Fact-checked by Grok 2 weeks ago

Oxide

An oxide is a binary chemical compound consisting of one or more oxygen atoms bonded to one or more atoms of another , typically formed through the reaction of oxygen with metals, nonmetals, or other substances. Oxides are classified into several types based on their chemical behavior and reactivity: acidic oxides, which are typically formed by nonmetals and react with bases to produce salts and water (e.g., CO₂ and SO₂); basic oxides, formed by metals and react with acids to form salts and water (e.g., MgO and CaO); amphoteric oxides, which exhibit both acidic and basic properties and can react with either acids or bases (e.g., Al₂O₃ and ZnO); and neutral oxides, which do not display acidic or basic characteristics (e.g., N₂O and ). The properties of oxides vary widely depending on their composition and structure; metal oxides are often crystalline solids with high melting points, serving as thermal insulators or conductors, while oxides are frequently gases or liquids at and contribute to atmospheric phenomena like . Oxides play a critical role in and daily life, with applications including ceramics and refractories (e.g., alumina in abrasives), catalysts (e.g., in ammonia synthesis), pigments and coatings (e.g., in paints and sunscreens), and electronics (e.g., in semiconductors).

Definition and Classification

Definition

An oxide is a binary chemical compound composed of oxygen and one other chemical element, formed typically through the reaction of that element with oxygen gas. These compounds are central to inorganic chemistry and include examples such as water (H₂O), carbon dioxide (CO₂), and calcium oxide (CaO). Oxides are ubiquitous in nature, with oxygen accounting for about 46% of the Earth's crust by weight, predominantly bound in oxide minerals like silicates and oxides of metals such as iron and aluminum. They form essential components of geological minerals, contribute to atmospheric constituents like carbon dioxide and sulfur dioxide, and play critical roles in biological systems through compounds involved in respiration, photosynthesis, and cellular processes. Oxides are distinguished from related oxygen-containing compounds by the oxidation state of oxygen, which is typically -2 in oxides (as O²⁻), whereas peroxides feature an O-O with oxygen in the -1 state (O₂²⁻), and superoxides contain the O₂⁻ ion with oxygen also at -1 but in a different bonding configuration. The term "oxide" originated in the late , coined by French chemists Louis-Bernard Guyton de Morveau and around 1790, blending "" (oxygen) with the suffix "-ide" from "acide" (acid) to denote compounds of oxygen with other elements.

Types of Oxides

Oxides are classified in various ways, primarily based on the elemental composition and their chemical reactivity. This categorization helps in understanding their roles in chemical reactions and applications.

Classification by Composition

Metal oxides are binary compounds formed between metals and oxygen, typically exhibiting and solid states at ; they often display properties due to the electropositive nature of metals. A representative example is (Na₂O), which reacts vigorously with to form . Non-metal oxides, in contrast, involve non-metals and oxygen, resulting in covalent bonds and frequently gaseous or volatile forms with acidic tendencies. (CO₂) serves as a classic example, dissolving in to produce . Metalloid oxides bridge these categories, showing variable bonding and reactivity; (SiO₂), for instance, forms a network covalent structure and behaves as an , resisting attack but reacting slowly with strong bases like .

Classification by Acid-Base Properties

Oxides are further categorized according to their behavior in acid-base reactions, reflecting their ability to donate or accept protons or electrons. Basic oxides, predominantly from metals in low oxidation states, react with acids to yield salts and water; (CaO), known as quicklime, exemplifies this by neutralizing to form . Acidic oxides, usually from non-metals, combine with water or bases to generate acids; (SO₃) reacts with water to produce , a key industrial process. Amphoteric oxides exhibit dual reactivity, interacting with both acids and bases, which is common in oxides of metals near the boundary; aluminum oxide (Al₂O₃) dissolves in acids to form aluminum salts and in bases to yield aluminates. Neutral oxides lack significant acid-base reactivity and are often from elements like carbon or nitrogen; (CO) does not react with acids or bases under standard conditions.

Mixed and Complex Oxides

Mixed oxides incorporate multiple metal cations or behave as combinations of simpler oxides, leading to unique properties like enhanced stability or catalytic activity. Magnetite (Fe₃O₄) is a well-known mixed oxide, effectively a composite of iron(II) oxide (FeO) and iron(III) oxide (Fe₂O₃), used in magnetic applications. Complex oxides feature intricate structures with multiple cations arranged in specific lattices; spinels, with the formula AB₂O₄ where A and B are divalent and trivalent metals respectively, represent this class, as seen in magnesium aluminate (MgAl₂O₄), which adopts a cubic close-packed oxygen framework with cations in tetrahedral and octahedral sites. These structures often arise in naturally occurring minerals and synthetic materials for electronics and ceramics.

Rare Types: Suboxides and Mixed-Valence Oxides

Suboxides are uncommon oxides where the oxygen content is deficient relative to the stoichiometric ratio, often resulting in non-stoichiometric compositions and metallic properties; (nominally FeO but actually Fe_{1-x}O) illustrates this, featuring iron vacancies that impart semiconducting behavior. Mixed-valence oxides contain the same metal ion in more than one within the , enabling delocalization and interesting electronic properties; lead tetroxide (Pb₃O₄), or red lead, contains both Pb²⁺ and Pb⁴⁺, functioning as a of PbO and PbO₂ and used historically in paints and batteries. These types are less prevalent but significant in like catalysts and devices.

Stoichiometry and Nomenclature

Stoichiometric Variations

Oxides frequently exhibit ideal stoichiometry, where the ratio of metal atoms to oxygen atoms follows simple integer proportions determined by the oxidation state of the metal. For instance, magnesium oxide adopts a 1:1 ratio in the formula MgO, reflecting the +2 oxidation state of magnesium. Similarly, iron(III) oxide has a 2:3 ratio in Fe2O3, corresponding to the +3 oxidation state of iron. For metals in the +3 oxidation state, such as aluminum, the typical stoichiometry is M2O3, as seen in Al2O3. In contrast, non-stoichiometric oxides deviate from these integer ratios, often due to intrinsic defect structures such as cation vacancies or interstitial oxygen atoms. A prominent example is , formulated as Fe1-xO, where x typically ranges from 0.05 to 0.16, arising primarily from iron vacancy defects that maintain charge balance through higher-valence iron ions. This non-stoichiometry is particularly common in oxides, where structural imperfections allow for compositional flexibility without altering the overall crystal significantly. The degree of stoichiometry in oxides is influenced by external conditions and intrinsic properties of the elements involved. Temperature plays a key role, as higher temperatures can promote defect formation or annealing, altering the oxygen-to-metal ratio; for example, in nickel oxide, heating reduces defects to approach stoichiometry. Oxygen partial pressure affects the equilibrium composition, with higher pressures favoring oxygen incorporation and lower pressures leading to metal excess, as observed in wüstite where defect concentration varies directly with oxygen activity. Additionally, the variable oxidation states of transition metals enable multiple stable stoichiometries, such as the coexistence of FeO-like and Fe2O3-like phases under different conditions. Non-stoichiometry can be generally illustrated by the formula MxOy, where the ratio x/y is not an integer, reflecting the presence of defects that disrupt perfect periodicity.

Naming Conventions

The International Union of Pure and Applied Chemistry (IUPAC) recommends systematic nomenclature for oxides, particularly binary compounds, where the name consists of the electropositive element followed by "oxide," with multiplicative prefixes such as "di-" or "tri-" used to denote stoichiometry when necessary. For metals exhibiting variable oxidation states, the oxidation number is indicated in Roman numerals in parentheses after the element name, ensuring clarity; for example, FeO is named iron(II) oxide, while Fe₂O₃ is iron(III) oxide. This stock system is preferred in modern chemical literature to avoid ambiguity. Traditional names, retained for some common oxides, employ Latin-derived roots with suffixes "-ous" for the lower and "-ic" for the higher; for instance, Cu₂O is cuprous oxide (copper(I)), and CuO is cupric oxide ((II)). Although these are accepted where unambiguous, IUPAC encourages the systematic approach for consistency across compounds. Special cases deviate from standard binary naming to reflect distinct oxygen bonding or . Peroxides, containing the O₂²⁻ , are named with the "peroxide" suffix, as in for H₂O₂ or for Na₂O₂. Superoxides, containing the O₂⁻ , are named with the "superoxide" suffix, as in for KO₂. Ozonides, featuring the O₃⁻ , use the "ozonide" ending, exemplified by potassium ozonide (KO₃). Suboxides, with oxygen content below typical , incorporate a "suboxide" descriptor, such as for C₃O₂. Oxide formulas are typically empirical for ionic compounds (e.g., CuO for cupric oxide) but molecular for covalent ones (e.g., CO₂ for ), reflecting the simplest whole-number ratio versus the actual atomic composition. In complex oxides involving multiple metals or polyatomic components, names combine cation elements with "oxide" and prefixes for , such as calcium titanium oxide for CaTiO₃ (or more precisely, calcium titanium trioxide), where polyatomic oxo-groups are implied in additive if coordination is specified.

Formation and Preparation

Metal Oxides

Metal oxides are commonly synthesized through direct oxidation, where metals are burned in an oxygen atmosphere to form the corresponding oxides. This method is particularly effective for alkali and alkaline earth metals due to their high reactivity. For instance, lithium metal reacts vigorously with oxygen to produce lithium oxide, as described by the balanced equation: $4\text{Li} + \text{O}_2 \rightarrow 2\text{Li}_2\text{O} This reaction occurs upon ignition and results in a white solid product, highlighting the exothermic nature of direct oxidation processes for less stable metals. Thermal decomposition represents another key high-temperature approach for preparing metal oxides, often applied to precursors like carbonates or hydroxides. In this process, the precursor is heated to drive off volatile components, leaving behind the oxide. A classic example is the of to produce quicklime (), which decomposes at around 900°C according to: \text{CaCO}_3 \rightarrow \text{CaO} + \text{CO}_2 This endothermic reaction is widely used in lime production and requires controlled heating to achieve complete decomposition without the product excessively. Similar decompositions apply to metal hydroxides, such as yielding . On an industrial scale, is integral to ceramic production, as seen in the for extracting alumina from . ore is digested in solution to form , which is then precipitated as aluminum and calcined at temperatures exceeding 1000°C to yield pure alumina (Al₂O₃). This high-temperature step removes water and impurities, producing a fine white powder essential for aluminum and refractories. Specific high-temperature methods are employed for metal oxides, exemplified by the production of zirconia (ZrO₂) from zircon sand (). dissociation or processes decompose zircon at temperatures above 1800°C, liberating silica and forming zirconia via: \text{ZrSiO}_4 \rightarrow \text{ZrO}_2 + \text{SiO}_2 This yields a stable, high-melting-point oxide used in ceramics and thermal barrier coatings, underscoring the emphasis on elevated temperatures to overcome the stability of precursors.

Non-Metal Oxides

Non-metal oxides are primarily prepared through direct reactions between non-metals and oxygen, often involving processes that yield gaseous or volatile products. These methods contrast with those for metal oxides by emphasizing high-temperature combination s that produce molecular rather than extended solids. Combustion of non-metals in oxygen is a common route, where the element is ignited in air or pure oxygen to form the oxide, with the completeness of the depending on oxygen availability and temperature. A representative example is the of carbon, which produces under complete oxidation conditions according to the : \mathrm{C + O_2 \rightarrow CO_2} Incomplete , such as in limited oxygen environments, yields instead: \mathrm{2C + O_2 \rightarrow 2CO} These reactions release significant heat and are fundamental to understanding fuel oxidation, though they are exothermic and self-sustaining once initiated. Similarly, reacts with oxygen upon burning to form , typically requiring elevated temperatures to ensure efficient combination: \mathrm{S + O_2 \rightarrow SO_2} This reaction occurs readily when is ignited in air, producing a characteristic and dense fumes of SO₂ gas, which is a key step in sulfur-based industrial chemistry. In laboratory settings, non-metal oxides like (N₂O₅) are prepared via dehydration of the corresponding acids using dehydrating agents such as (P₄O₁₀). For instance, concentrated is dehydrated at low temperatures: $4 \mathrm{HNO_3} + \mathrm{P_4O_{10}} \rightarrow 2 \mathrm{N_2O_5} + 4 \mathrm{HPO_3} This method yields N₂O₅ as a white, unstable solid that decomposes readily, highlighting the volatility and reactivity of many non-metal oxides. The process must be controlled to avoid explosive decomposition, and it remains a standard preparative technique for anhydrides of oxyacids. On an industrial scale, the production of sulfur trioxide (SO₃) exemplifies catalytic methods in the contact process for sulfuric acid manufacturing. Sulfur dioxide, obtained from burning sulfur or roasting sulfide ores, is oxidized with oxygen over a vanadium pentoxide (V₂O₅) catalyst at approximately 400–450°C and 1–2 atm pressure: \mathrm{2SO_2 + O_2 \xrightarrow{V_2O_5} 2SO_3} The V₂O₅ catalyst, often supported on silica and promoted with alkali metal salts, facilitates the reaction by lowering the activation energy through a redox mechanism involving vanadium oxidation states, achieving conversions exceeding 99% in multiple stages to optimize equilibrium yields. This process is highly efficient and accounts for the majority of global sulfuric acid production, underscoring the role of catalysis in scaling non-metal oxide synthesis.

Structure and Bonding

Molecular Structures

Molecular oxides, particularly those formed by non-metals, exhibit covalent bonding where atoms share electrons to achieve stable electron configurations. In , the molecule features a between carbon and oxygen, consisting of one sigma and two s, resulting in a linear ./09:_Covalent_Bonding/9.08:_Coordinate_Covalent_Bond) Similarly, displays two double bonds, each comprising one sigma and one , with the central carbon atom undergoing sp hybridization to form the linear \ce{O=C=O} structure. This hybridization allows the carbon atom's 2s and one 2p orbital to mix, creating two sp hybrid orbitals that align linearly for optimal bonding. The Valence Shell Electron Pair Repulsion ( effectively predicts the geometries of these molecular oxides by considering the repulsion between electron pairs around the central atom. For (SO₂), the central atom has three electron domains—two bonding pairs to oxygen and one —leading to a bent molecular with a bond angle of approximately 119 degrees./09:_Molecular_Geometry_and_Covalent_Bonding_Models/9.02:VSEPR-_Molecular_Geometry) In contrast, (SO₃) possesses three bonding domains and no s on the , resulting in a trigonal planar with bond angles of 120 degrees, facilitated by sp² hybridization of the central atom. Other notable examples include (N₂O), a linear with the \ce{N≡N-O} or forms, commonly used as an gas due to its pharmacological properties. (Cl₂O) adopts a bent configuration, with the oxygen atom as the central atom bonded to two atoms and possessing two lone pairs, akin to water's VSEPR classification (AX₂E₂). Spectroscopic techniques such as and provide evidence for these bond strengths and molecular geometries by identifying characteristic vibrational modes. For instance, the asymmetric stretching mode of CO₂ appears as a strong absorption at 2349 cm⁻¹, reflecting the high bond strength of the C=O double bonds, while the symmetric stretch is Raman active but inactive due to . In CO, the C≡O produces a prominent band at 2143 cm⁻¹, indicating its exceptional strength compared to single or double bonds. For SO₂, bands around 1362 cm⁻¹ (asymmetric stretch) and 1151 cm⁻¹ (symmetric stretch) confirm the bent structure and varying bond orders influenced by . These vibrational frequencies correlate directly with bond strengths, as higher wavenumbers signify stronger bonds due to greater force constants./4:_Infrared_Spectroscopy/4.1:_Introduction_to_Infrared_Spectroscopy)

Crystalline Structures

Many solid-state oxides, especially metal oxides, adopt ionic crystalline lattices where metal cations and oxygen anions are arranged in a regular, three-dimensional array to maximize electrostatic interactions and maintain charge balance. A prominent example is the rock salt (NaCl-type) structure, observed in monovalent oxides like those of alkaline earth metals (e.g., MgO and CaO), featuring a face-centered cubic arrangement of alternating cation and anion layers, with each ion coordinated to six nearest neighbors of the opposite type./06%3A_Structures_and_Energetics_of_Metallic_and_Ionic_solids/6.11%3A_Ionic_Lattices/6.11A%3A_Structure_-Rock_Salt(NaCl)) Sodium oxide (Na₂O), due to its 2:1 cation-to-anion stoichiometry, instead forms an antifluorite lattice, with O²⁻ ions in a face-centered cubic array and Na⁺ ions occupying all tetrahedral interstitial sites, resulting in cubic coordination for oxygen and tetrahedral for sodium. The rutile structure, exemplified by TiO₂, consists of slightly distorted hexagonal close-packed layers of oxide ions with Ti⁴⁺ cations occupying half the octahedral interstices, forming chains of edge-sharing TiO₆ octahedra along the c-axis in a tetragonal unit cell. In contrast, the corundum structure of Al₂O₃ features oxygen anions in a hexagonal close-packed arrangement, with Al³⁺ cations filling two-thirds of the octahedral sites, leading to a trigonal symmetry and strong directional bonding. Non-metal oxides like silica (SiO₂) typically form covalent network structures, where silicon-oxygen bonds create extended three-dimensional frameworks rather than discrete ionic units. , the most stable polymorph at ambient conditions, comprises corner-sharing SiO₄ tetrahedra arranged in helical chains parallel to the c-axis, with each bonded to four oxygens and each oxygen bridging two silicons, yielding a trigonal with open channels. polymorphs exhibit a more open covalent network, with SiO₄ tetrahedra connected in a diamond-like , alternating between cubic (β-cristobalite) and tetragonal (α-cristobalite) forms, resulting in lower compared to quartz due to larger Si-O-Si bond angles. Defects in ionic oxide s play a crucial role in their electrical properties, particularly by enabling ion migration. Schottky defects involve the formation of paired cation and anion vacancies to preserve charge neutrality, as seen in rock salt-type oxides like MgO, where the energy cost is balanced by the 's flexibility./Crystal_Lattices/Lattice_Defects/Schottky_Defects) Frenkel defects, conversely, occur when a cation (or sometimes anion) is displaced from its lattice site to an position, creating a vacancy-interstitial pair; this is prevalent in oxides with significant size disparity between ions, such as ZnO or certain zirconates./Crystal_Lattices/Lattice_Defects/Frenkel_Defects) These intrinsic defects contribute to ionic by providing pathways for ; for instance, in (ZrO₂ doped with Y₂O₃), substitution of Zr⁴⁺ by Y³⁺ generates oxygen vacancies (akin to Schottky-type defects) that facilitate O²⁻ migration, achieving conductivities up to 0.1 S/cm at 1000°C suitable for electrolytes. Polymorphism in oxides arises from different stable lattice arrangements under varying conditions, influencing stability and reactivity. Titanium dioxide (TiO₂) demonstrates this with its anatase and rutile phases: rutile features tetragonal with edge-sharing TiO₆ octahedra forming linear chains and a denser packing ( ~4.25 g/cm³), while anatase has body-centered tetragonal with more distorted octahedra sharing edges and corners in a zigzag pattern, leading to a lower (~3.89 g/cm³) and higher . The anatase phase is metastable and transforms to the thermodynamically stable rutile upon heating above ~600°C, a process driven by volume contraction and minimization.

Physical and Chemical Properties

Physical Properties

Oxides exhibit a range of physical states depending on their and bonding nature. Most metal oxides are solids at , characterized by strong ionic or covalent networks that result in high s; for example, (MgO) is a crystalline solid with a of 2825 °C. In contrast, many non-metal oxides exist as gases under standard conditions due to weak intermolecular forces, such as (CO₂), which has a low sublimation point of -78.5 °C and a of 1.98 g/L at . Liquid oxides are uncommon at ambient conditions but can occur in molten forms at elevated temperatures, particularly in silicate-based glasses where network structures allow fluidity upon heating. Density and vary significantly between ionic and molecular oxides, reflecting differences in atomic packing and bond strength. Ionic metal oxides, such as MgO, typically display high around 3.58 g/cm³ and moderate to high hardness, with MgO registering 5.5–6.0 on the due to its rock-salt crystal structure. Covalent network oxides like (SiO₂) in form also exhibit substantial density (2.65 g/cm³) and exceptional hardness (Mohs 7), approaching that of in some polymorphs, which contributes to their use in abrasives. Molecular oxides, however, have lower densities; CO₂ gas, for instance, is only 1.53 times denser than air at standard conditions. Color and optical properties of oxides are influenced by electronic transitions and defects in their structures. Many metal oxides appear white or colorless in pure form, such as zinc oxide (ZnO), a white powder that scatters light effectively due to its wide bandgap. Others display distinct colors from d-d transitions or charge transfer, including black copper(II) oxide (CuO), resulting from its monoclinic crystal lattice absorbing visible light. Transparency is notable in certain oxides like aluminum oxide (Al₂O₃) in its form, which is colorless and transmits light from the (down to ~150 nm) through the near-infrared (~5 μm) with minimal absorption. Thermal stability is a hallmark of many refractory oxides, enabling applications in high-temperature environments. (ThO₂), for example, possesses one of the highest points among oxides at 3300 °C (3573 K), attributed to its and strong Th-O bonds that resist . This stability contrasts with less robust molecular oxides like CO₂, which sublimes readily at low temperatures.

Acid-Base Properties

Oxides are classified into acidic, basic, amphoteric, and neutral categories based on their interactions with , acids, and bases, which reflect their position in the periodic table and bonding characteristics. Metal oxides from the left side of the periodic table tend to exhibit basic properties, while non-metal oxides from the right side are typically acidic, with transitional behavior in the middle. Basic oxides, primarily those of and metals, react with to form hydroxides that act as bases and also neutralize acids. For example, reacts with to produce :
\ce{Na2O + H2O -> 2NaOH}
Additionally, , or quicklime, reacts with to form and :
\ce{CaO + 2HCl -> CaCl2 + H2O}
Acidic oxides, commonly formed by non-metals, combine with to yield acids and react with bases to form salts. , for instance, forms upon dissolution in :
\ce{CO2 + H2O -> H2CO3}
reacts with to produce and :
\ce{SO3 + 2NaOH -> Na2SO4 + H2O}
Amphoteric oxides display dual behavior, acting as bases toward acids and as acids toward bases, often seen in oxides of metals near the boundary like aluminum. Aluminum oxide dissolves in to form aluminum chloride:
\ce{Al2O3 + 6HCl -> 2AlCl3 + 3H2O}
It also reacts with to form :
\ce{Al2O3 + 2NaOH -> 2NaAlO2 + H2O}
Neutral oxides do not exhibit significant acid-base reactivity with , acids, or bases. Examples include (N₂O) and (CO), which remain largely unreactive in such contexts.

Reactions

Reduction of oxides involves the removal of oxygen to recover the elemental form of the metal or non-metal, a critical process in and . This is achieved through various methods, each suited to the and reactivity of the oxide. The feasibility of these reductions is governed by , where the standard change (ΔG°) determines whether the proceeds spontaneously under given conditions. Thermal reduction using carbon is a widely employed method for less stable metal oxides, particularly in iron production. In the blast furnace process, iron(III) oxide (Fe₂O₃) reacts with carbon to produce metallic iron and carbon monoxide:
\ce{Fe2O3 + 3C -> 2Fe + 3CO}
This reaction occurs at high temperatures (around 1500–2000°C), where coke serves as the carbon source and reducing agent, facilitated by the hot blast of air that generates CO as an intermediate reductant. The process is exothermic in later stages, contributing to the furnace's heat balance, and has been the cornerstone of steelmaking since the 18th century.
For more reactive or less stable oxides, hydrogen gas provides a cleaner reducing agent, avoiding carbon-related impurities. A representative example is the reduction of copper(II) oxide:
\ce{CuO + H2 -> Cu + H2O}
This reaction proceeds at moderate temperatures (200–300°C) and is often studied for its mechanistic insights, involving surface adsorption and stepwise oxide phase transformations from CuO to Cu₂O and then to Cu metal. Hydrogen reduction is advantageous in producing high-purity metals and is explored for sustainable alternatives to carbon-based methods.
Electrochemical reduction enables the extraction of metals from highly stable oxides that resist thermal methods. The Hall-Héroult process, developed in 1886, electrolyzes alumina (Al₂O₃) dissolved in molten cryolite (Na₃AlF₆) at approximately 950°C. At the cathode, aluminum ions are reduced to molten aluminum:
\ce{Al^3+ + 3e^- -> Al}
while at the carbon anode, oxygen ions form CO₂ via reaction with the anode: overall simplifying to
\ce{Al2O3 -> 2Al + 3/2 O2}
(with oxygen combining with carbon). This method consumes significant energy (about 13–15 kWh/kg Al) but is essential for aluminum production, accounting for over 95% of global output.
Recent advancements include hydrogen direct reduction (HyDR) processes for iron oxides, which use pure instead of carbon to produce (DRI) with minimal CO₂ emissions. As of 2025, pilot plants and commercial facilities, such as those by HYBRIT and H2 Green Steel, demonstrate feasibility at temperatures around 800–1000°C, supporting the transition to low-carbon steelmaking. The thermodynamic viability of these reductions is visualized using the , which plots the standard free energy of formation (ΔG°) for oxide formation reactions against . Lines with steeper negative slopes indicate greater oxide stability; a reductant line (e.g., for C to CO or H₂ to H₂O) must lie below the oxide line for spontaneous reduction. For instance, carbon can reduce iron oxides above ~700°C, but not alumina at practical temperatures, explaining the preference for electrochemical methods in aluminum extraction. This diagram, first constructed by Harold Ellingham in 1944, remains a fundamental tool for predicting reduction processes across metals.

Hydrolysis and Dissolution

Basic metal oxides react with through to form the corresponding metal hydroxides, which are typically sparingly soluble bases. For example, undergoes a slow reaction:
\ce{MgO + H2O -> Mg(OH)2}
This process is limited by the low of Mg(OH)₂ and often requires elevated temperatures or prolonged exposure to achieve significant conversion. In contrast, oxides of metals, such as , react more readily and exothermically with to produce strongly basic solutions of hydroxides.
Acidic non-metal oxides hydrolyze vigorously with water to yield oxyacids. , for instance, forms :
\ce{SO3 + H2O -> H2SO4}
Similarly, hydrolyzes to :
\ce{P4O10 + 6H2O -> 4H3PO4}
These reactions are highly exothermic and proceed rapidly, reflecting the covalent nature and high of the non-metals involved.
Amphoteric oxides, such as , exhibit dual behavior by dissolving in either acidic or aqueous solutions. In acidic media, ZnO acts as a :
\ce{ZnO + 2H+ -> Zn^{2+} + H2O}
In alkaline conditions, it behaves as an , forming soluble ions:
\ce{ZnO + 2OH- + H2O -> [Zn(OH)4]^{2-}}
This pH-dependent solubility arises from the intermediate of , allowing the oxide to coordinate with either H⁺ or OH⁻. Network-forming oxides like silica (SiO₂) are largely insoluble in and most acids but dissolve slowly in due to the formation of stable fluorosilicates:
\ce{SiO2 + 4HF -> SiF4 + 2H2O}
The reaction is kinetically hindered and typically requires concentrated for practical rates.
Solubility trends for basic oxides show an increase down a group in the periodic table, driven by decreasing and increasing ionic size of the metal cation. In Group 2, for example, the hydroxides derived from oxide —such as Mg(OH)₂, Ca(OH)₂, Sr(OH)₂, and Ba(OH)₂—exhibit progressively higher , with Ba(OH)₂ being the most soluble and producing solutions with values approaching 14. Alkali metal oxides () generally display even greater than their Group 2 counterparts due to lower . The stability of oxides in aqueous environments is highly -dependent, as depicted in Pourbaix diagrams, which plot the predominant species (oxide, , or dissolved ions) against and . These thermodynamic maps reveal stability regions where oxides persist, such as the passivation zone for aluminum oxide in neutral to basic conditions, and dissolution domains in extreme , aiding predictions of or in varied aqueous media.

Applications and Significance

Industrial Applications

Oxides play a critical role in materials used to line high-temperature furnaces, where their and resistance to chemical attack are essential. Alumina (Al₂O₃) refractories, often in the form of high-alumina bricks with 80-85% Al₂O₃ content, are widely employed in blast furnaces, stoves, and reverberatory furnaces for aluminum due to their ability to withstand temperatures up to 1700°C and resist erosion. (MgO) refractories, prized for their high above 2800°C and basicity that neutralizes acidic slags, are integral to steelmaking processes in basic oxygen furnaces and furnaces, particularly in high-wear zones like converter linings. In ceramics, silica (SiO₂) serves as the primary component in , comprising approximately 70-75% of soda-lime flat formulations, which account for the majority of global flat glass output and enable the formation of durable, transparent sheets through the process. Certain oxides are indispensable as pigments and catalysts in industrial coatings and chemical processes. Titanium dioxide (TiO₂), the most widely used white pigment, imparts opacity and brightness to paints, plastics, and paper, with global annual production exceeding 5 million metric tons to meet demand in these sectors. Iron(III) oxide (Fe₂O₃), often as micaceous or synthetic variants, functions as a barrier pigment in anticorrosive paints, enhancing rust-proofing by forming a lamellar structure that impedes moisture and oxygen diffusion on metal surfaces. In , (CaO), derived from , acts as a flux in to remove impurities such as silica (SiO₂) and by forming a molten , typically (CaSiO₃), which floats atop the molten steel and is easily separated, thereby purifying the metal and improving its quality. Oxides also enable key electronic components through their dielectric properties. Zinc oxide (ZnO), in polycrystalline ceramic form, is the core material in varistors, which exhibit nonlinear voltage-current characteristics due to grain-boundary Schottky barriers, providing surge protection in electrical circuits against overvoltages. (BaTiO₃), a ferroelectric ceramic with a high constant exceeding 1000, is fundamental to multilayer ceramic s, where it allows compact energy storage in electronic devices like smartphones and power supplies.

Environmental and Modern Uses

Oxides play significant roles in environmental processes, both as contributors to ecological challenges and as natural components of filtration systems. (CO₂), an , is a primary responsible for , with emissions from and industry reaching 37.4 gigatonnes (Gt) in 2024, exacerbating through enhanced . (SO₂), emitted mainly from combustion, reacts with atmospheric and oxidants to form , contributing to that acidifies soils, lakes, and forests, harming aquatic life and . In contrast, (SiO₂), abundant in soils as and silicates, enhances natural by improving soil and acting as a medium in aquifers and riverbeds to trap sediments and pollutants, thereby purifying and . In pollution control, basic oxides like (CaO) are widely used in (FGD) processes at coal-fired power plants, where it reacts with SO₂ to form (CaSO₃), reducing emissions by up to 95% and mitigating formation: \ce{CaO + SO2 -> CaSO3}. Similarly, metal oxides such as alumina (Al₂O₃) serve as supports for (Pt) catalysts in automotive catalytic converters, enabling the oxidation of (CO) and hydrocarbons while reducing nitrogen oxides (NOₓ), thereby lowering urban levels. Modern applications of oxides increasingly leverage for sustainable technologies. (TiO₂) nanoparticles excel in for , where under UV light they generate to degrade organic pollutants like dyes and pharmaceuticals; advancements in the include doping with metals or coupling with to extend activity into visible light, achieving near-complete degradation in pilot-scale systems. oxides, such as (SrTiO₃), function as efficient electron transport layers in hybrid cells, contributing to power conversion efficiencies exceeding 25% by 2025 through improved charge extraction and reduced recombination, as demonstrated in tandem configurations. Recent developments highlight innovative uses of oxide nanostructures in and . Zinc oxide (ZnO) quantum dots, with their tunable bandgap, have advanced light-emitting diodes (LEDs) by serving as electron transport layers, enabling external quantum efficiencies up to 33.7% in blue QLEDs through operando recrystallization techniques that enhance charge injection. In cancer therapy, magnetite (Fe₃O₄) nanoparticles enable magnetic , where alternating magnetic fields induce localized heating to 42–45°C, selectively killing tumor cells; post-2020 clinical trials, including first-in-human studies for advanced , have confirmed safety and feasibility when combined with .

References

  1. [1]
    Oxides - Chemistry LibreTexts
    Jun 30, 2023 · Oxides are chemical compounds with one or more oxygen atoms combined with another element (e.g. Li2O). Oxides are binary compounds of oxygen ...
  2. [2]
    Oxide Compound - an overview | ScienceDirect Topics
    Oxide materials are the chemical compounds composed of at least one oxygen atom in combination with other atom(s). Oxygen atoms associate/bond with metals ( ...
  3. [3]
    Metal Oxide - an overview | ScienceDirect Topics
    Metal oxides are widely used in daily life, such as copper oxide, alumina, iron oxide, and so on. Copper oxide (CuO) is a kind of black oxide of copper, ...
  4. [4]
    Technical and Economic Importance of Oxides - Wiley Online Library
    Mar 25, 2018 · The applications of oxides are associated with their properties. The key properties are mechanical, electrical and magnetic, thermal, and ...<|control11|><|separator|>
  5. [5]
    Visualizing the Abundance of Elements in the Earth's Crust
    Dec 7, 2021 · #1: Oxygen ... Oxygen is by far the most abundant element in the Earth's crust, making up 46% of mass—coming up just short of half of the total.
  6. [6]
    Oxide - Etymology, Origin & Meaning
    Originating from French chemists Guyton de Morveau and Lavoisier in 1790, oxide means a compound of oxygen with another element.
  7. [7]
    Classification Of Oxides - BYJU'S
    Classification Of Oxides · Acidic oxides · Basic oxides · Amphoteric oxides · Neutral oxides ...
  8. [8]
    Oxides - Definition, Types, Valency & Metallic Character - Chemistry
    Metals have a general tendency to form basic oxides and non-metals have a general tendency to form acidic oxides. As we move down the group in a periodic table, ...
  9. [9]
    8.7: Spinel, Perovskite, and Rutile Structures - Chemistry LibreTexts
    Sep 27, 2021 · There are also mixed spinels, which are intermediate between the normal and inverse spinel structure. Some spinel and inverse spinel AB ...
  10. [10]
    Mixed Valence in Chemistry - SpringerLink
    Mixed valency is readily recognized by the chemical formula itself (Pb3 O4, TinO2n−1), but there are solids which show apparent integral oxidation states.
  11. [11]
    [PDF] Metal Oxide Surfaces and Their Interactions with Aqueous Solutions ...
    ample of this for R-Fe2O3 (hematite). UPS spectra are shown for a stoichiometric R-Fe2O3 (0001) surface, where all cations are Fe(III). When the surface is.
  12. [12]
    [PDF] Metal oxide redox chemistry for chemical looping processes
    Nov 6, 2018 · oxides, prominent examples of which are Fe2O3 and. CuO. These active materials are often complemented by dopants, supports (for example ...
  13. [13]
    Defect structure of ferrous oxide O | Phys. Rev. B
    Jun 1, 1995 · We have investigated the defect structure of nonstoichiometric wüstite F e 1 − x O as a function of temperature and oxygen partial pressure ...
  14. [14]
    The complex nonstoichiometry of wüstite Fe 1- z O: Review and ...
    The NaCl face centered cubic structure of wüstite is characterized by a high rate of iron vacancies, and the existence of a certain proportion of iron ions ...
  15. [15]
    The effect of stoichiometry on the structural, thermal and ... - NIH
    Thus, highly defective nickel oxide at lower temperature approaches stoichiometric nickel oxide as the heating temperature increases, presumably, due to change ...
  16. [16]
    Insights into defect cluster formation in non-stoichiometric wustite ...
    Feb 14, 2025 · The defect concentration in Fe1−xO is extremely high, and the interactions of these defects within the solid phase are very complex.
  17. [17]
    [PDF] Nomenclature of Inorganic Chemistry | IUPAC
    Chemical nomenclature must evolve to reflect the needs of the community that makes use of it. In particular, nomenclature must be created to describe new ...
  18. [18]
    [PDF] Brief Guide to the Nomenclature of Inorganic Chemistry | IUPAC
    Note that the 'ido' ending is now used for halide and oxide ligands as well. By convention, a single coordinated hydrogen atom is always considered anionic and ...
  19. [19]
    [PDF] Study of Calcination-Carbonation of Calcium Carbonate in Different ...
    May 2, 2011 · The calcination reaction (CaCO3 = CaO + CO2) was studied at four temperatures: 800oC, 900oC, 950oC and 1000oC for each of the three media, CO2, ...
  20. [20]
    [PDF] Chapter 2. Production and Processing of Aluminum
    Tropical monohydrate bauxite grades yielding 35–55% Al2O3 will no doubt continue to be the most favored aluminum ores for many decades. Laterite rocks similar ...
  21. [21]
    Zirconia: Synthesis and Characterization - IntechOpen
    Alkali fusion and plasma thermal dissociation are the important decomposition methods adopted widely to convert zircon to zirconia.
  22. [22]
    Occurrence, Preparation, and Compounds of Oxygen – Chemistry
    Most nonmetals react with oxygen to form nonmetal oxides. Depending on the available oxidation states for the element, a variety of oxides might form. Fluorine ...
  23. [23]
    11.6: Combustion Reactions - Chemistry LibreTexts
    Mar 20, 2025 · The products of the combustion of hydrocarbons are carbon dioxide and water. Many hydrocarbons are used as fuel because their combustion ...
  24. [24]
    An Introduction to the Combustion of Carbon Materials - 2022
    May 30, 2022 · Carbon combustion generates both CO and CO2 as primary products as discussed above. A simple concentration boundary layer is represented in ...
  25. [25]
    Burning of Sulfur to Produce Sulfur Dioxide
    The sulfur reacts with oxygen in the flask to form sulfur dioxide gas, which then dissolves in the water, forming sulfurous acid. The solution turns to a yellow ...
  26. [26]
    Deville rebooted &#x2013; practical N2O5 synthesis - RSC Publishing
    May 9, 2024 · After Deville' discovery, the most widely used method of synthesising N2O5 is the dehydration of nitric acid with phos- phorous pentoxide (P4O ...<|separator|>
  27. [27]
    The Contact Process - Chemistry LibreTexts
    Jan 29, 2023 · Step 1: Make sulfur dioxide · Step 2: Convert sulfur dioxide into sulfur trioxide (the reversible reaction at the heart of the process) · Step 3: ...
  28. [28]
    Covalent Bond - BYJU'S
    Each oxygen atom shares its two electrons with carbon, and therefore there are two double bonds in CO2. CO2 Molecule with Double Covalent Bond. Oxygen Molecule: ...
  29. [29]
    Hybridization of Carbon dioxide- Structure of CO2 and Properties
    Covalent bonds involve the sharing of electrons between atoms. In the case of CO 2, each oxygen atom shares two electrons with the central carbon atom. Double ...
  30. [30]
    Molecular Geometry, Lewis Structure, and Bond Angle of SO3
    May 13, 2023 · So, the hybridization in SO3 is sp2. VSEPR theory is an accurate method of predicting molecular geometry. According to this theory, the bond ...
  31. [31]
    Nitrous Oxide | N2O | CID 948 - PubChem
    Nitrous Oxide | N2O | CID 948 - structure, chemical names, physical and chemical properties, classification, patents, literature, biological activities, ...
  32. [32]
    Chlorine monoxide | Cl2O | CID 24646 - PubChem - NIH
    Dichlorine monoxide is formed by passing chlorine diluted with an inert gas over the surface of a slowly stirred 50% NaOH solution at about 5 °C. The exit gas ...
  33. [33]
    Carbon dioxide - the NIST WebBook
    This IR spectrum is from the Coblentz Society's evaluated infrared reference spectra collection. ... Data from NIST Standard Reference Database 69: NIST Chemistry ...
  34. [34]
  35. [35]
  36. [36]
    mp-2352: Na2O (cubic, Fm-3m, 225) - Materials Project
    Na2O is Fluorite structured and crystallizes in the cubic Fm-3m space group. The structure is three-dimensional. Na1+ is bonded to four equivalent O2- atoms ...
  37. [37]
    TiO 2 – Rutile: Interactive 3D Structure - ChemTube3D
    The rutile structure adopted by one polymorph of TiO2: (a) the buckled close-packed layers of oxide ions, arrowed, with titanium cations in half the octahedral ...
  38. [38]
    Corundum structure Al2O3 - ChemTube3D
    The corundum structure, as adopted by Al2O3, with cations occupying two-thirds of the octahedral holes between layers of close-packed oxide ions. Al2O3 corundum.
  39. [39]
    [PDF] 1 Chapter 12: Structures of Ceramics Outline Introduction Crystal ...
    □ Silica, silicon oxide (SiO2),. • Three crystal structures: quartz, cristobalite, and tridymite. • Open structure, not close-packed, low density. Cristobalite ...
  40. [40]
    A review on recent status and challenges of yttria stabilized zirconia ...
    Oct 27, 2019 · This paper is the first review which focused on the recent status and challenges of YSZ electrolyte towards lowering the operating temperature.Summary · INTRODUCTION · CURRENT STATUS OF...
  41. [41]
    Prediction of TiO 2 Nanoparticle Phase and Shape Transitions ...
    According to the present anal., anatase becomes more stable than rutile when the particle size decreases below ∼14 nm. The calcd. phase boundary between ...
  42. [42]
    Magnesium Oxide | MgO | CID 14792 - PubChem
    5 Melting Point. 5072 °F (NIOSH, 2024). CAMEO Chemicals. 2825 °C. Haynes, W.M. (ed.) CRC Handbook of Chemistry and Physics. 91st ed. Boca Raton, FL: CRC Press ...
  43. [43]
    Carbon Dioxide | CO2 | CID 280 - PubChem
    Absolute density, gas at 101.325 kPa at 0 °C: 1.9770 kg/cu m; relative ... IDENTIFICATION AND USE: Carbon dioxide (CO2) is a colorless gas and liquid ...
  44. [44]
    Physical State and Structures of Oxides
    The Physical States and Structures of Oxides · The Gaseous State · The Liquid State · The Solid State.
  45. [45]
    Properties: Silica - Silicon Dioxide (SiO2) - AZoM
    Silica - Silicon Dioxide (SiO2) ; Fracture Toughness, 0.62, 0.67 ; Hardness, 4500, 9500 ; Loss Coefficient, 8e-006, 2e-005 ; Modulus of Rupture, 110, 200 ...
  46. [46]
    Copper oxide (CuO) | CuO | CID 164827 - PubChem - NIH
    Cupric oxide, or copper (II) oxide, is an inorganic compound with the chemical formula CuO. Cupric oxide is used as a precursor in many copper-containing ...<|separator|>
  47. [47]
    Optical properties of Sapphire - SHINKOSHA Crystals for a bright ...
    Pure sapphire is colorless and transparent, transmitting light from UV to IR. It has birefringence, and can transmit up to 150nm with few oxygen defects.
  48. [48]
    Thermal transport in thorium dioxide - ScienceDirect.com
    Thorium oxide (ThO2), also known as thoria, has one of the highest melting points among all oxides (3573.15 K) [5]. Consequently, ThO2 is used in light ...
  49. [49]
    [PDF] 12.5 Iron And Steel Production 12.5.1 Process Description - EPA
    Iron oxides, coke and fluxes react with the blast air to form molten reduced iron, carbon monoxide (CO), and slag. The molten iron and slag collect in the ...
  50. [50]
    Dual mechanisms in hydrogen reduction of copper oxide
    Mar 26, 2024 · The study of the reduction of copper oxide (CuO) by hydrogen (H2) is helpful in elucidating the reduction mechanism of oxygen carriers.
  51. [51]
    The Aluminum Smelting Process and Innovative Alternative ... - NIH
    May 8, 2014 · Alumina is dissolved in the electrolyte and is electrolyzed at the cathode to form molten aluminum. There is a high ambient temperature in the ...
  52. [52]
    [PDF] Ellingham Diagrams - MIT
    The Ellingham diagram shown is for metals reacting to form oxides (similar diagrams can also be drawn for metals reacting with sulfur, chlorine, etc., but the ...
  53. [53]
    [PDF] Chemical Reactions
    Reaction of a nonmetal oxide with water produces an oxyacid in which the nonmetal is in the same oxidation state as in the oxide you started with. Both of these ...
  54. [54]
    Acid-base Behavior of the Oxides - Chemistry LibreTexts
    Jun 30, 2023 · Non-metal oxide acidity is defined in terms of the acidic solutions formed in reactions with water—for example, sulfur trioxide reacts with ...
  55. [55]
    Chemical Reactivity - Grandinetti Group
    If a non-metal oxide dissolves in water, it will form an acid. Non-Metal Oxide + Water → Acid. For example,. SO3(g) + H2 ...
  56. [56]
    7.8A: Amphoteric Behavior - Chemistry LibreTexts
    Jan 15, 2023 · Amphoterism depends on the oxidation state of the oxide. For example, zinc oxide (ZnO) reacts with both acids and with bases: In acid: ZnO + 2H+ ...<|separator|>
  57. [57]
    Dissolution of silicate minerals by hydrofluoric acid - ACS Publications
    An new experimental study on the geochemical behaviors of metals in fluorine-bearing hydrothermal fluids based on the Monel alloy solubility method.
  58. [58]
    Solubility of the hydroxides, sulphates and carbonates of the Group ...
    Hydroxides become more soluble, sulfates and carbonates become less soluble down Group 2, except barium carbonate is slightly more soluble than strontium ...<|separator|>
  59. [59]
    4.6: Pourbaix Diagrams - Chemistry LibreTexts
    May 3, 2023 · Pourbaix Diagrams plot electrochemical stability for different redox states of an element as a function of pH.
  60. [60]
    Alumina and Alumina Refractories - IspatGuru
    Jan 25, 2015 · Refractories with 80 % and 85 % Al2O3 were originally developed for use in aluminum smelting and holding furnaces. It is rare that they find ...
  61. [61]
    Refractory & Steelmaking - Magnesia Supplier
    Due to its excellent corrosion resistance, refractory grade fused magnesia is used in high wear areas in steel making, eg, basic oxygen and electric arc ...
  62. [62]
    The Chemistry of Glass - Pilkington
    The main constituent of Flat Glass is SiO2 (silica sand). This has a high melting temperature in the region of 1700 degrees C and its state at this temperature ...
  63. [63]
    Titanium Dioxide Market Size, Share, Growth & Forecast 2035
    The global Titanium Dioxide market was 5418 thousand tonnes in 2024, expected to reach 9394 thousand tonnes in 2035, with a 5.18% CAGR.
  64. [64]
    Iron oxides in anticorrosive paints - Zona de Pinturas
    Jan 12, 2021 · Below we analyze the way in which natural iron oxides act as "barrier" pigments in the formulation of anticorrosive paints.
  65. [65]
    How lime enhances steel production - Lhoist
    It is used predominantly as a fluxing agent during the ironmaking and steelmaking processes to promote slag formation which is essential to capture and remove ...
  66. [66]
    The physics of zinc oxide varistors - AIP Publishing
    Oct 1, 1977 · This paper presents a physical description of the action of ZnO varistors, which are complex ceramic bodies of ZnO grains sintered in an oxide flux.
  67. [67]
    Ceramic Capacitors FAQQWhat is the mechanism of the changing of ...
    In high dielectric constant ceramic capacitors, at present BaTiO3 (barium titanate) is used as the principal component of the ceramic. As shown below, BaTiO3 ...
  68. [68]
    Fossil fuel CO<sub>2</sub> emissions increase again in 2024
    Nov 13, 2024 · The 2024 Global Carbon Budget projects fossil carbon dioxide (CO 2 ) emissions of 37.4 billion tonnes, up 0.8% from 2023.
  69. [69]
    What is Acid Rain? | US EPA
    Mar 4, 2025 · Acid rain results when sulfur dioxide (SO2) and nitrogen oxides (NOX) are emitted into the atmosphere and transported by wind and air currents.
  70. [70]
    4500-SiO2 SILICA - Standard Methods For the Examination of Water ...
    The average abundance of silica in different rock types is 7% to 80%, in typical soils 50% to 80%, and in surface and groundwater 14 mg/L.
  71. [71]
    [PDF] Flue Gas Desulphurization in Circulating Fluidized Beds. - HAL
    Oct 15, 2019 · Limestone and/or lime are used to enable SO2 gas to chemically react with CaCO3 or CaO forming calcium sulphate (CaSO4) and calcium sulphite ( ...
  72. [72]
    Modified Titanium dioxide-based photocatalysts for water treatment
    TiO2 photocatalysts have attracted great attention as a promising water purification technology due to their environmentally friendly and efficient properties.
  73. [73]
    Perovskite Oxide SrTiO3 as an Efficient Electron Transporter for ...
    Nov 19, 2014 · The authors report on a significant power conversion efficiency improvement of perovskite solar cells from 8.81% to 10.15% due to insertion of ...Introduction · Results and Discussion · Conclusions · Supporting Information
  74. [74]
    Operando ZnO recrystallization for efficient quantum-dot light ...
    May 15, 2025 · When used as the electron transport layer in QLEDs, recrystallized ZnO NPs enhance the external quantum efficiency from 17.2% to 33.7% compared ...
  75. [75]
    Extended Treatment with Magnetic Nanoparticle Hyperthermia in a ...
    Extended Treatment with Magnetic Nanoparticle Hyperthermia in a “First-in-Human” Trial Demonstrates Safety and Feasibility in an Advanced Cervical Cancer Case ...