Fact-checked by Grok 2 weeks ago

Chemical compound

A chemical compound is a pure chemical substance composed of two or more different chemical elements that are chemically bonded together in a fixed ratio by mass. These bonds result in a distinct set of physical and chemical properties that differ from those of the individual elements involved. Compounds can only be broken down into their constituent elements through chemical reactions, not physical separation methods like or . Chemical compounds are broadly classified by the nature of their chemical bonds and composition. Ionic compounds form when one element transfers electrons to another, creating positively and negatively charged ions that are held together by electrostatic attractions, as seen in salts like (NaCl). Covalent compounds, also known as molecular compounds, arise from the sharing of electrons between atoms, resulting in discrete molecules such as (H₂O) or (CH₄). Additionally, compounds are often divided into and inorganic categories: compounds contain carbon atoms bonded primarily to , oxygen, , or other light elements and include the complex molecules essential to life, while inorganic compounds encompass all others, such as salts, minerals, and oxides. The study and synthesis of chemical compounds underpin nearly all areas of chemistry, enabling the development of materials, pharmaceuticals, and fuels that drive modern science and industry. From simple binary compounds like (CO₂) to complex polymers, they represent the vast diversity of matter and facilitate reactions that sustain biological systems and technological innovations.

Fundamental Concepts

Definition

A chemical compound is an electrically neutral substance composed of two or more different chemical elements, with their atoms present in definite ratios and held together by chemical bonds, resulting in a distinct identity and properties unlike those of its constituent elements. These ratios typically ensure that the compound has a constant composition by mass, distinguishing it from mixtures where proportions can vary, though some non-stoichiometric compounds exhibit variable composition. Representative examples include water (H_2O), formed by hydrogen and oxygen in a 2:1 atomic ratio, which displays unique properties like high surface tension and solvent capabilities absent in the pure elements. Similarly, sodium chloride (NaCl), a compound of sodium and chlorine in a 1:1 ratio, exists as a stable ionic crystal with high melting point and electrical conductivity in solution, differing markedly from the reactive metals and gas of its components. A key characteristic of most chemical compounds is their adherence to the , which states that every sample of a given compound contains the same elements in the same fixed proportion by mass, irrespective of its origin or preparation method, although exceptions exist in non-stoichiometric compounds such as certain oxides like Fe_{0.95}O; this principle was first articulated by in 1794. Compounds also exhibit homogeneity at the molecular or ionic level, meaning they are uniform throughout in structure and composition, enabling predictable behavior in chemical reactions. The concept of a chemical compound has evolved from empirical observations relying on macroscopic properties, such as consistent analytical yields and behaviors, to modern quantum mechanical descriptions that elucidate through interactions and molecular orbitals. This progression provides a foundational framework for understanding how elements combine to form stable entities with emergent properties.

Distinction from Elements and Mixtures

Chemical compounds differ from pure elements in both composition and chemical behavior. Elements are pure substances composed solely of one type of , such as diatomic oxygen (O₂), and cannot be chemically decomposed into simpler substances. Compounds, by contrast, consist of two or more different elements chemically bonded together in fixed ratios, and they can be broken down into their elemental components only through chemical reactions. This decomposability underscores the fundamental transformation involved in forming compounds from elements. Compounds are also distinct from mixtures, which involve physical combinations of two or more substances without chemical bonding or fixed proportions, as seen in air—a blend of nitrogen (N₂), oxygen (O₂), argon, and other gases. Mixtures can be separated into their components using physical methods like filtration, evaporation, or distillation, whereas compounds require chemical reactions, such as electrolysis or thermal decomposition, for separation. This variability in composition and ease of physical separation highlights mixtures as non-chemical unions, unlike the stable, uniform structure of compounds. A hallmark of compounds is their exhibition of properties entirely different from those of their constituent elements, providing a practical test for identification. Sodium, a soft metal that explodes violently in , and , a poisonous greenish-yellow gas, combine to form (NaCl), a white crystalline solid commonly known as table salt that is inert in water and essential for human health. Mixtures, however, display the additive properties of their components without such emergent characteristics. The , formulated by in 1803, further differentiates by stating that if two form more than one , the masses of one element combining with a fixed mass of the other are in small whole-number ratios, reflecting discrete atomic combinations. This contrasts with mixtures, where proportions vary arbitrarily and do not follow such ratios. This principle ties to the fixed composition of , as per the .

Historical Development

Ancient and Pre-modern Ideas

In ancient Greek philosophy, Empedocles (c. 494–434 BCE) proposed that all matter consists of four indivisible "roots" or elements—earth, air, fire, and water—which combine and separate under the opposing forces of Love (attraction) and Strife (repulsion) to form the diversity of substances observed in nature. These combinations were not seen as true fusions but as mechanical mixtures retaining the identity of their elemental components, lacking any notion of fixed proportions or chemical bonding. Aristotle (384–322 BCE) later refined this framework, adopting the four elements while arguing that their mixtures could produce substances with emergent properties distinct from the originals, such as flesh or bone, though he maintained that elements could be recovered unchanged upon decomposition. This Aristotelian view of qualitative changes through elemental blending profoundly shaped medieval European and Islamic thought, where it underpinned early speculations on material transformations without empirical verification. During the and into the , alchemical traditions built on these ideas, pursuing of base metals into through "compositions" of substances, though without an understanding of atomic structure. (c. 721–815 CE), known as in the Latin West, classified materials into spirits, metals, and stones, and described experimental preparations of metal alloys and salts, viewing metals as blends of (combustible principle) and mercury (fluid principle) in varying proportions to achieve balance (mīzān). His works emphasized , , and other processes to manipulate these compositions, but treated them as holistic unions rather than discrete particles. Later, (1493–1541) advanced alchemical medicine by positing three principles—, mercury, and (tria prima)—as the foundational "compounds" of all matter, advocating chemical preparations like (an tincture) as therapeutic agents to restore bodily balance. Pre-modern observers empirically noted uniform substances like alloys and salts, often through metallurgical and practical arts, but conflated them with elemental forms without recognizing fixed compositional ratios. For instance, ancient bronzes—alloys of and tin—were recognized as homogeneous materials stronger than their components, yet attributed to divine or elemental harmony rather than quantifiable mixing, as documented in Roman texts by . Salts, such as those derived from seawater , were similarly valued for their consistency in or preservation, but seen as natural essences indistinguishable from pure elements in alchemical schemes. The 16th and 17th centuries marked a transition via iatrochemistry, where alchemical pursuits shifted toward practical "compositions" for medical use, laying groundwork for systematic experimentation. Proponents like and his followers prepared mineral-based remedies, challenging Galenic humoral theory with chemical interventions, such as antimony purges, to treat diseases as imbalances of corpuscular matter. This empirical focus on reproducible preparations elevated chemistry's status among natural philosophers, fostering institutions like the Royal Society and paving the way for in the ensuing era.

17th to 19th Century Advances

The 17th century marked a pivotal shift in understanding chemical compounds through the work of Robert Boyle, whose 1661 publication The Sceptical Chymist challenged traditional alchemical views by proposing that matter consists of "corpuscles" or particles of elements that unite to form compounds, rejecting the ancient four-element theory of earth, air, fire, and water. Boyle emphasized experimental verification over philosophical speculation, arguing that compounds are distinct "corpuscles of one or more kinds" produced by the mechanical union of simpler particles, as demonstrated through his analyses of substances like salts and metals. This corpuscularian hypothesis laid the groundwork for viewing compounds as uniform entities arising from specific combinations, influencing subsequent empirical approaches in chemistry. Building on Boyle's foundations in the late , advanced the concept of compounds as stable products formed when elements combine in definite ways, often accompanied by measurable changes in weight, such as during where substances gain mass by uniting with oxygen. In his 1789 treatise Traité élémentaire de chimie, Lavoisier formalized the , stating that "in every operation an equal quantity of matter exists both before and after the operation," which provided quantitative evidence that compounds retain the total mass of their elemental constituents without creation or destruction. For instance, Lavoisier showed that involves oxygen combination, transforming elements into compounds like oxides, thereby redefining chemical reactions as rearrangements rather than qualitative alterations. Joseph Proust's contributions in the further solidified the uniformity of compounds through his , established via meticulous experiments demonstrating that elements in a compound always occur in fixed mass ratios regardless of preparation method or source. Proust's 1799 analysis of copper carbonate, for example, revealed identical proportions of , carbon, and oxygen in both synthetic and naturally occurring samples, countering variable composition theories and confirming compounds' consistent elemental makeup. This law underscored that compounds like copper carbonate maintain invariable , such as approximately 5 parts to 2 parts by weight, providing empirical support for their predictable nature. John Dalton's , outlined in his 1808 A New System of Chemical Philosophy, posited that chemical compounds form when atoms of different elements unite in simple, whole-number ratios, reviving and quantifying earlier corpuscularian ideas with a modern framework. Dalton illustrated this with , initially proposing a 1:1 ratio of to oxygen atoms (later revised to 2:1), and introduced the , observing that elements like carbon and oxygen form compounds such as and where oxygen masses are in a 1:2 ratio for fixed carbon. These principles explained compounds as discrete aggregates of indivisible atoms, enabling the prediction of compositional ratios and distinguishing them from mere mixtures. In the , contributed to the dissemination of these emerging ideas through his educational writings, which popularized Boyle's emphasis on compounds' uniformity and experimental rigor in philosophical and scientific texts aimed at broader audiences. Watts' works, such as those integrating into moral and intellectual improvement, helped embed the concept of compounds as consistent unions of elements within 18th-century educational , bridging elite with public understanding.

20th Century Refinements

In the early , revolutionized the understanding of chemical bonding by proposing the theory in his 1916 paper, which described covalent bonds in compounds as shared pairs of electrons between atoms, thereby explaining the stability of molecules like and . This model introduced the , positing that atoms tend to achieve eight electrons in their valence shells to mimic configurations, providing a foundational framework for predicting compound formation beyond classical atomic ratios. Building on Lewis's ideas, extended the octet theory in 1919, applying it systematically to and atomic structure, which clarified in coordination compounds and influenced the development of complex molecular architectures. Langmuir's work emphasized the role of octets in determining compound stability, bridging organic and . In the 1930s, advanced these concepts further by developing the scale, a quantitative measure of an atom's ability to attract s in a , allowing chemists to assess in compounds like . Pauling also introduced hybrid orbital theory, explaining directional ing in compounds such as through the mixing of atomic orbitals, as detailed in his seminal book The Nature of the Chemical Bond and the Structure of Molecules and Crystals. Spectroscopic techniques provided empirical validation for these theoretical refinements, with and William Lawrence Bragg's pioneering in the 1910s revealing the ionic lattice structures of compounds like , confirming extended three-dimensional arrangements rather than discrete molecules. Post-1940s advancements in (NMR) spectroscopy, independently developed by and Edward M. Purcell, enabled the determination of molecular structures in solution by detecting nuclear spin interactions, revolutionizing the analysis of organic compounds. Similarly, infrared (IR) spectroscopy matured in the 1940s and 1950s with commercial instruments, allowing identification of functional groups and vibrational modes in molecular compounds through absorption patterns. The also saw the broadening of compound definitions to include organometallics and polymers, recognizing their distinct chemical identities. Organometallic compounds, featuring direct carbon-metal bonds, gained prominence after the 1951 discovery of , which exemplified stable sandwich structures and spurred research into complexes as viable chemical entities. Hermann Staudinger's macromolecular theory in the established polymers as giant covalent compounds composed of repeating units, validated by his 1953 work on substances like . A key example is the 1950s development of Ziegler-Natta catalysts by and , which enabled stereospecific of olefins into and , demonstrating controlled synthesis of polymeric compounds.

Classification by Structure and Bonding

Covalent Compounds

Covalent compounds are chemical substances formed by , in which atoms share pairs of electrons to achieve stable electron configurations, typically between atoms. According to the International Union of Pure and Applied Chemistry (IUPAC), a is defined as a region of relatively high between nuclei that arises at least partly from the sharing of electrons, resulting in an attractive force and a characteristic internuclear distance. These bonds commonly occur among , as their similar electronegativities favor electron sharing over transfer, leading to the formation of discrete molecules or extended networks. Covalent bonds vary in multiplicity based on the number of shared electron pairs: single bonds involve one pair (e.g., the Cl–Cl bond in gas, Cl₂), double bonds involve two pairs (e.g., the O=O bond in oxygen gas, O₂), and triple bonds involve three pairs (e.g., the N≡N bond in nitrogen gas, N₂). Covalent compounds can manifest as molecular structures, consisting of discrete units like (CH₄), which adopts a tetrahedral around the central carbon atom due to four equivalent single bonds, or (CO₂), which is linear with two double bonds. In contrast, network solids feature an extended of covalent bonds, such as , where each carbon atom is tetrahedrally bonded to four others in a three-dimensional array, or (SiO₂), a silicon-oxygen tetrahedral network. The of covalent bonds depends on the difference between bonded atoms: nonpolar bonds occur when electronegativities are similar, as in H₂ where s are shared equally, while polar bonds arise from unequal sharing, creating a , as in HCl with partial negative charge on and partial positive on . Molecular shapes in covalent compounds are predicted by Valence Shell Electron Pair Repulsion (, which posits that electron pairs around a central atom arrange to minimize repulsion; for instance, (NH₃) exhibits a pyramidal due to three bonding pairs and one on , resulting in bond angles of approximately 107°. Representative examples of covalent compounds include organic molecules like hydrocarbons (e.g., , C₂H₆, with C–C single and C–H bonds) and inorganic ones like (SiO₂)./04%3A_Covalent_Bonding_and_Simple_Molecular_Compounds/4.07%3A_Organic_Chemistry) Covalent compounds predominate in , where carbon's ability to form stable chains and rings underpins vast molecular , and in , where they constitute the majority of biomolecules such as proteins, , and nucleic acids, enabling the complex structures essential for .

Ionic Compounds

Ionic compounds are chemical substances composed of positively and negatively charged ions bound together by strong electrostatic attractions, typically resulting from the interaction between metals and nonmetals. These compounds form extended network structures rather than discrete molecules, distinguishing them from covalent compounds where electrons are shared. The ionic bond arises from the complete transfer of electrons, leading to oppositely charged ions that stabilize each other through Coulombic forces. The formation of ionic compounds involves the transfer of one or more electrons from a metal atom, which has low , to a atom with high . This creates cations (positive ions) from the metal and anions (negative ions) from the , achieving stable electron configurations akin to . For instance, in , the sodium atom donates its to , forming \ce{Na+} and \ce{Cl-} ions. Ionic compounds crystallize into infinite three-dimensional lattices, where cations and anions are arranged in repeating patterns to maximize electrostatic attraction and minimize repulsion. In the rock salt structure, exemplified by NaCl, each cation is surrounded by six anions in an octahedral coordination, forming a face-centered cubic array. The cesium chloride structure, in contrast, adopts a body-centered cubic arrangement, with each \ce{Cs+} coordinated to eight \ce{Cl-} ions, which is favored when the cation is larger relative to the anion. The chemical formulas of ionic compounds are empirical, reflecting the simplest ratio of cations to anions required for electrical neutrality in the , such as \ce{NaCl} or \ce{CaCO3}. Unlike molecular compounds, these formulas do not represent units but the of the entire . in follows general rules, such as the principle that "like dissolves like," where polar molecules hydrate ions to disrupt the ; for example, most salts are soluble, while many sulfates and carbonates are not. Representative examples of ionic compounds include simple salts like (\ce{KCl}) used in fertilizers and table salt (\ce{NaCl}), metal oxides such as (\ce{MgO}) employed in refractories, and compounds incorporating polyatomic ions, like the ion (\ce{SO4^2-}) in (\ce{CaSO4 \cdot 2H2O}), a used in construction. Polyatomic ions, such as or , function as single anionic units within the lattice despite consisting of multiple atoms bonded covalently internally. The strong ionic interactions in these lattices impart high melting points, as significant energy is required to overcome the electrostatic forces; for example, \ce{NaCl} has a melting point of 801°C, and \ce{MgO} exceeds 2800°C. In the solid state, ionic compounds are electrical insulators because ions are fixed in position, but they conduct effectively when molten or dissolved in , as the mobile ions carry charge. Some ionic bonds exhibit partial covalent character due to the polarization of electron clouds by highly charged ions.

Intermetallic Compounds

Intermetallic compounds are solid-state materials composed of two or more metallic elements, characterized by fixed stoichiometric ratios, ordered structures, and that distinguishes them from disordered solid solutions or random alloys. Unlike solid solutions where atoms substitute randomly in a , intermetallics form distinct phases with long-range atomic , such as the face-centered cubic in where and atoms occupy specific sublattices. This ordered arrangement arises from thermodynamic stability at specific compositions, often revealed through phase diagrams. The bonding in intermetallic compounds involves delocalized valence electrons shared across the metallic framework, similar to pure metals, but the precise ordering introduces directional preferences and partial covalent character that enhance stability and differ from the isotropic bonding in elemental metals. This hybrid nature results from variations in and atomic size among the metals, leading to concentrations along certain bonds. The systematic recognition of intermetallics as a class emerged in the early , driven by studies that mapped stable compounds in binary metal systems; for instance, the work of researchers like William Hume-Rothery in the and established empirical rules for their formation, including atomic size factors (differences exceeding 15% favor compounds over solutions), valency mismatches, and electronegativity contrasts that promote ordered phases when solid is limited. Representative examples include Zintl phases, such as NaTl, where an electropositive like sodium transfers electrons to form polyanionic networks of post-transition metals (e.g., tetrahedral Tl₄⁸⁻ units), exhibiting semiconductor-like behavior due to the localized anionic structures within a metallic matrix. Another class is the , exemplified by MgCu₂, which adopt close-packed cubic or hexagonal structures (e.g., Fd-3m) maximizing metal-metal coordination and appearing in over 300 binary systems. These compounds are typically harder and more brittle than disordered alloys, with Vickers hardness values around 350 kg/mm² and low (e.g., 1-2 MPa·m¹/²), attributed to restricted motion in the ordered lattice; however, they enable specialized uses, such as in superconductors like Nb₃Sn, which achieves critical temperatures up to 18 K and supports high magnetic fields in applications like MRI magnets.

Coordination Complexes

Coordination complexes, also known as coordination compounds, consist of a central or surrounded by a set of ligands that donate pairs to form coordinate bonds./09:Coordination_Chemistry_I-_Structure_and_Isomers/9.05:_Coordination_Numbers_and_Structures) This was first systematically explained by in 1893, who proposed that ligands occupy specific positions around the metal, leading to defined geometries such as octahedral for 6, as demonstrated through his studies on (III) ammine complexes. Werner's theory distinguished between primary valences (ionic bonds to counterions) and secondary valences (coordinate bonds to ligands), confirming octahedral geometry in compounds like [Co(NH₃)₆]Cl₃, where six ligands arrange around the (III) . The , defined as the number of donor atoms from s directly bound to the central metal, typically ranges from 4 to 6 for complexes, influencing the overall geometry./09:Coordination_Chemistry_I-_Structure_and_Isomers/9.05:_Coordination_Numbers_and_Structures) Common geometries include tetrahedral or square planar for coordination number 4 and octahedral for 6, with the latter being prevalent due to efficient packing around the metal ./05:Coordination_Chemistry_I-_Structures_and_Isomers/5.03:_Coordination_Numbers_and_Structures) Bonding in these complexes primarily involves sigma donation from ligand lone pairs into empty metal orbitals, forming a , often supplemented by where filled metal d-orbitals donate electrons to empty ligand pi* orbitals, strengthening the metal-ligand interaction as seen in metal carbonyls./05:Coordination_Chemistry_and_Crystal_Field_Theory/5.05:%CE%A0-Bonding_between_Metals_and_Ligands) Isomerism arises from different ligand arrangements, leading to geometric isomers such as cis and trans forms in square planar or octahedral complexes./Coordination_Chemistry/Structure_and_Nomenclature_of_Coordination_Compounds/Isomers/Stereoisomers:_Geometric_Isomers_in_cis-platin) For example, in the square planar (II) complex [Pt(NH₃)₂Cl₂], the cis isomer () has chloride ligands adjacent, while the trans isomer has them opposite, with the cis form exhibiting distinct reactivity due to its geometry./Coordination_Chemistry/Structure_and_Nomenclature_of_Coordination_Compounds/Isomers/Stereoisomers:_Geometric_Isomers_in_cis-platin) Optical isomerism occurs in octahedral complexes with bidentate ligands, such as [Co(en)₃]³⁺, where enantiomers are non-superimposable mirror images resulting from chiral ligand arrangements. Nomenclature follows IUPAC rules, where ligands are named first in alphabetical order with prefixes indicating number (e.g., tetraammine for four NH₃), followed by the metal with its oxidation state in Roman numerals, and counterions suffixed. For instance, [Cu(NH₃)₄]SO₄ is named tetraamminecopper(II) sulfate, reflecting four ammonia ligands around Cu²⁺ and sulfate as the anion. Representative examples include , where iron(II) is coordinated by four atoms in a and one side chain, enabling reversible oxygen binding. Another is the EDTA complex [Fe(EDTA)]⁻, a hexadentate chelate used for sequestering metal ions in and due to its strong binding affinity. Coordination complexes play vital roles in catalysis, such as [RhCl(PPh₃)₃] for reactions; in dyes, like for pigments; and in medicine, with as an anticancer agent targeting DNA./19%3A_Transition_Metals_and_Coordination_Chemistry/19.02%3A_Coordination_Chemistry_of_Transition_Metals) Their tunable properties stem from the metal-ligand interactions, enabling applications in and therapeutics.

Chemical Bonding and Interactions

Intramolecular Bonding

Intramolecular bonding refers to the strong interactions that hold atoms together to form the structural units of chemical compounds, such as ions in ionic lattices, , or atoms in metallic solids. These bonds are typically characterized by energies ranging from hundreds to thousands of kilojoules per , far stronger than intermolecular forces. The primary types include ionic, covalent, and metallic bonds, with additional variants like coordinate bonds appearing in certain complexes. Ionic bonds form through electrostatic attraction between oppositely charged s, typically a cation and an anion, resulting in a lattice structure in ionic compounds. The strength of this attraction is quantified by , which represents the released when gaseous s combine to form the solid or the required to separate them. A basic expression for the between ion pairs is given by the U = \frac{k q_1 q_2}{r}, where k is Coulomb's constant, q_1 and q_2 are the charges, and r is the distance between ion centers; for extended s, this is modified by a to account for the full array of interactions. Covalent bonds arise from the sharing of pairs between atoms, allowing each to achieve a stable configuration, often following the . The indicates the number of shared pairs: a involves one pair (order 1), a two pairs (order 2), and a three pairs (order 3), with higher orders generally leading to shorter, stronger bonds. Atomic orbitals hybridize to form bonds with optimal overlap; for example, in (CH₄), the carbon atom undergoes sp³ hybridization, combining one s and three p orbitals to create four equivalent sp³ orbitals that form bonds with 1s orbitals. Metallic bonds occur in metals and alloys, where valence electrons are delocalized, forming a "sea" of mobile electrons surrounding a of positive metal ions; this model explains properties like malleability and . arises from band theory, in which orbitals overlap to form continuous bands; in metals, the overlap or have a narrow gap, allowing electrons to move freely under an . Other intramolecular bonds include coordinate (or dative) bonds, found in coordination complexes, where both electrons in the shared pair are donated by one atom (a ligand) to a central metal ion, forming a covalent bond with directional character. Hydrogen bonds can also act as strong intramolecular interactions in specific cases, such as the base pairing in DNA, where adenine-thymine pairs form two hydrogen bonds and guanine-cytosine pairs form three, stabilizing the double helix structure. Bond energies, or dissociation energies, measure the strength required to break a homolytically into neutral radicals; for instance, the average C-H is 413 kJ/mol. Factors influencing include (higher order increases strength), (shorter bonds are stronger), and difference between atoms (greater differences lead to more polar bonds, which can slightly weaken the covalent character but enhance ionic contributions in polar covalent bonds).

Intermolecular Forces

Intermolecular forces are the relatively weak electrostatic attractions between s or molecular units in chemical compounds, distinct from the stronger intramolecular bonds that hold atoms within a . These forces play a crucial role in determining the bulk behavior of substances, such as their phase transitions and interactions in mixtures. Van der Waals forces represent a broad category of intermolecular attractions, encompassing both London dispersion forces and dipole-dipole interactions. London dispersion forces arise from temporary fluctuations in electron distribution that create instantaneous s, leading to attractions between nonpolar molecules; their strength increases with molecular size and surface area, as observed in iodine (I₂) where larger, more polarizable molecules exhibit stronger dispersion effects. Dipole-dipole interactions, on the other hand, occur between molecules possessing permanent dipole moments due to uneven charge distribution, such as in acetone (CH₃COCH₃), where the partial positive and negative charges align to stabilize adjacent molecules. Hydrogen bonding is a particularly strong form of dipole-dipole interaction that forms when a hydrogen atom, covalently bonded to a highly electronegative atom like nitrogen, oxygen, or fluorine, interacts with the lone pair of electrons on another electronegative atom in a neighboring molecule. Common examples include water (H₂O), hydrogen fluoride (HF), and ammonia (NH₃), where these bonds create network-like structures that significantly enhance intermolecular cohesion compared to similar compounds lacking such capability, as evidenced by water's anomalously high boiling point relative to hydrogen sulfide (H₂S)./Physical_Properties_of_Matter/Atomic_and_Molecular_Properties/Intermolecular_Forces/Specific_Interactions/Hydrogen_Bonding) Ion-dipole forces involve attractions between an and the partial charges on a polar , commonly observed in aqueous solutions where ions like those from (NaCl) become surrounded by hydration shells formed by molecules orienting their dipoles toward the ion. Pi-pi stacking refers to the noncovalent interactions between electron-rich pi clouds of aromatic rings, often resulting in parallel or displaced alignments that stabilize structures in compounds like dimers. These intermolecular forces collectively govern the aggregation of molecules into liquids or solids and dictate patterns, exemplified by dissolves like" , where polar compounds dissolve in polar solvents due to compatible dipole-dipole or hydrogen bonding interactions, while nonpolar substances favor dispersion-dominated solvents.

Properties and Reactivity

Physical Properties

The physical properties of chemical compounds encompass measurable characteristics such as , melting and boiling points, , , electrical conductivity, and spectroscopic features, which are determined by atomic arrangement and intermolecular interactions. These properties provide insights into a compound's under various conditions without altering its . The for a chemical compound at standard conditions depends on the strength of forces holding particles together; ionic compounds, characterized by strong electrostatic attractions between ions, typically exist as crystalline , while covalent molecular compounds often appear as gases or liquids due to weaker intermolecular forces. For instance, (NaCl), an ionic compound, is a at , whereas (CH₄), a covalent molecular compound, is a gas. diagrams graphically represent the between , , and gas phases as functions of temperature and pressure, showing transition points like the where all three phases coexist. Melting and boiling points reflect the energy required to overcome intermolecular forces or lattice energies; network covalent solids like silicon dioxide (SiO₂) exhibit exceptionally high melting points, around 1710°C, owing to extensive covalent bonding in a three-dimensional , whereas simple molecular compounds like have low boiling points, approximately -162°C (111.7 ), due to minimal van der Waals interactions. Ionic compounds generally display high melting points as well, though lower than network solids, because of the robust ion-ion attractions in their . Solubility in solvents like follows the principle that "like dissolves like," where polar or ionic compounds dissolve readily in polar solvents through interactions such as bonding; (C₂H₅OH), for example, is fully miscible with at all proportions. Nonpolar molecular compounds, however, show low in . For gases, is governed by , which quantifies the proportionality between the gas's over the liquid and its concentration in : S = k_H \cdot P, where S is , P is , and k_H is the constant. Density varies with packing efficiency and atomic masses; ionic solids often have higher densities than molecular solids because of their compact lattices, while metals can exhibit even greater densities due to close-packed structures. Electrical conductivity in ionic solids is negligible in the crystalline form due to fixed ions but increases dramatically when melted or dissolved, as mobile ions carry charge; molecular compounds are typically nonconductive unless ionized. Metals, by contrast, conduct electricity efficiently in solid form via delocalized electrons. Spectroscopy serves as a key tool for identifying compounds by probing molecular vibrations and electronic transitions; () spectroscopy detects bond stretching and bending in the 4000–400 cm⁻¹ range, revealing functional groups like C=O at around 1700 cm⁻¹. Ultraviolet-visible (UV-Vis) spectroscopy, operating in the 200–800 nm range, identifies electronic excitations in conjugated systems or transition metals, aiding in structural elucidation. These techniques, often used in tandem, confirm compound identity without destructive analysis.

Chemical Reactions and Stability

Chemical compounds undergo a variety of reactions that alter their composition, including formation through combination processes where simpler substances, often elements, unite to produce more complex structures. A representative example is the of from and oxygen gases via the $2H_2 + O_2 \rightarrow 2H_2O, which releases 285.8 kJ of per of formed under standard conditions. This reaction exemplifies direct from elements, highlighting how formation drives release and compound . Such processes are in both natural and industrial contexts, where elemental precursors combine under controlled conditions to yield stable products. Decomposition reactions represent the reverse, breaking compounds into simpler constituents and typically requiring energy input, rendering them endothermic. For water, electrolysis achieves this decomposition: $2H_2O \rightarrow 2H_2 + O_2, absorbing an equivalent amount of energy to that released during formation. Thermal decomposition provides another pathway, as seen in metal carbonates like calcium carbonate, which upon heating above 800°C yields calcium oxide and carbon dioxide: CaCO_3 \rightarrow CaO + CO_2. Displacement reactions, a common type, involve one element supplanting another in a compound, such as zinc displacing copper in copper(II) sulfate: Zn + CuSO_4 \rightarrow ZnSO_4 + Cu. Many such reactions, including displacements, are redox processes involving electron transfer between species, where oxidation states change as electrons move from reducing to oxidizing agents. The stability of chemical compounds arises from both thermodynamic and kinetic factors. Thermodynamically, the change (\Delta G = \Delta H - T\Delta S) determines reaction spontaneity; a negative \Delta G signifies a thermodynamically favorable, stable state at constant temperature and pressure, balancing enthalpy (\Delta H) and entropy (\Delta S) contributions. Kinetically, stability depends on the (E_a), the minimum energy barrier reactants must surmount for reaction to proceed; high E_a confers kinetic inertness, preventing rapid decomposition even if thermodynamically unstable. An illustrative case is (XeF₂), a stable noble gas compound synthesized in , which challenged assumptions of inertness and demonstrated conditional stability through strong Xe-F bonds despite xenon's low reactivity. These principles underpin practical applications, such as the Haber-Bosch process for industrial synthesis: N_2 + 3H_2 \rightleftharpoons 2NH_3, conducted at and temperature over an iron catalyst to overcome kinetic barriers and achieve yields sufficient for global production. Conversely, low reactivity poses environmental challenges, as seen in plastics like , which degrade slowly via photolysis and oxidation, persisting for decades to centuries in ecosystems due to high energies for .

References

  1. [1]
    Elements and compounds | CPD article - RSC Education
    Sep 3, 2018 · A compound is a substance formed from two or more elements chemically joined (bonded) together. A compound has a definite chemical composition ...
  2. [2]
    Atoms, Molecules, and Compounds - University of Hawaii at Manoa
    Molecules of compounds have atoms of two or more different elements. For example, water (H2O) has three atoms, two hydrogen (H) atoms and one oxygen (O) atom.
  3. [3]
    Chemical Compounds
    Chemical bonds are generally divided into two fundamentally different kinds: ionic and covalent. In reality, however, the bonds in most substances are neither ...
  4. [4]
    [PDF] Chemical Compounds
    Chemical Compounds. Chapter 3. Page 2. Molecular Compounds. Molecular Formulas exact number of atoms of each type in compound example - H. 2. O. ▫ inorganic ...
  5. [5]
    Chapter 1: Introduction to Organic Compounds and Covalent Bonding
    May 17, 2023 · Organic compounds are made from carbon and contain mostly hydrogen. They may also contain other elements such as oxygen, nitrogen, sulfur, ...
  6. [6]
    Ab Initio Machine Learning in Chemical Compound Space
    Aug 13, 2021 · Chemical compound space (CCS), the set of all theoretically conceivable combinations of chemical elements and (meta-)stable geometries that ...
  7. [7]
    EDITORIAL: Chemical Compound Space Exploration by Multiscale ...
    Aug 12, 2024 · According to various estimates, the number of possible chemical compounds is in the range of 1018–10200. A popular estimate of 1060 refers ...
  8. [8]
    What Are Elements and Compounds? | Journal of Chemical Education
    May 3, 2012 · The IUPAC definition of compounds is very broad, and it also seems to take issue with the concept that a compound has to have an exact whole ...
  9. [9]
    Scientist of the Day - Joseph Louis Proust, French Chemist
    Sep 26, 2018 · 26, 1754. In 1794, Proust published a paper, "Researches on Prussian blue," in which he first set down the law of definite proportions, or ...Missing: citation | Show results with:citation
  10. [10]
  11. [11]
    Review of Elements, Compounds, and Mixtures
    Elements are substances with one kind of atom, compounds are formed by combining elements, and mixtures have varying compositions and are easily separated.
  12. [12]
    CHE 111 Help - Elements, Compounds, and Mixtures - AACC OLA4
    Elements are pure substances with one type of atom. Compounds have multiple types of atoms combined. Mixtures contain more than one pure substance, retaining ...Missing: distinction | Show results with:distinction
  13. [13]
    CH105: Chapter 3 - Ionic and Covelent Bonding - Chemistry
    Nevertheless, the compound has properties completely different from either elemental sodium (a chemically reactive metal) or elemental chlorine (a poisonous, ...
  14. [14]
    3.1: Compounds Display Constant Composition
    Sodium is an extremely reactive soft metal that cannot be exposed to air or water. Chlorine is a deadly gas. The compound sodium chloride is a white solid which ...Missing: NaCl | Show results with:NaCl
  15. [15]
    John Dalton – the man and the myth - RSC Publishing
    Dec 23, 2021 · The seven rules essentially lead to the law of multiple proportions (also known as Dalton's law) which states that if two elements form more ...
  16. [16]
    Empedocles - Stanford Encyclopedia of Philosophy
    Sep 26, 2019 · Aristotle credits Empedocles with being the first to distinguish clearly these four elements (Metaphysics. 985a31–3). However, the fact that the ...
  17. [17]
    HISTORY OF SCIENCE - WEST - csbsju
    ARISTOTLE - around 350 BC. most famous Greek scientific philosopher, with an influence that lasted 2000 yr. ELEMENTS: developed Empedocles 4 elements - roots ...
  18. [18]
    Jabir ibn Hayyan - PMC - NIH
    Jabir is credited with the introduction of experimental methodology into alchemy and the invention of several chemical processes used in modern chemistry.Missing: Paracelsus compositions
  19. [19]
    What's in a Name? - Science History Institute
    Jābir saw alchemy as an experimental science based on his philosophical theory of balance (mīzān). For example, Jābir understood all metals to be a blend of the ...
  20. [20]
    Chemistry | Inters.org
    He expanded on the Arabic doctrine that two principles, sulphur and mercury, were the roots of all things by adding a third principle, salt. Paracelsus also ...<|separator|>
  21. [21]
    [PDF] Ancient and Historic Metals - Getty Museum
    Pliny lists five types of bronze alloys: (1) campana, an alloy used for vases and utensils; (2) an alloy similar to the previous one, used for the same purposes ...
  22. [22]
    Metals & Alloys - The Chemistry of Art
    Many of the early alloys were discovered by accident in the process of smelting metals that employed ores contaminated with other elements, but even in the ...
  23. [23]
    Selling the Chemical Universe - Mixing Medicines - NCBI Bookshelf
    Chemical medicine, also known as iatrochemistry or spagyric medicine, posed the earliest serious challenge to the centuries-old Galenic and Hippocratic ...
  24. [24]
    “Sooty Empiricks” and Natural Philosophers: The Status of ...
    Jul 30, 2010 · This article argues that during the seventeenth century chemistry achieved intellectual and institutional recognition, starting its ...
  25. [25]
    Elements and Atoms: Chapter 2 Robert Boyle, a Sceptical Chymist
    Minerals are examples of relatively pure compounds which were chemically analyzed in Boyle's time. Homogeneous mixtures such as the gases which comprise air ...
  26. [26]
    The Project Gutenberg eBook of The Sceptical Chymist, by Robert ...
    Transcriber's Notes. This e-book was prepared from a facsimile of the 1661 first edition and contains spelling, capitalization, and punctuation ...
  27. [27]
    In retrospect: The Sceptical Chymist - Nature
    Jan 5, 2011 · Robert Boyle's The Sceptical Chymist (1661) falls into the last category. Widely celebrated as a landmark, it remains misrepresented, misunderstood and unread.
  28. [28]
    Antoine Laurent Lavoisier The Chemical Revolution - Landmark
    For the first time, the Law of the Conservation of Mass was defined, with Lavoisier asserting that "... in every operation an equal quantity of matter exists ...
  29. [29]
    Law of Conservation of Matter (Antoine Lavoisier)
    In his famous textbook, Trait lmentaire de Chimie, which was published in 1789, Lavoisier reported that water was roughly 85% oxygen and 15% hydrogen by weight.Missing: compounds combination
  30. [30]
    Proust copper carbonate: teaching notes - Le Moyne
    What Proust did with these data was show that this synthetic copper carbonate had the same composition as copper carbonates derived from natural mineral sources ...Missing: definite 1794
  31. [31]
    Chapter 7 Atoms of Definite Weight: Dalton - Le Moyne
    By weight, the ratio of oxygen to hydrogen in water is 7.94:1; the ratio of nitrogen to hydrogen in ammonia is 4.63:1. Now Dalton decided to use ...
  32. [32]
    [PDF] THE TEACHING OF SCIENCE IN ENGLISH DISSENTING ...
    This thesis attempts a new assessment of the place of science teaching in the English Dissenting Academies. It examines the approach to science teaching, ...
  33. [33]
    [PDF] Isaac Watts : reason, passion and the revival of religion - CORE
    Isaac Watts was a dissenting minister, theologian, philosopher, hymn writer and poet in the first half of the eighteenth century. Despite exercising significant ...Missing: chemistry | Show results with:chemistry
  34. [34]
    The Structure of Atoms and the Octet Theory of Valence - PNAS
    BY IRVING LANGMUIR. RESEARCH LABORATORY, GENERAL ELECTRIC COMPANY, SCHENECTADY, NEW YORK. Read before the Academy, April 29, 1919. In a paper soon to be ...Missing: original | Show results with:original
  35. [35]
    The Rise of Instruments during World War II | Analytical Chemistry
    The NMR instrument was the offspring of the wartime development of radar. The story of Varian Associates―the company that made ...<|separator|>
  36. [36]
  37. [37]
    CH103 - Chapter 5: Covalent Bonds and Introduction to Organic ...
    Some pure elements exist as covalent molecules. Hydrogen, nitrogen, oxygen, and the halogens occur naturally as the diatomic (“two atoms”) molecules H2, N2 ...
  38. [38]
    11.6 The Solid State of Matter – Chemistry Fundamentals
    Covalent network solid include crystals of diamond, silicon, some other nonmetals, and some covalent compounds such as silicon dioxide (sand) and silicon ...
  39. [39]
    Types of Covalent Bonds: Polar and Nonpolar
    Electrons are shared differently in ionic and covalent bonds. Covalent bonds can be non-polar or polar and react to electrostatic charges.
  40. [40]
    Chemical bonds | Chemistry of life | Biology (article) - Khan Academy
    Covalent bonds are more common than ionic bonds in the molecules of living organisms. For instance, covalent bonds are key to the structure of carbon-based ...Missing: prevalence | Show results with:prevalence
  41. [41]
    Ionic Compounds | manoa.hawaii.edu/ExploringOurFluidEarth
    The 1+ charge of the sodium is balanced by the 1– charge of the chlorine. The resulting sodium chloride (NaCl) compound is neutrally charged (Fig. 2.24 C).
  42. [42]
    [PDF] Atomic and Electronic Structure of Solids - Physics Courses
    Left: the rock-salt, NaCl, structure, in which the ions form a simple cubic lattice with each ion surrounded by six neighbors of the opposite type. Right: the ...
  43. [43]
    [PDF] Chapter 12: Structures & Properties of Ceramics
    CsCl Type Crystal Structures. 939.0. 181.0. 170.0. Cl. Cs. = = -. + r r. Adapted from Fig. 12.3,. Callister & Rethwisch 8e. Cesium Chloride structure: ∴ Since ...
  44. [44]
    5.3 Molecular and Ionic Compounds - Maricopa Open Digital Press
    However, the formulas for some ionic compounds containing polyatomic ions are not empirical formulas. For example, the ionic compound sodium oxalate is ...
  45. [45]
    [PDF] 111 Exam I Outline 16'
    SOLUBILITY RULES FOR IONIC COMPOUNDS. Ion contained in the Compound. Solubility. Exceptions. Group IA. Soluble. NH4. +. Soluble. C2H3O2. -. Soluble. NO3. -.
  46. [46]
    Ionic Compounds Containing Polyatomic Ions
    Polyatomic ions are ions which consist of more than one atom. For example, nitrate ion, NO 3 -, contains one nitrogen atom and three oxygen atoms.
  47. [47]
    Table of Polyatomic Ions - gchem
    Table of Polyatomic Ions ; ammonium, NH · nitrite, NO ; phosphite, PO · phosphate, PO · hydrogen phosphate, HPO ; acetate acetate, CH3COO · C2H3O · carbonate, CO ...
  48. [48]
    8.9: Physical Properties of Ionic Compounds - Chemistry LibreTexts
    Mar 20, 2025 · This page discusses the distinct physical properties of ionic compounds, highlighting their high melting points, hardness, brittleness, ...Missing: formation authoritative
  49. [49]
    Physical Properties of Ionic Compounds - CK12-Foundation
    Ionic compounds have high melting points due to the strong ionic crystal lattices, for example, sodium chloride has a melting point of about 800°C. Ionic ...Missing: authoritative | Show results with:authoritative
  50. [50]
    Intermetallics. Structures, Properties, and Statistics - PMC - NIH
    Intermetallics are solid-state compounds exhibiting metallic bonding, defined stoichiometry and ordered crystal structure. It is now a century since it was ...
  51. [51]
    [PDF] INTERMETALLIC-BASED HIGH-TEMPERATURE MATERIALS ...
    INTERMETALLIC COMPOUNDS. An intermetallic compounds' can be defined as an ordered alloy phases formed between two metallic. elements,z where an alloy phase is ...
  52. [52]
    Bonding Analyses in the Broad Realm of Intermetallics
    Based on the Löwdin charges and ICOBI, the Cu–Hf and Hf–Hf contacts show a rather delocalized character, as expected for metallic bonds, while the nature of ...Introduction · Results and Discussion · Concluding Remarks · Author Information
  53. [53]
    Beyond Hume-Rothery Rules | Accounts of Materials Research
    Aug 18, 2023 · The Hume-Rothery rules for solid solution formation are described as follows: (i) a 15% similarity in size and (ii) a similarity in electrochemical nature must ...
  54. [54]
    Zintl Phases: From Curiosities to Impactful Materials
    Sep 4, 2023 · Zintl phases exhibit a combination of ionic and covalent bonding and are typically semiconductors. Zintl phases provide a palette for iso- and aliovalent ...What's in a Name? · Transition-Metal-Containing... · Rare-Earth Zintl Phases
  55. [55]
    mp-1038: MgCu2 (Cubic, Fd-3m, 227) - Materials Project
    MgCu₂ is Cubic Laves structured and crystallizes in the cubic Fd̅3m space group. Mg is bonded in a 12-coordinate geometry to twelve equivalent Cu atoms.
  56. [56]
    Physical And Mechanical Properties Of Intermetallic Compounds ...
    Nov 30, 2011 · In every solder joint, between the solder and the substrate, a layer is present containing one or more intermetallic compounds [1,2,3]. ... The ...
  57. [57]
    Superconductivity - American Magnetics Inc.
    Type I superconductors are comprised of pure metals, whereas Type II superconductors are comprised primarily of alloys or intermetallic compounds.
  58. [58]
    Alfred Werner: the well-coordinated chemist - Chemistry World
    Feb 25, 2013 · Alfred Werner's careful experiments led to the discovery of what came to be known as coordination bonds. Mike Sutton looks at how he pointed the way to ...
  59. [59]
    [PDF] Alfred Werner s Coordination Chemistry - Chem 251
    Oct 11, 2011 · [16] Only after his group proved the chirality of certain coordination compounds did his theory finally become accepted, and Werner was awarded ...
  60. [60]
  61. [61]
    Nomenclature of Coordination Compounds | CK-12 Foundation
    Write the chemical formulas for the following coordination compounds: tetraamminecopper(II) sulfate, sodium hexachloridocobaltate(III). A coordination compound ...
  62. [62]
    Coordination Compounds | CK-12 Foundation
    Sep 30, 2025 · Coordination Number (CN): This is the total number of ligand donor atoms directly bonded to the central metal ion in a complex. Only sigma bonds ...
  63. [63]
  64. [64]
    Medicinal Applications of Coordination Complexes - IOPscience
    Coordination complexes are used as anti-cancer, anti-bacterial, and antivirus drugs, and are effective against diseases like Alzheimer's and malaria.
  65. [65]
    8.2 Ionic Bonding
    The electrostatic attraction energy between ions of opposite charge is directly proportional to the charge on each ion (Q 1 and Q 2 in Equation 8.1). Thus ...Missing: formula | Show results with:formula
  66. [66]
    Lattice Energy - Ionic Compounds
    The lattice energy of NaCl is the energy given off when Na + and Cl - ions in the gas phase come together to form the lattice of alternating Na + and Cl - ions ...
  67. [67]
    Lattice Energies in Ionic Solids
    The lattice energy of nearly any ionic solid can be calculated rather accurately using a modified form of Equation 8.1.
  68. [68]
    Hybridization - FSU Chemistry & Biochemistry
    For sp3 hybridized central atoms the only possible molecular geometry is tetrahedral. If all the bonds are in place the shape is also tetrahedral. If there are ...
  69. [69]
    [PDF] Chapter 9: Theories of Chemical Bonding
    In methane, one 2s and three 2p orbitals “mix,” or hybridize, forming four equal-energy hybrid orbitals (Figure 9.6). Figure 9.6 Orbital hybridization (sp3) in ...
  70. [70]
  71. [71]
    Complex Ions
    The coordination number is the number of places on the metal ion where ligands are bound. The bond between the metal ion and the ligand, where the ligand ...<|separator|>
  72. [72]
    7.4 Valence Bond Theory – Chemistry Fundamentals
    The strength of a covalent bond depends on the extent of overlap of the orbitals involved. Orbitals that overlap extensively form bonds that are stronger than ...
  73. [73]
  74. [74]
    [PDF] Chapter 11 Intermolecular Forces - MSU chemistry
    All weak intermolecular forces are called: van der Waals forces. Page 8. van der Waals Forces. Two major forms: • Dipole–dipole interactions. – Hydrogen bonding.Missing: pi- pi
  75. [75]
    Intermolecular Properties - Chemistry 301
    Define the three major intermolecular forces (IMF) that can exist in condensed phases: dipole-dipole, H-bonding, and dispersion (London, van der Waals, induced ...Missing: stacking | Show results with:stacking
  76. [76]
    Intermolecular π–π Stacking Interactions Made Visible - PMC - NIH
    Dec 4, 2020 · This simple experiment can be used to visually demonstrate the existence and the effect of noncovalent intermolecular π–π stacking interactions.
  77. [77]
    Physical and Chemical Properties – Chemistry - UH Pressbooks
    Familiar examples of physical properties include density, color, hardness, melting and boiling points, and electrical conductivity.
  78. [78]
    [PDF] Chemistry - University of North Georgia
    examples of physical properties include density, color, hardness, melting and boiling points, and electrical conductivity. We can observe some physical ...
  79. [79]
    CH105: Chapter 4 - The Shape and Characteristics of Compounds
    4.3 Shape of Ionic Compounds. In chemistry, an ionic compound is a chemical compound comprised ions held together by electrostatic forces, termed ionic bonding.
  80. [80]
    Ionic and Molecular Compounds and Bonding
    Chlorine is poisonous, but sodium chloride is essential to life; sodium atoms react vigorously with water, but sodium chloride simply dissolves in water. Fig.
  81. [81]
    Methane - the NIST WebBook
    Methane (CH4) has a molecular weight of 16.0425. It's also known as marsh gas, methyl hydride, and biogas. Its normal boiling point is 111.7 K.Normal boiling point · References
  82. [82]
    11.5 Phase Diagrams – Chemistry Fundamentals - UCF Pressbooks
    At 40 °C, water at low pressure is a vapor; at pressures higher than about 75 torr, it converts into a liquid. At -40 °C, water goes from a gas to a solid as ...
  83. [83]
    [PDF] SRM - SAFETY DATA SHEET
    May 22, 2015 · Melting point/freezing point (ºC):. 1710 (3110 ºF). Density (g/mL):. 2.32 as Specific gravity. Vapor Pressure (mmHg):. 0 at 20 ºC. Vapor Density ...
  84. [84]
    Ethanol | CH3CH2OH | CID 702 - PubChem - NIH
    Ethanol | CH3CH2OH or C2H6O | CID 702 - structure, chemical names, physical and chemical properties, classification, patents, literature, ...
  85. [85]
    [PDF] THE SOLUBILITY OF GASES IN LIQUIDS
    Henry's Law is a limi~ing law but can sometimes be used for converting solubility data from the experimental pressure to a partial gas pressure of 1 atmosphere, ...
  86. [86]
    Conductivity (Electrical Conductance) and Water - USGS.gov
    In chemical terms, salts are ionic compounds composed of cations (positively charged ions) and anions (negatively charged ions).
  87. [87]
    [PDF] Ionic conductivity of solid mixtures - NASA Technical Reports Server
    The electrical conductivity of several ionic salts, both in the solid and liquid state, have been investigated as a function of temperature (refs. 1 and 2).
  88. [88]
    Infrared Spectroscopy - MSU chemistry
    Infrared spectrometers, similar in principle to the UV-Visible spectrometer described elsewhere, permit chemists to obtain absorption spectra of compounds.
  89. [89]
    UV-Visible Spectroscopy - MSU chemistry
    Visible wavelengths cover a range from approximately 400 to 800 nm. The longest visible wavelength is red and the shortest is violet.
  90. [90]
    Energy Changes Accompanying Chemical Reactions
    The reverse reaction, the formation of one mole of water from hydrogen and oxygen (equation d), is exothermic and releases 285.8 kJ energy.
  91. [91]
    Electrolysis of Water - Energy Foundations for High School Chemistry
    Therefore, the separation of water into its elements—the reverse reaction—must be endothermic, the quantity of energy equal to the quantity released when water ...
  92. [92]
    The Thermal Stability of the Nitrates and Carbonates
    Jun 30, 2023 · All the carbonates in this group undergo thermal decomposition to the metal oxide and carbon dioxide gas. The term "thermal decomposition" ...
  93. [93]
    23.1: Direct Redox Reactions - Chemistry LibreTexts
    Mar 20, 2025 · When a strip of zinc metal is placed into a blue solution of copper (II) sulfate (figure below), a reaction immediately begins as the zinc strip ...
  94. [94]
    Oxidation-Reduction Reactions - Chemistry LibreTexts
    Aug 29, 2023 · An oxidation-reduction (redox) reaction is a type of chemical reaction that involves a transfer of electrons between two species.
  95. [95]
    19.5: Gibbs Free Energy - Chemistry LibreTexts
    Jul 25, 2023 · Both ΔH and TΔS are temperature dependent, but the lines have opposite slopes and cross at 373.15 K at 1 atm, where ΔH = TΔS. Because ΔG = ΔH − ...
  96. [96]
    6.2.3.3: The Arrhenius Law - Activation Energies
    Feb 12, 2023 · In other words, the higher the activation energy, the harder it is for a reaction to occur and vice versa.Effects of Enzymes on... · Activation Enthalpy, Entropy... · Calculation of Ea using...
  97. [97]
    (PDF) Neil Bartlett and the First Noble-Gas Compound - ResearchGate
    Aug 7, 2025 · It was first synthesized in 1962 in a landmark experiment by Neil Bartlett that finally disproved the dogmatic assumption that noble gases ...
  98. [98]
    The Haber Process - Chemistry LibreTexts
    Jan 29, 2023 · The process combines nitrogen from the air with hydrogen derived mainly from natural gas (methane) into ammonia.
  99. [99]
    Degradation Rates of Plastics in the Environment - ACS Publications
    Feb 3, 2020 · This Perspective summarizes the existing literature on environmental degradation rates and pathways for the major types of thermoplastic polymers.