaceutical chemistry, materials science, and coordination chemistry. Its electronic structure and reactivity patterns make it an essential subject of study in heterocyclic chemistry. IntroductionPyrimidine constitutes a fundamental class of nitrogen-containing heterocyclic compounds in organic chemistry. As one of the three diazine isomers, alongside pyrazine and pyridazine, pyrimidine features a six-membered aromatic ring containing two nitrogen atoms in meta-relationship. The systematic name 1,3-diazine precisely describes its molecular architecture. First synthesized in laboratory conditions by Gabriel and Colman in 1900 through reduction of 2,4,6-trichloropyrimidine, the parent compound has since become recognized as a cornerstone of heterocyclic chemistry. The pyrimidine ring system demonstrates exceptional versatility in chemical behavior and serves as the structural foundation for numerous derivatives of biological and industrial significance. Its electronic configuration places it among the π-deficient heteroaromatic systems, profoundly influencing its physical properties and chemical reactivity. Molecular Structure and BondingMolecular Geometry and Electronic StructurePyrimidine crystallizes in a planar hexagonal geometry with bond lengths demonstrating aromatic character. X-ray diffraction studies reveal C–N bond distances of approximately 1.337 Å and C–C bond distances of 1.390 Å, consistent with delocalized π-electron systems. The molecular point group symmetry is C2v, with the principal C2 axis bisecting the ring through the C2–C5 bond. Bond angles within the ring measure approximately 116° at carbon atoms and 124° at nitrogen atoms, deviating slightly from ideal hexagonal geometry due to the presence of heteroatoms. Molecular orbital theory describes pyrimidine as possessing a Hückel aromatic system containing six π-electrons, fulfilling the 4n+2 rule for aromaticity. The nitrogen atoms adopt sp2 hybridization, with lone pairs occupying orbitals in the molecular plane. These lone pairs contribute minimal electron density to the aromatic π-system, resulting in significant π-deficiency. Quantum mechanical calculations indicate highest occupied molecular orbital (HOMO) energy of -9.3 eV and lowest unoccupied molecular orbital (LUMO) energy of -0.9 eV, characterizing pyrimidine as an electron-poor aromatic system. Proton NMR spectroscopy confirms the aromatic nature through deshielded proton signals between δ 7.0 and 8.8 ppm. Chemical Bonding and Intermolecular ForcesCovalent bonding in pyrimidine follows typical aromatic patterns with bond orders intermediate between single and double bonds. The C–N bonds exhibit partial double bond character with bond dissociation energies estimated at 110 kcal·mol−1. The molecular dipole moment measures 2.33 D, oriented from the center of the ring toward the nitrogen atoms, reflecting the asymmetric electron distribution. Intermolecular forces in pyrimidine include significant dipole-dipole interactions due to the substantial molecular polarity. The compound demonstrates limited hydrogen bonding capability, acting primarily as a hydrogen bond acceptor through its nitrogen lone pairs. The nitrogen atoms have hydrogen bond acceptor capacities characterized by β values of approximately 0.90 in the Kamlet-Taft solvatochromic parameters. Van der Waals forces contribute significantly to crystal packing, with the planar molecules stacking in parallel arrangements with interplanar distances of 3.4–3.6 Å. These stacking interactions facilitate the formation of stable crystalline phases with melting behavior sensitive to purity. Physical PropertiesPhase Behavior and Thermodynamic PropertiesPyrimidine exists as a colorless liquid at room temperature with a characteristic odor reminiscent of pyridine. The compound exhibits a melting point range of 20.0–22.0 °C and boils at 123.0–124.0 °C under standard atmospheric pressure. The enthalpy of fusion measures 10.5 kJ·mol−1, while the enthalpy of vaporization is 38.5 kJ·mol−1. The heat capacity at 25 °C is 142.3 J·mol−1·K−1 for the liquid phase. The temperature dependence of vapor pressure follows the Antoine equation with parameters A=4.132, B=1420, and C=-63.5 for the range 20–124 °C. The density of liquid pyrimidine decreases linearly from 1.028 g·cm−3 at 20 °C to 0.989 g·cm−3 at 100 °C. The refractive index nD20 measures 1.499, with temperature coefficient dn/dT = -4.5×10−4 K−1. Surface tension at 20 °C is 38.5 mN·m−1 and viscosity measures 1.24 mPa·s at the same temperature. Spectroscopic CharacteristicsInfrared spectroscopy of pyrimidine reveals characteristic aromatic C–H stretching vibrations at 3075 cm−1 and ring stretching vibrations between 1600–1400 cm−1. The out-of-plane C–H bending vibrations appear at 810 cm−1 and 705 cm−1, consistent with meta-disubstituted aromatic systems. Proton NMR spectroscopy in CDCl3 displays three distinct signals: a doublet at δ 8.82 ppm (J = 5.1 Hz) assigned to H2/H6, a triplet at δ 7.19 ppm (J = 5.1 Hz) assigned to H4/H5, and a doublet at δ 7.05 ppm (J = 5.1 Hz) assigned to H3/H5. Carbon-13 NMR shows signals at δ 157.8 ppm (C2/C6), δ 157.2 ppm (C4), and δ 122.5 ppm (C3/C5). UV-Vis spectroscopy exhibits absorption maxima at λmax = 243 nm (ε = 2,040 M−1·cm−1) and 298 nm (ε = 430 M−1·cm−1) in hexane solution, corresponding to π→π* transitions. Mass spectrometry demonstrates a molecular ion peak at m/z 80 with characteristic fragmentation patterns including loss of HCN (m/z 53) and ring cleavage fragments. Chemical Properties and ReactivityReaction Mechanisms and KineticsPyrimidine undergoes electrophilic aromatic substitution preferentially at the 5-position, the site of highest electron density in the π-deficient system. Nitration with mixed acid requires vigorous conditions (100 °C, 4 hours) yielding 5-nitropyrimidine with 65% efficiency. Halogenation proceeds similarly with bromine in acetic acid at 80 °C giving 5-bromopyrimidine. The rate constant for bromination measures 2.3×10−4 M−1·s−1 at 25 °C, significantly slower than benzene derivatives. Nucleophilic substitution occurs readily at the 2-, 4-, and 6-positions. Reaction with ammonia at 150 °C yields 2-aminopyrimidine with second-order kinetics (k = 1.8×10−3 M−1·s−1). Alkylation with organolithium reagents proceeds through addition-elimination mechanisms with rate-determining addition steps. Hydrogenation over platinum catalyst at 100 °C and 3 atm pressure affords tetrahydropyrimidine with ΔH = -45 kJ·mol−1. The compound demonstrates thermal stability up to 400 °C, above which decomposition yields hydrogen cyanide and acetylene as primary products. Acid-Base and Redox PropertiesPyrimidine functions as a weak base with pKa = 1.10 for its conjugate acid in aqueous solution at 25 °C. Protonation occurs exclusively at N1, with the protonated species exhibiting increased aromatic character. The basicity constant follows the Hammett equation with σm = 0.88 for the nitrogen heteroatoms. The redox behavior shows irreversible oxidation at +1.85 V versus SCE in acetonitrile and reduction at -1.12 V, corresponding to electron addition to the LUMO. Electrochemical reduction proceeds through a one-electron transfer forming a radical anion with half-life of 0.3 milliseconds in DMF. The compound demonstrates stability in acidic media below pH 2 but undergoes gradual hydrolysis in strong acid at elevated temperatures. In alkaline conditions, pyrimidine remains stable up to pH 12, with no significant decomposition observed over 24 hours at 25 °C. Oxidation with peracids yields N-oxide derivatives selectively at N1 with second-order rate constants of 0.15 M−1·s−1 for mCPBA in chloroform. Synthesis and Preparation MethodsLaboratory Synthesis RoutesThe classical synthesis of pyrimidine involves the reduction of 2,4,6-trichloropyrimidine, itself prepared from barbituric acid. Modern laboratory preparations employ cyclocondensation reactions of 1,3-dicarbonyl compounds with nitrogen-containing reagents. The most efficient method utilizes formamide as both solvent and reactant at 180 °C under autogenous pressure, yielding pyrimidine in 45–50% conversion after 8 hours. Alternative synthetic routes include the reaction of ethyl acetoacetate with amidines in the presence of phosphorus oxychloride, following Pinner's original methodology. This approach affords 4-methylpyrimidine derivatives that can be dehydrogenated to the parent compound. More recent methodologies involve transition-metal catalyzed cyclizations, such as the palladium-catalyzed reaction of N-vinylamides with carbonitriles, providing pyrimidine in 65% yield with excellent purity. Purification typically employs fractional distillation under reduced pressure (bp 40–42 °C at 15 mmHg) or recrystallization from hydrocarbon solvents at -20 °C. Industrial Production MethodsIndustrial production of pyrimidine employs catalytic vapor-phase reactions of aliphatic precursors over zeolite catalysts. The most economical process involves ammonolysis of pentane-1,5-diol with ammonia over ZSM-5 catalyst at 400 °C, achieving 35% molar yield with 90% selectivity. Annual global production estimates range from 500–1000 metric tons, primarily for pharmaceutical intermediate synthesis. Process optimization focuses on catalyst development to improve yield and reduce energy consumption. Modern plants utilize fluidized-bed reactors with integrated product recovery systems minimizing waste generation. The production cost analysis indicates raw material costs comprising 60% of total expenses, with catalyst regeneration contributing significantly to operational expenditures. Environmental considerations require treatment of gaseous ammonia emissions and recycling of organic solvents, primarily achieved through absorption systems and distillation recovery units. Analytical Methods and CharacterizationIdentification and QuantificationGas chromatography with flame ionization detection provides the most reliable method for pyrimidine quantification, using a polar stationary phase (polyethylene glycol) with temperature programming from 60 °C to 200 °C at 10 °C·min−1. Retention indices relative to n-alkanes measure 825 on DB-WAX columns. The method demonstrates a linear range of 0.1–100 mg·mL−1 with detection limit of 0.05 mg·mL−1 and quantitation limit of 0.15 mg·mL−1. High-performance liquid chromatography with UV detection at 254 nm utilizing C18 reverse-phase columns with aqueous methanol mobile phases (20–80% gradient) achieves separation from common impurities. Mass spectrometric identification employs electron impact ionization with characteristic fragments at m/z 80 (M+•), 53 (M+•-HCN), and 26 (HCN+•). Infrared spectroscopy provides complementary identification through fingerprint region absorption between 900–700 cm−1. Purity Assessment and Quality ControlPyrimidine purity specification for research applications typically requires ≥99.0% by GC analysis. Common impurities include 4-methylpyrimidine (≤0.3%), pyrazine (≤0.2%), and formamide (≤0.5%). Water content determination by Karl Fischer titration specifies ≤0.1% for anhydrous grades. Residual solvent analysis by headspace GC limits methanol to ≤100 ppm and formamide to ≤50 ppm. Stability testing indicates no significant decomposition when stored under nitrogen atmosphere in amber glass containers at -20 °C for 24 months. Accelerated aging studies at 40 °C and 75% relative humidity demonstrate 98% recovery after 3 months. Quality control protocols include periodic verification of melting point, refractive index, and spectroscopic characteristics against reference standards. Applications and UsesIndustrial and Commercial ApplicationsPyrimidine serves as a fundamental building block in pharmaceutical synthesis, particularly for antiviral and anticancer agents. Approximately 75% of commercial pyrimidine production is directed toward drug intermediate manufacture, including zidovudine and fluorouracil synthesis. The compound functions as a ligand in coordination chemistry, forming stable complexes with transition metals including platinum(II), palladium(II), and ruthenium(II). In materials science, pyrimidine derivatives incorporate into polymeric systems as photoactive components and charge transport materials. The market for pyrimidine-based electronic materials has grown substantially, with annual demand increasing 8–10% over the past decade. Agricultural applications include synthesis of fungicidal and herbicidal compounds, though these represent a smaller market segment. Global consumption patterns show predominance in North America and Europe, accounting for 80% of total demand. Research Applications and Emerging UsesPyrimidine continues to attract significant research attention in supramolecular chemistry as a hydrogen-bonding motif in self-assembling systems. Its coordination behavior with lanthanide ions enables development of luminescent materials with quantum yields up to 0.45. Recent investigations explore pyrimidine as a precursor for carbon nitride materials through polymerization and pyrolysis sequences. Electrochemical studies focus on pyrimidine-containing redox-active systems for energy storage applications. Catalytic research utilizes pyrimidine derivatives as ligands in asymmetric synthesis, achieving enantiomeric excesses above 95% in hydrogenation reactions. Patent analysis indicates growing intellectual property activity in pyrimidine-based metal-organic frameworks for gas storage and separation applications. Historical Development and DiscoveryThe history of pyrimidine chemistry begins with the isolation of uric acid derivatives in the early 19th century, though the parent compound remained unknown. In 1879, Grimaux reported the first laboratory synthesis of a pyrimidine derivative, barbituric acid, from urea and malonic acid. Systematic investigation commenced with Adolf Pinner's work in 1884, who condensed ethyl acetoacetate with amidines to produce various substituted pyrimidines. Pinner introduced the name "pyrimidin" in 1885, drawing analogy to pyridine while recognizing the compound's relationship to pyridine and pyrazine. The unsubstituted parent compound was first isolated in 1900 by Gabriel and Colman through reduction of 2,4,6-trichloropyrimidine with zinc dust. Structural elucidation progressed through early X-ray crystallography studies in the 1930s, confirming the molecular geometry and aromatic character. Theoretical understanding advanced significantly with the application of molecular orbital theory in the 1950s, providing explanation for its unique electronic properties and reactivity patterns. ConclusionPyrimidine represents a fundamental heterocyclic system with distinctive electronic properties arising from its π-deficient aromatic character. The compound exhibits physical properties consistent with its molecular structure, including moderate polarity, limited basicity, and characteristic spectroscopic signatures. Its chemical behavior demonstrates preferential reactivity at the electron-deficient positions through nucleophilic substitution and at the 5-position through electrophilic attack. Synthetic methodologies have evolved from classical condensation reactions to modern catalytic processes, though industrial production remains challenging due to modest yields. Analytical characterization relies heavily on chromatographic and spectroscopic techniques capable of distinguishing pyrimidine from its isomers and common impurities. Applications continue to expand beyond traditional pharmaceutical uses into materials science and coordination chemistry. Future research directions likely focus on developing more efficient synthetic routes, exploring advanced materials applications, and further elucidating reaction mechanisms through computational and experimental approaches. | |||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||
Chemical Compound Properties DatabaseThis database contains physical properties and alternative names for thousands of chemical compounds. In chemical formula you may use:
The database includes melting points, boiling points, densities, and alternative names collected from various chemical sources. What are compound properties?Chemical compound properties include physical characteristics such as melting point, boiling point, and density, which are important for chemical identification and applications. Alternative names help identify the same compound when referenced by different naming conventions.How to use this tool?Enter a chemical formula (like H2O) or compound name (like water) to look up available properties and alternative names. The tool will search through the database and display any available physical properties and known alternative names for the compound. | |||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||
