Fact-checked by Grok 2 weeks ago

Chromium

Chromium is a chemical element with the atomic number 24 and symbol Cr, classified as a transition metal in group 6 of the periodic table. It is a hard, lustrous, steel-gray solid that takes a high polish, exhibits high melting and boiling points, and demonstrates exceptional resistance to corrosion due to the formation of a stable oxide layer on its surface. With an atomic mass of 51.996 u, density of 7.19 g/cm³ at room temperature, electron configuration [Ar] 3d⁵ 4s¹, a melting point of 1907 °C, and a boiling point of 2671 °C, chromium is valued for its durability and versatility in industrial applications. Discovered in 1797 by French chemist through the analysis of the mineral (PbCrO₄), chromium derives its name from the Greek word chroma, meaning "color," owing to the vivid hues of its compounds, such as reds, greens, and yellows. The element occurs naturally in the at an average concentration of about 140 parts per million, primarily extracted from ore (FeCr₂O₄), with major production sites including , , , and . Approximately 90% of mined chromium is used in metallurgical processes, particularly to produce alloys like the common 18-8 grade (18% chromium, 8% ), which enhances strength, hardness, and resistance to oxidation and wear. Other significant uses include for decorative and protective coatings on metals, pigments in paints and ceramics (e.g., and green), catalysts in chemical reactions, and refractories in high-temperature furnaces. In its most stable , +3 (trivalent chromium or Cr(III)), the element plays an essential biological role as a , potentiating insulin action to facilitate , , and , with recommended daily intakes of 20–35 micrograms for adults. Chromium is poorly absorbed in the body (typically 0.5–2.5%), but deficiencies have been linked to impaired glucose tolerance and potential contributions to conditions like , though evidence for supplementation benefits remains mixed. Conversely, the hexavalent form (Cr(VI)) is highly and carcinogenic, particularly when inhaled as dust or fumes, causing respiratory issues, skin allergies, and increased risk; it is regulated under environmental standards due to its mobility and oxidizing properties. Overall, chromium's dual nature—as both a vital material and a with concerns—underscores its importance in modern technology and health sciences.

Physical properties

Atomic properties

Chromium () is a with 24 and of 51.9961 u. As a in group 6 of the periodic table, its atomic structure features 24 electrons arranged in the [Ar] ⁵ 4s¹, where the half-filled subshell contributes to its stability and unique bonding properties. The term symbol for neutral chromium is ⁷S₃, reflecting the high-spin arrangement of its five unpaired electrons. The first of chromium, required to remove the 4s , is 652.9 /, while the second , involving removal from the 3d subshell, is significantly higher at 1590.6 /; these values highlight the relative ease of losing the outer s compared to the more tightly bound d electrons. radii for chromium include a calculated value of 166 and an empirical value of 140 , with the measured at 139 ; these dimensions influence its ability to form compact metallic lattices.
PropertyValueUnit
Atomic number24-
Atomic weight51.9961u
Electron configuration[Ar] 3d⁵ 4s¹-
Ground state term symbol⁷S₃-
First ionization energy652.9kJ/mol
Second ionization energy1590.6kJ/mol
Calculated atomic radius166pm
Empirical atomic radius140pm
Covalent radius139pm
The high of 2180 K and of 2944 K for chromium arise from the robust strengthened by the delocalized electrons in its 3d⁵ 4s¹ configuration, enabling high thermal stability. In its solid metallic form, chromium adopts a body-centered cubic with a of 288.5 pm at , a arrangement that accommodates its atomic size and electron distribution for efficient packing.

Bulk properties

Chromium metal exhibits a density of 7.19 g/cm³ at 20 °C, making it a relatively dense transition metal suitable for applications requiring structural integrity. Its linear thermal expansion coefficient is 4.9 × 10^{-6} K^{-1}, indicating low dimensional change under temperature variations, which contributes to its stability in thermal environments. Mechanically, chromium is renowned for its exceptional hardness, with a Mohs hardness of 9.0 and Vickers hardness of 1060 MPa, positioning it among the hardest pure metals. It possesses a tensile strength of up to 415 MPa, though its ductility is limited; while malleable at elevated temperatures, it remains brittle at room temperature, often showing negligible elongation before fracture. Electrically, chromium has a resistivity of 124.9 nΩ·m at , reflecting moderate compared to other metals. Magnetically, it displays antiferromagnetic ordering below 38 °C (its Néel ) and transitions to above this threshold, a behavior stemming from its that leads to antiparallel spin alignment in the bulk . Thermally, chromium demonstrates a conductivity of 93.9 W/(m·K) and a of 448 J/(kg·K), values that underscore its efficiency in heat dissipation while requiring moderate energy for temperature changes. A key feature enhancing its durability is the formation of a thin surface passivation layer, consisting of Cr₂O₃ approximately 1 nm thick, which self-heals and acts as a barrier to further oxidation and corrosion in ambient conditions.

Isotopes

Chromium has four isotopes: ^{50}Cr, ^{52}Cr, ^{53}Cr, and ^{54}Cr. Their relative atomic masses are 49.94604183(94) u, 51.94050623(63) u, 52.94064815(62) u, and 53.93887916(61) u, respectively, with natural abundances of 4.345(13)%, 83.789(18)%, 9.501(17)%, and 2.365(7)%. The of chromium, 51.9961(6), is the abundance-weighted average of these isotopic masses. These stable isotopes are nuclides, formed through and incorporated into the solar system during its formation, as indicated by their consistent compositions in primitive meteorites. Among the approximately 20 known radioactive , ^{51}Cr is notable with a of 27.7010(11) days, decaying primarily via to ^{51}V; it is widely used as a tracer in medical applications, such as labeling red blood cells to assess survival and circulation. Another example is ^{48}Cr, with a of 21.56(3) hours, also decaying via to ^{48}V. The nuclear properties of key isotopes include zero spin for ^{50}Cr, ^{52}Cr, and ^{54}Cr, while ^{53}Cr has a nuclear spin of $3/2 and a magnetic moment of -0.47454 \mu_N. Neutron capture cross-sections for chromium isotopes vary significantly by energy and isotope; for instance, evaluations from thermal to 20 MeV energies are critical for applications in reactors, with ^{52}Cr showing resonance structures influencing capture rates up to several barns in the keV range. Chromium isotopes exhibit in geological processes, driven by reactions and mineral-fluid interactions, such as during subduction-zone or microbial , resulting in \delta^{53}Cr variations up to several per mil that trace Earth's oxygenation history.

Chemical properties and compounds

Oxidation states

Chromium exhibits a range of oxidation states from −2 to +6, though the most common and ones in compounds are +3 and +6, with +2 also notable in certain contexts. The elemental form represents the Cr(0) state, a d⁶ s⁰ in the metallic , which is under conditions. In the +2 state, chromium adopts a d⁴ , resulting in labile complexes that are prone to or further oxidation due to the high-spin nature in aqueous environments. The +3 state is the most overall, featuring a d³ that is particularly favored in octahedral coordination, where it forms inert complexes common in aqueous solutions. The +6 state, with a d⁰ , is strongly oxidizing and typically occurs in oxyanions such as chromate (CrO₄²⁻) or dichromate (Cr₂O₇²⁻). The of these states arises from electronic factors in coordination complexes. For Cr(III), the d³ configuration leads to a half-filled t₂g orbital set in octahedral fields, providing significant crystal field stabilization energy (CFSE) of −1.2 Δ₀ and kinetic inertness due to the lack of low-lying excited states for . This half-filled subshell enhances thermodynamic compared to neighboring states like d⁴ (Cr(II)) or d² (Cr(IV)), which lack such . In contrast, Cr(VI) is conferred by strong Cr–O multiple bonds in oxyanions, where the empty d orbitals facilitate π-backbonding, though these species are highly reactive toward . Redox behavior between these states is characterized by standard reduction potentials that reflect their relative stabilities. For instance, the Cr³⁺/Cr²⁺ has E° = −0.424 V, indicating that Cr(II) is a strong reductant and readily oxidizes to Cr(III) in aerobic conditions. Conversely, the reduction of dichromate to Cr(III) in acidic media, Cr₂O₇²⁻ + 14H⁺ + 6e⁻ → 2Cr³⁺ + 7H₂O, has E° = +1.33 V, underscoring the powerful oxidizing nature of Cr(VI). These potentials highlight the thermodynamic preference for Cr(III) as an intermediate state in cycles. Less common oxidation states include +1 (d⁵), +4 (d²), and +5 (d¹), which are unstable and typically require stabilizing ligands or solid-state matrices. Cr(I) compounds, such as certain organometallic complexes, are rare and obtained via reduction of lower states. Cr(IV) appears in oxides like CrO₂, a ferromagnetic material used in magnetic recording tapes. Cr(V), exemplified by the tetraperoxo complex [CrO₈]³⁻ in K₃CrO₈, often generated as an intermediate in Cr(VI) reductions. In general, chromium ions in various oxidation states form coordination complexes via the reaction Cr^{n+} + m L → [CrL_m]^{n+}, where L represents ligands such as water, halides, or oxo groups, and m is the coordination number (often 6 for octahedral geometry). The electronic configuration influences ligand field splitting, dictating reactivity, color, and magnetic properties across these states.

Key compounds

Chromium forms a variety of compounds across different oxidation states, with key examples illustrating its coordination chemistry and reactivity. In the zero oxidation state, chromium hexacarbonyl, Cr(CO)6, is a representative organometallic compound. It features an octahedral structure where the chromium atom is surrounded by six carbon monoxide ligands, enabling it to serve as a precursor in organochromium synthesis. This white crystalline solid is prepared by the direct reaction of chromium metal with carbon monoxide under high pressure, and it sublimes at room temperature but decomposes violently upon heating above 130 °C, incompatible with strong oxidants like chlorine. In the +2 oxidation state, chromium(II) chloride, CrCl2, exemplifies the reducing nature of lower-valent chromium species. This air-sensitive, steel-gray crystalline solid is synthesized by reducing chromium(III) chloride with hydrogen at 500–800 °C or by reacting chromium metal with anhydrous hydrogen chloride gas. It exhibits strong reducing properties, rapidly oxidizing in moist air to form chromium(III) compounds and liberating hydrogen gas when dissolved in water; its solubility is high in water but low in alcohols and ethers. The +3 oxidation state is the most stable for chromium, with chromium(III) oxide, Cr2O3, being a prominent example. This green, crystalline solid adopts a corundum-like structure and is amphoteric, dissolving in both strong acids to form chromic salts and strong bases to yield chromites. It is prepared by roasting chromium(III) hydroxide or by thermal decomposition of ammonium dichromate, resulting in a refractory material with a melting point of 2435 °C and negligible solubility in water. Hexavalent chromium compounds are powerful oxidants, often featuring tetrahedral or octahedral coordination. , Na2CrO4, is a , hygroscopic crystalline with a tetrahedral chromate anion (CrO42-), prepared by roasting ore with and lime, followed by leaching and crystallization. , K2Cr2O7, appears as orange-red triclinic crystals containing the dichromate ion (Cr2O72-), which consists of two edge-sharing tetrahedra and acts as a strong ; it is obtained by reacting with . , H2CrO4, exists primarily in as dark purplish-red crystals when concentrated, formed by dissolving in or , and it equilibrates with ions depending on : $2 \text{CrO}_4^{2-} + 2 \text{H}^+ \rightleftharpoons \text{Cr}_2\text{O}_7^{2-} + \text{H}_2\text{O} This equilibrium shifts toward the yellow tetrahedral chromate in basic conditions and the orange octahedral dichromate in acidic media. Among other notable compounds, chromyl chloride, CrO2Cl2, is a volatile, dark red fuming liquid in the +6 oxidation state, featuring a tetrahedral structure with two oxo and two chloro ligands. It is synthesized by treating chromium(VI) oxide with concentrated hydrochloric acid in the presence of sulfuric acid to remove water, and it hydrolyzes vigorously with water to produce chromic acid and hydrochloric acid, while decomposing in light to chromium(IV) oxide and chlorine.

Occurrence and production

Natural occurrence

Chromium is the 21st most abundant in , with an average concentration of 140 . This abundance places it among the more common transition metals, though it is significantly depleted in the oceanic environment, where concentrations average approximately 0.2–0.5 ppb. These low levels reflect chromium's siderophile and lithophile tendencies, limiting its mobility in aqueous systems relative to its crustal distribution. The primary ore mineral for chromium is , with the FeCr₂O₄, belonging to the and typically containing 46–49% Cr₂O₃. forms in ultramafic igneous rocks through magmatic , often appearing as layered seams or podiform masses in complexes. These deposits are enriched in chromium due to the mineral's stability in high-temperature, reducing conditions during mantle-derived . Major chromium deposits are concentrated in a few large layered intrusions. The Bushveld Complex in accounts for over 80% of global chromium reserves, hosting vast stratiform chromitite layers within its mafic-ultramafic sequence. Significant additional reserves occur in the Great Dyke of , a linear intrusion with high-grade chromite reefs, and the Stillwater Complex in , , which contains economically viable podiform and stratiform chromitite. Chromium appears in trace amounts in other minerals beyond chromite, such as (PbCrO₄), a rare lead chromate found in oxidized lead deposits, and chrome ochre, a hydrated form of Cr₂O₃ occurring in weathered ultramafic terrains. In extraterrestrial sources, chromium is notably enriched in chondrites, which exhibit high bulk concentrations and unique partitioning into sulfides under reduced conditions, while lunar basalts display relatively low chromium contents, typically around 2,000 ppm in Apollo samples, reflecting derivation from a depleted mantle source.

Extraction and production methods

Chromium is primarily produced industrially from ore, the main source of the metal, through processes that yield alloys or high-purity metal for specialized applications. The predominant commercial method involves carbothermic reduction in submerged furnaces to produce , an iron-chromium alloy typically containing 50–70% chromium. ore is mixed with as the reductant and fluxes such as to form , then smelted at temperatures of 1,600–2,000°C. The key reduction reaction simplifies to Cr₂O₃ + 2C → 2Cr + , occurring within the chromite matrix (FeCr₂O₄), with the process consuming 3,000–4,000 kWh per of . Global chromite ore mine production reached an estimated 47 million metric tons as of 2024, primarily for smelting, with as the leading producer at 21 million metric tons (45% share), followed by at 8 million metric tons (17% share) and at 6.5 million metric tons (14% share). output reached 18.51 million metric tons as of 2024, driven largely by demand for . For pure chromium metal, aluminothermic reduction is employed on oxidized or , reacting the ore with aluminum powder in a refractory-lined vessel at 2,000–2,200°C via FeCr₂O₄ + 2Al → + 2Cr + Al₂O₃, yielding a crude chromium-iron that is leached and separated. This batch process is exothermic and suitable for smaller-scale production. High-purity chromium (>99.9%) is produced electrolytically by dissolving (CrO₃) in a , such as a of chromium chloride and , and applying current to deposit chromium on a at elevated temperatures. This method achieves purities up to 99.99% and consumes approximately 10–15 kWh per kg of metal. Further refining for ultra-high purity may involve to remove volatile impurities under reduced pressure.

History

Discovery and isolation

In 1797, French chemist Louis Nicolas Vauquelin discovered the element chromium while examining samples of Siberian red lead, a vibrant orange mineral sourced from the Ural Mountains in Russia and later identified as crocoite (PbCrO₄). By dissolving the ore in hydrochloric acid to precipitate lead chloride and subsequently processing the solution, Vauquelin isolated chromium oxide (Cr₂O₃), recognizing it as a new element distinct from previously known substances. He presented his discovery to the French Academy of Sciences that same year, highlighting the oxide's remarkable properties in forming intensely colored compounds across various oxidation states. The name "chromium" derives from the Greek word chroma, meaning "color," reflecting these vivid hues observed in its derivatives, such as green emerald-like pigments. The following year, in 1798, Vauquelin reported the first isolation of metallic chromium by reducing the with (carbon) in a high-temperature , yielding an impure but form of the metal. This early process, while not producing high-purity chromium, represented a significant advancement in isolation techniques during the late . The discovery was closely linked to the mineralogical studies of lead-bearing ores prevalent in and at the time, where occurred as an accessory mineral in lead deposits, sparking interest among European chemists in and their oxides.

Early industrial developments

The commercialization of pigment, lead chromate (PbCrO₄), began in 1809 following its synthesis by French chemist Nicolas-Louis Vauquelin, who recognized its potential as a vibrant yellow alternative to traditional pigments like for paints and dyes. This pigment quickly gained traction in the art and industrial sectors due to its bright hue and opacity, marking one of the earliest widespread commercial applications of chromium compounds in the pigment industry. Early experiments in chromium emerged in the 1850s, with the first known electrodeposit achieved in 1855 by scientist Dr. B. Geuther using baths, laying the groundwork for later industrial processes that enhanced metal durability and corrosion resistance. Although practical commercialization of occurred in the 1920s, these initial developments in and highlighted chromium's potential for surface treatments in . In the late , advancements in alloys incorporating chromium for began to take shape, with English metallurgists and patenting an acid-resistant iron-chromium alloy containing 30-35% chromium and 2% in 1872. This was followed by significant progress in the , including metallurgist Philipp Monnartz's and patents on chromium-nickel (Cr-Ni) s, which demonstrated improved passivation and to atmospheric , as detailed in his 1911 publication on iron-chromium alloys. Monnartz's work, patented in in 1910 alongside William Borchers, established key compositions for what would become modern s. Concurrently, British metallurgist patented the first practical in 1913, featuring 12.8% chromium, which proved highly resistant to and was initially applied in and firearms. Post-1900 global trade in chromium shifted dramatically as supplies moved from Russian chromite—dominant until 1827—to emerging deposits in , (now ), and the , driven by rising industrial demand in and the . By the early 20th century, these regions accounted for the majority of exports, supporting the growth of production for alloy manufacturing. further accelerated demand, with chromium essential for high-temperature alloys in jet engines and military hardware, prompting the establishment of new plants in Allied nations to secure supplies amid shortages from occupied territories. This wartime surge, particularly for superalloys in , underscored chromium's strategic importance and fueled post-war industrial expansion.

Applications

Metallurgical applications

Chromium plays a pivotal role in metallurgical applications, primarily as an alloying in production to impart resistance, , and strength. The majority of chromium—approximately 90% of global consumption—is used in , where serves as the key intermediate to introduce the element into molten steel. This addition enhances properties such as hardenability, allowing for improved responses and overall durability in structural applications. In production, chromium content typically ranges from 10% to 30%, forming a passive layer on the surface that prevents and oxidation. Austenitic grades, like the widely used 18/8 variety (18% chromium and 8% ), offer excellent and are non-magnetic, making them suitable for equipment and architectural uses. Ferritic stainless steels contain about 17% chromium, providing magnetic properties and good formability for automotive exhaust systems, while martensitic types with 11–18% chromium achieve high hardness through , ideal for and valves. Approximately 180 kg of chromium is incorporated per tonne of stainless steel, depending on the grade. Beyond , chromium is essential in high-strength low-alloy steels, such as chromoly (e.g., AISI 4130 with 0.8–1.1% chromium and 0.15–0.25% ), which provides superior tensile strength and fatigue resistance for pipelines and components. In superalloys like 718 (17–21% chromium), it contributes to oxidation resistance at elevated temperatures, enabling use in blades and engines. is typically added during processes like the basic oxygen furnace for carbon steels or argon oxygen decarburization for stainless grades, optimizing chromium recovery and alloy uniformity while enhancing the steel's ability to form for greater toughness. Chrome plating, an process using a bath (Cr⁶⁺ reduced to metallic Cr⁰), deposits thin layers of 0.5–100 µm for wear resistance and low friction on tools, hydraulic cylinders, and engine parts. Hard chrome plating, in particular, yields a of up to 1000 , extending component life in demanding industrial environments, though it accounts for a smaller fraction of overall chromium use compared to bulk alloying.

Pigments and coatings

Chromium-based pigments have been widely used for their vibrant colors and durability in paints, coatings, and ceramics. , chemically lead chromate (PbCrO₄), is a bright pigment historically produced by the precipitation of lead nitrate solution with sodium chromate, followed by filtration and to achieve the desired particle size and hue. This compound provides excellent opacity and but has been largely phased out in many applications due to its . Similarly, chromate (ZnCrO₄), another pigment, appears as a yellow-green powder and is valued for its anticorrosive properties, particularly as a primer in protective coatings for metals, where it inhibits formation through the release of chromate ions. For green coloration, chromium(III) oxide (Cr₂O₃), often called chrome green or emerald green, serves as a key inorganic pigment due to its intense hue, chemical inertness, and exceptional thermal stability up to 1700°C, making it suitable for high-temperature applications. This pigment is synthesized by the reduction of sodium dichromate with sulfur or carbon at elevated temperatures and is extensively employed in ceramics for glazes, in paints for industrial and architectural coatings, and in plastics for color stability under harsh conditions. Its non-toxicity compared to hexavalent chromium compounds has contributed to its sustained popularity. In protective coatings, chromates play a role in surface treatments beyond pigmentation. Chromic acid anodizing of aluminum forms a thin, porous layer sealed with chromate solutions, enhancing resistance while preserving electrical , commonly used in and automotive parts. Decorative , involving of a thin chromium layer over , imparts a shiny, reflective finish for aesthetic appeal in consumer goods like automobile trim and household fixtures, also providing moderate wear resistance. The use of lead chromate pigments has significantly declined since the early 2000s due to their carcinogenic and toxic effects from and lead exposure, prompting regulatory actions such as the Union's REACH restrictions starting in 2007, with inclusion in the authorization list in 2012 (sunset date 2015), and ongoing authorizations challenged in court (e.g., 2019 ruling). Alternatives like have emerged, offering similar yellow tones with lower toxicity and comparable durability in paints and coatings. Globally, pigments account for approximately 2-3% of total chromium consumption, with annual production of Cr₂O₃ estimated at around 100,000 metric tons to meet demand in these sectors.

Other uses

Elemental chromium finds application as a in various industrial and dehydrogenation processes. Raney chromium, a porous form of the metal, serves as an efficient heterogeneous for reactions, particularly in the conversion of organic compounds such as nitriles to amines and in the production of intermediates for pharmaceuticals and chemicals. This is prepared by aluminum from a chromium-aluminum , resulting in high surface area that enhances reactivity under mild conditions. Additionally, chromia-alumina catalysts, consisting of supported on alumina, are widely used in the dehydrogenation of to styrene, a key step in production, operating at high temperatures around 600–650°C with selectivities exceeding 90%. In electronics, thin films of elemental chromium are employed as underlayers in hard disk drive media to promote epitaxial growth of magnetic cobalt alloys, improving signal-to-noise ratios and storage densities in longitudinal recording systems. These films, typically 10–50 nm thick, provide a textured surface that aligns magnetic domains, contributing to the reliability of devices. Chromium-based thin films also function as resistors in integrated circuits, offering stable resistance values up to several kiloohms per square due to their low of resistance and compatibility with processes. For optical applications, evaporated chromium films serve as reflective coatings on mirrors, achieving approximately 70% reflectivity at 700 in the , which makes them suitable for astronomical telescopes and precision instruments requiring durable, non-tarnishing surfaces. These coatings provide a balance of reflectivity and to substrates, often enhanced with protective overcoats to prevent oxidation. Chromium contributes to wood preservation through mixtures like chromated copper arsenate (CCA), where it acts as a fixative to bind copper and arsenic compounds into the wood matrix, protecting against fungal decay and insect damage in outdoor structures such as decks and utility poles. However, its use has significantly declined since the early 2000s due to environmental and health concerns over arsenic leaching, leading to regulatory restrictions by the U.S. Environmental Protection Agency that phased out CCA for most residential applications by 2004. In niche applications, elemental chromium is alloyed into dental prosthetics, comprising up to 20–30% of cobalt-chromium alloys to enhance resistance, , and mechanical strength for crowns, bridges, and partial . These alloys exhibit high and low in the oral , supporting long-term durability.

Applications of compounds

Chromium compounds find extensive applications in various industrial processes, leveraging their chemical reactivity and stability. One of the primary uses is in the leather tanning industry, where chromium(III) sulfate, often in the form of basic Cr₂(SO₄)₃, serves as a key tanning agent. This compound cross-links collagen fibers in animal hides, enhancing the leather's durability, water resistance, and hydrothermal stability by forming coordinate bonds with the protein structure. Approximately 80–90% of global leather production relies on trivalent chromium salts for this purpose, making it the dominant method due to its efficiency and cost-effectiveness compared to vegetable or synthetic alternatives. Chromates, particularly sodium chromate (Na₂CrO₄), are widely employed as inhibitors in industrial settings. In cooling systems for power plants and facilities, chromate-based formulations passivate metal surfaces, such as and , to prevent pitting and by forming a protective layer. These inhibitors are effective at concentrations as low as 100–500 and have been a standard in open recirculating systems since the mid-20th century, though their use is declining due to concerns. In applications, chromate pigments in primers, such as strontium chromate, provide sacrificial protection for aluminum alloys in fuselages and wings by releasing Cr(VI) ions that inhibit anodic reactions at scratches or defects. In photography, chromium compounds have historically enabled alternative printing techniques. Early processes utilized emulsions for direct positive images, where exposure to light decomposed the compound to produce metallic silver deposits. More contemporarily, is central to gum bichromate printing, a non-silver invented in the , in which the dichromate sensitizes a gum arabic and mixture, hardening exposed areas to create multi-layered color prints upon water development. This method remains popular among photographers for its archival quality and tonal range. Chromic acid (H₂CrO₄), derived from (CrO₃), has been used in wood preservatives, often as a component in formulations like (CCA) to protect timber from fungal decay and insect infestation by penetrating cell structures and fixing metallic ions. These treatments were common for utility poles, marine pilings, and outdoor structures until regulatory restrictions. In 2003, the U.S. Environmental Protection Agency facilitated the voluntary phase-out of CCA for most consumer residential uses, such as decks and playground equipment, due to arsenic leaching risks, though industrial applications persist under controlled conditions. In , (C₁₄H₁₄N₄O) is a for the colorimetric detection of (Cr(VI)) in environmental and industrial samples. The method involves acidifying the sample to 1–2, adding the to form a 1,5-diphenylcarbazide-Cr(VI) complex, and measuring absorbance at 540 nm using , with a around 0.1 µg/L. This technique, outlined in EPA Method 7196A, is valued for its specificity and simplicity in monitoring Cr(VI) contamination in and .

Biological role

Nutritional requirements

Chromium is recognized as an essential trace element in human nutrition, primarily due to its role in enhancing insulin action through the oligopeptide chromodulin, a low-molecular-weight chromium-binding factor that potentiates insulin receptor activation and improves glucose metabolism. This mechanism involves chromodulin binding chromic ions (Cr(III)) in response to insulin signaling, thereby amplifying downstream effects such as increased tyrosine kinase activity and glucose uptake in cells. However, the essentiality of chromium in humans remains controversial, with some authorities noting no clear deficiency symptoms in healthy populations and no reliable methods to assess status. The element's essentiality was established based on observations of metabolic disturbances in chromium-deficient animal models and human cases, underscoring its necessity for carbohydrate and lipid metabolism. Due to insufficient data to determine an Estimated Average Requirement (), no Recommended Dietary Allowance (RDA) has been established for chromium; instead, Adequate Intake () levels are used to guide nutritional needs. For adults aged 19–50 years, the AI is 35 µg/day for men and 25 µg/day for women, based on median dietary intakes observed in healthy populations. These values reflect estimates from the Institute of Medicine's 2001 Dietary Reference Intakes , which considered balance studies and typical U.S. consumption patterns to prevent adverse effects. Pregnant and lactating women have slightly higher AIs of 30 µg/day and 45 µg/day, respectively, to account for increased metabolic demands. Chromium deficiency is rare in individuals consuming a varied diet but has been documented in specific scenarios, such as long-term total parenteral nutrition (TPN) without chromium supplementation. In the 1970s, reports from patients on prolonged TPN revealed symptoms including severe glucose intolerance, peripheral neuropathy, elevated serum lipids, and unexplained weight loss, all of which were reversed upon chromium administration. These cases linked deficiency to impaired insulin sensitivity and abnormal lipid profiles, highlighting chromium's role in maintaining metabolic homeostasis. Although uncommon in oral nutrition, marginal deficiencies may contribute to glucose intolerance and hyperlipidemia in at-risk groups like those with diabetes. Dietary absorption of chromium is inefficient, typically ranging from 0.4% to 2.5% of ingested amounts, influenced by its and dietary factors. Absorption occurs primarily in the via passive diffusion for inorganic forms, with enhanced by ascorbic acid, which reduces Cr(III) and facilitates uptake. Conversely, phytates—found in grains and —can inhibit absorption by forming insoluble complexes with chromium and other minerals. These interactions emphasize the importance of balanced meal composition for optimal chromium utilization. As of 2025, the levels established by the remain unchanged, reflecting stable consensus on safe intake thresholds amid limited new data on requirements. Emerging research continues to investigate organic chromium complexes, such as chromium picolinate, for potentially improved bioavailability and metabolic effects in insulin-resistant populations.

Sources and supplementation

Dietary sources of chromium include a variety of foods, though levels can vary significantly due to soil content, growing conditions, and processing methods. Selected sources include: (1 cup): 7.5–8.0 µg; (½ cup): 1.1 µg; (1 whole): 1.8 µg; whole-wheat bread (1 slice): 1.0 µg; (3 oz): 2.0 µg; brewer's (1 tablespoon): 3.3 µg; (1 cup): 1.0 µg. Meat is a modest source, with offering around 2 µg per 3 oz serving. Brewer's is among the richer sources, though content varies by product. The of chromium from food sources is generally low, ranging from 0.4% to 2.5%, but it tends to be higher from products like compared to plant-based foods due to differences in chemical form and interactions with dietary components such as phytates in plants that can inhibit . Food can substantially reduce chromium content; for example, refining leads to significant losses of minerals including chromium, with studies indicating up to 70% reduction in certain micronutrients during milling as and —where minerals concentrate—are removed. Chromium supplements are available in several forms, with chromium picolinate (chemical formula C₁₈H₁₂CrN₃O₆) being the most common due to its purported enhanced absorption. Other forms include chromium chloride and chromium nicotinate. Typical doses in supplements range from 200 to 1,000 µg per day, often exceeding dietary needs, and are frequently included in multivitamins or targeted products. In the United States, the (FDA) has required nutrition labeling on most packaged foods since the Nutrition Labeling and Education Act of 1990, which initially established a (DV) for chromium at 120 µg based on older reference intakes. This was updated in 2016 to 35 µg for adults and children aged 4 and older to align with current on adequate intake. Foods providing 20% or more of the (7 µg or more) can be labeled as a "good source" of chromium. The global market for chromium supplements reached approximately $680 million in 2024, driven largely by inclusion in weight-loss and metabolic support products, reflecting consumer interest in its role in nutrient metabolism.

Health effects and research

Research on chromium's health effects began in the 1950s when Klaus Schwarz and colleagues identified trivalent chromium (Cr(III)) as an essential component of the glucose tolerance factor (GTF), a complex that enhances insulin action and improves glucose metabolism in animal models. This discovery highlighted chromium's role in potentiating insulin sensitivity, facilitating glucose uptake in cells, and maintaining carbohydrate metabolism. Subsequent studies confirmed that GTF, containing Cr(III) bound to nicotinic acid and amino acids, ameliorates impaired glucose tolerance induced by chromium deficiency. In the , the approved a in 2011 stating that chromium contributes to normal , based on evidence linking adequate intake to insulin function and energy utilization from carbohydrates, fats, and proteins. However, the U.S. has not approved chromium as a for any medical condition; it is regulated only as a with a qualified allowing limited statements on chromium picolinate and reduced risk of and , but without endorsement for treatment. Despite early promise, major health organizations have found insufficient evidence to support chromium supplementation for specific therapeutic uses. The recommends against chromium use for lowering blood glucose in people with , citing inconsistent results from clinical trials showing only minor or negligible improvements in glycemic control. A 2013 Cochrane of randomized controlled trials concluded that chromium picolinate supplementation does not significantly reduce body weight or body fat in or obese adults, with low-quality evidence indicating no meaningful effect over 6 months. Similarly, the International Society of Sports Nutrition's 2018 review update states that evidence for chromium as an ergogenic aid to enhance athletic performance or is lacking, as most studies show no benefits on strength, endurance, or muscle gains. Regarding cancer risks, hexavalent chromium (Cr(VI)) compounds are classified by the International Agency for Research on Cancer as Group 1 carcinogens, with sufficient evidence of causing lung cancer via inhalation in occupational settings due to their oxidative DNA damage. In contrast, Cr(III) supplements at low doses (up to 1,000 mcg daily) are considered safe, with no established upper intake level and rare adverse effects like mild gastrointestinal upset, as supported by extensive safety data from human trials. Recent research as of 2025 has explored Cr(III) supplementation for (PCOS), yielding mixed results; a 2025 of randomized trials found benefits in reducing and improving lipid profiles in some women with PCOS, but other studies report inconsistent effects on hormonal balance and ovarian function. Emerging investigations into nanotechnology-based delivery systems for Cr(III), such as encapsulation to enhance , show promise for targeted applications in glucose metabolism disorders like PCOS, though clinical trials remain preliminary and focused on improving without .

Safety and environmental considerations

Toxicity of chromium species

Chromium exhibits varying depending on its , with (Cr(VI)) being highly toxic and carcinogenic, while trivalent chromium (Cr(III)) demonstrates low toxicity at typical exposure levels. Cr(VI) is classified as a by the International Agency for Research on Cancer, primarily due to its ability to penetrate membranes and induce genotoxic effects, whereas Cr(III) is poorly absorbed and generally non-genotoxic. Hexavalent chromium enters cells through anion transporters such as the sulfate/phosphate system and is rapidly reduced intracellularly by agents like ascorbate and glutathione to lower valent intermediates (Cr(V) and Cr(IV)), generating reactive oxygen species (ROS) that cause oxidative stress. This reduction process leads to the formation of DNA adducts, including ternary Cr(III)-DNA-His and Cr(III)-DNA-Cys complexes, as well as DNA-protein crosslinks and double-strand breaks, which are key mechanisms in its carcinogenicity. Inhalation of Cr(VI) compounds, such as those in welding fumes or chromate production, is associated with lung and nasal cavity cancers, while ingestion can result in tumors of the oral cavity and small intestine. Trivalent chromium, in contrast, has limited , with only about 1% absorbed via or , and is primarily excreted in , making it far less toxic than Cr(VI). Although not genotoxic, high doses of Cr(III) can cause skin and , particularly in occupational settings like cement handling where soluble Cr(III) compounds may sensitize the . Exposure to chromium species occurs mainly through , , and dermal contact. is the primary route for Cr(VI) in industrial environments, such as or , while arises from contaminated , and dermal happens during activities like or handling treated materials. The (OSHA) sets a (PEL) of 5 µg/m³ for Cr(VI) as an 8-hour time-weighted average to mitigate respiratory risks. For , the Environmental Protection Agency (EPA) establishes a maximum contaminant level (MCL) of 100 µg/L for total chromium in . Acute effects of Cr(VI) include respiratory irritation manifesting as cough, dyspnea, and wheezing from inhalation, and severe gastrointestinal symptoms such as abdominal pain, vomiting, and hemorrhage following high-dose ingestion. Dermal contact with Cr(VI) can cause burns, ulceration, and non-allergic irritation. Chronic exposure to Cr(VI) heightens cancer risk and may lead to pulmonary sensitization or renal damage, while Cr(III) primarily induces allergic reactions like redness and swelling in sensitive individuals. For Cr(III), chronic high-dose exposure from supplements has been linked to potential rhabdomyolysis or anemia, though such cases are rare. Treatment for chromium toxicity focuses on supportive care, including , fluid and management, and symptomatic relief such as oxygen for respiratory distress. In acute Cr(VI) , interventions like ascorbic acid to enhance reduction or for severe cases have been employed, while with agents such as EDTA shows limited efficacy and is not routinely recommended.

Environmental impacts and regulations

Chromium pollution primarily originates from industrial activities, including tannery effluents containing trivalent chromium (Cr(III)), from ore processing, and waste streams rich in (Cr(VI)). Tanneries use in processing, releasing Cr(III)-contaminated that can leach into and if untreated. operations generate , which contain Cr(III) bound to iron oxides, leading to long-term near extraction sites. processes, involving baths, emit Cr(VI) mists and sludges, with uncontrolled facilities releasing approximately 90 grams of chromium per megagram of product plated. In aquatic environments, Cr(VI) exhibits high ecotoxicity due to its and , bioaccumulating in organisms with factors of 125–200 in bivalves and polychaetes, which disrupts food webs and reduces . Acute toxicity to shows 96-hour LC50 values ranging from 17,600 µg/L () to 249,000 µg/L (), though sensitive species experience effects at lower chronic concentrations around 0.29–44 µg/L, impairing growth and reproduction in like . In contrast, Cr(III) is less mobile, rapidly precipitating as hydroxides in neutral waters, limiting its but still posing risks in acidic conditions where LC50 for can reach 3,330 µg/L in soft . Remediation strategies target chromium's speciation and mobility, with phytoremediation using hyperaccumulators like , which uptake 48–58 µg of chromium per plant primarily in roots, converting Cr(VI) to less toxic Cr(III) forms bound to organic acids such as and . Microbial reduction employs bacteria like and species to enzymatically reduce Cr(VI) to insoluble Cr(III) under aerobic or anaerobic conditions, achieving up to 90% removal in contaminated soils and waters. These biological methods offer cost-effective alternatives to chemical treatments, enhancing site restoration while minimizing secondary pollution. Regulatory frameworks address chromium's environmental risks through emission controls and . Under the EU's REACH regulation, Annex XVII restricts Cr(VI) in specific articles such as (≥3 mg/kg since 2015). The 2025 ECHA proposal sets worker exposure limits of 0.1–5 µg/m³ for Cr(VI) depending on the process and environmental emission caps for uses like , aiming to prevent up to 17 tonnes of annual releases, with an 18-month transition period. In the US, the Toxic Substances Control Act (TSCA) requires reporting of Cr(VI) manufacturing, processing, and use volumes under the Chemical Data Reporting rule, enabling risk assessments for persistent releases. Globally, the Convention's technical guidelines promote environmentally sound management of wastes contaminated with Cr(VI), treating them as hazardous to curb transboundary pollution. A notable case study is the Hudson County Chromate Chemical Production Waste Sites in , USA, where over 2 million tons of chromate residue from historical and dye manufacturing contaminated soils and sediments near the estuary since the early 1900s. Cleanup efforts, overseen by the Department of since the 1980s, have remediated many residential areas and reduced exposure through landfilling and stabilization, with ongoing restoration of key sites as of 2025. Recent advancements in practices, including chromium from industrial wastes, offer energy savings of up to 70-80% compared to , aligning with sustainable resource loops and reducing demands.