Fact-checked by Grok 2 weeks ago

Chemical decomposition

Chemical decomposition is a type of in which a single compound breaks down into two or more simpler substances, such as elements or other compounds, by the breaking of chemical bonds. The general equation for such reactions is AB → A + B, where AB represents the original compound and A and B are the products. These reactions typically require an input of to initiate and proceed, often in the form of , , , or a , as the bonds in the reactant are stronger than those formed in the products. Decomposition reactions are classified into several types based on the energy source or conditions: thermal decomposition, which uses heat to break down compounds like calcium carbonate (CaCO₃) into calcium oxide (CaO) and carbon dioxide (CO₂); electrolytic decomposition, or electrolysis, where electrical energy splits molecules such as water (H₂O) into hydrogen (H₂) and oxygen (O₂); photodecomposition, triggered by light, as seen in the breakdown of hydrogen peroxide (H₂O₂) into water (H₂O) and oxygen (O₂) under sunlight; and catalytic decomposition, accelerated by a catalyst without being consumed, for instance, the enzyme catalase in yeast speeding up H₂O₂ decomposition. In everyday life and industry, chemical decomposition plays a crucial role in processes like , where sodium hydrogen carbonate (NaHCO₃) decomposes at around 80°C to release CO₂ for leavening dough; bleaching and disinfection using H₂O₂, which decomposes to produce nascent oxygen; and production, such as powered by renewable sources to generate fuel. These reactions are fundamental in chemistry, enabling the synthesis of materials, waste breakdown, and , though they must often be controlled to prevent hazards like explosions from rapid gas release.

Fundamentals

Definition

Chemical decomposition is a type of in which a single compound breaks down into two or more simpler products, such as elements or other compounds. This process fundamentally involves the cleavage of chemical bonds within the reactant molecules, leading to the formation of new, less complex substances. Decomposition reactions exhibit varied thermodynamic characteristics, being either endothermic or exothermic depending on the specific compound and conditions involved. While some decompositions release , many require an external input—such as , , or —to initiate and sustain the bond-breaking process, overcoming the barrier. The earliest observations of chemical decomposition trace back to ancient , where processes like —heating substances to produce residues or gases—were noted as breakdowns of materials. These empirical insights were formalized in 18th-century chemistry by , who demonstrated the in decomposition reactions through precise experiments, such as the thermal breakdown of mercury oxide, establishing a quantitative foundation for understanding such changes.

General Reaction Equation

The general reaction equation for chemical decomposition represents the breakdown of a single compound into two or more simpler substances, typically or other s. This is symbolically expressed as: AB \rightarrow A + B where AB denotes the reactant , and A and B are the products, which may be or simpler s. Variations of this equation account for decompositions yielding more than two products, such as: ABC \rightarrow A + B + C In all cases, the equation must adhere to the law of , requiring balanced where the total number of atoms of each element remains constant on both sides of the reaction. Decomposition reactions are the reverse of or reactions, which combine simpler substances into a more complex compound (e.g., A + B \rightarrow AB); thus, the equation inverts this process by separating the bonded entities. The energy aspect of decomposition is denoted by the change, \Delta H, which indicates whether the reaction is endothermic (\Delta H > 0) or exothermic (\Delta H < 0). Most decomposition reactions are endothermic because they involve breaking chemical bonds, which requires energy input, and typically fewer new bonds are formed in the products compared to those broken in the reactant; the net \Delta H reflects the difference in bond energies, where bond breaking absorbs energy and any bond formation releases it.

Types

Thermal Decomposition

Thermal decomposition refers to a chemical reaction in which a compound breaks down into simpler substances upon the application of heat, serving as a key process in various chemical transformations. This type of decomposition occurs when thermal energy supplies the necessary activation to overcome the energy barrier for bond cleavage, applicable to solids, liquids, or gases. The mechanism involves the absorption of heat, which increases the kinetic energy of molecules or atoms, facilitating the rupture of chemical bonds and leading to the formation of new, more stable products. Several factors influence the rate and extent of thermal decomposition, including temperature thresholds specific to each compound's thermal stability, the presence of catalysts that lower the activation energy, and adherence to the for reaction kinetics. Thermal stability varies among compounds; for instance, carbonates generally decompose at elevated temperatures to yield metal oxides and carbon dioxide, with stability increasing down group 2 of the due to decreasing polarizing power of the cation. Catalysts, such as metal oxides or acids, can reduce the activation energy required, thereby accelerating the process at lower temperatures. The decomposition rate k is described by the : k = A e^{-E_a / RT} where A is the pre-exponential factor, E_a is the activation energy, R is the gas constant, and T is the absolute temperature; this empirical relation has been justified for solid-state thermal decompositions through statistical mechanics derivations./Descriptive_Chemistry/Elements_Organized_by_Block/1_s-Block_Elements/Group_2_Elements:_The_Alkaline_Earth_Metals/1Group_2:_Chemical_Reactions_of_Alkali_Earth_Metals/The_Thermal_Stability_of_the_Nitrates_and_Carbonates) A prominent application of thermal decomposition is in calcination processes, where nitrates or carbonates of metals are heated to produce pure metal oxides, often used in ceramics and catalyst preparation. This method leverages controlled heating to drive the endothermic decomposition, resulting in the release of gases like nitrogen oxides or carbon dioxide, while forming stable oxide residues. Such processes highlight the practical utility of thermal decomposition in achieving high-purity materials through precise temperature control.

Electrolytic Decomposition

Electrolytic decomposition, commonly referred to as , involves the use of electric current to drive the non-spontaneous decomposition of an into its constituent elements or simpler compounds at the electrodes of an . In this process, the electrolyte—a solution or molten salt that conducts electricity via —facilitates ion migration between the (negative electrode, site of ) and (positive electrode, site of ). At the cathode, cations accept electrons to form neutral atoms or molecules, while at the anode, anions donate electrons, leading to the release of gases or deposition of materials. This mechanism relies on the application of an external voltage to overcome the energy barrier of the decomposition reaction, converting electrical energy into chemical change. A representative example is the electrolysis of water in an aqueous electrolyte, where the overall decomposition follows the balanced equation: $2H_2O(l) \rightarrow 2H_2(g) + O_2(g) The cathode reaction is $2H_2O + 2e^- \rightarrow H_2 + 2OH^-, producing hydrogen gas, and the anode reaction is $4OH^- \rightarrow O_2 + 2H_2O + 4e^-, yielding oxygen gas. Theoretically, this process requires a minimum cell voltage of 1.23 V at standard conditions (298 K, 1 atm), derived from the standard Gibbs free energy change of the reaction (\Delta G^\circ = -nFE, where n = 2 moles of electrons). In practice, higher voltages (typically 1.5–2.0 V) are needed due to kinetic barriers. The extent of decomposition is quantitatively described by , formulated by in 1832–1834. The first law states that the mass m of a substance decomposed or deposited at an electrode is directly proportional to the total electric charge Q passed through the electrolyte (m \propto Q, where Q = I \cdot t, with I as current and t as time). The second law asserts that for a fixed charge, the masses of different substances produced are proportional to their chemical equivalent weights (molar mass divided by the number of electrons transferred per formula unit). These laws enable prediction of yields; for instance, one of charge (approximately 96,485 C, the charge of one mole of electrons) liberates one equivalent of substance, such as 1 g-equivalent of hydrogen (1 g of H₂) from . Practical efficiency of electrolytic decomposition is limited by overpotential—the additional voltage required beyond the theoretical value to achieve appreciable reaction rates due to slow electron transfer, gas bubble formation, or concentration gradients—and by the properties of electrode materials. Overpotentials can exceed 0.3–0.5 V at typical current densities (0.1–1 A/cm²), reducing energy efficiency to 60–80% in many systems. Electrode choice is critical; inert materials like lower overpotential by providing active sites for reactions, while cost-effective alternatives such as or carbon-based electrodes are used in alkaline media to balance performance and durability. This technique has historical significance in element discovery; in 1807, Humphry Davy isolated metallic sodium for the first time by electrolyzing molten sodium hydroxide (NaOH), using a voltaic pile to pass current through the electrolyte, with mercury as a cathode to capture the reactive metal. The cathode reaction was Na^+ + e^- \rightarrow Na, demonstrating electrolysis's role in decomposing stable compounds to reveal new elements.

Photodecomposition

Photodecomposition, also known as photodissociation, is a type of chemical decomposition triggered by the absorption of light, particularly in the ultraviolet (UV) range, where photons provide sufficient energy to excite electrons and initiate bond breaking in molecules. This process adheres to the Grotthuss-Draper law, the first law of photochemistry, which states that only light absorbed by a substance can induce a photochemical reaction; unabsorbed wavelengths pass through without effect. The mechanism begins with the absorption of a photon, promoting an electron from the ground state to an excited electronic state, often leading to homolytic or heterolytic bond dissociation due to the resulting instability. This excitation occurs rapidly, typically on the femtosecond timescale, following the , where nuclear coordinates remain unchanged during the vertical transition. A classic example is the photodecomposition of silver chloride (AgCl), a silver halide commonly used in photography, which occurs under UV light according to the reaction: $2\text{AgCl} \rightarrow 2\text{Ag} + \text{Cl}_2 Here, UV photons (wavelengths around 200-400 nm) are absorbed by the AgCl lattice, exciting electrons and generating electron-hole pairs that reduce Ag⁺ to metallic silver atoms while oxidizing Cl⁻ to chlorine, forming a latent image on photographic film. The efficiency of this process is quantified by the quantum yield (Φ), defined as the number of molecules decomposed per photon absorbed, which for primary photodissociation events is ideally 1 but can vary due to secondary reactions or recombination; in AgCl, Φ is near unity for initial dissociation but decreases with aggregation of silver clusters. Wavelength plays a critical role, as shorter UV wavelengths (e.g., below 300 nm) deliver higher energy (E = hc/λ) to surpass bond dissociation energies, typically 200-500 kJ/mol for common covalent bonds, making UV more effective than visible light for initiating decomposition in many compounds. Photodecomposition underlies practical phenomena such as the fading of dyes, where UV exposure leads to chromophore bond cleavage and color loss in textiles or inks, often via radical intermediates that propagate degradation. In photography, the selective decomposition of silver halides by light forms the basis of image development, with the released halogens (e.g., Cl₂ from AgCl) contributing to the sensitivity of the medium to specific wavelengths. These quantum-driven processes highlight photodecomposition's role in both destructive fading and controlled applications like imaging technology.

Other Types

Catalytic decomposition involves the use of a catalyst to accelerate the breakdown of a compound by lowering the activation energy required for the reaction, without the catalyst being consumed in the process. This type of decomposition facilitates reactions that might otherwise proceed too slowly at ambient conditions. A classic example is the decomposition of hydrogen peroxide (\ce{H2O2}) into water and oxygen gas, catalyzed by manganese dioxide (\ce{MnO2}):
\ce{2H2O2 ->[MnO2] 2H2O + O2}
This reaction demonstrates how the catalyst provides an alternative pathway with reduced energy barriers.
Hydrolytic decomposition, also known as hydrolysis, is a process where water molecules react with a compound to break its chemical bonds, resulting in the formation of simpler products. In this reaction, water acts as a reactant, often leading to the cleavage of salts into their constituent acids and bases or the breakdown of larger molecules like esters and amides. For instance, certain salts undergo hydrolysis in aqueous solutions to produce acidic or basic species, depending on the nature of the ions involved. This mode is particularly relevant in aqueous environments and biological systems. Other decomposition modes include explosive processes, which serve as auxiliary triggers often overlapping with primary types like thermal or electrolytic. Explosive decomposition occurs rapidly, releasing significant energy and gases, as seen in , where the compound breaks down violently upon initiation, producing , , , and oxygen. These mechanisms highlight how external factors can initiate or enhance chemical breakdown beyond conventional energy inputs.

Examples

Inorganic Compounds

Inorganic compounds provide numerous examples of chemical decomposition, often serving as key processes in industrial applications and educational demonstrations. These reactions typically involve the breakdown of stable salts, oxides, or hydrides into simpler components, such as metals, gases, or other inorganic products, and can be triggered by heat, electricity, or catalysts. A classic instance of is the breakdown of , found in , which occurs at high temperatures to produce (quicklime) and gas, as represented by the equation: \ce{CaCO3 -> CaO + CO2} This process, known as calcination, is central to lime production, where limestone is heated in kilns to around 900–1000°C, yielding quicklime used in construction, steelmaking, and water treatment. In electrolytic decomposition, water undergoes splitting into hydrogen and oxygen gases when an electric current is passed through it, following the overall reaction: \ce{2H2O -> 2H2 + O2} This requires energy input to overcome the stability of the H-O bonds, producing at the and oxygen at the , and is a foundational method for generating these gases in and settings. The decomposition of into and oxygen gas illustrates a reaction that proceeds slowly at but is greatly accelerated by catalysts such as or enzymes, according to: \ce{2H2O2 -> 2H2O + O2} This exothermic process releases oxygen rapidly in the presence of a catalyst, making it useful for demonstrating catalysis and in applications like disinfection and propulsion systems. A visually dramatic example of thermal decomposition is the reaction of ammonium dichromate, often performed as the "volcanic fountain" demonstration, where the orange crystals ignite to produce a fountain-like eruption of green chromium(III) oxide ash, nitrogen gas, and water vapor, as shown in: \ce{(NH4)2Cr2O7 -> Cr2O3 + N2 + 4H2O} This self-sustaining exothermic reaction, initiated at around 180–225°C, highlights the vigorous nature of certain inorganic decompositions and is commonly used in chemistry education to illustrate redox processes.

Organic Compounds

Organic compounds, characterized by their carbon skeletons and diverse functional groups, undergo chemical decomposition through various mechanisms, often yielding complex mixtures of products due to multiple reaction pathways involving C-C, C-H, and heteroatom bonds. This complexity arises from the structural intricacy of organic molecules, leading to fragmented chains, volatile organics, and residues that differ markedly from the more predictable outcomes in inorganic decompositions. A prominent example is the of glucose (C₆H₁₂O₆) during , where the breaks down at elevated temperatures in the absence of oxygen to produce (a carbon-rich solid residue) and a variety of gases such as , CO₂, and H₂O, alongside condensable vapors. This process highlights the fragmentation of the ring and side chains, resulting in a heterogeneous product distribution that includes bio-oil precursors. Photodecomposition of key pigment in , occurs under light exposure and breaks the molecule into simpler derivatives, disrupting its conjugated and leading to loss of color and . This light-induced exemplifies how electromagnetic energy can initiate oxidative in sensitive chromophores. Hydrolytic is illustrated by the of esters with , typically under acidic or basic conditions, following the general RCOOR' + H₂O → RCOOH + R'OH, where the ester linkage is cleaved to yield a and an . This is fundamental in and , emphasizing the role of in disrupting polar bonds. The decomposition of synthetic organic polymers, such as (PVC), during thermal or environmental exposure releases (HCl) gas through dehydrochlorination, contributing to and formation. This hazardous byproduct underscores the environmental challenges posed by degradation, where chain scission generates both volatile acids and conjugated polyene residues.

Biological Contexts

In biological systems, chemical decomposition plays a pivotal role in metabolic processes and the breakdown of , enabling , , and the maintenance of cellular and . One fundamental example is the enzymatic of (ATP), where ATP is decomposed into (ADP) and inorganic phosphate (P_i) to release for cellular activities such as and . This reaction, catalyzed by enzymes like ATPases, is exergonic and provides approximately 30.5 kJ/mol of under standard physiological conditions, powering nearly all energy-requiring processes in living organisms. Another key instance of biological involves the of proteins by proteases, which cleave bonds to break down polypeptides into smaller and free , facilitating , , and regulatory signaling. Proteases, such as and in the digestive system or ubiquitin-proteasome systems in cells, perform this hydrolytic with high specificity, ensuring the of for new or energy production. This process is essential for nutrient absorption in the gut and the degradation of damaged proteins to prevent cellular dysfunction. Microbial biodegradation represents a broader form of decomposition in living systems, where bacteria and fungi enzymatically break down complex organic polymers like cellulose into simpler compounds, ultimately yielding carbon dioxide (CO_2) and water (H_2O) under aerobic conditions. Cellulases produced by microbes such as *Trichoderma reesei* hydrolyze the \beta$-1,4-glycosidic bonds in cellulose, enabling the decomposition of plant cell walls in soils and ruminant guts, which supports carbon cycling and biomass conversion. This process exemplifies hydrolytic decomposition adapted for ecological roles. Decomposition in biological contexts also drives nutrient cycling, as microbes and detritivores break down dead organisms and organic detritus, releasing essential elements like , , and carbon back into the for uptake by and other life forms. For instance, saprotrophic fungi and decompose leaf litter, mineralizing organic compounds into inorganic ions that replenish and sustain primary productivity in ecosystems. This recycling prevents depletion and maintains the balance of biogeochemical cycles.

Applications

Industrial Processes

One prominent industrial application of chemical decomposition is the thermal decomposition of to produce quicklime, a critical step in . In this process, (CaCO₃) is heated in large-scale at temperatures around 900–1000°C, undergoing to yield (CaO) and (CO₂) according to the reaction CaCO₃ → CaO + CO₂. The resulting quicklime is then slaked with to form hydrated lime or directly used in production, where it reacts with silicates to create the binding agent in . This decomposition is endothermic and requires significant energy input, typically from fossil fuels, making it a major source of industrial CO₂ emissions. Another key industrial process involving decomposition is the electrolytic decomposition of to produce gas, which serves as a clean fuel for applications like fuel cells in transportation and power generation. electrolysis occurs in electrolyzers, where an passes through (often in an alkaline or setup), splitting H₂O into (H₂) at the and oxygen (O₂) at the . This method is scalable for industrial , with advancements in efficiency driven by renewable electricity sources to minimize carbon footprints. Globally, accounts for a growing share of output, supporting the transition to hydrogen-based systems. In , catalytic cracking represents a vital process for converting heavy hydrocarbons into lighter, more valuable products. Long-chain hydrocarbons from crude oil or residues are heated in the presence of catalysts like zeolites at 450–550°C and moderate , breaking C-C bonds to produce alkenes, , and shorter alkanes such as those used in and feedstocks, exemplified by petroleum → alkenes + H₂. Fluidized-bed reactors enable continuous operation in refineries, enhancing yield and selectivity compared to thermal cracking. This process is essential for meeting demand for transportation fuels and olefins in plastics production. The scale of these decomposition-based industries underscores their economic importance; for instance, annual global production of quicklime exceeds 400 million tons, almost entirely dependent on controlled of to support and sectors.

Analytical Techniques

Analytical techniques leverage chemical decomposition to identify and quantify components in samples by exploiting the specific conditions under which break down, such as , voltage, or exposure, allowing for precise measurement of decomposition products or mass changes. Gravimetric analysis via thermal decomposition is a classical method for determining the content of volatile compounds, particularly carbonates, where the sample is heated to induce decomposition and release carbon dioxide (CO₂), with the mass loss corresponding directly to the carbonate amount. For instance, in the analysis of limestone or soil samples, calcium carbonate (CaCO₃) decomposes according to the reaction CaCO₃ → CaO + CO₂ at temperatures around 800–900°C, enabling quantification by the difference in sample weight before and after heating in a controlled furnace. This technique provides high accuracy for inorganic analysis, with detection limits often below 0.1% by mass, and is widely used in geochemistry and materials testing. Thermogravimetric analysis (TGA) extends this principle by continuously monitoring mass changes as a function of , revealing decomposition profiles that characterize , purity, and . In , a sample is heated in a (e.g., inert or oxidative air), and the instrument records events corresponding to , , or oxidation steps; for example, polymers like exhibit distinct decomposition onset temperatures around 400–500°C, allowing identification of additives or contaminants. Coupled with techniques like , provides kinetic data on energies for , typically in the range of 100–300 kJ/mol for organic , aiding in and applications. Electrolytic decomposition plays a key role in voltammetric methods for detecting trace metal ions, where an applied potential drives the reduction and deposition of metals onto an electrode, followed by oxidative stripping that quantifies the ions through current peaks. Anodic stripping voltammetry (ASV), a prominent example, involves preconcentration of metal ions like lead (Pb²⁺) or (Cd²⁺) at negative potentials (e.g., -0.8 V vs. Ag/AgCl), then scanning to more positive potentials to induce decomposition and measure stripping currents proportional to concentration, achieving detection limits as low as . This approach is particularly valuable for environmental and water quality monitoring due to its sensitivity and portability.

Environmental Impact

Chemical decomposition processes can have profound environmental consequences, particularly through the release of greenhouse gases and other pollutants that contribute to and atmospheric degradation. In the cement industry, thermal decomposition of (CaCO₃) during produces (CaO) and releases (CO₂), accounting for approximately 60-66% of the sector's total CO₂ emissions. production as a whole is responsible for about 8% of global anthropogenic CO₂ emissions, exacerbating . Photodecomposition of chlorofluorocarbons (CFCs) in the , such as the breakdown of CFCl₃ into (Cl) radicals and CFCl₂ upon exposure to ultraviolet radiation, plays a critical role in . These radicals catalyze the destruction of (O₃) molecules, leading to the formation of the Antarctic ozone hole and increased ultraviolet radiation reaching Earth's surface. This process, first detailed in photochemical models, has resulted in significant stratospheric loss since the mid-20th century due to widespread CFC use in refrigerants and aerosols. In waste management, controlled thermal decomposition via incineration reduces solid waste volume by up to 90%, but it generates emissions of toxic pollutants including dioxins, furans, heavy metals like mercury, and particulate matter that harm air quality and human health. These incineration byproducts persist in the environment, contributing to soil and water contamination near facilities, and life-cycle assessments indicate that such processes can increase overall toxic releases compared to landfilling in some scenarios. Anaerobic decomposition of organic waste in landfills produces (CH₄), a potent , through microbial breakdown in oxygen-limited conditions, with landfill gas typically comprising about 50% . Globally, landfills account for roughly 11% of anthropogenic , which in turn represent about 16-20% of total human-induced when measured in CO₂-equivalent terms. This release intensifies forcing, as its is 28-34 times that of CO₂ over a 100-year period.

Comparison with Synthesis

Chemical decomposition reactions represent the breakdown of a single compound into two or more simpler substances, typically requiring energy input as they are endothermic processes that involve breaking chemical bonds. In contrast, synthesis reactions combine two or more reactants—often elements or simpler compounds—to form a more complex product, frequently releasing energy as they are exothermic due to the formation of new bonds. This fundamental opposition highlights their roles as inverse processes in chemical transformations, with decomposition simplifying molecular structures while synthesis increases complexity. A key difference lies in their directionality regarding molecular and thermodynamic favorability: yields simpler products from complex and often drives an increase in due to greater molecular , such as when solids or liquids produce gases. , conversely, builds from elemental or simple forms, with typically decreasing as ordered structures form, though the exothermicity often makes these reactions spontaneous under standard conditions. In reaction networks, serves as a precursor step for subsequent ; for instance, in the , thermal cracking decomposes larger hydrocarbons into monomers like , which are then polymerized to form plastics such as . Regarding equilibria, applies inversely between the two: in synthesis reactions where the number of gas moles decreases, increasing shifts the toward products, whereas in reactions producing more gas moles, favors the reactant side to counteract the volume increase. This contrast underscores how external stresses like or can selectively promote one direction over the other in reversible systems, aiding control in .

Reversible Decomposition

Reversible decomposition refers to chemical processes where a breaks down into simpler substances, but the can proceed in the reverse to reform the original under altered conditions, establishing a . This bidirectional nature distinguishes it from irreversible decompositions, allowing the system to respond to external changes. A classic example is the thermal of , which sublimes upon heating above approximately 340°C, dissociating into and gases, but recombines upon cooling. The is represented as: \mathrm{NH_4Cl(s) \rightleftharpoons NH_3(g) + HCl(g)} This equilibrium demonstrates thermal reversibility, where the forward decomposition is endothermic and the reverse combination is exothermic. Another prominent case involves phosphorus pentachloride, which decomposes in the gas phase into phosphorus trichloride and chlorine, with the equilibrium shifting toward greater decomposition at higher temperatures due to the endothermic nature of the forward reaction. The process is depicted as: \mathrm{PCl_5(g) \rightleftharpoons PCl_3(g) + Cl_2(g)} The equilibrium constant increases with temperature; for example, K_p ≈ 38 at 600 K and is larger at 700 K. The direction of reversible decompositions is governed by Le Chatelier's principle, which predicts that the system adjusts to counteract changes in conditions such as temperature or pressure. For instance, increasing temperature favors the endothermic decomposition direction, while elevating pressure shifts the equilibrium toward the side with fewer gas moles, promoting recombination. In the Haber-Bosch process for ammonia synthesis, reversible decomposition of ammonia ($2\mathrm{NH_3(g)} \rightleftharpoons \mathrm{N_2(g)} + 3\mathrm{H_2(g)}) occurs as a side reaction within the equilibrium, and process controls like temperature and pressure adjustments leverage Le Chatelier's principle to minimize decomposition and optimize forward synthesis yield.

Decomposition in Equilibria

In chemical equilibria, refers to the reversible breakdown of a into simpler components, where the forward reaction is balanced by the reverse association, resulting in a dynamic with constant concentrations of over time. This is governed by the , which quantifies the extent of under given conditions. Unlike irreversible decompositions, these equilibria allow systems to respond to perturbations by shifting according to , maintaining overall stability. A key example occurs in the dissociation equilibria of weak acids in aqueous solutions, where the acid partially decomposes into its conjugate base and a proton: \text{HA} \rightleftharpoons \text{H}^+ + \text{A}^- The equilibrium constant, known as the K_a, measures the degree of this decomposition; smaller K_a values indicate weaker acids with less extensive breakdown, such as acetic acid where only about 1% dissociates at typical concentrations. This partial decomposition establishes an equilibrium hydrogen ion concentration that determines the solution's without complete . In the context of phase changes, decomposition can intersect with transitions like , though most sublimations are purely physical processes without bond breaking, such as (solid CO₂) converting directly to CO₂ gas. However, certain cases involve chemical decomposition that mimics ; for , heating leads to reversible breakdown into gaseous products: \text{NH}_4\text{Cl(s)} \rightleftharpoons \text{NH}_3\text{(g)} + \text{HCl(g)} This decomposition occurs because the solid's bonds break, releasing molecules that can recombine upon cooling, distinguishing it from physical shifts by involving chemical changes in molecular structure. Buffer solutions exemplify the practical role of decomposition equilibria in maintaining system properties, particularly . A buffer typically consists of a weak and its conjugate , where the acid's partial decomposition provides H⁺ ions that neutralize added bases, while the base accepts protons from added acids, shifting the equilibrium to resist changes. For instance, in a buffer, the equilibria involving H₂PO₄⁻ decomposition help stabilize around 7.4 in biological and chemical applications, with the buffer capacity peaking near the pK_a of the acid. This controlled decomposition ensures minimal fluctuation, often within 1 unit, even with moderate additions of strong acids or bases. In , equilibrium contributes to pollution dynamics, notably through (N₂O₅), which undergoes reversible thermal breakdown: $2\text{N}_2\text{O}_5 \rightleftharpoons 4\text{NO}_2 + \text{O}_2 This equilibrium, with an increasing with temperature as the is endothermic, generates NO₂ and contributes to O₂ production that participate in night-time oxidation cycles, influencing tropospheric levels and photochemical formation in urban areas. The process links to broader NOx chemistry, where N₂O₅ on aerosols further produces , exacerbating and visibility reduction in events.

References

  1. [1]
    Composing and Decomposing Matter - University of Hawaii at Manoa
    Some compounds are decomposed by heating or by exposure to sunlight. For example, calcium carbonate (lime stone) decomposes into calcium oxide (quick lime) and ...
  2. [2]
    [PDF] 9. Chemical Reactions
    2) Decomposition: One substance is broken down into two or more, simpler substances. C12H22O11(s) 12C(s) + 11H2O(g) Pb(OH)2(s) PbO(s) + H2O(g) 2Ag2O(s) 4Ag(s) ...
  3. [3]
    Decomposition in daily life | Feature - RSC Education
    Jan 29, 2023 · In a decomposition reaction, substances are broken up into simpler parts (elements or simpler compounds), usually with heat, light or electricity.Missing: definition | Show results with:definition
  4. [4]
    Lesson 6.5: A Catalyst and the Rate of Reaction
    Jul 30, 2024 · This kind of change is a chemical reaction called decomposition. The decomposition of hydrogen peroxide is slow and is not usually noticeable.Instructions · 4 Explore · 5 Extend<|control11|><|separator|>
  5. [5]
    11.5: Decomposition Reactions - Chemistry LibreTexts
    Mar 20, 2025 · A decomposition reaction is a reaction in which a compound breaks down into two or more simpler substances.Decomposition Reactions · Exercise 11 . 5 . 1 · Example 11 . 5 . 1...
  6. [6]
    Decomposition Reactions - Chemistry LibreTexts
    Jun 12, 2023 · The official meaning of decomposition is a little bit more specific, and means a reaction in which one chemical splits into two or more chemicals.Missing: definition | Show results with:definition
  7. [7]
    Endothermic Reaction Vs Exothermic - Housing Innovations
    May 9, 2025 · Examples of endothermic reactions include the decomposition of calcium carbonate and the production of ammonia. Examples of exothermic reactions ...
  8. [8]
    4.5: Composition, Decomposition, and Combustion Reactions
    Jul 9, 2023 · A decomposition reaction produces multiple products from a single reactant. Combustion reactions are the combination of some compound with ...
  9. [9]
    Philosophy of Chemistry
    Mar 14, 2011 · In the 18th and early 19th centuries, chemical analyses of substances consisted in the decomposition of substances into their elemental ...
  10. [10]
    Lavoisier
    The first breakthrough in the study of chemical reactions resulted from the work of the French chemist Antoine Lavoisier between 1772 and 1794.
  11. [11]
    Lavoisier - Rare and Manuscript Collections - Cornell University
    Included among the manuscripts are laboratory notes from his dramatic experiments on the decomposition and recomposition of water, which helped to demonstrate ...
  12. [12]
    Decomposition Reaction - BYJU'S
    General Reaction Format. The general format of a decomposition reaction is provided below. AB -> A + B. Where AB is the parent molecule (reactant) and A & B ...
  13. [13]
    Decomposition Reaction: Definition, Examples, & Applications
    Aug 24, 2020 · General Equation. The following equation represents a general decomposition reaction. AB → A + B. When writing an actual reaction, the ...What is a Decomposition... · Examples of Decomposition... · How to Balance...
  14. [14]
    Composition, Decomposition, and Combustion Reactions
    A composition reaction produces a single substance from multiple reactants. A decomposition reaction produces multiple products from a single reactant.
  15. [15]
    Types of Chemical Reactions - Let's Talk Science
    Oct 14, 2021 · Decomposition reactions are essentially the reverse of synthesis reactions. This type of reaction involves a compound that breaks apart into ...
  16. [16]
    Decomposition & Synthesis Reactions - The Physics Classroom
    A decomposition reaction has a form that is opposite of a synthesis reaction. In synthesis reactions, there are two or more reactants and one product. In ...
  17. [17]
    5.4: Some Types of Chemical Reactions - Chemistry LibreTexts
    Aug 4, 2019 · Decomposition Reactions. A decomposition reaction is the reverse of a combination reaction. In a decomposition reaction, a single substance ...
  18. [18]
    Bond Energies - Chemistry LibreTexts
    Apr 3, 2025 · Energy is released to generate bonds, which is why the enthalpy change for breaking bonds is positive. Energy is required to break bonds. Atoms ...Bond Breakage and Formation · Example 2 : Generation of...
  19. [19]
    Bond enthalpies (article) | Thermodynamics | Khan Academy
    Enthalpy of reaction. During chemical reactions, the bonds between atoms may break, reform or both to either absorb or release energy. The result is a change ...
  20. [20]
    7.1 Bond Energy ad Enthalpy of Reaction
    Bond energies can be used to estimate the energy change of a chemical reaction. When bonds are broken in the reactants, the energy change for this process is ...
  21. [21]
    Thermal Decomposition Kinetics - an overview | ScienceDirect Topics
    The kinetics for thermal decomposition can be very complex and the Arrhenius equation describes the pyrolysis reaction rate based on temperature, mass ...
  22. [22]
    Origin and Justification of the Use of the Arrhenius Relation ... - MDPI
    Origin and Justification of the Use of the Arrhenius Relation to Represent the Reaction Rate of the Thermal Decomposition of a Solid. by. Benjamin Batiot.
  23. [23]
    [PDF] Template-free Synthesis of Mesoporous and Crystalline Transition ...
    Mar 4, 2020 · The method is based on the thermal crystalline transformation from basic carbonates to mesoporous metal oxides. The crystal interconversion not ...<|control11|><|separator|>
  24. [24]
    Water electrolysis: from textbook knowledge to the latest scientific ...
    May 16, 2022 · Water electrolysis – literally the decomposition of water under the action of electricity – was first performed by using static electricity by ...
  25. [25]
    Review article Alternate water electrolysis - ScienceDirect.com
    The theoretical decomposition voltage of water electrolysis is 1.23 V under standard conditions at 298 K [12]. The electrode involving OER dominates the cell ...
  26. [26]
    Electrolytic Cells
    Faraday's law of electrolysis can be stated as follows. The amount of a substance consumed or produced at one of the electrodes in an electrolytic cell is ...
  27. [27]
    [PDF] CHAPTER 7 LECTURE NOTES 7.1. Faraday's Laws of Electrolysis
    To produce w grams or (w/E) equivalents of an element by electrolysis, we need (w/E) × F Coulombs of charge to pass through the solution. Mathematically, Q = ...Missing: explanation | Show results with:explanation
  28. [28]
    How to Reliably Report the Overpotential of an Electrocatalyst
    It is highly important to report the electrocatalytic activity, especially the η value, through a unified and reliable method.
  29. [29]
    Sodium » historical information - WebElements Periodic Table
    Sodium was first isolated in 1807 by Sir Humphry Davy, who made it by the electrolysis of very dry molten sodium hydroxide, NaOH.
  30. [30]
    A pinch of sodium | Nature Chemistry
    Nov 23, 2011 · Davy discovered sodium in 1807 by isolating it from sodium hydroxide through electrolysis, and in 1811 he gave chlorine its name after ...
  31. [31]
    Photodecomposition - an overview | ScienceDirect Topics
    In the process, the photochemical mechanism of photolysis is divided into three stages: (1) the adsorption of light which excites electrons in the molecule, (2 ...
  32. [32]
    Making a photographic print using silver halides - RSC Education
    The decomposition of the silver halides in light is a photochemical redox reaction in which an electron is transferred from the halide ion to the silver ion ...<|control11|><|separator|>
  33. [33]
    quantum yield (Q04991) - IUPAC Gold Book
    Quantum yield is the number of defined events occurring per photon absorbed by the system. Strictly, it applies only for monochromatic excitation.
  34. [34]
    Photodegradation and photostabilization of polymers, especially ...
    UV radiation causes photooxidative degradation which results in breaking of the polymer chains, produces free radical and reduces the molecular weight.
  35. [35]
    Parameters that affect the photodegradation of dyes and pigments in ...
    Unstable dyes typically fade more easily under Visible (Vis) radiation, whereas dyes of high light fastness usually require UV radiation as it is of higher ...
  36. [36]
    Chemistry and the Black and White Photographic Process
    When the film is exposed to photons (discrete packets of light energy), the silver halides decompose to form an invisible, yet developable, latent image.
  37. [37]
    14.7: Catalysis - Chemistry LibreTexts
    Jul 7, 2023 · Catalysts are substances that increase the reaction rate of a chemical reaction without being consumed in the process.
  38. [38]
    Catalytic Decomposition Demonstration Sheet
    ... many such cycles). This is the definition of a catalyst: a substance that increases the rate of a chemical reaction without being consumed in the reaction.
  39. [39]
    5.4: Hydrolysis Reactions - Chemistry LibreTexts
    Mar 18, 2025 · Hydrolysis reactions are the reverse of condensation reactions. In a hydrolysis reaction, a larger molecule forms two (or more) smaller molecules and water is ...Hydrolysis of Esters · Hydrolysis of Amides · Career Focus: Athletic Trainer · Solution
  40. [40]
    Hydrolysis: Definition and Examples (Chemistry) - ThoughtCo
    Dec 8, 2019 · Hydrolysis is a type of decomposition reaction where one of the reactants is water; and typically, water is used to break chemical bonds in the other reactant.Hydrolysis Examples · Salt · Acid-Base
  41. [41]
    Nitroglycerin | C3H5(NO3)3 | CID 4510 - PubChem - NIH
    4 - Materials that in themselves are readily capable of detonation or explosive decomposition or explosive reaction at normal temperatures and pressures.
  42. [42]
    [PDF] Chemical Effects of Nuclear Decay
    Further, if the nuclear decay has occurred by means of alpha or beta decay, it will result in the transmutation of one element into another. At the moment ...<|control11|><|separator|>
  43. [43]
    [PDF] Manufacture of lime - NIST Technical Series Publications
    limestone into lime through the agency of heat. The term limestone is used to describe a class of rocks var}dng in com- position from pure calcium carbonate ...
  44. [44]
  45. [45]
    Ammonium dichromate volcano | Demonstration - RSC Education
    Try this demonstration to create a mini volcanic eruption illustrating the decomposition of ammonium dichromate. Includes kit list and safety instructions.
  46. [46]
    Ammonium dichromate | Cr2H8N2O7 | CID 24600 - PubChem
    3.2.​​ Ammonium dichromate can undergo thermal decomposition and produce a violent conflagration. It is a strong oxidizing agent and flammable solid.
  47. [47]
    Product distribution from fast pyrolysis of glucose-based carbohydrates
    Classification of pyrolysis products into three lumped categories (tar or bio-oil, char and gas) has become a standard over the years. Such classification, of ...
  48. [48]
    The Alpha–Bet(a) of Glucose Pyrolysis - ACS Publications
    Among the solid and gaseous products from α-d-glucose, char was 10.26 ± 3.90 ... Though the classification of pyrolysis products into overall tar, char ...
  49. [49]
    Products of chlorophyll photodegradation. I. Detection and separation
    Aug 10, 2025 · The porphyrin macrocycles were broken due to photodegradation. Therefore, the transmittance in the red region of oil samples was constantly ...Missing: breaks | Show results with:breaks
  50. [50]
    15.9: Hydrolysis of Esters - Chemistry LibreTexts
    Jan 31, 2022 · Acidic hydrolysis of an ester gives a carboxylic acid and an alcohol. Basic hydrolysis of an ester gives a carboxylate salt and an alcohol.Missing: RCOOR' + RCOOH + R'OH
  51. [51]
    Poly(vinyl chloride) degradation: identification of acidic ... - Nature
    Jul 30, 2025 · Chemical degradation leads to the release of acidic gases, notably hydrogen chloride (HCl), which not only compromises the integrity of the PVC ...
  52. [52]
    Physiology, Adenosine Triphosphate - StatPearls - NCBI Bookshelf
    Through metabolic processes, ATP becomes hydrolyzed into ADP, or further to AMP, and free inorganic phosphate groups. The process of ATP hydrolysis to ADP ...
  53. [53]
    Proteases: Multifunctional Enzymes in Life and Disease - PMC - NIH
    Proteases are the efficient executioners of a common chemical reaction: the hydrolysis of peptide bonds (16). Most proteolytic enzymes cleave α-peptide bonds ...
  54. [54]
    Chapter 7: Catalytic Mechanisms of Enzymes - Chemistry
    Proteolytic enzymes (also termed peptidases, proteases and proteinases) are capable of hydrolyzing peptide bonds in proteins. They can be found in all living ...
  55. [55]
    Degradation of Cellulose Derivatives in Laboratory, Man-Made ... - NIH
    Jun 28, 2022 · Complete biodegradation of cellulose ultimately results in carbon dioxide and water under aerobic conditions and carbon dioxide, methane ...
  56. [56]
    Background - SERC (Carleton)
    Jan 13, 2021 · As these plant residues decompose, they release the elements and nutrients contained in them back to the soil. These enter into the nutrient ...<|control11|><|separator|>
  57. [57]
    Biogeochemical Cycles in Plant–Soil Systems - PubMed Central - NIH
    Plants absorb nutrients/elements from the soil to fuel their metabolic activities, while microbial communities decompose organic matter, releasing nutrients/ ...
  58. [58]
    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 ...<|separator|>
  59. [59]
    Lime Cycle - an overview | ScienceDirect Topics
    A lime kiln is a kiln used to produce quicklime by the calcination of limestone. The chemical equation representing this reaction is: CaCO3 + heat = CaO + CO2.
  60. [60]
    Production of hydrogen - U.S. Energy Information Administration (EIA)
    Jun 23, 2023 · Electrolysis uses electricity to produce hydrogen. Electrolysis is a process that splits hydrogen from water using an electric current.Hydrogen Explained... · Hydrogen Production · Electrolysis Uses...
  61. [61]
    Hydrogen Production and Delivery | Hydrogen and Fuel Cells - NREL
    Feb 10, 2025 · One solution is to produce hydrogen through the electrolysis—splitting with an electric current—of water and to use that hydrogen in a fuel cell ...
  62. [62]
    Cracking and related refinery processes
    Cracking, as the name suggests, is a process in which large hydrocarbon molecules are broken down into smaller and more useful ones.
  63. [63]
    Houdry Process for Catalytic Cracking - American Chemical Society
    The Houdry process conserved natural oil by doubling the amount of gasoline produced by other processes. It also greatly improved the gasoline octane rating, ...Development of the Houdry... · Impact of the Houdry Process
  64. [64]
    Lime Market Size, Share, Analysis & Growth Report, 2033
    Oct 21, 2025 · In 2023, global lime production was estimated at approximately 430 million metric tons. This substantial output underscores the material's ...
  65. [65]
    Thermogravimetric Analysis
    Thermogravimetric analysis (TGA) is a method of thermal analysis that measures weight changes in a material as a function of temperature.
  66. [66]
    Soil carbon determination by thermogravimetrics - PMC - NIH
    Feb 12, 2013 · Thermal decomposition of most carbonates is quite separate and obvious from other constituents occurring over 600 °C (Kasozi, Nkedi-Kizza & ...
  67. [67]
    [PDF] High temperature properties and decomposition of inorganic salts ...
    Since thekinetics of carbonate decomposition have beenmore extensively studied than those of any other class of inorganic salts, an examination of the ...
  68. [68]
    Thermal and Chemical Characterization of Non-Metallic Materials ...
    Thermogravimetric analysis (TGA) is widely employed in the thermal characterization of non-metallic materials, yielding valuable information on ...
  69. [69]
    New Prospects in the Electroanalysis of Heavy Metal Ions (Cd, Pb ...
    Electrochemical techniques, more specifically voltammetry, due to its properties, is a promising method for the simultaneous detection of heavy metal ions. This ...
  70. [70]
    [PDF] development of novel electrochemical sensor to detect heavy metals
    Apr 15, 2025 · Anodic stripping voltammetry is a popular electrochemical method for the quantitative determination of metal ions because of its high ...
  71. [71]
    Oxygen and chlorine isotopic fractionation during perchlorate ...
    Feb 28, 2007 · These data indicate that isotope ratio analysis will be useful for documenting perchlorate biodegradation in soils and groundwater. The ...Missing: decomposition | Show results with:decomposition<|control11|><|separator|>
  72. [72]
    Isotope ratio method: state-of-the-art of forensic applications to ...
    The isotope ratio method has promise as a useful tool in microbial forensics to establish the origin and transmission route of a particular microbial strain.Missing: decomposition | Show results with:decomposition
  73. [73]
    Cement production and its decarbonization - thyssenkrupp
    Around 40% comes from the combustion of fossil fuels to heat the kilns, while 60% is released during the thermal decomposition of limestone into carbon dioxide ...
  74. [74]
    Towards decarbonization of cement industry: a critical review of ...
    Jul 4, 2025 · Cement production accounts for 8% of global CO2 emissions, necessitating its deep decarbonization.
  75. [75]
    CFCs and their substitutes in stratospheric ozone depletion.
    Mario Molina, showed that CFCs could be a major source of inorganic chlorine in the stratosphere following their photolytic decomposition by ultra-violet (UV) ...Missing: photodecomposition | Show results with:photodecomposition
  76. [76]
    [PDF] Chlorofluorocarbons and Ozone Depletion
    Photodissociation of the chlorofluoromethanes in the stratosphere produces significant amounts of chlorine atoms, and leads to the destruction of atmospheric ...Missing: photodecomposition radicals
  77. [77]
    Ozone Milestones - JPL Atmospheric Chemistry
    CFCs and Ozone Depletion Linked. 1974: Photochemical theory of ozone destruction by chlorine radicals is proposed. Initial concern focuses on ozone loss due ...
  78. [78]
    Burned: Why Waste Incineration Is Harmful - NRDC
    Jul 19, 2021 · For example, toxics like PFAS, dioxins, and mercury compounds are found in the environment, people, and marine mammals in the Arctic, far from ...Missing: decomposition | Show results with:decomposition
  79. [79]
    Systematic review on environmental impact assessment of ...
    This paper presents a systematic review of the environmental implications of incineration technologies through life cycle assessment (LCA), as applied to MSW ...
  80. [80]
    Comprehensive characterization and environmental implications of ...
    Nov 22, 2024 · During incineration, wastes undergo composition and decomposition reactions, altering their physical and chemical properties. Solid ash residues ...
  81. [81]
    Basic Information about Landfill Gas | US EPA
    Sep 12, 2025 · Landfill gas (LFG) is a natural byproduct of the decomposition of organic material in landfills. LFG is composed of roughly 50 percent methane.
  82. [82]
    [PDF] Waste Management - Intergovernmental Panel on Climate Change
    7. Page 16. 600. Waste Management. Chapter 10. 10.4.2 CH4 management at landfills. Global CH4 emissions from landfills are estimated to be. 500–800 MtCO2-eq/yr ...
  83. [83]
    [PDF] Global Methane Emissions and Mitigation Opportunities
    Methane is the second most abundant GHG after carbon dioxide (CO2), accounting for 14 percent of global emissions.
  84. [84]
    Chemical Reactivity - MSU chemistry
    Thus, if the products have a greater total bond energy than the reactants the reaction will be exothermic, and the opposite is true for an endothermic reaction.
  85. [85]
    CH104: Chapter 5 - Chemical Reactions - Chemistry
    A general equation for decomposition reactions is: AB → A + B. For example, the equation below is a decomposition reaction that occurs when Sodium ...
  86. [86]
    [PDF] 5 Types Of Chemical Reactions
    These reaction types are synthesis, decomposition, single displacement, double displacement, and combustion. Each type exhibits distinct patterns in ...
  87. [87]
    [PDF] Chapter 19 Principles of Reactivity: Entropy and Free Energy
    While exothermic reactions are almost always spontaneous, the entropy does play a role, and,should the entropy increase greatly enough, cause the reaction to be ...
  88. [88]
    [PDF] CHAPTER 19 SPONTANEOUS CHANGE: ENTROPY AND GIBBS ...
    3A (M) The entropy change for the reaction is expressed in terms of the standard entropies of the reagents.
  89. [89]
    Introduction to polymers: 3.4 The petrochemical industry | OpenLearn
    3.1 Primary sources of synthetic polymers · 3.2 Petrochemical processing · 3.2.1 Thermal cracking · 3.2.2 Ethane cracking · 3.3 Petrochemical intermediates and ...
  90. [90]
    Shifting Equilibria: Le Châtelier's Principle – Chemistry
    Le Châtelier's principle can be used to predict changes in equilibrium concentrations when a system that is at equilibrium is subjected to a stress.
  91. [91]
    [PDF] Chemical Equilibrium
    According to Le Chatelier's principle, we can expect a reaction to respond to a lowering of temperature by releasing heat and to respond to an increase of tem-.
  92. [92]
    Le Chatelier's Principle Fundamentals - Chemistry LibreTexts
    Jan 29, 2023 · An action that changes the temperature, pressure, or concentrations of reactants in a system at equilibrium stimulates a response that partially ...
  93. [93]
    Reversible reactions - AQA - GCSE Combined Science Revision
    It breaks down when heated, forming ammonia and hydrogen chloride. When these two gases are cool enough, they react together to form ammonium chloride again.
  94. [94]
    Reversible reactions examples explained NH4Cl <=> NH3 + HCl ...
    On heating strongly above 340oC, the white solid ammonium chloride, thermally decomposes into a mixture of two colourless gases ammonia and hydrogen chloride.
  95. [95]
    For the decomposition reaction PCl5(g) ⇌ PCl3(g) + Cl2(g) - Pearson
    For the decomposition reaction PCl5(g) ⇌ PCl3(g) + Cl2(g), Kp = 381 at 600 K and Kc = 46.9 at 700 K. (a) Is the reaction endothermic or exothermic? Explain.
  96. [96]
    [PDF] ANSWERS: Equilibria AS90310 2004-2009 & AS91166 2012
    The equilibrium can be represented as: PCl5(g) Ý PCl3(g) + Cl2(g). (a) Complete the equilibrium constant expression for this reaction. Kc = [PCl3(g)][Cl2(g)].
  97. [97]
    The Haber Process - Chemistry LibreTexts
    Jan 29, 2023 · This page describes the Haber Process for the manufacture of ammonia ... The reaction is reversible and the production of ammonia is exothermic.
  98. [98]
  99. [99]
    Some Special Types of Equilibria – Introductory Chemistry
    This is the case for all weak acids and bases. An acid dissociation constant, Ka, is the equilibrium constant for the dissociation of a weak acid into ions.
  100. [100]
    [PDF] Sublimation
    The reverse process of sublimation is deposition or desublimation, in which a substance passes directly from a gas to a solid phase. Sublimation has also ...