Properties of ZnO (Zinc oxide):
Alternative NamesZinc white, calamine, philosopher's wool, Chinese white, flowers of zinc, zinca Elemental composition of ZnO
Related compounds
Sample reactions for ZnO
Zinc Oxide (ZnO): Chemical CompoundScientific Review Article | Chemistry Reference Series
AbstractZinc oxide (ZnO) is an inorganic compound with the chemical formula ZnO, characterized as a white, insoluble powder with a molecular mass of 81.406 g·mol−1. The compound crystallizes primarily in the hexagonal wurtzite structure (space group P63mc) with lattice parameters a = 3.2495 Å and c = 5.2069 Å. ZnO exhibits amphoteric behavior, dissolving in both acids and bases, and demonstrates a wide direct band gap of 3.2–3.4 eV at room temperature. This semiconductor material manifests strong room-temperature luminescence, high electron mobility (~180 cm2·V−1·s−1), and pronounced piezoelectric properties. Industrial production exceeds 105 tons annually through French (indirect), American (direct), and wet chemical processes. Applications span rubber vulcanization, ceramic glazes, ultraviolet radiation protection, varistors, and gas sensors. The compound's unique combination of electrical, optical, and mechanical properties establishes its significance in materials science and industrial chemistry. IntroductionZinc oxide represents a fundamental inorganic compound occupying a critical position in both industrial chemistry and materials science. Classified as a II-VI semiconductor, ZnO demonstrates exceptional versatility across applications ranging from rubber manufacturing to advanced optoelectronics. The compound occurs naturally as the mineral zincite, though most commercial material undergoes synthetic production. Historical records indicate early medicinal use in Indian and Greek civilizations, with systematic production developing in Europe during the 18th century. Modern understanding recognizes ZnO's unique combination of wide band gap semiconductor behavior, piezoelectricity, and photocatalytic activity. These properties derive from its electronic structure and chemical bonding characteristics, which have been extensively characterized through X-ray diffraction, spectroscopic methods, and theoretical calculations. The compound's stability under various environmental conditions and relative non-toxicity further contribute to its widespread technological utilization. Molecular Structure and BondingMolecular Geometry and Electronic StructureZinc oxide crystallizes predominantly in the hexagonal wurtzite structure (B4 type) belonging to space group P63mc (C6v4) with two formula units per unit cell. The structure features tetrahedral coordination around both zinc (Zn2+) and oxygen (O2−) ions, with bond angles of 109.5° characteristic of sp3 hybridization. The wurtzite structure lacks inversion symmetry, resulting in non-centrosymmetric properties including piezoelectricity and pyroelectricity. Experimental lattice parameters measure a = 3.2495 Å and c = 5.2069 Å at room temperature, yielding a c/a ratio of approximately 1.602 which slightly deviates from the ideal hexagonal value of 1.633. The cubic zincblende structure (B3 type) represents a metastable polymorph obtainable through epitaxial growth on substrates with cubic symmetry. This structure belongs to space group F43m (Td2) and also exhibits tetrahedral coordination. At pressures exceeding approximately 10 GPa, ZnO undergoes phase transition to the rocksalt (NaCl-type) structure with octahedral coordination. Electronic structure calculations based on density functional theory reveal a direct band gap at the Γ point of the Brillouin zone. The valence band maximum derives primarily from oxygen 2p orbitals, while the conduction band minimum consists mainly of zinc 4s orbitals. The compound's ionicity, quantified by the Phillips ionicity scale, measures approximately 0.616, intermediate between covalent and ionic bonding extremes. Chemical Bonding and Intermolecular ForcesChemical bonding in ZnO exhibits predominantly ionic character with partial covalent contributions. The electronegativity difference between zinc (1.65) and oxygen (3.44) according to the Pauling scale suggests approximately 63% ionic character. Bond lengths measure 1.977 Å in the wurtzite structure, with Madelung constants of 1.641 for wurtzite and 1.638 for zincblende polymorphs. The binding energy calculated from Born-Haber cycles approximates 17.3 eV per formula unit. Intermolecular forces in solid ZnO include strong ionic interactions between Zn2+ and O2− ions, with Coulombic attraction dominating the cohesive energy. Van der Waals interactions contribute minimally to crystal stability due to the compound's high lattice energy of approximately -3961 kJ·mol−1. The polar nature of Zn-O bonds creates electrically charged zinc and oxygen planes perpendicular to the c-axis, with surface dipoles measuring approximately 0.8 eV per unit cell along the [0001] direction. The compound's polarity influences surface properties and defect chemistry, particularly regarding oxygen vacancies and zinc interstitials which commonly act as n-type dopants. Nonstoichiometric compositions Zn1+xO occur at elevated temperatures, with x reaching 7×10−5 at 800°C under atmospheric oxygen pressure. Physical PropertiesPhase Behavior and Thermodynamic PropertiesZinc oxide appears as a white microcrystalline powder with density measuring 5.606 g·cm−3 at 20°C. The compound exhibits thermochromic behavior, changing from white to yellow upon heating above 300°C due to formation of nonstoichiometric Zn1+xO and reverting to white upon cooling. Melting occurs with decomposition at 1974°C under standard atmospheric pressure, while sublimation becomes significant above 1700°C. The standard enthalpy of formation (ΔfH○298) measures -350.46 ± 0.27 kJ·mol−1, with Gibbs free energy of formation (ΔfG○298) of -320.5 kJ·mol−1. Heat capacity (Cp) follows the Debye model with values of 40.26 J·mol−1·K−1 at 298 K, increasing to 57.51 J·mol−1·K−1 at 1000 K. Entropy (S○298) measures 43.65 ± 0.40 J·mol−1·K−1. Thermal conductivity demonstrates anisotropy between directions parallel and perpendicular to the c-axis, with values of 0.6 W·cm−1·K−1 at room temperature. The linear thermal expansion coefficient measures 4.75×10−6 K−1 along the a-axis and 2.92×10−6 K−1 along the c-axis between 20–800°C. Solubility in water is extremely limited, measuring 0.0004% (4 mg·L−1) at 17.8°C. The compound exhibits amphoteric dissolution behavior, dissolving readily in acids to form zinc salts and in strong bases to form zincate ions [Zn(OH)4]2−. Refractive indices demonstrate birefringence with no = 2.013 and ne = 2.029 at 589 nm. Magnetic susceptibility measures -27.2×10−6 cm3·mol−1, characteristic of diamagnetic behavior. Spectroscopic CharacteristicsInfrared spectroscopy reveals characteristic Zn-O stretching vibrations between 380–580 cm−1, with the most intense absorption at 418 cm−1 corresponding to the E1(TO) mode. Raman active modes include A1 + E1 + 2E2 vibrations, with the E2high mode at 437 cm−1 serving as a fingerprint for wurtzite structure. Photoluminescence spectra exhibit near-band-edge emission at approximately 380 nm (3.26 eV) with a full width at half maximum of 120 meV at room temperature, accompanied by broad visible emission bands centered around 500–600 nm attributed to defect states. UV-Vis spectroscopy demonstrates strong absorption below 400 nm with an absorption coefficient exceeding 105 cm−1 at 3.4 eV. The exciton binding energy measures 60 meV, significantly higher than thermal energy at room temperature (26 meV), facilitating efficient excitonic emission. X-ray photoelectron spectroscopy shows Zn 2p3/2 and 2p1/2 peaks at 1021.8 eV and 1044.9 eV respectively, with O 1s peaks at 530.2 eV (lattice oxygen) and 531.5 eV (surface hydroxyl groups). Chemical Properties and ReactivityReaction Mechanisms and KineticsZinc oxide demonstrates amphoteric reactivity, functioning as both a weak base and weak acid. Reaction with mineral acids proceeds rapidly at room temperature with dissolution rates exceeding 10-4 mol·m-2·s-1 in 1 M HCl. The dissolution mechanism involves protonation of surface oxygen atoms followed by detachment of zinc ions. Reaction with strong bases such as sodium hydroxide forms tetrahydroxozincate ions [Zn(OH)4]2− with equilibrium constants log K = 16.8 at 25°C. Thermal decomposition occurs above 1974°C according to the equilibrium ZnO(s) ⇌ Zn(g) + ½O2(g) with equilibrium constant log Kp = -6.24 at 1200°C. Carbothermal reduction with carbon proceeds at temperatures above 950°C via ZnO(s) + C(s) → Zn(g) + CO(g) with activation energy of 180 kJ·mol-1. Reaction with hydrogen sulfide at 230–430°C produces zinc sulfide: ZnO + H2S → ZnS + H2O, with reaction rates following Langmuir-Hinshelwood kinetics. Surface reactions with fatty acids form corresponding zinc carboxylates, with stearic acid exhibiting reaction half-times of approximately 30 minutes at 120°C. Cement formation with zinc chloride produces basic zinc chlorides approximating Zn(OH)Cl, with setting times ranging from 3–10 minutes depending on concentration. Reaction with phosphoric acid forms hopeite (Zn3(PO4)2·4H2O) through dissolution-precipitation mechanisms. Acid-Base and Redox PropertiesThe point of zero charge for ZnO occurs at pH 8.7–9.5, with surface protonation dominating below this range and deprotonation above. The compound functions as a solid-state Lewis acid, catalyzing various organic transformations including Meerwein-Ponndorf-Verley reductions and Knoevenagel condensations. Redox properties include standard reduction potential E°(ZnO/Zn) = -1.26 V versus standard hydrogen electrode, indicating moderate oxidizing power. Electrochemical behavior in aqueous solutions shows corrosion potential of -0.96 V versus SCE in neutral solutions, with anodic dissolution following Ecorr = -1.05 - 0.06 pH V. The flat band potential measures -0.8 V versus NHE at pH 7, with donor densities typically ranging from 1017 to 1019 cm-3 for undoped material. Photocatalytic activity manifests through band gap excitation (3.2 eV) generating electron-hole pairs with recombination lifetimes of 100–500 ps. Synthesis and Preparation MethodsLaboratory Synthesis RoutesLaboratory synthesis of zinc oxide employs several approaches depending on desired morphology and purity. Precipitation methods involve addition of alkaline solutions to zinc salt solutions, typically using zinc nitrate or zinc acetate with sodium hydroxide or ammonium carbonate. Controlled precipitation at pH 6.5–7.5 yields amorphous zinc hydroxide, which converts to crystalline ZnO upon calcination at 300–600°C. Hydrothermal synthesis utilizes aqueous zinc salts in autoclaves at temperatures of 100–200°C under autogenous pressure, producing well-defined nanocrystals with sizes from 10–100 nm. Vapor-phase transport methods employ chemical vapor deposition with zinc metal vapor and oxygen at temperatures of 800–1000°C, yielding single crystals up to several cubic centimeters. Physical vapor deposition including pulsed laser deposition and sputtering produces thin films with thickness control down to monolayer precision. Sol-gel processing uses zinc alkoxide precursors such as zinc acetate dihydrate in alcoholic solutions, with hydrolysis and condensation reactions forming gel networks that transform to ZnO upon thermal treatment. Industrial Production MethodsIndustrial production predominantly follows three processes: French (indirect), American (direct), and wet chemical methods. The French process vaporizes metallic zinc at temperatures above 907°C in graphite crucibles, with zinc vapor oxidizing exothermically upon contact with air. This method produces high-purity ZnO (99.5–99.9%) with particle sizes of 0.1–5 μm and annual capacity exceeding 60,000 tons worldwide. The American process employs zinc-containing ores or smelter by-products reduced carbothermally with anthracite coal at 1000–1200°C, followed by oxidation of zinc vapor. This method yields material with lower purity (94–98% ZnO) due to impurity carry-over, but remains economically advantageous for certain applications. Wet chemical processes precipitate basic zinc carbonate or hydroxide from purified zinc sulfate solutions, followed by calcination at 800°C. This route produces material with controlled morphology and surface area, particularly for specialized applications. Analytical Methods and CharacterizationIdentification and QuantificationQualitative identification of ZnO utilizes X-ray diffraction with characteristic peaks at 2θ = 31.8° (100), 34.4° (002), and 36.3° (101) for wurtzite structure. Fourier-transform infrared spectroscopy shows strong absorption between 400–500 cm-1 assigned to Zn-O stretching vibrations. Quantitative analysis employs complexometric titration with ethylenediaminetetraacetic acid (EDTA) using Eriochrome Black T indicator, with detection limits of 0.1 mg·L-1. Atomic absorption spectroscopy provides quantitative determination with detection limits of 0.01 mg·L-1 at 213.9 nm wavelength. Inductively coupled plasma optical emission spectrometry achieves detection limits below 0.001 mg·L-1 with multi-element capability. X-ray fluorescence spectroscopy permits non-destructive analysis with precision of 0.1% for major components. Thermogravimetric analysis monitors decomposition above 1800°C with mass loss corresponding to oxygen release. Purity Assessment and Quality ControlIndustrial specifications typically require ZnO content exceeding 99.0% for most applications, with limits on impurities including lead (<100 ppm), cadmium (<10 ppm), and iron (<50 ppm). Pharmaceutical grades according to USP monographs require absence of arsenic and heavier metals beyond specified limits. Surface area measurement via nitrogen adsorption follows BET methodology, with typical values of 3–10 m2·g-1 for French process material and 10–50 m2·g-1 for precipitated grades. Particle size distribution analysis employs laser diffraction or sedimentation methods, with median diameters ranging from 0.2–1.0 μm for standard grades. Electron paramagnetic resonance spectroscopy detects paramagnetic defects including oxygen vacancies and transition metal impurities at concentrations down to 1014 spins·g-1. Electrical characterization measures resistivity from 10-1 to 106 Ω·cm depending on doping and stoichiometry. Applications and UsesIndustrial and Commercial ApplicationsRubber manufacturing consumes 50–60% of global ZnO production, primarily as an activator in sulfur vulcanization. The compound reduces vulcanization time and improves crosslink density through complex formation with stearic acid and reaction with accelerator molecules. Ceramic applications utilize ZnO as a flux in glazes and frits, reducing melting temperatures by 100–200°C and modifying thermal expansion coefficients to prevent crazing. Photocopier papers historically employed ZnO as a photoconductive coating, though this application has largely been superseded by organic photoconductors. Varistor production utilizes sintered ZnO ceramics with additions of bismuth, cobalt, and manganese oxides, exhibiting non-ohmic behavior with switching voltages of 20–300 V·mm-1. Gas sensors exploit changes in electrical conductivity upon adsorption of reducing or oxidizing gases, with detection limits of 1–100 ppm for compounds including hydrogen sulfide and nitrogen dioxide. Research Applications and Emerging UsesTransparent conducting oxides represent an active research area, with aluminum-doped ZnO films achieving resistivities of 2×10-4 Ω·cm and optical transparency exceeding 90% in the visible spectrum. Piezoelectric applications include energy harvesting devices utilizing ZnO nanowires with output voltages up to 10 V under mechanical deformation. Photocatalytic water treatment demonstrates degradation rates of organic pollutants exceeding 90% within 60 minutes under UV irradiation. UV photodetectors based on ZnO nanostructures exhibit responsivities of 100–1000 A·W-1 at 370 nm wavelength with response times below 100 ms. Light-emitting diodes employing ZnO as both n-type layer and active region demonstrate electroluminescence across the ultraviolet to visible spectrum, though efficient p-type doping remains challenging. Lithium-ion battery anodes based on ZnO nanostructures show capacities of 500–1000 mAh·g-1 with cycle stability improvements through composite formation. Historical Development and DiscoveryEarly utilization of zinc compounds dates to ancient civilizations, with Indian medical texts from 500 BC describing pushpanjan (likely ZnO) as eye salve and wound treatment. Greek physician Dioscorides documented medicinal use in the 1st century AD, while Roman metallurgists produced brass through cementation processes involving zinc ores around 200 BC. Systematic production developed in India between the 12th–16th centuries, transferring to China in the 17th century before reaching Europe. The first European zinc smelter established in Bristol (1743) enabled larger-scale production, with Louis-Bernard Guyton de Morveau proposing zinc white as a lead white pigment replacement in 1782. Edme-Jean Leclaire industrialized zinc white paint production in Paris (1845), with manufacturing spreading throughout Europe by 1850. The French process developed by Leclaire remains a major production method today. Semiconductor properties received significant attention beginning in the 1950s, with detailed characterization of optical and electrical properties emerging through the 1970s. Nanoscale ZnO research expanded rapidly from the 1990s onward, driven by advances in synthesis and characterization techniques. ConclusionZinc oxide represents a multifunctional inorganic compound with unique combinations of semiconductor, piezoelectric, and photocatalytic properties. The wurtzite crystal structure with its non-centrosymmetric arrangement enables applications requiring piezoelectric response and polarization effects. Wide band gap semiconductor behavior facilitates ultraviolet optoelectronics and transparent electronics, while amphoteric chemical reactivity supports diverse catalytic and chemical processing applications. Industrial production methods continue to evolve toward higher purity and controlled morphology, particularly for nanostructured materials. Ongoing research addresses challenges including reproducible p-type doping, enhanced photocatalytic efficiency, and integration into hybrid devices. The compound's established industrial importance coupled with emerging applications in energy harvesting, sensing, and electronics ensures continued scientific and technological relevance across multiple disciplines. | ||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||
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. | ||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||
