Cadmium selenide (CdSe) is an inorganic binary compound of cadmium and selenium, recognized as a prototypical II-VI semiconductor with a direct bandgap of 1.74 eV at 300 K, enabling efficient light absorption and emission in the visible spectrum.[1][2] It occurs naturally as the rare mineral cadmoselite and is synthesized industrially through methods such as direct combination of elements or precipitation from solutions, typically yielding a grayish-black to red crystalline powder or lumps.[3][4] The material exhibits polymorphism, adopting either a hexagonal wurtzite structure or a cubic zincblende structure depending on preparation conditions, with key physical properties including a density of 5.81 g/cm³ and a melting point of 1258 °C.[5][6]As an n-type semiconductor, CdSe is prized for its tunable optical properties, particularly in nanoscale forms like quantum dots, where particle size controls emission wavelengths from green to red, facilitating applications in light-emitting diodes (LEDs), photovoltaics, and biomedical imaging.[6][1] Bulk CdSe also serves as a redpigment in artists' paints and ceramics due to its vibrant color and stability, though its use has declined owing to cadmium's toxicity.[2] Additionally, thin films of CdSe are employed in photoresistors, solar cells as absorber layers, and infrared-transparent windows for optical instruments.[6][2]Despite these advantages, CdSe poses significant health and environmental hazards as a cadmium-containing compound, classified as toxic by ingestion, inhalation, and skin contact, with potential for bioaccumulation and carcinogenicity; stringent regulations limit its production and disposal worldwide.[7][4]Research continues to explore safer alternatives and encapsulation techniques to mitigate risks while preserving its technological utility.[6]
Properties
Crystal structure
Cadmium selenide (CdSe) primarily adopts two polymorphic crystal structures: the wurtzite (hexagonal) form, which is thermodynamically stable at ambient conditions including room temperature, and the zincblende (cubic) form, which is metastable and typically synthesized under kinetic control such as lower-temperature or specific growth conditions.[8][9] In both structures, Cd and Se atoms are arranged in a tetrahedral coordination, characteristic of II-VI semiconductors, with the wurtzite phase belonging to the space group P6₃mc and the zincblende phase to F̅43m.[8]The lattice parameters for the wurtzite structure are a = 4.299 Å and c = 7.015 Å, yielding a c/a ratio of approximately 1.633 that reflects its close-packed hexagonal arrangement.[9] For the zincblende structure, the lattice parameter is a = 6.05 Å, corresponding to an isotropic cubic lattice.[8] These parameters determine the unit cell volume and atomic spacing, influencing phonon modes and thermal expansion behaviors in the material.In bulk CdSe, both polymorphs exhibit a direct bandgap of approximately 1.74 eV at room temperature, located at the Γ point of the Brillouin zone, which underpins its semiconductor properties for optoelectronic applications.[10] The crystal structure modulates the electronic band structure subtly: the cubic zincblende form has degenerate valence bands due to its higher symmetry, whereas the hexagonal wurtzite structure introduces crystal-field splitting of the top valence band (on the order of tens of meV), leading to anisotropic effective masses and slight differences in optical transitions.[11] This structural influence on band edges and carrier dynamics is foundational to understanding CdSe's behavior in devices, though the direct gaps remain similar across polymorphs.Regarding stability, the wurtzite phase is favored thermodynamically at standard pressures and temperatures, while the zincblende phase can be obtained under controlled synthesis, such as ion-exchange processes; the transition is reversible but kinetically hindered, preventing easy interconversion at ambient conditions without external stimuli like pressure or temperature gradients.[9] High-pressure studies further reveal that both can transform to a rocksalt phase above 2-3 GPa, but wurtzite recovers upon decompression, highlighting its robustness.[12]
Physical and optical properties
Cadmium selenide (CdSe) in its wurtzite form possesses a density of 5.81 g/cm³. The melting point is 1258 °C. CdSe is insoluble in water but soluble in acids such as hydrochloric acid.[7][13][9]The wurtzite crystal structure contributes to CdSe's direct bandgap of 1.74 eV at 300 K, influencing its optical absorption onset in the visible to near-infrared range. Bulk CdSe exhibits a refractive index of approximately 2.66 at 500 nm. The absorption coefficient is on the order of 10⁴ cm⁻¹ near the band edge (around 720 nm), facilitating efficient visible light harvesting. Photoluminescence in bulk CdSe occurs primarily near the band edge, emitting in the red to near-infrared spectrum under optical excitation.[9][14][15][16]Thermal properties of wurtzite CdSe are anisotropic due to its hexagonal symmetry. The coefficient of linear thermal expansion is 2.45 × 10⁻⁶ K⁻¹ parallel to the c-axis and 4.4 × 10⁻⁶ K⁻¹ perpendicular to the c-axis at 25 °C. The specific heat capacity at constant pressure follows Cp = 48.46 + 5.87 × 10⁻³ T – 58154 T⁻² J mol⁻¹ K⁻¹, yielding approximately 0.258 J g⁻¹ K⁻¹ at 300 K.[13][17][9]As an intrinsic n-type semiconductor in its undoped bulk form, CdSe has a low carrier concentration at room temperature, rising to 6 × 10¹³ cm⁻³ at 800 K due to thermal generation across the bandgap. Electron mobility in undoped bulk CdSe is 660 cm² V⁻¹ s⁻¹ at 300 K, while hole mobility is about 40 cm² V⁻¹ s⁻¹.[9][18]
Chemical properties
Cadmium selenide (CdSe) features a polar covalent bond between cadmium and selenium, characterized by partial ionic character. The electronegativity difference of 0.86 between Cd (1.69) and Se (2.55) results in approximately 16% ionic character, calculated using Pauling's formula: % ionic character = 100 × (1 - e^(-0.25 × (Δχ)^2)), where Δχ is the electronegativity difference. This mixed bonding nature contributes to CdSe's semiconductor properties while influencing its reactivity.[19]/07:_Chemical_Bonding_and_Molecular_Structure/7.01:_Ionic_and_Covalent_Bonding)CdSe demonstrates notable reactivity under thermal conditions. When heated in air, it oxidizes to cadmium oxide (CdO) and elemental selenium, as evidenced in studies of Cd-rich CdSe quantum dots where partial oxidation produces CdO alongside residual CdSe phases. This process highlights the compound's susceptibility to oxidative environments at elevated temperatures.[20]In terms of stability, CdSe is resistant to dilute acids and practically insoluble in water, but it decomposes in stronger acidic media. Exposure to concentrated nitric acid (HNO₃) or aqua regia leads to dissolution, releasing cadmium ions. Similarly, in hot concentrated hydrochloric acid (HCl), CdSe undergoes decomposition via the reaction:\text{CdSe} + 2\text{HCl} \rightarrow \text{CdCl}_2 + \text{H}_2\text{Se}This reaction indicates the compound's vulnerability to protonation and selenide displacement under aggressive acidic conditions.[21][22]CdSe also exhibits solubility in complexing agents like cyanide or thiosulfate solutions, where released Cd²⁺ ions form stable complexes such as the tetracyanocadmate(II) ion, [Cd(CN)₄]²⁻. This behavior facilitates its dissolution in ammoniacal or alkaline media containing excess cyanide, underscoring the role of coordination chemistry in its chemical transformations.
Synthesis
Industrial production
Cadmium selenide's industrial production emerged in the early 20th century, driven initially by demand for vibrant pigments in paints, ceramics, and glass coloration, with commercial cadmium red (a CdS-CdSe solid solution) available from 1919 onward.[23] For pigment-grade material, a common method involves the aqueous precipitation of cadmium salts, such as cadmium sulfate, with alkali selenosulfates (e.g., sodium selenosulfate) at 60–100°C, followed by hydrolysis and filtration to yield a precipitate of 95–100% purity suitable for calcination into the final pigment.[24]For semiconductor applications requiring bulk crystalline CdSe, the primary method is the high-pressure vertical Bridgman (HPVB) process, which synthesizes the compound via direct combination of cadmium and selenium elements in a sealed quartz crucible under inert argon atmosphere at elevated temperatures (typically initiating reaction around 700–800°C to generate vapors and form the melt, with subsequent growth at higher melt temperatures up to ~1350°C).[25] This technique prevents dissociation of the volatile compound under high pressure (20–100 atm) and enables directional solidification for large single-crystal boules. An alternative route for precursor synthesis involves reacting cadmium acetate or similar salts with hydrogen selenide gas in controlled conditions to form CdSe powder, which can then be consolidated.[6]Purification to semiconductor-grade purity (>99.99%) is achieved through vacuum distillation to remove volatile impurities or high-pressure vertical zone melting (HPVZM), where a molten zone is traversed along the ingot under argon to segregate contaminants.[25]Industrial batch processes, such as those using HPVB furnaces, typically yield up to several kilograms of material per run, such as crystal tapes up to approximately 1 kg, to meet demands in optoelectronics.[25]
Laboratory methods
Cadmium selenide (CdSe) is synthesized in laboratory settings through small-scale, precise techniques that prioritize purity and structural control for research applications, yielding typically milligrams to grams of material. These methods allow researchers to tailor the polymorph—wurtzite or zincblende—by varying reaction conditions such as temperature, pH, and precursor stoichiometry. Purity is routinely assessed using spectroscopic methods like UV-Vis absorption and X-raydiffraction to confirm phase and composition.[26]A widely used wet chemical approach involves precipitation from aqueous solutions of cadmium salts, such as cadmium chloride (CdCl₂) or cadmium sulfate (CdSO₄), and selenium sources like sodium selenosulfate (Na₂SeSO₃) or sodium selenite (Na₂SeO₃) reduced in situ.[27] The reaction proceeds under inert atmosphere at temperatures of 25–180°C, often in an autoclave for hydrothermal variants, with pH adjusted to 9–12 using agents like NaOH or hydrazine to promote nucleation and prevent aggregation.[27] Stoichiometric ratios near 1:1 for Cd:Se favor complete reaction, resulting in polycrystalline powders after centrifugation and washing; this method's simplicity enables rapid production but requires careful handling to minimize impurities from hydrolysis.[27] Recent advances include green synthesis methods at room temperature using safer precursors to minimize toxicity and environmental impact.[28]The organometallic route, a seminal high-temperature pyrolysis method, reacts dimethylcadmium (Cd(CH₃)₂) with trioctylphosphine selenide (TOPSe, prepared from elemental Se and trioctylphosphine) injected into a coordinating solvent like tri-n-octylphosphine oxide (TOPO).[26] Performed at 230–260°C under argon, the rapid injection induces burst nucleation, followed by controlled growth at slightly lower temperatures (e.g., 250°C) for 10–30 minutes, with precursor stoichiometry and dwell time dictating the wurtzite phase formation and monodispersity.[26] Yields reach up to several grams after precipitation with methanol and purification via size-exclusion chromatography, emphasizing optical quality verified by narrow emission linewidths.[26]Vapor-phase techniques like metalorganic chemical vapor deposition (MOCVD) deposit CdSe films or powders using volatile precursors such as dimethylcadmium adducts and selenium alkyls (e.g., diethylselenide) or H₂Se, transported in a hydrogen carrier gas.[29] Reactions occur at 300–500°C on substrates like GaAs under low pressure (10–100 Torr), with growth rates of 0.1–1 μm/h controlled by precursor flow ratios (Cd:Se ≈ 1:1) and temperature gradients to stabilize the hexagonal wurtzite structure.[29] This method adapts industrial direct synthesis principles to inert-gas laboratory reactors, producing epitaxial layers with high crystalline purity confirmed by reflection high-energy electron diffraction.[29]
Nanomaterials
CdSe nanoparticles
Cadmium selenide (CdSe) nanoparticles typically range in size from 2 to 20 nm, a scale at which quantum confinement effects begin to significantly modify their electronic and optical properties relative to bulk CdSe, which has a bandgap of approximately 1.74 eV.[30][31] In this regime, the nanoparticles exhibit size-dependent absorption and emission spectra, with smaller particles showing blue-shifted bandgaps due to enhanced confinement of excitons within the reduced volume.[26]Colloidal synthesis represents the primary route for producing CdSe nanoparticles, involving the injection of precursors into a heated solvent to nucleate and grow particles under controlled conditions. The classic hot-injection method, developed by Bawendi, Steigerwald, and Alivisatos in the early 1990s and recognized in the 2023 Nobel Prize in Chemistry for quantum dot discoveries, uses cadmium oxide or acetate and trioctylphosphine selenide (TOPSe) precursors in a mixture of trioctylphosphine oxide (TOPO) and hexadecylphosphonic acid (HDA) at temperatures around 300°C, yielding nearly monodisperse spheres with narrow size distributions.[26] More environmentally friendly alternatives employ oleic acid as a surfactant and stabilizer, often in conjunction with octadecene, to facilitate nucleation at lower temperatures (200–250°C) while maintaining high crystallinity and uniformity.[32] These colloidal approaches allow precise control over reaction parameters like temperature, precursor concentration, and injection rate to tune particle size and yield.Surface passivation is essential in CdSe nanoparticle synthesis to minimize defects, prevent aggregation, and inhibit oxidation, with ligands such as TOPO or oleic acid binding to the particle surface during growth. TOPO, a bulky phosphine oxide, forms a steric barrier that stabilizes the nanoparticles in nonpolar solvents, reducing interparticle interactions and preserving photoluminescence quantum yields above 50% for sizes around 3–5 nm.[33]Oleic acid, a carboxylic acidsurfactant, similarly caps the surface through carboxylate coordination, enabling dispersion in organic media and protecting against oxidative degradation by limiting access to atmospheric oxygen.[34]Morphological control of CdSe nanoparticles extends beyond spherical shapes, enabling the formation of rods, plates, or other anisotropic structures through seeded growth techniques. In seeded growth, pre-synthesized spherical CdSe seeds (2–4 nm) serve as templates for epitaxial deposition of additional CdSe layers, where the choice of surfactants and growthtemperature directs preferential elongation along specific crystallographic axes; for instance, a TOPO/HDA mixture at 250–280°C promotes rod formation with aspect ratios up to 1:20.[35] This method yields high-aspect-ratio nanorods or platelets while maintaining uniform diameters, facilitating applications requiring directional charge transport.[36]Despite these advances, CdSe nanoparticles face stability challenges, including photooxidation under ambient light and oxygen exposure, which degrades the surface and quenches luminescence, as well as aggregation in solution due to insufficient steric repulsion. Photooxidation proceeds via reactive oxygen species attacking selenium atoms, leading to up to 50% loss in quantum yield within hours of illumination for bare CdSe particles.[37] Aggregation is mitigated by optimizing ligand density and chain length, ensuring colloidal stability in solvents like toluene for months. Key strategies for enhancing overall stability include overcoating with a wider-bandgap ZnS shell (1–3 monolayers thick), which passivates the core surface, reduces photooxidation rates by over an order of magnitude, and improves environmental resistance without altering the confinement regime.[38][31]
Quantum dots
Cadmium selenide quantum dots (CdSe QDs) exhibit pronounced quantum confinement effects when their dimensions are reduced to the nanoscale, typically below 10 nm, leading to discrete energy levels and size-dependent electronic properties. In this regime, the confinement of charge carriers within the potential well of the nanocrystal increases the effective bandgap compared to bulk CdSe. The quantum confinement model is described by the effective mass approximation, where the bandgap energy E is given byE = E_{\text{bulk}} + \frac{\hbar^2 \pi^2}{2 r^2 \mu},with E_{\text{bulk}} as the bulk bandgap, r the radius of the quantum dot, \hbar the reduced Planck's constant, and \mu the reduced effective mass of the electron-hole pair. This formulation, originally derived by Louis Brus in 1984 for small semiconductor crystallites and recognized in the 2023 Nobel Prize in Chemistry, predicts a blueshift in the absorption and emission spectra as particle size decreases, enabling precise tuning of optical properties.[39]The tunable emission of CdSe QDs arises directly from this confinement, allowing photoluminescence (PL) wavelengths to span the visible spectrum: bulk CdSe emits in the red (~700 nm), while particles as small as 2 nm emit in the blue (~450-500 nm). This size-dependent emission is characterized by narrow full-width at half-maximum (FWHM) values of 20-40 nm, providing high color purity suitable for display and imaging applications. The monodispersity achieved in synthesis ensures consistent emission profiles across ensembles.[31]A key synthesis route for high-quality, monodisperse CdSe QDs is the hot-injection method, involving the rapid injection of a selenium precursor (e.g., Se dissolved in trioctylphosphine) into a hot solution of cadmium oxide (CdO) in coordinating solvents like trioctylphosphine oxide (TOPO) or oleic acid/octadecene mixtures at temperatures around 250-300°C. This kinetic approach separates nucleation and growth phases, yielding dots with size distributions less than 5%, and sizes controlled by reaction time and temperature.To improve stability and luminescence efficiency, CdSe QDs are often encapsulated in a wide-bandgap shell, such as ZnS, forming core-shell structures like CdSe/ZnS. The ZnS shell passivates surface trap states, reducing non-radiative recombination and boosting PL quantum yields up to 80% while maintaining size-tunable emission. These structures exhibit enhanced photostability under prolonged excitation.[31]Characterization of CdSe QDs relies on techniques like transmission electron microscopy (TEM) for direct size measurement and confirmation of core-shell morphology, revealing spherical particles with diameters from 2-6 nm. Photoluminescencespectroscopy assesses emission wavelength, quantum yield, and FWHM, correlating optical data with theoretical confinement models.[31]
Applications
Electronics and optoelectronics
Cadmium selenide (CdSe) serves as a key II-VI semiconductor material in electronics and optoelectronics due to its direct bandgap of approximately 1.74 eV at room temperature, enabling efficient absorption and emission in the visible spectrum.[2] This property makes CdSe suitable for applications requiring high photosensitivity and tunable optoelectronic responses, particularly in thin-film devices where its high absorption coefficient facilitates compact architectures. In photovoltaic cells, CdSe functions as an absorber layer in thin-film configurations, with reported power conversion efficiencies reaching 1.88% for devices fabricated via rapid thermal evaporation on fluorine-doped tin oxide substrates.[40] When integrated into tandem structures, such as CdSe/CdTe heterojunctions, CdSe enhances wide-bandgap absorption for the top cell, contributing to overall device performance; for instance, early studies on CdSe as a 1.7 eV absorber paired with lower-bandgap bottom cells like CIGS have demonstrated open-circuit voltages up to 350 mV and short-circuit current densities around 7 mA/cm² after optimization with CdS buffer layers and annealing.[41] Electrodeposition emerges as a scalable thin-film deposition technique for CdSe in these photovoltaic applications, involving cathodic reduction from aqueous solutions of CdSO₄ and SeO₂ on conductive substrates like FTO, yielding uniform films with low resistivity and high photosensitivity suitable for large-area fabrication.[42]In light-emitting diodes (LEDs) and lasers, CdSe quantum dots (QDs) play a prominent role in visible light emission, leveraging their size-tunable bandgap for precise color control in displays. CdSe/ZnS core-shell QDs, for example, enable full-color QLEDs with emissions spanning yellow, cyan, violet, and white, achieving high pixel resolution (≈10 μm, 1278 PPI) through patterned arrays fabricated via asymmetric wettability interfaces.[43] In micro-LEDs, CdSe QDs act as color converters, transforming blue light into narrow-band red and green emissions (FWHM ≈20–30 nm) with photoluminescence quantum yields exceeding 90%, covering over 90% of the Rec. 2020 color gamut and supporting brightness levels up to 100,000 nits for high-contrast displays.[44] These devices exhibit external quantum efficiencies around 20% for red and green QD-based electroluminescent LEDs, outperforming traditional phosphors in color purity and stability. For photodetectors, CdSe's direct bandgap imparts high responsivity in the UV-visible range, with nanostructured thin films demonstrating values up to 4.9 A/W and detectivity of 7.9 × 10¹¹ Jones under visible illumination, attributed to efficient carrier generation and low dark current.[45] Annealed CdSe films further optimize these metrics, achieving detectivity up to 1.27 × 10¹⁰ Jones and response times as low as 12 ms.[46]Performance in CdSe-based devices is underpinned by carrier dynamics, including minority carrier lifetimes of approximately 0.1 μs in bulk and thin-film CdSe, which support effective charge transport.[47] In alloyed systems like Cd(Se)Te used in optoelectronic layers, lifetimes extend to hundreds of nanoseconds, enhancing diffusion lengths and enabling minority carrier collection over distances critical for photovoltaic and photodetector efficiency. Electrodeposited CdSe films exhibit carrier concentrations influenced by doping (e.g., Cu), with improved lifetimes and diffusion lengths due to reduced defect densities, facilitating rapid charge separation in devices.[48] These metrics underscore CdSe's viability in high-responsivity photodetectors and efficient emitters, where long diffusion lengths (on the order of microns in optimized films) minimize recombination losses.[42]
Pigments and other uses
Cadmium selenide (CdSe), often combined with cadmium sulfide as cadmium sulfoselenide (Cd(S,Se)), serves as a key pigment for producing vibrant red-orange hues in ceramics, glazes, and paints, owing to its inherent optical properties that absorb in the blue-green spectrum.[49][50] These pigments are valued for their thermal stability, maintaining color integrity in applications subjected to temperatures up to 300°C, such as high-firing ceramic enamels and specialty coatings.[51][52]Historically, CdSe-based pigments emerged in the mid-19th century as part of the cadmium color family, with cadmium red—derived from CdS1-xSex—patented around 1892 and commercialized by 1910 for artists' materials, replacing toxic vermilion (mercuric sulfide) in oil and acrylic paints.[23][53][54] This development enabled intense, lightfast reds in fine art and industrial finishes, with European Union sales of cadmium paints alone reaching approximately 40 tonnes annually as of 2018.[55]Beyond pigments, CdSe finds application in phosphors for fluorescent lamps, where it acts as a color-converting dye to produce warm tones, such as in gold or yellow-tinted tubes.[56] In biomedical contexts, bulk CdSe particles have been explored as non-quantum imaging tags for cellular labeling and tissue staining, leveraging their fluorescence for in vitro visualization.[57]Global production of CdSe and related cadmium pigments for non-electronic uses, primarily ceramics and paints, is estimated at 10–100 tonnes per year, a decline from a 1975 peak of 6,000 tonnes due to regulatory pressures.[55] In response to environmental and health regulations, such as the EU's REACH restrictions, industry has shifted toward less toxic alternatives, including mixed-oxide perovskites like CaTaO2N and LaTaON2 solid solutions, which offer comparable yellow-red shades without heavy metals.[58]
Occurrence and safety
Natural occurrence
Cadmium selenide occurs naturally as the rare mineral cadmoselite (CdSe), which crystallizes in the hexagonal wurtzite structure and typically appears as black to pale gray opaque grains or disseminations.[59] This mineral forms primarily under reducing, alkaline diagenetic conditions in sedimentary strata, such as sandstones, where it acts as a cementing agent in interstitial spaces.[60] Cadmoselite was first described in 1957 from the Ust'Uyuk vanadium-selenium-uranium deposit in the Turan District of Tuva, Russia, its type locality.[60]Known occurrences of cadmoselite are limited and scattered, including sites in Russia (such as the Kondoberezhskaya deposit in Karelia and the Kudriavy volcano on Iturup Island), the Czech Republic (Příbram uranium and base-metal district), Argentina (Sierra de Cacho in La Rioja Province), Canada (Ontario), and Jordan.[59] It is often associated with sphalerite (ZnS) in zinc-bearing deposits, where trace amounts of CdSe may substitute within the lattice or occur as minor phases, particularly in selenium-enriched environments like those linked to uranium or base-metal ores.[59] Examples include sphalerite-rich zones in sedimentary-hosted zinc deposits, though pure cadmoselite grains remain exceptional.Due to its extreme rarity, cadmoselite contributes negligibly to global cadmium resources, representing far less than 1% of total cadmium production, which is predominantly derived from cadmium impurities in sphalerite during zinc refining processes.[61]Cadmium from these natural selenide sources is not extracted as pure CdSe but co-recovered alongside zinc and other metals in roasting-leaching-electrowinning operations at zinc smelters.[62]
Health and environmental hazards
Cadmium selenide (CdSe) presents substantial health hazards primarily attributable to its cadmium component, which can release toxic Cd²⁺ ions upon degradation or exposure. Cadmium and its compounds, including those derived from CdSe, are classified by the International Agency for Research on Cancer (IARC) as Group 1 carcinogens, carcinogenic to humans, based on sufficient evidence of lung cancer from occupational inhalation exposures.[63] The carcinogenicity stems from Cd²⁺ ions interfering with DNA repair and inducing oxidative stress, leading to chronic effects such as kidney damage and prostate cancer with prolonged exposure.[64] Acute inhalation of CdSe dust or fumes can cause pneumonitis, pulmonary edema, and severe respiratory distress, as observed in cadmium compound exposures during industrial handling.[65]Human exposure to CdSe occurs predominantly through inhalation of fine dust generated in production and processing facilities, as well as via dermal absorption or incidental ingestion during handling.[7]Leaching of cadmium from CdSe-based pigments in paints or coatings can also contribute to indirect exposure through skincontact or environmental contamination.[66] For acute oral toxicity, cadmium compounds exhibit an LD50 of approximately 200 mg/kg in rats, reflecting moderate lethality that underscores the need for careful ingestion prevention.[67]Environmentally, CdSe contributes to cadmium pollution, which bioaccumulates in aquatic organisms such as fish and invertebrates, facilitating biomagnification through the food web and posing risks to higher trophic levels including humans.[68] In soil, CdSe nanoparticles degrade over time, releasing cadmium ions that bind strongly to organic matter and clay, resulting in persistence with a half-life exceeding 10 years and limited natural attenuation.[69] This longevity exacerbates soil contamination from industrial effluents or discarded electronics, inhibiting microbial activity and plant growth while enabling uptake into crops.[70]Regulatory frameworks aim to curb these risks through exposure limits and usage restrictions. The U.S. Occupational Safety and Health Administration (OSHA) establishes a permissible exposure limit (PEL) of 5 µg/m³ for airborne cadmium, applicable to CdSe operations to prevent respiratory and systemic effects.[71] In the European Union, the REACH regulation restricts cadmium selenide in consumer products, including electrical and electronic equipment, prohibiting concentrations ≥0.01% by weight to minimize environmental release and human contact; however, as of May 2024, the RoHS Directive includes exemptions for CdSe in downshifting quantum dots for display lighting applications, extended pending further review.[72][73]Risk mitigation strategies for CdSe, particularly in quantum dot formulations, involve encapsulation within polymer shells or silica coatings to inhibit Cd²⁺ ion release and reduce bioavailability. Such encapsulation has demonstrated significant reduction in cytotoxicity toward mammalian cells and aquatic species compared to uncoated particles.[74] Despite CdSe's low water solubility, which curtails immediate leaching in aqueous environments, long-term degradation necessitates these protective measures.[3]