Fact-checked by Grok 2 weeks ago

Calcination

Calcination is a pyrometallurgical involving the treatment of materials at high temperatures, typically above 900°C, to drive off volatile components such as , chemically bound , or without causing or of the substance. This endothermic reaction often results in , phase transitions, or the purification of the material, producing a more stable or residue known as calcine. The is conducted in controlled atmospheres, such as in the absence of oxygen to prevent oxidation, and is distinguished from by its focus on rather than sulfation. Historically, calcination originated from the Latin term calcinare, referring to the heating of () to produce quicklime () by releasing CO₂, a reaction exemplified by CaCO₃(s) → CaO(s) + CO₂(g) at temperatures exceeding 900°C. Over time, the term has broadened beyond lime production to encompass similar treatments in various chemical and metallurgical contexts, evolving with industrial advancements in the 19th and 20th centuries to include ore beneficiation and material synthesis. In modern applications, calcination plays a critical role in industries such as cement manufacturing, where is decomposed into for clinker production, and in alumina extraction via the , converting aluminum hydroxide to alumina at 1300–1500°C through reactions like 2Al(OH)₃(s) → Al₂O₃(s) + 3H₂O(g). It is also essential in and production, enhancing material properties like purity, , and reactivity by removing impurities and volatiles. Common equipment includes rotary kilns for continuous processing, which allow precise control of temperature profiles and retention times—often 6 hours to several days depending on the material—and calciners for uniform heating of fine particles. Key process parameters, including (typically 800–1340°C), , and of gases like CO₂, significantly influence reaction efficiency and product quality, with higher temperatures yielding harder, less reactive limes suitable for uses. Environmental considerations are prominent, as calcination emits substantial CO₂, contributing to efforts in carbon capture and sustainable designs.

Fundamentals

Definition and Process Overview

Calcination is a treatment applied to solid materials, involving heating at high temperatures—typically in the range of 800–1400°C—within a , often with limited oxygen or air, to induce chemical changes such as , phase transformations, or the removal of volatile components, all without causing the material to melt. This is fundamental in and for altering the physical and chemical properties of substances like ores, minerals, and precursors, enhancing attributes such as stability, reactivity, or purity. The key steps in calcination include preheating the solid feedstock in a or to gradually elevate the , maintaining the heat under controlled conditions to drive the desired reactions—such as the of carbonates or hydroxides into oxides—while closely monitoring to prevent or excessive , and finally cooling the product to stabilize it. Resulting products often include stable oxides; for instance, the calcination of follows the general reaction: \mathrm{CaCO_3 \rightarrow CaO + CO_2} where yields quicklime and gas, though detailed mechanisms are governed by reaction kinetics and . Calcination is distinct from related thermal processes: unlike , which typically occurs in an oxidizing atmosphere with ample air to convert sulfides into oxides or sulfates, calcination emphasizes in a limited-oxygen to avoid such oxidation. In contrast to , which primarily bonds particles through diffusion at high temperatures without significant , calcination focuses on inducing compositional changes like or decarbonation while avoiding particle . Common equipment includes rotary kilns, which rotate to ensure uniform heating and material movement at temperatures up to 1400°C; reactors, utilizing gas flow to suspend fine particles for efficient at 800–1200°C; and shaft kilns, employing vertical gravity-fed designs for continuous operation in the 900–1300°C range.

Etymology

The term "calcination" derives from the Medieval Latin verb calcināre, meaning "to heat" or "to burn to lime," which itself stems from the Late Latin calcīna (lime) and the classical Latin noun calx (lime, limestone, or burnt stone). This root reflects the process's ancient association with heating limestone to produce lime, a practice central to early construction and metallurgy. The word entered English in the as calcinen or similar forms, borrowed via calciner (to calcine), with the earliest recorded use of "calcination" appearing around 1386 in the works of , where it denoted the fiery purification of substances. Initially tied to alchemical contexts of reducing metals or minerals to a powdery "" through intense heat, the term's meaning evolved to encompass broader thermal treatments for . Etymologically linked to "calcium," the chemical element isolated from lime compounds and named in 1808 by from Latin calx, the word also connects to "quicklime" (unslaked lime, or calx viva), underscoring its origins in lime production. The Latin calx traces further to kháliks (χάλιξ), meaning "pebble" or "small stone," highlighting the linguistic progression from natural stone to processed lime. Over time, "calcination" shifted semantically from alchemical "purgation by fire" to its modern denotation of controlled in chemistry.

Historical Development

Ancient and Medieval Uses

The practice of calcination originated in the as early as 12,000 BCE, where early communities produced quicklime by heating in simple hearths or kilns to create binders for plasters and mortars. Archaeological evidence from sites like in reveals lime plasters used in wall coatings and flooring as early as 6000 BCE, indicating controlled calcination processes for construction and symbolic purposes, such as modeling human skulls. Similarly, at Aşıklı Höyük in central , dated to circa 7000 BCE, lime-based plasters containing calcined materials demonstrate the technology's role in stabilizing mud-brick structures and creating durable surfaces. By the time of , around 2600 BCE, calcined materials including and were integral to monumental . mortar served as a key component for binding massive limestone blocks in the pyramids at , while plasters provided weather-resistant coatings. This application extended the utility of calcination beyond basic shelter to large-scale engineering feats, enhancing structural integrity in arid environments. In the Roman era, from the BCE onward, calcination produced for opus caementicium, a hydraulic mixed with pozzolanic additives like , enabling the construction of enduring infrastructure such as aqueducts, harbors, and the . Archaeological remnants of Roman kilns and samples from sites like confirm on-site calcination, which facilitated imperial expansion by supporting vast building projects that conveyed water over long distances and housed . Medieval advancements refined calcination techniques, with improved kiln designs emerging in both and the to boost efficiency and output for diverse applications. In the Islamic realm, early structures like the in (built 670 CE and expanded in the ) incorporated calcined mortars for enhanced durability and water resistance, while Umayyad-period kilns in the , dated to the 7th–8th centuries, featured advanced draft systems for consistent high-temperature burning. European developments paralleled this, as seen in castles from the , where calcined supported rapid fortification efforts. These innovations extended calcination to glassmaking, where quicklime acted as a stabilizer in alkali- recipes prevalent in 9th–10th century Abbasid , and to , aiding in the production of glazed wares through lime-based fluxes that improved firing and surface quality.

Alchemical Significance

In , practiced from approximately 300 BCE to 1700 CE, calcination constituted the inaugural phase of the Magnum Opus, or Great Work, embodying the stage of blackening that signified decomposition and the dissolution of base elements through intense heat, ultimately yielding a or ash as the purified residue. This process not only dismantled physical substances but also symbolized the alchemist's inner confrontation with chaos and ego, facilitating toward higher states of perfection. The eighth-century polymath , often regarded as a foundational figure in Islamic , detailed calcination as an essential technique for metal purification, involving high-temperature to separate and identify constituent principles such as and mercury, thereby enabling the refinement of base materials. In the sixteenth century, extended calcination's application to medicinal , employing fire in sealed vessels to incinerate impurities from metals like silver () and gold (), extracting noble essences for therapeutic elixirs that promoted bodily purification and vitality. Alchemical practitioners executed calcination through repeated heating of metals or minerals in specialized self-regulating furnaces known as athanors, which sustained gentle, constant warmth to volatilize volatile components while preserving essential principles, frequently in conjunction with philosophical mercury to isolate pure quintessences. Symbolically, in calcination served as the supreme purifying agent, mirroring the alchemist's spiritual journey toward by incinerating material dross and unveiling latent divinity.

Transition to Modern Chemistry

In the late , revolutionized the understanding of calcination by reinterpreting it through rigorous experimentation, particularly his studies on the calcination of mercury conducted between 1774 and 1778. Observing that mercury gained weight upon forming its () when heated in air, and that the "fixed air" (dephlogisticated air, later identified as oxygen) diminished in volume, Lavoisier concluded that calcination involved the fixation of oxygen by the metal rather than the release of phlogiston as proposed by earlier theories. This empirical framework shifted calcination from a mystical alchemical process to a quantifiable , laying the groundwork for oxygen-based theory. The saw accelerated advancements in calcination studies, driven by the Industrial Revolution's demand for efficient thermal processes in materials production. Chemists began conducting precise volumetric and gravimetric analyses of reactions, with contributing key observations in the 1810s and 1830s on the of carbonates, such as (), where he noted that the presence of or could influence the rate and completeness of calcination to form oxides. These investigations emphasized controlled heating conditions and gas evolution, providing foundational data for scaling chemical processes beyond laboratory settings. By the early , calcination had been fully incorporated into modern and , enabling predictive modeling of its underlying mechanisms. The introduction of by in 1889 offered a quantitative tool to describe the minimum energy required for bond breaking in decompositions, such as those yielding metal oxides from carbonates or hydroxides; this concept was rapidly applied to calcination , revealing how thresholds govern feasibility and . Such integrations transformed calcination into a cornerstone of , with standardized equations for and phase changes. A pivotal milestone occurred with the development of the periodic table in , where explicitly linked calcination-derived formulas to elemental classification, using properties like valency and stability (e.g., MO, M₂O₃ patterns) to predict undiscovered elements and their behaviors under thermal treatment. This contextualized calcination as essential for revealing in formation across groups, solidifying its role in systematic .

Chemical Principles

Thermal Decomposition Reactions

Thermal decomposition reactions in calcination represent a fundamental non-oxidative process where inorganic compounds, such as carbonates and hydroxides, undergo endothermic breakdown upon heating, driven by the absorption of thermal energy to sever chemical bonds and liberate volatile gases like CO₂ or H₂O. This mechanism proceeds via a topotactic decomposition pathway, particularly for carbonates, where the crystal lattice of the precursor facilitates the nucleation and growth of the resulting oxide phase while releasing the gas in a controlled manner. The endothermic nature ensures that sufficient heat input is required to overcome the lattice energy, preventing spontaneous reaction at ambient conditions and allowing precise control over the transformation. A primary example is the decomposition of magnesium carbonate, where (MgCO₃) converts to (MgO) and :
\ce{MgCO3 -> MgO + CO2}
This reaction reaches around 400°C under standard conditions, with the forward decomposition favored at higher temperatures due to the Le Chatelier principle shifting the balance away from the gaseous CO₂ product. Similarly, hydroxide dehydration, such as that of aluminum hydroxide (Al(OH)₃, often in form), yields alumina (Al₂O₃) and :
\ce{2Al(OH)3 -> Al2O3 + 3H2O}
This stepwise process typically initiates between 200–500°C, involving intermediate or transition aluminas before stabilizing as α-Al₂O₃ at 1100–1300°C.
The kinetics of these decompositions follow the , k = A e^{-E_a / RT}, where the rate constant k depends exponentially on T, with E_a quantifying the energy barrier for bond rupture. For (CaCO₃) —a for carbonate calcination—E_a is approximately 200 kJ/, reflecting the of the CO₃²⁻ and the need for substantial thermal activation. Key influencing factors include , which affects and surface area exposure (smaller particles accelerate due to shorter diffusion paths for volatiles), and the surrounding atmosphere, where inert gases like N₂ promote faster rates by minimizing reverse compared to CO₂-rich environments. During calcination, these reactions induce changes from the precursor solid (e.g., hydrated or carbonated ) to a more stable , often preserving nanoscale in the product to maintain reactivity. Careful is essential to avoid , where excessive heat causes particle coalescence and pore collapse, which would reduce the 's surface area and catalytic potential; thus, calcination protocols typically limit dwell times and temperatures to below the sintering threshold for the target .

Oxidation and Redox Processes

In the context of calcination processes involving oxygen, the procedure entails heating ores or metal compounds in a with limited air supply to facilitate the oxidation of impurities while avoiding complete or or sulfation. This variant contrasts with full , which employs excess air to ensure thorough oxidation and volatilization of as SO₂. The limited oxygen environment promotes selective reactions that convert lower-valence metal into stable oxides, enhancing the material's purity for subsequent processing without excessive energy input or unwanted side products. Key reactions in such calcination include the oxidation of lower-valence , exemplified in preparation for , where undergoes oxidation to :
$4\text{Fe}_3\text{O}_4 + \text{O}_2 \rightarrow 6\text{Fe}_2\text{O}_3
This process strengthens pellet structure by forming a dense layer, often conducted at 1100–1200°C with limited air to balance oxidation and prevent over-sintering. These reactions highlight the nature, where metals increase in , driven by oxygen as the .
Thermodynamically, these oxidation processes are governed by the change (ΔG), which must be negative for spontaneity: ΔG = ΔH - TΔS < 0 at elevated temperatures. For metal oxide oxidations, the enthalpy (ΔH) is typically negative due to strong metal-oxygen bonds formed, while (ΔS) contributions from gas evolution further favor the reaction at high temperatures, as illustrated in Ellingham diagrams where oxidation lines slope downward. The of oxygen plays a critical role; lower pO₂ in limited-air calcination shifts the toward , preventing excessive exothermicity and allowing precise control over reaction extent. Unlike pure reactions, which are predominantly endothermic and focus on volatile release without oxygen involvement, oxidation-inclusive calcination incorporates exothermic contributions from steps, potentially reducing overall energy requirements and promoting of particles into cohesive structures. This exothermicity can elevate local temperatures, accelerating and layer formation, though careful air control is essential to avoid .

Industrial Applications

Lime and Cement Production

Calcination plays a pivotal role in production by thermally decomposing (CaCO₃) in s, typically at temperatures between 900°C and 1000°C, to produce quicklime (CaO) and (CO₂) via the reaction CaCO₃ → CaO + CO₂. This process yields approximately 56% quicklime by mass from pure , though actual yields depend on the mineral's purity, which must be at least 50% CaCO₃ for commercial viability. is quarried, crushed, screened, and preheated before entering the , where it descends countercurrent to rising hot gases from fuel , ensuring efficient and complete . Rotary kilns dominate lime production, accounting for about 90% of output due to their suitability for continuous, high-volume operations. These long, inclined cylinders rotate slowly while tumbles through, achieving uniform calcination; alternative vertical shaft kilns offer higher but lower throughput. Energy consumption for lime calcination generally ranges from 3 to 5 GJ per , primarily from fuels like or , with vertical kilns consuming less due to better heat recovery. , the primary product, serves as a in to remove impurities such as silica and during basic oxygen and processes. In cement production, calcination integrates with clinker formation by heating a mixture of and clay (or other siliceous materials) in rotary at around 1450°C, where CaCO₃ decomposes into CaO that reacts with silica, alumina, and iron oxides to form clinker nodules. The raw mix is prepared by grinding and blending, then preheated and partially calcined in a preheater tower before entering the for complete ; modern dry-process with precalciners enhance efficiency by performing up to 90% of calcination outside the main . Post-calcination, the hot clinker is rapidly cooled and ground with 3-5% to produce , which controls setting time. Energy use for clinker production in these preheater/precalciner typically ranges from 2.6 to 3.5 GJ per ton, reflecting optimizations in heat recovery and fuel efficiency. is essential for , binding aggregates in for buildings, roads, and .

Metallurgical Calcination

Metallurgical calcination serves as a critical pre-reduction step in the extraction and refining of metals from ores, involving the thermal treatment of ore concentrates to convert them into more reactive forms, typically oxides, by heating in the absence or limited presence of air. This process removes volatile impurities such as moisture, carbon dioxide, and organic matter, while enhancing the ore's reducibility for subsequent smelting or leaching operations. Performed at temperatures generally between 500°C and 800°C, calcination prevents fusion or sintering of the material, ensuring a porous structure that facilitates later reduction. In the Bayer process for aluminum production, calcination is applied to the precipitated aluminum hydroxide derived from digestion, dehydrating it to yield pure alumina (Al₂O₃) at temperatures around 1000–1300°C in rotary or fluidized beds, though initial preparation may involve lower-temperature calcination to remove bound water from . Similarly, , particularly carbonates or hydrates, undergoes calcination prior to direct reduction processes, where heating decomposes the and improves its for hydrogen or reduction, achieving over 60% CO₂ emission reductions compared to traditional methods. These steps prepare the by eliminating volatiles that could interfere with reduction efficiency. A specialized variant, dead roasting, employs low-temperature calcination (typically 600–800°C) to oxidize ores into oxides without producing molten SO₂ or causing material , allowing controlled removal as gas while retaining the metal oxide for further processing. For instance, in concentrates like (CuFeS₂), dead roasting converts sulfides to oxides such as CuO and Fe₂O₃ via oxidation reactions producing SO₂ gas, followed by reduction to metal. This method minimizes environmental impacts from emissions and prepares the calcine for hydrometallurgical or pyrometallurgical recovery. Prominent examples include production, where (ZnS) concentrates are dead roasted at 900–1000°C to form ZnO calcine, which is then leached in for electrolytic refining, achieving high-purity with efficient sulfur elimination. In laterite processing, calcination at 600–800°C dehydrates the and partially reduces iron oxides, concentrating for subsequent high-pressure leaching or , as demonstrated in beneficiation studies where post-calcination recovers up to 80% of values. These applications underscore calcination's role in improving reactivity and process economics in non-ferrous metallurgy.

Applications in Ceramics and Materials

Calcination plays a pivotal role in production by transforming raw clays into reactive materials suitable for advanced applications. In particular, the calcination of , with the \ce{Al2Si2O5(OH)4}, at temperatures between 600°C and 800°C yields , a highly reactive pozzolanic material used in supplementary cementitious formulations. This process involves the thermal dehydroxylation of , resulting in an amorphous that enhances the durability and strength of pozzolanic cements when blended with . 's stems from its ability to react with in cement hydration, forming additional gels that improve long-term performance in structures. In catalyst , calcination is essential for activating and achieving the desired crystalline structures in metal oxides and s. For instance, calcination of titanyl sulfate (\ce{TiOSO4}) through and subsequent heating produces (\ce{TiO2}) nanoparticles with high surface area, widely used as photocatalysts for . Similarly, in , calcination at around 500°C removes structure-directing agents from hydrothermal , stabilizing the microporous and enhancing catalytic selectivity in processes like cracking. These steps ensure phase purity and thermal stability, critical for industrial catalysts in and emission control applications. For , calcination provides high-temperature treatment to attain phase purity and structural integrity in advanced composites. In nanomaterials, such as \ce{LaFeO3}, calcination at 650–1150°C eliminates impurities and promotes single-phase formation, improving optoelectronic properties for and applications. For -based materials, thermal calcination of graphene oxide at 300–700°C reduces oxygen functional groups, yielding reduced graphene oxide with enhanced electrical conductivity and mechanical strength for use in devices. This controlled preserves nanoscale morphology while optimizing performance in composites. Post-2000 innovations have introduced microwave-assisted calcination to enhance in s and materials . This method uses electromagnetic heating to achieve uniform temperature distribution, reducing by up to 50% compared to conventional furnaces through faster reaction kinetics and minimized heat loss. In synthesis, calcination at lower overall temperatures accelerates phase transformations in clays and oxides, enabling scalable production of high-purity materials with reduced environmental impact. These developments, including hybrid -conventional systems, have been applied to and clay calcination, demonstrating improved yield and sustainability in nanomaterial fabrication.

References

  1. [1]
    [PDF] Primary Metal Production
    In current practice, calcination refers to any process where the material is heated to drive off volatile organics, CO2, chemically bound water, or similar ...
  2. [2]
    What is Calcination? - FEECO International
    Calcination is the process of heating a solid material in order to cause chemical separation of its components. The diversity of chemical separation lends ...Missing: authoritative | Show results with:authoritative
  3. [3]
    Calcination of Limestone - IspatGuru
    Feb 15, 2014 · Calcination or calcining is a thermal treatment process to bring about a thermal decomposition. The process takes place below the melting point of the product.
  4. [4]
    Calcination - an overview | ScienceDirect Topics
    Calcination is a process of heating a substance under controlled temperature and in a controlled environment. This process is known to improve the chroma, ...
  5. [5]
  6. [6]
  7. [7]
    Effect of heating rate on the kinetics of limestone calcination
    Nov 1, 2023 · This paper reports on the effect of heating rates between 0.1 °C/s and 205.0 °C/s on the thermodynamics and kinetics of limestone calcination in air.
  8. [8]
    What is the difference between calcining and roasting? Key insights ...
    24 mei 2025 · The main difference is the chemical change: roasting involves oxidation, while calcination involves decomposition. Key Points: What is the ...
  9. [9]
    Understanding the Calcination Process - Louisville Dryer Company
    May 13, 2025 · Types of Calcination Equipment · Rotary Calciners (Rotary Kiln Calciners) · Shaft Calciners · Fluidized Bed Calciners.
  10. [10]
    Rotary Kiln - an overview | ScienceDirect Topics
    The rotary kiln is used in many solid processes, including drying, incineration, heating, cooling, humidification, calcination and reduction. This widespread ...
  11. [11]
    CALCINE Definition & Meaning - Merriam-Webster
    Word History. Etymology. Verb. Middle English calcenen, from Medieval Latin calcinare, from Late Latin calcina lime, from Latin calc-, calx. First Known Use.
  12. [12]
    CALCINE definition in American English - Collins Dictionary
    Word origin. [1350–1400; ME ‹ ML calcināre to heat, orig. used by alchemists] ... Word origin. C14: from Medieval Latin calcināre to heat, from Latin calx lime.
  13. [13]
    calcination, n. meanings, etymology and more | Oxford English ...
    OED's earliest evidence for calcination is from around 1386, in the writing of Geoffrey Chaucer, poet and administrator. calcination is a borrowing from Latin. ...
  14. [14]
    Calcium - Etymology, Origin & Meaning
    Originating from Latin calx meaning "limestone," calcium, coined in 1808 by Sir Humphry Davy, is a metallic element essential in chemistry and biology.
  15. [15]
  16. [16]
    CALCINATION Definition & Meaning - Merriam-Webster
    The meaning of CALCINATION is the act or process of calcining : the state of being calcined.
  17. [17]
    [PDF] A Short History of the Use of Lime as a Building Material
    It was known that the Phoenicians made use of lime mortar and plaster. They were the first to develop the cocciopesto mortars that became widespread throughout ...
  18. [18]
    [PDF] a study of neolithic plaster materials - DSpace@MIT
    Specimens from Asikli Huyuk in central Anatolia (c. 7000 B.C.E.) and Pre- pottery Neolithic B Jericho (c. 6000 B.C.E.) are shown to be lime plasters containing ...
  19. [19]
    History of lime in mortar - Graymont
    The earliest documented use of lime as a construction material was approximately 4000 B.C. when it was used in Egypt for plastering the pyramids(ref. ii). The ...
  20. [20]
    Historic Concrete Science: Opus Caementicium to “Natural Cements”
    Feb 3, 2023 · Romans advised using pure limestone for the production of lime (de Architectura 2.6.3) to further activate reactions with pozzolanic aggregates ...
  21. [21]
    Glass from Islamic Lands - The Metropolitan Museum of Art
    Oct 1, 2002 · At the time of the Arab conquest in the seventh century A.D., glassmaking had flourished in Egypt and western Asia for more than two millennia ...
  22. [22]
    What Have We Learned from the Recent Historiography of Alchemy?
    Alchemy has a great deal to teach historians of science. The lessons range from the nature and influence of Aristotelian natural philosophy and scholasticism ...
  23. [23]
    Jabir Ibn Hayyan and Islamic Golden Era Alchemists - About Islam
    Feb 15, 2017 · Jabir Ibn Hayyan and Islamic Golden Era Alchemists. Makers of ... calcination and crystallization. At least two treatises attributed ...
  24. [24]
    Coelum philosophorum by Paracelsus
    ### Summary of Calcination in Paracelsus' *Coelum Philosophorum*
  25. [25]
    Alchemical Glossary: The Chymistry of Isaac Newton Project
    Calcination: A chemical operation involving roasting a substance in an open dish over a hot fire. The product of calcination is referred to as a calx or ...
  26. [26]
    Antoine Laurent Lavoisier (1743-1794) - Le Moyne
    Memoir on the Nature of the Principle which Combines with Metals during their Calcination and which Increases their Weight[1] ... mercury, and I succeeded in ...
  27. [27]
    On the decomposition of carbonate of lime by heat
    ON THE DECOMPOSITION OF CARBONATE OF LIME BY HEAT. M. Gay-Lussac observes, that it has long been supposed that the calcination of limestone is accelerated by ...
  28. [28]
    [PDF] A History of Chemistry
    This short history of chemistry is dedicated essentially to a description of the evolution of ideas in chemistry during the nineteenth century. The ...
  29. [29]
    [PDF] The origin and status of the Arrhenius equation
    The classic paper (3) which is the main focus of attention here, appeared in 1889 under the title, "Uber die Reaktionsgeschwindigkeit bei der Inuersion uon ...
  30. [30]
    History of the Periodic Table - Chemistry LibreTexts
    Mar 8, 2023 · Dmitri Ivanovich Mendeleev proposed the periodic law behind his periodic table compiling. ... Formula of oxide, Eb2O3, Sc2O3. Density of oxide ...
  31. [31]
    Kinetics of endothermic decomposition reactions. 2. Effects of the ...
    ... Thermal Decomposition of Inorganic Carbonates in an Inert Gas Atmosphere: ZnCO3 and CaCO3. ... A new kinetic model of calcite thermal decomposition.
  32. [32]
    (PDF) Thermal Decomposition of Calcite: Mechanisms of Formation ...
    A novel model for the thermal decomposition of calcite that explains how decarbonation occurs at the atomic scale via a topotactic mechanism.
  33. [33]
    The Thermal Stability of the Nitrates and Carbonates
    Jun 30, 2023 · Enthalpy changes. The enthalpy changes for the decomposition of the various carbonates indicate that the reactions are strongly endothermic ...
  34. [34]
    An Experimental Study of the Decomposition and Carbonation of ...
    Oct 16, 2025 · Magnesium carbonate (MgCO3) has a turning temperature of 396 °C, a theoretical potential to store 1387 J/g and is low cost (~GBP 400/1000 kg).
  35. [35]
    Low‐temperature synthesis of α‐Al2O3 powder aided by ball milling ...
    Jun 15, 2023 · α-Al2O3 is usually prepared by calcination of aluminum hydroxide (Al(OH)3) and boehmite (AlOOH), which are low costs for industrial production.
  36. [36]
    Kinetic modeling in a fixed-bed reactor - ScienceDirect
    The activation energy of CaCO3 decomposition reaction was found equal to 216 ± 29 kJ/mol derived by the UCM and 181 ± 9 for CGM fitting, according to the ...
  37. [37]
    Limestone Calcination Kinetics in Microfluidized Bed ... - MDPI
    Dec 17, 2022 · In addition, due to several factors affecting the calcination of CaCO3, such as particle size, temperature, CO2 concentration, and experimental ...
  38. [38]
    Roasting - an overview | ScienceDirect Topics
    Roasting is a process that drives out unwanted sulfur and carbon from sulfide/carbonate ore in an oxidizing environment, leaving an oxide.<|control11|><|separator|>
  39. [39]
    Mechanisms in oxidation and sintering of magnetite iron ore green ...
    Apr 9, 2008 · When a pellet starts to oxidize, a shell of hematite is formed around the pellet while the core still is magnetite. Dilatation curves were ...Missing: redox | Show results with:redox
  40. [40]
    [PDF] Ellingham Diagrams - MIT
    The Gibbs free energy (∆G) of a reaction is a measure of the thermodynamic driving force that makes a reaction occur. A negative value for ∆G indicates ...Missing: calcination | Show results with:calcination
  41. [41]
    Heating Limestone: A Major CO₂ Culprit in Construction - USGS.gov
    Aug 20, 2024 · When limestone (CaCO₃) is heated in a kiln, it undergoes a chemical reaction called calcination. This process produces calcium oxide (CaO) and ...
  42. [42]
    [PDF] 11.17 Lime Manufacturing 11.17.1 Process Description
    The limestone is charged at the top and is calcined as it descends slowly to discharge at the bottom of the kiln. A primary advantage of vertical kilns over ...
  43. [43]
    How Lime is Made - The National Lime Association
    Calcining – the kiln fuel is burned in the preheated air from the cooling zone and, as the limestone moves down the kiln, the heat turns the limestone into ...
  44. [44]
    [PDF] Final Technical Report - OSTI.GOV
    Feb 10, 2012 · Therefore, the energy consumption for the lime industry is one fifth the consumption for the cement industry. Secondly, Microwave Assist.
  45. [45]
    Applications - MPA Lime
    Lime is principally used for removing impurities in both the basic oxygen process and electric arc furnaces. Dolime is also used in steelmaking as a repair ...
  46. [46]
    [PDF] Technical Support Document for Process Emissions from Cement
    Jan 28, 2009 · formed clinker and from calcination of carbonates that formed clinker kiln dust (CKD). Additional process-related CO2 emissions may arise ...
  47. [47]
    [PDF] Energy Efficiency Improvement and Cost Saving Opportunities for ...
    Aug 8, 2013 · In a dry kiln cement plant, electricity use is typically broken down as follows (ECRA, 2009): 38% cement grinding, 24% raw material grinding, ...
  48. [48]
    Bayer Process - an overview | ScienceDirect Topics
    "The Bayer process is defined as a method for alumina recovery that involves digesting crushed bauxite in strong sodium hydroxide solutions at high ...
  49. [49]
    Iron Carbonate Beneficiation Through Reductive Calcination - NIH
    Direct iron carbonate reduction through reductive calcination in a hydrogen atmosphere is a high‐potential candidate for environmentally benign pig iron ...
  50. [50]
    Roasting of Sulfide Minerals - Wiley Online Library
    By and large, roasting of zinc sulfide is carried out at and above 800. C to produce zinc oxide. This is known as dead roasting. The rate of oxidation is ...
  51. [51]
    Roasting Copper Sulfides Chemistry - 911Metallurgist
    Feb 4, 2021 · In this process, the dead roasting of sulfide is followed by the reduction of oxide to metallic copper by a reducing agent such as hydrogen.
  52. [52]
    Calcination of zinc sulfide concentrates - Google Patents
    Because Zinc sulfide, which is the most common zinc mineral found in zinciferous ores, cannot be reduced directly, these concentrates are roasted or calcined to ...
  53. [53]
    Calcination of low-grade laterite for concentration of Ni by magnetic ...
    In this study, a low-grade nickeliferous laterite ore was first calcinated and then processed by using a wet magnetic separator in order to recover nickel.
  54. [54]
    [PDF] METAKAOLIN A Highly Reactive Natural Pozzolan
    – Metakaolins are typically used at levels between 8% to 12% replacement of cement. – The use of metakaolins can increase the expected life of concrete in ...
  55. [55]
    [PDF] Mineralogical Properties of Kaolin and Metakaolin from Selected ...
    Results from the DTA showed that kaolin can be calcined to get metakaolin at temperature ranging from 700°C - 850°C.
  56. [56]
    Characterization of net-zero pozzolanic potential of thermally ... - NIH
    Nov 2, 2023 · Metakaolin is produced from calcination of Kaolin clay, one of ... Hydration and characteristics of metakaolin pozzolanic cement pastes.
  57. [57]
    [PDF] Clay Calcination Technology: State-of-the-Art Review by the RILEM ...
    While calcined clays, mainly metakaolin, have been used as SCMs and in the production of cement in many places around the world [8, 11-16], advances in ...
  58. [58]
    Preparation of Calcined kaolinite/TiO 2 composites by titanyl sulfate ...
    Jul 15, 2025 · Calcined kaolinite (CK)/TiO2 composites were synthesized through hydrolysis precipitation of titanyl sulfate (TiOSO4) and calcination.
  59. [59]
    Sol-Gel Synthesis of TiO 2 from TiOSO 4 (Part 2) - MDPI
    The sol-gel process was used to create titanium dioxide (TiO2) nanoparticles, a nanocrystalline semiconductor. How several synthesis factors, ...
  60. [60]
    Methodological review of zeolite synthesis from industrial waste and ...
    The respective H-form of the zeolites were obtained via ion exchange with 1 M NH4NO3 solution and calcination at 500 °C for 4 h. Notably, zeolite seeds ...
  61. [61]
    New progress in zeolite synthesis and catalysis - Oxford Academic
    Benefiting from the large interlayer space, the removal of the OSDA by calcination created a meso/microporous structure, which was effective in depressing the ...Abstract · SYNTHESIS OF ZEOLITES... · SYNTHESIS AND CATALYTIC...
  62. [62]
    Perovskite oxides as a new family of tunable CO 2 sorbents
    Feb 26, 2025 · It was therefore determined that the most suitable calcination temperature is 650 °C, as it allows us to achieve both phase purity and a high ...
  63. [63]
    Effect of Calcination Temperature on the Stability of the Perovskite ...
    These findings suggest that calcination at 1150 °C is the minimum required temperature to achieve phase purity, whereas calcination at 1250 °C enhances the ...
  64. [64]
    Effect of calcination temperature on characteristics of reduced ...
    Coconut shell is one of the carbon sources in nature that can be converted into reduced Graphene Oxide (rGO). There are several methods of rGO synthesis ...
  65. [65]
    Calcination of reduced graphene oxide decorated TiO2 composites ...
    Reduced graphene oxide (RGO)–TiO 2 composites were synthesized hydrothermally. Then the samples were calcined at 450 °C for 30 min under air atmosphere to ...
  66. [66]
    Energy Efficiency in the Microwave-Assisted Solid-State Synthesis of ...
    The results show that, when using a properly designed microwave applicator, the specific energy consumption can be significantly lowered compared to ...Missing: post- | Show results with:post-
  67. [67]
    Advanced Ceramic Materials Sintered by Microwave Technology
    Microwave sintering represents an interesting opportunity at consolidating advanced ceramic materials with a reduced processing time and energy consumption.2. Microwave Sintering... · 2.3. Theoretical Aspects In... · 2.4. Microwave Systems For...
  68. [68]
    Microwave-assisted ceramic calcination and sintering process
    ... Microwave-assisted ceramic calcination and sintering process: an alternative heat treatment to save energy. Alexander Lourenço Pessoa, Rodolfo Izquierdo ...
  69. [69]
    [PDF] Microwave-assisted calcination of spodumene for efficient, low-cost ...
    To utilize the advantages of microwave-assisted calcination over conventional heating, this work studied the ef- fect of microwave power on phase ...