Metalloid
A metalloid is a chemical element with properties that are intermediate between those of typical metals and nonmetals, often exhibiting semiconductor behavior rather than full conductivity or insulation.[1] In the periodic table, metalloids occupy a diagonal "staircase" band that separates the metals on the left from the nonmetals on the right, typically comprising six elements: boron (B), silicon (Si), germanium (Ge), arsenic (As), antimony (Sb), and tellurium (Te).[2] While classifications can vary, some sources include polonium (Po) and astatine (At) as additional metalloids due to their borderline characteristics, bringing the total to eight.[3] These elements generally appear metallic in luster but are brittle rather than malleable or ductile, and they form covalent crystal structures rather than metallic lattices.[2] Their electronegativities fall between those of metals and nonmetals, leading to amphoteric behavior in compounds—capable of acting as either acids or bases—and they rarely form simple monatomic ions.[2] Unlike metals, metalloids conduct electricity poorly at room temperature but can be doped to enhance conductivity, making them essential for semiconductor applications.[1] Metalloids play critical roles in modern technology and materials science; for instance, silicon is the foundation of the semiconductor industry, enabling transistors, microchips, and photovoltaic cells, while boron contributes to heat-resistant borosilicate glass used in laboratory equipment and cookware.[2] Arsenic and antimony find applications in alloys and flame retardants, and germanium is valued in fiber optics and infrared detectors.[3] Their unique electronic properties also make them important in emerging fields like thermoelectric materials and nanotechnology.[1]Definitions and Classification
Core Definitions
Metalloids are chemical elements that exhibit properties intermediate between those of metals and nonmetals, such as luster, hardness, and reactivity.[4] Their electrical conductivity is higher than that of typical nonmetals but lower than that of metals, enabling them to function as semiconductors in many applications.[4] This intermediate behavior extends to other traits, including the formation of amphoteric oxides that react with both acids and bases, and a tendency to form brittle solids rather than ductile ones.[5] The classification of elements as metalloids often relies on expert judgment due to the absence of universal agreement on precise boundaries, with different sources recognizing varying numbers of such elements based on contextual properties. Observed values for recognized metalloids include first ionization energies typically ranging from 750 to 1000 kJ/mol, which fall between the generally lower values for metals and the higher values for nonmetals.[6] Similarly, electronegativity on the Pauling scale is commonly between 1.9 and 2.2 for metalloids, distinguishing them from metals (generally below 1.9) and nonmetals (above 2.2).[7] For their semiconducting nature, metalloids possess band gaps of approximately 0.1 to 3 eV, allowing controlled electron excitation under moderate energy inputs, unlike the zero band gap of metals or the large gaps exceeding 5 eV in insulators.[8] These defining characteristics position metalloids along a diagonal line separating metals and nonmetals in the periodic table, highlighting their transitional role in chemical behavior.[4]Classification Criteria
Classification of elements as metalloids relies on a combination of positional, physical, and chemical criteria, though these are not rigidly defined and often overlap with adjacent categories in the periodic table. One primary positional criterion involves elements located in the p-block, particularly those forming a diagonal band separating metals from nonmetals, reflecting their intermediate electronegativity and bonding behaviors that bridge metallic and nonmetallic characteristics.[4] Physical properties provide quantitative thresholds for identification, such as densities typically ranging from 2.3 to 6.7 g/cm³ for the common metalloids, which position them between the low densities of nonmetals (often below 2 g/cm³) and the higher densities of metals (frequently above 7 g/cm³). Melting points generally fall between 450°C and 2100°C, lower than many metals but higher than most nonmetals, contributing to their solid state under standard conditions with variable thermal stability. Metalloids often exhibit luster ranging from dull to metallic and are characteristically brittle in hardness, contrasting with the ductility of metals and the softness or gaseous nature of many nonmetals.[9] Chemically, metalloids are distinguished by their tendency to form amphoteric oxides, which can react as either acids or bases depending on conditions, unlike the predominantly basic oxides of metals or acidic oxides of nonmetals. They display variable oxidation states, often spanning positive and negative values due to their intermediate electronegativities, and preferentially engage in covalent bonding rather than the ionic bonding dominant in metals. These traits underscore their hybrid reactivity, as seen in elements like boron and silicon. No single criterion universally applies, as metalloids defy strict boundaries, leading to ongoing debates in classification; surveys of metalloid lists indicate consistency with multi-criteria approaches, typically identifying 5 to 10 elements depending on the emphasis on physical, chemical, or positional factors.Historical Development
Etymology and Terminology
The term "metalloid" derives from the Greek metallon (μέταλλον), meaning "metal," and eidos (εἶδος), meaning "form" or "kind," connoting elements that resemble metals. It was introduced in chemistry by Swedish chemist Jöns Jacob Berzelius in 1811, who applied it to nonmetallic elements capable of forming oxyanions, referencing behaviors akin to metals.[10] In the early 19th century, the term initially described ore-like substances or nonmetals with certain metallic affinities, often in the context of mineralogy and early chemical analysis. By the 1830s, its usage evolved to emphasize chemical classification, referring specifically to elements displaying hybrid properties between metals and nonmetals, such as variable conductivity and bonding behaviors. Alternative terms include "semimetals," which highlights electronic band structures akin to semiconductors in physics contexts, and "border elements," underscoring their transitional position. "Chalcogens" serves as a misnomer for some borderline cases like selenium and tellurium, which belong to group 16 but occasionally exhibit metalloid traits despite being predominantly nonmetallic. The International Union of Pure and Applied Chemistry (IUPAC) offers no formal definition of metalloid, reflecting ongoing debates in classification; however, "metalloid" remains the preferred term in chemical literature over "semimetal" to distinguish it from solid-state physics applications. This nomenclature aids in contextualizing their placement along the periodic table's metal-nonmetal divide.Evolution of Recognition
In the early 19th century, chemists began distinguishing elements with properties intermediate between metals and nonmetals, laying the groundwork for the concept of metalloids. British chemist Humphry Davy and French chemists Joseph Louis Gay-Lussac and Louis Jacques Thénard independently isolated impure boron in 1808 by chemically reducing boric acid with potassium, recognizing its ambiguous characteristics—such as poor conductivity and brittleness—that set it apart from typical metals like sodium and potassium, which he had also isolated around the same time.[11][12] Similarly, Swedish chemist Jöns Jacob Berzelius introduced the term "metalloid" in 1811 to describe nonmetallic elements capable of forming oxyanions, akin to metals, exemplified by his work on silicon, which he isolated in purer form by 1824.[10] These efforts highlighted elements that defied strict categorization, prompting a shift from binary metal-nonmetal views toward a more nuanced understanding. By the mid-19th century, the periodic table formalized this recognition. Dmitri Mendeleev's 1869 table arranged elements by atomic weight and grouped those with similar properties, prominently featuring the p-block "stair-step" diagonal line separating metals from nonmetals, where metalloids like silicon, arsenic, and antimony resided.[13] This arrangement emphasized recurring trends in reactivity and bonding, positioning metalloids as transitional elements in the p-block that exhibited variable oxidation states and semiconductor-like behaviors, influencing subsequent classifications. The 20th century brought refinements driven by technological advances, particularly the discovery of semiconducting properties in the 1940s. Researchers at Bell Laboratories, including Russell Ohl, identified the p-n junction in silicon in 1940, enabling its use in radar detectors during World War II, while germanium's semiconducting qualities led to the first transistor in 1947, elevating both elements' status as quintessential metalloids.[14] Arsenic and antimony, known since ancient times for their use in bronze alloys—arsenic hardening copper tools as early as 3000 BCE and antimony in Egyptian cosmetics and metallurgy from the third millennium BCE—were retroactively affirmed as metalloids due to their intermediate electrical conductivities and toxicity profiles.[15][16] Contemporary debates continue to shape metalloid classification, influenced by solid-state physics since the 1950s. The rise of the semiconductor industry highlighted band gap structures, leading to polonium's occasional inclusion as a metalloid post its 1898 discovery, based on its metallic luster and semiconducting potential despite intense radioactivity, though many exclude it due to post-transition metal traits. Carbon, conversely, is firmly excluded owing to its nonmetallic properties—insulator behavior in diamond form, covalent bonding dominance, and lack of metallic luster—despite occasional allotropes like graphite showing partial conductivity.[17] These discussions, rooted in electronic band theory from solid-state physics, underscore metalloids' role in bridging metallic and nonmetallic domains without rigid boundaries.[10]Position in the Periodic Table
Location and Boundaries
Metalloids are primarily located in the p-block of the periodic table, occupying a diagonal region along the "staircase" or zigzag line that extends from boron in group 13, period 2, through silicon (group 14, period 3), germanium (group 14, period 4), arsenic (group 15, period 4), antimony (group 15, period 5), to tellurium (group 16, period 5).[18][19] This positioning spans groups 13 through 16 and periods 2 through 6, reflecting their intermediate nature between metals and nonmetals.[19] Visually, the staircase line on standard periodic tables serves as a boundary separating metals, which dominate the left side, from nonmetals on the right, with metalloids clustered along this irregular diagonal divide.[19] This line typically begins between beryllium and boron in period 2 and zigzags downward to between bismuth and polonium in period 6, though exceptions exist, such as aluminum in group 13, period 3, which lies adjacent to the boundary but is classified as a post-transition metal rather than a metalloid.[20][19] The theoretical boundaries of metalloids in the periodic table are defined by trends in metallic character, which decreases from left to right across a period due to increasing effective nuclear charge and decreasing atomic radius, making electron loss more difficult, while metallic character increases down a group as atomic size grows and valence electrons are farther from the nucleus.[21] These gradients place metalloids in the transitional zone where neither metallic nor nonmetallic dominance is clear, often aligning with elements exhibiting semiconducting behavior or mixed bonding properties.[21] Variations in metalloid classification can arise from elemental allotropes, as seen with arsenic, where the stable gray allotrope displays metalloid characteristics such as a layered structure and semimetallic conductivity, while the yellow allotrope behaves more like a nonmetal with molecular tetrahedral units.[22][23] Such polymorphic forms influence whether an element is included or excluded from the metalloid category in certain contexts, though the gray form is conventionally recognized as the metalloid phase.[22]Alternative Classifications
One alternative classification of metalloids emphasizes their electronic structure, particularly the presence of a small band gap in their solid-state energy bands, typically ranging from approximately 0.1 to 2 eV. This distinguishes them from metals, which have no band gap (0 eV) and exhibit free electron conduction, and from insulators, which have larger band gaps exceeding 3 eV, limiting electron mobility. Elements like silicon and germanium exemplify this category, with band gaps of about 1.1 eV and 0.7 eV at room temperature, respectively, enabling semiconducting behavior that aligns with metalloid characteristics.[24][25][26] Another approach groups metalloids based on reactivity, focusing on their amphoteric behavior, where their oxides react with both acids and bases. For instance, silicon dioxide (SiO₂) dissolves in hydrofluoric acid (HF) to form silicon tetrafluoride and water (SiO₂ + 4HF → SiF₄ + 2H₂O) and in hot, concentrated sodium hydroxide (NaOH) to form sodium silicate (SiO₂ + 2NaOH → Na₂SiO₃ + H₂O), demonstrating dual acidic and basic properties typical of metalloid oxides. This reactivity bridges metallic (basic oxides) and nonmetallic (acidic oxides) tendencies, providing a chemical criterion for classification.[27] There is notable overlap between metalloids and post-transition metals, with elements like gallium sometimes included in broader metalloid groupings due to their intermediate properties, such as poor metallic luster and variable conductivity, though standard classifications exclude them as true post-transition metals. This ambiguity arises from the lack of sharp boundaries in the p-block, where gallium's position near the metal-nonmetal divide leads to occasional metalloid attribution in specific contexts.[28] Since the 1970s, modern alternatives have employed quantum mechanical models, such as pseudopotential theory, to analyze borderline cases by approximating valence electron interactions and unifying covalent (nonmetallic) and metallic bonding descriptions. This approach, advanced through empirical chemical pseudopotentials, helps resolve ambiguities in elements exhibiting hybrid bonding, like those along the periodic table's dividing line, by calculating electronic structures without full core electron treatment.[29]Physical and Chemical Properties
Key Physical Properties
Metalloids are characterized by electrical properties that position them as semiconductors, with conductivity intermediate between metals and nonmetals. Their inherent low conductivity arises from an energy band gap that prevents free electron movement at room temperature, but this can be modulated through doping—introducing impurities to form n-type semiconductors (excess electrons from group 15 elements like phosphorus) or p-type semiconductors (electron deficiencies or "holes" from group 13 elements like boron).[30] Representative band gaps include 1.11 eV for silicon and 0.66 eV for germanium at 300 K, values that enable precise control over charge carrier generation and mobility in response to temperature or electric fields.[25] Thermally and mechanically, metalloids exhibit moderate thermal conductivity, typically higher than nonmetals but substantially lower than metals, facilitating heat dissipation without the efficiency of pure metallic conductors. They are brittle solids lacking the ductility and malleability of metals, often fracturing under stress rather than deforming plastically. Mohs hardness values vary across the group, ranging from 2.25 for tellurium to 9.3 for boron, with many falling between 3.5 and 7 for elements like arsenic and silicon. Luster also differs, from the dull appearance of boron to the more metallic sheen observed in tellurium.[23][31] At room temperature, all commonly recognized metalloids exist as solids with densities between 2.3 g/cm³ for silicon and 6.7 g/cm³ for antimony, values that reflect their intermediate atomic packing compared to the denser, more malleable metals. This density range underscores their structural rigidity without the high mass efficiency of lighter nonmetals or the compactness of heavy metals.[32] Optically, metalloids demonstrate properties suited to light interaction, including transparency in thin films—such as silica's high transmittance in visible and ultraviolet wavelengths—and photoconductivity, where illumination generates charge carriers to boost electrical conductivity. These traits stem from their band structure, allowing photon absorption to bridge the energy gap and influence electronic behavior.[23]Key Chemical Properties
Metalloids exhibit predominantly covalent bonding, forming extended network structures similar to nonmetals, though with varying degrees of metallic character depending on the element. For instance, silicon adopts a diamond-like cubic lattice where each atom is tetrahedrally coordinated to four others via strong covalent Si-Si bonds, resulting in a semiconductor with directional bonding that contrasts with the delocalized electrons in true metals.[2] In terms of reactivity, metalloid oxides often display amphoteric properties, capable of reacting with both acids and bases due to their intermediate electronegativity. Arsenic trioxide (As₂O₃), for example, dissolves in strong bases to form arsenites and in acids to form arsenates, exemplifying this dual behavior.[33] Elements such as arsenic, antimony, and tellurium commonly exhibit +3 and +5 oxidation states in their compounds, reflecting their p-block position and ability to form stable oxyanions or covalent halides in these valences.[34] Metalloids form volatile and often highly toxic hydrides that differ markedly from the interstitial or saline hydrides of metals, as these are covalent molecular gases prone to flammability and extreme reactivity. Arsine (AsH₃), the hydride of arsenic, is a colorless, flammable gas with an odor of garlic, notorious for its acute toxicity through hemolysis even at low concentrations, unlike the more stable metallic hydrides.[34][35] In coordination chemistry, metalloids frequently adopt tetrahedral geometries in their compounds, driven by sp³ hybridization, and many act as Lewis acids due to electron deficiency. Boranes, such as diborane (B₂H₆), feature three-center two-electron bonds and readily coordinate with Lewis bases like ammonia to form stable adducts, achieving an octet configuration around boron and highlighting their acidic character.[36][37]Recognized Metalloid Elements
Boron
Boron (atomic number 5) is a prototypical metalloid element in group 13 of the periodic table, exhibiting intermediate properties between metals and nonmetals.[38] It occurs naturally with two stable isotopes, ^{10}B (approximately 20% abundance) and ^{11}B (approximately 80% abundance), where ^{10}B possesses a high thermal neutron absorption cross-section of 3837 barns, making boron compounds essential in nuclear applications such as control rods for reactors.[39] As a semiconductor, boron displays unique structural and bonding characteristics that distinguish it from neighboring elements like aluminum and carbon. Elemental boron exists in multiple allotropes, primarily amorphous and crystalline forms. Amorphous boron appears as a brown powder with less ordered atomic arrangement, while crystalline forms, such as the β-rhombohedral phase, feature complex icosahedral B_{12} units linked in a three-dimensional network, contributing to its stability and rigidity.[40] These icosahedral structures result in variable bonding environments, with boron atoms achieving electron deficiency through multicenter bonds rather than traditional two-center bonds. Key physical properties of boron underscore its metalloid nature. Crystalline boron exhibits extreme hardness, rated at 9.3 on the Mohs scale, surpassed only by diamond and a few other materials, due to its covalent icosahedral framework.[41] It has a low density of 2.34 g/cm³, which is lighter than many metals yet denser than typical nonmetals like carbon in graphite form.[42] As an intrinsic semiconductor, boron has a wide band gap of 1.50–1.56 eV, enabling applications in high-temperature electronics where silicon would fail.[42] Boron's chemical behavior is dominated by electron-deficient compounds, exemplified by boranes, which feature three-center two-electron bonds in B-H-B bridges. These hydrides, such as diborane (B_2H_6), deviate from the octet rule, leading to cluster-like structures with delocalized electrons that facilitate unique reactivity in synthesis and catalysis.[43] A prominent industrial compound is borax (Na_2B_4O_7·10H_2O), a hydrated sodium tetraborate mineral used extensively in glass manufacturing to lower melting points and improve durability, as well as in detergents for water softening and in metallurgy as a flux.[44] These applications highlight boron's role in bridging inorganic chemistry with practical materials science.Silicon
Silicon (Si), with atomic number 14 and belonging to group 14 of the periodic table, is classified as a metalloid and stands as the most abundant such element in Earth's crust, comprising approximately 27.7% by mass.[45] This abundance underscores its foundational role in geological formations, primarily occurring in silicate minerals. In its elemental form, silicon exhibits a diamond cubic crystal structure, where each atom is tetrahedrally coordinated to four others via covalent bonds, contributing to its brittle, grayish crystalline appearance.[46] As a semiconductor, silicon possesses an indirect band gap of 1.12 eV at room temperature, enabling controlled electrical conductivity that is moderate in its intrinsic state but tunable for technological applications.[47] Chemically, silicon demonstrates intermediate properties between metals and nonmetals, forming stable compounds such as silicon dioxide (SiO₂), which constitutes quartz and serves as a primary glass former due to its tetrahedral network structure.[48] It also produces silanes, including silane (SiH₄), a colorless, flammable gas analogous to methane but with Si-H bonds that are more reactive toward hydrolysis.[49] These hydride compounds highlight silicon's ability to form catenated structures, though less stably than carbon due to weaker Si-Si bonds. For electronic applications, silicon requires high purity, achieved through zone refining, a process that melts and recrystallizes the material in a controlled manner to segregate impurities, yielding purities exceeding 99.9999% (6N).[50] To enhance conductivity, silicon is doped with group 15 elements like phosphorus (P) for n-type semiconductors, introducing excess electrons, or group 13 elements like boron (B) for p-type, creating electron deficiencies or "holes."[51] Additionally, silicon exists in an amorphous allotrope, lacking long-range order, which is deposited as thin films for solar cells due to its higher absorption coefficient compared to crystalline forms, despite lower efficiency.[52]Germanium
Germanium is a metalloid element with atomic number 32 and chemical symbol Ge, belonging to group 14 of the periodic table directly below silicon. It exhibits semiconductor behavior with an indirect band gap of 0.67 eV at 300 K, which is narrower than silicon's 1.12 eV, enabling better performance in certain high-speed applications while maintaining parallels to silicon in lattice compatibility for alloy semiconductors. Germanium has a density of 5.323 g/cm³, more than twice that of silicon at 2.329 g/cm³, contributing to its use in compact optical components.[53][25][54] The element's existence was predicted in 1871 by Dmitri Mendeleev as eka-silicon, an undiscovered analog to silicon based on periodic trends in atomic weight and properties. It was discovered in 1886 by Clemens Winkler, a German chemist analyzing the mineral argyrodite (Ag₈GeS₆), from which he isolated the new element through chemical separation and spectroscopic confirmation. Winkler named it germanium in honor of his native country, Germany.[53][14] Elemental germanium crystallizes in the diamond cubic structure at standard conditions, similar to silicon, but under high pressure or in nanocrystalline forms, it can adopt a tetragonal ST-12 phase with distorted tetrahedral coordination. Its primary oxide, germanium dioxide (GeO₂), possesses a tetragonal rutile structure and displays amphoteric character, reacting with acids to form germanium(IV) salts and with bases to yield germanate ions such as [GeO₃]²⁻. Germanium also forms organogermanium compounds, or organogermanes, featuring stable C–Ge bonds that enable applications in catalysis and bioactive materials, analogous to organosilicon chemistry but with distinct reactivity due to germanium's larger atomic size.[55][56][57][58] Germanium has five stable isotopes, of which ⁷⁴Ge is the most abundant with a natural occurrence of 36.5%. These isotopes, particularly in enriched form, are utilized in high-purity germanium detectors for experiments probing neutrinoless double beta decay of ⁷⁶Ge, a rare process that would indicate neutrinos are their own antiparticles and violate lepton number conservation if observed. In early electronics, germanium was the material of choice for the first transistors developed in 1947, offering higher electron mobility than silicon for point-contact devices in radar and amplification circuits. Its optical transparency from 2 to 14 µm makes it essential for infrared applications, including lenses and windows in thermal imaging and night-vision systems.[53][59][14][60]Arsenic
Arsenic (As) is a chemical element with atomic number 33 and is positioned in group 15 of the periodic table, known as the pnictogens, where it exhibits metalloid characteristics intermediate between nonmetals like phosphorus and metals like antimony.[22][61] This placement underscores its dual nature, displaying poor electrical conductivity in its elemental form yet forming versatile compounds used in semiconductors and other applications. Arsenic's electron configuration allows it to form three covalent bonds, contributing to its reactivity and ability to adopt multiple oxidation states, primarily +3 and +5, which influence its chemical behavior.[22] Arsenic exists in several allotropes, each with distinct structures and properties that highlight its metalloid versatility. The most stable and common form is gray arsenic, a brittle, metallic-appearing crystalline solid that behaves as a semimetal due to partial overlap of its valence and conduction bands near the T and L points in the Brillouin zone, resulting in intrinsic electrical conductivity.[62] Yellow arsenic, a molecular allotrope consisting of As₄ tetrahedra, is unstable and waxy, resembling a nonmetal in its low density and reactivity. Black arsenic, an amorphous form produced by rapid quenching, is less ordered and also semiconductor-like, though less studied for practical uses. These allotropes demonstrate arsenic's adaptability, with gray arsenic's semimetallic properties contrasting sharply with the insulating nature of its group 15 neighbor phosphorus.[63] Arsenic's compounds exemplify its chemical versatility but also its notorious toxicity, posing significant health risks. Arsenic trioxide (As₂O₃), a white powder, is highly toxic, interfering with cellular respiration and enzyme function, and has historically been employed in pesticides such as lead arsenate for crop protection, though its use has declined due to environmental concerns.[64] Arsine (AsH₃), a colorless, flammable gas with a garlic-like odor, is even more acutely toxic, causing severe hemolysis upon inhalation by binding to hemoglobin.[65] These properties have necessitated careful handling in industrial contexts. Arsenic has a long historical record, with the element first isolated around 1250 AD by Albertus Magnus through heating soap with arsenic trisulfide, marking an early milestone in alchemy and chemistry.[22] In 1836, chemist James Marsh developed the Marsh test, a sensitive method involving hydrogen gas to detect arsenic traces by producing a characteristic black deposit, revolutionizing forensic toxicology for poisoning cases.[66] More recently, arsenic trioxide has found a therapeutic role; in 2000, the U.S. FDA approved it under the trade name Trisenox for treating relapsed or refractory acute promyelocytic leukemia, where it induces cancer cell differentiation and apoptosis at controlled doses.[67] This approval highlights arsenic's paradoxical utility in medicine despite its inherent toxicity.Antimony
Antimony (Sb) is a metalloid element with atomic number 51 and belonging to group 15 of the periodic table.[68][69] It exhibits a rhombohedral crystal structure in its stable metallic form.[70] As a semimetal, antimony features a small band overlap of approximately 0.2 eV, contributing to its semiconducting-like behavior in certain applications.[71] Chemically, it shares toxico-chemical similarities with arsenic, its lighter group 15 analog, including analogous compound formation and biological effects.[72] Key properties of antimony include the amphoteric nature of its trioxide (Sb₂O₃), which reacts with both acids and bases to form salts.[73] Stibine (SbH₃), a hydride of antimony, is highly toxic, causing severe hemolytic effects and organ damage upon exposure due to its reactivity and similarity to arsine.[74] Antimony also demonstrates high thermal stability, with its oxides maintaining structural integrity up to elevated temperatures around 600–700°C before phase transitions occur.[75] Antimony exists in multiple allotropes, including the stable metallic gray form, which is lustrous and brittle, and a metastable explosive yellow allotrope formed under specific low-temperature oxidation conditions.[76] The metallic gray allotrope is particularly valued in alloys, such as type metal (typically 15–20% antimony with lead and tin), where it enhances hardness, reduces shrinkage during casting, and improves durability for printing applications.[77] Antimony compounds have been utilized since ancient times, with evidence of their use as cosmetics like kohl (primarily antimony sulfide) dating back to predynastic Egypt around 3100 BC for eye makeup and medicinal purposes.[78] The elemental form was likely first isolated around 300 BC, as referenced in early metallurgical texts, though pure isolation techniques were refined much later.[76] Industrially, antimony's role in alloys and compounds underscores its importance, particularly in lead-acid batteries and flame-retardant formulations derived from Sb₂O₃, leveraging its stability and reactivity.[73]Tellurium
Tellurium (Te) is a metalloid element with atomic number 52 and belonging to group 16 of the periodic table, positioned among the chalcogens but exhibiting metallic characteristics that distinguish it from lighter nonmetals in the group.[79] It was discovered in 1782 by Austrian mineralogist Franz Joseph Müller von Reichenstein while analyzing gold ore from a mine in Transylvania (modern-day Romania), initially mistaking it for an impure form of antimony; the element was later isolated and named in 1798 by Martin Heinrich Klaproth after the Latin word tellus, meaning "earth," reflecting its terrestrial association.[80] Tellurium is extremely rare in Earth's crust, occurring at concentrations of approximately 1 to 5 parts per billion, making it one of the scarcest elements and primarily sourced as a byproduct of copper and lead refining.[81] In its elemental form, tellurium adopts a hexagonal (trigonal) crystal structure with spiral chains of atoms, imparting a brittle, silvery-white appearance and metallic luster despite its semiconducting nature.[82] It functions as a narrow-bandgap semiconductor with an indirect bandgap of approximately 0.33 eV at room temperature, enabling specialized electrical properties that bridge metallic and nonmetallic behaviors.[83] Chemically, tellurium forms oxides like tellurium dioxide (TeO₂), a conditional glass former that requires rapid cooling or modifiers to produce stable amorphous materials with high refractive indices, useful in optical applications.[84] Tellurides, such as cadmium telluride (CdTe), exemplify its role in compound semiconductors, where tellurium contributes to photovoltaic and infrared detection technologies due to tunable electronic properties.[85] However, tellurium and its compounds exhibit high toxicity, with elemental tellurium causing systemic effects upon ingestion or inhalation, including garlic-like breath odor from metabolites and potential neurological damage. Tellurium has eight naturally occurring isotopes, six of which (¹²⁰Te, ¹²²Te, ¹²³Te, ¹²⁴Te, ¹²⁵Te, ¹²⁶Te) are stable, while ¹²⁸Te and ¹³⁰Te possess extremely long half-lives (7.7 × 10²⁴ years and 8.2 × 10²¹ years, respectively) and are effectively stable for most purposes.[86] The isotope ¹²⁸Te, in particular, supports studies of weak interactions and double beta decay in geological samples, aiding in the dating of ancient materials through analysis of decay products. This chalcogen-like chemistry combined with metallic conductivity underscores tellurium's borderline status as a metalloid, enabling niche roles in materials science where scarcity limits broader adoption.[82]Borderline or Less Recognized Elements
Carbon
Carbon, with atomic number 6, belongs to group 14 of the periodic table and is characterized by its ability to form multiple allotropes that exhibit a wide range of physical properties.[87] These allotropes include diamond, an electrical insulator with a wide indirect band gap of 5.47 eV due to its tetrahedral sp³-hybridized covalent bonding network; graphite, a semimetal with layered sp²-hybridized structure and anisotropic electrical conductivity on the order of 10⁴ to 10⁵ S/m along the basal plane; fullerenes such as C₆₀ buckyballs, which are molecular semiconductors with a band gap around 1.7–2.3 eV; and graphene, a single-layer form of graphite that acts as a zero band gap semimetal, where the conduction and valence bands touch at the Dirac points, enabling high electron mobility exceeding 200,000 cm²/V·s.[88][89][90][91] All these forms are unified by strong covalent carbon-carbon bonds, with bond lengths varying from 1.54 Å in diamond to 1.42 Å in graphene. This variability in bonding hybridization (sp³, sp²) underlies carbon's diverse electrical behaviors, from insulating to conducting. Despite its predominantly nonmetallic character, carbon's status as a metalloid is debated due to the semimetallic properties of certain allotropes like graphite and graphene, which bridge metallic and nonmetallic conduction without a finite band gap in the latter.[92] In Vernon's 2013 analysis, carbon is described as a "near metalloid" because graphite shows semimetallic conduction along its basal plane but fails to satisfy key metalloid criteria, such as an electronegativity between 1.9 and 2.2 or an ionization potential of 750–1000 kJ/mol, appearing in only 9% of surveyed metalloid classifications. Bulk carbon does not display typical metalloid traits like intermediate hardness or amphoteric oxides, reinforcing its primary nonmetal designation. However, at the nanoscale, structures like graphene exhibit metallic-like conductivity, with variable carrier density tunable by gating, distinguishing it from core group 14 metalloids such as silicon.[91] The metalloid-like aspects of carbon are thus allotrope-specific and nanoscale-dependent, with no uniform bulk behavior qualifying it as a recognized metalloid, though its electronic versatility continues to inspire applications in advanced materials.[93]Aluminum
Aluminum (Al), atomic number 13, occupies group 13 of the periodic table and is classified as a post-transition metal. Despite this standard categorization, its position near the metal-nonmetal boundary contributes to a borderline status, with some properties echoing those of metalloids. In historical classifications, such as those from early 20th-century texts, aluminum was occasionally grouped with metalloids due to its ambiguous traits, though modern consensus firmly places it among metals.[94]/06%3A_The_Periodic_Table/6.07%3A_Metalloids) A key metalloid-like feature is the amphoteric behavior of its oxide, Al₂O₃, which reacts with both strong acids like HCl and bases like NaOH, dissolving to form salts in each case. This dual reactivity contrasts with typical metal oxides, which are basic, and highlights aluminum's intermediate chemical nature. Additionally, while bulk aluminum demonstrates strong metallic conductivity at approximately 3.5 × 10⁷ S/m—comparable to many metals—its naturally forming oxide layer acts as a nonmetallic electrical insulator, providing corrosion resistance but altering surface electrical properties. With a low density of 2.7 g/cm³, aluminum's lightweight yet robust profile further underscores its utility, though these traits position it on the metallic side of the divide.[95][96][97][98][99] Aluminum's role in semiconducting materials adds to the debate, as it serves as a p-type dopant in alloys, introducing acceptor levels that modify electrical behavior in substrates like silicon carbide. This application leverages its group 13 kinship with the core metalloid boron, enabling controlled conductivity tuning akin to metalloid functions in electronics. Overall, while not a recognized metalloid, aluminum's blend of metallic luster, conductivity, and occasional nonmetal-like reactivity keeps it relevant in discussions of elemental boundaries.[100][101]Selenium
Selenium (Se) is a chemical element with atomic number 34, positioned in group 16 of the periodic table alongside oxygen, sulfur, and tellurium as one of the chalcogens.[102] It was discovered in 1817 by Swedish chemist Jöns Jacob Berzelius during the analysis of a sulfuric acid production residue from a copper refinery, where he isolated the element and named it after Selene, the Greek goddess of the moon, owing to its close chemical resemblance to tellurium.[103] Selenium manifests in multiple allotropes, with the thermodynamically stable gray form displaying metallic luster and photoconductive behavior, whereas the red allotrope acts as an electrical insulator.[104] The gray selenium, characterized by a helical chain structure in a hexagonal lattice, possesses an indirect band gap of approximately 1.8 eV, which facilitates its application as a photoconductor in xerographic processes for imaging reproduction.[105] Additionally, selenium dioxide (SeO₂) behaves as an acidic oxide, reacting with water to produce selenous acid (H₂SeO₃).[106] The metalloid classification of selenium remains contentious, as it is predominantly regarded as a nonmetal due to its high electronegativity and nonmetallic chemical tendencies, yet its gray allotrope exhibits metalloid traits.[107] According to Vernon's criteria for metalloids—which emphasize a standard-state semiconductor band structure, electronegativity in the intermediate range of about 1.9–2.2, and amphoteric oxide formation—selenium falls short overall, with its Pauling electronegativity of 2.55 aligning more closely with nonmetals and its oxide displaying purely acidic properties.[107][102] However, the semiconductor nature of gray selenium underscores its borderline position, akin to other debated elements in the p-block. Selenium also shares toxicological profiles with tellurium, inducing similar adverse effects in biological systems through bioaccumulation and disruption of enzymatic processes.[108]Polonium
Polonium (Po) is a chemical element with atomic number 84, situated in group 16 of the periodic table, making it the heaviest chalcogen.[109] It was discovered in 1898 by Marie and Pierre Curie during their investigations of pitchblende, and named after Marie's native Poland to honor her heritage.[110] As the core homolog of tellurium, polonium shares similar valence electron configurations but exhibits distinct behavior due to its radioactivity.[109] The element is highly radioactive, with all its isotopes unstable; the most prevalent, polonium-210, undergoes alpha decay with a half-life of 138 days.[111] In its alpha allotrope, polonium adopts a metallic structure with a density of approximately 9.1 g/cm³ and a narrow band gap of about 0.2 eV, which imparts semiconductor-like electronic properties.[112][113] Chemically, it forms oxides such as polonium dioxide (PoO₂), consistent with its group position, and serves as an alpha emitter in applications like static eliminators.[114] Polonium's classification is debated: it is often regarded as a post-transition metal due to its metallic luster and conductivity, yet its borderline electrical properties lead some models to categorize it as a metalloid. Its extreme scarcity underscores its rarity, with global production estimated at less than 100 grams per year, primarily as polonium-210 from neutron irradiation of bismuth.[114] This limited availability restricts detailed studies of its properties.[114]Astatine
Astatine (At) is a chemical element with atomic number 85 and is positioned in group 17 of the periodic table, directly below iodine.[115] It was first synthesized in 1940 by Dale R. Corson, Kenneth R. MacKenzie, and Emilio Segrè at the University of California, Berkeley, through the bombardment of bismuth with alpha particles.[116] Unlike the stable halogens above it, astatine is highly radioactive, with all isotopes decaying rapidly; the most stable, astatine-210, has a half-life of approximately 8.1 hours.[115] This short-lived nature, combined with its synthetic production, has limited direct experimental study, leading to properties that are largely extrapolated from theoretical models and trace quantities. The classification of astatine as a metalloid remains highly debated due to its position at the boundary between nonmetals and metals in the periodic table. While it shares halogen-like reactivity, such as forming hydrogen astatide (HAt) analogous to other group 17 hydrides, computational studies suggest it may exhibit more metallic characteristics than iodine, which is unequivocally a nonmetal.[117] Density functional theory (DFT) calculations predict that solid astatine at atmospheric pressure would be monatomic and metallic, with no band gap between valence and conduction bands, contrasting with iodine's molecular, insulating structure.[118] Earlier estimates proposed a semiconducting phase with a band gap of about 0.7 eV for diatomic astatine molecules, but relativistic effects favor a metallic state under standard conditions.[118] Chemically, astatine displays some amphoteric tendencies; for instance, the astatate ion (AtO₃⁻) can coprecipitate with both acidic and basic insoluble salts, indicating behavior intermediate between halogens and metalloids.[119] However, without bulk samples—due to its instability and rarity— these properties remain predictive rather than empirically confirmed. Astatine's extreme scarcity underscores its elusive status; estimates indicate that less than a few micrograms have ever been produced artificially through nuclear reactions, far below 1 gram in total.[120] This paucity arises from the need for particle accelerators to generate it via bismuth irradiation, yielding only trace amounts per run. Despite these challenges, astatine, particularly the isotope astatine-211 (half-life 7.2 hours), holds promise in targeted alpha therapy for cancer treatment.[115] Its alpha-emitting decay delivers high-energy particles over short ranges, enabling precise tumor destruction while minimizing damage to surrounding healthy tissue, as demonstrated in preclinical and early clinical studies labeling biomolecules for selective uptake in malignant cells. As of 2025, astatine-211 has advanced to first-in-human clinical trials, including a study showing good tolerability and preliminary efficacy in patients with radioiodine-refractory differentiated thyroid cancer, with the first labeled compounds shipped in the U.S. for blood cancer trials and research exploring combinations with immunotherapy to enhance effectiveness.[121][122][123] Ongoing research continues to focus on improving production and conjugation methods to further advance these clinical applications.[124]Practical Applications
Semiconductors and Electronics
Metalloids, particularly silicon, dominate the semiconductor industry due to their tunable electrical properties, enabling the fabrication of integrated circuits and transistors that form the backbone of modern electronics. The invention of the transistor in 1947 by John Bardeen, Walter Brattain, and William Shockley at Bell Laboratories revolutionized electronics by replacing bulky vacuum tubes with compact, efficient semiconductor devices, initially using germanium but quickly transitioning to silicon for its superior thermal stability and abundance.[125] Silicon's role expanded in the 1960s with the development of integrated circuits, allowing billions of transistors to be etched onto a single chip, powering computers, smartphones, and countless devices.[126] In photovoltaics, silicon-based solar cells achieve commercial efficiencies of up to 25% (as of November 2025), converting sunlight to electricity at scale for renewable energy applications, with ongoing advancements pushing laboratory records to 27.81% for single-junction cells (as of April 2025) and up to 33% for perovskite-silicon tandems (as of June 2025).[127][128][129] Germanium, another key metalloid, played a pivotal role in early transistor development, with the first point-contact transistor demonstrated in 1947 at Bell Labs using a germanium crystal, marking the birth of solid-state electronics.[130] Although largely supplanted by silicon in general computing, germanium remains essential in niche high-performance applications, such as photodetectors for fiber-optic communications, where its sensitivity to infrared wavelengths enables efficient signal detection in telecommunications networks.[131] These detectors leverage germanium's direct bandgap and high carrier mobility to convert optical signals into electrical ones with minimal loss, supporting high-speed data transmission over long distances.[132] III-V compound semiconductors, incorporating metalloids like arsenic in gallium arsenide (GaAs), offer superior electron mobility—up to six times faster than in silicon—enabling high-frequency operation in devices such as light-emitting diodes (LEDs) and microwave amplifiers.[133] GaAs LEDs emit light efficiently due to their direct bandgap, making them ideal for displays, optical communication, and sensing applications where silicon's indirect bandgap results in poor luminescence.[134] Doping in these materials involves introducing group II elements (e.g., zinc) for p-type conductivity or group VI elements (e.g., sulfur) for n-type, creating p-n junctions that control carrier flow and enable device functionality, though challenges like dopant diffusion limit high-density integration.[135] Looking ahead, silicon-germanium (SiGe) alloys enhance chip performance by straining the silicon lattice to boost carrier speeds, finding use in high-speed bipolar transistors for RF applications in 5G and automotive radar systems.[136] These alloys enable transistors operating at frequencies exceeding 400 GHz at room temperature, supporting next-generation computing and communication technologies.[137] The global semiconductor market, driven by metalloid-based innovations, is projected to reach $701 billion in sales by the end of 2025, underscoring their economic impact.[138]Alloys and Materials
Metalloids play a crucial role in alloy development, where their addition to metals enhances mechanical properties such as strength, hardness, and resistance to corrosion or wear, without compromising other desirable traits like conductivity or ductility. Silicon, antimony, boron, and tellurium are among the most prominent metalloids used in structural and industrial alloys, contributing to applications in construction, manufacturing, and energy storage. These elements form intermetallic compounds or solid solutions that refine microstructures and improve performance under mechanical stress. Silicon is widely incorporated into iron-based and aluminum-based alloys to modify their processing and properties. In steel production, ferrosilicon, an alloy containing 75% silicon and 25% iron, serves as a key deoxidizer by reacting with oxygen impurities to form silicon dioxide slag, thereby preventing defects like porosity and improving the steel's cleanliness and mechanical integrity.[139] This addition also enables precise control of silicon content to enhance strength and elasticity in structural steels. In aluminum casting alloys, silicon contents typically range from 5% to 23%, promoting excellent fluidity during pouring and reducing shrinkage, which allows for complex shapes in automotive pistons and engine blocks while maintaining low density and good corrosion resistance.[140][141] Antimony strengthens lead-based alloys, particularly in electrochemical and printing applications. In lead-acid batteries, antimony-lead grids (typically 2-6% Sb) provide superior mechanical stability and resistance to corrosion under cyclic charging, enabling deeper discharge cycles and longer service life compared to antimony-free alternatives.[142] For type metal, an alloy of antimony and tin (often 10-25% Sb with 50-80% Sn), antimony increases hardness and low-melting characteristics, ensuring sharp, durable impressions in historical printing processes.[143] Boron enhances wear resistance in nickel-based hardfacing alloys applied to surfaces exposed to abrasion, such as in mining equipment or turbine components. Additions of 1-3% boron in nickel-chromium-boron-silicon alloys promote the formation of hard boride phases, like nickel borides, which significantly boost hardness (up to 60 HRC) and reduce material loss during sliding or impact wear.[144] These alloys are deposited via thermal spraying or welding, offering a cost-effective overlay that extends component lifespan in harsh environments.[145] Tellurium improves the workability of copper alloys, making them suitable for precision machining in electrical and plumbing fittings. Copper-tellurium alloys (CDA 145, with 0.5-0.8% Te) exhibit a machinability rating of 85-90%, far superior to pure copper's 20%, due to tellurium's role in forming soft inclusions that act as chip breakers during cutting, while preserving high electrical conductivity (over 90% IACS).[146] Additionally, tellurium enhances ductility, allowing the alloy to withstand forming operations without cracking, which is essential for applications like screw-machine products and connectors.[147]Catalysts and Biological Roles
Metalloids serve as effective catalysts in several key chemical reactions due to their unique electronic properties that facilitate bond activation and intermediate stabilization. Boron compounds, in particular, have revolutionized organic synthesis through hydroboration, a process discovered by Herbert C. Brown in the 1950s that enables the selective addition of borane (BH₃) across carbon-carbon double bonds in alkenes, yielding organoboranes that can be oxidized to alcohols with anti-Markovnikov orientation and syn stereochemistry. This reaction proceeds under mild conditions, often at room temperature, and is tolerant of many functional groups, making it indispensable for stereoselective synthesis in pharmaceuticals and natural products. Brown's pioneering work on hydroboration-oxidation earned him the 1979 Nobel Prize in Chemistry, shared with Georg Wittig for complementary phosphorus-based methods.[148][149] In biological systems, metalloids fulfill essential roles that underpin structural integrity and metabolic processes. Silicon is crucial for biomineralization in diatoms, unicellular algae that form elaborate silica-based exoskeletons (frustules) within specialized silica deposition vesicles, enabling these organisms to thrive in aquatic environments and contribute approximately 20% of global primary productivity. The process involves silicon transporters that uptake silicic acid, followed by polycondensation into amorphous silica nanostructures templated by organic matrices, which provide mechanical protection and aid in nutrient cycling. Boron, meanwhile, acts as an essential micronutrient in plants, where it stabilizes cell walls by cross-linking rhamnogalacturonan-II pectins, supports pollen tube elongation for reproduction, and facilitates membrane function and hormone signaling; deficiencies manifest as stunted growth, brittle stems, and reduced seed set, affecting crops like alfalfa and citrus.[150][151] Medically, certain metalloids exploit their toxicity to pathogens while harnessing narrow therapeutic windows for human benefit. Antimony compounds, such as sodium stibogluconate (Pentostam), have been a cornerstone treatment for leishmaniasis since the 1940s, administered intravenously or intramuscularly at 20 mg Sb/kg/day for 10–20 days to visceral and cutaneous forms caused by Leishmania parasites; the drug inhibits trypanothione reductase, disrupting the parasite's antioxidant defense and redox metabolism. Similarly, arsenic trioxide (Trisenox) received FDA approval in 2000 for relapsed or refractory acute promyelocytic leukemia (APL), where it targets the PML-RARα fusion protein to promote degradation, induce differentiation, and trigger apoptosis in leukemic promyelocytes, achieving complete remission rates of 60–80% with minimal myelosuppression when combined with all-trans retinoic acid. Despite these applications, arsenic and antimony's toxicity—arsenic causing acute gastrointestinal distress, peripheral neuropathy, and chronic risks like skin cancer and cardiovascular disease via oxidative stress and DNA damage, while antimony induces cardiotoxicity, pancreatitis, and hepatic effects through similar mechanisms—necessitates precise dosing to maintain efficacy within safe margins, as evidenced by their classification as carcinogens by the International Agency for Research on Cancer.[152][153][154][155]Other Uses
Metalloids find specialized applications in glass production, where boron enhances the thermal properties of borosilicate glasses such as Pyrex. The incorporation of boron oxide (B₂O₃) into the glass matrix lowers the coefficient of thermal expansion, providing exceptional resistance to thermal shock and making it suitable for laboratory equipment and cookware that withstands rapid temperature changes.[156] Similarly, silicon dioxide (SiO₂) forms the basis of fused silica glass, an amorphous material prized for its low thermal expansion, high chemical resistance, and transparency across a wide spectrum, used in high-precision optics and laboratory vessels.[157] In optics, germanium is a key metalloid for infrared (IR) components due to its high transparency in the mid- to long-wave IR range of 2–14 μm, enabling its use in lenses for thermal imaging systems, night-vision devices, and spectroscopy.[158] Tellurium contributes to phase-change materials in rewritable optical storage media, such as CD-RW and DVD-RW discs, where alloys like germanium-antimony-tellurium (GST) switch between amorphous and crystalline states under laser irradiation to enable data recording and erasure.[159] Antimony sulfide (Sb₂S₃), also known as antimony trisulfide, is employed in pyrotechnics for its role in creating glittering and shimmering effects in fireworks, as it decomposes during combustion to produce reflective particles that enhance visual sparkle.[160] Historically, arsenic compounds, particularly arsenic sulfides like orpiment (As₂S₃), served as vivid yellow pigments in paints and dyes from ancient times through the 19th century, valued for their bright color despite toxicity concerns that later led to their decline.[161] Antimony trioxide (Sb₂O₃) acts as a crucial flame retardant synergist in plastics, particularly when combined with halogenated compounds, where it promotes the formation of volatile antimony halides that inhibit flame spread by interfering with radical chain reactions in the gas phase.[162] This application is significant in industries producing electronics housings, textiles, and automotive parts, with the global antimony trioxide flame retardant market valued at approximately $1.25 billion in 2024.[163]Occurrence and Production
Natural Abundance
Metalloids exhibit a wide range of abundances in Earth's crust, reflecting their geochemical behaviors and incorporation into silicate minerals. Silicon is the second most abundant element in the crust, comprising approximately 27.7% by mass, primarily due to its prevalence in silicate structures that dominate crustal rocks.[45] In contrast, boron is far less common, with an average crustal concentration of around 10 parts per million (ppm), while tellurium is extremely rare at approximately 1 part per billion (ppb).[164][80]| Element | Crustal Abundance (by mass) | Primary Form in Crust |
|---|---|---|
| Silicon | 27.7% | Silicates (e.g., quartz, feldspars) |
| Boron | 10 ppm | Borates |
| Germanium | 1.5 ppm | Sulfides (e.g., in zinc ores), silicates |
| Arsenic | 1.8 ppm | Sulfides (e.g., arsenopyrite) |
| Antimony | 0.2 ppm | Sulfides (e.g., stibnite) |
| Tellurium | 1 ppb | Sulfides (trace) |