Main-group element
Main-group elements, also known as representative elements, are the chemical elements belonging to the s- and p-blocks of the periodic table, encompassing groups 1, 2, and 13 through 18.[1] These elements include the highly reactive alkali metals (group 1) and alkaline earth metals (group 2), as well as a diverse array of metalloids, nonmetals, and post-transition metals in groups 13–18, such as the pnictogens (group 15), chalcogens (group 16), halogens (group 17), and noble gases (group 18).[1] Their electron configurations, with valence electrons primarily in s and p orbitals, lead to predictable chemical behaviors and bonding patterns that form the foundation of much of inorganic and organic chemistry.[2] Main-group elements dominate the composition of the Earth's crust, accounting for over 93% of its mass through abundant species like oxygen (46.6%), silicon (27.7%), aluminum (8.1%), calcium (3.6%), sodium (2.8%), potassium (2.6%), and magnesium (2.1%).[3] They exhibit wide-ranging properties, from the extreme reactivity of alkali metals that ignite in air to the inertness of noble gases, and play critical roles in biological systems (e.g., carbon and nitrogen in organic molecules) and industrial processes (e.g., chlorine in disinfectants, silicon in semiconductors).[1] Periodic trends in atomic size, ionization energy, electronegativity, and redox potentials across these groups enable systematic predictions of their reactivity and compound formation.[2]Definition and Classification
Definition
Main-group elements, also known as representative elements, are the chemical elements located in the s- and p-blocks of the periodic table, encompassing groups 1, 2, and 13 through 18. These elements fill their outermost electron shells using only s and p orbitals, in contrast to the d-block transition metals and f-block inner transition metals, which involve d and f orbitals, respectively. This classification highlights their role in exhibiting straightforward valence behaviors and forming the foundational structure of chemical periodicity.[4] The term "main-group" emerged from early 20th-century periodic table notations, particularly the long form, where these elements formed continuous vertical columns (traditionally labeled as A subgroups) uninterrupted by the inserted transition series in the short form. This distinction was formalized around the time transition metals were explicitly named in 1921, emphasizing the primary, representative nature of the s- and p-block elements in demonstrating periodic trends. The synonymous term "representative elements" underscores their prevalence and typical chemical behaviors, as they constitute the majority of elements in the universe and Earth's crust.[5][6] The main-group elements comprise 44 known members from atomic number 1 (hydrogen) to 88 (radium), excluding the superheavy synthetic elements (atomic numbers 113–118). They are distributed as follows:| Group | Elements (Atomic Numbers) |
|---|---|
| 1 (Alkali metals) | H (1), Li (3), Na (11), K (19), Rb (37), Cs (55), Fr (87)* |
| 2 (Alkaline earth metals) | Be (4), Mg (12), Ca (20), Sr (38), Ba (56), Ra (88)* |
| 13 | B (5), Al (13), Ga (31), In (49), Tl (81) |
| 14 | C (6), Si (14), Ge (32), Sn (50), Pb (82) |
| 15 | N (7), P (15), As (33), Sb (51), Bi (83) |
| 16 | O (8), S (16), Se (34), Te (52), Po (84) |
| 17 (Halogens) | F (9), Cl (17), Br (35), I (53), At (85) |
| 18 (Noble gases) | He (2), Ne (10), Ar (18), Kr (36), Xe (54), Rn (86) |
Position in the Periodic Table
Main-group elements are positioned in the s-block and p-block of the periodic table, which together comprise the representative elements. The s-block includes groups 1 and 2, situated on the extreme left side of the table, encompassing the alkali metals (group 1) and alkaline earth metals (group 2), along with hydrogen in group 1. The p-block occupies groups 13 through 18 on the right side, including elements such as the boron group (13), carbon group (14), nitrogen group (15), chalcogens (16), halogens (17), and noble gases (18). This arrangement visually separates main-group elements from the central and lower portions of the table, highlighting their role in forming the table's outer framework.[7][8] The block classification of main-group elements derives from the filling of their outermost electron orbitals. In the s-block, the valence electrons occupy the ns orbital, where n represents the principal quantum number of the period. For the p-block, valence electrons fill the np orbitals, leading to electron configurations that end in s¹ or s² for s-block elements and varying np¹ to np⁶ for p-block elements. This orbital-based categorization distinguishes main-group elements from others, as it emphasizes the involvement of s and p electrons in their chemical behavior.[7][9] In comparison, the d-block, known as transition metals, spans groups 3 through 12 in the central region, where valence electrons fill (n-1)d orbitals alongside ns. The f-block consists of the lanthanides (period 6) and actinides (period 7), positioned below the main body, with valence electrons in (n-2)f orbitals. Main-group elements are thus excluded from these inner transition and transition categories, focusing solely on s- and p-block occupancy to underscore their distinct periodic properties.[7][8] Periodic table formats vary, with the conventional 18-column layout contracting the f-block into separate rows for compactness, while the extended 32-column format integrates the f-block inline to reflect full orbital sequences. Despite these differences, the positions of s-block (groups 1-2) and p-block (groups 13-18) elements remain consistent across both formats, preserving the structural integrity of main-group classifications.[10][11]Physical Properties
Atomic and Electronic Structure
Main-group elements, comprising the s-block and p-block of the periodic table, exhibit characteristic electron configurations that determine their chemical behavior. The s-block elements (groups 1 and 2) follow the general pattern [ \text{noble gas core} ] ns^1 for group 1 and [ \text{noble gas core} ] ns^2 for group 2, where n is the principal quantum number corresponding to the valence shell. For example, lithium (Li, atomic number 3) has the configuration [\ce{He}] 2s^1, sodium (Na, atomic number 11) is [\ce{Ne}] 3s^1, and magnesium (Mg, atomic number 12) is [\ce{Ne}] 3s^2. These configurations arise from the filling of the outermost s orbital with one or two electrons after the inner shells are complete.[12][13] The p-block elements (groups 13–18) have valence electron configurations of the form [ \text{noble gas core} ] ns^2 np^{1-6}, where the p subshell accommodates up to six electrons, leading to a total of 3 to 8 valence electrons. Representative examples include aluminum (Al, atomic number 13) as [\ce{Ne}] 3s^2 3p^1, silicon (Si, atomic number 14) as [\ce{Ne}] 3s^2 3p^2, phosphorus (P, atomic number 15) as [\ce{Ne}] 3s^2 3p^3, sulfur (S, atomic number 16) as [\ce{Ne}] 3s^2 3p^4, chlorine (Cl, atomic number 17) as [\ce{Ne}] 3s^2 3p^5, and argon (Ar, atomic number 18) as [\ce{Ne}] 3s^2 3p^6. For heavier periods, thallium (Tl, atomic number 81) is [\ce{Xe}] 4f^{14} 5d^{10} 6s^2 6p^1, and lead (Pb, atomic number 82) is [\ce{Xe}] 4f^{14} 5d^{10} 6s^2 6p^2. These patterns reflect the sequential addition of electrons to s and p orbitals within the valence shell.[12][13] The orbital filling in main-group elements adheres to the Aufbau principle, which dictates that electrons occupy orbitals of lowest energy first, progressing from s to p subshells within each principal level. This principle ensures that the 1s orbital fills before 2s, followed by 2p, then 3s and 3p, and so on, without involvement of d or f orbitals in the valence shells of these elements. Exceptions are minimal for main-group atoms, as the energy ordering of s and p orbitals remains straightforward across periods.[14][13] A notable feature in heavier p-block elements is the inert pair effect, where the ns^2 electrons become less reactive due to relativistic stabilization and poor shielding by inner d and f electrons, favoring lower oxidation states. For instance, thallium exhibits a stable +1 state (using only the 6p electron) over +3, as in TlCl, while lead prefers +2 (retaining the 6s² pair) in compounds like PbCl₂ rather than +4. This effect strengthens down groups 13–15, influencing the stability of oxidation states two units below the group valence.[15][16]Trends in Physical Properties
The atomic radius of main-group elements exhibits systematic trends influenced by electron configuration and nuclear charge. Across a period, atomic radius decreases from left to right due to increasing effective nuclear charge, which pulls electrons closer to the nucleus without adding new shells. Down a group, atomic radius increases as additional electron shells are added, outweighing the effect of increased nuclear charge. For s-block elements, metallic radii are typically used, while for p-block, covalent or van der Waals radii apply depending on bonding. In Group 1, radii range from 152 pm for lithium to 265 pm for cesium, illustrating the downward increase.[17][18]| Element | Atomic Radius (pm, metallic) |
|---|---|
| Li | 152 |
| Na | 186 |
| K | 231 |
| Rb | 244 |
| Cs | 265 |
Chemical Properties
Valence and Bonding
Main-group elements possess valence electrons in their outermost s and p orbitals, ranging from 1 to 8 electrons, which primarily govern their chemical reactivity and bonding preferences.[24] These valence electrons determine the elements' tendency to achieve a stable electron configuration, often by adhering to the octet rule, wherein atoms form bonds to attain eight electrons in their valence shell, mimicking the noble gas configuration.[25] This rule is particularly applicable to main-group elements, as their valence shells lack the d-orbital involvement seen in transition metals, leading to predictable bonding patterns based on electron gain, loss, or sharing.[26] The bonding behavior of main-group elements varies by group and electronegativity differences with partners. S-block elements, with low ionization energies, typically form ionic bonds by transferring valence electrons to highly electronegative atoms, as exemplified by sodium chloride (NaCl), where Na donates its single valence electron to Cl, resulting in Na⁺ and Cl⁻ ions.[27] In contrast, p-block elements often engage in covalent bonding through electron sharing to complete their octet, such as in methane (CH₄), where carbon shares four pairs of electrons with four hydrogens.[24] Metallic bonding predominates within the elemental forms of main-group metals, particularly in s-block groups, where delocalized valence electrons create a "sea" of electrons holding positive ions together, contributing to their conductivity and malleability.[27] Oxidation states in main-group elements reflect the number of valence electrons lost or gained during bonding. For s-block elements, oxidation states are fixed and equal to their group number; Group 1 elements, like sodium, consistently exhibit +1 due to loss of their single ns¹ electron. P-block elements display variable oxidation states spanning positive and negative values, depending on bonding context; for instance, carbon commonly shows +4 in carbon dioxide (CO₂) and -4 in methane (CH₄), while nitrogen ranges from -3 in ammonia (NH₃) to +5 in nitric acid (HNO₃).[28] In p-block elements, valence orbitals often hybridize to optimize bond geometry and achieve the octet. The sp³ hybridization occurs when one s and three p orbitals mix to form four equivalent orbitals, as in methane (CH₄), enabling tetrahedral geometry.[29] Sp² hybridization, using one s and two p orbitals, produces three planar orbitals, seen in boron trifluoride (BF₃) with trigonal planar arrangement around boron.[30] Sp hybridization, involving one s and one p orbital, results in two linear orbitals, facilitating triple bonds in compounds like acetylene (C₂H₂).[31] These hybridizations enhance orbital overlap, stabilizing covalent bonds in p-block species.[30]Reactivity Patterns
Main-group elements display distinct reactivity patterns influenced by their position in the periodic table, particularly within the s- and p-blocks. In the s-block (Groups 1 and 2), reactivity generally increases down each group due to progressively lower ionization energies, which facilitate easier loss of valence electrons and formation of cations. For instance, the alkali metals in Group 1 exhibit heightened reactivity toward water as one descends from lithium to cesium, with cesium reacting explosively while lithium reacts more mildly.[32] Similarly, in Group 2, the alkaline earth metals show increasing reactivity with water down the group, from beryllium's inertness to barium's vigorous hydrogen evolution.[33] In contrast, p-block reactivity trends vary by group; for the halogens in Group 17, reactivity decreases down the group owing to larger atomic radii, reduced electronegativity, and weaker X-X bonds, making fluorine the most reactive oxidant while iodine is the least.[34] Key reaction types for main-group elements include redox processes and acid-base reactions, reflecting their electron-donating or -accepting tendencies. Redox reactions are prominent in s-block elements, where metals act as strong reducing agents; a classic example is the reaction of sodium with water, producing hydrogen gas and sodium hydroxide:$2\mathrm{Na}(s) + 2\mathrm{H_2O}(l) \rightarrow 2\mathrm{NaOH}(aq) + \mathrm{H_2}(g)
This exothermic reaction underscores the metals' ability to reduce water to hydrogen.[32] Acid-base reactions are common with p-block and s-block oxides, which exhibit amphoteric or basic character; for example, magnesium oxide from Group 2 reacts with hydrochloric acid to form a salt and water:
\mathrm{MgO}(s) + 2\mathrm{HCl}(aq) \rightarrow \mathrm{MgCl_2}(aq) + \mathrm{H_2O}(l)
Such reactions highlight the basic nature of early main-group oxides toward protic acids.[35] Hydride formation provides another insight into reactivity patterns across the main groups. Elements in Groups 1 and 2, being highly electropositive, form ionic hydrides where hydrogen adopts the hydride ion (H⁻), as seen in sodium hydride (NaH), which reacts violently with water to liberate hydrogen gas.[36] In Groups 13 through 16, less electropositive elements produce covalent hydrides through shared electron pairs, exemplified by methane (CH₄) from carbon and ammonia (NH₃) from nitrogen, which vary in stability and volatility based on molecular structure.[36] Elements in Group 17 form covalent binary hydrides (HX), which are acidic gases or liquids, while elements in Group 18 (noble gases) do not form stable binary hydrides under standard conditions due to their low reactivity.[36] The stability of main-group compounds, particularly toward thermal decomposition, follows predictable trends tied to ionic size and lattice energy. In Group 2, carbonates exhibit increasing thermal stability down the group; magnesium carbonate (MgCO₃) decomposes readily upon heating to yield magnesium oxide and carbon dioxide, whereas barium carbonate (BaCO₃) requires much higher temperatures for decomposition due to the larger, less polarizing Ba²⁺ cation stabilizing the carbonate ion.[37] This pattern arises from decreasing charge density of the cations, reducing their ability to polarize and destabilize the carbonate anion.[37]
S-block Elements
Group 1: Alkali Metals
The Group 1 elements of the periodic table, collectively known as the alkali metals, include lithium (Li), sodium (Na), potassium (K), rubidium (Rb), caesium (Cs), and francium (Fr). These elements are highly reactive, monovalent metals that exist primarily as +1 cations in their compounds due to their low ionization energies. Francium, the heaviest member, is extremely rare and radioactive, with its most stable isotope having a half-life of only 22 minutes, limiting detailed study of its properties.[38] The alkali metals exhibit distinctive physical properties, including exceptional softness and low densities. They are malleable and ductile, allowing them to be cut with a knife, and their fresh surfaces appear silvery-white before rapidly tarnishing in air.[39] Lithium, in particular, has a density of 0.534 g/cm³, making it the least dense metal and enabling it to float on water, while sodium (0.968 g/cm³) and potassium (0.862 g/cm³) also have densities below that of water.[40] These properties stem from their large atomic radii and weak metallic bonding, with melting points decreasing down the group from 180.5°C for lithium to an estimated 27°C for francium.[41] Chemically, the alkali metals are renowned for their vigorous reactions with water, producing hydrogen gas and metal hydroxides that yield alkaline solutions. The general reaction is represented by the equation: $2\mathrm{M} + 2\mathrm{H_2O} \rightarrow 2\mathrm{MOH} + \mathrm{H_2} where M denotes an alkali metal.[42] Lithium reacts steadily with cold water, fizzing gently and moving across the surface if placed on filter paper, while sodium reacts more vigorously, melting into a molten ball that skims rapidly; potassium exhibits even greater intensity, often igniting the hydrogen produced.[42] Flame tests provide a diagnostic tool for identifying these elements, as their compounds impart characteristic colors to a flame: crimson-red for lithium, intense yellow for sodium, lilac for potassium, red-violet for rubidium, and blue for caesium.[43] Key compounds of the alkali metals include sodium chloride (NaCl), which occurs naturally as the mineral halite and serves as the primary source of sodium, essential for food preservation and chemical manufacturing.[44] Potassium nitrate (KNO₃), known as saltpeter, is historically significant for gunpowder production and is now widely used in fertilizers to supply potassium and nitrogen to plants.[44] Lithium plays a pivotal role in lithium-ion batteries, where lithium ions intercalate between graphite anodes and metal oxide cathodes during charge-discharge cycles, enabling high energy density for applications in portable electronics and electric vehicles.[45] An notable anomaly in Group 1 is the diagonal relationship between lithium and magnesium, arising from their comparable charge-to-radius ratios and polarizing power despite belonging to adjacent groups.[46] This similarity manifests in shared behaviors, such as the formation of stable nitrides (Li₃N and Mg₃N₂) and comparable solubilities for certain salts, like the low solubility of lithium carbonate and magnesium carbonate in water.[46]Group 2: Alkaline Earth Metals
The Group 2 elements of the periodic table, known as the alkaline earth metals, consist of beryllium (Be), magnesium (Mg), calcium (Ca), strontium (Sr), barium (Ba), and radium (Ra). These metals are silvery-white, malleable, and ductile solids at room temperature, with beryllium being notably hard and the others softer. Compared to the Group 1 alkali metals, they possess higher densities—ranging from 1.85 g/cm³ for beryllium to 5.0 g/cm³ for radium—and higher melting points, such as 650 °C for magnesium and 842 °C for calcium, due to their smaller atomic radii, +2 oxidation state, and enhanced metallic bonding from the additional valence electron.[47][48] Radium, however, is radioactive, with its most stable isotope radium-226 having a half-life of 1600 years, and exists only in trace amounts in nature.[49] The alkaline earth metals exhibit moderate reactivity, less vigorous than that of the alkali metals, primarily forming divalent cations (M²⁺) in ionic compounds. Their reactions with water are slower and depend on the element: beryllium and magnesium do not react with cold water, but magnesium reacts with hot water or steam to yield magnesium hydroxide and hydrogen gas via Mg + 2H₂O → Mg(OH)₂ + H₂. Calcium reacts slowly with cold water to produce calcium hydroxide and hydrogen, while strontium and barium react more readily at room temperature, generating the corresponding hydroxides (M(OH)₂) and hydrogen gas. These elements also burn in air to form basic oxides (MO), which react with water to produce the hydroxides, contributing to their classification as "alkaline earths" from early observations of their insoluble, alkaline soil residues.[50]/Descriptive_Chemistry/Elements_Organized_by_Block/1_s-Block_Elements/Group__2_Elements:_The_Alkaline_Earth_Metals/2.1:_Chemical_Properties_of_the_Alkaline_Earth_Metals) Notable compounds include calcium carbonate (CaCO₃), the primary component of limestone and a key mineral in sedimentary rocks, and magnesium sulfate (MgSO₄·7H₂O), commonly known as Epsom salt for its use in baths and as a laxative. Beryllium deviates from the group's ionic character due to its small ionic radius and high charge density, leading to covalent bonding in compounds like beryllium chloride (BeCl₂), which exhibits tetrahedral coordination and volatility. Beryllium is highly toxic, with inhalation of its dust or fumes causing berylliosis, a chronic lung disease, and it is classified as a human carcinogen by the International Agency for Research on Cancer.[51]/12:Group_2-_The_Alkaline_Earth_Metals/12.03:_The_Group_2_elements) Biologically, calcium is vital for human bone and tooth mineralization, forming hydroxyapatite crystals, and serves as a signaling ion in muscle contraction and nerve transmission. Magnesium is central to chlorophyll's structure in plants, facilitating photosynthesis by stabilizing the porphyrin ring, and in humans, it acts as a cofactor in over 300 enzymes involved in ATP hydrolysis and DNA replication.[52][53][54]P-block Elements
Groups 13–16: Representative Metals, Metalloids, and Nonmetals
Groups 13 through 16 of the periodic table encompass the early p-block elements, transitioning from metalloids and nonmetals to representative metals, with properties influenced by increasing atomic size and metallic character down each group.[55] These elements exhibit diverse bonding behaviors, including covalent and ionic characteristics, and play key roles in materials science due to their semiconducting and catalytic properties.[56] Group 13, known as the boron group, includes boron (B), aluminum (Al), gallium (Ga), indium (In), thallium (Tl), and the synthetic nihonium (Nh). Boron is a metalloid with high melting point and covalent network structure, while aluminum is a lightweight metal widely used in alloys.[56] Gallium, indium, and thallium show increasing density and lower melting points, with thallium exhibiting toxicity. Nihonium, element 113, is highly radioactive with a half-life of seconds, and its properties are largely predicted based on relativistic effects. A notable chemical feature is the amphoteric behavior of aluminum hydroxide, Al(OH)<sub>3</sub>, which dissolves in both acids and bases to form aluminates or alums.[55] Gallium arsenide (GaAs) is a key III-V semiconductor material with a direct bandgap of 1.42 eV, enabling efficient light emission in LEDs and lasers. In Group 14, the carbon group, elements carbon (C), silicon (Si), germanium (Ge), tin (Sn), lead (Pb), and flerovium (Fl) demonstrate a progression from nonmetal to metal. Carbon's unique ability to catenate forms long chains and rings, underpinning organic chemistry, and its allotropes include insulating diamond and conductive graphite.[57] Silicon and germanium are metalloids used in semiconductors, with silicon forming the basis of integrated circuits due to its abundance and stable oxide layer. Tin and lead are soft metals; lead's +2 oxidation state is stabilized in applications like lead-acid batteries, where Pb and PbO<sub>2</sub> electrodes facilitate reversible reactions in sulfuric acid electrolyte, providing high surge current capacity.[57] Flerovium, element 114, is synthetic and volatile, with predicted metallic properties influenced by relativistic stabilization of the 7s electrons. Group 15 elements, the pnictogens, comprise nitrogen (N), phosphorus (P), arsenic (As), antimony (Sb), bismuth (Bi), and moscovium (Mc). Nitrogen exists primarily as stable N<sub>2</sub>, requiring energy-intensive fixation processes like the Haber-Bosch method to convert it to ammonia for fertilizers. Phosphorus has allotropes including reactive white phosphorus (P<sub>4</sub> tetrahedra) and polymeric red phosphorus, with white form igniting spontaneously in air. Arsenic and antimony are metalloids with toxic compounds; arsenic trioxide causes acute poisoning by inhibiting enzymes, while antimony compounds affect cardiac function. Bismuth stands out as the most metallic, with a low thermal conductivity and rhombohedral crystal structure, used in low-melting alloys.[58] Moscovium, element 115, is a synthetic superheavy element that is highly radioactive, with the longest-lived isotope (Mc-289) having a half-life of about 0.8 seconds; its properties are predicted, including a likely solid metal state and stabilization of the +1 oxidation state due to the inert pair effect.[59] The chalcogens in Group 16 include oxygen (O), sulfur (S), selenium (Se), tellurium (Te), polonium (Po), and livermorium (Lv). Oxygen's O<sub>2</sub> molecule is paramagnetic due to two unpaired electrons in its ground state, enabling its role in combustion and respiration.[60] Sulfur participates in biogeochemical cycles, oxidizing to sulfate in aerobic environments and reducing to sulfide anaerobically, influencing global redox balance. Selenium and tellurium find applications in electronics; selenium in photocells due to photoconductivity, and tellurium in thermoelectric devices for efficient cooling.[61] Polonium is a rare, intensely radioactive element with no stable isotopes (longest half-life ~138 days for Po-209), exhibiting metallic appearance but chalcogen-like chemistry in +2 and +4 oxidation states; it is used in antistatic devices and as an alpha-particle source.[62] These elements exhibit oxidation states ranging from -2 in oxides and chalcogenides to +6 in sulfate and selenate, reflecting their variable valence.[63] Livermorium, element 116, is synthetic with a short half-life, predicted to be more metallic than lighter homologs due to relativistic effects. A recurring trend in heavier elements of Groups 13–16 is the inert pair effect, where the ns<sup>2</sup> electron pair becomes reluctant to participate in bonding due to poor shielding by d and f electrons, stabilizing lower oxidation states. In thallium, the +1 state (e.g., Tl<sup>+</sup>) is more stable than +3, as seen in the endothermic disproportionation of TlCl. Similarly, in lead, Pb<sup>2+</sup> predominates over Pb<sup>4+</sup>, evident in the stability of PbO over PbO<sub>2</sub> in aqueous solutions.[64] This effect intensifies down the groups, altering reactivity and compound stability.[65]Group 17: Halogens
The halogens, comprising Group 17 of the periodic table, include the elements fluorine (F), chlorine (Cl), bromine (Br), iodine (I), astatine (At), and tennessine (Ts).[66] These elements exist primarily as diatomic molecules (X₂) and exhibit a range of physical states at standard temperature and pressure: fluorine and chlorine are pale yellow and greenish-yellow gases, respectively, bromine is a reddish-brown liquid, and iodine forms violet-gray crystals as a solid.[66] Astatine, a radioactive element, is presumed to be a solid with metallic properties, while tennessine, the heaviest and fully synthetic member, has theoretical properties consistent with group trends but remains poorly characterized due to its short half-life.[66] The vivid colors of the halogens arise from electronic transitions in their molecules, becoming darker down the group as wavelengths shift toward the visible spectrum.[66] Halogens are highly electronegative nonmetals that readily gain one electron to form X⁻ anions, driving their reactivity as oxidizing agents.[66] Reactivity decreases down the group—from fluorine, the most reactive element, to astatine and tennessine—as atomic size increases and the ability to attract electrons diminishes, reflecting broader p-block trends in electron affinity and ionization energy.[66] This trend enables displacement reactions, where a more reactive halogen oxidizes the halide of a less reactive one, such as chlorine liberating bromine from bromide ions.[66] Interhalogen compounds form between dissimilar halogens, often with the larger halogen in a positive oxidation state; for example, iodine heptafluoride (IF₇) features iodine in the +7 state and adopts a pentagonal bipyramidal geometry.[67] Chlorine gas (Cl₂) demonstrates bleaching action through oxidation of colored organic compounds, releasing hypochlorite species that disrupt chromophores in dyes and stains.[66] Several halogen compounds play critical roles in chemistry and biology. Hydrogen fluoride (HF) is a weak acid with a pKₐ of approximately 3.17, owing to the strong H–F bond and limited dissociation in water despite fluorine's high electronegativity.[68] Sodium hypochlorite (NaClO) serves as the active oxidizing agent in household bleach, where it decomposes to release hypochlorous acid (HOCl) for disinfection and whitening.[69] Molecular iodine (I₂) is an essential component of thyroid hormones thyroxine (T₄) and triiodothyronine (T₃), which regulate metabolism, growth, and development; iodine deficiency impairs hormone synthesis, leading to goiter and developmental disorders.[70] Fluorine stands out for its exceptional reactivity, capable of forming compounds with even noble gases like xenon. This was demonstrated by Neil Bartlett in 1962, who synthesized the first noble gas compound, xenon hexafluoroplatinate (XePtF₆), through the reaction of xenon with platinum hexafluoride (PtF₆), challenging the inertness of noble gases. This breakthrough led to the synthesis of xenon difluoride (XeF₂) shortly thereafter.[71] This reactivity stems from fluorine's low dissociation energy and high electron affinity, enabling it to break stable bonds that other halogens cannot.[66]Group 18: Noble Gases
The noble gases, comprising helium (He), neon (Ne), argon (Ar), krypton (Kr), xenon (Xe), radon (Rn), and oganesson (Og), occupy Group 18 of the periodic table and are characterized by their monatomic gaseous state under standard conditions.[72] These elements exhibit closed-shell electron configurations, with helium possessing $1s^2 and the others following the general form ns^2 np^6, where n is the principal quantum number, resulting in a stable octet of valence electrons that imparts exceptional chemical inertness.[72] This full valence shell minimizes their tendency to form bonds, distinguishing them from more reactive p-block elements like the halogens in Group 17.[73] Oganesson, element 118, is a fully synthetic superheavy element with an extremely short half-life (~0.7 ms for Og-294) and predicted to be a noble gas, though relativistic effects may reduce its inertness and alter its physical state.[74] Physically, the noble gases are nonpolar and exhibit weak intermolecular forces, leading to extremely low boiling points that increase gradually down the group due to rising atomic masses. For instance, helium has the lowest boiling point of any element at 4.2 K, making it unique for applications requiring extreme low temperatures.[75] Argon, the most abundant noble gas in Earth's atmosphere at approximately 0.93% by volume, exemplifies their prevalence in trace amounts, while others like neon (0.0018%) and krypton (0.0001%) occur in even smaller quantities.[76] Radon, being radioactive with a half-life of about 3.8 days for its most stable isotope ^{222}\mathrm{Rn}, is the rarest and poses environmental health risks due to its emission from uranium decay in soils and rocks.[77] Despite their general inertness, noble gases heavier than neon can form compounds under specific conditions, challenging early assumptions of complete nonreactivity. In 1962, the synthesis of xenon hexafluoroplatinate (XePtF₆) by Neil Bartlett marked the first stable noble gas compound, followed by xenon difluoride (XeF₂) and krypton difluoride (KrF₂), both achieved through direct reaction with fluorine gas at elevated temperatures and pressures.[78] These fluorides demonstrate limited oxidation states for xenon (+2 in XeF₂) and krypton (+2 in KrF₂), with xenon forming additional species like XeF₄ and XeF₆. Lighter gases like helium and neon remain entirely unreactive, while argon's compounds are transient and rare.[78] The noble gases find diverse applications leveraging their inertness and physical properties. Helium serves as a cryogenic coolant for superconducting magnets in MRI machines and particle accelerators due to its low boiling point.[77] Neon is used in illuminated signs and displays, where its red-orange glow under electric discharge provides vibrant lighting. Argon acts as a shielding gas in welding to prevent oxidation of metals, and it is also employed in incandescent bulbs to extend filament life. Krypton and xenon enhance flash lamps for photography and lasers, with xenon's high light output in arc lamps. Radon, however, is primarily noted for its radioactivity, serving in radiation therapy for cancer treatment but posing significant inhalation hazards in homes, leading to mitigation through ventilation standards.[77]Occurrence and Production
Natural Abundance
The cosmic abundance of main-group elements reflects primordial nucleosynthesis and stellar processes, with hydrogen and helium dominating the baryonic mass of the universe at approximately 74% and 24% by mass, respectively, together comprising nearly 98% of the total. Oxygen, a key p-block element, ranks third in overall abundance by mass at about 1%, synthesized primarily in massive stars and supernovae.[79][80] In Earth's crust, main-group elements form the bulk of silicate and oxide minerals, with oxygen the most prevalent at 46.6% by mass, followed by silicon at 27.7% and aluminum at 8.1%. These abundances underscore the crust's siliceous nature, though some main-group elements like francium occur in extreme rarity, with estimates of only 20–30 grams present globally due to its short half-life and decay chain origins in uranium ores. Iron, a transition metal, contextualizes this at 5.0% by mass.[81][82][81]| Element | Abundance (% by mass) | Group |
|---|---|---|
| Oxygen | 46.6 | 16 (p-block) |
| Silicon | 27.7 | 14 (p-block) |
| Aluminum | 8.1 | 13 (p-block) |
| Iron* | 5.0 | Transition |