Fact-checked by Grok 2 weeks ago

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. 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). 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. 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%). 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). Periodic trends in atomic size, ionization energy, electronegativity, and redox potentials across these groups enable systematic predictions of their reactivity and compound formation.

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 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. 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 , emphasizing the primary, representative nature of the s- and p-block elements in demonstrating . The synonymous term "representative elements" underscores their prevalence and typical chemical behaviors, as they constitute the majority of elements in the and . The main-group elements comprise 44 known members from atomic number 1 () to 88 (), excluding the superheavy synthetic elements (atomic numbers 113–118). They are distributed as follows:
GroupElements (Atomic Numbers)
1 (Alkali metals)H (1), (3), (11), (19), Rb (37), Cs (55), Fr (87)*
2 (Alkaline earth metals)Be (4), (12), Ca (20), Sr (38), Ba (56), (88)*
13B (5), Al (13), Ga (31), In (49), Tl (81)
14C (6), Si (14), Ge (32), Sn (50), Pb (82)
15N (7), P (15), As (33), Sb (51), Bi (83)
16O (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)
*Note: Francium (87) and radium (88) are included despite their radioactivity, as they complete groups 1 and 2 among the naturally occurring elements; superheavy analogs are not listed here. Helium is placed in group 18 (p-block) by convention, though its configuration is s².

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 in group 1. The p-block occupies groups 13 through 18 on the right side, including elements such as the (13), (14), nitrogen group (15), chalcogens (16), (17), and (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. The block classification of main-group derives from the filling of their outermost orbitals. In the s-block, the occupy the ns orbital, where n represents the principal of the . For the p-block, fill the np orbitals, leading to configurations that end in s¹ or s² for s-block and varying np¹ to np⁶ for p-block . This orbital-based categorization distinguishes main-group from others, as it emphasizes the involvement of s and p in their chemical behavior. 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. Periodic table formats vary, with the conventional 18-column contracting the f-block into separate rows for compactness, while the extended 32-column 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.

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. 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. The orbital filling in main-group elements adheres to the , which dictates that electrons occupy orbitals of lowest first, progressing from to subshells within each principal level. This principle ensures that the 1 orbital fills before 2, followed by 2, then 3 and 3, and so on, without involvement of or f orbitals in the valence shells of these elements. Exceptions are minimal for main-group atoms, as the energy ordering of and orbitals remains straightforward across periods. A notable feature in heavier p-block elements is the , 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, 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 . The of main-group elements exhibits systematic trends influenced by and nuclear charge. Across a , atomic radius decreases from left to right due to increasing , which pulls electrons closer to the 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 to 265 pm for cesium, illustrating the downward increase.
ElementAtomic Radius (pm, metallic)
Li152
Na186
K231
Rb244
Cs265
Ionization energy, the energy required to remove an from a gaseous , follows complementary trends: it generally increases across a as decreases and rises, making electron removal harder, and decreases down a group due to larger radii and greater electron shielding. The first ionization energy for sodium is 496 kJ/mol, reflecting its relatively low value in the s-block, while for it is 1681 kJ/mol, highlighting the high value in the upper p-block. These trends stem from the stability of electron configurations, such as half-filled or fully filled p subshells in later periods. Electronegativity, a measure of an atom's ability to attract electrons in a , on the Pauling scale shows low values in the s-block (0.8 for and to 1.0 for ) due to large atomic sizes and low nuclear attraction, and high values in the upper p-block (3.0 for and to 4.0 for ) arising from small radii and high . This scale, derived from differences, underscores the transition from metallic to nonmetallic character across the main groups. Melting and boiling points of main-group elements vary with bonding type and atomic/molecular structure, showing no uniform trend but distinct patterns by group. Group 1 metals have low melting points (e.g., 181°C for decreasing to 28°C for cesium) and boiling points (e.g., 1342°C for to 671°C for cesium), attributable to weak from single valence electrons and increasing atomic size. In contrast, Group 14 elements like carbon exhibit exceptionally high values in solid forms, with subliming around 3642°C and melting near 4000°C under pressure, due to strong covalent network bonding. These patterns reflect the shift from delocalized metallic bonds in s-block to localized covalent bonds in p-block nonmetals.

Chemical Properties

Valence and Bonding

Main-group elements possess electrons in their outermost s and p orbitals, ranging from 1 to 8 electrons, which primarily govern their chemical reactivity and preferences. These electrons determine the elements' tendency to achieve a stable , often by adhering to the , wherein atoms form bonds to attain eight electrons in their valence shell, mimicking the configuration. This rule is particularly applicable to main-group elements, as their valence shells lack the d-orbital involvement seen in metals, leading to predictable patterns based on gain, , or sharing. 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. 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. 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. 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 (CO₂) and -4 in (CH₄), while ranges from -3 in (NH₃) to +5 in (HNO₃). 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 (CH₄), enabling tetrahedral geometry. Sp² hybridization, using one s and two p orbitals, produces three planar orbitals, seen in (BF₃) with trigonal planar arrangement around . Sp hybridization, involving one s and one p orbital, results in two linear orbitals, facilitating triple bonds in compounds like acetylene (C₂H₂). These hybridizations enhance orbital overlap, stabilizing covalent bonds in p-block species.

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. 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. 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. Key reaction types for main-group elements include processes and acid-base reactions, reflecting their electron-donating or -accepting tendencies. reactions are prominent in s-block elements, where metals act as strong reducing agents; a classic example is the reaction of sodium with , producing gas and :
$2\mathrm{Na}(s) + 2\mathrm{H_2O}(l) \rightarrow 2\mathrm{NaOH}(aq) + \mathrm{H_2}(g)
This underscores the metals' ability to reduce to . Acid-base reactions are common with p-block and s-block oxides, which exhibit amphoteric or basic character; for example, from Group 2 reacts with to form a and :
\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.
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. 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. 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. The stability of main-group compounds, particularly toward , follows predictable trends tied to ionic size and . In Group 2, exhibit increasing thermal stability down the group; magnesium carbonate (MgCO₃) decomposes readily upon heating to yield and , whereas (BaCO₃) requires much higher temperatures for decomposition due to the larger, less polarizing Ba²⁺ cation stabilizing the . This pattern arises from decreasing of the cations, reducing their ability to polarize and destabilize the anion.

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. 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. Lithium, in particular, has a density of 0.534 , making it the least dense metal and enabling it to float on , while sodium (0.968 ) and potassium (0.862 ) also have densities below that of . These properties stem from their large atomic radii and weak , with melting points decreasing down the group from 180.5°C for to an estimated 27°C for . 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. 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. 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. Key compounds of the alkali metals include (NaCl), which occurs naturally as the mineral and serves as the primary source of sodium, essential for and chemical manufacturing. (KNO₃), known as saltpeter, is historically significant for production and is now widely used in fertilizers to supply and to plants. 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. An notable anomaly in Group 1 is the between and magnesium, arising from their comparable charge-to-radius ratios and polarizing power despite belonging to adjacent groups. 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 and magnesium carbonate in water.

Group 2: Alkaline Earth Metals

The Group 2 elements of the periodic table, known as the alkaline earth metals, consist of (Be), magnesium (Mg), calcium (Ca), strontium (Sr), barium (Ba), and (Ra). These metals are silvery-white, malleable, and ductile solids at , with 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 to 5.0 g/cm³ for —and higher melting points, such as 650 °C for magnesium and 842 °C for calcium, due to their smaller atomic radii, +2 , and enhanced from the additional . , however, is radioactive, with its most stable isotope radium-226 having a of 1600 years, and exists only in trace amounts in nature. 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 to yield and gas via Mg + 2H₂O → Mg(OH)₂ + H₂. Calcium reacts slowly with cold water to produce and , while strontium and react more readily at , generating the corresponding hydroxides (M(OH)₂) and gas. These elements also burn in air to form basic oxides (MO), which react with to produce the hydroxides, contributing to their classification as "alkaline earths" from early observations of their insoluble, alkaline residues./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 (CaCO₃), the primary component of and a key mineral in sedimentary rocks, and (MgSO₄·7H₂O), commonly known as Epsom for its use in baths and as a . Beryllium deviates from the group's ionic character due to its small and high , leading to covalent bonding in compounds like (BeCl₂), which exhibits tetrahedral coordination and volatility. Beryllium is highly toxic, with inhalation of its dust or fumes causing , a chronic lung disease, and it is classified as a by the International Agency for Research on Cancer./12:Group_2-_The_Alkaline_Earth_Metals/12.03:_The_Group_2_elements) Biologically, calcium is vital for bone and tooth mineralization, forming crystals, and serves as a signaling in and transmission. Magnesium is central to chlorophyll's structure in plants, facilitating by stabilizing the ring, and in humans, it acts as a cofactor in over 300 enzymes involved in and .

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. These elements exhibit diverse bonding behaviors, including covalent and ionic characteristics, and play key roles in due to their semiconducting and catalytic properties. Group 13, known as the , includes (B), aluminum (Al), (Ga), (In), (Tl), and the synthetic (Nh). is a with high and covalent network structure, while aluminum is a lightweight metal widely used in alloys. , , and show increasing density and lower melting points, with exhibiting toxicity. , element 113, is highly radioactive with a 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. (GaAs) is a key III-V material with a direct bandgap of 1.42 , enabling efficient light emission in LEDs and lasers. In Group 14, the , elements carbon (C), (Si), (Ge), tin (Sn), lead (Pb), and flerovium (Fl) demonstrate a progression from to metal. Carbon's unique ability to catenate forms long chains and rings, underpinning , and its allotropes include insulating and conductive . and are metalloids used in semiconductors, with forming the basis of integrated circuits due to its abundance and stable oxide layer. Tin and lead are soft metals; lead's +2 is stabilized in applications like lead-acid batteries, where Pb and PbO<sub>2</sub> electrodes facilitate reversible reactions in electrolyte, providing high surge current capacity. , element 114, is synthetic and volatile, with predicted metallic properties influenced by relativistic stabilization of the 7s electrons. Group 15 elements, the pnictogens, comprise (N), (P), (As), (Sb), (Bi), and (Mc). exists primarily as stable N<sub>2</sub>, requiring energy-intensive fixation processes like the Haber-Bosch method to convert it to for fertilizers. has allotropes including reactive white (P<sub>4</sub> tetrahedra) and polymeric red , with white form igniting spontaneously in air. and are metalloids with toxic compounds; causes acute poisoning by inhibiting enzymes, while compounds affect cardiac function. stands out as the most metallic, with a low thermal conductivity and rhombohedral , used in low-melting alloys. , element 115, is a synthetic that is highly radioactive, with the longest-lived (Mc-289) having a of about 0.8 seconds; its properties are predicted, including a likely solid metal state and stabilization of the +1 due to the . The chalcogens in Group 16 include oxygen (O), (S), (Se), (Te), (Po), and (Lv). Oxygen's O<sub>2</sub> molecule is paramagnetic due to two unpaired electrons in its , enabling its role in and . participates in biogeochemical cycles, oxidizing to in aerobic environments and reducing to anaerobically, influencing global balance. and find applications in electronics; in photocells due to , and in thermoelectric devices for efficient cooling. is a rare, intensely radioactive element with no stable isotopes (longest ~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. These elements exhibit oxidation states ranging from -2 in oxides and chalcogenides to +6 in and selenate, reflecting their variable valence. , element 116, is synthetic with a short , predicted to be more metallic than lighter homologs due to relativistic effects. A recurring trend in heavier elements of Groups 13–16 is the , 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 , the +1 state (e.g., <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. This effect intensifies down the groups, altering reactivity and compound stability.

Group 17: Halogens

The , comprising Group 17 of the periodic table, include the elements (F), (Cl), (Br), iodine (I), (At), and (Ts). These elements exist primarily as diatomic molecules (X₂) and exhibit a range of physical states at : and are pale yellow and greenish-yellow gases, respectively, is a reddish-brown liquid, and iodine forms violet-gray as a solid. , a radioactive element, is presumed to be a solid with metallic properties, while , the heaviest and fully synthetic member, has theoretical properties consistent with group trends but remains poorly characterized due to its short half-life. The vivid colors of the arise from electronic transitions in their molecules, becoming darker down the group as wavelengths shift toward the . Halogens are highly electronegative nonmetals that readily gain one to form X⁻ anions, driving their reactivity as oxidizing agents. Reactivity decreases down the group—from , the most reactive element, to and —as atomic size increases and the ability to attract electrons diminishes, reflecting broader p-block trends in and . This trend enables displacement reactions, where a more reactive oxidizes the of a less reactive one, such as liberating from ions. Interhalogen compounds form between dissimilar , often with the larger in a positive ; for example, iodine heptafluoride (IF₇) features iodine in the +7 state and adopts a pentagonal bipyramidal . gas (Cl₂) demonstrates bleaching action through oxidation of colored organic compounds, releasing species that disrupt chromophores in dyes and stains. Several halogen compounds play critical roles in chemistry and biology. (HF) is a weak acid with a pKₐ of approximately 3.17, owing to the strong H–F bond and limited dissociation in despite fluorine's high . (NaClO) serves as the active oxidizing agent in household bleach, where it decomposes to release (HOCl) for disinfection and whitening. Molecular iodine (I₂) is an essential component of thyroxine (T₄) and (T₃), which regulate , growth, and development; iodine deficiency impairs hormone synthesis, leading to goiter and developmental disorders. Fluorine stands out for its exceptional reactivity, capable of forming compounds with even like . This was demonstrated by Neil Bartlett in 1962, who synthesized the first , (XePtF₆), through the reaction of with hexafluoride (PtF₆), challenging the inertness of . This breakthrough led to the synthesis of (XeF₂) shortly thereafter. This reactivity stems from fluorine's low dissociation energy and high , enabling it to break stable bonds that other cannot.

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. 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. This full valence shell minimizes their tendency to form bonds, distinguishing them from more reactive p-block elements like the halogens in Group 17. 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. Physically, the are nonpolar and exhibit weak intermolecular forces, leading to extremely low that increase gradually down the group due to rising atomic masses. For instance, has the lowest of any at 4.2 K, making it unique for applications requiring extreme low temperatures. , the most abundant in Earth's atmosphere at approximately 0.93% by volume, exemplifies their prevalence in trace amounts, while others like (0.0018%) and (0.0001%) occur in even smaller quantities. , being radioactive with a of about 3.8 days for its most stable ^{222}\mathrm{Rn}, is the rarest and poses risks due to its emission from in soils and rocks. 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. 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. 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. 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.

Occurrence and Production

Natural Abundance

The cosmic abundance of main-group elements reflects primordial and stellar processes, with and dominating the baryonic mass of the at approximately 74% and 24% by mass, respectively, together comprising nearly 98% of the total. Oxygen, a p-block , ranks third in overall abundance by mass at about 1%, synthesized primarily in massive stars and supernovae. In , main-group elements form the bulk of and minerals, with oxygen the most prevalent at 46.6% by mass, followed by at 27.7% and aluminum at 8.1%. These abundances underscore the crust's siliceous nature, though some main-group elements like occur in extreme rarity, with estimates of only 20–30 grams present globally due to its short and origins in ores. Iron, a , contextualizes this at 5.0% by mass.
ElementAbundance (% by mass)Group
Oxygen46.616 (p-block)
Silicon27.714 (p-block)
Aluminum8.113 (p-block)
Iron*5.0Transition
*Included for context; not main-group. The atmosphere highlights gaseous main-group elements, dominated by at 78.08%, oxygen at 20.95%, and at 0.93% by volume in dry air. , a major hydrospheric reservoir, features sodium (30.6% of total salts), (55%), and magnesium (3.7%) as principal ions, accounting for over 90% of dissolved solids at an average of 35 g/kg. Geochemical cycles maintain the distribution of carbon, nitrogen, and sulfur across Earth's spheres. The carbon cycle interconnects atmospheric CO₂ (reservoir ~750 GtC), oceanic dissolved inorganic carbon (~38,000 GtC), and lithospheric carbonates (~60,000,000 GtC) through fluxes like weathering, sedimentation, and volcanism, regulating long-term atmospheric composition. The nitrogen cycle, with the atmosphere holding ~3,900,000 GtN as N₂, involves fixation (converting N₂ to bioavailable forms), nitrification, and denitrification, cycling ~140 TgN annually via microbial and lightning processes in soils and waters. The sulfur cycle links atmospheric SO₂ (from volcanism and oxidation), oceanic sulfate (~1,300,000 GtS), and sedimentary sulfides through reduction in anoxic environments and oxidation in oxic ones, with global fluxes around 300 MtS per year influencing acidity and mineral formation.

Extraction Methods

Main-group elements, particularly the highly reactive metals in the s-block and many p-block elements, are typically isolated through energy-intensive processes such as or thermal reduction, owing to their strong affinity for oxygen and tendency to form stable compounds in nature. These methods prioritize separation from abundant ores like halides or oxides, often requiring molten salts to enable ion mobility and prevent water-related side reactions. For s-block elements, electrolysis of molten chlorides is the dominant industrial approach for and metals. is produced commercially via the , where molten , mixed with to lower the to about 600°C, undergoes ; at the , ions are reduced to molten metal ($2Na^+ + 2e^- \rightarrow 2Na), while ions are oxidized to gas at the ($2Cl^- \rightarrow Cl_2 + 2e^-). The overall cell reaction is $2NaCl \rightarrow 2Na + Cl_2, yielding high-purity that floats and is collected separately from the denser . metal, however, faces greater challenges due to its higher reactivity and , making direct molten-salt impractical; instead, it is obtained through of with vapor at 870–980°C: Na (g) + KCl (l) \rightarrow NaCl (l) + K (g), followed by to separate the more volatile . is extracted by of molten at around 800°C, often derived from via (CaO) production and subsequent chlorination; the cathodic yields metal (Ca^{2+} + 2e^- \rightarrow Ca), with gas as a . In the p-block, extraction varies by group but frequently involves electrolytic or chemical oxidation processes tailored to the element's ore form. Aluminum, the most produced p-block metal, is isolated using the Hall-Héroult process, where purified alumina (Al₂O₃) is dissolved in molten (Na₃AlF₆) at 950–980°C and electrolyzed; aluminum ions reduce to molten metal at the carbon (Al^{3+} + 3e^- \rightarrow Al), while oxygen reacts with the carbon to form CO₂ (C + O_2 \rightarrow CO_2), resulting in the net reaction $2Al_2O_3 + 3C \rightarrow 4Al + 3CO_2. This process consumes significant , about 13–15 kWh per kg of aluminum, highlighting its energy intensity. (HNO₃), essential for nitrogen compounds, is synthesized industrially via the starting from (derived from ); ammonia is oxidized over a platinum-rhodium catalyst at 850–900°C to ($4NH_3 + 5O_2 \rightarrow 4NO + 6H_2O), followed by further oxidation to ($2NO + O_2 \rightarrow 2NO_2) and absorption in water to form ($3NO_2 + H_2O \rightarrow 2HNO_3 + NO). is recovered from underground deposits using the , where superheated water (about 160°C) is injected via concentric pipes to melt the sulfur, which is then forced to the surface by compressed air, achieving recoveries of up to 90% without . Extraction challenges are pronounced for certain reactive or specialized p-block elements. Beryllium, sourced from beryl (Be₃Al₂Si₆O₁₈), requires energy-intensive processing: the ore is crushed, roasted with fluoride salts to form soluble beryllium fluoride, and then extracted via solvent or precipitation methods, often involving high-temperature sintering up to 1650°C to decompose silicates, consuming substantial thermal energy due to beryllium's stable oxide layer. Radon, a noble gas, is isolated from the radioactive decay of radium in uranium ores, but handling poses severe challenges due to its short half-life (3.8 days for ²²²Rn) and alpha-emitting decay products, necessitating shielded containment and ventilation to mitigate inhalation risks during emanation and purification. Environmental considerations have driven process improvements, particularly in halogen production. Chlorine, a key byproduct of s-block electrolyses like the , was historically produced via mercury-cell chlor-alkali processes using aqueous NaCl, where mercury cathodes amalgamated sodium, leading to significant mercury emissions and ; these cells, responsible for up to 40–50% of mercury use in alone, have been globally phased out since the in favor of membrane-cell , reducing mercury releases by over 99% and minimizing ecological impacts like in waterways.

Applications and Biological Role

Industrial and Technological Uses

Main-group elements play a pivotal role in various , leveraging their unique chemical properties for large-scale production and technological innovations. Among the and alkaline earth metals, sodium is essential in the manufacture of s and detergents through the process, where reacts with fats and oils to produce . Sodium compounds, such as , are also critical in , acting as a to lower the of silica and facilitate the formation of silicate glass. Aluminum, from , is widely used in applications due to its lightweight, high-strength alloys, which provide corrosion resistance and structural integrity in fuselages, wings, and engine components. These alloys, often containing small amounts of , magnesium, or , enable significant weight reductions while maintaining durability under extreme conditions. , another element, is fundamental in technology, where it serves as the base material doped with impurities like or to create n-type or p-type semiconductors, enabling the fabrication of transistors and integrated circuits essential for . Nonmetals from groups 14–17 are integral to materials and . Carbon is indispensable in production, where it acts as a in the to convert into and as an alloying element to impart strength and hardness to , supporting , automotive, and sectors. Nitrogen, a group 15 element, is primarily utilized in fertilizers via the Haber-Bosch process, which synthesizes from atmospheric and under high pressure and temperature: \ce{N2 + 3H2 -> 2NH3}, enabling the production of ammonium-based compounds that boost agricultural yields globally. Halogens, particularly from group 17, are key in the production of (PVC), where reacts with ethylene to form , which polymerizes into PVC used in pipes, cables, and packaging. Noble gases from group 18 find niche applications due to their inertness. is employed in high-intensity arc lamps for applications like cinema projectors, automotive headlights, and medical , where its high-pressure discharge produces bright, continuous white light with excellent color rendering. , valued for its low , is used to cool superconducting magnets in (MRI) machines, maintaining temperatures near to enable strong, stable magnetic fields for high-resolution imaging. Emerging technologies highlight the evolving uses of main-group elements. Lithium from group 1 has seen explosive demand in lithium-ion batteries since the post-2010 electric vehicle boom, where it serves as the key component in cathodes and electrolytes, powering with high and enabling the global shift toward sustainable transportation. Similarly, (), a compound of and , is revolutionizing light-emitting diodes (LEDs), particularly in blue and white LEDs for displays, lighting, and , due to its wide bandgap and high efficiency in converting electricity to light.

Role in Biology and Environment

Main-group elements play critical roles in biological systems, forming the foundation of life processes through their incorporation into biomolecules and cellular functions. Carbon (C), hydrogen (H), oxygen (O), nitrogen (N), (P), and sulfur (S)—collectively known as —constitute approximately 98% of living matter and are essential for all organisms, serving as building blocks for carbohydrates, proteins, nucleic acids, and . is vital for energy transfer in ATP and genetic storage in and , while sulfur supports via like and . (Na) and (K) maintain membrane potentials and osmotic balance, with K universally required across all life forms for enzyme activation and nerve impulses in animals, and Na essential for extracellular fluid regulation in animals and certain microbes. Calcium (Ca) functions in signaling pathways, , and such as formation in vertebrates, and magnesium (Mg) acts as a cofactor in over 300 enzymes, including those involved in via . While beneficial at physiological levels, certain main-group elements exhibit toxicity when in excess, impacting health and ecosystems. (F) strengthens and prevents dental caries at low doses, but excessive intake during tooth development causes , characterized by enamel hypomineralization and discoloration ranging from mild white streaks to severe pitting. P-block elements like arsenic (As) and lead (Pb) are potent environmental pollutants; inorganic As from contaminated water and soil disrupts and , leading to cancer and , while Pb interferes with neurological development and synthesis, causing cognitive impairments even at low chronic exposures. Main-group elements are integral to environmental cycles that regulate Earth's atmosphere and . Oxygen's triatomic form, (O₃), forms in the via photochemical reactions with O₂, absorbing harmful and protecting from DNA damage. Sulfur participates in the global , where volcanic emissions of SO₂ oxidize to (H₂SO₄), contributing to that acidifies soils and water bodies, harming aquatic life and forests. Nitrogen from synthetic fertilizers, which surged over 20-fold since 1950, drives in freshwater and coastal systems, promoting algal blooms that deplete oxygen and create hypoxic "dead zones." In biogeochemical contexts, main-group elements link biological and geological processes. Silicon (Si) is indispensable for diatoms, marine phytoplankton that form opal frustules and account for up to 40% of oceanic , thereby controlling the marine through uptake and export to sediments. Iodine (I) cycles through marine environments via microbial transformations, concentrating in and to support hormone synthesis in higher trophic levels, thus facilitating its transfer up the marine food chain to terrestrial consumers.

Historical Development

Early Discoveries

The earliest known uses of main-group elements trace back to prehistoric and ancient civilizations, where carbon and were recognized for their practical properties. Carbon, in the form of produced by heating wood in low-oxygen conditions, has been utilized since at least 30,000 BCE, as evidenced by charred remains found in early hominid caves, serving as fuel for and a for . , appearing as bright yellow native deposits, was collected and used as early as 5000 BCE in regions like the for medicinal, ritualistic, and purposes, often referred to as "" in ancient texts due to its flammable nature. One of the earliest modern discoveries of a main-group element was in 1669 by German alchemist , who isolated the element by distilling residues from evaporated urine, obtaining a waxy, phosphorescent white substance that glowed in the dark; initially thought to be connected to the , it was later recognized as a distinct element and named from the Greek for "light-bearer." metals such as sodium and were extracted from wood and plant ashes through leaching processes well before 1800 CE; ancient Egyptians and Romans produced () and soda ash () by burning vegetation and dissolving the residues in water, employing these for glassmaking, production, and textiles as documented in historical chemical practices. Key experiments in the late advanced the identification of main-group elements, particularly nonmetals. In 1778, French chemist isolated and named oxygen (from the Greek for "acid producer") through careful combustion studies, heating mercuric oxide to release the gas and demonstrating its role in respiration and burning, which overturned the and established oxygen as a fundamental element. Building on this era's progress, Swedish chemist achieved the first isolation of in 1824 by heating potassium metal with potassium fluorosilicate (K₂SiF₆), yielding impure amorphous silicon that he purified further, confirming it as an element distinct from previously known compounds like silica. The 18th and 19th centuries saw the discovery of several reactive through innovative chemical manipulations. was first produced in 1774 by apothecary , who reacted with (pyrolusite) to liberate the greenish-yellow gas, initially mistaking it for a compound but noting its bleaching and oxidizing properties. emerged in 1811 from the work of French chemist Bernard Courtois, who, while extracting sodium and salts from ash for production, added excess and observed violet vapors condensing into dark crystals; this new element was later confirmed by . , the most challenging due to its extreme reactivity, was finally isolated in 1886 by French chemist via of a solution of in anhydrous using a platinum-iridium apparatus, earning him the in 1906 for this breakthrough. Among the radioactive main-group elements, discoveries occurred in the early through studies of nuclear decay. Radon was identified in 1900 by German physicist Friedrich Ernst Dorn as a radioactive gas emanating from compounds, observed during investigations of radium's and initially termed "radium emanation" for its alpha-particle emission. , the heaviest , was discovered in 1939 by French physicist at the Curie Institute while purifying samples; she detected an unexpected beta-emitting impurity that precipitated like cesium, confirming element 87 through spectroscopic analysis of its salts.

Integration into Periodic Table

The integration of main-group elements into the periodic table began with early attempts to identify patterns among them based on chemical similarities. In 1829, observed that certain groups of three elements, known as triads, exhibited comparable properties, with the atomic weight of the middle element approximately equaling the average of the other two; a prominent example was the alkali metals , sodium, and , which shared reactive behaviors with . This insight highlighted regularities in main-group elements but was limited to a few sets. Building on this, John Newlands in 1865 proposed the law of octaves, arranging known elements by increasing atomic weights and noting that every eighth element, akin to musical octaves, displayed similar properties; main-group elements like the alkali metals and fitted this pattern, though the scheme faltered for heavier elements and faced criticism for oversimplification. Dmitri Mendeleev's 1869 periodic table marked a pivotal advancement, organizing all known elements by atomic weight into rows and columns where main-group elements formed distinct vertical families based on valence electrons and chemical analogies. Mendeleev distinguished main groups (subgroups A) from transitional subgroups (B) by emphasizing valence similarities, placing alkali metals, alkaline earths, and pnictogens in primary families while accommodating variable valences in others. His table predicted undiscovered main-group elements, such as eka-aluminum (later gallium, discovered in 1875) and eka-silicon (germanium, 1886), with accurate forecasts of their properties like density and valence, validating the framework for main-group classification. Mendeleev's system initially left a gap for undiscovered inert gases, which was filled by the late 19th-century discoveries of the noble gases; argon was identified in 1894 by Lord Rayleigh and William Ramsay as a 1% component of air unexplained by nitrogen and oxygen, leading to the terrestrial isolation of helium in 1895 and neon, krypton, and xenon in 1898, establishing group 18 (originally group 0) as the unreactive main-group family completing the octet rule and periodicity. In the modern era, the International Union of Pure and Applied Chemistry (IUPAC) formalized the periodic table's structure in the 1980s, adopting the 1–18 group numbering system that clearly delineates main-group elements in groups 1, 2, and 13–18, reflecting their s- and p-block electron configurations. This notation resolved ambiguities in earlier systems and extended to synthetic superheavy main-group elements, such as nihonium (group 13, atomic number 113) and tennessine (group 17, atomic number 117), officially included in 2016 after IUPAC verification of their synthesis and properties. Post-Mendeleev anomalies, like the unexpected stability of lower oxidation states in heavier main-group elements (e.g., Tl(I) over Tl(III)), were later explained by the inert pair effect, proposed by Nevil Sidgwick in 1927; this relativistic phenomenon stabilizes the ns² electron pair, reducing its participation in bonding and aligning observed trends with quantum theory.