Fact-checked by Grok 2 weeks ago

Silicon

Silicon is a chemical element with atomic number 14 and chemical symbol Si, classified as a tetravalent metalloid and semiconductor that occurs in two allotropes: amorphous and crystalline forms. Crystalline silicon appears as a hard, brittle solid with a blue-grey metallic lustre and a density of 2.329 g/cm³ at room temperature, where it exists as a solid with a melting point of 1414 °C and a boiling point of 3265 °C. It is relatively inert but reacts with halogens and dilute alkalis, and is resistant to most acids except hydrofluoric acid. As the second most abundant element in the Earth's crust after oxygen, silicon constitutes approximately 27.7% by weight, primarily occurring in silicate minerals such as quartz (SiO₂) and various aluminosilicates that form rocks, sand, and clay. Its three stable isotopes are ²⁸Si (92.2%), ²⁹Si (4.7%), and ³⁰Si (3.1%), with atomic mass averaging 28.085 u; silicon also has several radioactive isotopes, including ³²Si with a half-life of about 172 years. First isolated in impure form in 1823 by Swedish chemist Jöns Jacob Berzelius through the reduction of silica with potassium fluorosilicate, pure silicon was later produced in 1854 by French chemist Henri Étienne Sainte-Claire Deville via heating silicon tetrachloride with sodium. Silicon's semiconducting properties, stemming from its four valence electrons and ability to be doped with impurities to control electrical conductivity, make it indispensable in modern technology, particularly for fabricating integrated circuits, transistors, microprocessors, and photovoltaic solar cells that power electronics and renewable energy systems. Beyond electronics, it is a key component in producing ferrosilicon alloys for steelmaking, silicones for lubricants and medical implants, and silica-based materials for glass, ceramics, and abrasives like silicon carbide. Although not essential for human biology, silicon is present in trace amounts in the body and plays structural roles in diatoms, sponges, and plant tissues.

History

Discovery and early isolation

In 1787, French chemist Antoine Lavoisier proposed that silica (silicon dioxide), a common component of rocks and minerals, was likely the oxide of a previously unidentified chemical element, marking the first recognition of silicon's potential existence as a distinct substance. This insight built on earlier observations of silica's refractory properties, which resisted decomposition under heat and acids, suggesting it was not a simple earth but a compound. Attempts to isolate the element soon followed, but silicon's strong affinity for oxygen proved challenging. In 1808, British chemist Humphry Davy tried to reduce silica using electrolysis and other methods, producing impure residues but failing to obtain the pure element; he nonetheless proposed the name "silicium," derived from the Latin word for flint, silex. French chemists Joseph Louis Gay-Lussac and Louis Jacques Thénard also attempted isolation around 1811 by reacting potassium with silicon tetrafluoride, yielding a dark powder that contained silicon but was contaminated with impurities. The successful isolation came in 1823–1824, when Swedish chemist Jöns Jacob Berzelius heated potassium fluosilicate (K₂SiF₆) with excess potassium metal, producing amorphous silicon as a brown powder. Berzelius confirmed its elemental nature through further analysis, noting its refractory behavior—resistant to oxidation and melting only at high temperatures—and distinguishing it from metals and other nonmetals. He renamed it "silicon" to emphasize its non-metallic character, retaining the root from silex. In 1854, French chemist Henri Étienne Sainte-Claire Deville produced crystalline silicon by heating silicon tetrachloride with sodium. Subsequent early 19th-century experiments by Berzelius and others, including reactions with halogens to form volatile compounds like silicon tetrachloride, solidified silicon's status as a unique element in the periodic system. Prior to the isolation of elemental silicon, its compounds, particularly silica, had long been utilized in practical applications such as ceramics, pottery glazes, and glassmaking, dating back to ancient civilizations for tools, vessels, and ornaments. These uses exploited silica's durability and heat resistance, though the pure element's scarcity limited its immediate exploitation until later refinements.

Development of semiconductors

The development of silicon as a semiconductor material began with key discoveries in crystal purification and junction formation during the early 20th century. In 1940, Russell Ohl at Bell Laboratories accidentally discovered the p-n junction while investigating silicon crystals for use in radar detectors during World War II; he observed that a crack in a silicon boule created a rectifying barrier that exhibited photovoltaic effects, leading to a patent for a light-sensitive device in 1941. This breakthrough demonstrated silicon's potential for electronic rectification, though initial impurities limited its performance compared to germanium. Advancements in purification were essential to harness silicon's properties. In 1952, William Pfann at Bell Labs invented the zone melting (or zone refining) technique, which involved passing a narrow molten zone along a semiconductor rod to segregate impurities, achieving purity levels below one part per million—critical for reproducible device fabrication. This method, initially applied to germanium, was quickly adapted for silicon, enabling the growth of high-quality single crystals via the Czochralski process refined around the same period. These techniques addressed silicon's higher melting point and greater affinity for oxygen, making it viable for semiconductor applications. The transistor's invention marked a pivotal shift toward silicon. In December 1947, William Shockley, John Bardeen, and Walter Brattain at Bell Labs demonstrated the first point-contact transistor using germanium, earning them the 1956 Nobel Prize in Physics; however, silicon's superior thermal stability and abundance prompted rapid research to transition the technology. By 1954, Morris Tanenbaum at Bell Labs fabricated the first silicon junction transistor, achieving higher operating temperatures up to 150°C compared to germanium's limits. Independently, Gordon Teal's team at Texas Instruments produced the first commercial silicon transistor that year using grown-junction methods, enabling reliable high-power applications and spurring industry adoption. Early commercial exploitation focused on diodes and rectifiers. Building on Ohl's work, companies like General Electric and Sylvania began producing silicon rectifiers in the early 1950s for power conversion in industrial and military uses, offering greater efficiency and durability than vacuum tubes or selenium alternatives. By mid-decade, these devices powered early computers and telecommunications equipment, laying the groundwork for silicon's dominance in electronics.

The Silicon Age

The term "Silicon Age" emerged in the late 20th century to describe the era dominated by silicon-based technologies, paralleling historical periods like the Iron Age in marking a transformative shift in materials and society. This designation gained prominence in the 1990s as silicon's role in electronics became ubiquitous, fundamentally altering computing, communications, and daily life. The concept is closely tied to the rise of Silicon Valley, a region in California's Santa Clara Valley that began coalescing as a tech hub in the 1950s with the establishment of firms like Shockley Semiconductor Laboratory in 1955 and Fairchild Semiconductor in 1957, fostering innovation in silicon devices during the 1960s. By the late 1960s, the area's concentration of semiconductor activity had earned it the nickname "Silicon Valley," symbolizing the start of widespread silicon industrialization post-World War II. A pivotal milestone in silicon-based computing arrived in 1971 with Intel's introduction of the 4004 microprocessor, the first commercially available single-chip CPU, which integrated 2,300 transistors and enabled programmable logic on silicon wafers. This innovation marked the transition from discrete components to integrated systems, accelerating the digital revolution by making computing more accessible and compact. The semiconductor industry's growth exploded thereafter, with global sales surpassing $1 billion in 1966 and reaching $627 billion by 2024, propelled by Moore's Law—coined by Gordon Moore in his 1965 Electronics magazine article, which predicted the number of transistors on a chip would double approximately every year, later revised to every two years. This exponential scaling drove cost reductions and performance gains, fueling the proliferation of personal computers and consumer electronics. The 1980s brought the VLSI (very-large-scale integration) revolution, where advancements in design methodologies, notably from Carver Mead and Lynn Conway's 1980 textbook, enabled millions of transistors per chip, revolutionizing chip complexity and production efficiency. This era solidified silicon's dominance in integrated circuits, supporting the expansion of computing into businesses and homes. The 2000s witnessed a mobile computing boom, ignited by devices like the 2007 iPhone, which spiked demand for advanced silicon processors and memory, integrating computing into portable gadgets and spurring applications in telecommunications and multimedia. Global production dynamics shifted dramatically over decades, with Asia emerging as the dominant force; by 2025, the region accounted for over 70% of worldwide semiconductor manufacturing capacity, led by Taiwan, South Korea, and China, despite ongoing diversification efforts elsewhere. This concentration arose from investments in fabrication facilities starting in the 1970s, enabling cost-effective scaling and supply chain integration that powered the industry's growth into a multi-hundred-billion-dollar ecosystem.

Properties

Physical and atomic properties

Silicon is a chemical element with atomic number 14 and the electron configuration [Ne] 3s² 3p², featuring a covalent radius of 111 pm. These atomic characteristics position silicon within group 14 of the periodic table, where it forms tetrahedral covalent bonds due to its four valence electrons. At room temperature, silicon exists as a solid with a density of 2.329 g/cm³. It transitions to a liquid at its melting point of 1414 °C and vaporizes at a boiling point of 3265 °C. The stable allotrope of silicon adopts a diamond cubic crystal structure, characterized by a lattice parameter of approximately 0.543 nm, and displays a grayish metallic luster. Silicon exhibits a linear thermal expansion coefficient of 2.6 × 10⁻⁶ K⁻¹ and a specific heat capacity of 0.71 J/g·K. At the electronic level, it possesses an indirect band gap of 1.12 eV, a property that underlies its classification as a semiconductor and influences its electrical behavior.

Electrical and optical properties

Silicon exhibits intrinsic semiconducting behavior with a high resistivity of approximately $2.3 \times 10^5 Ω·cm at 300 K, arising from its indirect bandgap that limits thermal generation of charge carriers. This resistivity can be tuned over several orders of magnitude through doping: phosphorus introduces donor levels for n-type conduction, increasing electron concentration, while boron creates acceptor levels for p-type conduction, enhancing hole concentration. In pure silicon, charge carrier mobilities at 300 K reach 1400 cm²/V·s for electrons and 450 cm²/V·s for holes, reflecting the material's low effective masses and minimal scattering in the diamond cubic lattice. The static dielectric constant of silicon is 11.7, enabling effective electric field screening in devices, while its dielectric breakdown strength is approximately 0.3 MV/cm, beyond which avalanche multiplication occurs. Optically, silicon features an absorption edge near 1.1 eV, tied to its indirect bandgap of 1.12 eV, resulting in low absorption for photon energies below this threshold and high transparency in the infrared spectrum beyond about 1.1 μm. This infrared transparency underpins its application in infrared detectors, where it efficiently transmits longer wavelengths while absorbing visible light. Thermoelectrically, silicon displays a Seebeck coefficient that ranges from roughly -100 μV/K in n-type material to +100 μV/K in p-type, modulated by doping concentration which alters the Fermi level position relative to the band edges.

Chemical properties

Silicon exhibits tetravalent character, utilizing its four valence electrons to form covalent bonds, primarily in a tetrahedral geometry similar to carbon but with distinct preferences due to atomic size and electronegativity differences. The Si-Si bond strength averages 226 kJ/mol, which is relatively weak compared to carbon-carbon bonds, while the Si-O bond is exceptionally strong at 452 kJ/mol, favoring oxygen-containing structures over extended silicon chains. In contrast, the Si-H bond energy is 318 kJ/mol, enabling some hydridic reactivity but limiting stable molecular analogs to hydrocarbons. This bonding profile contributes to silicon's tendency to form extended network solids rather than discrete molecular compounds, unlike carbon, which readily produces stable liquid-phase organics; silicon lacks analogous stable liquid compounds and instead polymerizes into robust covalent lattices. Silicon demonstrates notable chemical stability under ambient conditions, resisting most acids due to the formation of a thin, impervious SiO₂ passivation layer upon exposure to oxygen, which inhibits further corrosion. However, it dissolves readily in hydrofluoric acid (HF), where the reaction proceeds via formation of soluble fluorosilicates, driven by the high stability of Si-F bonds: Si + 4HF → SiF₄ + 2H₂ (gaseous), though in aqueous media it forms H₂SiF₆. Similarly, silicon reacts with hot, concentrated alkalis such as NaOH to produce silicates and hydrogen: Si + 2NaOH + H₂O → Na₂SiO₃ + 2H₂. Silicon's high affinity for oxygen leads to spontaneous oxidation above approximately 900°C, where it reacts with O₂ to form SiO₂, a process central to thermal oxidation in semiconductor fabrication. The oxide of silicon, SiO₂, displays amphoteric behavior, albeit weakly, reacting with strong acids like HF to form silicon tetrafluoride (SiO₂ + 4HF → SiF₄ + 2H₂O) and with strong bases such as hot concentrated NaOH to yield silicates (SiO₂ + 2NaOH → Na₂SiO₃ + H₂O). This dual reactivity arises from the polarized Si-O bonds, allowing SiO₂ to act as a Lewis acid toward bases while being susceptible to fluoride attack, distinguishing it from purely acidic nonmetal oxides like CO₂.

Isotopes

Silicon has three stable isotopes: ^{28}Si, ^{29}Si, and ^{30}Si, with no stable isotope at mass number 31. The most abundant is ^{28}Si, comprising 92.223% of naturally occurring silicon, followed by ^{29}Si at 4.685% and ^{30}Si at 3.092%. Their respective atomic masses are 27.97692653465(44) u for ^{28}Si, 28.97649466490(52) u for ^{29}Si, and 29.973770136(23) u for ^{30}Si. These isotopic abundances contribute to silicon's standard atomic weight of [28.084, 28.086]. Silicon has over 20 known radioactive isotopes, most with half-lives of less than a day. Among them, ^{32}Si is notable for its relatively long half-life; measurements vary between approximately 100 and 276 years, with a recent value of 153 years, during which it undergoes beta decay to ^{32}P with a maximum energy of 0.21 MeV. Produced cosmogenically in the Earth's atmosphere through spallation of argon by cosmic rays, ^{32}Si serves as a tracer for geological and hydrological processes, enabling dating of groundwater and recent sediments up to about 1,000 years old. The isotope ^{29}Si, with its nuclear spin of 1/2 and natural abundance of approximately 4.7%, is particularly useful in nuclear magnetic resonance (NMR) spectroscopy for studying the structure of silicon-containing compounds, as its spin properties yield sharp, interpretable signals without quadrupolar broadening.
IsotopeAtomic Mass (u)Natural Abundance (%)StabilityNotes
^{28}Si27.97692653465(44)92.223(19)StableMost abundant
^{29}Si28.97649466490(52)4.685(8)StableSpin 1/2; used in NMR
^{30}Si29.973770136(23)3.092(11)Stable-
^{32}Si~31.974<0.0001 (cosmogenic)RadioactiveHalf-life ~153 y (measurements vary 100-276 y); β⁻ decay

Occurrence

Abundance in the universe and Earth

Silicon is the eighth most abundant element in the universe by mass fraction, accounting for approximately 0.065% of the total elemental mass, and ranks seventh by atomic abundance. It is primarily synthesized through the oxygen-burning process in the cores of massive stars, where temperatures reach about 2 billion kelvin, fusing lighter elements like neon and oxygen into silicon and other intermediates such as sulfur and magnesium. This nucleosynthetic pathway contributes significantly to silicon's prevalence in cosmic dust, interstellar medium, and stellar atmospheres. In the solar system, silicon's abundance in the Sun's photosphere is estimated at a mole fraction of about 3.5 × 10^{-5} (or logarithmic abundance A(Si) = 7.51), making it one of the dominant heavy elements after oxygen and carbon. This value aligns closely with cosmic averages and reflects the element's incorporation during the Sun's formation from a protostellar nebula enriched by previous stellar generations. On Earth, silicon is the second most abundant element in the crust, comprising 27.7% by mass, predominantly bound as silicon dioxide (SiO₂) in igneous, sedimentary, and metamorphic rocks. During planetary differentiation, silicon exhibits moderately siderophile behavior, partitioning preferentially into the metallic core under high-pressure and high-temperature conditions, which results in a depletion of silicon in the mantle relative to chondritic meteorites (with bulk silicate Earth Si content ~15-20% lower than CI chondrites). This core sequestration leads to relative enrichment of silicon in the silicate mantle and crust, enhancing its availability in terrestrial rocks. Variations in silicon abundance appear in meteorites; for instance, enstatite chondrites show elevated silicon contents (up to ~18 wt% SiO₂ in some phases) due to their highly reduced formation environments, contrasting with ordinary chondrites at ~15-16 wt%.

Natural minerals and sources

Silicon primarily occurs in nature as a component of silicate minerals, which dominate the Earth's crust. Quartz (SiO₂), the most common silicon-bearing mineral, constitutes approximately 12% of the Earth's crust by volume and is found in a wide range of igneous, metamorphic, and sedimentary rocks. It forms colorless to smoky crystals in granitic pegmatites and is a key constituent of sandstones and quartzites. Other silicate minerals are even more prevalent, with feldspars comprising about 60% of the crust and serving as the single most abundant mineral group. Feldspars include alkali varieties like orthoclase (KAlSi₃O₈), which occurs in granites and syenites, and plagioclase types that dominate basaltic and andesitic rocks. Micas, such as muscovite and biotite, contribute to the silicate content in schists and granites through their layered structures incorporating silicon-oxygen sheets. Clays, including kaolinite and illite, form from the weathering of feldspars and other silicates, accumulating in soils and sedimentary deposits as fine-grained phyllosilicates. Amorphous forms of silica, lacking a crystalline structure, are also significant sources. Opal consists of hydrated amorphous silica (SiO₂·nH₂O) and forms colorful deposits in sedimentary environments or as infillings in voids. Diatomaceous earth arises from the accumulation of opal skeletons from diatoms, microscopic algae, creating porous, lightweight sediments in ancient lake and ocean basins. Native silicon, in its elemental form, is exceedingly rare on Earth but has been identified in extraterrestrial materials. Silicon carbide (SiC), a presolar mineral, occurs in carbonaceous chondrites like the Murray meteorite, preserved as interstellar grains within these primitive asteroids. Siliceous deposits also form through geothermal and volcanic activity. Hot springs in volcanic regions precipitate siliceous sinter, a low-density, porous material rich in opal, as silica supersaturates and deposits from cooling geothermal fluids. Volcanic sources contribute via rhyolitic lavas and ash, which are high in silica content, leading to extensive siliceous rock formations in caldera systems.

Production

Metallurgical processes

Metallurgical-grade silicon, with a purity of 98-99%, is primarily produced through high-temperature reduction processes starting from silica-rich raw materials such as quartz. The dominant method is carbothermic reduction, accounting for approximately 95% of global production, where silica (SiO₂) is reduced by carbon in submerged electric arc furnaces. This process operates at temperatures around 1900°C, yielding silicon metal along with carbon monoxide gas as a byproduct. The reaction can be represented as: \text{SiO}_2 + 2\text{C} \rightarrow \text{Si} + 2\text{CO} The process involves feeding quartz and carbonaceous reductants like coke, charcoal, or wood chips into the furnace, where the arc generates the necessary heat for the endothermic reaction. An alternative approach for achieving slightly higher purity levels is aluminothermic reduction, which uses aluminum as the reductant instead of carbon, producing silicon and alumina slag. This method follows the reaction: $3\text{SiO}_2 + 4\text{Al} \rightarrow 3\text{Si} + 2\text{Al}_2\text{O}_3 It is less common than carbothermic reduction but offers potential for reduced carbon emissions and integration with aluminum recycling streams, though it remains primarily experimental or used for alloy production. These processes are highly energy-intensive, requiring 10-13 kWh per kilogram of silicon produced, with electricity sourced mainly from hydropower in regions like Norway or coal in major producers like China. A key byproduct is ferrosilicon, an iron-silicon alloy formed when iron impurities are present, which is widely used in steel deoxidation to remove oxygen and improve castability. Global production of metallurgical-grade silicon reached an estimated 4.6 million metric tons in 2024, with projections for similar or slightly higher output in 2025, dominated by China at approximately 85% of the total.

High-purity silicon for electronics

High-purity silicon for electronics begins with metallurgical-grade silicon, which contains 0.5% to 1.5% impurities and serves as the starting material for further refinement. The primary method for producing electronic-grade polysilicon is the Siemens process, which involves converting metallurgical silicon into trichlorosilane (TCS) through the reaction
\ce{Si + 3HCl -> SiHCl3 + H2}
at approximately 300°C in a fluidized bed reactor. The TCS is then purified via fractional distillation to remove impurities with differing boiling points, followed by decomposition in a high-temperature reactor at 1150°C using hydrogen:
\ce{SiHCl3 + H2 -> Si + 3HCl}
This deposits silicon onto heated seed rods, yielding polysilicon rods up to 2 meters long.
An alternative to the batch-style Siemens deposition is the fluidized bed reactor (FBR) process, which enables continuous silicon deposition from silane gas (\ce{SiH4}) onto silicon particles, achieving up to 90% lower energy consumption compared to traditional methods. To produce single-crystal ingots suitable for semiconductors, purified polysilicon is melted and grown via the Czochralski process, the predominant method accounting for over 90% of semiconductor silicon production, involving the controlled pulling of a seed crystal from the melt to form large cylindrical ingots. The float-zone (FZ) process is an alternative niche method for growing high-purity single-crystal silicon, particularly for applications requiring ultra-low oxygen content, though it accounts for less than 10% of production compared to Czochralski. Advanced semiconductor applications require silicon at 11N purity (99.999999999%), achieved through these combined refining steps to minimize defects and ensure optimal electrical performance in integrated circuits.

Sustainable production methods

Efforts to enhance the sustainability of silicon production have intensified post-2020, focusing on reducing greenhouse gas emissions from energy-intensive processes like carbothermal reduction. Innovations include carbon capture and reuse technologies, renewable energy integration, material recycling, and alternative electrochemical methods, aiming to lower the sector's carbon footprint while maintaining production scalability. Elkem, a leading Norwegian silicon producer, launched a pilot project in 2023 at its Rana smelter to demonstrate carbon capture and storage (CCS) in silicon arc furnaces, achieving capture rates of up to 95% of CO₂ emissions during approximately 3,000 operational hours. This initiative, part of Elkem's broader Sicalo concept, involves recirculating process off-gas to reuse captured carbon as a reductant in the furnace, effectively eliminating nearly all direct CO₂ emissions from silicon production. Building on this, Elkem secured a €1.8 million EU grant in November 2024 to further develop carbon looping for zero-CO₂ silicon and manganese alloys, with demonstrations planned through 2025 in collaboration with SINTEF and NORCE. Additionally, a 2025 SINTEF-supported project at Elkem explores exhaust gas recirculation in silicon furnaces to boost CO₂ concentrations for more cost-effective CCS integration. In China, the world's largest silicon producer, hydropower has driven greener production, particularly in regions like Xinjiang. In June 2025, national silicon metal output rose 6.5% month-over-month to 327,700 metric tons, largely due to resumed operations in Xinjiang supported by stable hydropower supplies, alongside seasonal hydropower gains in Sichuan and Yunnan. This shift aligns with China's 2025 industrial green energy mandates, requiring polysilicon producers to source 25-70% of power from renewables, contributing to overall emission reductions in high-energy sectors. Recycling silicon from end-of-life solar panels and electronic waste represents a circular economy approach to sustainable supply. Advanced mechanical and chemical processes can recover over 90% of silicon from crystalline silicon-based photovoltaic modules, as demonstrated by companies like SolarCycle. The European Union's Waste Electrical and Electronic Equipment (WEEE) Directive enforces targets of at least 85% overall material recovery and 80% preparation for reuse by 2030, with compliance rates already reaching 88% in 2025 for solar waste streams. These efforts are projected to handle 400,000-1 million tons of annual PV waste in Europe by 2030, reclaiming high-purity silicon to offset virgin material demand. As a lab-scale alternative to traditional carbothermal reduction, electrochemical processes in molten salts offer a low-carbon pathway for silicon production. Post-2020 advancements include electrodeposition of high-purity (up to 99.9998%) crystalline silicon films from SiO₂ in CaCl₂-CaO melts at 850°C, achieving solar cell efficiencies of 3.1% without CO₂ generation. Techniques like the FFC Cambridge process variant have produced nanostructured silicon nanowires and nanotubes from nano-SiO₂ or CaSiO₃ in CaCl₂-based salts at 650-900°C, with yields up to 90% and applications in energy storage. These methods, reviewed in 2023, emphasize reduced energy use and emissions compared to conventional high-temperature smelting, though scaling remains a challenge.

Chemical compounds

Silanes, silicides, and halides

Silanes are binary compounds of silicon and hydrogen, analogous to hydrocarbons but exhibiting greater reactivity due to the weaker Si-H bond compared to C-H bonds. The simplest silane, monosilane (SiH₄), is a colorless, flammable gas with a boiling point of -112°C and a melting point of -185°C. It is pyrophoric, igniting spontaneously in air at concentrations above 1.4% due to exothermic oxidation forming silicon dioxide and water. SiH₄ hydrolyzes in the presence of water or moisture to produce hydrogen gas and silicon dioxide via the reaction SiH₄ + 2H₂O → SiO₂ + 4H₂, though the process is slow at room temperature and accelerates under basic conditions or elevated temperatures above 555°C. Commercially, SiH₄ is synthesized by the reaction of magnesium silicide (Mg₂Si) with dilute acids such as hydrochloric acid, yielding up to 50% SiH₄ along with higher silanes, or via reduction of silicon halides with alkali metal hydrides. Higher silanes, such as disilane (Si₂H₆), are less stable analogs with Si₂H₆ having a boiling point of -14°C; they are prepared similarly from Mg₂Si acidification (yielding ~30% Si₂H₆) or reduction of hexachlorodisilane with lithium aluminum hydride, and exhibit even greater pyrophoricity, decomposing thermally above 300°C. Silicides are intermetallic compounds formed between silicon and metals, often exhibiting semiconducting or metallic properties depending on composition and structure. Magnesium silicide (Mg₂Si) is an n-type semiconductor with a cubic antifluorite structure, valued for its thermoelectric applications due to a high Seebeck coefficient of approximately -487 µV/K and low thermal conductivity when nanostructured. It is synthesized via mechanochemical methods, such as high-energy ball milling of elemental Mg and Si powders under inert atmosphere, or by direct combination at elevated temperatures around 800°C. Iron silicide (FeSi) adopts a cubic B20 structure and is a hard, brittle intermetallic phase used in alloys to enhance mechanical strength and abrasion resistance, with an indentation hardness exceeding that of other stable iron silicides at room temperature. FeSi is produced through solid-state reactions of Fe and Si powders at temperatures above 1000°C or via selective chemical vapor deposition routes. Many silicides, including Mg₂Si and calcium silicide, react vigorously with acids to liberate silane gas; for instance, Mg₂Si + 4HCl → SiH₄ + 2MgCl₂ + 2H₂, a process exploited in laboratory silane generation but requiring careful handling due to the pyrophoric nature of the product. Silicon halides are tetrahedral compounds of the formula SiX₄ (X = F, Cl, Br, I), characterized by high reactivity toward nucleophiles owing to the polar Si-X bonds. Silicon tetrachloride (SiCl₄) is a colorless, fuming liquid with a pungent odor, boiling at 57.6°C, and adopts a tetrahedral geometry with Si-Cl bond lengths of approximately 2.02 Å. It is synthesized industrially by the direct chlorination of metallurgical silicon: Si + 2Cl₂ → SiCl₄, typically conducted at 600°C to ensure complete reaction and minimize side products. SiCl₄ reacts exothermically with water to form silicic acid and hydrochloric acid: SiCl₄ + 4H₂O → Si(OH)₄ + 4HCl, evolving significant heat and fumes. Silicon tetrafluoride (SiF₄) is a colorless gas boiling at -86°C, also tetrahedral, and highly reactive; it hydrolyzes in moist air to form silica and hydrogen fluoride but notably reacts with hydrofluoric acid to yield hexafluorosilicic acid: SiF₄ + 2HF → H₂SiF₆. SiF₄ is prepared by direct fluorination of silicon with fluorine gas or as a byproduct in phosphate processing, where it is captured and converted to H₂SiF₆.

Oxides and silicates

Silicon dioxide, SiO₂, is the most abundant silicon compound in the Earth's crust and exists in both amorphous and crystalline forms. The amorphous form, known as fused silica, lacks long-range order and is produced by melting and rapid cooling of quartz sand. Crystalline forms include quartz, which has a trigonal structure, and cristobalite, which adopts a tetragonal or cubic arrangement. The density of α-quartz is 2.65 g/cm³, while fused silica has a lower density of 2.20 g/cm³. Silicon dioxide exhibits polymorphism, with phase transitions influenced by temperature. The stable low-temperature form, α-quartz, reversibly transforms to the high-temperature β-quartz at 573°C, involving a change from trigonal to hexagonal symmetry without altering the basic tetrahedral SiO₄ units. At higher temperatures, around 1470°C, β-quartz can convert to β-cristobalite, a cubic polymorph, though this often requires impurities or specific conditions; upon cooling, β-cristobalite inverts to α-cristobalite near 200–270°C. These transitions are critical for understanding the behavior of silica in geological and industrial processes. The solubility of SiO₂ is low in neutral water, with quartz dissolving to approximately 6 ppm at 25°C under equilibrium conditions, limiting its transport in aqueous environments. However, solubility increases significantly in alkaline solutions, where SiO₂ reacts to form soluble silicate ions, such as in sodium hydroxide, enabling its dissolution for geochemical weathering and industrial applications. Thermally, SiO₂ demonstrates high stability, melting at 1713°C for the cristobalite form, which contributes to its use in refractories that withstand extreme temperatures in furnaces. Silicates, compounds derived from SiO₄⁴⁻ tetrahedra linked in various polymerized structures, constitute over 90% of the Earth's crust and are fundamental to rock formation. Isolated tetrahedra, or nesosilicates, occur in minerals like olivine ((Mg,Fe)₂SiO₄), common in ultramafic rocks. Chain silicates, or inosilicates, feature single chains in pyroxenes (e.g., MgSiO₃) or double chains in amphiboles, forming key components of basalts and metamorphic rocks. Sheet silicates, or phyllosilicates, consist of layered T-O structures in micas (e.g., muscovite, KAl₂(AlSi₃O₁₀)(OH)₂), which influence soil properties and sedimentary deposits. Framework silicates, or tectosilicates, involve fully connected three-dimensional networks, as in zeolites (aluminosilicates with open structures) and quartz itself, dominating granitic and volcanic rocks. These structural variations determine the mechanical, thermal, and chemical properties of geological materials.

Organosilicon compounds and silicones

Organosilicon compounds are a class of chemical substances featuring at least one silicon-carbon (Si-C) bond, which imparts stability due to the comparable electronegativities of silicon and carbon atoms. These compounds encompass a wide range of structures, including monomeric silanes and polymeric siloxanes, with silicones representing a prominent subclass of the latter. Silicones, specifically polydimethylsiloxanes with the repeating unit [(CH_3)_2SiO]_n, are renowned for their chemical inertness and thermal stability, maintaining structural integrity up to approximately 250°C without significant degradation. The synthesis of silicones typically begins with the hydrolysis of dimethyldichlorosilane, (CH_3)_2SiCl_2, in the presence of water, which yields a mixture of cyclic oligomers and linear silanol intermediates through condensation reactions. These oligomers then undergo ring-opening polymerization, often catalyzed by acids or bases, to form high-molecular-weight polydimethylsiloxane chains, enabling the production of versatile silicone materials like fluids, elastomers, and resins. Key properties of polydimethylsiloxanes include a notably low glass transition temperature of around -123°C, which confers exceptional flexibility even at cryogenic conditions, alongside hydrophobic behavior arising from low surface energy and a dielectric constant ranging from 2.3 to 4, making them suitable for insulating applications. Beyond polydimethylsiloxanes, other organosilicon compounds include general siloxanes with varying organic substituents on the Si-O-Si backbone and silanes such as trimethylsilane, (CH_3)_3SiH, which serve as volatile precursors in chemical vapor deposition processes. Recent advancements have focused on developing degradable silicon-based polymers to address environmental concerns with traditional non-biodegradable silicones, employing step-growth polymerization techniques that incorporate cleavable linkages for controlled breakdown under specific conditions.

Applications

Alloys and metallurgy

Silicon plays a crucial role in metallurgy as an alloying element that enhances the mechanical properties, castability, and resistance to oxidation in various metal matrices, particularly in iron, aluminum, and copper-based alloys. These alloys are produced using metallurgical-grade silicon, derived from the reduction of silica with carbon in electric arc furnaces. One of the most prominent silicon alloys is ferrosilicon (Fe-Si), which contains 10-90% silicon by weight and is essential in steel production. It serves primarily as a deoxidizer to remove oxygen from molten steel, preventing defects such as porosity, and as an alloying agent to improve strength, elasticity, and magnetic properties. The 75% silicon grade (FeSi 75) is particularly valued for its high deoxidizing efficiency and is widely used in the manufacture of low-alloy steels and cast irons. In aluminum alloys, silicon is a key component in silumin (Al-Si) series, typically ranging from 5-22% silicon, which are favored for their excellent castability in complex shapes. These hypoeutectic and hypereutectic alloys exhibit superior fluidity during casting, allowing for thin-walled components with reduced shrinkage and improved feeding characteristics. Common applications include automotive engine blocks and cylinder heads, where the silicon content enhances wear resistance and thermal conductivity while maintaining lightweight properties. The Al-Si phase diagram reveals a eutectic point at 12.6 wt% silicon and 577°C, where the liquid transforms into a mixture of aluminum and silicon phases, optimizing the microstructure for casting performance. This eutectic composition enables fine lamellar structures that balance strength and ductility in silumin alloys. Silicon-bronze (Cu-Si) alloys, containing 3-5% silicon, offer exceptional corrosion resistance due to the formation of a protective oxide layer, making them ideal for harsh environments. These alloys are commonly used in marine hardware such as propellers, fittings, and fasteners, where they withstand saltwater exposure without significant degradation. Their combination of high strength, weldability, and non-magnetic properties further supports applications in naval architecture and offshore equipment. The global market for silicon alloys underscores their industrial significance, valued at USD 12.8 billion in 2024 and projected to reach USD 19.4 billion by 2034, driven by demand in steelmaking, automotive, and marine sectors.

Electronics and semiconductors

Silicon's role in electronics and semiconductors is foundational, primarily due to its ability to be doped with impurities to create n-type and p-type semiconductors, enabling the formation of p-n junctions essential for devices like diodes and transistors. This doping process alters silicon's electrical conductivity, allowing precise control over charge carrier flow in integrated circuits. In complementary metal-oxide-semiconductor (CMOS) technology, silicon wafers are doped to fabricate metal-oxide-semiconductor field-effect transistors (MOSFETs), which serve as the building blocks for logic gates and memory cells in microprocessors and other chips. Advancements in CMOS scaling have pushed silicon-based transistors toward sub-2nm nodes by 2025, addressing the limitations of traditional planar designs through innovative architectures. Intel's RibbonFET technology, a gate-all-around nanosheet transistor, enables silicon CMOS scaling to gate lengths as small as 6nm while targeting 2nm-class nodes for production in 2025, improving drive current and reducing leakage for high-performance computing. This approach maintains silicon's dominance in logic scaling despite challenges like quantum tunneling, with complementary efforts from foundries like TSMC's N2 process incorporating similar nanosheet MOSFETs for AI and mobile applications. Silicon wafer production remains critical for the semiconductor supply chain, with 300mm wafers serving as the standard substrate for advanced nodes due to their larger area and higher yield potential. In Q3 2025, worldwide shipments of silicon wafers increased by 3% year-on-year to 3,313 million square inches, driven primarily by demand for 300mm wafers in AI, cloud infrastructure, and memory fabrication. This growth reflects sustained investments in fabrication capacity to support the escalating needs of electronics manufacturing. To overcome planar scaling limits, advanced packaging techniques like 3D stacking and chiplet integration have emerged as key enablers for silicon-based AI accelerators. These methods involve vertically stacking multiple silicon dies or modular chiplets—small, specialized silicon blocks—connected via high-density interconnects, allowing heterogeneous integration of compute, memory, and I/O components. NVIDIA's GPUs, such as the H100, exemplify this by using chiplet designs with 3D-stacked high-bandwidth memory (HBM3), enhancing data throughput for AI training while reducing overall power consumption and footprint in data centers. Silicon photonics extends silicon's utility beyond electronics into optical interconnects, integrating lasers and waveguides directly onto silicon substrates to enable high-speed data transmission. This co-integration leverages silicon's compatibility with CMOS fabrication processes, allowing hybrid photonic-electronic chips that convert electrical signals to optical ones for minimal conversion overhead. In data centers, such advancements reduce latency by up to 50% compared to traditional copper-based links, as demonstrated by NVIDIA's silicon photonics switches that embed photonic elements into network ICs for efficient AI workload scaling. The semiconductor industry, heavily reliant on silicon, reached $627 billion in global sales in 2024, with AI applications driving a 19% year-on-year growth. This expansion is projected to continue, with sales expected to reach approximately $701 billion in 2025. Emerging trends like software-defined silicon—customizable, programmable architectures—further optimize silicon chips for diverse AI tasks, enhancing flexibility without hardware redesigns.

Photovoltaics and solar energy

Silicon plays a central role in photovoltaic (PV) technology, particularly through crystalline silicon (c-Si), which dominates the solar cell market with approximately 95% share due to its balance of efficiency, scalability, and cost-effectiveness. Crystalline silicon encompasses monocrystalline and polycrystalline variants, where monocrystalline cells, produced from single-crystal ingots, achieve commercial module efficiencies up to 22% by leveraging high-purity wafers that minimize defects and enhance charge carrier mobility. Polycrystalline silicon cells, formed from multiple crystal grains, offer slightly lower efficiencies around 18% but have historically provided a cost advantage through simpler manufacturing processes, though their market presence has diminished in favor of advanced monocrystalline designs. Amorphous silicon (a-Si) represents a thin-film alternative to crystalline forms, deposited in non-crystalline layers to enable flexible and lightweight modules with efficiencies around 10%. While a-Si reduces material usage and production costs compared to c-Si, it suffers from stability issues, such as light-induced degradation (Staebler-Wronski effect), limiting long-term performance and adoption to niche applications like building-integrated photovoltaics. Advancements in cell architectures have significantly boosted silicon PV performance, with passivated emitter rear cell (PERC) technology adding a rear passivation layer to reduce recombination losses, and heterojunction (HJT) combining crystalline silicon with thin amorphous silicon layers for improved passivation and carrier collection. By 2025, these innovations have enabled lab-scale efficiencies exceeding 25%, exemplified by HJT modules reaching 25.44%. Further progress includes tandem cells stacking perovskites atop silicon bottom cells to harness a broader spectrum, achieving certified efficiencies over 30% and approaching the practical limit for silicon-based systems. Global silicon demand for PV applications is projected at about 1.4 million metric tons in 2025, driven by expanding installations, with the sector anticipating a compound annual growth rate (CAGR) of around 10.5% through 2030 amid rising renewable energy targets.

Batteries and energy storage

Silicon has emerged as a promising anode material for lithium-ion batteries due to its exceptionally high theoretical specific capacity of 4200 mAh/g, which is approximately ten times greater than that of conventional graphite anodes at 372 mAh/g. This capacity arises from silicon's ability to alloy with lithium to form Li4.4Si, enabling higher energy densities for applications in electric vehicles and portable electronics. However, a major challenge is the material's volume expansion of up to 300% during lithiation and delithiation cycles, which leads to mechanical stress, pulverization, and rapid capacity fading. To address these issues, researchers have developed nanostructured silicon forms, such as nanowires and porous particles, which accommodate volume changes more effectively through increased surface area and reduced diffusion lengths. Silicon-carbon (Si-C) and silicon-graphene (Si-Gr) composite hybrids further enhance stability by buffering expansion with the elastic properties of carbon matrices, improving cycling performance and electrical conductivity. These advancements have facilitated commercialization efforts, with production capacity for silicon anode materials projected to expand sixfold from 2024 to 2030 as companies scale up manufacturing. Silicon oxide (SiOx) anodes represent a practical variant, offering a balance between high capacity and moderated volume expansion compared to pure silicon, with capacities up to 2000 mAh/g depending on the oxygen content. These materials are increasingly integrated into electric vehicle batteries to boost energy density, as seen in production cells from major manufacturers that incorporate SiOx blends for enhanced range and faster charging. Beyond electrochemical storage, silicon serves as a phase change material in high-temperature thermal energy systems, leveraging its melting point of 1414°C and latent heat of fusion of approximately 1.8 MJ/kg to store significant thermal energy. In concentrated solar power (CSP) plants, molten silicon systems enable efficient dispatchable heat storage, converting intermittent solar input into continuous high-grade thermal output for electricity generation or industrial processes. The silicon anode sector is poised for substantial growth, with the global market for these battery materials forecasted to exceed US$15 billion by 2035, driven by demand for higher-performance lithium-ion batteries in transportation and grid applications.

Other industrial uses

Silicon carbide (SiC), a compound derived from silicon, serves as a key material in abrasives due to its exceptional hardness, rated at 9.5 on the Mohs scale, which enables effective material removal in demanding applications. This hardness, combined with its sharp crystalline structure, makes SiC ideal for manufacturing grinding wheels used in precision machining of metals, ceramics, and composites, where it provides superior cutting efficiency and durability compared to traditional abrasives like aluminum oxide. In industrial settings, these wheels are employed for tasks such as surface finishing in automotive and aerospace components, contributing to reduced processing times and improved workpiece quality. In ceramics and glass production, silicon dioxide (SiO₂) plays a foundational role, forming the primary constituent in materials valued for thermal stability and optical clarity. Borosilicate glass, such as that used in Pyrex laboratory ware, typically contains approximately 80% SiO₂, which imparts resistance to thermal shock and chemical corrosion, allowing it to withstand rapid temperature changes without fracturing. Complementing this, fused quartz—essentially pure SiO₂—finds extensive use in optics due to its high transparency across ultraviolet to infrared wavelengths and low thermal expansion coefficient, making it suitable for lenses, prisms, and windows in high-precision instruments like spectrometers and laser systems. These properties ensure minimal distortion in optical paths, supporting applications in scientific research and telecommunications. Silicon nanocrystals, often referred to as silicon quantum dots, are emerging in optoelectronic devices for their size-tunable photoluminescence, enabling emission wavelengths from approximately 400 nm in the blue-violet range to 1000 nm in the near-infrared. This tunability arises from quantum confinement effects, allowing precise control over color output in light-emitting diodes (LEDs) and display technologies, where they offer advantages in color purity and efficiency over conventional phosphors. In displays, silicon quantum dots enhance brightness and viewing angles, positioning them as a biocompatible alternative to heavy-metal-based quantum dots in consumer electronics. As a filler, fumed silica—nanoscale amorphous SiO₂ produced via flame hydrolysis—reinforces rubber compounds, particularly in tire manufacturing, by increasing tear resistance and overall mechanical integrity. When incorporated at levels of 10-30% by weight, it forms a network within the polymer matrix that disperses stress, reducing crack propagation and enhancing durability under dynamic loads, such as those experienced during vehicle operation. This application not only improves tire longevity but also contributes to better grip and reduced hysteresis, aligning with demands for high-performance elastomers in automotive and industrial contexts. Looking to 2025, silicon-based materials are gaining traction in additive manufacturing, particularly through advancements in silicone 3D printing techniques that enable the fabrication of flexible, biocompatible prototypes for medical devices and soft robotics. These methods leverage direct ink writing and vat photopolymerization to achieve resolutions below 100 μm, expanding applications in customized implants and seals. Concurrently, silicones are integrating into high-performance polymers, enhancing properties like thermal resistance and flexibility in composites for aerospace and electronics, with the silicone market projected to grow from $24.65 billion in 2024 to $28.73 billion in 2025 (over 16% growth for that year) and a longer-term CAGR of approximately 5-7% through 2030 driven by sustainable formulations. This trend underscores silicon's role in developing eco-friendly, high-strength materials for next-generation engineering challenges.

Biological role

Role in plants and diatoms

Silicon plays a crucial role in the structural integrity of diatoms, unicellular algae that form intricate silica-based cell walls known as frustules. These frustules are composed of biogenic opal, a hydrated form of amorphous silica (SiO₂·nH₂O), which provides rigid protection and support while allowing efficient light transmission for photosynthesis. In diatoms, biogenic silica constitutes a significant portion of the biomass, often comprising 20–50% of the dry weight through the deposition of this biogenic silica during cell division and growth. In siliceous sponges, silicon forms the basis of spicules, providing skeletal support through amorphous silica structures. In higher plants, particularly grasses, silicon accumulates as phytoliths—microscopic deposits of amorphous silica that reinforce cell walls and enhance mechanical strength. Grasses such as rice (Oryza sativa) are notable silicon accumulators, with silicon levels reaching up to 5% of the plant's ash content or 10% of shoot dry weight under optimal conditions, forming phytoliths that contribute to tissue rigidity and deter herbivory by abrading insect mouthparts. These phytoliths, deposited intracellularly and extracellularly, provide physical barriers that improve plant erectness and resistance to lodging, while also amplifying defenses against pests like stem borers. Plants take up silicon primarily as uncharged silicic acid (H₄SiO₄) from the soil solution through aquaporin channels in root cell membranes, followed by radial transport to the xylem and axial movement to shoots. Once in plant tissues, H₄SiO₄ rapidly polymerizes into silica gel (SiO₂·nH₂O) within cell walls and lumens, especially in silica-accumulating species, where it integrates with organic matrices like hemicellulose to form stable phytoliths. This polymerization process is pH-dependent and occurs at concentrations above 2 mM, creating a durable amorphous structure that enhances overall plant architecture. The incorporation of silicon confers multiple benefits to plants, including improved tolerance to abiotic and biotic stresses. In drought-prone environments, silicon supplementation strengthens cell walls, reduces transpiration losses, and boosts water use efficiency, leading to enhanced photosynthetic rates and root hydraulic conductance in crops like rice and maize. For disease resistance, silicon induces physical barriers and activates defense pathways, such as the production of phenolic compounds, which suppress pathogens like rice blast fungus (Magnaporthe oryzae) and reduce infection severity by up to 50% in treated plants. Field studies demonstrate that silicon fertilization can increase crop yields by 20–30%, as seen in rice under combined stress conditions, by mitigating oxidative damage and improving nutrient uptake. Diatoms drive the global silicon cycle in marine ecosystems, fixing approximately 240 Tmol of silicon per year through frustule formation, which accounts for the majority of biogenic silica production in the oceans and influences nutrient dynamics and carbon export.

Human nutrition and health effects

Silicon is obtained primarily through the diet, with average daily intakes estimated at 20–50 mg for adults in Western populations, derived mainly from sources such as drinking water, cereal grains, and beer. These plant-based foods contribute to silicon's dietary presence, often in forms that can be converted in the body. The most bioavailable form of silicon for human absorption is orthosilicic acid, a monomeric species that readily enters the bloodstream from the gastrointestinal tract. Research has proposed several potential benefits of silicon for human health, particularly in supporting bone and connective tissue integrity. Silicon appears to play a role in collagen cross-linking and the formation of glycosaminoglycans, which are essential for the structural stability of bones, skin, and cartilage. Pioneering studies by Edith Carlisle in the 1970s and 1980s demonstrated that silicon deprivation in animal models led to impaired bone mineralization and connective tissue development, suggesting a possible preventive role against conditions like osteoporosis. However, more recent reviews, including those up to 2024, indicate that while silicon supplementation may enhance bone mineral density in some human trials, its status as an essential nutrient remains unestablished due to insufficient evidence of strict physiological requirements. Silicon deficiency is rare in humans owing to its ubiquity in the diet, but experimental silicon depletion in chicks and rats has been associated with abnormalities such as thin, fragile skin, poor hair growth, and skull deformities. To address potential inadequacies, dietary supplements often provide silicon as choline-stabilized orthosilicic acid (ch-OSA), a form that maintains solubility and bioavailability; typical recommended doses range from 5 to 10 mg of silicon per day.

Microbial and environmental interactions

Silicon plays a pivotal role in marine ecosystems through its interactions with microbes, particularly diatoms, which are unicellular algae that construct their cell walls from biogenic silica. Diatoms are major primary producers in the ocean, contributing approximately 40% of total oceanic primary production and facilitating the biological carbon pump by exporting organic carbon to deeper waters. This process influences global carbon cycling and CO₂ drawdown, as diatom blooms sequester atmospheric CO₂ through photosynthesis and subsequent sinking of silica frustules. In high-nutrient, low-chlorophyll (HNLC) regions, such as the equatorial Pacific, silicon limitation can restrict diatom growth despite abundant macronutrients like nitrate and phosphate, leading to reduced primary production and altered carbon export dynamics. Bacteria further mediate silicon cycling in marine environments by accelerating the dissolution of diatom silica frustules after cell death. Natural assemblages of marine bacteria can increase silica dissolution rates from lysed diatoms by up to threefold, recycling silicic acid back into the water column and making it available for new diatom growth. This microbial dissolution is crucial for maintaining silicon bioavailability in surface waters, where at least 50% of biogenic silica produced by diatoms dissolves in the upper 100 meters of the ocean. In HNLC regions, such processes can exacerbate silicon limitation when combined with iron scarcity, limiting diatom proliferation and the efficiency of CO₂ sequestration. On terrestrial landscapes, microbes contribute to silicon cycling by enhancing the weathering of silicate minerals, which releases dissolved silicic acid and other nutrients into soils and rivers. Soil bacteria and fungi produce organic acids and enzymes that accelerate silicate dissolution, accounting for a small but significant portion—estimated at 1-3%—of the global silicon flux from land to oceans. This microbial weathering supports nutrient availability for ecosystems and influences downstream silicon delivery. However, anthropogenic activities, such as the construction of dams, disrupt this flux by promoting diatom blooms in reservoirs that deplete dissolved silicon upstream, reducing its transport to coastal oceans by up to 30% globally and potentially suppressing marine diatom productivity. Recent research in 2025 has highlighted the role of soil microbes in modulating silicon bioavailability, particularly in saline or contaminated environments. Studies show that certain bacterial communities enhance silicon solubilization from soil minerals, improving its uptake by plants while mitigating sodium toxicity through microbial detoxification processes. For instance, rhizosphere microbes can increase reactive silica availability, fostering plant resilience to abiotic stresses and altering soil microbial composition to favor silicon-mobilizing taxa. These findings underscore microbes' potential in enhancing silicon cycling for sustainable agriculture and ecosystem health.

Safety and environmental impact

Health hazards and toxicity

Inhalation of crystalline silica dust, a common form encountered in occupational settings such as mining and construction, leads to silicosis, a progressive lung disease characterized by fibrosis and inflammation of lung tissue. This condition arises from the deposition of respirable particles in the alveoli, triggering an inflammatory response that results in scarring and reduced lung function. The International Agency for Research on Cancer (IARC) classifies inhaled crystalline silica in the form of quartz or cristobalite as a Group 1 carcinogen, indicating sufficient evidence of carcinogenicity in humans, primarily linked to lung cancer. Regulatory agencies have established strict exposure limits to mitigate these risks. The Occupational Safety and Health Administration (OSHA) sets a permissible exposure limit (PEL) of 50 μg/m³ (0.05 mg/m³) for respirable crystalline silica, calculated as an 8-hour time-weighted average. Similarly, the National Institute for Occupational Safety and Health (NIOSH) recommends a recommended exposure limit (REL) of 0.05 mg/m³ as a time-weighted average for up to a 10-hour workday. Acute exposure to silicon dust can cause irritation to the eyes and skin, manifesting as redness, dryness, or discomfort upon direct contact. Chronic exposure, however, poses more severe threats, including the development of chronic obstructive pulmonary disease (COPD) and an increased risk of lung cancer, with relative risks estimated between 1.2 and 2.0 in exposed populations. In contrast, amorphous silica is generally less hazardous than its crystalline counterpart. It is approved as a food additive (E551) for use as an anticaking agent, with the European Food Safety Authority (EFSA) concluding no safety concerns at levels up to 2% in food products across all population groups. Concerns regarding silica nanoparticles include potential cytotoxicity through mechanisms such as oxidative stress, inflammation, lysosomal dysfunction, and cellular uptake. For instance, 2025 research reaffirms the safety of synthetic amorphous silica nanoparticles in food applications, showing no adverse effects from oral exposure.

Environmental considerations and regulations

The production of silicon via the carbothermal reduction process generates substantial greenhouse gas emissions, with total emissions reaching 10-12 tons of CO₂ per ton of silicon produced, primarily from electricity consumption and direct process reactions. This process contributes approximately 1-2% to global CO₂-equivalent emissions when combined with related aluminum production activities. Silicon purification, particularly in the production of high-purity grades for electronics and photovoltaics, requires significant water input, estimated at 20-50 cubic meters per ton, depending on the refining method and regional practices. These processes often involve hydrofluoric acid (HF) for etching and cleaning, leading to wastewater contamination with fluoride compounds that necessitate specialized treatment to prevent environmental release. Waste management in silicon-related industries includes handling spent potliners from aluminum production, which contain silicon compounds and can be recycled into ferrosilicon or ferro-silicate slag through processes like leaching and smelting, reducing landfill disposal. Silicon mining, primarily of quartzite and silica sand, disrupts local ecosystems by clearing vegetation and fragmenting habitats, leading to biodiversity loss in affected areas such as riverbeds and forests. However, recycling silicon from end-of-life products like solar panels can reduce the demand for new mining operations, mitigating habitat impacts. Regulatory frameworks address these environmental concerns through targeted measures. In the European Union, the REACH regulation and the Carcinogens and Mutagens Directive establish an occupational exposure limit of 0.1 mg/m³ for respirable crystalline silica dust to minimize emissions and health risks during handling. The United States' Toxic Substances Control Act (TSCA) requires reporting and recordkeeping for silicon nanomaterials manufactured at the nanoscale, ensuring oversight of potential environmental releases. In China, 2025 mandates under the industrial green development plan require polysilicon producers to source 25-70% of their energy from renewables, aiming to curb emissions and promote sustainable production. Efforts toward sustainable silicon production, such as carbon capture in carbothermal processes and enhanced recycling, are gaining traction to further align with global decarbonization goals.