Fact-checked by Grok 2 weeks ago

Polyatomic ion

A polyatomic ion is a covalently bonded group of two or more atoms that carries a net , functioning as a single charged unit in chemical reactions. These ions form when a neutral gains or loses one or more electrons, resulting in a stable charged species, such as by deprotonating an oxoacid (e.g., removing H⁺ from , HNO₃, to yield the nitrate ion, NO₃⁻). Common examples include the cation (NH₄⁺), anion (OH⁻), nitrate anion (NO₃⁻), anion (SO₄²⁻), and anion (PO₄³⁻), each with distinct names, formulas, and charges that must be memorized for accurate . Polyatomic ions are essential building blocks in , frequently appearing in ionic compounds, salts, acids, and bases where they pair with oppositely charged ions to achieve electrical neutrality. For instance, in compounds like (Ca(NO₃)₂), the formula uses parentheses around the polyatomic ion to indicate multiples, and naming follows the convention of stating the cation first followed by the anion name without alteration. Their prevalence in everyday substances—such as fertilizers (), detergents (), and biological systems ( in DNA)—underscores their practical significance, making recognition of their structures and behaviors crucial for understanding chemical properties and reactions.

Fundamentals

Definition and Characteristics

A is a charged composed of two or more atoms that are covalently bonded together, with the overall charge arising from the loss or gain of one or more electrons by the group. These ions exhibit covalent bonding among their internal atoms while interacting through ionic bonds with oppositely charged species in compounds, behaving as a single charged unit during chemical reactions and in ionic lattices. In aqueous solutions, polyatomic ions typically dissociate from their parent salts, maintaining their structural integrity and charge while participating in electrolytic processes. A common representation for oxyanions, a prevalent class of polyatomic ions, is the general formula \ce{XO_n^{m-}}, where X is a central atom, O represents oxygen atoms, n indicates the number of oxygen atoms, and m- denotes the overall negative charge; the charge is distributed across the structure due to the covalent bonds and delocalization, contributing to the ion's stability. The term "ion" for charged particles in was introduced by in 1834, based on his studies demonstrating the migration of charged particles in solutions. It was formalized within Svante Arrhenius's electrolytic dissociation theory in 1884, which explained how electrolytes in solution separate into ions, including polyatomic species, to conduct .

Distinction from Monatomic Ions

Polyatomic ions differ fundamentally from monatomic ions in their structure. Monatomic ions consist of a single atom that has gained or lost one or more s to achieve a net charge, such as the sodium (Na⁺), which is simply a sodium atom with one electron removed. In contrast, polyatomic ions are composed of two or more atoms covalently bonded together, forming a stable group that carries an overall charge, as seen in the sulfate (SO₄²⁻), where four oxygen atoms are bonded to a central atom. This multi-atom framework allows polyatomic ions to exhibit molecular properties absent in their monatomic counterparts. The formation processes of these ions also highlight their distinctions. Monatomic ions typically arise from straightforward between , where a atom loses electrons to form a cation or a gains them to form an anion, resulting in isolated charged . Polyatomic ions, however, often form through more complex mechanisms involving molecules, such as or . For instance, the ammonium ion (NH₄⁺) is generated by the of (NH₃) with a (H⁺), creating a charged molecular rather than altering a single . This process preserves the internal covalent bonds within the ion, distinguishing it from the atomic simplicity of formation. Behaviorally, polyatomic ions exhibit unique reactivity due to their composite nature. Unlike monatomic ions, which interact solely through their atomic charge in ionic reactions, polyatomic ions maintain their internal covalent bonds and act as intact units during chemical processes, such as in the formation of compounds like (NH₄Cl). Additionally, many polyatomic ions, particularly anions like (NO₃⁻), achieve enhanced stability through , where the charge is delocalized across multiple equivalent structures, lowering the overall energy compared to localized bonding in monatomic ions. This resonance stabilization contributes to the persistence of polyatomic ions in solutions and solids, contrasting with the purely electrostatic interactions of monatomic ions. Spectroscopic techniques further underscore these differences, particularly in identification. Monatomic ions lack internal bonds and thus display no vibrational modes in infrared (IR) spectra, appearing as featureless in such analyses. Polyatomic ions, however, produce distinct IR absorption bands corresponding to their vibrational modes, such as the stretching and bending of bonds within the ion—for example, the sulfate ion (SO₄²⁻) shows characteristic peaks around 1100 cm⁻¹ due to S–O asymmetric stretching. These vibrational signatures enable precise identification of polyatomic ions in mixtures, a capability not applicable to monatomic species.

Classification

Polyatomic Anions

Polyatomic anions are negatively charged ions consisting of two or more atoms bound together by covalent bonds, functioning as a single unit in chemical reactions. These ions typically feature a central atom surrounded by electronegative atoms such as oxygen, resulting in an overall negative charge due to an excess of electrons. They are prevalent in ionic compounds and play key roles in acid-base chemistry and aqueous solutions. Polyatomic anions are primarily classified by their atomic composition and structure. The most common category is oxyanions, which incorporate oxygen atoms bonded to a central atom, often from groups 15, 16, or 17 of the periodic table. Oxyanions are further grouped by the central atom; for example, sulfur-based oxyanions include (SO_4^{2-}) and (SO_3^{2-}), while phosphorus-based ones include (PO_4^{3-}) and phosphite (PO_3^{3-}). Halide-containing polyatomic anions, such as those with , feature a central atom bonded to oxygen, exemplified by (ClO^-) and (ClO_4^-). Carbon-based polyatomic anions, which integrate carbon into the framework, include (CN^-) and (CH_3COO^-). The negative charge of polyatomic anions originates from the addition of electrons to the molecular group, frequently through the of polyprotic acids. For instance, (H_2SO_4), a diprotic acid, undergoes stepwise to form the bisulfate ion (HSO_4^-) and then the sulfate ion (SO_4^{2-}). Similarly, (H_3PO_4), a triprotic acid, can lose up to three protons to yield the ion (PO_4^{3-}). This process reflects the acidic nature of the precursor, where protons are removed, leaving the anionic species. Stability in polyatomic anions, especially oxyanions, is influenced by the of the central atom, which determines the number of oxygen atoms and the charge distribution. Higher oxidation states of the central atom typically allow for more oxygen atoms, leading to multiple anionic forms for the same element, such as (ClO_4^-, chlorine +7) and (ClO_2^-, chlorine +3) in the chlorine series. This variation arises because oxygen's high stabilizes higher positive charges on the central atom through delocalized electron density.

Polyatomic Cations

Polyatomic cations are positively charged ions consisting of two or more atoms covalently bonded together, typically involving central atoms from group 15 or 16 elements of the periodic table, such as nitrogen or oxygen. These ions carry a net positive charge due to the loss or gain of electrons in a way that results in an overall cationic species, and they behave as single units in chemical reactions. Unlike the more abundant polyatomic anions, polyatomic cations are relatively rare in common chemical contexts. Prominent examples include the ammonium ion (\mathrm{NH_4^+}), formed from and ; the hydronium ion (\mathrm{H_3O^+}), involving oxygen and ; and organic variants like alkylammonium ions (\mathrm{RNH_3^+}, where R represents an ). These types highlight the prevalence of - and oxygen-based structures among known polyatomic cations. Polyatomic cations primarily form through protonation of neutral molecules bearing lone pairs of electrons. For instance, (\mathrm{NH_3}) accepts a proton to yield the ion: \mathrm{NH_3 + H^+ \rightarrow NH_4^+}. Likewise, (\mathrm{H_2O}) is protonated to form the ion: \mathrm{H_2O + H^+ \rightarrow H_3O^+}. This process is facilitated in aqueous environments but encounters stability limitations in non-aqueous media, where these cations may decompose or fail to persist due to reduced effects. In ionic compounds, polyatomic cations serve as counterions to anions, as exemplified by (\mathrm{NH_4Cl}), a where the ammonium cation balances the anion. Such compounds demonstrate the utility of polyatomic cations in forming stable crystalline structures despite their relative scarcity.

Nomenclature

Naming Conventions for Anions

The International Union of Pure and Applied Chemistry (IUPAC) provides systematic rules for naming polyatomic anions in inorganic , primarily through compositional, substitutive, and additive approaches that reflect the anion's composition, central , and . For oxyanions—those containing oxygen bonded to a central —the name is derived from the root of the central , with suffixes indicating the : the "-ate" suffix denotes the highest common or maximum oxygen content, while "-ite" is used for lower s or reduced oxygen. Prefixes modify these to denote extremes in oxygen content: "hypo-" for the lowest and "per-" for the highest. This convention is particularly evident in oxyanion series derived from the same central element, where sequential addition of oxygen atoms progresses the name accordingly. For the chlorine oxyanions, the series illustrates this progression based on increasing oxidation states from +1 to +7:
FormulaNameOxidation State of Cl
ClO^-Hypochlorite+1
ClO_2^-Chlorite+3
ClO_3^-Chlorate+5
ClO_4^-Perchlorate+7
These names correspond to the conjugate bases of the respective oxyacids (hypochlorous, chlorous, chloric, and perchloric acids) and are retained in IUPAC recommendations for common usage. For non-oxy polyatomic anions, naming follows similar substitutive or additive principles without oxygen-based suffixes. Simple examples include the hydroxide ion (OH^-), named directly from its parent hydride or functional group, and more complex heteropolyatomic ions like (SCN^-), which uses a substitutive name indicating the central atom and ligands. The 2005 IUPAC permits retained traditional names for well-established polyatomic anions (e.g., SO_4^{2-} instead of tetraoxidosulfate(2−)), balancing systematic rigor with practical continuity from historical nomenclature. These rules ensure names reflect the anion's structural basis, such as the number of oxo groups in oxyanions.

Naming Conventions for Cations

The naming of polyatomic cations generally follows IUPAC recommendations, which emphasize systematic or retained names based on the central atom or molecular structure, often appending indicators for the positive charge. For simple inorganic polyatomic cations, traditional names are commonly retained, such as for \mathrm{NH_4^+} and for \mathrm{H_3O^+}, where the name reflects the protonated form of the parent neutral species and , respectively. These names do not require additional charge indicators in isolation, as the cationic nature is implied, but in compounds, the cation name precedes the anion without modification. Organic polyatomic cations, particularly those derived from amines, are named substitutively by adding the suffix "-ium" to the parent hydride name, indicating or at . For example, \mathrm{CH_3NH_3^+} is named methanaminium, while more complex variants like \mathrm{(CH_3)_4N^+} become N,N,N-trimethylmethanaminium; retained names such as alkylammonium (e.g., methylammonium for \mathrm{CH_3NH_3^+}) are permitted for general use but not preferred in strict IUPAC contexts. This approach prioritizes the longest chain or principal function, with locants assigned to ensure the lowest numbers for the cationic center. Coordination polyatomic cations, common in complexes, are named using additive , where ligands are listed alphabetically before the central metal atom, which is followed by its in . Ligand names are modified (e.g., becomes ammine, becomes aqua), with multiplicative prefixes like di-, tri-, or bis- (for complex ligands) indicating ; for instance, [\mathrm{Co(NH_3)_6}]^{3+} is hexaamminecobalt(III). The overall complex is treated as the cation when positively charged, named before any counter anion in the full compound formula. Certain polyatomic cations retain historical common names as exceptions to systematic rules, particularly in . The diazonium group -\mathrm{N_2^+}, as in \mathrm{C_6H_5N_2^+}, is named benzenediazonium, where the suffix "-diazonium" is added directly to the parent name, reflecting its widespread use despite alternatives like diazenylium. Such retained names persist due to their established role in synthetic applications, but IUPAC encourages systematic for novel structures to ensure clarity and consistency.

Structure and Bonding

Covalent Bonding Mechanisms

In polyatomic ions, the atoms are primarily held together by covalent bonds, which form through the sharing of valence electrons between atoms to achieve stable electron configurations. These bonds are often polar covalent due to differences in electronegativity among the atoms involved, resulting in partial charges that contribute to the overall ionic character of the polyatomic unit when it interacts with other ions in compounds. For instance, in the hydroxide ion (\ce{OH^-}), the oxygen-hydrogen bond is polar, with oxygen pulling electrons more strongly, creating a partial negative charge on oxygen and a partial positive on hydrogen. Similarly, in the nitrate ion (\ce{NO3^-}) and ammonium ion (\ce{NH4^+}), the N-O and N-H bonds exhibit polarity from electronegativity differences of approximately 0.5–1.0 and 0.9, respectively. Lewis structures provide a key method for visualizing the arrangement of valence electrons and covalent bonds in polyatomic ions, helping to identify formal charges and bonding patterns. To construct a Lewis structure, one first calculates the total valence electrons—summing those from all atoms, adding electrons for negative charges, and subtracting for positive charges—then arranges atoms with the least electronegative (often central) atom bonded to surrounding atoms using single bonds, followed by distributing lone pairs to satisfy the octet rule where possible. Double or triple bonds are introduced if needed to complete octets, and the ion is enclosed in brackets with its charge. For the sulfate ion (\ce{SO4^2-}), with 32 valence electrons (6 from S, 24 from four O, plus 2 for the charge), the structure features sulfur as the central atom bonded to four oxygens, typically with two double bonds and two single bonds, resulting in formal charges of +2 on S, -1 on two O atoms, and 0 on the others to minimize overall charge separation. This representation highlights how covalent sharing accommodates the ion's charge while adhering to the octet rule for second-period elements. The Valence Shell Electron Pair Repulsion (VSEPR) theory extends this understanding by predicting the three-dimensional geometry of polyatomic ions based on the repulsion between electron pairs around the central atom, minimizing energy through optimal spacing. In VSEPR notation, geometries are classified as AX_m E_n, where A is the central atom, X denotes bonding pairs, and E lone pairs; for many common polyatomic ions without lone pairs on the central atom, shapes are regular polyhedra. The sulfate ion (\ce{SO4^2-}) adopts a tetrahedral geometry (AX_4), with bond angles of approximately 109.5°, as the four bonding pairs repel equally around sulfur. Likewise, the nitrate ion (\ce{NO3^-}) exhibits trigonal planar geometry (AX_3), with 120° bond angles, due to three equivalent bonding pairs around nitrogen. These geometries influence the ion's reactivity and symmetry. Bond order, defined as the average number of pairs shared between atoms, directly impacts bond lengths in polyatomic ions, with higher orders corresponding to shorter, stronger bonds due to greater overlap. In cases involving multiple bonding representations, such as the carbonate ion (\ce{CO3^2-}), the C-O bonds have a of 1.33, calculated as the average across three forms (one and two single bonds per structure), resulting in lengths intermediate between single (~143 pm) and double (~120 pm) C-O bonds. This partial multiple bonding enhances stability and shortens distances compared to purely single bonds, reflecting the shared .

Resonance and Stability

Resonance in polyatomic ions involves the delocalization of π electrons across multiple structures, known as resonance structures, which collectively describe the actual bonding more accurately than any form. This delocalization occurs when a single diagram cannot fully represent the electron , leading to equivalent bonds in ions such as the nitrate ion (\ce{NO3^-}), where three resonance structures depict the nitrogen-oxygen bonds as identical, with each N-O exhibiting a of \frac{4}{3}. The true structure of the ion is a , an average of these contributing forms that minimizes the overall energy of the system. To construct resonance structures, rules are applied: formal charge is calculated as valence electrons minus nonbonding electrons minus half of bonding electrons, prioritizing structures with minimal formal charges and negative charges on more electronegative atoms like oxygen. This approach ensures the hybrid reflects the observed and bond lengths, as seen in experimental data where all bonds in the hybrid are intermediate between single and double bonds. Resonance enhances the stability of polyatomic ions by distributing , which strengthens bonds through partial double-bond character and lowers the ion's reactivity compared to non-resonant analogs. The stabilization is quantified by , the difference between the hybrid's energy and that of the most , often resulting in shorter, stronger bonds; for example, in the (\ce{CO3^2-}), three equivalent lead to uniform C-O bond lengths of approximately 1.29 Å and reduced tendency for nucleophilic attack. Ions resembling aromatic systems, such as certain cyclic polyatomic ions, exhibit even greater due to cyclic delocalization akin to . However, resonance delocalization is limited in non-planar polyatomic ions, where poor overlap of p-orbitals hinders effective π sharing, resulting in less stabilization and bond equivalence. In such cases, the ion's structure more closely resembles a single dominant form rather than a .

Special Types

Zwitterions

Zwitterions are neutral molecules containing both positively and negatively charged functional groups, resulting in an overall electrical neutrality despite the internal charge separation. This dipolar structure arises primarily in amphoteric compounds, where the charges are separated within the same molecule, distinguishing zwitterions from polyatomic ions that carry a net charge. Zwitterions exhibit ionic behavior due to their charged moieties, influencing their and interactions in , though their net zero charge prevents them from behaving as typical electrolytes. The formation of zwitterions typically involves an intramolecular proton transfer within amphoteric molecules, such as , where the acidic carboxyl group (\ce{-COOH}) donates a proton to the basic amino group (\ce{-NH2}), yielding a anion (\ce{-COO^-}) and an cation (\ce{-NH3^+}). This process is pH-dependent, occurring predominantly under neutral or mildly basic conditions, as the favors the zwitterionic form when the solution aligns with the molecule's intrinsic tendencies. For instance, in , the simplest amino acid, the neutral form \ce{H2N-CH2-COOH} converts to the zwitterion ^+H3N-CH2-COO^- through this proton shift, a transformation supported by spectroscopic evidence like and NMR data. Key properties of zwitterions include high dipole moments stemming from the spatial separation of opposite charges, which enhances their polarity and solubility in polar solvents like water while conferring salt-like characteristics such as elevated melting points—for example, alanine's zwitterion melts around 300°C. Another critical property is the isoelectric point (pI), defined as the pH at which the molecule exists predominantly in its zwitterionic form with zero net charge, calculated as the average of the relevant pKa values for the ionizing groups. At the pI, zwitterions exhibit minimal electrophoretic mobility and altered solubility profiles compared to their charged forms. Common examples include amino acids like glycine and alanine, which display amphoteric behavior, and betaines such as glycine betaine (^+H3N-CH2-COO^- with a trimethylated ammonium), which form permanent zwitterions due to quaternary ammonium groups that prevent deprotonation. This amphoterism sets zwitterions apart from conventional polyatomic ions by enabling reversible ionization states responsive to environmental pH.

Polycharged Polyatomic Ions

Polycharged polyatomic ions are charged species composed of multiple atoms covalently bonded together, possessing an absolute charge magnitude greater than one due to the gain or loss of multiple electrons. These ions contrast with monocharged polyatomic ions by exhibiting enhanced electrostatic interactions, which influence their reactivity and stability. Examples include the sulfate ion (SO₄²⁻) and phosphate ion (PO₄³⁻), where the central atom coordinates with oxygen atoms, resulting in a net charge from multiple electron transfers. Such ions often form via stepwise processes involving polyprotic acids or bases, where protons are sequentially removed or added. In the case of (H₃PO₄), a triprotic acid, occurs in stages: first to dihydrogen phosphate (H₂PO₄⁻), then to (HPO₄²⁻), and finally to (PO₄³⁻), with each step governed by distinct acid dissociation constants that reflect the increasing difficulty of removing subsequent protons. This stepwise formation allows for the existence of intermediate polycharged species under varying conditions. Similarly, (H₂SO₄) yields the hydrogen sulfate ion (HSO₄⁻) and then the sulfate ion (SO₄²⁻). The presence of higher charges in these ions leads to significant electrostatic repulsion among the excess electrons (in anions) or electron deficiencies (in cations), which can destabilize the structure and promote reactivity. This repulsion is particularly pronounced in triply charged ions like PO₄³⁻, where the concentrated negative charge increases the ion's tendency to accept protons or coordinate with counterions. Stability is achieved through mechanisms such as , which delocalizes the charge across the ion's framework—for instance, in SO₄²⁻, multiple equivalent structures distribute the charge evenly among oxygen atoms—and , where solvent molecules (e.g., ) form a shell that screens the charge and reduces inter-ionic repulsions./Coordination_Chemistry/Complex_Ion_Chemistry/Acidity_of_the_Hexaaqua_Ions) Representative examples of polycharged polyatomic anions include the series derived from : H₂PO₄⁻ (charge -1, but part of the progression), HPO₄²⁻ (doubly charged), and PO₄³⁻ (triply charged), which play key roles in biological and environmental systems due to their variable states. Polycharged polyatomic cations are rarer and typically involve coordination complexes, such as the hexaqua ion ([Fe(H₂O)₆]³⁺), where the high +3 charge is partially offset by strong and ligand field stabilization, though it renders the highly acidic in aqueous media./Coordination_Chemistry/Complex_Ion_Chemistry/Acidity_of_the_Hexaaqua_Ions)

Examples and Applications

Common Examples

Polyatomic ions are groups of atoms that carry a net charge and behave as a single unit in chemical reactions. Among the most prevalent are anions such as (SO₄²⁻), which consists of a central atom surrounded by four oxygen atoms and carries a 2− charge, derived from the parent acid (H₂SO₄); a common salt is (Na₂SO₄). The ion (NO₃⁻), featuring a central atom bonded to three oxygen atoms with a 1− charge, originates from (HNO₃) and forms salts like (NH₄NO₃). (PO₄³⁻), with a central atom linked to four oxygen atoms and a 3− charge, comes from (H₃PO₄) and is found in compounds such as (Na₃PO₄). Cations among polyatomic ions include the (NH₄⁺), comprising a central atom bonded to four atoms with a 1+ charge, derived from the base (NH₃); it appears in salts like (NH₄NO₃). The (H₃O⁺), formed by a proton attaching to a and carrying a 1+ charge, arises from (H₂O) in acidic solutions. Less common but notable polyatomic ions include the cyanide ion (CN⁻), a linear diatomic anion with a 1− charge from hydrocyanic acid (HCN), often seen in (NaCN). The permanganate ion (MnO₄⁻), featuring a central atom bonded to four oxygen atoms with a 1− charge, derives from (HMnO₄) and is present in (KMnO₄). The following table summarizes these common polyatomic ions, including their formulas, charges, and parent acids or bases:
NameFormulaChargeParent Acid/Base
SO₄²⁻-2 (H₂SO₄)
NO₃⁻-1 (HNO₃)
PO₄³⁻-3 (H₃PO₄)
NH₄⁺+1 (NH₃)
HydroniumH₃O⁺+1 (H₂O)
CN⁻-1Hydrocyanic acid (HCN)
MnO₄⁻-1 (HMnO₄)

Practical Uses

Polyatomic ions play a crucial role in analytical chemistry, particularly in qualitative identification and quantitative titrations. In qualitative tests, polyatomic ions such as sulfate (SO₄²⁻) and carbonate (CO₃²⁻) are detected through precipitation reactions; for instance, adding barium chloride to a solution containing sulfate ions forms an insoluble barium sulfate precipitate, confirming the presence of the ion. Similarly, tests for halide ions, which can form polyhalide species like triiodide (I₃⁻), involve silver nitrate, where the formation of colored silver halide precipitates (e.g., pale yellow for chloride, yellow for bromide) allows differentiation. In titrations, polyprotic acids containing polyatomic anions, such as sulfuric acid (H₂SO₄) with its sulfate ion, exhibit multiple equivalence points due to stepwise proton donation, enabling precise determination of acid concentration through pH monitoring. In industrial applications, polyatomic ions are integral to s, detergents, and systems. (NH₄NO₃), featuring the (NH₄⁺) and (NO₃⁻) ions, serves as a high-nitrogen that provides essential nutrients for crop growth, with global production of approximately 21 million tons as of 2022, primarily for agricultural use. Phosphate ions (PO₄³⁻) in detergents act as builders to soften water and enhance cleaning efficiency by sequestering calcium and magnesium ions, though their use has declined due to environmental regulations. In lead-acid batteries, the ion (SO₄²⁻) from electrolyte participates in the electrochemical reaction, forming lead sulfate (PbSO₄) on the electrodes during discharge, which reverses upon charging to sustain the battery's operation in vehicles and uninterruptible power supplies. Biochemically, polyatomic ions are vital for , nutrient cycling, and . The phosphate ion is central to (ATP) and (ADP), where the transfer of the terminal phosphate group from ATP to ADP releases energy (approximately 30.5 kJ/mol under standard conditions) to drive cellular processes like and . Nitrate ions (NO₃⁻) are key intermediates in the , assimilated by plants via enzymes and reduced to for , supporting global primary productivity. The charged groups in zwitterionic forms of , such as the ammonium (NH₃⁺) and carboxylate (COO⁻) moieties, enable formation in proteins, contributing to their three-dimensional folding and functional stability in enzymes and structural components. Environmentally, polyatomic ions from human activities contribute to pollution challenges. Nitrate ions from fertilizers like leach into and surface waters, causing and health risks such as at concentrations above 10 mg/L, with agricultural runoff accounting for up to 70% of loading in some watersheds. Sulfate ions in , primarily from oxidation in the atmosphere, lower precipitation to below 5.6, leading to , forest decline, and damage by mobilizing toxic aluminum ions.

References

  1. [1]
    Rules for Naming Ionic Compounds Containing Polyatomic Ions
    Polyatomic ions are ions which consist of more than one atom. For example, nitrate ion, NO3-, contains one nitrogen atom and three oxygen atoms. The atoms in a ...
  2. [2]
    5.3 Polyatomic Ions
    Polyatomic ionsA group of two or more atoms that has a net electrical charge. are groups of atoms that bear a net electrical charge, although the atoms in a ...
  3. [3]
    Table of Polyatomic Ions - gchem
    Polyatomic ions are multiple atoms covalently bonded, stable as charged ions. Examples include ammonium (NH4+), nitrite (NO2-), and sulfate (SO42-).Missing: definition | Show results with:definition
  4. [4]
    Memorizing Polyatomic Ions
    Polyatomic ions are made up of more than one element and have an overall charge. Examples are SO42-, OH-, and NH4+. Polyatomic ions are important in chemistry ...Missing: definition | Show results with:definition
  5. [5]
    Polyatomic ions - (Physical Science) - Vocab, Definition, Explanations
    Definition. Polyatomic ions are ions that consist of two or more atoms bonded together, carrying a net charge due to the loss or gain of electrons.
  6. [6]
    Polyatomic ions (article) | Khan Academy
    Polyatomic ions are a type of covalent compound, not ionic compound. Recall that all atoms work to fulfil their valence shell. There are a few ways it can be ...
  7. [7]
    CH104: Chapter 3 - Ions and Ionic Compounds - Chemistry
    Polyatomic ions can be thought of in a very similar way to monoatomic ions, in that they are ionized by either gaining or losing electrons so that they carry a ...
  8. [8]
    Polyatomic Ions (M2Q6) – UW-Madison Chemistry 103/104 ...
    Polyatomic ions are electrically charged molecules, formed from more than one atom, acting as discrete units. Ocyanions are polyatomic ions with one or more ...
  9. [9]
  10. [10]
    CH150: Chapter 3 - Ions and Ionic Compounds - Chemistry
    Polyatomic ions are ions that form from multiple atoms that are covalently bonded together. Polyatomic ions behave as a single group when participating in ...
  11. [11]
    5.3 Molecular and Ionic Compounds - Maricopa Open Digital Press
    The ions that we have discussed so far are called monatomic ions, that is, they are ions formed from only one atom. We also find many polyatomic ions. These ...Missing: differences | Show results with:differences
  12. [12]
    Ammonium (PAMDB001595) - P. aeruginosa Metabolome Database
    It is formed by the protonation of ammonia (NH3). Ammonium is also a ... ammonium ion (CHEBI:28938 ); nitrogen hydride (CHEBI:28938 ); onium cation ...
  13. [13]
    [PDF] CHEM& 121 Workbook - Bellevue College
    ... charge; these ions are monatomic ions. Polyatomic ions are formed when a group of atoms collectively gain or lose electrons. Ions. Compound Formula ammonium.
  14. [14]
    6.3 Formal Charges and Resonance – Chemistry Fundamentals
    The actual electronic structure of the molecule (the average of the resonance forms) is called a resonance hybrid of the individual resonance forms. A double- ...
  15. [15]
    [PDF] variations in infrared spectra, molecular symmetry and site ... - RRuff
    Polyatomic molecules, such as the sulfate ion, SOa2-, perform compli- cated vibrational motions which are resolved into specific sets of vibra- tions or normal ...
  16. [16]
    [PDF] Infrared Spectroscopy of Buffer-Gas Cooled Molecules and ... - JILA
    Infrared absorption spectroscopy is a highly sensitive technique for probing molecular struc- ture, and when combined with cold molecular beam methods it ...
  17. [17]
    2.6 Ionic and Molecular Compounds - Chemistry 2e | OpenStax
    Feb 14, 2019 · We also find many polyatomic ions. These ions, which act as discrete units, are electrically charged molecules (a group of bonded atoms with an ...
  18. [18]
    Illustrated Glossary of Organic Chemistry - Sulfate
    The conjugate base formed by deprotonation of bisulfate ion, or by double deprotonation of sulfuric acid.
  19. [19]
    [PDF] Radiation-Induced Formation of Chlorine Oxides and Their Potential ...
    Mar 18, 2013 · Perchlorates are extremely stable, making them suitable as ... chloride ion to perchlorate in aqueous solution.10 This pathway.Missing: chlorite | Show results with:chlorite
  20. [20]
    Polyatomic Cation - an overview | ScienceDirect Topics
    Polyatomic cations are defined as positively charged species composed of two or more atoms, which can include elements such as sulfur, selenium, or tellurium, ...Missing: reliable | Show results with:reliable
  21. [21]
    3.4: Polyatomic Ions and Formulae for Ionic Compounds
    Sep 11, 2024 · The ammonium ion contains one nitrogen and four hydrogen's that collectively bear a +1 charge. The hydronium ion contains one oxygen and three ...Missing: protonation | Show results with:protonation
  22. [22]
    Common Polyatomic Ion with Charge +1 - BYJU'S
    List of Common Polyatomic Ions. Common Polyatomic Ion with Charge +1. Formula. Name. [NH4]+. Ammonium. [H3O]+. Hydronium. Common Polyatomic Ion with Charge -1 ...
  23. [23]
    Ammonium | Formula, Symbol & Structure - Lesson - Study.com
    Lesson Summary. To sum up, ammonium is a positively-charged, polyatomic ion that forms when neutral ammonia is protonated by an acid. Protonation is taking on ...Missing: hydronium | Show results with:hydronium
  24. [24]
    Hydronium ion - American Chemical Society
    Sep 28, 2020 · Hydronium ion (H3O+) is protonated water, formed when acids add to water. It's the simplest form of oxonium ions.
  25. [25]
    The Hydronium Ion - Chemistry LibreTexts
    Jan 29, 2023 · Free Hydrogen Ions do not Exist in Water. The hydrogen ion in aqueous solution is no more than a proton, a bare nucleus.
  26. [26]
    [PDF] Nomenclature of Inorganic Chemistry | IUPAC
    The present book supersedes not only Red Book I but also, where appropriate,. Nomenclature of Inorganic Chemistry II, IUPAC Recommendations 2000 (Red Book II).
  27. [27]
    [PDF] Brief Guide to the Nomenclature of Inorganic Chemistry | IUPAC
    Additive nomenclature was developed in order to describe the structures of coordination entities, or complexes, but this method is readily extended to other ...
  28. [28]
    [PDF] PDF - IUPAC nomenclature
    The concept of preferred IUPAC names as applied to radicals and ions is based on the following principles. (1) substitutive nomenclature based on carbane ...
  29. [29]
    Nomenclature of Coordination Complexes - Chemistry LibreTexts
    Jun 30, 2023 · A last little side note: when naming a coordination compound, it is important that you name the cation first, then the anion. You base this on ...Introduction · Rule 1: Anionic Ligands · Rule 2: Neutral Ligands · Rule 4: The Metals
  30. [30]
    Covalent Bonding – Chemistry
    The atoms in polyatomic ions, such as OH–, NO 3 − , and NH 4 + , are held together by polar covalent bonds. However, these polyatomic ions form ionic compounds ...
  31. [31]
    Lewis structures - CHEM 101
    Nov 1, 2023 · Lewis structures are structural formulas for molecules and polyatomic ions that represent all valence electrons.
  32. [32]
    9.1 Predicting the Geometry of Molecules and Polyatomic Ions
    We can use the VSEPR model to predict the geometry of most polyatomic molecules and ions by focusing on only the number of electron pairs around the central ...
  33. [33]
    Molecular Geometry
    ... polyatomic ions. We will use a model called the Valence Shell Electron-Pair Repulsion (VSEPR) model that is based on the repulsive behavior of electron-pairs.
  34. [34]
    [PDF] The Shapes of Molecules - Web Pages
    – The bond order is (2+1+1)/3 = 1.33. Example: Benzene, C6H6. • The two ... – Polyatomic molecules where the bond dipoles do not cancel each other. H2O ...Missing: length | Show results with:length
  35. [35]
    Resonance - Chemistry LibreTexts
    Jan 29, 2023 · Resonance is a way of describing delocalized electrons within certain molecules or polyatomic ions where the bonding cannot be expressed by a single Lewis ...
  36. [36]
    7.5: The Predicted Stabilities of Resonance Contributors
    Oct 20, 2020 · In general, molecules with multiple resonance structures will be more stable than one with fewer and some resonance structures contribute more to the stability ...
  37. [37]
  38. [38]
    Natural zwitterionic betaine enables cells to survive ultrarapid ... - NIH
    Nov 22, 2016 · Zwitterionic betaine is well known for its highly hydrophilic property; it can strongly bind water molecules via ionic solvation effects due to ...
  39. [39]
    6.17: Polyatomic Ions - Chemistry LibreTexts
    Jun 20, 2023 · Polyatomic ions, common in any lab, contain several atoms covalently bonded together. Often, these ions are charged and combine with metals ...
  40. [40]
  41. [41]
  42. [42]
    Acid Names and Anions
    SO · sulfate ion, ClO ; SO · sulfite ion, ClO ; NO · nitrate ion, ClO ; NO · nitrite ion, ClO ; PO · phosphate ion, HSO ...Missing: parent | Show results with:parent
  43. [43]
    Acid-Base Pairs, Strength of Acids and Bases, and pH
    Acids and Bases ; NaHSO4 (sodium hydrogen sulfate), 1.4 ; H2SO3 (sulfurous acid), 1.5 ; H3PO4 (phosphoric acid), 1.5 ; HF (hydrofluoric acid), 2.1 ; CH3CO2H (acetic ...Missing: parent | Show results with:parent
  44. [44]
    [PDF] Table of Common Polyatomic Ions | Bellevue College
    Table of Common Polyatomic Ions. +. 4. NH. Ammonium. +. OH. 3. Hydronium. -. OH. Hydroxide. -. 2. 3. CO. Carbonate. -. 3. HCO. Hydrogen carbonate.
  45. [45]
    21.2 The Chemistry of Hydrogen
    If the lone pair of electrons belongs to an oxygen atom of a water molecule, the result is the hydronium ion (H 3O +). ... ion, or form a covalent bond or polar ...
  46. [46]
    Elements, compounds, and mixtures
    Predict the formula of the compound that forms when magnesium metal ... permanganate, HMnO4, permanganic acid. CrO42-, chromate, H2CrO4, chromic acid. ClO3 ...
  47. [47]
    6: Qualitative Analysis of Group I Ions (Experiment)
    Sep 22, 2021 · In qualitative analysis, the ions in a mixture are separated by selective precipitation. Selective precipitation involves the addition of a ...
  48. [48]
    Testing for halide ions - Chemguide
    This page describes and explains the tests for halide ions (fluoride, chloride, bromide and iodide) using silver nitrate solution followed by ammonia solution.
  49. [49]
    Titrations of polyprotic acids (video) - Khan Academy
    Dec 9, 2021 · Titrating a polyprotic acid with a strong base produces a pH curve with as many equivalence points as there are acidic protons on the acid.
  50. [50]
    [PDF] 8.3 Ammonium Nitrate - U.S. Environmental Protection Agency
    High density prills, granules, and crystals are used as fertilizer, grains are used solely in explosives, and low density prills can be used as either. 8.3.2 ...
  51. [51]
    The contribution of household chemicals to environmental ...
    Detergent P contributes to eutrophication impacts from household sewage effluent. Approximately, 25% of P in wastewater is generated from detergents.
  52. [52]
    Lead Acid Batteries | PVEducation
    For example, in the lead acid battery, sulfate ions changes from being in solid form (as lead sulfate) to being in solutions (as sulfuric acid). If the lead ...<|separator|>
  53. [53]
    ATP/ADP - Chemistry LibreTexts
    Jul 4, 2022 · ATP is used for energy transfer in the cell, and it interconverts to ADP by adding or removing a phosphate group. ATP is the primary energy ...
  54. [54]
    The Nitrogen Cycle: Processes, Players, and Human Impact - Nature
    Nitrification is the process that converts ammonia to nitrite and then to nitrate and is another important step in the global nitrogen cycle. Most ...
  55. [55]
    26.1: Structures of Amino Acids - Chemistry LibreTexts
    Sep 30, 2024 · Amino acids exist in aqueous solution primarily in the form of a dipolar ion, or zwitterion (from the German zwitter, meaning “hybrid”).<|separator|>
  56. [56]
    Fertilizers and nitrate pollution of surface and ground water
    Mar 31, 2021 · Nitrate pollution of ground and surface water bodies all over the world is generally linked with continually increasing global fertilizer nitrogen (N) use.
  57. [57]
    National look at nitrate contamination of Ground Water
    Mar 20, 2025 · The risk of ground-water contamination by nitrate depends both on the nitrogen input to the land surface and the degree to which an aquifer is vulnerable.
  58. [58]
    Effects of Acid Rain | US EPA
    Mar 19, 2025 · In the atmosphere, SO2 and NOX gases can be transformed into sulfate and nitrate particles, while some NOX can also react with other pollutants ...