Fact-checked by Grok 2 weeks ago

Catenation

Catenation is the property by which atoms of an element form covalent bonds with other atoms of the same element to create long chains, rings, or other extended structures. This phenomenon is most pronounced in carbon, where it enables the formation of diverse hydrocarbons and the vast array of compounds essential to life and , due to the strength of carbon-carbon bonds and carbon's tetravalency. Other elements, such as , , , and , also exhibit catenation to varying degrees, though typically forming shorter chains or more limited structures compared to carbon. For instance, silicon forms polymeric silicates with Si-O-Si linkages but has weaker Si-Si bonds, limiting pure silicon catenation; sulfur creates cyclic S8 rings and chains; and phosphorus yields structures like red phosphorus with P-P bonds. The tendency for catenation depends on factors like , atomic size, and , with smaller atoms and stronger homonuclear bonds favoring longer chains. In carbon, the high C-C (approximately 348 kJ/mol) and ability to form stable multiple bonds (double and triple) allow for nearly limitless chain lengths, underpinning the structural diversity of molecules from simple alkanes like (C2H6) to complex polymers. In contrast, heavier group 14 elements like have larger atomic radii, leading to longer, weaker bonds that destabilize extended chains beyond a few atoms. Catenation in non-carbon elements often results in inorganic polymers or allotropes, such as and for carbon, or catenated in chemistry, influencing material properties like conductivity, , and reactivity. This property has profound implications across chemistry: in , it explains the existence of millions of carbon-based compounds; in , it contributes to the formation of semiconductors (e.g., the covalent Si-Si network in elemental ). Research continues to explore catenation in heavier elements and synthetic constructs, such as catenated metal-organic frameworks, to develop with tailored properties.

Fundamentals of Catenation

Definition

Catenation is the process by which atoms of the same form covalent bonds with one another, resulting in the creation of chains, rings, or extended networks, a property generally exhibited by elements possessing a of at least two. This self-linking capability arises from the strength and stability of homonuclear covalent bonds, allowing atoms to connect in various configurations without involving different elements. In contrast to , which entails the bonding of distinct molecular monomers to produce macromolecules, catenation is confined to the formation of homonuclear linkages solely among identical atoms. This distinction highlights catenation as a fundamental atomic-level phenomenon rather than a between pre-formed units, though it can contribute to the formation of polymeric structures in certain contexts. The structural manifestations of catenation encompass linear chains, as seen in conceptual polyatomic sequences; cyclic formations, such as closed rings; and more complex extended networks, including two-dimensional layers or three-dimensional frameworks. These arrangements provide a basis for diverse molecular architectures, particularly in elements capable of multiple bonding modes. The concept of catenation was first systematically explored in the context of during the mid-19th century, with key contributions from chemist in his seminal 1858 publication on the constitution and transformations of chemical compounds. Primarily associated with p-block elements, catenation underpins much of the structural complexity and reactivity diversity in both inorganic and domains.

Key Factors

The extent of catenation is primarily determined by the strength of homonuclear bonds (E–E) relative to heteronuclear bonds (E–X, where X is another element like hydrogen). High E–E bond energies provide the enthalpic driving force for chain formation, as the cumulative stability from multiple such bonds often outweighs the energy of alternative bonds, even if individual E–E bonds are slightly weaker. For example, the C–C bond dissociation energy is 348 kJ/mol, compared to 413 kJ/mol for C–H, yet long carbon chains are stable due to the overall lattice or molecular energy minimization./Chemical_Bonding/Fundamentals_of_Chemical_Bonding/Bond_Energies) In heavier p-block elements, weaker E–E bonds, such as the Si–Si bond at 222 kJ/mol, reduce this favorability, limiting chain lengths./Chemical_Bonding/Fundamentals_of_Chemical_Bonding/Bond_Energies) Atomic size and electronegativity significantly influence catenation by affecting orbital overlap and bond polarity. Smaller atoms facilitate superior overlap of valence orbitals, leading to stronger sigma bonds essential for chains. The effectiveness of this overlap is quantified by the overlap integral for sigma bonds: S = \int \psi_A \psi_B \, d\tau where \psi_A and \psi_B represent the atomic orbitals of the bonding atoms, and higher S values correlate with stronger bonds; larger atomic radii result in more diffuse orbitals, diminishing S./Chemical_Bonding/Valence_Bond_Theory/Orbital_Overlap) Moderate electronegativity (typically 2.5–3.0) supports balanced covalent character without excessive ionicity, as seen in carbon ( 2.55), promoting self-bonding over ionic alternatives./21%3A_The_p-Block_Elements) Orbital hybridization enables the directional bonding required for extended catenation. In elements capable of sp³ hybridization, such as carbon, the four equivalent hybrid orbitals form strong, tetrahedrally arranged sigma bonds, allowing linear, branched, or cyclic structures with optimal 109.5° angles for minimal strain./Fundamentals/Hybrid_Orbitals/sp%253A_Hybridization) sp² and sp hybridizations further support planar or linear chains in unsaturated systems, enhancing versatility. Heavier elements with poorer hybridization efficiency exhibit reduced catenation due to less effective orbital mixing. Thermodynamic stability of catenated species depends on both enthalpic and entropic factors. The catenation enthalpy (ΔH_cat), defined as the energy change for E–E bond formation relative to monomeric or clustered forms, is more negative for elements with strong E–E bonds, favoring chains over discrete molecules. Entropy contributions arise from increased rotational and vibrational in flexible chains versus rigid clusters, further stabilizing extended structures at higher temperatures. However, for , ΔH_cat is less favorable than for carbon due to weaker Si–Si bonds and comparable Si–H strengths, making catenation thermodynamically marginal. Kinetic factors, including activation energies for bond cleavage and formation, influence the persistence of catenated structures. Catenated chains often face higher activation barriers for because breaking multiple E–E bonds requires overcoming cumulative energies, providing kinetic even if are close to . In contrast, monomeric forms may have lower barriers for association, but for elements with weak E–E bonds, these kinetic advantages diminish, leading to easier fragmentation. Periodic trends show catenation decreasing down p-block groups, driven by increasing atomic size and decreasing E–E bond strengths. As atomic radius grows (e.g., from C at 77 pm to Si at 117 pm), orbital overlap weakens, and bond energies drop (C–C 348 kJ/mol vs. Si–Si 222 kJ/mol), reducing the propensity for long chains. Electronegativity also declines, shifting toward metallic character and less covalent self-bonding. This trend is evident across groups 14–16, with carbon showing maximal catenation and heavier homologues limited to short chains or rings./21%3A_The_p-Block_Elements)

Catenation in Group 14 Elements

Carbon

Carbon exhibits the most pronounced catenation among all elements, forming extended chains, rings, and networks that underpin the vast diversity of compounds and materials. This ability arises from the strength and versatility of carbon-carbon bonds, enabling structures ranging from simple hydrocarbons to complex macromolecules. For instance, in polymers such as , carbon chains can extend to thousands of atoms, with molecular weights typically between 200,000 and 500,000 g/mol corresponding to approximately 14,000 to 35,000 carbon atoms per chain. Rings like (C6H6) demonstrate cyclic catenation, while three-dimensional networks, as in , connect each carbon to four others in a tetrahedral arrangement. The versatility of carbon catenation is closely tied to its hybridization states, which dictate the geometry and bonding in different structures. In sp3-hybridized carbons, as seen in alkanes like n-hexane (C6H14), tetrahedral geometry supports single C-C bonds forming linear or branched chains. Sp2 hybridization occurs in alkenes and aromatic compounds, such as the layered sheets in , where planar trigonal carbons form double bonds and delocalized π-systems. In sp-hybridized forms, like (HC≡CH), linear chains feature triple bonds, allowing for rigid, elongated structures. This adaptability enables carbon to construct a wide array of molecular architectures beyond simple chains. Carbon's allotropes represent extreme manifestations of catenation, showcasing its capacity for zero-, one-, two-, and three-dimensional bonding. Diamond consists of an infinite tetrahedral network of sp3-hybridized carbons, providing exceptional hardness due to the uniform C-C connectivity. features sp2-hybridized carbons in hexagonal layers, where strong in-plane catenation contrasts with weak interlayer van der Waals forces. Fullerenes, such as (C60), form closed polyhedral cages with 60 carbons linked by alternating single and double bonds, while carbon nanotubes roll sheets into seamless one-dimensional tubes, exhibiting lengths up to millimeters with diameters of a few nanometers. The thermodynamic favorability of carbon catenation stems from the high stability of C-C bonds, with a bond energy of approximately 83 kcal/mol, which surpasses many alternatives like C-O or C-N bonds in longevity and strength for extended structures. This stability drives the preference for catenated forms in both synthetic and natural systems, notably in biomolecules where carbon backbones form the structural core; for example, the polypeptide chains in proteins rely on sequential C-C and C-N linkages from amino acid residues, and the deoxyribose sugar in DNA provides a five-carbon chain linking phosphate groups. Although carbon also participates in inorganic catenates, such as the networked Si-C bonds in silicon carbide (SiC), its pure elemental forms dominate catenation phenomena. Recent advances include the 2016 synthesis of ultralong linear carbon chains exceeding 6,000 atoms stabilized within double-walled carbon nanotubes, with a 2024 study demonstrating low-temperature synthesis methods to enhance production scalability; density functional theory (DFT) calculations have confirmed their enhanced stability.

Silicon

Silicon displays moderate catenation, forming Si-Si bonds in linear, cyclic, and networked structures, though less extensively than carbon due to weaker bonding and greater reactivity. This property is exemplified in silanes (Si_nH_{2n+2}) and substituted derivatives like polysilanes, where chain formation is limited by thermodynamic instability. Unsubstituted silanes exhibit limited chain lengths, with stable linear compounds known up to n-hexasilane (Si_6H_{14}); longer chains decompose readily at . In contrast, polysilanes with alkyl or aryl substituents can achieve degrees of up to approximately 1000 atoms, as the bulky groups sterically hinder cyclization and oxidation, enhancing and stability. The Si-Si bond of 222 kJ/mol is significantly weaker than the Si-H bond at 318 kJ/mol, promoting thermal , hydrogen elimination, and tendencies toward cyclization or cross-linking above 200–300 °C. Representative structures include linear (H_3Si-SiH_3, Si_2H_6), the simplest catenated with a staggered conformation and Si-Si of about 235 pm; cyclic compounds like ((SiH_2)6, Si_6H{12}), a puckered ring analogous to but more reactive; and networked homoatomic frameworks in certain oligomers, distinct from the Si-O-Si linkages in siloxanes that define polymers. These structures highlight silicon's capacity for sigma-bonded chains, though practical applications often rely on substituted variants. Polysilanes find use as photoresists in microlithography for semiconductor fabrication, where UV irradiation cleaves Si-Si bonds to enable pattern transfer with resolutions below 100 nm, and as precursors for silicon carbide ceramics via pyrolysis. The extended Si-Si backbone facilitates sigma-electron delocalization, yielding semiconducting properties with band gaps of 3–4 eV and conductivities up to 10^{-2} S/cm upon doping, suitable for optoelectronic devices. This catenation-driven conjugation contrasts with carbon's pi-delocalization but enables unique photoconductivity. Catenation ability diminishes down Group 14 due to increasing atomic size and decreasing bond strengths, with forming primarily short chains like digermane (Ge_2H_6) and rarely beyond n=3 under ambient conditions. Recent advances include the organometallic synthesis of stable Si_{20} rings as silafulleranes encapsulating ions, achieved via of chlorosilanes and stabilization by silyl substituents, opening pathways to cluster nanomaterials.

Catenation in Group 15 Elements

Nitrogen

Nitrogen exhibits limited catenation compared to other group 15 elements, primarily forming short chains due to the inherent instability of extended nitrogen-nitrogen linkages. The most stable nitrogen species is the diatomic , featuring a strong , but catenation manifests in forms such as the linear azide ion (N₃⁻), which consists of three nitrogen atoms connected by two bonds. This ion is a key example of a short catenated structure, often encountered in salts like (NaN₃). Similarly, (HN₃) represents a catenated molecule with a linear N₃ unit attached to , known for its explosive properties and role as a precursor in nitrogen chemistry. The restricted tendency for catenation in nitrogen arises from the weakness of N-N single bonds and the preference for multiple bonds driven by nitrogen's high electronegativity. The N-N single bond energy is approximately 167 kJ/mol, significantly lower than the N=N double bond at 418 kJ/mol, making extended chains energetically unfavorable due to lone-pair repulsions and poor orbital overlap in single bonds. Nitrogen's Pauling electronegativity of 3.04 further promotes the formation of strong multiple bonds, such as the N≡N triple bond in N₂ (bond energy 942 kJ/mol), over weaker single-bonded chains. In contrast to phosphorus, which forms stable rings and chains owing to stronger P-P bonds (around 201 kJ/mol), nitrogen favors compact structures like triples, limiting catenation to brief sequences. Catenated nitrogen structures include both linear and cyclic forms, though extended polymers remain elusive. Linear polyazides, such as (AgN₃), feature polymeric chains of N₃ units in the solid state, with the ligand adopting a linear (N-N-N angle ~180°). Cyclic catenates are exemplified by tetrazoles, which contain four atoms in a five-membered ring fused with carbon, providing stability through . Unlike carbon's infinite chains, nitrogen's catenation does not extend to polymeric solids under ambient conditions. Applications of nitrogen catenation leverage these short, energetic structures in explosives and pharmaceuticals. Lead azide (Pb(N₃)₂) serves as a primary in detonators, valued for its and rapid decomposition to N₂ gas. In pharmaceuticals, catenated nitrogen appears in heterocycles like , a six-membered with two adjacent nitrogen atoms, used in drugs such as antihypertensives. These uses highlight the high-energy release from breaking weak N-N bonds. Recent advancements include the 2001 synthesis of the N₅⁺ cation, a rare pentanitrogen chain achieved through the reaction of N₂F⁺ with HN₃, demonstrating fleeting stability in ionic compounds. Computational studies in 2025 have predicted the stability of N₁₀ molecular crystals under , suggesting potential for polynitrogen phases above several GPa, though experimental realization remains challenging. Catenation ability decreases down the group, with heavier elements like forming layered structures analogous to black but with more metallic character.

Phosphorus

Phosphorus demonstrates a pronounced tendency for catenation, forming diverse allotropes and compounds characterized by stable P-P bonds that enable cyclic, chain-like, and layered structures. White phosphorus exists as discrete P₄ tetrahedra, in which each phosphorus atom connects to three others via single bonds in a tetrahedral , representing a basic cyclic catenated unit. Red phosphorus adopts an amorphous polymeric form comprising interconnected chains and rings, including abundant five-membered rings and more complex clusters that can extend to sizes involving over 100 atoms, contributing to its relative stability compared to the molecular white form. Black phosphorus features a layered orthorhombic structure with puckered sheets of catenated atoms arranged in continuous six-membered rings, akin to a two-dimensional network where intralayer P-P bonds predominate. The strength of the P-P single bond, approximately 201 kJ/mol, underpins phosphorus's robust catenation, allowing for extended structures that are far more stable than those of nitrogen, where the weaker N-N bond (167 kJ/mol) limits chain lengths to typically fewer than four atoms. This trend reflects the increasing catenation ability down Group 15, with phosphorus exhibiting the maximum propensity among nonmetallic members, while arsenic shows similar layered catenation but with more character. In polyphosphides, phosphorus forms notable catenated motifs, such as the P₇⁴⁻ rings in Na₃P₇, which adopt a planar cycloheptaphosphanide structure stabilized by alkali metal cations. Simple phosphanes illustrate linear catenation, as in (P₂H₄), a colorless featuring a direct P-P bond analogous to but with a longer due to phosphorus's larger atomic size. More complex polymeric forms include Hittorf's phosphorus (violet phosphorus), which consists of infinite tubular chains arranged in a monoclinic with pentagonal cross-sections formed by fused five- and seven-membered rings. These catenated structures have practical applications; for instance, the polymeric chains in red provide the controlled reactivity needed for safety matches, where it ignites upon without . In fertilizers, catenates in like elemental phosphorus are processed into phosphates essential for crop nutrition, enhancing plant growth through improved phosphorus availability in soil. Although the DNA backbone relies on P-O-P linkages in phosphate diesters for its polymeric chain, catenation via P-P bonds is prominent in synthetic and inorganic compounds rather than biological polymers. Recent advances include the theoretical and experimental insights into stable P₁₂ icosahedral clusters, which mimic fullerene-like structures and offer potential as building blocks for novel .

Catenation in Group 16 Elements

Sulfur

Sulfur exhibits one of the most versatile forms of catenation among the chalcogens, forming stable rings, chains, and polymers due to the relatively strong S-S single bonds with an energy of approximately 266 kJ/mol. This bond strength, combined with the availability of low-lying 3d orbitals that facilitate flexible bonding geometries, enables sulfur to create extended structures ranging from cyclic molecules to infinite chains. Unlike lighter elements like oxygen, sulfur's larger atomic size reduces bond strain in catenated forms, promoting linear and helical configurations over smaller rings. The most stable and common ring form is the S8 crown-shaped cyclooctasulfur, featuring eight sulfur atoms in a puckered with S-S bond lengths of about 2.06 and bond angles near 108°. In the liquid state above its , sulfur undergoes to form catenasulfur, consisting of helical chains that can reach lengths up to 10^6 atoms, contributing to the high of molten . These chains represent a with smaller rings and ends, allowing for reversible catenation. Polymeric sulfur, denoted as S∞, features infinite linear or helical chains that persist in the solid phase, providing elasticity and strength to the material. Sulfur's allotropes highlight its catenative diversity: rhombic sulfur (α-S8) is the stable orthorhombic form composed of stacked S8 rings, while monoclinic sulfur (β-S8) adopts a similar cyclic structure but with different crystal packing, stable between 95.5°C and 119°C. Plastic sulfur, formed by rapid quenching of molten , consists of entangled long chains that impart a fibrous, rubber-like , though it slowly reverts to S8 over time. Additional catenated structures include open-chain sulfanes such as H2S7 and ions like Na2Sx, where x typically ranges from 2 to 5, forming flexible anionic chains. These ionic species are key in electrochemical systems. Catenation in sulfur underpins practical applications, particularly in . In vulcanized rubber, short sulfur chains (S cross-links) bridge polymer strands, enhancing durability and elasticity by restricting chain slippage. This process, discovered in the , relies on sulfur's ability to form multiple S-S bonds. In lithium-sulfur batteries, long catenated sulfur chains in copolymer cathodes improve capacity retention (up to 1005 mAh/g over 100 cycles) and cycling stability by mitigating shuttling. The viscoelastic properties of polymeric sulfur, arising from entangled long chains, further enable its use in high-impact composites. Recent advances include the synthesis of S12 rings under , which remain stable at ambient conditions, expanding the range of accessible cyclic allotropes.

Oxygen, Selenium, and Tellurium

Oxygen displays limited catenation, primarily forming diatomic O₂ molecules and the triatomic ozone (O₃), which features a bent structure with an O-O-O ring-like arrangement. Longer chains are unstable due to the weak O-O single bond energy of 142 kJ/mol and oxygen's high electronegativity of 3.44 on the Pauling scale, which favors multiple bonding and oxidation states over extended homoatomic linkages. Brief examples of catenation appear in peroxides, such as hydrogen peroxide (H₂O₂), where an O-O single bond connects two oxygen atoms, but these compounds are prone to decomposition and do not form stable polymers. Selenium exhibits moderate catenation, forming cyclic Se₈ rings in its monoclinic allotrope and infinite helical chains in gray (hexagonal) , the most stable form. Polyselenide anions, such as those in salts like Na₂Seₙ, can extend up to chains of about 10 selenium atoms, though longer structures are less common than in analogs. The Se-Se energy of 172 kJ/mol supports these structures but is weaker than the S-S bond, limiting extensive polymerization. Tellurium shows even weaker catenation tendencies, with elemental tellurium primarily adopting a structure of infinite helical chains in its hexagonal allotrope, resembling gray selenium but with more metallic character due to weaker interchain interactions. Discrete rings like Te₈ appear in certain polycationic compounds, such as [Te₈]²⁺, but these are not dominant in the bulk element. Polonium, the heaviest chalcogen, exhibits negligible catenation, behaving as a post-transition metal with a simple cubic lattice and no significant homoatomic bonding chains. In Group 16, catenation peaks at , is limited in the lighter oxygen, and decreases in the heavier elements , , and , primarily due to increasing atomic size, which weakens M-M s and favors over covalent chains in heavier members. This trend highlights homo-catenation's reliance on optimal bond strength and minimal steric repulsion, with oxygen's high further restricting chain formation compared to less electronegative congeners. Applications of catenated forms include oxygen's role in peroxides for bleaching and disinfection, leveraging the reactive O-O bond. For , amorphous polymeric forms—consisting of disordered chains—serve as photoconductors in photocopiers and xerographic imaging, where exposure to generates charge carriers for electrostatic transfer.

Catenation in Other Elements

Boron

Boron, a metalloid element, demonstrates catenation through the formation of B-B bonds in electron-deficient clusters and limited chain structures, primarily in boranes and borides, where bonding often involves multi-center interactions to compensate for its electron deficiency. Unlike carbon's ability to form extended saturated chains, boron's catenation is constrained by its tendency to adopt polyhedral geometries stabilized by three-center two-electron (3c-2e) bonds. This electron-deficient nature arises from boron's three valence electrons, leading to structures that prioritize cluster stability over long linear polymers. In , cluster catenation is exemplified by (B₂H₆), which features a dimeric structure analogous to ethene but with two B-H-B bridging bonds instead of a direct B-B . These bridges consist of 3c-2e bonds, where each involves two atoms and one , allowing the molecule to achieve octet satisfaction despite overall ; the terminal B-H bonds are conventional two-center two-electron (2c-2e) bonds. Larger neutral boranes, such as closo-dodecaborate dianion ([B₁₂H₁₂]²⁻), form icosahedral clusters with 12 atoms at the vertices connected by 30 B-B edges, each approximately 1.77 Å long, exemplifying delocalized bonding across the for enhanced stability. Chain-like catenation appears in polyboranes such as (B₁₀H₁₄), which adopts an open nido structure with segments resembling short chains bridged by hydrogens, though the overall motif remains cluster-dominated. polymers, including polyborazylene ((BNH)ₙ), incorporate B-B linkages within a of B-N bonds, serving as precursors to ceramics via , where catenated units contribute to the framework's thermal resilience. The B-B strength, approximately 290 kJ/mol, supports these assemblies but is lower than carbon's C-C (348 kJ/mol), and relies on 3c-2e interactions to delocalize electrons across deficient sites. In metal borides, boron forms linear or extended structures; for instance, magnesium diboride (MgB₂) features quasi-two-dimensional boron sheets with catenated hexagonal rings, akin to but with partial ionic character from magnesium, enabling at 39 K. Cyclic catenation occurs in boron subhalides, such as the rhomboidal B₄ ring in certain vapor-phase species or stabilized complexes like B₄(NCy₂)₄, where puckered four-membered rings exhibit alternating B-B bond lengths around 1.7-1.8 Å, stabilized by or ligands. These catenated boron structures underpin practical applications, notably in boron neutron capture (BNCT), where polyhedral and carboranes deliver ¹⁰B isotopes to tumor cells for selective via the ¹⁰B(n,α)⁷Li reaction, achieving tumor-to-normal tissue ratios up to 20:1 in preclinical models. In high-temperature ceramics, borides like ZrB₂ and HfB₂, featuring extended B-B networks, serve as ultra-high-temperature materials (melting points >3000 °C) for hypersonic vehicles and re-entry shields, offering oxidation resistance up to 2000 °C. Recent advances include 2023 developments in carborane-based chain assemblies for , such as self-assembled monolayers cross-linked into boron-rich 2D sheets for enhanced thermal and shielding in electronics.

Hydrogen

Hydrogen primarily exists as the diatomic molecule H₂ under standard conditions, where two hydrogen atoms are linked by a strong with a dissociation energy of 436 kJ/mol. However, catenation— the formation of chains or networks of hydrogen atoms—manifests in rare polyatomic species, beginning with the H₃⁺, a triangular structure that serves as the simplest example of hydrogen catenation. This ion, with bond lengths of approximately 0.87 Å and D₃ₕ symmetry, arises from the reaction H₂⁺ + H₂ → + H and plays a pivotal role in chemistry by initiating chains leading to more complex molecules. Polyhydrides, such as those incorporating longer hydrogen chains, extend this concept but remain largely theoretical or observable only under extreme conditions. Exotic forms of hydrogen catenation appear in high-pressure solid phases, where molecular transitions to atomic or chain-like structures. For instance, phase IV of , proposed at pressures around 150–200 GPa and temperatures near 300 K, features low-symmetry arrangements dominated by zigzag chains of hydrogen atoms connected via symmetric bonds, as revealed by structure-searching simulations. Computational studies also predict hydrogen chains embedded in metallic hosts, such as in alkali polyhydrides, where hydrogen sublattices form extended one-dimensional structures stabilized by pressure-induced . These configurations contrast with ambient-pressure , highlighting catenation's dependence on extreme to overcome inherent instabilities. Despite the robustness of individual H-H bonds, extended hydrogen chains exhibit significant instability due to the atoms' low coordination numbers (typically 1–2 in chains versus 4 in three-dimensional networks) and high reactivity, which promote or recombination into diatomic molecules. This fragility limits observable catenation to transient or confined environments, such as discharges where H₃⁺ and higher ions form briefly before decaying. Notable examples include theoretical catena-hydrides in alkali metals, where polyhydrides like LiH₂ feature chain-like hydrogen motifs stabilized above 100 GPa via partial covalent bonding between H atoms, diverging from the typical ionic character of s-block hydrides. In astrophysical plasmas, such as those in molecular clouds, H₃⁺ acts as a for transient polymers, facilitating ion-molecule reactions that build organic precursors. This represents an anomaly for elements, where catenation is enabled by ionic charge delocalization promoting covalent H-H links, unlike the non-catenating salts common at ambient pressures. Recent simulations of superionic phases in -rich systems, such as 2024 studies on high-pressure ices, suggest potential for short H₅-like proton chains amid diffusive hydrogen networks, though experimental confirmation remains elusive.

Halogens

Halogens exhibit minimal catenation compared to other p-block elements, primarily forming stable diatomic molecules such as F₂, Cl₂, Br₂, and I₂ under standard conditions. These diatomic species dominate due to the relatively weak homonuclear bonds and high reactivity of the elements, limiting chain formation to short polyhalide anions in ionic environments. Catenation in extends beyond diatomics mainly in polyhalide ions, where additional halide ions coordinate to form short linear or bent chains. Representative examples include the linear [I₃]⁻, which consists of a central iodine atom bridged symmetrically to two terminal iodines, and analogous [Br₃]⁻ species. Longer chains occur up to [I₅]⁻, featuring a V-shaped or zigzag arrangement with nearly linear I–I–I units. These polyhalides form in solutions containing excess halide and molecular halogen, such as I₂ in solution yielding [I₃]⁻. The tendency for catenation decreases down the group, influenced by bond strength trends: the F–F is 159 /, weaker due to lone-pair repulsion, while the I–I is 151 / but limited by increasing and steric hindrance that destabilizes longer chains. and form shorter polyhalides like [Cl₃]⁻ and [Br₃]⁻, but extended structures are rarer for owing to its high and reactivity. Polyhalide structures are typically asymmetric chains with varying bond lengths, as seen in [I₃]⁻ where terminal I–I bonds are longer than the central one, reflecting partial multiple-bond character in the core. Homocatenation rarely forms cycles; however, exceptional cases like tetrameric (IF₄)₄ rings exist, though these involve mixed halogens and highlight the instability of purely homonuclear cyclic forms. The high reactivity of , particularly , contributes to the instability of catenated species, which decompose readily in the gas or aqueous solutions unless stabilized by counterions or solvents. Homo-catenation remains rare outside polyhalide ions, with interhalogen compounds providing some stabilization but not extending homonuclear chains significantly. Applications of halogen catenation include the use of [I₃]⁻ in iodine-based disinfectants, where the form in solutions like enhances solubility and antimicrobial efficacy against and viruses. The [I₃]⁻– complex, forming a inclusion compound, serves as a visual indicator for iodine detection in and medical testing. Recent advances include the 2011 synthesis of stable [Br₇]⁻ chains in ionic liquids, where pyramidal polybromide anions were isolated as discrete units in compounds like [(Ph)₃PBr][Br₇], demonstrating extended catenation for under controlled conditions. This work expanded the known structural diversity of polyhalides, enabling their study in non-aqueous media.

Semimetallic Elements

Semimetallic elements in the p-block, such as , , and , exhibit catenation through the formation of extended covalent networks that contribute to their semiconducting properties. adopts a diamond-like three-dimensional where each atom is tetrahedrally coordinated by four Ge-Ge bonds, forming a continuous covalent that underpins its use in . In contrast, and form layered structures; gray arsenic consists of puckered hexagonal sheets of As atoms linked by As-As bonds within layers, with weaker van der Waals interactions between layers, resulting in a rhombohedral . displays a similar rhombohedral arrangement with puckered layers of Sb atoms connected by Sb-Sb bonds, though the bonding becomes progressively more metallic down the group. maintains a related layered but with increased metallic character, where Bi-Bi bonds exhibit partial delocalization, diminishing the extent of catenation compared to lighter homologs. The bond strengths in these networks reflect the trend toward weaker catenation in heavier elements. The average Ge-Ge bond energy is approximately 188 kJ/mol, while As-As bonds are around 146 kJ/mol and Sb-Sb bonds about 121 kJ/mol, enabling stable semiconducting frameworks but limiting long-chain formation in bismuth due to its lower Bi-Bi bond energy. These energies support the formation of extended structures rather than discrete molecules, distinguishing semimetals from lighter nonmetals. In Zintl phases, such as KGe₁₇, catenation manifests as isolated chains or clusters of germanium atoms, where electropositive potassium cations balance the anionic Ge networks, highlighting polyanionic catenation in intermetallic compounds. Across the heavier p-block, catenation evolves from molecular clusters in lighter elements to two- and three-dimensional covalent networks with increasing metallic character, driven by larger atomic sizes and poorer orbital overlap. This shift enhances electrical conductivity while preserving semimetallic structures. Applications leverage these networks: germanium's catenated is central to devices due to its tunable bandgap and high . and contribute to through homoatomic bonding in alloys like Bi₂Te₃, where layered catenation aids low thermal conductivity and high Seebeck coefficients, though pure elemental forms emphasize network stability over compounds. Recent advances include the 2021 growth of stable gray –tin alloys via chemical vapor transport, yielding layered structures with enhanced bandgap properties that extend catenation to alloyed nanowires for potential .

References

  1. [1]
  2. [2]
    [PDF] Organic chemistry – sOme Basic PrinciPles and Techniques - ncert
    In the previous unit you have learnt that the element carbon has the unique property called catenation due to which it forms covalent bonds with other carbon ...
  3. [3]
    The Chemistry of Oxygen and Sulfur
    The tendency of an element to form bonds to itself is called catenation (from the Latin catena, "chain"). Because sulfur forms unusually strong S-S single bonds ...
  4. [4]
    Catenation - Chemistry LibreTexts
    Jun 30, 2023 · Catenation of chemical bonds leads to the formation of inorganic polymers. However, inorganic polymers are mostly solids in the form of crystals.
  5. [5]
    Catenation Definition and Examples in Chemistry - Science Notes
    Nov 22, 2022 · Catenation is the chemical linking of atoms of the same element into chains or rings. It is similar to polymerization, except the monomer units are atoms of an ...
  6. [6]
    Ueber die Constitution und die Metamorphosen der chemischen ...
    Ueber die Constitution und die Metamorphosen der chemischen Verbindungen und über die chemische Natur des Kohlenstoffs. August Kekulé. Heidelberg.
  7. [7]
    [PDF] Chemistry of s and p-block elements
    In general, compounds that are formed when hydrogen is reacted with non-metals are called covalent hydrides. The compounds share a covalent bond and are either ...<|control11|><|separator|>
  8. [8]
    Polyethylene - Chemistry LibreTexts
    Jan 22, 2023 · Linear polyethylene is normally produced with molecular weights in the range of 200,000 to 500,000, but it can be made even higher. ...Missing: length | Show results with:length
  9. [9]
    8.7.1: The Group 14 Elements and the Many Allotropes of Carbon
    Apr 27, 2023 · When subjected to extremes of heat and pressure, graphite converts to its more dense diamond allotrope, in which all the carbons are tetrahedral ...Missing: catenation | Show results with:catenation
  10. [10]
    5.2: Orbital Hybridization Theory - Chemistry LibreTexts
    May 8, 2020 · Alkanes are hydrocarbons where all the carbon atoms are sp3-hybridized, all bonds are single bonds, and all carbons are tetrahedral. Methane is ...Missing: sp2 catenation
  11. [11]
    Carbon Allotrope - an overview | ScienceDirect Topics
    Carbon allotropes: diamond, graphite, C60 molecule, and carbon nanotube. ... Fullerenes (C60) and carbon nanotubes have similar bonds to graphite but the ...<|separator|>
  12. [12]
    21.5: Group 14: The Carbon Family - Chemistry LibreTexts
    Jul 12, 2023 · Unlike carbon, catenated silicon hydrides become thermodynamically less stable as the chain lengthens. Thus straight-chain and branched ...
  13. [13]
    Carbon | Introduction to Biology - Lumen Learning
    The functional groups in a macromolecule are usually attached to the carbon backbone at one or several different places along its chain and/or ring structure.
  14. [14]
    One-dimensional carbon chains encapsulated in hollandite - Nature
    Oct 3, 2023 · Shi et al. established a synthesis route for preparing carbyne chains of more than 6000 atoms inside double-walled carbon nanotubes. First- ...
  15. [15]
  16. [16]
    Silane - ChemicalBook
    It is now known as a fact that the longest chain is Si6H14. Its chemical character is more active than alkane, for it's easy for Si-H bond to break than the C-H ...Missing: length | Show results with:length
  17. [17]
    Polysilanes - an overview | ScienceDirect Topics
    Polysilanes are chains, rings, and 3D network polymers of silicon that contain the Si single bond Si bond in the polymer backbone.
  18. [18]
    Common Bond Energies (D - Wired Chemist
    Common Bond Energies (D. ) ; Si-Si, 222, 233 ; Si-N · 355 ; Si-O · 452, 163 ; Si-S · 293, 200.
  19. [19]
    Disilane - the NIST WebBook
    Disilane · Formula: H6Si · Molecular weight: 62.2186 · IUPAC Standard InChI: InChI=1S/H6Si2/c1-2/h1-2H3 Copy. InChI version 1.06 · IUPAC Standard InChIKey: ...
  20. [20]
    Polysilane - an overview | ScienceDirect Topics
    Polysilanes have found applications as photoresists in microlithography and free radical photoinitiators in polymerization of unsaturated hydrocarbons.
  21. [21]
    Possible Products of the End-On Addition of N3- to N5+ and Their ...
    Because homoleptic polynitrogen compounds with more than three nitrogen atoms are nearly unknown and generally not very stable, we investigated possible ...Missing: catenation HN3
  22. [22]
    High-energy-density materials: synthesis and characterization of N5 ...
    Aug 7, 2025 · In the area of polynitrogen chemistry, the synthesis and characterization of the novel N5+ cation from N2F+ and HN3 is described. View. Show ...Missing: catenation | Show results with:catenation
  23. [23]
    22.5: Group 15: The Nitrogen Family - Chemistry LibreTexts
    Jul 12, 2023 · In group 15, nitrogen and phosphorus behave chemically like nonmetals, arsenic and antimony behave like semimetals, and bismuth behaves like a metal.Missing: catenation | Show results with:catenation
  24. [24]
    Why does nitrogen show catenation properties less than phosphorus?
    Oct 6, 2017 · Catenation is much more common in phosphorous compounds than in nitrogen compounds. This is because of the relative weakness of the N−N single bond.Missing: comparison | Show results with:comparison
  25. [25]
    Can Catenated Nitrogen Compounds with Amine-like Structures ...
    Aug 17, 2023 · 3.4. Bond Dissociation Enthalpy. Here, the bond dissociation enthalpy (BDE) is used to further describe the molecular thermodynamic stability.
  26. [26]
    The explosive potential of nitrogen compounds - Chemistry World
    Mar 16, 2011 · As well as being used to make explosives, tetrazoles are used in many industrial applications, from catalysis to medicinal chemistry. The ...Missing: catenation AgN3 polyazides pyridazine pharmaceuticals
  27. [27]
    Computational prediction of a stable all-nitrogen molecular crystal N10
    Aug 25, 2025 · However, as pressure increased, MH-N10 surpassed N8 in stability above 3.32GPa and outperformed the. V-phase bispentazole N10 above 7.04 GPa.
  28. [28]
    S8 properties
    The covalent bonding in octasulfur involves electron pair sharing between sulfur atoms with bond dissociation energies of approximately 265 kJ/mol for S–S bonds ...
  29. [29]
    [PDF] Elemental Sulfur - CalTech GPS
    Large, polymeric molecules occur both in the solid and liquid phase, while ions are formed only in ionic solution. Each group will be discussed in connection.
  30. [30]
    [PDF] 1.2 Elemental Sulfur - Thieme Connect
    Solid poly-catena-sulfur occurs in many forms: as rubbery S, plastic (x)S, lamina S, fibrous (w , k),, polymeric (F), and insoluble (w)S. Furthermore, ...
  31. [31]
    Thermochemistry of Sulfur-Based Vulcanization and of ... - MDPI
    This paper not only proposes a simplified mechanism of vulcanization and devulcanization but also proposes an analytical method to check the devulcanization ...
  32. [32]
    Inverse Vulcanization of Elemental Sulfur to Prepare Polymeric ...
    Feb 20, 2014 · These materials exhibit enhanced capacity retention (1005 mAh/g at 100 cycles) and battery lifetimes over 500 cycles at a C/10 rate.
  33. [33]
    Laboratory infrared spectra and fragmentation chemistry of sulfur ...
    Jul 15, 2024 · Here, we report the laboratory far-infrared spectra of sulfur allotropes and examine their fragmentation pathways.
  34. [34]
    Selenium Element Facts - Chemicool
    Structure: long, helical chains (crystalline hexagonal), Se8. rings (crystalline monoclinic). Hardness: 2.0 mohs. Specific heat capacity, 0.32 J g-1 K-1. Heat ...
  35. [35]
    [PDF] The Bicyclic Polyselenium Cation Se102+ in the Structure of Se10 ...
    May 8, 2010 · Bond lengths / Å: Se(1)–Se(2) 2.421(2), Se(1)–Se(6). 2.423(2), Se(1)–Se(7) 2.447(2), Se(2)–Se(3) 2.261(2), Se(3)–Se(4) 2.457(2), Se(4)–Se(5).
  36. [36]
    Selenium containing macrocycles: transformation between Se–N/Se ...
    The mild stimulating condition can be attributed to the low bond energy (172 kJ mol −1 ) and weak rupture force (1100 ± 300 pN) [8,82]. In addition, the ...Missing: Se8 Se10 gray
  37. [37]
    Structure and dynamics in amorphous tellurium and Te clusters
    Apr 9, 2012 · The stable forms show a systematic change from diatomic molecules (O) through rings, chains, and helices (S, Se, Te) to the only simple cubic ...Missing: polonium negligible
  38. [38]
    [Te8][NbOCl4]2 containing an infinite chain-like {[Te-Te-Te-(Te5)]2+ ...
    The novel catena-octatellurium(2+) cation is composed of Te(5) rings that are linked via Te(3) units [Te-Te = 2.6455 (18)-2.8164 (19) Å]. The composition and ...
  39. [39]
    Selenium, Tellurium and Polonium - ScienceDirect.com
    Selenium and tellurium share to a limited extent sulfur's great propensity for catenation (see allotropy of the elements, polysulfanes, halides, etc.).
  40. [40]
    Photoelectric Device - an overview | ScienceDirect Topics
    Although sulfur shows a great tendency toward catenation, selenium and tellurium show much less tendency in this regard. Some structures containing chains and ...
  41. [41]
    Photocopying - an overview | ScienceDirect Topics
    Later, selenium was replaced by more reliable synthetic amorphous polymeric films; here we have another major industrial application of amorphous (glassy) ...
  42. [42]
    Boron Hydride - an overview | ScienceDirect Topics
    The boron hydrides, or polyboranes, are the original cluster compounds as well as the first known family of electron-deficient, 'non-classical' covalent ...
  43. [43]
    Update 1 of: Chemistry of the Carba- closo -dodecaborate(−) Anion ...
    In their stable redox form, deltahedral closo boranes thus have the formula BnHn2–, and the twelve-vertex icosahedral cluster, B12H122–, is the most common.
  44. [44]
    [PDF] Properties of Borane, the BH3 Molecule - MIT
    Diborane, B2H6, is a colorless gas forming explosive mixtures with air ... Pauling suggesting the same structure with B-H-B bridge bonds (A Diborane ...
  45. [45]
    [PDF] Diborane – B2H6
    Consequently, two B–H–B bridging bonds are formed, each of which consists of two electrons (Figure 2), forming what are called three- center-two-electron bonds ...Missing: BHB | Show results with:BHB
  46. [46]
    Progress in three-dimensional aromatic-like closo-dodecaborate
    Oct 1, 2021 · Among the closed borane family, closo-[B12H12]2− (closo-dodecaborate) is an icosahedral anion cluster that is stable for a long time.
  47. [47]
    Quantum Mechanical Design and Structure of the Li@B 10 H 14 ...
    An innovative type of lithium decahydroborate (Li@B10H14) complex with a basketlike complexant of decaborane (B10H14) has been designed using quantum ...Missing: polyboranes | Show results with:polyboranes
  48. [48]
    Polymer-Derived Boron Nitride: A Review on the Chemistry ... - NIH
    Polyborazylene delivered BN by pyrolysis under argon or ammonia in the temperature range 900–1450 °C in excellent chemical (89%–99%) and ceramic yield (84%–93%) ...
  49. [49]
    Boron nitride fibers processed from poly(borazinylamine) solutions
    Polymer-Derived Boron Nitride: A Review on the Chemistry, Shaping and Ceramic Conversion of Borazine Derivatives. Materials 2014, 7 (11) , 7436-7459. https ...
  50. [50]
    Topological superconducting phase in high-T c superconductor MgB ...
    May 3, 2019 · MgB2 has the AlB2-type centrosymmetric crystal structure with the space group P6/mmm (191). As shown in Fig. 1a, it is a layered structure with ...
  51. [51]
    Structural stability of a bent tetra(amino)tetraborane ring across four ...
    Mar 13, 2025 · Here, we present a simple and efficient method for the preparation of the cyclic tetraborane B4(NCy2)4. Strikingly, its puckered tetraborane ...
  52. [52]
    Boron Neutron Capture Therapy: A Review of Clinical Applications
    Feb 26, 2021 · Boron neutron capture therapy (BNCT) is an emerging treatment modality aimed at improving the therapeutic ratio for traditionally difficult ...
  53. [53]
    The formulation of polyhedral boranes for the boron neutron capture ...
    The early promise of boron neutron capture therapy as a method for the treatment of cancer has been inhibited by the inherent toxicity associated with ...
  54. [54]
    Review on ultra-high temperature boride ceramics - ScienceDirect
    This paper comprehensively reviews the various routes/techniques, including the advanced ones, as adopted for the synthesis and densification of the di-borides.
  55. [55]
    Carborane Nanomembranes | ACS Nano - ACS Publications
    We report on the fabrication of a boron-based two-dimensional (2D) material via electron irradiation-induced cross-linking of carborane self-assembled ...
  56. [56]
    Bond Enthalpies - Chemistry 301
    (See Bond Energy table at the bottom of this page) A H-H bond enthalpy (BE) is 436 kJ/mol, a C=C bond is 602 kJ/mol, a C-C bond is 346 kJ/mol, and a C-H BE is ...
  57. [57]
    Trihydrogen is the most important ion you've never heard of - C&EN
    Jun 8, 2019 · This unearthly molecule consists of three protons arranged in an equilateral triangle, sharing two electrons among them.Missing: H3+ | Show results with:H3+
  58. [58]
    Mechanisms and time-resolved dynamics for trihydrogen cation (H3 ...
    Jul 5, 2017 · Formed via an efficient ion-neutral bi-molecular reaction, H2 + + H2 → H3 + + H, trihydrogen cations can be commonly found in hydrogen plasmas.
  59. [59]
    Superionicity, disorder, and bandgap closure in dense hydrogen ...
    Sep 1, 2021 · Structure-searching methods suggest a sequence of low-symmetry phases dominated by zigzag chains of symmetric hydrogen bonds: a triclinic P 1 ¯ ...
  60. [60]
    Synthesis of sodium polyhydrides at high pressures - Nature
    Jul 28, 2016 · Polyhydrides may provide chemical means to pre-compress hydrogen molecules and facilitate the creation of metallic superconducting hydrogen at ...
  61. [61]
    [PDF] Molecular Dynamics in Dense Hydrogen
    Jul 26, 1993 · Early studies of the atomic phase. [4] have shown a prominence of anisotropic structures with low coordination up to very high pressures (~ 100.Missing: IV | Show results with:IV
  62. [62]
    Prediction of Stable Ground-State Lithium Polyhydrides under High ...
    May 11, 2017 · Accurate electronic structure calculation with GW approximation indicates that LiH, LiH2, LiH7, and LiH9 are insulative up to at least 208 GPa, ...Missing: polyhydride | Show results with:polyhydride
  63. [63]
    Phase transition kinetics of superionic H2O ice phases revealed by ...
    Sep 23, 2024 · H2O transforms to two forms of superionic (SI) ice at high pressures and temperatures, which contain highly mobile protons within a solid ...Missing: H5 chains
  64. [64]
    Polyhalogen and Polyinterhalogen Anions from Fluorine to Iodine
    May 15, 2019 · This Review deals with the evolving field of polyhalogen chemistry, specifically polyhalogen anions (polyhalides).
  65. [65]
    Catenated compounds in Group 17—Polyhalides - ResearchGate
    The present chapter offers an overview of polyanions of the Group 17 as examples of catenation. The main knowledge obtained from the rich literature on ...
  66. [66]
    Thermodynamics of Iodine Solubility and Triiodide Ion Formation in ...
    Iodine release and recovery, influence of polyiodide anions on electrical conductivity and nonlinear optical activity in an interdigitated and interpenetrated ...
  67. [67]
  68. [68]
    Preparation of Triiodide Resin Using Potassium Iodide and ... - MDPI
    Triiodide resin has an antimicrobial effect on bacteria in water. In the traditional TR manufacturing method, potassium iodide (KI) and crystalline I2 are ...
  69. [69]
    18.3: Structure and General Properties of the Metalloids
    Oct 27, 2022 · There are several allotropes of arsenic with the most stable being layer like and containing puckered sheets of arsenic atoms. Each arsenic atom ...
  70. [70]
    Arsenic, Antimony and Bismuth - ScienceDirect
    Arsenic and antimony are classed as metalloids or semimetals, and bismuth is a typical β subgroup (post-transition element) metal like tin and lead.
  71. [71]
    22.8: The Other Group 15 Elements: P, AS, Sb, and Bi
    Sep 8, 2020 · In group 15, nitrogen and phosphorus behave chemically like nonmetals, arsenic and antimony behave like semimetals, and bismuth behaves like a metal.
  72. [72]
    [PDF] Zintl Phases - eScholarship
    Sep 26, 2023 · The Zintl concept provides a rationale based on structure and bonding that can be employed in designing new compounds with exciting techno-.
  73. [73]
    21.3: The Elements of Group 15 (The Pnicogens)
    Jul 4, 2022 · In group 15, nitrogen and phosphorus behave chemically like nonmetals, arsenic and antimony behave like semimetals, and bismuth behaves like a metal.Missing: catenation | Show results with:catenation
  74. [74]
    Semi-metals as potential thermoelectric materials | Scientific Reports
    Jun 29, 2018 · In this work, we propose semi-metals with large asymmetry between conduction and valence bands as an alternative class of thermoelectric materials.
  75. [75]
    Bandgap opening in layered gray arsenic alloy - ResearchGate
    May 30, 2025 · Herein, we report the growth of high quality gray arsenic–phosphorus–tin (g-AsPSn) alloys via a simple one-step chemical vapor transport process ...