Fact-checked by Grok 2 weeks ago

Cadmium selenide

Cadmium selenide (CdSe) is an inorganic binary compound of and , recognized as a prototypical II-VI with a bandgap of 1.74 eV at 300 K, enabling efficient light absorption and emission in the . It occurs naturally as the rare cadmoselite and is synthesized industrially through methods such as combination of or from solutions, typically yielding a grayish-black to crystalline powder or lumps. The material exhibits polymorphism, adopting either a hexagonal structure or a cubic zincblende structure depending on preparation conditions, with key physical properties including a of 5.81 g/cm³ and a of 1258 °C. 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 , facilitating applications in light-emitting diodes (LEDs), , and biomedical imaging. Bulk CdSe also serves as a in artists' paints and ceramics due to its vibrant color and stability, though its use has declined owing to cadmium's . Additionally, thin films of CdSe are employed in photoresistors, cells as absorber layers, and infrared-transparent windows for optical instruments. 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 and carcinogenicity; stringent regulations limit its production and disposal worldwide. continues to explore safer alternatives and encapsulation techniques to mitigate risks while preserving its technological utility.

Properties

Crystal structure

Cadmium selenide (CdSe) primarily adopts two polymorphic crystal structures: the (hexagonal) form, which is thermodynamically stable at ambient conditions including , and the zincblende (cubic) form, which is metastable and typically synthesized under kinetic control such as lower-temperature or specific growth conditions. 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 P6₃mc and the zincblende phase to F̅43m. The parameters for the structure are a = 4.299 and c = 7.015 , yielding a c/a ratio of approximately 1.633 that reflects its close-packed hexagonal arrangement. For the zincblende structure, the parameter is a = 6.05 , corresponding to an isotropic cubic . These parameters determine the unit cell volume and atomic spacing, influencing phonon modes and behaviors in the material. In bulk CdSe, both polymorphs exhibit a direct bandgap of approximately 1.74 at , located at the Γ point of the , which underpins its properties for optoelectronic applications. The modulates the subtly: the cubic zincblende form has degenerate valence bands due to its higher symmetry, whereas the hexagonal 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. 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 phase is favored thermodynamically at standard and , while the zincblende phase can be obtained under controlled , such as ion-exchange processes; the is reversible but kinetically hindered, preventing easy interconversion at ambient conditions without external stimuli like or gradients. 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.

Physical and optical properties

Cadmium selenide (CdSe) in its form possesses a of 5.81 g/cm³. The is 1258 °C. CdSe is insoluble in but soluble in acids such as . The crystal 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 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. in bulk CdSe occurs primarily near the band edge, emitting in the red to near-infrared spectrum under optical excitation. Thermal properties of CdSe are anisotropic due to its hexagonal symmetry. The coefficient of linear 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 at constant 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. As an intrinsic n-type in its undoped bulk form, CdSe has a low carrier concentration at , rising to 6 × 10¹³ cm⁻³ at 800 due to generation across the bandgap. in undoped bulk CdSe is 660 cm² V⁻¹ s⁻¹ at 300 , while hole mobility is about 40 cm² V⁻¹ s⁻¹.

Chemical properties

Cadmium selenide (CdSe) features a polar between and , characterized by partial ionic character. The difference of 0.86 between (1.69) and (2.55) results in approximately % ionic character, calculated using Pauling's formula: % ionic character = 100 × (1 - e^(-0.25 × (Δχ)^2)), where Δχ is the difference. This mixed bonding nature contributes to CdSe's properties while influencing its reactivity./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 (CdO) and elemental , as evidenced in studies of Cd-rich CdSe quantum dots where produces CdO alongside residual CdSe phases. This process highlights the compound's susceptibility to oxidative environments at elevated temperatures. 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 (HNO₃) or leads to dissolution, releasing cadmium ions. Similarly, in hot concentrated (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 and selenide displacement under aggressive acidic conditions. CdSe also exhibits solubility in complexing agents like or 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 , underscoring the role of coordination chemistry in its chemical transformations.

Synthesis

Industrial production

selenide's industrial production emerged in the early , driven initially by demand for vibrant pigments in paints, ceramics, and coloration, with commercial cadmium red (a CdS-CdSe ) available from onward. For pigment-grade material, a common method involves the aqueous precipitation of salts, such as , with alkali selenosulfates (e.g., sodium selenosulfate) at 60–100°C, followed by and filtration to yield a precipitate of 95–100% purity suitable for into the final . 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). 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. Purification to semiconductor-grade purity (>99.99%) is achieved through to remove volatile impurities or high-pressure vertical (HPVZM), where a molten zone is traversed along the under to segregate contaminants. 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 .

Laboratory methods

Cadmium selenide (CdSe) is synthesized in settings through small-scale, precise techniques that prioritize purity and structural control for 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 . Purity is routinely assessed using spectroscopic methods like UV-Vis absorption and to confirm and . A widely used wet chemical approach involves from aqueous solutions of cadmium salts, such as cadmium chloride (CdCl₂) or cadmium sulfate (CdSO₄), and selenium sources like sodium selenosulfate (Na₂SeSO₃) or (Na₂SeO₃) reduced . The reaction proceeds under inert atmosphere at temperatures of 25–180°C, often in an for hydrothermal variants, with adjusted to 9–12 using agents like NaOH or to promote and prevent aggregation. Stoichiometric ratios near 1:1 for Cd:Se favor complete reaction, resulting in polycrystalline powders after and ; this method's simplicity enables rapid production but requires careful handling to minimize impurities from . Recent advances include green synthesis methods at using safer precursors to minimize toxicity and environmental impact. The organometallic route, a seminal high-temperature method, reacts (Cd(CH₃)₂) with trioctylphosphine (TOPSe, prepared from elemental and trioctylphosphine) injected into a coordinating like tri-n-octylphosphine (TOPO). Performed at 230–260°C under , the rapid injection induces burst , followed by controlled growth at slightly lower temperatures (e.g., 250°C) for 10–30 minutes, with precursor and dictating the phase formation and monodispersity. Yields reach up to several grams after precipitation with and purification via , emphasizing optical quality verified by narrow emission linewidths. Vapor-phase techniques like metalorganic chemical vapor deposition (MOCVD) deposit CdSe films or powders using volatile precursors such as adducts and selenium alkyls (e.g., diethylselenide) or H₂Se, transported in a carrier gas. Reactions occur at 300–500°C on substrates like GaAs under low pressure (10–100 ), 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 structure. This method adapts industrial direct synthesis principles to inert-gas laboratory reactors, producing epitaxial layers with high crystalline purity confirmed by reflection high-energy .

Nanomaterials

CdSe nanoparticles

Cadmium selenide (CdSe) nanoparticles typically range in size from 2 to 20 , a scale at which quantum confinement effects begin to significantly modify their electronic and relative to CdSe, which has a bandgap of approximately 1.74 . In this regime, the nanoparticles exhibit size-dependent and spectra, with smaller particles showing blue-shifted bandgaps due to enhanced confinement of excitons within the reduced volume. 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 and recognized in the 2023 for discoveries, uses or acetate and trioctylphosphine selenide (TOPSe) precursors in a of trioctylphosphine oxide (TOPO) and hexadecylphosphonic acid (HDA) at temperatures around 300°C, yielding nearly monodisperse spheres with narrow size distributions. More environmentally friendly alternatives employ as a and , often in conjunction with octadecene, to facilitate at lower temperatures (200–250°C) while maintaining high crystallinity and uniformity. 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 binding to the particle surface during growth. TOPO, a bulky , forms a steric barrier that stabilizes the nanoparticles in nonpolar solvents, reducing interparticle interactions and preserving quantum yields above 50% for sizes around 3–5 nm. , a , similarly caps the surface through coordination, enabling in media and protecting against oxidative by limiting access to atmospheric oxygen. Morphological control of CdSe nanoparticles extends beyond spherical shapes, enabling the formation of , plates, or other anisotropic structures through seeded techniques. In seeded , pre-synthesized spherical CdSe seeds (2–4 nm) serve as templates for epitaxial deposition of additional CdSe layers, where the choice of and directs preferential along specific crystallographic axes; for instance, a TOPO/HDA at 250–280°C promotes rod formation with aspect ratios up to 1:20. This method yields high-aspect-ratio nanorods or platelets while maintaining uniform diameters, facilitating applications requiring directional charge transport. Despite these advances, CdSe nanoparticles face stability challenges, including photooxidation under ambient light and oxygen exposure, which degrades the surface and quenches , as well as aggregation in solution due to insufficient steric repulsion. Photooxidation proceeds via attacking atoms, leading to up to 50% loss in within hours of illumination for bare CdSe particles. Aggregation is mitigated by optimizing density and chain length, ensuring colloidal stability in solvents like 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 , and improves environmental resistance without altering the confinement regime.

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 by E = 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. The tunable emission of CdSe QDs arises directly from this confinement, allowing (PL) wavelengths to span the : bulk CdSe emits in the (~700 ), while particles as small as 2 emit in the (~450-500 ). This size-dependent emission is characterized by narrow full-width at half-maximum (FWHM) values of 20-40 , providing high color purity suitable for and applications. The monodispersity achieved in ensures consistent emission profiles across ensembles. A key synthesis route for high-quality, monodisperse CdSe QDs is the hot-injection method, involving the rapid injection of a precursor (e.g., Se dissolved in trioctylphosphine) into a hot solution of (CdO) in coordinating solvents like trioctylphosphine oxide (TOPO) or /octadecene mixtures at temperatures around 250-300°C. This kinetic approach separates and phases, yielding dots with size distributions less than 5%, and sizes controlled by reaction time and temperature. To improve stability and efficiency, CdSe QDs are often encapsulated in a wide-bandgap , such as ZnS, forming core- structures like CdSe/ZnS. The ZnS passivates surface states, reducing non-radiative recombination and boosting quantum yields up to 80% while maintaining size-tunable emission. These structures exhibit enhanced photostability under prolonged excitation. Characterization of CdSe QDs relies on techniques like (TEM) for direct size measurement and confirmation of core- morphology, revealing spherical particles with diameters from 2-6 nm. assesses emission wavelength, , and FWHM, correlating optical data with theoretical confinement models.

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. 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. 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. 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. 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 ) through patterned arrays fabricated via asymmetric wettability interfaces. In micro-LEDs, CdSe QDs act as color converters, transforming into narrow-band and emissions (FWHM ≈20–30 nm) with quantum yields exceeding 90%, covering over 90% of the color gamut and supporting brightness levels up to 100,000 nits for high-contrast displays. These devices exhibit external quantum efficiencies around 20% for and QD-based electroluminescent LEDs, outperforming traditional phosphors in color purity and . For photodetectors, CdSe's direct bandgap imparts high 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. Annealed CdSe films further optimize these metrics, achieving detectivity up to 1.27 × 10¹⁰ Jones and response times as low as 12 ms. 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. 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 efficiency. Electrodeposited CdSe films exhibit carrier concentrations influenced by doping (e.g., ), with improved lifetimes and diffusion lengths due to reduced defect densities, facilitating rapid charge separation in devices. These metrics underscore CdSe's viability in high-responsivity and efficient emitters, where long diffusion lengths (on the order of microns in optimized films) minimize recombination losses.

Pigments and other uses

Cadmium selenide (CdSe), often combined with as cadmium sulfoselenide (Cd(S,Se)), serves as a key for producing vibrant red-orange hues in s, s, and paints, owing to its inherent that absorb in the spectrum. These s are valued for their thermal stability, maintaining color integrity in applications subjected to temperatures up to 300°C, such as high-firing enamels and specialty coatings. 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. 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. Beyond pigments, CdSe finds application in phosphors for fluorescent lamps, where it acts as a color-converting to produce warm tones, such as in gold or yellow-tinted tubes. In biomedical contexts, bulk CdSe particles have been explored as non-quantum imaging tags for cellular labeling and tissue staining, leveraging their for visualization. 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. 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 .

Occurrence and safety

Natural occurrence

Cadmium selenide occurs naturally as the rare mineral cadmoselite (CdSe), which crystallizes in the hexagonal structure and typically appears as black to pale gray opaque grains or disseminations. 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. Cadmoselite was first described in 1957 from the Ust'Uyuk vanadium-selenium-uranium deposit in the Turan District of , , its type locality. Known occurrences of cadmoselite are limited and scattered, including sites in (such as the Kondoberezhskaya deposit in and the Kudriavy volcano on Island), the (Příbram uranium and base-metal district), (Sierra de Cacho in La Rioja Province), (Ontario), and . It is often associated with (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 or base-metal ores. Examples include -rich zones in sedimentary-hosted zinc deposits, though pure cadmoselite grains remain exceptional. Due to its extreme rarity, cadmoselite contributes negligibly to global resources, representing far less than 1% of total production, which is predominantly derived from impurities in during refining processes. from these natural selenide sources is not extracted as pure CdSe but co-recovered alongside and other metals in roasting-leaching-electrowinning operations at smelters.

Health and environmental hazards

Cadmium selenide (CdSe) presents substantial health hazards primarily attributable to its component, which can release toxic Cd²⁺ ions upon degradation or exposure. and its compounds, including those derived from CdSe, are classified by the International Agency for Research on Cancer (IARC) as carcinogens, carcinogenic to humans, based on sufficient evidence of from occupational exposures. The carcinogenicity stems from Cd²⁺ ions interfering with and inducing , leading to chronic effects such as kidney damage and with prolonged exposure. Acute of CdSe dust or fumes can cause , , and severe respiratory distress, as observed in compound exposures during industrial handling. Human exposure to CdSe occurs predominantly through of fine generated in and processing facilities, as well as via dermal absorption or incidental during handling. of from CdSe-based pigments in paints or coatings can also contribute to indirect exposure through or environmental . For acute oral toxicity, compounds exhibit an LD50 of approximately 200 mg/kg in rats, reflecting moderate lethality that underscores the need for careful prevention. Environmentally, CdSe contributes to pollution, which bioaccumulates in organisms such as and , facilitating through the and posing risks to higher trophic levels including humans. In soil, CdSe nanoparticles degrade over time, releasing ions that bind strongly to and clay, resulting in persistence with a exceeding 10 years and limited natural attenuation. This longevity exacerbates from industrial effluents or discarded electronics, inhibiting microbial activity and plant growth while enabling uptake into crops. Regulatory frameworks aim to curb these risks through exposure limits and usage restrictions. The U.S. (OSHA) establishes a (PEL) of 5 µg/m³ for airborne , applicable to CdSe operations to prevent respiratory and systemic effects. In the , 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 Directive includes exemptions for CdSe in downshifting quantum dots for display lighting applications, extended pending further review. 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 . Such encapsulation has demonstrated significant reduction in toward mammalian cells and aquatic species compared to uncoated particles. Despite CdSe's low water solubility, which curtails immediate in aqueous environments, long-term necessitates these protective measures.