Fact-checked by Grok 2 weeks ago

Oxidation state

The oxidation state of an atom is a measure of its degree of oxidation within a or , defined as the hypothetical charge it would carry if all bonds were fully ionic, with electrons assigned to the more electronegative atom in heteronuclear bonds and divided equally in homonuclear bonds. This integer value, which can be positive, negative, or zero, provides a standardized way to track distribution and changes during chemical reactions. Oxidation states are fundamental in chemistry, where an increase in an atom's oxidation state signifies oxidation (electron loss), and a decrease signifies (electron gain). They enable the balancing of equations by identifying without explicitly counting s, and they inform predictions of reactivity, such as in and . In coordination chemistry, oxidation states of central metal atoms are essential for naming complexes and understanding their magnetic, spectroscopic, and biological properties. Assignment of oxidation states follows IUPAC rules, including that the sum of oxidation states in a is zero and equals the ion's charge in charged ; common conventions assign +1 to (except in metal hydrides), -2 to oxygen (except in peroxides and superoxides), and 0 to elements in their form. These states also correlate with physical trends, such as acidity in oxoacids and stability in high-oxidation-state s, aiding in the design of materials and pharmaceuticals.

Introduction and Definition

Overview

The oxidation state of an atom in a is defined as the hypothetical charge it would possess if all bonds were completely ionic, with electrons assigned to the more electronegative atom in heteronuclear bonds and shared equally in homonuclear bonds. This concept serves as a tool to track the degree of electron loss or gain, facilitating the analysis of in chemical reactions. The idea of oxidation states originated in the late with Lavoisier's work on and reactions involving oxygen, initially framing oxidation strictly as combination with oxygen, but it has since evolved into a broader framework applicable to all processes and beyond. By the early , chemists like Wendell Latimer refined it into a systematic method for assigning numerical values to atoms in compounds, extending its utility far past oxygen-centric reactions. Oxidation states are essential in for predicting an element's reactivity, compound stability, and systematic , as they reveal how electrons are distributed in bonding. For instance, in (NaCl), sodium has an oxidation state of +1 and -1, reflecting complete in this ionic compound, while in potassium permanganate (KMnO₄), reaches +7, indicating its strong oxidizing power due to high . In practical applications, oxidation states underpin processes in , such as in lithium-ion batteries where transition metals like change oxidation states (e.g., from +4 to +3) during charge-discharge cycles, driving flow for power generation. They also explain , like the rusting of iron where the metal oxidizes from 0 to +2 or +3 states in the presence of oxygen and , leading to material degradation. Additionally, in , tracking oxidation states helps assess pollutant behavior, such as the mobility of species where the +6 state is more toxic and soluble than +3.

Formal Definition

The oxidation state, also known as oxidation number, of an atom is defined by the International Union of Pure and Applied Chemistry (IUPAC) as the charge it would bear if all heteronuclear bonds were converted to ionic bonds, with electrons assigned entirely to the more electronegative atom, while homonuclear bonds are divided equally between the atoms. This ionic approximation provides a formal electron-counting to describe the degree of oxidation of an atom within a . For a central atom in a coordination , the oxidation state is specifically the charge it would bear if all ligands were removed along with the electron pairs shared with it, in the ionic model where the ligands are assigned the shared electron pairs (equivalent to heterolytic of the bonds). In practice, this is calculated using the OS = z − ∑q_i, where OS is the oxidation state of the central atom, z is the overall charge of the coordination , and ∑q_i is the sum of the charges assigned to the ligands (with ligands having q_i = 0 and charged ligands contributing their ionic charge). Key assumptions underlying this definition include treating all bonds as fully ionic regardless of their actual and ignoring the covalent sharing of electrons, which distinguishes oxidation state from the actual partial atomic charges derived from quantum chemical calculations. Despite its utility, the oxidation state has inherent limitations as it serves primarily as a bookkeeping tool for tracking distribution in processes rather than reflecting the true electronic structure or oxidation level of an atom. It does not account for delocalized electrons or effects accurately and can yield fractional or ambiguous values in certain cases, emphasizing its role as a conceptual rather than a physical . The oxidation state differs from , another electron-assignment concept, in that formal charge assumes equal sharing of bonding electrons (dividing each pair equally) and assigns all lone-pair electrons to the atom, whereas oxidation state prioritizes by awarding both electrons of a heteronuclear to the more electronegative partner. This distinction makes oxidation state particularly suited for analyzing reactions, while formal charge is more useful for assessing stability.

Determination Methods

Basic Rules for Simple Compounds

The basic rules for assigning oxidation states in simple compounds assume an ionic model, where electrons are fully transferred from less electronegative to more electronegative atoms, providing a straightforward way to determine the hypothetical charge on each atom. These rules are particularly applicable to ionic compounds and simple molecular species without delocalized bonding or complex electron sharing. A fundamental principle is that the sum of the oxidation states of all atoms in a neutral compound equals zero, whereas in a , the sum equals the ion's charge. This conservation rule ensures consistency across the and serves as the basis for solving for unknown states. For monatomic ions, the oxidation state simply matches the ion's charge, such as +1 for Na⁺ or -2 for O²⁻. Certain elements have fixed oxidation states in simple compounds due to their and common bonding behaviors. , the most electronegative element, always exhibits an oxidation state of -1 in its compounds. Oxygen typically has an oxidation state of -2, though exceptions occur in peroxides (where it is -1) and compounds like OF₂ (where it is +2). is usually +1 when bonded to nonmetals but -1 in metal hydrides, such as NaH. Alkali metals () are always +1, and alkaline earth metals (Group 2) are always +2, reflecting their tendency to lose one or two electrons, respectively. To assign oxidation states in binary compounds, follow a step-by-step process without considering partial differences: first, identify elements with known fixed states (e.g., metals or ); second, apply the sum rule to solve for the remaining element's state; third, verify the total equals zero (or the charge). For example, in , is +1 (known rule), so two H atoms contribute +2; thus, oxygen must be -2 to sum to zero. In , each is +1 (total +2), each O is -2 (total -8 for four O), so S must be +6 to balance to zero. For the SO₄²⁻, four O contribute -8, so S is +6, and the total is -2, matching the 's charge. These rules facilitate quick assignment in straightforward cases, such as binary ionic compounds like NaCl (Na +1, Cl -1) or CaO (Ca +2, O -2), emphasizing the ionic approximation for educational and purposes.

Advanced Algorithms for Bonding

In compounds featuring covalent or delocalized bonds, where simple ionic approximations fail, advanced s rely on differences and structural representations to assign oxidation states systematically. The primary , as recommended by IUPAC, treats each bond individually: for heteronuclear bonds, both electrons are assigned to the more electronegative atom; for homonuclear bonds or bonds between atoms of equal electronegativity, the electrons are split equally between the atoms. Lone pair electrons are fully assigned to their respective atoms. The oxidation state (OS) of an atom is then calculated as OS = V - S, where V is the number of electrons in the neutral atom, and S is the total number of electrons assigned to that atom from lone pairs and bonds. This electronegativity-based assignment provides a consistent framework for molecules with mixed bonding. For instance, in (H₂O), oxygen (V = 6) receives both electrons from each O-H bond (adding 4 electrons) plus its two lone pairs (4 electrons), yielding S = 8 and OS = 6 - 8 = -2, while each (V = 1) receives no bonding electrons and has no lone pairs, giving OS = +1. In cases of equal electronegativity, such as the O-O bonds in (H₂O₂), the bond electrons are shared equally, resulting in OS = -1 for each oxygen./Electrochemistry/Redox_Chemistry/Oxidation_State/Oxidation_States_II) A complementary algorithm applies to Lewis structures by summing the contributions from bond orders. In a Lewis representation, each bond order contributes proportionally to the electron assignment: full bonds (order 1) follow the electronegativity rule, while fractional orders (from resonance) are handled by averaging across structures. For ozone (O₃), one resonance form depicts a central oxygen double-bonded to one terminal oxygen and single-bonded to the other, with formal charges indicating OS values of 0 for the double-bonded terminal oxygen, +1 for the central oxygen, and -1 for the single-bonded terminal oxygen. Averaging over the two equivalent resonance structures yields an effective OS of +1 for the central oxygen and -0.5 for each terminal oxygen, maintaining a molecular total of zero./Electrochemistry/Redox_Chemistry/Oxidation_State/Oxidation_States_II) The bond graph method extends this approach by modeling the molecule as a graph where vertices represent atoms and edges represent bonds with assigned orders, facilitating computation for complex or delocalized systems. In benzene (C₆H₆), each carbon atom connects to two carbons (bond order 1.5 each) and one hydrogen (bond order 1). With equal electronegativity in C-C bonds, each carbon receives 1.5 electrons per C-C bond (total 3 from two bonds); for the C-H bond, carbon (more electronegative) receives both electrons (2 total). Thus, V = 4 for carbon, S = 5, yielding OS = -1 per carbon, with the average across the ring reflecting the symmetric delocalization but consistent individual assignments. These algorithms include important caveats. In fully symmetric homonuclear diatomics like Cl₂, each (V = 7) has three lone pairs (6 electrons) and shares one bond equally (1 electron assigned), resulting in S = 7 and OS = 0 for both atoms, rendering the assignment arbitrary yet balanced. Resonance structures necessitate averaging oxidation states to capture delocalization, as in , avoiding overemphasis on any single form./Electrochemistry/Redox_Chemistry/Oxidation_State/Oxidation_States_II) Fundamentally, oxidation states derived from these methods are not quantum mechanical observables, as they rely on classical ionic approximations rather than distributions from wavefunctions; however, they remain valuable tools for predicting reactivity and organizing chemical trends./17:_Electrochemistry/17.03:_Defining_Oxidation_States)

Use in Redox Balancing

In reactions, changes in oxidation states reflect the transfer of electrons between , where an increase in oxidation state corresponds to oxidation (loss of electrons) and a decrease to (gain of electrons). Balancing these reactions requires ensuring that the total number of electrons lost equals the total gained, which is achieved by quantifying oxidation state changes. This principle underpins both the oxidation number method and the integration of oxidation states within the half-reaction method, allowing for systematic equation balancing in various media. The oxidation number method directly uses changes in oxidation states to balance equations. The process begins by assigning oxidation numbers to all elements in the unbalanced equation to identify the oxidizing and reducing agents based on state increases and decreases, respectively. Next, the equation's atoms (except and oxygen) are balanced using coefficients determined by the of the oxidation number changes per atom; for instance, if one gains 5 s (a decrease of 5 in oxidation state) and another loses 1 per atom, a of 5 is applied to the latter to equalize . and oxygen atoms are then balanced by adding H⁺ (or H₂O in conditions) and H₂O, followed by verifying overall charge balance. This method is particularly straightforward for reactions where oxidation state shifts are clear and simplifies tracking electron flow without initially separating half-reactions. A representative example is the reaction between permanganate ion (MnO₄⁻) and iron(II) ion (Fe²⁺) in acidic medium, yielding Mn²⁺ and Fe³⁺. Manganese changes from +7 in MnO₄⁻ to +2 in Mn²⁺ (5-electron gain), while iron changes from +2 to +3 (1-electron loss per ion). To balance, the coefficient for Fe²⁺ is set to 5, resulting in the equation MnO₄⁻ + 5 Fe²⁺ + 8 H⁺ → Mn²⁺ + 5 Fe³⁺ + 4 H₂O after adjusting for oxygen and hydrogen. This demonstrates how equalizing the total oxidation state change (5 units) ensures conservation of electrons. Oxidation states also integrate seamlessly into the method, where they first identify the oxidation and half-reactions by highlighting state changes in the oxidizing and reducing agents. Each is balanced separately: atoms other than and oxygen are equalized, oxygen by adding H₂O, by adding H⁺ (in acidic media) or ⁻ (in basic), and charge by adding s based on the oxidation state difference. The balanced half-reactions are then combined by multiplying to equalize s and simplifying. In acidic media, H⁺ and H₂O are used; in basic, ⁻ is added to neutralize H⁺, converting it to H₂O. This approach, informed by initial oxidation state assignments, is versatile for complex reactions involving multiple transfers.

Variations and Ambiguities

Nominal and Average States

Nominal oxidation states represent the most stable or common values assigned to elements in typical compounds, serving as simplified approximations for educational purposes and predicting chemical behavior. These states are determined using rules, such as assigning -2 to oxygen and +1 to in binary compounds, to balance the overall charge. For example, carbon exhibits a nominal oxidation state of +4 in (CO₂), where the two oxygen atoms each have -2, resulting in a molecule./Electrochemistry/Redox_Chemistry/Oxidation_States_(Oxidation_Numbers)) Similarly, nitrogen has a nominal oxidation state of -3 in (NH₃), as the three hydrogens contribute +1 each./Electrochemistry/Redox_Chemistry/Oxidation_States_(Oxidation_Numbers)) Average oxidation states, on the other hand, are computed for equivalent atoms within polyatomic species by distributing the total charge equally among them, which is particularly relevant in symmetric clusters or delocalized systems. In the neutral cluster S₈, the overall charge of zero divided by eight sulfur atoms yields an average oxidation state of 0 for each sulfur. Likewise, in (P₄O₁₀), the ten oxygen atoms contribute a total of -20 charge, so the four atoms share +20, giving an average oxidation state of +5 per phosphorus atom. This method of averaging applies to metal clusters as well, where delocalized electrons prevent distinct assignments to individual atoms; for instance, in certain clusters like Au₂₀ supported by ligands, the average oxidation state of atoms is 0, reflecting the neutral overall structure. Nominal states thus provide simplifications for straightforward compounds in teaching and , while average states offer a precise charge for complex, symmetric assemblies without implying fractional values for individual atoms.

Ambiguous and Fractional Cases

In compounds exhibiting , such as the nitrate ion (NO₃⁻), the oxidation state of is formally assigned as +5, based on the standard rules where oxygen is -2 and the ion's overall charge is -1. However, resonance delocalizes the π electrons across the three N-O bonds, resulting in equivalent bond lengths and partial charges on the oxygen atoms rather than distinct single and double bonds, which challenges the strict ionic approximation underlying formal oxidation states. Computational methods, such as , reveal partial charges on nitrogen around +0.7 to +1.0 and on each oxygen approximately -0.9 to -1.0, highlighting the ambiguity between formal assignments and actual electron distribution. Fractional oxidation states arise in systems like metal clusters or mixed-valence compounds where electrons are delocalized, preventing assignments to individual s. For instance, in the mercurous (Hg₂²⁺), each mercury is formally +1 due to the symmetric Hg-Hg bond, but in larger mercury clusters such as Hg₃²⁺ or Hg₄²⁺, the average oxidation state per mercury becomes +⅔ or +½, respectively, reflecting shared electrons across the cluster. Similarly, in (FeS₂), iron is +2 and the (S₂) unit is -2, assigning -1 to each sulfur . These cases underscore that fractional states are averages over symmetric units rather than true non-integer charges on single atoms. Physical techniques like X-ray photoelectron spectroscopy (XPS) and Mössbauer spectroscopy provide insights into effective oxidation states by measuring binding energies or hyperfine interactions, which reflect the local electron density rather than formal assignments. XPS detects shifts in core-level binding energies; for example, in iron compounds, Fe²⁺ shows a 2p₃/₂ peak around 709-710 eV, while Fe³⁺ is at 711 eV, allowing quantification of mixed states in ambiguous systems. Mössbauer spectroscopy, particularly for ⁵⁷Fe, uses isomer shifts to distinguish oxidation states based on s-electron density, as seen in delocalized iron clusters where effective states deviate from formal ones. Importantly, these methods probe physical electron distributions, whereas oxidation states remain a formal construct for bookkeeping electrons in bonding, emphasizing the distinction between conceptual and experimental realities. Arbitrary assignments occur in allotropes or species with variable bonding, such as dioxygen (O₂), where oxygen is formally 0 as an elemental form with a , but in peroxides like H₂O₂, the O-O leads to an oxidation state of -1 per oxygen atom to balance the structure under IUPAC rules. This convention treats the linkage as a reduced form compared to oxides (-2), aiding and tracking, though actual bond orders and charges may vary slightly due to .

Multiple States in Elements and Metals

Many elements, particularly those in the d-block, exhibit a wide range of oxidation states due to the availability of d-electrons for bonding, allowing states from +2 up to +7 or higher in some cases. This variability arises because transition metals can lose both ns and (n-1)d electrons, leading to multiple stable configurations in different compounds. For instance, manganese (Mn) commonly displays oxidation states of +2, +3, +4, +6, and +7, as seen in compounds like Mn²⁺ ions in aqueous solutions, MnO₂ (+4), and the permanganate ion MnO₄⁻ (+7). These states enable manganese's role in diverse chemical behaviors, from reducing agents in lower states to strong oxidants in higher ones. The stability of these oxidation states in transition metals is influenced by factors such as successive ionization energies, which increase for higher states, making them less favorable unless stabilized by specific ligands, and ligand field stabilization energy (LFSE) from crystal field theory, which provides extra stability to certain d-electron configurations. Higher oxidation states tend to be more oxidizing due to the higher charge density on the metal ion, which polarizes surrounding ligands and facilitates electron acceptance. In bulk metals, the oxidation state is typically 0, reflecting metallic bonding where electrons are delocalized, but in alloys, surfaces, or catalytic environments, variable states emerge; for example, platinum (Pt) cycles between Pt(0) in zero-valent nanoparticles and Pt(IV) in oxidized forms during catalysis, enhancing reactivity in processes like hydrogenation. Elements in the p-block also show multiple oxidation states, though generally fewer than d-block metals, with nitrogen (N) exemplifying a range from -3 to +5 due to its high electronegativity, which favors both electron gain and sharing. In ammonia (NH₃), nitrogen is in the -3 state, while in nitric acid (HNO₃), it reaches +5; stability trends indicate that higher oxidation states are more stable for elements with greater electronegativity, like nitrogen at the top of group 15, whereas lower states become prominent down the group due to the inert pair effect. The oxidation states of elements exhibit systematic variations across the periodic table, closely tied to their electronic configurations and valence electron availability. In the s-block (Groups 1 and 2), elements display fixed low positive oxidation states corresponding to their group number: +1 for alkali metals and +2 for alkaline earth metals, as these atoms readily lose their single or two s-electrons to achieve noble gas configurations./01%3A_General_Concepts_and_Trends/1.01%3A_Fundamental_Properties_-_Oxidation_State) This stability arises from the low ionization energies and high reactivity of these electropositive metals, limiting higher states due to the absence of accessible d-orbitals. In the p-block (Groups 13–18), oxidation states generally increase with group number, reflecting the progressive filling of p-orbitals. For instance, Group 15 elements can exhibit states from -3 (gaining three electrons to form anions like ) to +5 (losing five electrons in compounds like phosphates), with intermediate states such as +3 common in oxoanions./Descriptive_Chemistry/Main_Group_Reactions/Reactions_of_Main_Group_Elements_with_Oxygen) This trend stems from the ability of p-block elements to either share or transfer electrons based on bonding partners, with maximum positive states equal to the number of electrons (group number − 10) and maximum negative states up to 8 minus the number of electrons (18 − group number) for nonmetals. Transition metals in the d-block (Groups 3–12) show more variable oxidation states, often ranging from +2 to the group number (e.g., up to +7 for in Group 7), due to the involvement of both s and d electrons in bonding, allowing multiple electron loss options./Descriptive_Chemistry/Elements_Organized_by_Block/3d-Block_Elements/1b_Properties_of_Transition_Metals/Oxidation_States_of_Transition_Metals) Across periods, oxidation states tend to reach higher maxima in heavier rows because of relativistic effects and poorer shielding by inner d- or f-electrons, which increase and stabilize higher oxidation states. For example, while 3d transition metals like achieve +6, the 4d and 5d analogs ( and ) more readily form stable +6 states in compounds due to larger atomic sizes accommodating higher charges without excessive repulsion./Descriptive_Chemistry/Elements_Organized_by_Block/3d-Block_Elements/1b_Properties_of_Transition_Metals/General_Trends_among_the_Transition_Metals) Higher oxidation states become less stable across a row from left to right but more accessible down a group. An important exception occurs in heavier p-block elements, where the stabilizes lower oxidation states by two units compared to the group maximum; for in Group 13, the +1 state is preferred over +3 because the 6s electrons are reluctant to participate in bonding due to increased relativistic stabilization and poor overlap with ligands. These trends are further influenced by , where elements with higher electronegativity values favor elevated positive oxidation states in compounds with less electronegative partners, as the shifts toward the more electronegative atom, formalizing higher oxidation numbers for the counterpart./08%3A_Chemistry_of_the_Main_Group_Elements/8.01%3A_General_Trends_in_Main_Group_Chemistry/8.1.02%3A_Electronegativity_increases_and_radius_decreases_towards_the_upper_left_of_the_periodic_table_with_electron_withdrawing_substituents_and_with_oxidation_state) For instance, fluorine's high electronegativity (4.0 on the Pauling scale) consistently results in -1 states, forcing partners like oxygen in OF₂ to +2. This interplay underscores how electronic structure and periodic positioning dictate the feasibility and prevalence of oxidation states in chemical bonding.

Common Oxidation States List

The common oxidation states of elements are influenced by their electronic configuration and position in the periodic table, with s-block elements showing the most predictable values due to their tendency to lose electrons completely. These states are fundamental in predicting compound formation and reactivity./Descriptive_Chemistry/Elements_Organized_by_Block/s-Block_Elements) For the s-block, alkali metals () typically exhibit +1 oxidation state in their compounds, as they lose their single s-electron to form stable ions, while alkaline earth metals (group 2) show +2 by losing both s-electrons. , associated with this block, displays +1 in acids and -1 in hydrides. Rare deviations, such as Be forming covalent bonds, are noted but uncommon./Descriptive_Chemistry/Elements_Organized_by_Block/s-Block_Elements/Group_2:_General_Properties)
ElementCommon Oxidation StatesExamples
Li+1LiCl ()
Na+1NaOH ()
K+1KF ()
Be+2BeO ()
Mg+2MgSO₄ ()
Ca+2CaCO₃ ()
Sr+2SrCl₂ ()
Ba+2Ba(NO₃)₂ ()
H-1, +1NaH (), HCl ()
The p-block elements display greater variability in oxidation states, ranging from negative values for nonmetals forming anions to positive states up to +8 for heavier elements like , reflecting their ability to gain, lose, or share p-electrons in diverse bonding scenarios. Common states often align with group numbers for maximum oxidation, but lower states are frequent due to incomplete octet satisfaction.
ElementCommon Oxidation StatesExamples
B+3BCl₃ (boron trichloride)
C-4, +4CH₄ (methane), CO₂ (carbon dioxide)
N-3, +3, +5NH₃ (ammonia), HNO₂ (nitrous acid), HNO₃ (nitric acid)
O-2H₂O (water)
F-1HF (hydrogen fluoride)
Al+3Al₂O₃ (aluminum oxide)
Si+4, -4SiO₂ (silicon dioxide), SiH₄ (silane)
P-3, +3, +5PH₃ (phosphine), P₄O₆ (phosphorus trioxide), H₃PO₄ (phosphoric acid)
S-2, +4, +6H₂S (hydrogen sulfide), SO₂ (sulfur dioxide), H₂SO₄ (sulfuric acid)
Cl-1, +1, +5, +7NaCl (sodium chloride), HClO (hypochlorous acid), HClO₃ (chloric acid), HClO₄ (perchloric acid)
Xe+2, +4, +6, +8XeF₂ (xenon difluoride), XeF₄ (xenon tetrafluoride), XeO₃ (xenon trioxide), XeO₄ (xenon tetroxide)
d-block transition metals exhibit multiple oxidation states due to the availability of both s and d electrons for bonding, with common values from +2 to +7, often stabilizing through ligand interactions in coordination compounds. The zero state is typical in metallic forms, and higher states are more common in early transition metals./Descriptive_Chemistry/Elements_Organized_by_Block/d-Block_Elements)
ElementCommon Oxidation StatesExamples
Sc+3Sc₂O₃ (scandium oxide)
Ti+3, +4Ti₂O₃ (titanium(III) oxide), TiO₂ (titanium dioxide)
V+2, +3, +4, +5VO (vanadium(II) oxide), V₂O₃ (vanadium(III) oxide), VO₂ (vanadium(IV) dioxide), V₂O₅ (vanadium pentoxide)
Cr+3, +6Cr₂O₃ (chromium(III) oxide), CrO₃ (chromium trioxide)
Mn+2, +4, +7MnO (manganese(II) oxide), MnO₂ (manganese dioxide), KMnO₄ (potassium permanganate)
Fe+2, +3FeO (iron(II) oxide), Fe₂O₃ (iron(III) oxide)
Co+2, +3CoO (cobalt(II) oxide), Co₂O₃ (cobalt(III) oxide)
Ni+2NiO (nickel(II) oxide)
Cu+1, +2Cu₂O (copper(I) oxide), CuO (copper(II) oxide)
Zn+2ZnS (zinc sulfide)
The f-block elements, including lanthanides and actinides, generally favor +3 oxidation states owing to the stability of f⁰ or f¹⁴ configurations after losing the s-electrons, though actinides can reach higher states up to +6 or +7 due to greater involvement of 5f orbitals in bonding. Rare earths show limited variability, while actinides display more diverse chemistry./Descriptive_Chemistry/Elements_Organized_by_Block/f-Block_Elements)
ElementCommon Oxidation StatesExamples
+3La₂O₃ (lanthanum oxide)
+3, +4Ce₂O₃ (cerium(III) oxide), CeO₂ (cerium(IV) dioxide)
+3NdCl₃ (neodymium chloride)
+2, +3EuCl₂ (europium(II) chloride), Eu₂O₃ (europium(III) oxide)
+3Lu₂O₃ (lutetium oxide)
+3AcCl₃ (actinium chloride)
+4ThO₂ (thorium dioxide)
U+3, +4, +5, +6UO₂ (uranium(IV) oxide), UO₂(NO₃)₂ (uranyl nitrate for +6)
+3, +4, +5, +6PuO₂ (plutonium(IV) dioxide), PuO₃ (plutonium(VI) oxide)
Am+3, +4, +5, +6AmF₃ (americium(III) fluoride), AmO₂ (americium(IV) oxide), AmO₂(NO₃) (americium(V) nitrate), AmO₂²⁺ (americium(VI) ion)

Applications and Nomenclature

Role in Naming Compounds

The Stock system, also known as the oxidation number system, employs Roman numerals in parentheses following the name of the element to denote its oxidation state in compounds, particularly for metals exhibiting variable oxidation states. This approach ensures unambiguous identification of the specific compound when multiple possibilities exist, such as distinguishing FeCl₂ as iron(II) chloride from FeCl₃ as iron(III) chloride. In the of coordination compounds, ligands are named first in , followed by the central metal atom with its oxidation state indicated by a Roman numeral in parentheses if the state is not deducible from the overall charge. For anionic es, the ending "-ate" is used; for example, the [Fe(CN)₆]⁴⁻ is named hexacyanidoferrate(II), where "ferrate" indicates the iron-centered anion and (II) specifies the +2 oxidation state of iron. Cationic complexes retain the metal name, as in [Co(NH₃)₆]³⁺ hexaamminecobalt(III). This systematic ordering prioritizes clarity in describing the and metal oxidation state. For oxoanions, IUPAC nomenclature uses suffixes "-ate" for the highest oxidation state and "-ite" for the next lower, with prefixes "hypo-" for even lower states and "per-" for the highest when four oxyanions are possible. Thus, is (chlorine in +1 oxidation state), is (+3), is (+5), and is (+7), reflecting the progressive increase in oxygen content and oxidation state. These conventions derive from the central atom's oxidation state relative to oxygen's -2 assignment./03%3A_Ionic_Compounds/3.03%3A_Names_and_Formulas_of_Ionic_Compounds/3.3.03%3A_Naming_Ionic_Compounds) The integration of oxidation states into modern IUPAC evolved from earlier, less precise systems—such as "ferric" for +3 iron and "" for +2—to the system, adopted internationally to eliminate ambiguity in naming compounds of elements with multiple oxidation states. This shift, formalized in IUPAC recommendations, promotes consistency across inorganic and coordination chemistry.

Utility in

Oxidation states play a crucial role in electrochemistry by enabling the prediction of redox potentials and reaction feasibility, which underpin the design and optimization of electrochemical systems. In these processes, changes in oxidation states determine the direction and driving force of electron transfer, influencing energy storage, catalytic efficiency, and material stability. For instance, the voltage of an electrochemical cell arises from the difference in standard reduction potentials associated with specific oxidation state transitions, allowing engineers to select materials that maximize performance while minimizing side reactions. In battery technologies, particularly lithium-ion batteries, oxidation state variations drive the cell voltage and capacity. During discharge, lithium metal or intercalated lithium undergoes oxidation from the 0 to +1 state at the , as in the reaction Li → Li⁺ + e⁻ within hosts, while at the , s like or undergo from higher to lower oxidation states (e.g., Co⁴⁺ to Co³⁺) to accommodate Li⁺ insertion. These state changes directly correlate with the , typically around 3.7 V for LiCoO₂-based cells, and guide the development of high-energy-density cathodes by predicting stability limits and degradation pathways, such as transition metal dissolution due to over-oxidation. In catalysis, oxidation state cycling facilitates multi-electron transfers essential for reactions like water oxidation and organic transformations. Ruthenium complexes, for example, cycle through states from Ru(II) to higher valences, including Ru(VIII) in ruthenium tetroxide-mediated oxidations of alcohols to aldehydes or ketones, enabling selective and efficient catalysis under mild conditions. This cycling is leveraged in enzymatic mimics and heterogeneous catalysts, where the accessibility of multiple states enhances turnover frequencies, as seen in Ru-based water oxidation catalysts that achieve rates up to 10 s⁻¹ by stabilizing high-valent intermediates. Corrosion processes are predicted using oxidation state stability through Pourbaix diagrams, which map potential-pH regions where specific states predominate. For iron, corrosion proceeds via oxidation from Fe(0) in metallic form to Fe(III) in (Fe₂O₃·nH₂O), predominant in neutral to acidic aqueous environments above approximately -0.4 V vs. SHE, leading to rapid material loss at rates exceeding 1 mm/year in unprotected . These diagrams, derived from Nernst equations and solubility data, inform corrosion-resistant alloy design by identifying passivation zones, such as Fe(II)/Fe(III) oxide layers that stabilize at > 9./04:_Redox_Stability_and_Redox_Reactions/4.06:_Pourbaix_Diagrams) In contemporary applications as of 2025, oxidation states derived from (DFT) computations accelerate the screening of novel materials for batteries and electrocatalysts. DFT models predict state-dependent potentials with errors below 0.2 V for and metal-complex electrolytes, enabling high-throughput evaluation of thousands of candidates to identify , high-voltage systems like quinone-based mediators. integrations with DFT further refine these predictions, incorporating oxidation state evolution to forecast cycling stability in next-generation solid-state batteries.

Historical Evolution

Early Conceptual Foundations

The concept of oxidation state traces its roots to the late , when proposed his oxygen theory of acids and bases. Lavoisier posited that oxygen was the essential component of all acids, serving as the "acidifying principle" that imparted acidic properties to compounds, while bases were substances capable of neutralizing acids by combining with oxygen. This theory implied a rudimentary notion of positive and negative equivalents in chemical combinations, where oxygen acted as a negative entity combining with positive metals or radicals to form compounds, laying early groundwork for understanding valence-like behaviors in processes. In the , advanced these ideas through his theory of electrochemical , which viewed chemical compounds as aggregates of electropositive and electronegative atoms held together by electrostatic attraction. Berzelius classified elements based on their electrical affinities, with metals as electropositive and nonmetals as electronegative, providing a framework for assigning relative charges in compounds that prefigured modern oxidation numbers. Concurrently, the development of s—introduced by Jeremias Benjamin Richter in the 1790s and refined by Berzelius—quantified the combining capacities of elements, where the equivalent weight of an element was its atomic weight divided by its , directly linking to variable oxidation behaviors in reactions. This system enabled chemists to balance equations and predict compound formation based on stoichiometric equivalents, emphasizing the role of oxygen or other electronegative elements in altering elemental "states." Early reflected these evolving concepts, particularly in naming oxides and acids where oxidation levels varied. In the reform led by Lavoisier and collaborators like Guyton de Morveau, systematic names for oxygen-containing compounds were established, but for transition metals exhibiting multiple oxidation states, the suffixes "-ic" and "-ous" were adopted to distinguish higher and lower states, respectively—such as "ferric" for (III) and "ferrous" for (II) in oxides and salts. This convention, rooted in Latin roots and extended by French chemists in the early 1800s, facilitated communication about variations without explicit numerical assignment, influencing the naming of acids like sulfuric (higher oxygen) versus sulfurous (lower oxygen). A significant precursor to formal oxidation state assignments emerged around with N. Lewis's introduction of shared electron pairs and the in his seminal paper "The Atom and the Molecule." Lewis depicted atoms as achieving stability by completing an octet of electrons through bonding, implying ionic or covalent models where electrons were formally assigned to atoms, enabling the calculation of formal charges that paralleled oxidation numbers in simple cases. This electron-based perspective bridged 19th-century dualistic ideas with quantum understandings, setting the stage for precise tracking in compounds.

Development of Modern Framework

In the early , the concept of oxidation states evolved from qualitative descriptions toward a more formalized system, particularly during the and 1940s. This period saw the integration of oxidation states with emerging theories of chemical bonding, such as developed by and others, which provided a framework for understanding how atoms achieve different oxidation states through shared electron pairs and hybridization. The term "oxidation number" was officially introduced by Wendell M. Latimer in 1938 in his seminal book The Oxidation States of the Elements and Their Potentials in Aqueous Solutions, where it was defined as a numerical value representing the degree of oxidation of an atom based on differences in bonds. This formalization, building on earlier work by chemists like G.N. Lewis, allowed for systematic assignment in complex compounds and aligned with the growing use of quantum mechanical models to explain multiple oxidation states in transition metals. By the 1970s, the International Union of Pure and Applied Chemistry (IUPAC) began standardizing to clarify distinctions between related concepts. The 1971 IUPAC recommendations for explicitly defined oxidation numbers as formal charges assuming complete in bonds, while distinguishing them from coordination numbers, which denote the number of donor atoms attached to a central atom. This separation was crucial for coordination compounds, where confusion had previously arisen, and it emphasized oxidation numbers' role in balancing rather than geometric arrangement. From the to the , IUPAC refined oxidation state assignments to address challenges in organometallic compounds and metal s, where traditional rules often led to ambiguities due to delocalized electrons and covalent character. The IUPAC recommendations for nomenclature of organometallic compounds of transition elements noted the ambiguities in assigning oxidation states, particularly in species with metal-carbon bonds, and advised against using formal oxidation numbers in their nomenclature. These refinements recognized the utility of average (fractional) oxidation states in symmetric cluster compounds, where delocalized electrons lead to non-integer values. Post-2013 developments have increasingly incorporated to probe the limitations of formal oxidation states, highlighting their lack of direct physical correspondence to distribution. The 2016 IUPAC comprehensive definition reinforced the ionic approximation model but acknowledged critiques regarding its applicability to delocalized systems. Recent has proposed "effective oxidation states" derived from wavefunction or analyses, such as using Bader's atoms-in-molecules partitioning to assign states based on actual charge accumulation rather than formal rules; for instance, a 2020 study on coordination compounds, including complexes, demonstrated how effective states resolve borderline cases in organometallics where formal assignments fail. Similarly, topological approaches in ionic conductors have critiqued traditional states by linking them to charge pathways, revealing dynamic variations not captured by static formalisms. Building on this, a 2022 study integrated topological theory with oxidation states to analyze charge in ionic conductors, demonstrating how nontrivial in non-stoichiometric systems reveals variations not captured by traditional static assignments.