Fact-checked by Grok 2 weeks ago

Inert-pair effect

The inert pair effect refers to the empirical observation that heavier elements in groups 13–17 of the periodic table often exhibit stable oxidation states two units lower than the group's maximum valence, as the ns² electrons in their valence shell tend to remain chemically inert and unshared in compounds. This phenomenon is most pronounced in post-transition metals and heavier p-block elements, where the preference for lower oxidation states increases down a group due to relativistic effects and poor shielding by inner d- or f-electrons, which raise the ionization energy of the ns² pair. The effect arises primarily from the decreasing bond energies down a group, as larger atomic sizes lead to longer bond lengths and weaker orbital overlap, making the energy cost of promoting ns electrons to form higher oxidation states unfavorable compared to the stabilization gained from additional bonds. For instance, in group 13, thallium(I) compounds like Tl₂O are more stable than thallium(III) ones, contrasting with aluminum's exclusive +3 state, while in group 14, lead(II) is favored over lead(IV), as seen in PbCl₂. In group 15, bismuth(III) dominates over bismuth(V), often featuring stereochemically active lone pairs that distort molecular geometries, such as in BiI₃. These trends not only explain anomalous stability in heavier elements but also influence structural chemistry, where the inert pair can act as a lone pair affecting bond angles and coordination numbers, as in the bent structure of SnCl₂.

Definition and Overview

Core Concept

The p-block elements are characterized by a valence electron configuration of ns^2 np^{1-6}, where the np electrons primarily determine the group , while the ns^2 pair can participate in bonding to achieve higher oxidation states. This configuration underpins the variable oxidation states observed in these elements, ranging from the group valence down to lower states as increases. The inert-pair effect describes the progressively greater stability of the ns^2 electron pair in heavier p-block elements, rendering it less available for covalent bonding and promoting the adoption of oxidation states two units lower than the expected group maximum. This trend is prominent in groups 13 through 16, with the preference for the lower oxidation state intensifying down each group for the heavier congeners. For instance, in group 13, the +1 state becomes more stable than +3 for the heaviest member. A primary outcome of the inert-pair effect is the unusual thermodynamic of these lower oxidation states relative to their lighter analogs, which alters the behavior and chemical reactivity of the resulting compounds. This stability often enhances the reducing power of lower-valent species while diminishing the viability of higher-valent forms, influencing overall compound properties in p-block chemistry.

Historical Development

The inert-pair effect was first recognized in the early through empirical observations of unusual stability in lower oxidation states of heavy p-block elements, particularly in studies of and during the and . Nevil Sidgwick introduced the term "inert pair" in 1927 to describe the reluctance of the 6s² in to participate in bonding, explaining the stability of Tl(I) over the expected +3 state, as detailed in his seminal work on electronic theories of valency. Similar patterns were noted in lead(II) compounds, where the +2 predominated despite the group valency of +4, attributing this to the inertness of the ns² electrons in heavier elements. By the mid-20th century, chemists began formalizing the concept and integrating it into broader , with key contributions emphasizing its role in p-block chemistry. In , Russell S. Drago provided a quantitative framework by analyzing promotion energies and bond dissociation energies, demonstrating how the inert pair contributes to the energetic preference for lower valences down the groups. This work helped embed the effect within periodic table patterns, highlighting its increasing prominence from aluminum to in group 13 and analogous trends in groups 14–16. The 1960s marked significant milestones in experimental confirmation, leveraging and to validate the effect's manifestations. Thermodynamic studies, building on Drago's approach, quantified the stability of lower oxidation states through and calculations for compounds like Pb(II) and Sn(II). Spectroscopic techniques, including early on tin and lead halides, revealed electronic environments consistent with inert ns² pairs, providing direct evidence of their non-participation in bonding. From the 1980s onward, refined the understanding, shifting from purely empirical observations to a theoretical model incorporating relativistic effects. Pioneering calculations by Pekka Pyykkö in 1979 demonstrated how relativistic stabilization of the 6s orbital enhances the inert-pair tendency in heavy elements. Kenneth S. Pitzer's concurrent work further linked this to broader chemical properties, such as the lanthanide contraction's influence on bonding. Subsequent studies in the 1990s and 2000s, including those by Peter Schwerdtfeger and Martin Kaupp, used to model and orbital contractions, solidifying relativity's role in the effect's evolution.

Theoretical Basis

Relativistic Influences

The inert-pair effect in heavy p-block elements arises primarily from relativistic influences, which become significant for atoms with atomic numbers Z > 50. In these elements, inner-shell electrons move at velocities approaching a substantial fraction of the , leading to an increase in their effective mass according to . This relativistic mass increase contracts the s-orbitals, as the higher effective mass pulls electrons closer to the , while p-orbitals experience a countervailing due to reduced and poorer shielding by the contracted s-electrons. These orbital distortions deepen the energy separation between the ns and np orbitals, with the ns orbitals stabilizing more than the np orbitals. Consequently, the energy gap ΔE = E(np) - E(ns) widens progressively down each group, making the ns² less available for bonding and promoting lower oxidation states. For instance, in group 13, this stabilization favors the +1 state over +3 in heavier elements. The implications of the , which incorporates into , quantify these effects; for (Z = 81), scalar relativistic contributions raise the 6s-electron binding energy by approximately 10–20%, enhancing the inertness of the 6s² pair. Supporting computational evidence comes from Dirac-Fock calculations, which demonstrate the inertness of the ns² pair in elements such as (Z = 83) and lead (Z = 82) by showing substantial relativistic stabilization of the s-orbitals relative to p-orbitals. These calculations confirm that the effect intensifies with increasing Z, aligning with observed chemical trends in the p-block.

Electronic Structure Factors

In heavier p-block elements, the inert-pair effect arises partly from challenges in orbital hybridization, where the ns and np orbitals exhibit increasingly disparate radial distributions down a group. The ns orbitals become more diffuse and penetrate less effectively into the bonding region, leading to reduced overlap with np orbitals and diminished s-electron participation in covalent bonds. This results in weaker hybridization for higher oxidation states, favoring structures where the ns² electrons remain non-bonding. For instance, in compounds like PbX₄, the sp³ hybridization is ineffective due to these size mismatches, stabilizing the +2 state instead. Ionization energy trends further contribute to the inert-pair effect by making the removal of ns² electrons progressively more costly in heavier elements. While first ionization energies generally decrease down a group due to increasing atomic size, the energy required to remove the ns² pair from the +1 ion (corresponding to the second and third ionization steps for achieving +3) shows an anomalous increase for the heaviest members, such as thallium. For group 13, the sum of the second and third ionization energies (required to remove the ns² pair from the +1 ion to achieve the +3 state) increases from 4524 kJ/mol for indium to 4849 kJ/mol for thallium, reflecting tighter binding of the 6s electrons despite the larger size, due to poorer shielding by inner d electrons and diffuse orbital character. This elevates the energetic penalty for promoting ns electrons to higher oxidation states. The preference for over covalent in lower oxidation states of heavy p-block elements also stems from their lower and larger ionic radii, which reduce the tendency for s-electron sharing. In these elements, the +1 or +2 states form more stable ionic compounds with electronegative ligands, as the large size accommodates the charge without significant strain, whereas higher states demand covalent bonding that is hindered by poor orbital overlap. This shift is evident in (I) halides, which are predominantly ionic, contrasting with the more covalent aluminum halides. Non-relativistic computational models, such as Hartree-Fock methods, quantify the contribution of these size and overlap effects to the inert-pair phenomenon, estimating that they account for approximately 20–30% of the overall inertness in 5p and 6p elements independent of relativistic influences. These calculations reveal that the energetic cost of s-p and formation decreases down the group due to diffuse orbitals, with promotion energies for to the +3 state around 541 kJ/mol outweighing average energies. Natural atomic orbital analyses further support this by showing reduced s-character in bonds of heavier analogs, like PH₃ compared to NH₃, emphasizing non-relativistic steric and overlap factors.

Chemical Manifestations

In Group 13 Elements

In group 13 elements, the inert pair effect manifests as a progressive stabilization of the +1 over the group oxidation state of +3 as increases, due to the reluctance of the ² electrons to participate in bonding. For the lighter elements and aluminum, the +3 oxidation state is overwhelmingly stable, as seen in their common compounds like AlCl₃ and Al₂O₃, where the 3s² electrons are readily ionized. In contrast, for heavier elements like , , and especially , the +1 state gains thermodynamic favor, with thallium exhibiting greater stability for Tl⁺ than Tl³⁺. This trend arises from the increasing and poor shielding by d and f electrons, which contract the ns orbital and strengthen the ns² pair. A key indicator of this shift is the standard for the Tl³⁺/Tl⁺ , E° = +1.25 V, which corresponds to a highly negative ΔG° ≈ -241 kJ/mol for the reduction Tl³⁺ + 2e⁻ → Tl⁺, rendering Tl³⁺ a strong oxidant that readily converts to Tl⁺ in . For aluminum, no analogous stable +1 state exists, and the for Al³⁺/Al is -1.66 V, highlighting the absence of significant inert pair influence. This thermodynamic preference explains the instability of Tl³⁺ compounds relative to Tl⁺ analogs; for instance, Tl₂O and TlCl are stable, while Tl₂O₃ decomposes to Tl₂O and O₂ upon heating. Illustrative examples include the stability of thallium(I) carbonate (Tl₂CO₃), which persists under conditions where the hypothetical thallium(III) carbonate (Tl₂(CO₃)₃) would decompose, reflecting the inert pair's role in favoring lower-valent species. Tl⁺ acts as a soft acid, preferring coordination with soft bases like or over hard ones, consistent with the polarized 6s² 's reduced availability for bonding. In compounds like TlCl, the Tl⁺ adopts coordination geometries where the lone pair is stereochemically inactive, maintaining ionic structures without distortion from the ns² electrons.

In Group 14 Elements

In group 14 elements, the inert-pair effect manifests as a progressive stabilization of the +2 oxidation state relative to the +4 state down the group, from carbon and silicon to germanium, tin, and lead. For carbon and silicon, the +4 oxidation state dominates due to the effective participation of all valence electrons in bonding, with +2 states being unstable and rarely observed. In contrast, germanium exhibits a viable but less stable +2 state alongside the preferred +4, while tin shows comparable stability for both +2 and +4 states, and lead strongly favors the +2 state, where the ns² electrons remain largely inert. This trend is exemplified in the relative stabilities of oxides: for lead, PbO (Pb²⁺) is thermodynamically more stable than PbO₂ (Pb⁴⁺), which decomposes readily or acts as a strong , reflecting the reluctance of the 6s² pair to participate in . Similarly, SnCl₂ adopts a pyramidal in the gas phase, where the on Sn²⁺ is stereochemically active, distorting the structure from trigonal planar and influencing its reactivity as a . In aqueous solutions, Pb²⁺ ions are highly stable and serve as mild reducing agents, readily oxidizing to Pb⁴⁺ only under specific conditions. The ease of reduction of Pb⁴⁺ to Pb²⁺ is quantified by the standard E° ≈ +1.69 V for Pb⁴⁺ + 2e⁻ → Pb²⁺, indicating that Pb⁴⁺ is a potent oxidant and underscoring the inert-pair stabilization of the +2 state. This property is industrially exploited in lead-acid batteries, where the stability of Pb²⁺ enables reversible electrochemical reactions between Pb, PbO₂, and PbSO₄ in electrolyte, powering applications from vehicles to uninterruptible power supplies.

In Group 15 Elements

In group 15 elements, the inert pair effect manifests as a decreasing stability of the +5 and increasing stability of the +3 down the group from and to and . For , the +5 state is stable in compounds such as nitrates (NO₃⁻) and (HNO₃), while forms stable PCl₅; for the heavier elements, +3 becomes preferred; forms stable SbCl₃ but SbCl₅ only under specific conditions, and exhibits stable BiCl₃ while BiCl₅ is highly unstable and decomposes readily. Representative examples illustrate this trend in hydrides and other compounds. (SbH₃) is less thermally stable than (PH₃), with the order of hydride stability decreasing as NH₃ > PH₃ > AsH₃ > > BiH₃ due to weakening M–H bonds. Sb³⁺ compounds, such as Sb₂O₃, are common and stable, while primarily occurs in the +3 in medicinal applications, including used for gastrointestinal treatments. This preference for +3 arises from bond energy trends, where the average bond dissociation energies for group 15–V bonds (e.g., to or ) weaken down the group, dropping approximately 20% from to and favoring retention of the ns² over involvement in +5 bonding. Sb(III) leverages its stereochemically active in catalytic applications, particularly oxidation reactions; for instance, Sb₂O₃–CuO nanocomposites efficiently catalyze the photooxidation of p-nitrophenol via mechanisms.

In Group 16 Elements

In group 16 elements, the inert-pair effect manifests more subtly than in groups 13–15, primarily influencing the relative stabilities of the +4 and +2 oxidation states relative to +6 as increases, due to the increasing reluctance of the ns² electrons to participate in bonding. For oxygen and , the +6 state dominates in compounds like (SO₄²⁻) and (H₂SO₄), with the -2 state prevalent in oxides and sulfides, showing minimal impact from the inert-pair effect. follows a similar , with stable +6 species such as SeF₆ and +4 in SeO₂, though the +4 state begins to gain slight thermodynamic favor. In , the +4 is highly stable (e.g., in TeO₂ and TeCl₄), while +6 compounds like TeF₆ exist but TeCl₆ is unstable, decomposing to TeCl₄ and Cl₂, reflecting the emerging inert-pair influence that weakens higher oxidation states. The +2 state appears in compounds such as TeCl₂, which is known but prone to into elemental and TeCl₄, underscoring the partial stabilization of the lower state. Tellurium(IV) compounds, like TeO₂, serve as mild oxidants in certain reactions, further highlighting the practical dominance of this state. Polonium exhibits the strongest inert-pair effect in the group, with the +2 state (e.g., PoCl₂) and +4 state (e.g., PoO₂) being the most stable, while +6 species like PoF₆ are rare and unstable. Polonium(II) compounds are uncommon due to the element's tendency toward higher states in solution but can be isolated and remain stable under inert atmospheres to avoid oxidation. The thermodynamic favorability of reducing +6 to +4 or +2 increases down the group; for instance, the enthalpy change for such reductions in tellurium and polonium is more exothermic compared to lighter chalcogens, driven by higher s-electron promotion energies (e.g., ~600 kJ/mol for heavier analogs) and weaker bonding in higher states. Polonium's alpha decay, with a half-life of 138 days for ²¹⁰Po, can indirectly affect observations of the +2 state by generating decay products that alter the chemical environment in samples.

Structural and Steric Consequences

Lone Pair Behavior

In compounds influenced by the inert pair effect, the ns² on heavier p-block elements, such as Tl⁺ and Pb²⁺, is tightly bound due to relativistic stabilization, which contracts the s-orbital and increases the experienced by these electrons. This binding renders the pair largely non-bonding, concentrating high near the while minimizing its participation in covalent , thereby exhibiting low chemical reactivity. The resulting stability enhances the persistence of lower oxidation states, distinguishing these systems from lighter congeners where the s electrons more readily engage in hybridization or . The ns² 's steric behavior varies between inactivity and activity depending on the coordination environment and ligand field. In many cases, it behaves as stereochemically inactive, occupying minimal space within the and allowing for relatively symmetric geometries without significant distortion, as the pair remains compact and directed away from ligands. However, when active, the lone pair exerts steric repulsion against bonding pairs, leading to asymmetric coordination and polyhedral distortions, such as hemidirected geometries in Tl⁺ and Pb²⁺ complexes. This duality arises from the pair's and its interaction with surrounding , influencing overall molecular symmetry without direct bonding involvement. Spectroscopic methods confirm the localization and behavior of the ns² lone pair. In ²⁰⁵Tl NMR studies of thallium(I) compounds, chemical shifts correlate directly with the stereochemical activity of the 6s² pair, with greater shielding observed when the lone pair is more isolated and less involved in orbital mixing, providing evidence of its non-bonding localization. Similarly, ²⁰⁷Pb NMR spectra of lead(II) species, such as in PbO and related oxides, reveal anisotropic chemical-shift tensors that reflect the pair's compact nature and its influence on the electronic environment, supporting the inert pair's role in stabilizing the +2 state through reduced s-electron delocalization. The inert ns² imparts reducing properties to the stabilized lower oxidation states by making promotion to higher valence states energetically unfavorable. For example, Tl⁺ acts as a mild , capable of slow oxidation to Tl³⁺ in aqueous environments, as the tightly held 6s² electrons resist involvement in further bonding. This reactivity underscores the pair's electronic isolation, enhancing the thermodynamic preference for the +1 state relative to +3, consistent with observed potentials across group 13.

Examples in Molecular Geometry

The inert-pair effect manifests in molecular geometries through the stereochemical activity or inactivity of the ns² in heavier p-block elements, leading to distinct coordination preferences. In cases where the is stereochemically inactive, it occupies space without significantly distorting the , resulting in higher symmetry geometries. For instance, (I) fluoride (TlF) adopts an orthorhombic (distorted rock salt type) where each Tl⁺ is coordinated to six F⁻ at approximately 2.6 Å, consistent with the inactive 6s² allowing for higher coordination without substantial distortion. Similarly, lead(II) compounds often avoid square planar geometries typical of d⁸ or d¹⁰ analogs, favoring instead lower coordination numbers or hemidirected bonding patterns due to the inactive 6s² , which minimizes steric repulsion without active distortion. In contrast, when the inert-pair lone pair is stereochemically active, it exerts repulsive forces that distort the molecular geometry, akin to a VSEPR electron pair. A classic example is tin(II) chloride (SnCl₂), which adopts a pyramidal AX₂E geometry in the gas phase, with a bent structure around the Sn²⁺ center (Cl–Sn–Cl angle ~95°), driven by the active 5s² lone pair occupying an equatorial position in the trigonal pyramidal arrangement. This distortion is evident in the solid state as well, where polymeric chains form with trigonal-pyramidal coordination featuring short terminal Sn–Cl bonds (~2.3 Å) and longer bridging ones (~2.7 Å). For antimony(III) chloride (SbCl₃), the active 5s² lone pair induces stereochemical distortion in its orthorhombic structure, resulting in trigonal pyramidal coordination (AX₃E) with Sb–Cl bond lengths of ~233 pm and Cl–Sb–Cl angles of ~97°. In solid-state structures, the inert-pair lone pair directs layered architectures with asymmetric coordination. Lead(II) oxide (PbO) in its tetragonal α-form (litharge) features a layered structure where each Pb²⁺ ion exhibits 4+2 coordination: four short Pb–O bonds (~2.23 Å) in the layer plane and two longer axial bonds (~2.48 Å) perpendicular to it, with the 6s² lone pair pointing out of the layer to avoid steric clashes and stabilize the puckered sheets. X-ray diffraction studies of bismuth(III) complexes further illustrate lone pair repulsion effects, often causing elongation and angular distortions. In dithiocarbamate complexes like [Bi(Me₂DTC)₃]₂, the active 6s² lone pair leads to distorted octahedral geometries with Bi–S bond angles compressed to 80–90° in the equatorial plane, reflecting repulsion that elongates axial bonds (up to 3.0 Å) and creates a "void" for the lone pair.

Deviations from the Trend

The inert pair effect is negligible in lighter p-block elements such as , carbon, , and oxygen, where relativistic contributions to the ns orbital contraction are minimal and the ns electrons readily participate in due to strong np orbital involvement. For these elements (Z < 30), the energy separation between ns and np orbitals remains small, typically less than 5 eV, which is insufficient to confer inertness to the ns² pair and promote lower oxidation states. Environmental factors can weaken or reverse the inert pair trend by stabilizing higher oxidation states. For instance, the highly electronegative fluorine ligand favors the +3 state in thallium despite the general preference for +1, as seen in the stability of TlF₃, where the strong Tl–F bonds compensate for the energetic cost of promoting ns electrons./08:_Chemistry_of_the_Main_Group_Elements/8.06:Group_13(and_a_note_on_the_post-transition_metals)/8.6.02:_Heavier_Elements_of_Group_13_and_the_Inert_Pair_Effect) Similarly, conditions with high oxidation potentials can enforce higher states across the p-block. Notable anomalies occur within groups where the expected stabilization of the lower state fails. In group 13, Ga(I) compounds are thermodynamically unstable and prone to disproportionation into Ga(0) and Ga(III) species, contrasting with the stable Tl(I) due to the less pronounced inert pair in lighter elements like gallium. In group 16, polonium deviates by favoring the +4 oxidation state over +2 in aqueous solutions, where Po(IV) is the most stable form, reflecting weaker inert pair influence compared to tellurium or selenium.

Comparisons with Other Effects

The inert-pair effect shares some underlying relativistic origins with the lanthanide contraction but differs in scope and manifestation within the periodic table. The lanthanide contraction arises from the poor shielding of 4f electrons combined with relativistic contraction of s and p orbitals, resulting in unexpectedly small atomic and ionic radii for elements in the 5d and 6d series following the lanthanides, which affects trends across d-block and f-block transitions. In comparison, the inert-pair effect is confined to heavier p-block elements (groups 13–16), where relativistic effects preferentially stabilize the valence ns² electrons, increasing the energy required to ionize them and favoring lower oxidation states (e.g., +1 for Tl, +2 for Sn and Pb). Although both phenomena involve relativistic s-orbital contraction, the lanthanide contraction primarily influences atomic sizes and intermetallic properties globally, whereas the inert-pair effect specifically alters valence electron participation and oxidation chemistry in the p-block. Unlike the inert-pair effect, which governs oxidation state stability through electronic stabilization, diagonal relationships in the periodic table stem from similarities in physical and chemical properties between elements positioned diagonally, such as and or Be and , due to comparable charge-to-radius ratios and electronegativities that counteract the usual group and period trends. The inert-pair effect provides an indirect explanation for certain anomalies in these relationships by promoting lower oxidation states in heavier analogs, which can mimic the chemistry of diagonally related lighter elements through similar effective ionic charges (e.g., the +3 state of resembling the +2 state of Be in polarizing power and reactivity). However, diagonal relationships are fundamentally driven by ionic size and electronegativity balances rather than relativistic s-electron inertness, and the inert-pair effect does not apply to the s-block pairs like where such relationships are most pronounced. Fajans' rules describe the tendency toward covalent character in ostensibly ionic compounds through polarization of anions by small, highly charged cations, emphasizing factors like cation size, charge, and anion polarizability to predict bond type without invoking electronic orbital specifics. The inert-pair effect complements this by generating lower-oxidation-state cations (e.g., Bi³⁺ instead of Bi⁵⁺) that possess higher charge densities due to the retained ns² pair, thereby increasing their polarizing power per Fajans' criteria and enhancing covalency in resulting compounds, as seen in the more covalent nature of PbCl₂ compared to PbCl₄. Nonetheless, the inert-pair effect is a distinct relativistic phenomenon focused on valence s-electron reluctance, whereas Fajans' rules operate on geometric and electrostatic principles applicable across the periodic table, independent of relativistic stabilization. In the modern context of superheavy elements (Z > 100), the inert-pair effect must be differentiated from the dominant role of spin-orbit coupling, which becomes exceptionally strong and inverts orbital energy orders (e.g., 7p_{1/2} below 7s in element 114). For these elements, the traditional inert-pair stabilization of ns² is overshadowed by spin-orbit splitting of np orbitals, leading to predicted chemical inertness and low oxidation states driven by j-coupled configurations rather than simple s-pair isolation, as exemplified by the enhanced stability of the +2 state in beyond p-block trends. This distinction highlights how spin-orbit effects in superheavies represent an extreme relativistic regime, extending but not synonymous with the inert-pair mechanism observed in lighter elements.

References

  1. [1]
  2. [2]
  3. [3]
    p-Block Elements - BYJU'S
    The general electronic external configuration for p block components is ns2np(1−6). (n−1)d(1−10)ns(0−2) is the general electronic outer configuration of d block ...Missing: ns2 | Show results with:ns2
  4. [4]
    P block Elements, Properties and Uses - Unacademy
    The valence shell configuration ranges from ns2 np1 to ns2 np6. Except for Helium, the general electronic configuration of p-block elements is ns2np1-6.
  5. [5]
    The Inert Pair Effect: An Analysis Using the Chemdex Database - MDPI
    Abstract. The presence of a so-called 'inert pair effect' in the chemistry of the 6p elements has been recognised for almost a century.
  6. [6]
    Frontiers in molecular p-block chemistry: From structure to reactivity
    Feb 1, 2019 · Whereas the inert pair effect makes species such as In+ stable and Lewis basic (nucleophilic), univalent group 13 elements such as B(I) and ...
  7. [7]
    Oxidation state trends in Group 4 - Chemguide
    Carbon monoxide is a strong reducing agent because it is easily ... This is often known as the inert pair effect - and is dominant in lead chemistry.
  8. [8]
    The Electronic Theory Of Valency : Nevil Vincent Sidgwick
    Oct 26, 2006 · The Electronic Theory Of Valency. by: Nevil Vincent Sidgwick ... PDF download · download 1 file · SINGLE PAGE PROCESSED TIFF ZIP download.
  9. [9]
  10. [10]
    [PDF] Relativistic effects on the chemistry of heavier elements
    Oct 31, 2023 · For example, by considering the effects of lanthanide contraction, inert pair effect observed for the heavier congeners of the p-block elements ...
  11. [11]
    [PDF] Relativity and the Periodic System of Elements
    Then the 6s-6p hybrid- ization is energetically less favorable and an ( 6 ~ ) ~. “inert pair” is formed. We are now able to attribute this. “relativistic ...
  12. [12]
    Relativistic Dirac-Fock expectation values for atoms with Z = 1 to Z ...
    This paper presents numerical relativistic Dirac-Fock calculations for atoms Z=1 to Z=120, including total atom energies, orbital binding energies, and ...Missing: inert | Show results with:inert
  13. [13]
  14. [14]
  15. [15]
    Relativistic Effects and the Chemistry of the Heaviest Main-Group ...
    The heaviest main-group elements (mercury through radon and their heavier congenors) often show markedly different chemical properties than their lighter ...
  16. [16]
    Thermodynamic Evaluation of the Inert Pair Effect - ACS Publications
    Thermodynamic Evaluation of the Inert Pair Effect. Click to copy article link ... Chemistry and Stability of Lead( IV ) Compounds. Angewandte Chemie ...Missing: volatility | Show results with:volatility
  17. [17]
    Group 14: General Properties and Reactions - Chemistry LibreTexts
    Jun 30, 2023 · The Group 14 elements tend to adopt oxidation states of +4 and, for the heavier elements, +2 due to the inert pair effect. carbon in periodic ...Introduction · Carbon · Tin · Lead
  18. [18]
    The following standard reduction potentials have been ... - Chegg
    Dec 12, 2022 · Question: The following standard reduction potentials have been determined for the aqueous chemistry of lead: Pb4+(aq) + 2e- Pb2+(aq) E° = 1.690 ...
  19. [19]
    Why BiCl5 isn't stable? - Chemistry Stack Exchange
    Feb 16, 2019 · Well, your inert pair effect makes the pair harder to use - one needs stronger oxidant to oxidate Bi (III) to Bi (V) and therefore Bi (V) ...Missing: bismuth BiCl3
  20. [20]
    [Odia] BiCl(3), is more stable than BiCl(5) Why? - Doubtnut
    Bi has little tendency to form pentahalides because +5 oxidation state of Bi is much less stable than +3 oxidation state due to inert pair effect.
  21. [21]
    Explain the thermal stability of hydrides of group 15 elements. - CK-12
    The thermal stability of the hydrides of group 15 elements decreases as we move down the group. The order of thermal stability is NH3 > PH3 > AsH3 > SbH3 > BiH3 ...Missing: PH3 AsH3 inert pair effect
  22. [22]
    Current and Potential Applications of Bismuth-Based Drugs - PMC
    In general the three 6p electrons are involved in bond formation and in the majority of compounds Bi is in the +3 oxidation state. Such compounds possess a so- ...
  23. [23]
    Recent advances in heavier group 15 (P, As, Sb, Bi) radical chemistry
    Feb 25, 2025 · This effect has been observed with the Pn–H bond dissociation enthalpies for the PnH3 series (Pn = P: 81.4; As: 74.6; Sb: 63.3; Bi: 51.8 kcal ...
  24. [24]
    Applications of Antimony in Catalysis | ACS Organic & Inorganic Au
    This Review will focus on the application of antimony in photocatalysis, electrocatalysis, and other organic syntheses in the past 10 years.
  25. [25]
    Tellurium dichloride - Wikipedia
    Tellurium dichloride (TeCl2) is unstable with respect to disproportionation. Several complexes of it are known and well characterized. They are prepared by ...Missing: oxidation | Show results with:oxidation
  26. [26]
    Stability Study of Hypervalent Tellurium Compounds in Aqueous ...
    Aug 10, 2017 · The C–Te bond of organotellurium compounds is commonly considered unstable, disfavoring their applicability in biological studies. In this study ...Missing: TeCl6 | Show results with:TeCl6
  27. [27]
    [PDF] MASTER The Radiochemistry of Polonium - OSTI.gov
    Quadrivalent Polonium The +4 oxidation state is the stable state in solutions. The dioxide can be formed directly from the elements and forms spontaneously when ...
  28. [28]
    Efficient Lone-Pair-Driven Luminescence: Structure–Property ...
    Following the inert pair effect, the oxidative stability of ns2 cations increases moving down the periodic table, with 6s2 cations such as Pb2+ and Bi3+ ...
  29. [29]
    [PDF] Chemistry, Structure, and Function of Lone Pairs in Extended Solids
    Jan 1, 2022 · Important properties associated with s2 electron-derived lone pairs include their role in creating conditions favorable for ion transport, in ...
  30. [30]
    [PDF] Chemistry, Structure, and Function of Lone Pairs in Extended Solids
    Important properties associated with s2 electron-derived lone pairs include their role in creating conditions favorable for ion transport, in the formation and ...
  31. [31]
    [PDF] Lone Pairs, Hidden and Otherwise - UCSB MRL
    Chemistry counts: The role of the anion in Pb2+: [Xe]5d106s2. Stereochemically “active” lone pair: PbO. Stereochemically “inactive” or “silent” lone pair: PbS ...Missing: Tl+ | Show results with:Tl+
  32. [32]
    Structural effects of the Pb2+ 6s2 lone pair activity: Eccentricity
    Apr 15, 2025 · The stereochemical activity of the 6s2 lone pair has profound consequences on the coordination chemistry of metal ions like Tl+, Pb2+ and Bi3+.
  33. [33]
    Effect of Stereochemically Active Electron Lone Pairs on Magnetic ...
    Aug 2, 2023 · The results demonstrate the distinctive effects of stereochemically active lone pairs in modifying electronic structure near the Fermi level and for mediating ...
  34. [34]
    Correlation between the covalency and the thallium-205 nuclear ...
    A correlation between the chemical shift and the stereochemical activity of the 6s2 lone pair of TlI was established; the greater this activity, the greater the ...
  35. [35]
    NMR of minium, Pb 3 O 4 : evidence for the [Pb 2 ] 4+ ion and ...
    Solid Pb3O4 has been studied with 207 Pb nuclear magnetic resonance (NMR) spectroscopy. The 207 Pb NMR chemical-shift tensor of the Pb2+ site has principal ...
  36. [36]
    Periodicity – Chemistry - UH Pressbooks
    In general, the inert pair effect is important for the lower p-block elements. ... The stability of the 3+-oxidation state is another example of the inert pair ...
  37. [37]
    The Coordination Chemistry and Clustering of Subvalent Ga + and ...
    May 17, 2021 · The elements of the group 13 share a ns2np1 electron configuration ... ns-np separation between occupied ns2 HOMO and np0 LUMO in kJ ...
  38. [38]
    Insights into the Composition and Structural Chemistry of Gallium(I ...
    Oct 6, 2020 · ... gallium chemistry, each of which has advantages and disadvantages. True GaI halides are thermodynamically unstable, and while they can be ...
  39. [39]
    Review of Chemical and Radiotoxicological Properties of Polonium ...
    At zero oxidation degree, polonium is a noble metal, thermodynamically stable in aqueous solution in the absence of oxidant.
  40. [40]
    Perspective: Relativistic effects - AIP Publishing
    Apr 19, 2012 · fects, creating the 6s “inert pair” effect.15 A large number of such ... lanthanide contraction may be attributed to relativity. For de ...
  41. [41]
    RELATIVISTIC EFFECTS IN CHEMICAL SYSTEMS - Annual Reviews
    Jan 2, 2025 · For mercury one first notes that the anomalous volatility of the element is just the inert pair effect discussed above. ... Pitzer in 1979 ...
  42. [42]
  43. [43]
    [PDF] Chemistry of s and p-block elements
    They form coloured complexes. Relative stability of different oxidation states. 1. For group-1 elements the common and stable oxidation state is +1. The ...<|control11|><|separator|>
  44. [44]
    [PDF] p-BLOCK ELEMENTS GROUP-IIIA (13) BORON FAMILY
    is called inert pair effect. The inert pair effect increases gradually in ... in the covalent character (Fajans rules). Hence, in boron,. +3 oxidation ...
  45. [45]
    [PDF] 1 Chemical Bonding of Main-Group Elements - Wiley-VCH
    May 26, 2014 · The nevertheless reasonably inert character of elemental mercury is the consequence of an inert-pair effect also in the possible reaction ...
  46. [46]
    Relativity and the periodic table - Journals
    elsewhere for the non-relativistic [9] and relativistic [10] problems, ensures that the Fock operator ... Relativistic binding energy and bond length ...