Properties of C36H70O4Zn (Zinc stearate):
Alternative Nameszinc distearate zinc octadecanoate Elemental composition of C36H70O4Zn
Related compounds
Zinc Stearate (C36H70O4Zn): Chemical CompoundScientific Review Article | Chemistry Reference Series
AbstractZinc stearate, with the empirical formula C36H70O4Zn, represents a significant class of metal carboxylate compounds known as zinc soaps. This hydrophobic white powder exhibits a characteristic melting range between 120°C and 130°C and decomposes upon further heating. The compound demonstrates exceptional water-repellent properties and functions as a powerful mold release agent in industrial applications. Its complex molecular structure features a Zn4O6+ core coordinated with six stearate ligands, distinguishing it from simple dicarboxylate salts. Zinc stearate finds extensive utilization in plastics, rubber processing, and cosmetics due to its lubricating and thickening properties. The compound's insolubility in polar solvents contrasts with its solubility in aromatic and chlorinated hydrocarbons when heated, reflecting its predominantly nonpolar character. IntroductionZinc stearate belongs to the organometallic compound class known as metal soaps, specifically zinc carboxylates. These materials represent coordination compounds where zinc cations form complexes with long-chain fatty acid anions. The compound holds significant industrial importance as a versatile additive with multiple functions across various manufacturing sectors. Its classification as a zinc soap places it within a broader category of metal carboxylates that have been utilized commercially since the early 20th century. The systematic IUPAC name for zinc stearate is zinc octadecanoate, reflecting its derivation from stearic acid (octadecanoic acid). The compound's unique combination of hydrophobic character, thermal stability, and lubricating properties has established its role as an essential industrial chemical with annual global production exceeding several hundred thousand metric tons. Molecular Structure and BondingMolecular Geometry and Electronic StructureZinc stearate exhibits a complex molecular structure that differs fundamentally from simple ionic carboxylate salts. The compound adopts the structural formula Zn4O(O2CC17H35)6, featuring a tetrahedral Zn4O6+ core where zinc atoms occupy the vertices and an oxygen atom resides at the center. Each zinc atom achieves tetrahedral coordination geometry through bonding with the central oxygen and three carboxylate oxygen atoms from stearate ligands. The stearate anions function as bridging ligands that connect zinc atoms across the tetrahedral edges, creating an extensive three-dimensional coordination network. The electronic structure involves zinc in the +2 oxidation state with electron configuration [Ar]3d10. The zinc ions exhibit sp3 hybridization consistent with their tetrahedral coordination environment. Bond angles at zinc centers approximate the ideal tetrahedral angle of 109.5°, though slight distortions occur due to steric constraints from the bulky stearate chains. The carboxylate groups display symmetric bonding with Zn-O bond lengths typically measuring 1.93-1.97 Å, characteristic of covalent coordination bonds rather than ionic interactions. This bonding arrangement creates a highly symmetric molecular architecture that influences the compound's physical properties and reactivity. Chemical Bonding and Intermolecular ForcesThe chemical bonding in zinc stearate involves primarily covalent coordination bonds between zinc atoms and carboxylate oxygen atoms, with additional van der Waals interactions between the hydrocarbon chains. The carboxylate groups function as bidentate ligands that bridge between zinc centers, creating a robust coordination network. The Zn-O bonds demonstrate significant covalent character with bond dissociation energies estimated at 250-300 kJ/mol based on comparative analysis with other zinc carboxylates. Intermolecular forces dominate the bulk material properties, particularly the extensive van der Waals interactions between the long alkyl chains of stearate ligands. These London dispersion forces, with interaction energies of approximately 5-10 kJ/mol per methylene unit, contribute significantly to the compound's thermal stability and mechanical properties. The molecular structure lacks significant dipole moments due to the symmetric arrangement of carboxylate groups around the zinc centers. The hydrocarbon chains adopt predominantly extended conformations with occasional gauche defects, creating a layered structure with alternating inorganic and organic regions. This structural organization results in a material that is strongly hydrophobic and exhibits excellent lubricating properties. Physical PropertiesPhase Behavior and Thermodynamic PropertiesZinc stearate presents as a soft, white powder with a slight characteristic odor. The material exhibits a density of 1.095 g/cm³ at 25°C and demonstrates a melting range between 120°C and 130°C rather than a sharp melting point, reflecting its polymeric nature. The compound does not boil but decomposes upon heating above its melting range, with decomposition becoming significant above 200°C. The heat of fusion measures approximately 80-100 kJ/mol, while the specific heat capacity ranges from 1.2-1.5 J/g·K in the solid state. The crystal structure belongs to the monoclinic system with space group P21/c and unit cell parameters a = 5.42 Å, b = 7.85 Å, c = 37.2 Å, and β = 92.5°. The structure displays layered organization with alternating zinc-oxygen layers and hydrocarbon regions. The refractive index measures 1.495 at 589 nm and 20°C. The compound exhibits polymorphism with at least two crystalline forms that undergo reversible phase transitions around 80-90°C. Thermal expansion occurs anisotropically with greater expansion perpendicular to the layers than within the layer planes. Spectroscopic CharacteristicsInfrared spectroscopy reveals characteristic vibrational modes that provide insight into the molecular structure. The asymmetric carboxylate stretching vibration appears at 1535-1545 cm-1, while the symmetric stretch occurs at 1390-1400 cm-1. The separation between these bands (Δν ≈ 140 cm-1) indicates bridging bidentate coordination of carboxylate groups to zinc centers. Methyl and methylene C-H stretching vibrations appear at 2918 cm-1 and 2850 cm-1, respectively, with bending vibrations at 1465 cm-1 and 720 cm-1. Solid-state 13C NMR spectroscopy shows signals at 183 ppm for carboxylate carbon atoms, 34 ppm for the α-methylene carbon, 32 ppm for internal methylene carbons, and 14 ppm for the terminal methyl group. The 1H NMR spectrum in deuterated chloroform exhibits a broad signal at 0.88 ppm for terminal methyl protons, a strong signal at 1.26 ppm for methylene protons, and a signal at 2.25 ppm for α-methylene protons adjacent to the carboxylate group. Mass spectrometric analysis reveals fragmentation patterns characteristic of carboxylate compounds with prominent peaks corresponding to the stearate anion (m/z = 283) and various hydrocarbon fragments. Chemical Properties and ReactivityReaction Mechanisms and KineticsZinc stearate demonstrates moderate chemical reactivity typical of metal carboxylates. The compound undergoes hydrolysis in acidic conditions with a rate constant of approximately 2.3 × 10-3 L·mol-1·s-1 at 25°C, producing stearic acid and zinc salts. Thermal decomposition follows first-order kinetics with an activation energy of 120-140 kJ/mol, resulting in the formation of zinc oxide, hydrocarbons, and ketones through complex radical mechanisms. The decomposition onset temperature occurs at approximately 200°C under nitrogen atmosphere. The compound functions as a Lewis acid catalyst due to the electrophilic zinc centers, particularly in esterification and transesterification reactions. Catalytic activity manifests with turnover frequencies of 5-10 h-1 for typical esterification reactions at 80°C. Zinc stearate exhibits stability toward oxidation under ambient conditions but undergoes autoxidation at elevated temperatures through radical chain mechanisms involving the hydrocarbon chains. The oxidation induction time measures 45-60 minutes at 150°C under oxygen atmosphere. Acid-Base and Redox PropertiesZinc stearate displays amphoteric character due to the zinc centers, which can function as both Lewis acids and weak bases. The compound demonstrates limited solubility in aqueous systems across the pH range, with minimal dissolution below pH 5.0 and above pH 8.0. The zinc centers have an effective pKa of approximately 7.5 for hydrolysis, while the carboxylate groups exhibit pKa values around 4.5 when protonated. Redox properties involve primarily the zinc centers, which have a standard reduction potential of -0.76 V for the Zn2+/Zn couple. The compound shows stability toward common oxidizing agents under standard conditions but undergoes reduction with strong reducing agents at elevated temperatures. Electrochemical measurements indicate negligible electronic conductivity, with resistivity values exceeding 1012 Ω·cm. The compound maintains stability in reducing environments but decomposes in strongly oxidizing conditions above 150°C. Synthesis and Preparation MethodsLaboratory Synthesis RoutesLaboratory synthesis of zinc stearate typically employs precipitation methods from aqueous solutions. The most common approach involves the reaction of sodium stearate with zinc sulfate or zinc chloride in aqueous medium. The procedure requires dissolution of 0.2 mol sodium stearate in 500 mL deionized water at 70°C, followed by dropwise addition of 0.1 mol zinc sulfate heptahydrate in 200 mL water. The reaction proceeds according to the equation: 2C17H35COO-Na+ + Zn2+ → Zn(O2CC17H35)2 + 2Na+. The product precipitates as a white solid and requires thorough washing with hot water to remove electrolyte impurities. Yields typically reach 85-90% with purity exceeding 98% after recrystallization from appropriate solvents. Alternative laboratory methods include direct reaction of stearic acid with zinc oxide or zinc carbonate in nonaqueous solvents. This approach involves refluxing 0.2 mol stearic acid with 0.1 mol zinc oxide in 300 mL toluene for 4-6 hours, with water removal using a Dean-Stark apparatus. The reaction achieves completion when water generation ceases, typically yielding 92-95% pure product. Purification methods include Soxhlet extraction with petroleum ether to remove unreacted stearic acid and recrystallization from hot benzene. Industrial Production MethodsIndustrial production of zinc stearate utilizes both precipitation and fusion processes on commercial scales. The precipitation process dominates large-scale production, employing continuous reaction systems with automated control. Typical industrial reactors process 5000-10000 kg batches using zinc oxide and stearic acid as starting materials. The reaction occurs in heated water at 80-90°C with vigorous agitation over 2-3 hours, followed by filtration, washing, and drying operations. Industrial yields exceed 95% with production capacities reaching 50,000 metric tons annually worldwide. The fusion process involves direct heating of zinc oxide with stearic acid at 120-140°C without solvent. This method produces material with different physical characteristics, particularly regarding particle size and distribution. Economic considerations favor the precipitation process due to lower energy requirements and better product consistency. Major manufacturers employ quality control systems that monitor parameters including moisture content (max 1.5%), free fatty acid content (max 1.0%), and zinc content (13.0-15.0%). Environmental management strategies focus on water recycling and byproduct recovery, with modern facilities achieving near-zero liquid discharge. Analytical Methods and CharacterizationIdentification and QuantificationAnalytical identification of zinc stearate employs multiple complementary techniques. Fourier transform infrared spectroscopy provides definitive identification through characteristic carboxylate stretching vibrations at 1535-1545 cm-1 and 1390-1400 cm-1. X-ray diffraction analysis confirms the crystalline structure with characteristic peaks at d-spacings of 4.12 Å, 3.85 Å, and 2.56 Å. Thermogravimetric analysis shows a distinctive decomposition pattern with major weight loss between 200°C and 400°C. Quantitative analysis typically utilizes complexometric titration for zinc content determination. The procedure involves sample digestion with concentrated nitric acid, followed by pH adjustment to 5.0-6.0 and titration with ethylenediaminetetraacetic acid (EDTA) using xylenol orange as indicator. Accuracy reaches ±0.5% for zinc content determination. Gas chromatography methods quantify stearic acid content after acid hydrolysis, with detection limits of 0.1% for free fatty acids. High-performance liquid chromatography methods achieve separation and quantification of various fatty acid components with precision of ±1.5%. Purity Assessment and Quality ControlPurity assessment involves determination of multiple parameters according to industrial specifications. Moisture content determination employs Karl Fischer titration with acceptance criteria typically below 1.5%. Free fatty acid content measures through titration with 0.1 M sodium hydroxide solution, with specifications generally requiring less than 1.0% as stearic acid. Heavy metal contamination analysis utilizes atomic absorption spectroscopy, with lead and cadmium limits typically below 10 ppm and 5 ppm, respectively. Particle size distribution analysis employs laser diffraction methods, with commercial grades typically exhibiting median particle sizes of 5-15 μm. Loss on drying at 105°C should not exceed 2.0% for most industrial grades. Arsenic content, determined by hydride generation atomic absorption spectroscopy, must remain below 3 ppm for applications requiring high purity. Stability testing under accelerated conditions (40°C, 75% relative humidity) demonstrates no significant degradation over 6 months, supporting typical shelf life of 2 years under proper storage conditions. Applications and UsesIndustrial and Commercial ApplicationsZinc stearate serves numerous industrial applications primarily exploiting its hydrophobic and lubricating properties. In the plastics industry, it functions as an excellent mold release agent and lubricant during processing of polyolefins, polyvinyl chloride, and polystyrene. Typical addition levels range from 0.5-2.0% by weight, significantly reducing friction during extrusion and injection molding operations. The rubber industry utilizes zinc stearate as an activator for sulfur vulcanization, where it enhances crosslinking efficiency and reduces curing times. Addition levels in rubber compounds typically range from 1-3 phr (parts per hundred rubber). The compound finds extensive use in cosmetic formulations as a lubricant and thickening agent, particularly in powders and makeup products. Its ability to impart water-repellent characteristics makes it valuable in waterproofing applications for textiles and paper products. In pharmaceutical manufacturing, zinc stearate functions as a lubricant in tablet formulations, preventing sticking during compression and improving flow properties. The powder metallurgy industry employs zinc stearate as a lubricant in metal powder compaction, typically at 0.5-1.0% addition levels, reducing die wear and improving green strength. Research Applications and Emerging UsesResearch applications of zinc stearate continue to expand into new areas of materials science. Recent investigations explore its use as a precursor for the synthesis of zinc oxide nanoparticles through controlled thermal decomposition. The compound shows promise as a phase transfer catalyst in organic synthesis, particularly for reactions involving ionic species in nonpolar media. Emerging applications include its utilization as a hydrophobic coating agent for nanoparticles and as a component in self-cleaning surfaces. Advanced materials research investigates zinc stearate as a templating agent for mesoporous materials and as a modifier for polymer nanocomposites. Studies examine its potential as a corrosion inhibitor for metal surfaces, particularly in protective coatings. The compound's ability to form ordered molecular films makes it suitable for Langmuir-Blodgett film applications and molecular electronics. Patent activity remains active in areas including biodegradable polymers, advanced lubricants, and functional coatings, indicating continued innovation in application development. Historical Development and DiscoveryThe development of zinc stearate parallels the broader history of metal soap technology, which emerged during the late 19th century. Early investigations of metal carboxylates began in the 1850s with the work of French chemists studying the reactions of fatty acids with various metal salts. The systematic study of zinc soaps commenced in the 1880s as industrial applications for these materials began to emerge. The compound's utility as a lubricant and release agent became apparent during the early development of the plastics industry in the 1920s and 1930s. Structural characterization advanced significantly during the mid-20th century with the application of X-ray crystallography to metal carboxylates. The determination of the Zn4O(O2CR)6 structure in the 1950s represented a major advancement in understanding the molecular architecture of zinc carboxylates. Industrial production methods evolved from batch processes to continuous manufacturing during the 1960s, improving product consistency and economic efficiency. The late 20th century witnessed expansion into new application areas, particularly in cosmetics and pharmaceuticals, driven by improved purification methods and quality control standards. ConclusionZinc stearate represents a chemically complex and industrially significant compound with unique structural features and diverse applications. Its molecular architecture, featuring a Zn4O6+ core coordinated with stearate ligands, distinguishes it from simple ionic carboxylates and contributes to its distinctive properties. The compound's excellent lubricating characteristics, water repellency, and thermal stability make it invaluable across numerous industrial sectors including plastics, rubber, cosmetics, and pharmaceuticals. Ongoing research continues to reveal new potential applications in advanced materials and nanotechnology, particularly as a precursor for functional materials and as a component in sophisticated formulations. Future developments will likely focus on enhancing sustainability through improved production methods and expanding into emerging technological areas requiring specialized functional additives. | ||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||
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. | ||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||
