Fact-checked by Grok 2 weeks ago

Quantum dot

A quantum dot is a nanoscale particle, typically 2–10 nanometers in diameter, whose electronic and are governed by quantum confinement effects, resulting in a size-tunable bandgap that enables precise control over light absorption and emission across the visible and spectra. These zero-dimensional nanostructures, composed of materials such as (CdSe), (InP), or (PbS), behave as artificial atoms, with their energy levels quantized due to spatial confinement of electrons and holes within the particle's dimensions. First observed in the late by Aleksey Ekimov in matrices and theoretically explained by Brus in the , quantum dots gained practical synthesis methods in the early through Moungi Bawendi's colloidal techniques, earning the trio the 2023 for their foundational contributions to . Key properties of quantum dots include high quantum yields (often exceeding 90% in core-shell structures), narrow emission linewidths ( around 20–40 nm), and exceptional photostability compared to traditional dyes, making them brighter and more durable for and applications. Their solution-processability allows for facile integration into films, inks, or devices via methods like spin-coating or , while surface passivation with shells (e.g., ZnS) mitigates and enhances stability against environmental degradation. typically involves colloidal routes, such as hot-injection or continuous-flow reactors, enabling monodisperse populations with precise size control by varying reaction temperature, time, or precursor ratios. Quantum dots have revolutionized fields like optoelectronics, biomedicine, and energy harvesting; in displays, they serve as color converters in QLED televisions, achieving over 100% Rec. 2020 color gamut coverage and external quantum efficiencies surpassing 20%. In photovoltaics, their multiple exciton generation potential boosts solar cell efficiencies beyond the Shockley-Queisser limit, with power conversion efficiencies reaching 18.3% in perovskite quantum dot solar cells as of October 2025. Biomedically, biocompatible variants (e.g., carbon or silicon dots) enable multiplexed imaging, targeted drug delivery, and biosensing, though challenges like heavy-metal toxicity in cadmium-based dots necessitate greener alternatives for clinical translation. Emerging uses span quantum computing for spin-based qubits and photocatalysis for sustainable fuel production, underscoring their versatility in advancing quantum technologies.

Fundamentals

Definition and Characteristics

Quantum dots are zero-dimensional nanocrystals, typically with diameters in the range of 2–10 nm, that exhibit pronounced quantum mechanical effects due to and confinement within their limited spatial dimensions. These nanoscale particles, often referred to as "artificial atoms," display discrete energy levels rather than the continuous bands found in bulk materials, arising from the quantization of energy states in all three dimensions. Key characteristics of quantum dots include their high surface-to-volume , which influences their reactivity and , and a tunable bandgap that enables size-dependent optical and electronic properties. For instance, in (CdSe) quantum dots, increasing the diameter from approximately 2 nm to 6 nm shifts the emission color from blue (around 450–495 nm) to red (around 620–750 nm) due to the varying degree of quantum confinement. This tunability stems from the inverse relationship between particle size and confinement energy, allowing precise control over wavelengths across the . Common compositions for quantum dots include II–VI semiconductors such as CdSe and , III–V semiconductors like InP and GaAs, and IV–VI materials such as PbS, which provide diverse bandgap energies suitable for various applications. Alternatives to these inorganic semiconductors encompass carbon-based quantum dots and quantum dots, which offer and lower toxicity while retaining quantum confinement effects. In contrast to bulk semiconductors, where charge carriers move freely within continuous valence and conduction bands, quantum dots lose this extended band structure at the nanoscale, resulting in atomic-like discrete states that enhance their utility in and .

Quantum Confinement

Quantum confinement refers to the spatial restriction of charge carriers, specifically electrons and holes, within a material on the nanoscale, typically in three dimensions for quantum dots (QDs). This confinement arises when the dimensions of the QD are comparable to or smaller than the de Broglie of the carriers, leading to quantization of their levels akin to particles in a . The phenomenon is described using the effective mass approximation, where carriers are treated as having effective masses m_e^* and m_h^* within the , and the confining potential is modeled as an infinite spherical well for simplicity. This approximation captures the transition from bulk-like continuous bands to discrete atomic-like levels, fundamentally altering the electronic structure of QDs. The size dependence of the electronic properties is a hallmark of quantum confinement, most notably manifested in the widening of the bandgap energy as the QD radius r decreases. In the strong confinement regime, the bandgap energy E_g can be approximated by the relation E_g = E_{g,\text{bulk}} + \frac{\hbar^2 \pi^2}{2 r^2} \left( \frac{1}{m_e^*} + \frac{1}{m_h^*} \right)^{-1}, where E_{g,\text{bulk}} is the bulk bandgap, \hbar is the reduced Planck's constant, and the second term represents the kinetic energy contribution from single-particle confinement of electrons and holes. This parabolic $1/r^2 scaling arises from solving the Schrödinger equation for independent carriers in a spherical box, predicting a blueshift in the absorption and emission spectra with decreasing size—for instance, CdSe QDs exhibit emission wavelengths tunable from near-infrared to ultraviolet as radii vary from ~5 nm to ~2 nm. Quantum confinement manifests in three distinct regimes, delineated by the relationship between the QD radius r and the Bohr radius a_B, which is the scale of the electron-hole pair (typically 5–20 for common semiconductors like CdSe or InP). In the strong confinement regime (r < a_B), the confinement energy dominates over the Coulomb attraction between electron and hole, treating them as uncorrelated particles and yielding discrete energy levels primarily from single-particle quantization. The intermediate regime (r \approx a_B) features comparable confinement and Coulomb energies, requiring variational methods to account for partial correlation, resulting in modified spectral shifts. In the weak confinement regime (r > a_B), the behaves as a quasi-particle with the center-of-mass motion quantized, while the internal structure resembles the , leading to smaller energy shifts proportional to $1/r rather than $1/r^2. These regimes influence the optical spectra: strong confinement produces sharp, size-dependent peaks with pronounced blueshifts, while weak confinement yields broader features closer to . The binding , the required to dissociate the electron-hole pair, is enhanced in QDs due to reduced screening compared to the material. In larger semiconductors, the high constant \epsilon (often 10–12) effectively screens the interaction, yielding modest binding energies of ~10–50 meV. However, in QDs, the finite size and surface effects lower the effective \epsilon, particularly in confinement where the wavefunctions extend near the boundary, diminishing screening and increasing the binding to 100–300 meV for materials like CdSe. This enhancement stabilizes , sharpening lines and boosting quantum yields, though it is perturbative relative to the dominant confinement in small QDs.

Structures

Core/Shell Architectures

Core/shell architectures in quantum dots consist of a central core of one material coated with a of a different , designed to passivate surface defects and enhance confinement. For instance, a CdSe core is commonly encapsulated by a ZnS , where the wider-bandgap ZnS reduces non-radiative recombination at the surface by providing electronic passivation, thereby improving overall luminescence efficiency. This heterostructure confines both electrons and holes primarily within the core, leveraging quantum confinement effects unique to the nanoscale dimensions. These structures are classified into Type I and Type II based on alignment. In Type I core/ quantum dots, such as CdSe/ZnS, the has a larger bandgap than the core, confining both charge carriers to the core region and resulting in bright, core-like with enhanced stability. Conversely, Type II configurations, exemplified by CdTe/CdSe, feature offsets where the conduction of the core is above that of the (or vice versa for the ), promoting spatial separation of electrons and holes across the core- to facilitate charge transfer and applications requiring prolonged lifetimes. This separation in Type II systems allows energies below the bandgaps of the individual materials, tunable via thickness. To further mitigate surface-related issues, double-shell designs incorporate an intermediate layer, such as CdSe/ZnSe/ZnS, where the inner ZnSe provides matching to reduce at the core , and the outer ZnS offers additional passivation. These configurations significantly suppress photoblinking by minimizing recombination and can achieve quantum yields exceeding 90% through optimized shell thicknesses. Fabrication of core/shell quantum dots often encounters challenges from lattice mismatch between core and shell materials, inducing strain that can lead to defects or dislocations at the . For example, the ~12% mismatch between CdSe and ZnS promotes compressive strain in the core, potentially degrading performance unless mitigated. Graded interfaces, achieved by gradually varying the shell composition, alleviate this strain by distributing it across a transitional layer, enabling thicker, more stable shells without compromising structural integrity. Such architectures are typically realized through colloidal methods.

Alloyed and Doped Variants

Alloyed quantum dots represent a class of nanostructures where two or more semiconductor materials are combined within a single lattice to form ternary or quaternary compositions, enabling precise control over electronic and optical properties. For instance, CdSe_{1-x}S_x alloys allow for continuous tuning of the bandgap energy by varying the sulfur content x, which shifts the emission wavelength without altering the particle size, thus maintaining consistent quantum confinement effects. This compositional flexibility extends to quaternary alloys, such as those based on copper-indium-zinc-sulfide systems, which further broaden the tunable emission range while preserving structural uniformity. In contrast, doped quantum dots incorporate intentional impurities at low concentrations to modify intrinsic properties, often introducing localized luminescent centers or enhancing charge transport. A prominent example is Mn-doped ZnSe, where manganese ions (\mathrm{Mn}^{2+}) serve as radiative recombination sites, producing characteristic emission from the ^4T_1 \to ^6A_1 d-d with lifetimes in the millisecond range, distinct from the host's band-edge . Typical doping levels range from 0.1% to 5% atomic fraction relative to the host lattice, balancing enhanced functionality with minimal disruption to the overall . These variants offer significant advantages, including reduced toxicity through strategies like partial replacement of in alloys (e.g., ZnSe compositions with lower content) and expanded emission tunability for applications requiring specific colors. For example, alloyed InP/ZnSe structures achieve high-efficiency green emission around 530 nm while being cadmium-free, supporting environmentally friendlier alternatives to traditional II-VI dots. However, challenges persist, such as phase in alloys due to thermodynamic instabilities that can lead to compositional inhomogeneities and broadened emission linewidths, as well as difficulties in achieving uniform distribution to avoid clustering or surface . Alloyed and doped quantum dots can also be integrated into core/shell architectures to synergistically combine lattice mixing with passivation effects.

Synthesis

Colloidal Methods

Colloidal methods represent the predominant approach for synthesizing quantum dots (QDs) through wet-chemical routes in , enabling the of high-quality, monodisperse nanocrystals with tunable . These techniques leverage solution-phase reactions to control and growth, typically yielding QDs in the 1-10 nm size range suitable for quantum confinement effects. Unlike gas-phase or lithographic methods, colloidal occurs at relatively mild temperatures (200-350°C) in coordinating solvents, facilitating scalability and versatility for various materials such as CdSe, , and InP. The core process involves rapid followed by controlled in coordinating solvents, with the hot-injection technique being the most widely adopted for achieving narrow size distributions. In this method, a precursor solution is swiftly injected into a heated solvent, such as trioctylphosphine oxide (TOPO) for CdSe QDs, promoting burst nucleation due to the sudden . Subsequent proceeds at a slightly lower , allowing to refine particle sizes while maintaining monodispersity, often with size variations below 5%. This approach, pioneered for II-VI semiconductors, has been extended to III-V materials like InP, producing QDs with diameters as small as 2 nm. Precursors vary by synthesis environment: organometallic compounds, such as (CdMe₂) for and trioctylphosphine selenide for , are common in non-aqueous hot-injection routes, reacting in high-boiling solvents like TOPO or octadecene. In contrast, aqueous methods employ greener precursors, including acetate or chloride salts paired with chalcogenide sources like for sulfides, enabling room-temperature or hydrothermal reactions in with stabilizers such as mercaptoacetic acid. control in both variants is achieved primarily through reaction time and ; shorter times or lower temperatures favor smaller QDs (e.g., 2-4 nm after 10 seconds at 280°C for CdSe), while extended growth yields larger particles up to 8 nm. Post-synthesis processing often includes ligand exchange to enhance solubility and stability for specific applications. Initially capped with long-chain ligands like in non-polar solvents, QDs undergo phase transfer by replacing these with shorter thiols (e.g., 1,2-ethanedithiol or mercaptopropionic acid), improving aqueous dispersibility while preserving quantum yields above 50%. These batch processes typically achieve yields of up to several grams per reaction, sufficient for laboratory-scale production of materials like CdSe with high purity. For industrial scalability, continuous flow reactors have emerged as a key advancement, replacing batch hot-injection with automated, steady-state processes that enhance reproducibility and throughput. In these systems, precursors are mixed via microfluidic channels or tubular reactors, enabling gram-scale hourly production of QDs like with quantum yields exceeding 80%, while minimizing solvent use and waste. This transition supports applications in , including brief adaptations for core/shell architectures during growth. A recent innovation (as of 2024) involves using superheated molten inorganic salts, such as , as solvents for colloidal of III-V QDs, enabling access to previously challenging compositions at high temperatures, with improved efficiency and reduced reliance on solvents.

Advanced Fabrication Techniques

Plasma synthesis represents a physical for producing quantum dots through gas-phase , offering high purity and compatibility with dry processing workflows. In this approach, techniques such as of bulk precursors like ZnS generate ultrafine particles by vaporizing material in an inert atmosphere, followed by rapid condensation into nanocrystals. For instance, ablation of silicon targets in helium gas yields oxide-passivated Si quantum dots approximately 2.5 nm in diameter, with exciton-based at 810 nm and quantum yields of 3–5%, free from solution-phase contaminants. These advantages stem from the absence of , enabling scalable production of biocompatible dots suitable for applications requiring minimal impurities. Lithographic fabrication provides precise control over quantum dot positioning, contrasting with the nature of colloidal synthesis. Electron-beam lithography patterns substrates like GaAs/AlGaAs to define narrow pillars or arrays containing quantum-confined structures, achieving diameters as small as 10 nm through and deposition steps. Self-assembled quantum dots, meanwhile, emerge via the Stranski-Krastanov growth mode in , where lattice mismatch induces strain-driven islanding of InAs on GaAs substrates at temperatures of 400–530°C and coverages of 1.6–1.7 monolayers. This epitaxial process produces dislocation-free, type-I confined dots with densities up to 10^{10} cm^{-2} and tunable emission from near-infrared to visible wavelengths. Viral and electrochemical assembly methods utilize biological templates and potential-driven deposition to achieve oriented or layered quantum dot structures. Virus capsids from the , engineered with peptide fusions on pVIII coat proteins (e.g., J140 for affinity), template the nucleation of quantum dots by incubating phages in CdCl_2 and Na_2S solutions at low temperatures, forming 3–5 nm nanocrystals aligned along the direction into nanowire arrays. This biomimetic approach ensures epitaxial orientation and enables heterostructure formation with dual peptides, leveraging the virus's for scalable, phase-controlled synthesis. Electrochemical assembly, such as , deposits quantum dots layer-by-layer within porous templates like anodized alumina; for , sequential reduction of Cd^{2+} and reaction with S^{2-} fills pores to controlled depths, yielding uniform arrays with narrow size distributions. The hybrid electrochemical/chemical variant oxidizes pre-deposited metals before sulfide displacement, promoting epitaxial growth and tunable at lower cost than . Bulk methods facilitate large-scale quantum dot production through solid-state or solution-to-solid transformations, though often at the expense of size uniformity. Mechanochemical milling grinds precursors like CdCO_3 and S in a , inducing solid-state reactions to form quantum dots with average sizes of 3–4 nm, as confirmed by showing hexagonal phase. This solvent-free process supports kilogram-scale output without high temperatures, ideal for industrial applications despite broader polydispersity. Sol-gel involves and of metal alkoxides (e.g., for TiO_2 or ZnO dots), transitioning sol phases to gels that yield quantum-confined particles upon , providing a versatile route for oxide-based dots in bulk quantities. Compared to colloidal routes, these methods emphasize throughput over monodispersity, suiting less demanding structural needs.

Properties

Optical Behavior

Quantum dots (QDs) exhibit tunable absorption and emission spectra across the to near-infrared range, arising from quantum confinement that discretizes electronic states. Absorption features sharp ic peaks, with the first energy varying from approximately 2.0 to 2.7 (620–460 ) depending on size in materials like InP. occurs at longer wavelengths due to a , typically tens of meV, resulting from excitonic relaxation and fine-structure splitting. The emission linewidths are notably narrow, with (FWHM) values of 20–80 meV at for monodisperse samples, enabling high color purity. In passivated QDs, such as core-shell InP variants, quantum yields reach 50–80%, approaching near-unity in optimized structures. A key phenomenon in single-QD emission is fluorescence blinking, characterized by random intermittency between "on" and "off" states under continuous excitation. This arises from photoinduced charging, where carriers are trapped at or defects, leading to recombination and non-radiative decay. Blinking reduces photon output and complicates single-particle tracking, but it can be suppressed through shell engineering; higher-bandgap shells passivate traps, block charge transfer, and increase the "on" time fraction to over 95%. Recent mechanisms highlight and band-edge carrier dynamics, with suppression strategies like optimization further stabilizing emission. QDs also demonstrate nonlinear optical effects, including multi-photon , where size-tuned confinement enhances processes like three- and four-photon in CdSe structures under intense . This property supports applications in deep-tissue by enabling at longer wavelengths. In anisotropic QDs, such as rod-shaped variants, polarization-dependent responses emerge, with nonlinear susceptibilities varying based on orientation due to shape-induced asymmetry. Advances in carbon QDs have extended into the near-infrared, with second near-infrared (NIR-II) variants achieving wavelengths from 1000 to 1700 nm through defect engineering and formation. For instance, nitrogen-vacancy polarons in oxidized carbon QDs yield peaks at 1035 nm and at 840 nm under 808-nm , boosting utility in luminescent materials for biomedical theranostics. As of July 2025, techniques enabling from dark excitons in QDs have been developed, allowing access to otherwise optically inactive states and improving efficiency for advanced applications.

Electronic and Structural Properties

Quantum dots (QDs) exhibit a quantized structure due to three-dimensional spatial confinement, resulting in levels for electrons and holes rather than continuous found in bulk semiconductors. This quantization leads to a series of well-defined excitonic transitions, where the spacing between levels scales inversely with the QD , enabling size-tunable properties. The effective mass of charge carriers in QDs is modified by the confinement potential, often approximated using effective mass theory, which accounts for interactions and mismatches at interfaces to predict level spacing and carrier behavior. Furthermore, the (DOS) in QDs transitions from a parabolic in bulk materials to a series of delta-function-like peaks corresponding to the levels, influencing carrier statistics and transport characteristics. In single QDs, charge transport is dominated by the Coulomb blockade effect, where the addition of an requires overcoming the due to electrostatic repulsion, leading to quantized conductance steps observable at low temperatures. This phenomenon manifests as periodic oscillations in current-voltage characteristics, with the blockade region width determined by the QD and single-particle level spacing. In contrast, QD films exhibit hopping conduction as the primary transport mechanism, where carriers between localized states in adjacent QDs, facilitated by interdot and to reduce barriers. As interdot distance decreases or strengthens, the regime shifts from insulating to semiconducting hopping, with activation reflecting the disorder in the film. Structurally, QDs often display to minimize energy, involving atomic rearrangements or ligand-induced passivation that alters facet stability and electronic properties. For instance, in II-VI QDs like CdSe, vacancy formation on surfaces leads to reconstructions that delocalize frontier orbitals and widen the band gap, enhancing stability without introducing mid-gap traps. Crystal phase differences, such as (hexagonal) versus blende (cubic), result in distinct facet morphologies— favors {11̅20} and {10̅10} planes with zigzag anion arrangements, while blende exposes {110} facets—impacting and . Defects like cation or anion vacancies play a critical role in doping, as they can act as shallow donors or acceptors, compensating intentional dopants and modulating carrier concentration, though in materials like , such vacancies often tolerate without deep trap states. Scanning tunneling microscopy (STM) and spectroscopy (STS) provide atomic-scale insights into QD electronic structure by mapping the local (LDOS), revealing discrete peaks corresponding to confined levels and their spatial variation across facets or aggregates. In PbSe QDs, for example, spectra show quantized LDOS resonances that broaden in molecular aggregates due to coupling, confirming delocalization effects. Recent advances in heavy-metal-free InP QDs have achieved electron mobilities of 0.45 cm²/V·s through optimized surface chemistry, enabling improved charge in optoelectronic devices without toxic elements.

Theoretical Framework

Quantum Mechanical Models

The particle-in-a-box model serves as the simplest quantum mechanical framework for describing carrier confinement in quantum dots, treating the nanocrystal as an infinite that quantizes the energy levels of electrons and holes. In the one-dimensional infinite square well approximation, the energy eigenvalues are given by E_n = \frac{\hbar^2 \pi^2 n^2}{2 m^* L^2}, where \hbar is the reduced Planck's constant, n is the principal , m^* is the effective mass of the carrier, and L is the confinement length. For three-dimensional spherical quantum dots, the model extends to solutions of the radial involving spherical , yielding size-tunable discrete energy levels that explain the observed blueshift in and spectra with decreasing dot radius. This idealization assumes abrupt infinite barriers, which overestimates confinement energies, but it provides essential insight into the quantum confinement regime where binding energies exceed bulk values. Extensions to finite potential wells address the limitations of the infinite well by incorporating barrier penetration, leading to modified wavefunctions and energy levels solved via boundary matching or variational methods. In these models, the effective confinement potential is finite, allowing evanescent tails into the surrounding matrix, which reduces the bandgap shift compared to the infinite case and better aligns with experimental confinement effects in colloidal dots. More realistic band structures are captured by the tight-binding method, which discretizes the into orbitals and computes hopping integrals to derive and conduction edges in quantum dots. This approach accounts for multi-orbital overlaps and , enabling predictions of and optical transitions in materials like InAs or CdSe. Complementarily, the k·p method perturbs the bulk structure around high- points (e.g., Γ-point) to model confined states, incorporating effective masses, strain, and piezoelectric fields for accurate band-edge energies in or zincblende dots. Within these frameworks, treats doping effects by adding impurity potentials that shift local energy levels and introduce mid-gap states, quantifying how donor or acceptor concentrations alter carrier densities and recombination rates. Density functional theory (DFT) provides atomistic simulations of core/shell interfaces in quantum dots, resolving atomic-scale strain, charge transfer, and potential barriers that arise from lattice mismatch between core and shell materials like CdSe/ZnS. These calculations reveal how interface dipoles and alloying modify the electronic landscape, enhancing stability against recombination. DFT further elucidates by computing electron-phonon elements, showing how longitudinal optical phonons broaden spectral lines and facilitate non-radiative relaxation, with coupling strengths scaling inversely with dot size in perovskites and II-VI semiconductors. Advanced quantum mechanical treatments employ (TDDFT) to simulate real-time dynamics, capturing ultrafast processes such as charge separation and coherent interactions following photoexcitation. In TDDFT, time of the Kohn-Sham orbitals under external perturbations yields spectra and relaxation pathways, revealing how quantum confinement suppresses multiphoton processes in small dots. Recent 2024 DFT models for quantum dots, functionalized with biomolecules like , simulate hybrid structures for healthcare applications, predicting tunable electronic properties for and bioimaging by analyzing edge states and binding energies at interfaces.

Semiclassical and Classical Approaches

Semiclassical approaches to quantum dot systems approximate dynamics in ensembles or devices by extending classical drift-diffusion equations with quantum , such as density-gradient terms that account for tunneling and compressibility effects without solving the full . These models are essential when treating large-scale QD arrays, where quantum mechanical simulations become computationally infeasible. For example, the quantum-corrected Poisson-drift-diffusion framework has been applied to analyze transport and recombination in InAs/GaAs quantum dot lasers grown on , revealing how quantum confinement influences currents and profiles. In quantum dot films, effective medium theory provides a homogenized description of the , averaging the response of dispersed QDs within a host matrix to predict bulk . The Maxwell-Garnett formulation is widely used for this purpose, yielding the effective as \epsilon_{\rm eff} = \epsilon_m + 3f \frac{\epsilon_{\rm QD} - \epsilon_m}{\epsilon_{\rm QD} + 2\epsilon_m - f(\epsilon_{\rm QD} - \epsilon_m)}, where \epsilon_m is the matrix , \epsilon_{\rm QD} is the QD (often size-dependent due to confinement), and f is the QD . This approach, combined with tight-binding calculations for individual QD responses, enables efficient modeling of and in colloidal QD thin films for optoelectronic assessments. Classical models apply to larger quantum dots, where confinement energies are small relative to thermal scales, allowing treatment as spheres. governs electromagnetic scattering in this regime, providing exact solutions to for spherical particles with sizes comparable to the incident wavelength, typically QDs exceeding 10 nm. Scattering cross-sections computed via scale with the fourth power of radius in the limit but transition to for even larger particles, aiding design of light extraction in QD-based LEDs. To model fluorescence intermittency or in quantum dots, classical rate equations describe of excitonic states, emphasizing non-radiative recombination as the dominant mechanism. The rate for charged trions follows k_{\rm Auger, X^\pm} \propto R^{c} (with c \approx 3 for volume scaling and R the effective radius), while the biexciton rate is k_{\rm Auger, XX} = 2(k_{\rm Auger, X^+} + k_{\rm Auger, X^-}); these yield power-law distributions of on/off times by coupling to charging/discharging processes. Engineering shell thickness in core/shell QDs tunes these rates by up to an , suppressing for stable emission. Multiscale modeling bridges these approximations by employing for the QD core—capturing confinement and electronic states—while treating ligands and solvent classically via or continuum . This hybrid strategy simulates in photocatalytic applications, such as charge separation and transfer in QD-sensitized systems, where partitions the reactive region quantumly and the extended environment classically to predict reaction rates without excessive computational cost. Such semiclassical and classical methods are limited to QDs larger than approximately 10 , beyond which quantum confinement weakens and bulk-like behavior emerges, rendering full quantum treatments unnecessary or invalid for ensemble averages.

Applications

Quantum dots (QDs) have revolutionized optoelectronics by enabling tunable emission and detection across visible and near-infrared spectra, primarily through their integration into light-emitting diodes (LEDs) and photodetectors. In these devices, QDs serve as the active layer, leveraging their quantum confinement effects for high color purity and efficiency. This section focuses on QD-based LEDs (QLEDs), display technologies, and photodetectors, highlighting key structural and performance advancements. QLEDs typically feature a multilayer structure where the QD emissive layer is sandwiched between a hole transport layer (HTL) and an electron transport layer (ETL), facilitating balanced charge injection and recombination for . The HTL, often composed of organic materials like poly(9-vinylcarbazole), injects holes into the QD layer, while the ETL, such as zinc oxide nanoparticles, transports electrons, minimizing non-radiative losses. Recent optimizations in red QLEDs have achieved external quantum efficiencies (EQEs) exceeding 20%, with record values reaching 36.5% as of November 2025 through enhanced charge balance and reduced recombination. These efficiencies stem from doping strategies and improved passivation, enabling brighter operation at lower voltages. In display applications, QDs excel in color conversion for (LCD) backlights, where they down-convert LED light into pure and green emissions, expanding the color gamut to over 100% . Cadmium-free QDs, such as InP-based variants, have been commercially adopted since to comply with environmental regulations, offering comparable photoluminescence quantum yields while reducing . Self-emissive QLED televisions represent the next frontier, with 2025 advancements in eco-friendly InP QDs achieving brighter, more stable panels through core-shell architectures that enhance and thermal resilience. These InP systems provide wide color tunability without heavy metals, supporting sustainable large-area fabrication via . Recent developments include third-generation photonic quantum dots (P-QDs) potentially delivering up to 95% color space coverage. QD photodetectors leverage the strong absorption and tunable bandgap of materials like for short-wave (SWIR) detection, crucial for and sensing beyond 1 μm. QD devices exhibit high (up to ~0.8 A/W at 1300 nm) in optimized heterostructures with ETLs like ZnO, enabling sensitive SWIR response up to 1.7 μm with low dark currents. Heavy-metal-free alternatives, such as Ag₂Se QDs, have emerged in 2025 research, delivering comparable SWIR performance and improved for integrated systems. These non-toxic QDs address regulatory concerns while maintaining fast response times under low bias. Despite these advances, QLEDs face stability challenges under operational bias, including migration and interfacial that limit lifetimes to hundreds of hours. Recent perovskite-QD hybrids mitigate these issues by combining the broadband absorption of perovskites with QD emission control, extending spectral coverage and boosting device stability through passivation layers.

Energy Conversion and Storage

Quantum dots (QDs) have emerged as promising materials for energy conversion and applications due to their tunable bandgap, high surface area, and enhanced dynamics enabled by quantum confinement effects. In photovoltaic systems, QDs facilitate efficient light absorption across a broad spectrum, including regions, while in energy devices, their improves and capacity retention. These properties position QDs as key enablers for advancing technologies, with recent developments focusing on integration strategies to minimize recombination losses and enhance stability. In quantum dot-sensitized solar cells (QDSSCs), lead sulfide (PbS) QDs deposited on titanium dioxide (TiO2) electrodes have demonstrated power conversion efficiencies (PCE) approaching 13%, attributed to effective charge separation at the QD-TiO2 interface and extended light harvesting via multiple exciton generation. Hybrid architectures combining QDs with perovskites have further boosted performance; for instance, perovskite quantum dot solar cells (PQDSCs) achieved a certified PCE of 18.3% as of 2025 through optimized ligand exchange and defect passivation, enabling better charge extraction and reduced hysteresis. Integration of QDs with nanowires, such as ZnO nanowires interpenetrated with colloidal QDs, enhances charge collection by providing direct pathways for electron transport, resulting in improved external quantum efficiencies in near-infrared regions. For lithium-ion batteries, germanium (Ge) QDs serve as high-capacity anodes, leveraging their theoretical capacity of 1624 mAh/g and confinement-enhanced lithium diffusion kinetics to deliver reversible capacities exceeding 1000 mAh/g, as seen in Ge QDs embedded in nitrogen-doped carbon frameworks that retain 1042 mAh/g after 2000 cycles at C/2 rate. Silicon (Si) QDs similarly benefit from quantum confinement to mitigate volume expansion during lithiation, offering stable cycling with capacities around 1000-1500 mAh/g in composite anodes. These materials outperform traditional anodes by providing higher while maintaining structural integrity through nanoscale effects. In , () QDs excel as visible-light absorbers, with heterostructures like Cu2S@ achieving evolution rates of 10 mmol/g/h under irradiation, driven by efficient p-n junction formation that suppresses charge recombination. Advanced composites, such as coupled with MoS2, have reported rates up to 68.89 mmol/g/h with an apparent of 6.39% at 420 nm, highlighting the role of cocatalysts in promoting proton reduction. Recent green synthesis methods for QDs, including biomass-derived approaches, support sustainable ; for example, eco-friendly QDs produced via hydrothermal processes from natural precursors enable scalable, low-toxicity photocatalysts for long-term H2 generation as of 2025. Carbon QDs have gained traction in flexible supercapacitors, where their pseudocapacitive behavior and high conductivity contribute to via rapid adsorption and reactions. In gel polymer electrolytes incorporating carbon QDs, devices exhibit enhanced specific and cycling stability, with recent 2025 advancements achieving energy densities suitable for wearable applications through nitrogen doping and . These developments underscore the versatility of carbon QDs in bridging energy conversion with flexible storage solutions.

Biomedical Uses

Quantum dots (QDs) have emerged as versatile platforms in biomedical applications due to their tunable and ability to serve as multifunctional probes for , , and diagnostics. In , CdSe/ZnS core-shell QDs, modified with bioconjugates such as ubiquinone, enable high-resolution tracking of cellular processes with minimal , allowing for long-term monitoring in biological systems. Similarly, PbS QDs, with emissions in the near-infrared (NIR) range around 1000-1500 nm, facilitate deep-tissue by penetrating several millimeters into tissues with reduced autofluorescence and scattering, as demonstrated in studies simulating biological environments. In therapeutic applications, QDs function as antennae in (PDT) by absorbing light and transferring energy to generate , which induces oxidative damage to cancer cells; for instance, QDs like CdSe have shown efficient production with quantum yields up to 0.5 under visible light irradiation. Surface functionalization of QDs with targeting ligands, such as peptides or antibodies, further enables , where QDs conjugate with chemotherapeutic agents like for site-specific release in tumor microenvironments, improving efficacy while reducing systemic toxicity in models. For diagnostics, quantum dots (GQDs) have been integrated into biosensors for cancer detection, achieving sensitivities exceeding 95% in identifying tumor biomarkers through quenching mechanisms, as reported in 2024 studies on diagnostics. Recent advancements in 2025 include QD-infused nanocomposites for multiplexed assays, enabling simultaneous detection of multiple cancer biomarkers with enhanced signal amplification and limits of detection in the picomolar range, supporting point-of-care applications. Biocompatibility remains a key challenge, addressed through coatings on QDs, which sterically hinder protein adsorption and reduce clearance, thereby extending circulation times up to several hours compared to uncoated counterparts. Heavy-metal-free alternatives, such as , offer inherently safer profiles due to their composition from biocompatible carbon sources, exhibiting low (cell viability >90% at concentrations up to 100 μg/mL) and enabling applications in bioimaging and without heavy metal leaching concerns.

Emerging Materials Applications

Quantum dots (QDs) have shown significant promise in chemical sensing applications, particularly through films that detect gases via fluorescence quenching mechanisms. For instance, PbS QD-doped poly(3-hexylthiophene) films enable solution-processed NO₂ sensors with high responsivity to concentrations as low as 0.4 ppm at room temperature, leveraging the QDs' tunable bandgap for selective detection. Similarly, SnS₂ QDs in resistive configurations achieve room-temperature NO₂ sensing with enhanced sensitivity due to quantum confinement effects, offering rapid response times under ambient conditions. In photonic nose systems, QD-based layers integrated with AI algorithms facilitate multi-gas identification, improving selectivity in complex environments through pattern recognition of quenching profiles. Carbon QDs (CQDs) extend these capabilities to , where their and photostability enable detection of pollutants like and organic contaminants. CQDs synthesized from natural precursors serve as sorbents and fluorescent probes, allowing real-time tracking of parameters with limits of detection in the nanomolar range. Their role in sensing and volatile organic compounds highlights their versatility, as surface functionalization enhances adsorption and signal amplification for field-deployable devices. Overall, CQDs contribute to sustainable monitoring by integrating into low-cost, portable platforms that minimize environmental impact during deployment. In anticorrosion coatings, QDs enhance barrier properties and active protection mechanisms. Epoxy resins incorporating CuS and ZnS QDs demonstrate inhibition efficiencies exceeding 90% against chloride-induced corrosion on steel substrates, attributed to the QDs' ability to form passivation layers and block ionic pathways. These 2025 developments show improved mechanical adhesion and long-term durability, with electrochemical impedance spectroscopy confirming reduced corrosion rates by orders of magnitude compared to unfilled epoxies. For self-healing materials, Cu- and N-co-doped CQDs in waterborne epoxy coatings enable autonomous repair of micro-cracks through pH-responsive release of inhibitors, achieving self-healing efficiencies up to 85% while maintaining anticorrosion performance over extended immersion periods. This synergy arises from the QDs' catalytic activity in promoting polymerization at damage sites, extending coating lifespan in harsh conditions. In fundamental science, magnetic QDs advance by enabling control of spin-polarized currents at nanoscale interfaces. Carbon QDs with engineered magnetic edge states facilitate ultrafast manipulation, supporting applications in low-power spintronic devices through enhanced hot-carrier mobility and reduced decoherence. Hybrid systems coupling QDs to magnetic insulators generate currents via temperature gradients, offering insights into spin caloritronics for energy-efficient information processing. For , silicon QDs serve as qubits with fidelities exceeding 99%, achieved through industry-compatible fabrication on 300 mm wafers, paving the way for scalable arrays with error rates below thresholds. These single-electron qubits in leverage long times and CMOS integration, demonstrating gate fidelities above 99.9% in multi-qubit operations. Recent advances in 2025 emphasize bioresource-derived materials, such as carbohydrate-based CQDs synthesized from like , which offer green alternatives for and with high quantum yields and low toxicity. These CQDs, derived from sustainable sources, enable selective detection of analytes in , with turn-off responses tailored for environmental and applications. QD-infused nanocomposites further revolutionize diagnostics by amplifying signals in point-of-care devices, where core-shell QD-polymer hybrids achieve sub-femtomolar detection limits for biomarkers through . This integration enhances and , supporting rapid, non-invasive assays with minimal sample volumes.

Safety and Environmental Considerations

Health Risks

Quantum dots (QDs), particularly those based on cadmium such as CdSe and CdTe, pose health risks primarily through the leaching of toxic heavy metal ions like Cd²⁺, which can occur under physiological conditions and lead to cellular damage. This leaching disrupts mitochondrial function and causes DNA damage in various tissues. Additionally, QDs generate reactive oxygen species (ROS), inducing oxidative stress that promotes apoptosis, inflammation, and organ dysfunction, with cadmium-based QDs showing pronounced effects in neuronal cells. Cadmium ions themselves exhibit acute toxicity, with an oral LD50 of approximately 100 mg/kg in rodents, and QD degradation amplifies this risk by releasing bioavailable ions. Exposure to QDs can occur via , dermal contact, or , with potential for through damaged or mucous membranes during occupational handling or accidental release. is a key concern in settings, where aerosolized QDs may deposit in tissues and translocate to the bloodstream. is influenced by and shape; smaller QDs (e.g., <5 nm) exhibit higher cytotoxicity due to enhanced cellular uptake and greater surface area for ion release, while rod-shaped QDs demonstrate increased toxicity compared to spherical ones owing to higher aspect ratios that facilitate membrane interactions. In vitro studies reveal varying cytotoxicity depending on surface coatings, with IC50 values ranging from 0.044 mg/mL for PEG-OH coated CdSe/ZnS QDs in breast cancer cells to higher thresholds (e.g., ~1 mg/mL) for more stable coatings that limit ion release. In vivo rodent models, such as mice and rats, show dose-dependent effects like hepatotoxicity and oxidative damage at exposures of 10-80 mg/kg for CdSe QDs, leading to histopathological changes in liver and kidneys without a defined LD50 for intact QDs but mirroring cadmium ion lethality. Recent 2024 reviews highlight that carbon-based QDs present lower health risks than cadmium-based counterparts, exhibiting higher biocompatibility and minimal ROS induction, making them preferable for biomedical applications where safety is paramount. To mitigate these risks, core-shell designs (e.g., ) encapsulate the toxic core, reducing Cd²⁺ leaching and cytotoxicity by up to several fold in cellular assays. Ligand engineering, such as PEGylation or protein coatings like BSA, further stabilizes QDs, minimizes protein corona formation, and lowers uptake in non-target cells, thereby decreasing systemic toxicity. Regulatory frameworks in the European Union classify cadmium-containing QDs under REACH and directives, emphasizing risk assessments for leaching and exposure in consumer products to ensure safe handling.

Sustainability and Eco-Friendly Developments

Quantum dots (QDs), particularly those containing heavy metals like cadmium, exhibit significant environmental persistence in aquatic and terrestrial systems. In water, cadmium-based QDs such as CdSe cores dissolve slowly through oxidation, often requiring over 80 days in the absence of light, while carbon QDs can persist for decades in turbid conditions due to aggregation and sorption to natural organic matter. In soil, predicted concentrations from sludge or sewage treatment range from 0.0001 to 17 ng/kg, with stability influenced by pH and electrolytes, leading to limited mobility but potential long-term accumulation. Bioaccumulation of released metals, especially cadmium, occurs across trophic levels; for instance, CdSe QDs lead to cadmium uptake in aquatic organisms like Daphnia magna and Ceriodaphnia dubia, with biomagnification factors up to 1.4 from bacteria to protozoa, and high bioaccumulation factors such as 5211 in plants like Lemna minor. Incineration of QD-containing waste can concentrate cadmium in ashes, potentially exceeding U.S. Resource Conservation and Recovery Act (RCRA) limits for hazardous waste. Recent advancements in green synthesis methods address these concerns by employing aqueous-based processes and natural precursors, minimizing toxic solvents and energy use. For example, nitrogen-doped carbon quantum dots (N@CQDs) synthesized via microwave-assisted treatment of apricot (Prunus armeniaca) juice achieve a quantum yield of 37.1% while using only water as a solvent, enabling rapid (5-minute) production from renewable biomass and reducing environmental impact through green chemistry principles. Similarly, carbohydrate-derived carbon dots from sources like jackfruit seeds or banana peels via low-energy hydrothermal or microwave methods (e.g., 600 W for 90 seconds) yield fluorescent materials suitable for sensing and imaging, with non-toxic reagents like ethanol and citric acid replacing hazardous organics. These approaches, including 2025 developments in biomass pyrolysis from onion waste or lemon juice, maintain high optical performance while lowering carbon footprints compared to traditional organometallic routes. Lifecycle assessments highlight the shift toward heavy-metal-free QDs to mitigate e-waste toxicity and enhance recyclability. Carbon dots, derived from agro-waste, eliminate cadmium and reduce annual CO₂ emissions by up to 14,000 tons in applications like QD-LED televisions, contrasting with indium phosphide (InP) QDs that consume 150 tons of indium yearly. In 2024, heavy-metal-free InP and carbon-based QDs enabled simpler recycling in solar cells, complying with circular economy goals, though challenges persist in end-of-life degradability studies for carbon variants. Overall, these materials lower eco-toxicity in disposal phases, with biomass sourcing further decreasing reliance on finite resources. Regulatory frameworks under REACH and EPA guidelines enforce sustainability in QD nanomaterials through hazard assessments and reporting. Since 2020, REACH mandates registration of nanoforms, including detailed physicochemical characterization and exposure evaluations for substances like QDs, to ensure safe lifecycle management. The U.S. EPA, via the Toxic Substances Control Act (TSCA), has reviewed over 160 nanoscale material notices since 2005, requiring pre-manufacture notifications for novel QDs to evaluate environmental persistence and bioaccumulation risks. Sustainability metrics, such as balancing quantum yield (e.g., >30% in green InP QDs) against eco-costs like solvent use and metal content, guide compliance, promoting alternatives that achieve high efficiency with minimal environmental burden.

Historical Development

Following the foundational discoveries in the late 20th century, quantum dot research advanced rapidly in the with improvements in synthesis techniques. In 1993, and colleagues at introduced the hot-injection method for producing high-quality, monodisperse colloidal quantum dots, enabling precise size control and uniformity below 5% variation. This breakthrough facilitated scalable production and paved the way for practical applications. The late 1990s marked the transition to applied research. In 1998, the establishment of Quantum Dot Corporation (QDC) focused on commercializing quantum dots for biomedical and display technologies. That same year, researchers demonstrated quantum dots as biological labels, with studies by Bruchez et al. and Chan et al. showing their potential for due to tunable emission and photostability. The early 2000s saw innovations in quantum dot structures and devices. In 1996, Philippe Guyot-Sionnest's team developed core-shell quantum dots to enhance stability and quantum yield. By 2000, Paul Alivisatos' group at UC Berkeley synthesized rod-shaped quantum dots, expanding morphological diversity. A milestone in optoelectronics came in 2002 when Stephen Coe-Sullivan and colleagues fabricated the first quantum dot light-emitting diodes (QD-LEDs), achieving electroluminescence for potential display applications. Commercialization accelerated in the . Samsung and LG introduced quantum dot-enhanced LCD televisions in 2015, using CdSe/ZnS dots as color converters to achieve wider color gamuts. Concerns over toxicity spurred development of eco-friendly alternatives, such as (InP) and carbon dots. In 2019, -free quantum dots were integrated into consumer TVs, improving sustainability. The field continued to evolve into the 2020s, with the 2023 Nobel Prize in Chemistry recognizing Ekimov, Brus, and Bawendi's contributions, boosting global interest. Recent advances as of 2025 include perovskite quantum dots for high-efficiency LEDs and solar cells, with external quantum efficiencies exceeding 25% in some prototypes, and explorations in using dot-based spin qubits. These developments underscore quantum dots' ongoing impact across .

References

  1. [1]
    Quantum Dots and Applications - PMC - NIH
    Feb 18, 2020 · Abstract. It is the unique size-dependent band gap of quantum dots (QDs) that makes them so special in various applications.Missing: definition | Show results with:definition
  2. [2]
    Quantum Dots at Work - NSF Impacts
    Their ability to absorb and emit light with extraordinary precision makes them essential building blocks in electronics, photonics and quantum science. Over the ...Missing: definition | Show results with:definition
  3. [3]
    Quantum dots make it big at last | Nature Materials
    Feb 2, 2024 · The 2023 Nobel Prize in Chemistry acknowledged work that paved the way to a nanotechnology worthy of the name.
  4. [4]
    Review: Quantum Dot Light-Emitting Diodes - ACS Publications
    Feb 16, 2023 · Quantum dot light-emitting diodes (QD-LEDs) are one of the most promising self-emissive displays in terms of light-emitting efficiency, wavelength tunability, ...
  5. [5]
    Biomedical Applications of Quantum Dots: Overview, Challenges ...
    May 2, 2022 · Quantum dots (QDs) are semiconductors-based nanomaterials with numerous biomedical applications such as drug delivery, live imaging, and medical diagnosis.Preparation Of Qds · Live Cell Imaging And In... · Traceable Drug DeliveryMissing: definition | Show results with:definition
  6. [6]
    An Experimental Introduction to Colloidal Nanocrystals through InP ...
    Mar 28, 2023 · Quantum dots are colloidal semiconductor nanocrystals that display size-dependent electronic and optical properties.
  7. [7]
    Quantum Dots and Their Applications: What Lies Ahead?
    Jun 26, 2020 · 1,2) These semiconductor nanocrystals, with nanometer-sized diameters, exhibit quantum size effects in their optical and electronic properties. ...
  8. [8]
    Surface-Related Emission in Highly Luminescent CdSe Quantum Dots
    We report our experimental studies of surface-related emission in highly luminescent CdSe quantum dots (QDs) with controlled quantum yield and photooxidation.
  9. [9]
    Effect of Quantum and Dielectric Confinement on the Exciton ...
    Our analysis shows that the combined contributions from quantum and dielectric confinement can result in strong exciton−exciton repulsion with giant ...
  10. [10]
    (CdSe)ZnS Core−Shell Quantum Dots: Synthesis and ...
    In the smallest sizes of TOPO capped dots the color of the PL is normally dominated by broad deep trap emission and appears as faint white light. After ...
  11. [11]
    Type-II Quantum Dots: CdTe/CdSe(Core/Shell) and CdSe/ZnTe ...
    The type-II quantum dots reported here can emit at lower energies than the band gaps of comprising materials. The type-II emission can be tailored by the shell ...
  12. [12]
    Shell layer dependence of photoblinking in CdSe/ZnSe/ZnS ...
    Jan 5, 2011 · We report on the influence of shell addition, shell thickness, and shape of CdSe/ZnSe/ZnS colloidal quantum dots (QDs) on luminescence, ...
  13. [13]
    Shell Phase and Morphology Control for Emission Tuning in III–V ...
    Aug 5, 2025 · Epitaxial growth of shells on III–V semiconductor quantum dot (QD) cores yields improved fluorescence quantum efficiency and stability ...
  14. [14]
    Role of the Inner Shell Architecture on Quantum Yield and Blinking ...
    We show that the inner shell architecture between the CdSe core and the outer ZnS shell plays a significant role in both quantum yield and blinking dynamics.Missing: seminal | Show results with:seminal
  15. [15]
    Core-shell quantum dots: A review on classification, materials ...
    Nov 30, 2022 · The current review highlights the classifications and applications of CSQD systems based on semiconductor materials. The aspects related to the ...Missing: seminal | Show results with:seminal
  16. [16]
    Synthesis of Quaternary Semiconductor Nanocrystals with Tunable ...
    Tuning the composition of heavy metal-free quaternary quantum dots for improved photoelectrochemical performance. Journal of Materials Chemistry A 2021, 9 ...Experimental Section · I Crystalline Structure · Ii Composition And...Missing: seminal | Show results with:seminal
  17. [17]
    Optical properties of manganese-doped nanocrystals of ZnS
    Jan 17, 1994 · We report for the first time that doped nanocrystals of semiconductor can yield both high luminescent efficiencies and lifetime shortening at the same time.Missing: ZnSe dots
  18. [18]
    Frontier challenges in doping quantum dots - RSC Publishing
    We review the state-of-the-art doping techniques arising from colloidal synthesis methods. We first discuss the advantages and challenges involved in doping.<|control11|><|separator|>
  19. [19]
    Highly efficient green InP-based quantum dot light-emitting diodes ...
    May 30, 2022 · To acquire high-quality InP-based QDs, we designed a gradient alloyed inner shell ZnSexS1−x layer between InP cores and outer ZnS shell as shown ...
  20. [20]
    Synthesis and characterization of nearly monodisperse CdE (E ...
    Synthesis and characterization of nearly monodisperse CdE (E = sulfur, selenium, tellurium) semiconductor nanocrystallites
  21. [21]
    [PDF] Colloidal Synthesis of Semiconductor Quantum Dots toward Large ...
    Jan 29, 2018 · This review covers colloidal QD synthesis via chemical approaches, including hot-injection, noninjection, aqueous, and biosynthesis, and ...<|control11|><|separator|>
  22. [22]
    UV induced synthesis of starch capped CdSe quantum dots ...
    UV induced synthesis of starch capped CdSe quantum dots: Functionalization with thiourea and application in sensing heavy metals ions in aqueous solution.
  23. [23]
    Counterion-Mediated Ligand Exchange for PbS Colloidal Quantum ...
    Oct 29, 2015 · PbS quantum dots of 3.8 nm diameter capped with oleic acid and dispersed in chloroform were synthesized as reported previously using a hot- ...
  24. [24]
    Colloidal Quantum Dots for Solar Technologies - ScienceDirect.com
    Aug 10, 2017 · ... synthesis can produce 1 g of NCs in one batch. Because of its effectiveness, the hot-injection method has also been applied for the ...
  25. [25]
    Continuous flow synthesis of PbS/CdS quantum dots using ...
    Nov 19, 2024 · The synthesis uses continuous flow, substituted thioureas as sulfur source, and lead/cadmium oleate as metal precursors, achieving high-quality ...
  26. [26]
    Continuous Flow Aqueous Synthesis of Highly Luminescent AgInS2 ...
    Nov 23, 2022 · The presented method provides excellent reproducibility and high scalability, facilitating the large-scale production of highly luminescent AIS/ZnS QDs.
  27. [27]
    Laser-synthesized oxide-passivated bright Si quantum dots ... - Nature
    Apr 22, 2016 · We introduce ultrapure laser-synthesized Si-based quantum dots (QDs), which are water-dispersible and exhibit bright exciton PL in the window of relative ...
  28. [28]
    Strain-driven quantum dot self-assembly by molecular beam epitaxy
    Jul 16, 2020 · In this tutorial, we summarize the history of self-assembled QDs, outline some examples of quantum optics applications based on QDs, discuss the science that ...Missing: lithographic | Show results with:lithographic
  29. [29]
    Viral assembly of oriented quantum dot nanowires - PNAS
    The highly organized structure of M13 bacteriophage was used as an evolved biological template for the nucleation and orientation of semiconductor nanowires ...
  30. [30]
    Hybrid Electrochemical/Chemical Synthesis of Quantum Dots
    The “electrochemical/chemical method” (or “E/C method”) is a new wet chemical method for synthesizing semiconductor quantum dots on graphite surfaces.
  31. [31]
    Size-Dependent Optical Properties of InP Colloidal Quantum Dots
    Sep 6, 2023 · (21,22) Briefly, in most semiconductors the ground exciton state is a dark state, and in small sized II–VI and III–V CQDs strong spatial ...Missing: IV- | Show results with:IV-
  32. [32]
    Semiconductor quantum dots: Technological progress and future ...
    Aug 6, 2021 · Quantum confinement emerges when electrons are constrained to a domain comparable with their de Broglie wavelength. Quantum-confined structures ...Missing: seminal | Show results with:seminal
  33. [33]
    Suppressing the Fluorescence Blinking of Single Quantum Dots ...
    Sep 8, 2016 · The blinking behavior will reduce the photons generation rate, cause difficulty in single particle tracking, and degrade the performance of ...
  34. [34]
    Mechanisms and Suppression of Quantum Dot Blinking
    ### Summary of Recent Advances on Quantum Dot Blinking Suppression Mechanisms
  35. [35]
    Exploring the frontier: nonlinear optics in low dimensional materials
    Quantum dots, with their size-tunable electronic properties, excel in multi-photon absorption and frequency up-conversion, which are quintessential for ...
  36. [36]
    Recent Progress on Second Near-Infrared Emitting Carbon Dots in Biomedicine
    ### Summary of Advances in Second Near-Infrared Emitting Carbon Dots and Emission Wavelengths
  37. [37]
    Polaron engineering promotes NIR-II absorption of carbon quantum ...
    Jul 5, 2024 · This discovery could open a promising route for the polarons of organic semiconductor materials as NIR-II absorbers in nanomedical applications.
  38. [38]
    Discrete excitonic transitions in quantum-sized CdS particles
    Five discrete excitonic transitions were observed in quantum dots,i.e. quantum-sized CdS, as a function of particle size. The experimentally determined ...
  39. [39]
    Effective mass and dielectric constant mismatch effects in spherical ...
    Jan 9, 2007 · The role of effective mass and dielectric mismatches on chemical potentials and addition energies of many-electron multishell quantum dots ...
  40. [40]
    Density of States Measured by Scanning-Tunneling Spectroscopy ...
    The other discrete energy levels also have a fourfold higher multiplicity with respect to the corresponding energy levels of II-VI and III-V quantum dots (QDs).
  41. [41]
    Transport in quantum dots - ScienceDirect
    This is the Coulomb blockade, a single-electron charging effect. Only when the applied bias is sufficiently high can a carrier tunnel to the dot.
  42. [42]
    Transport in Single Quantum Dots: A Review from Linear Response ...
    Aug 2, 2025 · The Kondo resonance can overcome the Coulomb blockade, enabling ballistic conduction through the SET QD (see bottom panel of Fig. 3), inducing ...
  43. [43]
    Coulomb Blockade and Hopping Conduction in PbSe Quantum Dots
    Oct 3, 2005 · As the interdot coupling is increased, the system evolves from an insulating regime dominated by Coulomb blockade to a semiconducting regime, ...Abstract · Article Text
  44. [44]
    Electrical Transport in Colloidal Quantum Dot Films - PubMed
    May 3, 2012 · Transport by hopping may nevertheless lead to an increased mobility with decreasing temperature for some temperature range, and such behavior at ...
  45. [45]
    Surface Reconstructions in II–VI Quantum Dots | ACS Nano
    Here, we use DFT calculations to study the electronic properties of QDs as a function of size, up to a diameter of ∼4.5 nm.
  46. [46]
    Electronic and Structural Differences between Wurtzite and Zinc ...
    By establishing unreconstructed and defect-free surface facets for both Wz and Zb, we can reliably measure differences between valence and conduction band edges ...
  47. [47]
    Error (ACS Publications)
    - **Insufficient relevant content**: The requested URL (https://pubs.acs.org/doi/10.1021/jp511025g) returns a "Page Not Found" error, indicating the page is inaccessible or does not exist.
  48. [48]
  49. [49]
    Scanning Tunneling Spectroscopy of Individual PbSe Quantum Dots ...
    The electronic local density of states (LDOS) of single PbSe quantum dots (QDs) and QD molecules is explored using low-temperature scanning tunneling ...
  50. [50]
    Engineering the Surface Chemistry of Colloidal InP Quantum Dots ...
    Engineering the Surface Chemistry of Colloidal InP Quantum Dots for Charge Transport ... cm2/Vs (10.5 ± 2.4 cm2/Vs) and with current modulation of 105 ...
  51. [51]
    [PDF] QUANTUM DOTS – SEEDS OF NANOSCIENCE - Nobel Prize
    Oct 4, 2023 · The basic theoretical concept underlying quantum dots is referred to as the 'particle-in-a-box' ... model describing the effect of particle size ...Missing: seminal | Show results with:seminal
  52. [52]
    A Powerful Visual Aid for Introducing the Particle in a Box
    Semiconductor nanocrystals can be qualitatively modeled as a particle in a box. This paper presents methods of fabricating nanocrystals.Missing: seminal | Show results with:seminal
  53. [53]
    (PDF) Theoretical study of the quantum confinement effects on ...
    Aug 6, 2025 · Using particle in a box model, we have studied the quantum confinement effects on quantum dots (QDs). A mathematical equation for the confinement energy and ...
  54. [54]
    Tight-binding model for semiconductor nanostructures | Phys. Rev. B
    An empirical s c ⁢ p 3 a tight-binding (TB) model is applied to the investigation of electronic states in semiconductor quantum dots.Missing: seminal | Show results with:seminal
  55. [55]
    x ⁢ N / G a ⁢ N quantum dots using an eight-band k ∙ p model
    Oct 26, 2006 · The eight-band k ∙ p -model accounts for strain effects, piezoelectricity and pyroelectricity, and spin-orbit and crystal-field splitting.
  56. [56]
    First-principles study of the influence of different interfaces and core ...
    Jun 3, 2015 · We present such a study of CdSe/CdS core-shell NCs using density functional theory, where we focus on dependence of the properties of these NCs on core types ...
  57. [57]
    Quantifying the Size‐Dependent Exciton‐Phonon Coupling Strength ...
    Jan 15, 2024 · Exciton-phonon coupling is particularly relevant in quantum dots (QDs) of APbX3 lead-halide perovskite, a new class of semiconductors with ...
  58. [58]
    Time-dependent density functional theory of excitation energies of ...
    We study the low-energy excitations of two-dimensional elliptical quantum dots with closed-shell structures using time-dependent density functional theory ...
  59. [59]
    Investigating the electronic properties of edge glycine/biopolymer ...
    Sep 20, 2024 · This study systematically investigated four types of graphene quantum dots (GQDs) AHEX, ZTRI, ZHEX, and ATRI, and their interactions with glycine to form GQD- ...
  60. [60]
    Impact of carrier transport on the performance of QD lasers on silicon
    The operation of quantum dot lasers epitaxially grown on silicon is investigated through a quantum-corrected Poisson-drift-diffusion model.
  61. [61]
    Colloidal Quantum Dots Thin Films: A Theoretical Insight Into Optical ...
    Dec 27, 2024 · This article presents an analytical model of QD optical indices combining tight-binding simulations and effective medium theory.
  62. [62]
    Mie Theory - an overview | ScienceDirect Topics
    Mie theory is an analytical solution of Maxwell's equations for the scattering of electromagnetic radiation by particles of any size (also called Mie scattering) ...
  63. [63]
    Engineering Auger recombination in colloidal quantum dots ... - Nature
    Apr 15, 2019 · Auger recombination is the main non-radiative decay pathway for multi-carrier states of colloidal quantum dots, which affects performance of most of their ...
  64. [64]
    Multiscale QM/MM modelling of catalytic systems with ChemShell
    Apr 20, 2023 · We present an overview of recent applications of ChemShell to problems in catalysis and review new functionality introduced into the redeveloped ...
  65. [65]
    Spectroscopic and Device Aspects of Nanocrystal Quantum Dots
    Sep 28, 2016 · One of the defining features of a semiconductor is an energy gap (Eg) separating a fully occupied valence band from the unoccupied conduction ...
  66. [66]
    Degradation of 1.3 μm Quantum Dot Laser Diodes for Silicon ...
    Jul 17, 2024 · To support our assumption, we employed a quantum-corrected Poisson-drift-diffusion simulation tool to evaluate the carrier distribution and ...
  67. [67]
    Ubiquinone-quantum dot bioconjugates for in vitro and intracellular ...
    Mar 25, 2013 · Thus, Q2NS modified CdSe/ZnS QDs are suitable for use in some potential biomedical applications. ... Quantum dots for live cells, in vivo imaging ...
  68. [68]
    (PDF) PbS Quantum Dots for Near-Infrared Fluorescence Imaging
    In this phantom-based study, we assessed the imaging potential of lead sulfide (PbS) near-infrared quantum dots (QDs) as novel contrast agents for deep tissue ...
  69. [69]
    Semiconductor Quantum Dots for Photodynamic Therapy
    We are currently investigating singlet oxygen generation on less toxic QDs15 and are evaluating different surface-coating techniques to increase the 1O2 quantum ...
  70. [70]
    Quantum dots as targeted doxorubicin drug delivery nanosystems in ...
    Feb 8, 2021 · This study aimed at developing a new targeted drug delivery (TDD) system based on Ag–In–Zn–S quantum dots (QDs).
  71. [71]
    Emergence of Quantum Dots as Innovative Tools for Early Diagnosis ...
    May 9, 2024 · This review underscores the inmense potential of quantum dots (QDs) in advancing the diagnosis and treatment of triple negative breast cancer (TNBC).Abstract · Introduction · General Classification of... · QDs in Cancer Diagnosis and...
  72. [72]
  73. [73]
    [PDF] Synthesis, characterization and cytotoxicity of polyethylene glycol ...
    Jan 1, 2019 · (PEG) was con- sidered to be the best biocompatible material which is well known to increase the circulation time by decreasing the clearance of ...
  74. [74]
    Carbon Quantum Dots in Biomedical Applications: Advances ...
    Nov 28, 2024 · Their small size, biocompatibility, and ease of surface functionalization allow for the development of targeted drug delivery systems, enhancing ...
  75. [75]
    Solution-processed NO2 gas sensor based on poly(3-hexylthiophene)
    Sep 4, 2024 · A NO 2 sensor was fabricated by doping PbS QDs with poly(3-hexylthiophene) (P3HT). The as-developed sensor exhibited high responsivity to 100–0.4-ppm NO 2 gas.
  76. [76]
    Room Temperature Resistive NO2 Sensor Based on SnS2 Quantum ...
    Feb 24, 2025 · Quantum dots (QDs), measuring just a few nanometers in size, exhibit unique properties derived from quantum confinement and edge effects as well ...
  77. [77]
  78. [78]
    Carbon dots as sensors and sorbents in environmental monitoring ...
    This review aims to analyze the current knowledge about carbon dots for application in environmental monitoring and chemical analysis.
  79. [79]
    Illuminating Pollutants: The Role of Carbon Dots in Environmental ...
    This review critically analyzes the transformative role of CDs in environmental sensing and remediation. Highlighting their versatile applications.
  80. [80]
    The Role of Carbon Quantum Dots in Environmental Protection
    Jun 22, 2024 · This review will thoroughly examine how the several stages of CQDs life cycle can be exploited in the protection of the environment.Green Synthesis using Waste... · Pollution Management · Energy Generation
  81. [81]
    Study of anti-corrosion epoxy resin coatings with high corrosion ...
    A series of CuS and ZnS quantum dots (QD) coatings is prepared. Quantum dots effectively increase coatings' anti-corrosion performance.
  82. [82]
    Study of anti-corrosion epoxy resin coatings with ... - ResearchGate
    Aug 7, 2025 · Study of anti-corrosion epoxy resin coatings with high corrosion resistance and mechanical performance based on quantum dots.
  83. [83]
    Cu, N co-doped carbon quantum dots empowering waterborne ...
    Aug 22, 2025 · Self-healing coatings presented an ideal solution for addressing physical damage in protective layers. They could autonomously repair damaged ...
  84. [84]
    Cu, N co-doped carbon quantum dots empowering waterborne ...
    Aug 27, 2025 · Cu, N co-doped carbon quantum dots empowering waterborne epoxy coatings: synergistic anti-corrosion and self-healing performance. August 2025 ...
  85. [85]
    Manipulation of magnetic edge states in carbon quantum dots for ...
    Jul 1, 2025 · Magnetic edge states of carbon quantum dots enable ultrafast electron relaxation and hot-carrier mobility, offering a high-relaxivity MRI-guided ...
  86. [86]
    Spin caloritronics of a quantum dot coupled to a magnetic insulator ...
    Jul 2, 2025 · We investigate the generation of spin current through temperature gradients in a quantum dot-based hybrid system.
  87. [87]
    Industry-compatible silicon spin-qubit unit cells exceeding 99% fidelity
    Sep 24, 2025 · Among the many types of qubit presently being investigated for a future quantum computer, silicon spin qubits with millions of qubits on a ...
  88. [88]
    Single-Electron Spin Qubits in Silicon for Quantum Computing
    May 2, 2025 · This review focuses on single-spin qubits in silicon. First, we start with foundational spin qubit theory. Then, we discuss gate-defined quantum dots and donor ...
  89. [89]
    Recent advancements in application of carbohydrate-derived ...
    Oct 31, 2025 · The current review provides an extensive and up-to-date overview of carbohydrate-derived carbon quantum dots (CDCQDs) research from 2021 to 2025 ...
  90. [90]
    Quantum dot-infused nanocomposites: revolutionizing diagnostic ...
    The incorporation of QDs into nanocomposites enables them to achieve better diagnostic modes such as targeted delivery, signal amplification, and ...Missing: multiplexed | Show results with:multiplexed
  91. [91]
    Quantum Dot-Infused Nanocomposites: Revolutionizing Diagnostic ...
    Aug 17, 2025 · Supra-Quantum Dot Assemblies to Maximize Color-Based Multiplexed Fluorescence Detection with a Smartphone Camera ... quantum dots. September 2025 ...
  92. [92]
    A Review of in vivo Toxicity of Quantum Dots in Animal Models - PMC
    Dec 29, 2023 · It is reported that QDs exposure can induce abnormal inflammatory responses, increase oxidative stress generation, alter neuronal function and ...
  93. [93]
    Oxidative stress-induced neurotoxicity of quantum dots and ...
    In this review, we focus on elucidating the mechanisms and consequences of CNS oxidative stress induced by QDs.
  94. [94]
    The Effects of Cadmium Toxicity - MDPI
    Cadmium (Cd) is a toxic non-essential transition metal that poses a health risk for both humans and animals.Missing: LD50 | Show results with:LD50
  95. [95]
    A Toxicologic Review of Quantum Dots: Toxicity Depends on ... - NIH
    Inhalation exposures may pose potential risks given that QDs have been shown to be incorporated via endocytosis by a variety of cell types and may reside in ...Missing: shape | Show results with:shape
  96. [96]
    Dependence of Quantum Dot Toxicity In Vitro on Their Size ... - NIH
    Aug 9, 2022 · We demonstrated that QDs with the smallest size had the highest in vitro cytotoxicity. The most toxic QDs were characterized by a low negative surface charge.Missing: IC50 | Show results with:IC50
  97. [97]
    CARBON DOTS: Bioimaging and Anticancer Drug Delivery
    Jan 11, 2024 · CDs offer higher biocompatibility and lower toxicity than heavy metal-containing quantum dots, making them ideal for biological applications.
  98. [98]
    Surface Modification, Toxicity, and Applications of Carbon Dots to ...
    Compared with toxic metal-based QDs [53], CDs are much safer and eco-friendly for biomedical applications as they are free from heavy metals and toxic elements.
  99. [99]
    Towards sustainable quantum dots: Regulatory framework, toxicity ...
    This review aims to provide a clear overview of the European regulatory framework, focusing on horizontal and sector-specific regulations that govern ...
  100. [100]
    Assessing the Environmental Effects Related to Quantum Dot ...
    LC50s of many cadmium and selenium-based QDs are often lower (i.e., more toxic) for QDs without shells or surface modifications based on the overall QD ...
  101. [101]
    Synthesis of highly fluorescent green carbon quantum dots from ...
    Sep 12, 2025 · In this study, juice extracted from Prunus armeniaca (apricots) was employed as a natural precursor to produce N@CQDs using a one-step microwave ...
  102. [102]
    Introduction to Green Carbon Dots | ACS Symposium Series
    Apr 8, 2025 · The synthesis of green carbon dots involves the utilization of natural precursors, non-toxic reagents, and methods that reduce energy ...
  103. [103]
    Carbon Dots: A Review with Focus on Sustainability - Ren - 2024
    Jul 18, 2024 · The study quantifies why it is attractive that CDs can be synthesized with biomass as the sole starting material and be free from toxic and precious metals and ...
  104. [104]
    Research Progress of Heavy-Metal-Free Quantum Dot Light ... - MDPI
    May 9, 2024 · This article reviews the research on heavy-metal-free quantum dot (QD) materials and light-emitting diode (LED) devices.
  105. [105]
    Nanomaterials - ECHA - European Union
    As of 1 January 2020, explicit legal requirements under REACH apply for companies that manufacture or import nanoforms. These reporting obligations address ...
  106. [106]
    Fact Sheet: Nanoscale Materials | US EPA
    Aug 8, 2025 · This page provides information on how EPA treats nanoscale materials, particularily chemicals under the Toxic Substances Control Act (TSCA).
  107. [107]
    Environmentally friendly synthesis of quantum dots and their ...
    Aug 8, 2025 · Zhu et al. applied a low-temperature hot injection preparation method to synthesize super-efficient green luminescent CIS QDs (the synthesis ...Missing: seminal | Show results with:seminal