Fact-checked by Grok 2 weeks ago

Nanostructure

A nanostructure is defined as a or having at least one dimension in the range of 1 to 100 nanometers (ISO/TS 80004-1), positioning it between molecular and bulk scales where quantum effects become prominent. This size regime leads to exceptional properties, including a high surface-to-volume that enhances reactivity and strength, as well as quantum confinement that alters electrical, optical, and magnetic behaviors compared to their bulk counterparts. Nanostructures are classified by dimensionality into zero-dimensional (0D), one-dimensional (1D), two-dimensional (), and three-dimensional () categories, each offering distinct functionalities. 0D nanostructures, such as nanoparticles and quantum dots, confine electrons in all three dimensions, enabling tunable like size-dependent in CdSe quantum dots for biological . 1D nanostructures, including nanowires and carbon nanotubes, exhibit anisotropic growth and ballistic electron transport, supporting applications in and sensors. 2D nanostructures, exemplified by and dichalcogenides, provide exceptional electrical and flexibility, ideal for and devices. 3D nanostructures, such as nanocomposites and aerogels, integrate lower-dimensional components to achieve hierarchical structures with enhanced mechanical and thermal properties. The synthesis of nanostructures employs bottom-up approaches, like and , or top-down methods, such as , to precisely control size, shape, and composition. These materials have revolutionized fields including —where gold nanoparticles enable and diagnostics—energy technologies, such as efficient solar cells and batteries via nanostructured electrodes, and through photocatalytic nanostructures. Ongoing research focuses on scalability, biocompatibility, and integration into real-world devices to harness their full potential.

Definition and Fundamentals

Definition

A nanostructure is defined as a material or system whose structural elements—such as clusters, crystallites, or molecules—have at least one dimension in the range of 1 to 100 nanometers, a scale at which quantum confinement effects and pronounced surface properties significantly influence behavior. According to ISO/TS 80004-4:2011, nanostructured materials have an internal or surface structure with a significant fraction of features, grains, voids, or precipitates in the nanoscale. At this nanoscale, the discrete nature of atomic arrangements leads to phenomena like altered electronic band structures and increased dominance of surface atoms over bulk interior ones, distinguishing nanostructures from larger microscopic systems. In contrast to the broader category of , which includes any substance with nanoscale components regardless of organization, nanostructures emphasize deliberate, ordered atomic or molecular configurations that exploit nanoscale geometry for unique functionalities. This structural specificity allows for tailored properties not achievable in conventional materials, such as enhanced optical or mechanical responses due to the interplay of size and arrangement. A key principle underlying nanostructures is the dramatic increase in surface-to-volume ratio as dimensions shrink, which amplifies surface-driven effects like reactivity, catalytic activity, and capacity while enabling quantum effects such as size-dependent bandgap tuning. For instance, this ratio can rise exponentially with decreasing particle size, leading to behaviors where surface interactions govern overall material performance rather than volume-averaged properties. Basic forms include nanoparticles, nanotubes, and nanofilms, which illustrate how dimensional constraints at the nanoscale can yield emergent properties like improved or . These structures are often classified by the number of confined dimensions, from zero to three, to highlight their varying impacts on material characteristics.

Dimensionality and Classification

Nanostructures are classified according to the number of dimensions confined to the nanoscale (typically below 100 ), a criterion that governs the extent of quantum confinement on wavefunctions and profoundly influences their electronic, optical, and mechanical behaviors. This arises from quantum mechanical principles, where spatial restriction in one or more directions quantizes energy levels, modifies the , and enhances properties like bandgap tunability and mobility compared to bulk materials. The degree of confinement decreases from zero-dimensional (0D) to three-dimensional () structures, leading to progressively weaker quantum effects and behaviors that approach those of conventional solids. In 0D nanostructures, all three dimensions are confined, resulting in point-like structures such as quantum dots, where electron wavefunctions are fully restricted, producing discrete energy levels and strong size-dependent quantum effects that dominate optical absorption and emission spectra. This complete confinement amplifies quantum behaviors, enabling precise control over electronic properties through size variation. One-dimensional (1D) nanostructures confine two dimensions, yielding elongated forms like nanowires, with quantum effects primarily influencing transport along the unconfined length while maintaining in mechanical and electrical responses. The partial confinement here leads to behaviors such as enhanced in the free dimension, distinct from the isotropic nature of bulk counterparts. Two-dimensional (2D) nanostructures restrict one dimension, creating sheet-like architectures exemplified by , where quantum confinement perpendicular to the plane alters in-plane and mechanical strength, often resulting in Dirac-like electronic dispersion. This configuration produces behaviors like exceptional carrier velocities and tunable bandgaps under strain or layering. Three-dimensional (3D) nanostructures lack strict dimensional confinement but feature nanoscale components arranged in bulk-like volumes, such as nanocomposites, where quantum effects are minimal and properties emerge from collective interactions, bridging nanoscale enhancements with macroscopic functionality. Overall, dimensionality dictates the balance between quantum confinement and classical diffusion, with lower dimensions fostering emergent phenomena critical for advanced technologies.

Historical Development

Early Discoveries

The use of nanostructures dates back to ancient civilizations, where artisans inadvertently harnessed nanoscale materials for enhanced material properties. In , forged primarily between the 8th and 18th centuries in the and , the exceptional strength and distinctive patterns arose from nanoscale (iron carbide) nanowires and carbon nanotubes embedded within the metal matrix, as revealed by high-resolution electron microscopy analysis of a 17th-century saber. Similarly, medieval stained-glass windows from the 12th to 16th centuries in achieved their vibrant ruby-red hues through the incorporation of nanoparticles, typically around 30-50 nm in size, which scattered light in a size-dependent manner to produce the desired color without altering the glass's transparency. These early applications demonstrate serendipitous , where nanoscale structures improved aesthetics and durability long before their scientific identification. The 19th century marked the beginning of intentional scientific exploration of nanostructures, with Michael Faraday's pioneering work on colloidal metals laying foundational insights into nanoscience. In his 1857 Bakerian Lecture to the Royal Society, Faraday described the preparation of stable colloidal gold suspensions—finely divided gold particles dispersed in water—observing that these "ruby gold" solutions exhibited intense red colors due to the particles' submicron size, which influenced light interaction through absorption and scattering. Faraday's experiments, including the reduction of gold chloride with phosphorus to form particles averaging 5-10 nm, highlighted the dependence of optical properties on particle size and marked the first systematic study of what would later be recognized as nanomaterials, establishing him as a key figure in the origins of nanoscience. In the early , investigations into carbon-based materials further uncovered nanostructured forms through analysis of and related substances. , produced via incomplete combustion and consisting of aggregated carbon nanoparticles typically 10-100 in diameter, gained recognition around 1912 for its ability to reinforce rubber, enhancing tensile strength and wear resistance in tires—a property attributable to its nanoscale morphology. Early analyses during this period, including those of industrial carbon blacks derived from lampblack processes, revealed complex nanoscale carbon clusters and polyaromatic structures that served as precursors to later-identified fullerenes, providing inadvertent glimpses into carbon nanostructure formation without contemporary awareness of their atomic-scale organization.

Key Milestones in the 20th and 21st Centuries

The development of electron microscopy in the mid-20th century revolutionized the ability to visualize structures at the nanoscale, with transmission electron microscopes () achieving resolutions down to 1 nm by the 1950s and becoming widely available for applications through the 1980s. This technological advancement allowed researchers to observe atomic arrangements in solids, laying the groundwork for deliberate nanostructure studies. In 1959, physicist delivered his seminal lecture "There's Plenty of Room at the Bottom," envisioning the manipulation of individual atoms and molecules through miniaturized tools, which inspired future generations of nanoscientists. In 1974, Norio Taniguchi coined the term "" to describe processes for producing nanoscale structures, particularly in semiconductor manufacturing. A pivotal breakthrough occurred in 1981 with the invention of the (STM) by and at , enabling atomic-resolution imaging and manipulation of surfaces under vacuum conditions. This instrument, recognized with the 1986 , marked the emergence of nanofabrication tools capable of probing and altering matter at the atomic scale. In 1985, Harold W. Kroto, Robert F. Curl, and Richard E. Smalley discovered (C60), a stable, soccer-ball-shaped carbon molecule produced via laser vaporization of , ushering in the era of discrete carbon nanostructures. Their work, honored with the 1996 , demonstrated that carbon could form closed-shell architectures beyond and . The field advanced further in 1991 when at Corporation identified multi-walled carbon nanotubes using high-resolution TEM, revealing helical tubules with diameters of 1-10 nm synthesized from arc-discharge evaporation. These structures exhibited unique one-dimensional properties, sparking intense research into their synthesis and applications. In 1993, and colleagues reported the synthesis of single-walled carbon nanotubes, which consist of a single layer of rolled into a tube, offering even more pronounced one-dimensional properties. In 2000, the United States launched the (NNI), a coordinated federal program allocating nearly $500 million in its initial proposed budget for fiscal year 2001 to support interdisciplinary nanostructure research across agencies like NSF and , formalizing nanotechnology as a national priority. The 21st century saw continued innovation with the 2004 isolation of , a single atomic layer of carbon, by and at the using mechanical exfoliation from . This two-dimensional material, with exceptional electronic and mechanical qualities, earned them the 2010 and catalyzed the exploration of other atomically thin structures. In 2023, the was awarded to Moungi G. Bawendi, , and Alexei I. Ekimov for the discovery and synthesis of quantum dots, underscoring their tunable optoelectronic properties for applications in displays and . These milestones collectively transformed nanostructures from conceptual curiosities into a cornerstone of modern .

Types of Nanostructures

Zero-Dimensional Nanostructures

Zero-dimensional (0D) nanostructures are nanoscale materials confined in all three spatial dimensions, typically within 1 to 100 nm, resulting in point-like structures with isotropic confinement. This confinement distinguishes them from higher-dimensional nanostructures by producing fully quantized states rather than partial delocalization. Representative examples include quantum dots, such as (CdSe), and metallic nanoparticles, like (Au), both of which exhibit size-dependent behaviors unique to 0D systems. Colloidal synthesis methods dominate the of 0D nanostructures due to their ability to achieve precise control and monodispersity through solution-phase reactions. For CdSe quantum dots, the hot-injection —developed in seminal work by Bawendi and coworkers—involves injecting organometallic precursors into a heated coordinating , promoting rapid followed by controlled growth. This process allows tuning of particle diameters from 2 to 10 nm by adjusting reaction parameters like temperature and precursor concentration, yielding ensembles with narrow distributions essential for uniform properties. Similar colloidal approaches apply to Au nanoparticles, often using citrate reduction for stable, spherical particles. The defining physical properties of 0D nanostructures stem from quantum confinement, where charge carriers are restricted in all directions, leading to discrete electronic energy levels instead of continuous bands found in bulk materials. In quantum dots like CdSe, this manifests as a size-tunable bandgap, with the confinement energy \Delta E scaling inversely with the square of the radius r, as captured in the Brus equation: \Delta E = \frac{\hbar^2 \pi^2}{2 r^2} \left( \frac{1}{m_e^*} + \frac{1}{m_h^*} \right) - \frac{1.8 e^2}{\epsilon r}, where \hbar is the reduced Planck's constant, m_e^* and m_h^* are the effective masses of electrons and holes, e is the electron charge, and \epsilon is the constant; the dominant term for small r yields \Delta E \propto 1/r^2. For Au nanoparticles, quantum confinement enhances , enabling strong, tunable optical absorption peaks that shift with size from ~520 nm for 10 nm particles to longer wavelengths for larger ones. These characteristics position 0D nanostructures for emerging roles in and sensing, where size-dependent emission and plasmonic effects enable applications like light-emitting devices and bioimaging probes.

One-Dimensional Nanostructures

One-dimensional (1D) nanostructures are characterized by their elongated, wire-like geometry, where two dimensions are confined to the nanoscale (typically 1–100 nm), while the third dimension can extend to micrometers or longer, resulting in high aspect ratios often exceeding 1000:1. This linear morphology imparts anisotropic properties, such as directional and transport, distinguishing them from zero-dimensional structures that exhibit isotropic quantum confinement. Prominent examples of 1D nanostructures include carbon nanotubes (CNTs) and semiconductor nanowires. CNTs exist in single-walled (SWCNTs) and multi-walled (MWCNTs) forms, with SWCNTs consisting of a single sheet rolled into a seamless cylinder of ~1 nm diameter, while MWCNTs feature concentric layers. Semiconductor nanowires, such as (Si) and zinc oxide (ZnO), are solid crystalline rods with diameters of 5–100 nm and lengths up to several micrometers, enabling applications in and sensing due to their tunable conductivity. A key structural variation in SWCNTs arises from their , defined by the roll-up direction of the graphene lattice via the chiral vector (n,m), which determines whether the nanotube is metallic or semiconducting. Armchair (n=n, m=0) and certain chiral configurations exhibit metallic behavior with zero bandgap, while others are semiconducting; approximately two-thirds of SWCNTs are semiconducting based on random chirality distribution. For semiconducting SWCNTs, the bandgap E_g is inversely proportional to the d (in nm), approximated by the tight-binding model as E_g \approx \frac{0.8}{d} \quad (\text{eV}), leading to bandgaps of ~0.5–1.5 for typical diameters. The high aspect ratios of 1D nanostructures facilitate ballistic transport, where charge carriers travel without significant over mean free paths of tens to hundreds of micrometers at , particularly in metallic CNTs and clean nanowires. This phenomenon arises from reduced dimensionality, minimizing interactions and enabling quantum conductance quantized in units of $2e^2/h. In MWCNTs, has been observed with currents up to 1 mA per tube, limited primarily by rather than intrinsic . Template-assisted growth is a common fabrication method for 1D nanostructures, involving the use of nanoporous templates like anodic aluminum oxide (AAO) to confine material deposition and dictate uniform diameter and alignment. This bottom-up approach, often via or within template pores, yields high-density arrays of nanowires (e.g., or ZnO) with precise control over dimensions, facilitating scalable production for device integration.

Two-Dimensional Nanostructures

Two-dimensional nanostructures are atomically thin materials typically one or a few atomic layers thick, exhibiting sheet-like geometries where surface atoms dominate interactions and properties, leading to pronounced quantum confinement effects perpendicular to the plane. Unlike bulk materials, these structures display enhanced reactivity, tunable electronic band structures, and superior mechanical flexibility due to the prevalence of van der Waals bonding between layers and strong in-plane covalent interactions. This dimensionality imparts unique behaviors, such as high specific surface areas that amplify catalytic and sensing capabilities. Prominent examples include , a single layer of sp²-hybridized carbon atoms in a , first isolated in by mechanical cleavage from . dichalcogenides (TMDCs), such as (MoS₂), form layered crystals where individual monolayers consist of a plane sandwiched between layers, enabling facile exfoliation to yield 2D sheets with semiconductor characteristics. , a family of 2D carbides, nitrides, or carbonitrides (e.g., Ti₃C₂Tₓ), are derived from MAX phase precursors through selective etching, resulting in hydrophilic surfaces terminated with functional groups like -OH, -O, or -F that enhance dispersibility in aqueous media. Graphene exemplifies the mechanical prowess of 2D nanostructures, boasting an in-plane of approximately 1 TPa, which surpasses that of and enables applications in . Its electronic properties arise from charge carriers behaving as massless Dirac fermions near the Dirac points, governed by the linear E = \hbar v_F |k| where \hbar is the reduced Planck's constant, v_F \approx 10^6 m/s is the Fermi velocity, and k is the wavevector relative to the Dirac point; this relativistic-like spectrum yields ultrahigh exceeding 200,000 cm²/V·s at . In contrast, MoS₂ monolayers exhibit a direct bandgap of ~1.8 eV, facilitating strong , while demonstrate metallic conductivity up to 24,000 S/cm combined with pseudocapacitive due to their accordion-like layered . Isolation of 2D nanostructures often relies on mechanical exfoliation, a top-down approach using to repeatedly peel layers from bulk crystals, as initially demonstrated for to produce pristine, defect-free flakes up to micrometers in size. For scalable production, (CVD) enables epitaxial growth of continuous films; for instance, is synthesized by decomposing on foils at ~1000°C, yielding large-area sheets transferable to arbitrary substrates. Thickness profoundly influences properties: in MoS₂, single layers maintain a direct bandgap conducive to , whereas few-layer stacks (3–5 layers) transition to an indirect bandgap of ~1.2–1.5 eV, suppressing but enhancing charge transport; similar layer-dependent shifts occur in other TMDCs and , where interlayer coupling modulates conductivity and stability.

Three-Dimensional Nanostructures

Three-dimensional (3D) nanostructures represent hierarchical assemblies where nanoscale features extend across all three spatial dimensions, enabling volumetric integration of lower-dimensional building blocks such as nanoparticles, nanowires, or nanosheets into complex, bulk-like architectures. These structures often exhibit emergent properties arising from the ordered arrangement of their constituents, distinguishing them from isolated lower-dimensional forms by their ability to achieve macroscopic functionality through nanoscale precision. Prominent examples of nanostructures include nanocomposites, which disperse within a continuous matrix to form reinforced composites with enhanced mechanical and thermal properties. Aerogels, such as silica aerogels, exemplify ultra-lightweight porous networks formed by sol-gel processes followed by , achieving porosities exceeding 99% while maintaining structural integrity. Porous like mesoporous silica further illustrate this category, featuring ordered pore networks with diameters of 2-50 nm that facilitate high surface areas up to 1000 m²/g. Key architectures in nanostructures encompass core-shell configurations, where a nanoscale is encapsulated by a shell of different material, often assembled into larger superstructures for improved stability and multifunctionality; for instance, polymer-coated aerogels demonstrate mechanically robust core-shell assemblies. Dendrimers constitute highly branched, globular macromolecules with precise dendritic architectures, typically 1-10 in , synthesized via divergent or convergent methods to yield generations of repeating units around a central . Supercrystals, meanwhile, form as ordered lattices of colloidal nanocrystals, akin to atomic but on the nanoscale, with interparticle spacings of 1-10 that mimic symmetries like face-centered cubic arrangements. These architectures derive unique properties from their 3D organization, particularly enhanced that boosts catalytic efficiency by providing ample active sites and pathways; for example, mesoporous silica's tunable pores enable selective in reactions like cracking. Multifunctionality emerges from integrating lower-dimensional elements, such as embedding 1D nanowires into matrices to combine electrical with . In nanocomposites, this integration yields synergistic effects, like improved catalytic performance in porous carbon-metal hybrids due to stabilized dispersion. The assembly of nanostructures often relies on principles, where colloidal interactions drive the formation of lattices through evaporation-induced ordering or templating. Seminal work on supercrystal formation highlights how ligand-stabilized nanocrystals self-assemble into thermodynamically stable arrays, governed by van der Waals forces and electrostatic repulsion at the nanoscale. In synthesis, iterative branching promotes radial growth into compact shapes, while core-shell assemblies leverage sequential deposition to build hierarchical shells around cores. These bottom-up processes enable scalable production of ordered lattices with long-range periodicity, essential for applications in and sensing.

Synthesis Methods

Top-Down Fabrication Techniques

Top-down fabrication techniques involve the subtractive patterning of bulk materials to create nanostructures, typically through a combination of for defining patterns and for material removal. These methods draw from established manufacturing processes, enabling precise control over feature sizes down to the nanoscale. establishes a or on the surface, while subsequent transfers the pattern into the underlying material by selectively removing exposed regions. This approach is particularly suited for producing uniform, large-area nanostructures compatible with . Key lithographic techniques include , (EBL), and (NIL). employs light passed through a to expose a layer on the , creating soluble and insoluble regions that define the after ; resolution is diffraction-limited to around 100 nm but can reach 65-130 nm with deep variants and sub-10 nm with extreme systems. EBL uses a focused beam to directly write into an electron-sensitive resist without a , achieving resolutions below 10 nm due to the short de Broglie of , though it suffers from proximity effects where scattered blur nearby features. NIL, pioneered by Chou et al. in 1995, mechanically imprints a nanoscale into a polymer under heat or UV curing, replicating features as small as 10 nm with and throughput superior to EBL. These methods often produce one-dimensional nanowires or two-dimensional arrays by patterning lines or grids on substrates like . Etching processes, such as reactive ion etching (RIE) and focused ion beam (FIB) milling, are essential for transferring lithographic patterns into the bulk material. RIE combines chemical reactions from plasma-generated radicals with physical sputtering from accelerated ions, enabling anisotropic etching with aspect ratios exceeding 10:1 and resolutions around 10 nm, making it ideal for deep trenches in silicon nanostructures. FIB employs a finely focused gallium ion beam to sputter material directly, allowing maskless prototyping of features down to 5-10 nm but limited to small areas due to serial processing. These techniques evolved from microscale fabrication in the 1960s, driven by the semiconductor industry's push toward smaller features under Moore's law, transitioning to nanoscale precision in the late 1960s and 1970s with the advent of EBL and advanced etching for integrated circuits. Top-down methods offer significant advantages, including seamless compatibility with existing cleanroom infrastructure and the ability to produce defect-free, scalable nanostructures for , but they are hindered by high equipment costs—often millions of dollars for EBL systems—and low throughput for serial techniques like FIB, limiting their use to prototyping or low-volume production.

Bottom-Up Assembly Approaches

Bottom-up assembly approaches in nanostructure fabrication involve constructing materials from , molecular, or supramolecular building blocks, enabling precise control over at the nanoscale through or directed . These methods contrast with top-down techniques by relying on chemical and physical principles to build structures incrementally, often in solution or vapor phases, to achieve atomic-scale precision. Key techniques include (CVD), which decomposes gaseous precursors on a to grow nanostructures like carbon nanotubes and . In CVD for single-walled carbon nanotubes, hydrocarbons such as are catalytically decomposed over metal particles (e.g., iron or ) at temperatures around 700–1000°C, leading to tubular structures with diameters of 1–2 nm. Similarly, CVD on copper foils at 1000°C using enables large-area films, where carbon atoms diffuse across the metal surface and precipitate as a continuous 2D sheet upon cooling. For zero-dimensional nanoparticles, the sol-gel process hydrolyzes and condenses metal alkoxides in solution to form oxide networks, yielding uniform particles like silica nanoparticles with sizes tunable from 10–100 nm via control of precursor concentration and pH. techniques further exemplify bottom-up strategies; for instance, folds a long single-stranded DNA scaffold with short staple strands into custom 2D shapes (e.g., squares or stars) measuring 100 nm across, driven by base-pairing specificity at . Block copolymer , meanwhile, phase-separates immiscible blocks into periodic domains, such as cylinders or lamellae with 10–50 nm features, useful for templating inorganic nanostructures. The underlying mechanisms are governed by thermodynamic driving forces, primarily the minimization of surface , which favors compact structures over dispersed ones. In , this occurs via reversible non-covalent interactions like hydrogen bonding or van der Waals forces, allowing error correction during formation. For directed growth in CVD and sol-gel, kinetic factors such as precursor dictate the process; initiates when monomer concentration exceeds a critical , followed by diffusive attachment to growing clusters. Advantages of bottom-up assembly include atomic-level precision, potentially surpassing lithographic limits, and scalability through solution-based processing. Layer-by-layer deposition, for example, alternately adsorbs oppositely charged polyelectrolytes onto a , building films with nanometer-thick layers (e.g., 1–5 per bilayer) over large areas at ambient conditions, enabling multifunctional coatings like sensors or matrices. This method's versatility supports integration of diverse components, such as nanoparticles within matrices, without specialized equipment. For zero- and one-dimensional nanostructures, and are particularly critical. In (0D), the radius rate follows \frac{dR}{dt} \propto S - 1, where S is the , ensuring monodisperse sizes by separating (high S) from (low S) phases. For one-dimensional structures like nanowires or nanotubes, anisotropic favor along preferred crystallographic directions, with rates by catalysts that lower barriers, yielding lengths microns while maintaining nanoscale diameters. These principles enable high-yield production, as demonstrated in CVD where control via gas flow yields aligned nanotube arrays with densities exceeding 10^10 cm^{-2}.

Physical and Chemical Properties

Mechanical and Structural Properties

Nanostructures exhibit mechanical properties that differ markedly from their bulk counterparts due to size-dependent effects, where reduced dimensions lead to fewer defects and a higher proportion of surface atoms, which can enhance overall strength and elasticity through stiffening or softening depending on the material. In carbon nanotubes (CNTs), for instance, the absence of grain boundaries and dislocations results in tensile strengths approaching 150 GPa for multiwalled variants, far surpassing the ~1 GPa typical of high-strength steels, as measured via in-situ in a transmission electron microscope. This superior yield strength arises from the seamless cylindrical structure, minimizing sites for crack initiation. Similarly, two-dimensional (2D) materials like demonstrate exceptional elasticity, sustaining strains up to 25% before fracture while maintaining a of approximately 1 TPa, reflecting the robust sp² carbon bonding network. Structural properties at the nanoscale are dominated by atomic-scale arrangements, where distortions and defects such as vacancies and dislocations play amplified roles due to the high surface-to-volume ratio. Vacancies, for example, create local fields that distort the surrounding by up to several percent, altering propagation and mechanical stability in like metal oxides. Dislocations, often confined or partial in nanostructures, induce non-uniform gradients that can either strengthen the by impeding further slip or weaken it through localized concentrations, with effects scaling inversely with —becoming dominant below 10 . These defects are particularly pronounced in ion-implanted or synthesized nanostructures, where even low-dose causes persistent distortions observable via advanced . The scaling of in nanostructures exhibits size-dependent behavior, with surface effects potentially leading to either softening or stiffening of the effective stiffness; for ZnO nanowires, the modulus increases from ~140 GPa in bulk to ~220 GPa for 17 nm diameters, as determined by in-situ resonant testing in a electron microscope. Surface further influences this behavior, inducing intrinsic biaxial modeled by \sigma = \frac{\gamma}{h}, where \sigma is the , \gamma the surface , and h the thickness, leading to compressive or tensile strains in ultrathin films that modify elastic response. In 2D sheets, this equation highlights how atomic-scale thickness amplifies , potentially increasing modulus by 10-20% in freestanding . Mechanical and structural characterization of nanostructures relies on techniques like and (AFM), which probe local responses at the nanoscale. applies controlled loads via a tip to measure and through load-displacement curves, revealing, for example, the nonlinear elasticity of suspended membranes. AFM complements this by mapping surface topography and performing indentation with a sharp , enabling simultaneous visualization and quantification of defect-induced variations in stiffness across individual nanostructures.

Chemical Properties

Nanostructures often display enhanced chemical properties due to their high surface-to-volume ratio, which increases the proportion of surface atoms available for interactions. This leads to greater chemical reactivity compared to bulk materials, enabling applications in where nanoparticles exhibit superior activity and selectivity. For instance, metal nanoparticles like can show significantly higher catalytic efficiency for reactions such as oxidation due to undercoordinated surface sites. Additionally, the altered electronic structure from quantum confinement can modify chemical bonding and stability, influencing rates and in biological environments.

Electronic and Optical Properties

Nanostructures exhibit distinctive electronic properties arising from quantum confinement effects, where the spatial restriction of charge carriers to dimensions comparable to their de Broglie wavelength results in discrete levels rather than continuous bands observed in bulk materials. In zero-dimensional (0D) structures like quantum dots, this confinement quantizes states into atomic-like shells, altering the band and enabling phenomena such as single-electron charging. For one-dimensional (1D) nanostructures, such as nanowires or nanotubes, the deviates significantly from the bulk parabolic form; in 1D systems with parabolic dispersion, the g(E) follows g(E) \propto \frac{1}{\sqrt{E}}, leading to enhanced carrier densities near the band edges and van Hove singularities. Charge transport in these low-dimensional systems is dominated by quantum tunneling rather than classical , particularly in 0D and 1D configurations. In quantum dots, the effect manifests as a suppression of at low biases due to the discrete charging E_c = e^2 / 2C, where C is the dot , requiring an external bias to overcome the barrier for sequential electron tunneling. This quantization enables precise control over electron occupancy, with applications in single-electron transistors. In contrast to bulk semiconductors, where limits mean free paths to nanometers, 1D carbon nanotubes support over micrometer lengths at , as electrons traverse the structure without significant , yielding conductances quantized in units of $2e^2 / h. Optically, nanostructures display enhanced light-matter interactions due to these confinement effects. In metallic nanoparticles, resonances arise from collective oscillations of conduction electrons, with the resonance exhibiting dependence; for small particles, approximations show \omega \propto 1 / \sqrt{r} in certain quantum-corrected models accounting for electron spill-out and confinement. quantum dots, meanwhile, allow tuning of emission wavelength through variation, as quantum confinement increases the effective bandgap according to the particle-in-a-box model, shifting emission from to visible as diameters decrease from 10 nm to 2 nm. This tunability stems from the inverse relationship between confinement energy and nanostructure , enabling color-selective emitters with narrow linewidths.

Characterization Techniques

Imaging and Microscopy Methods

Scanning electron microscopy (SEM) is a widely used technique for visualizing the surface morphology and topography of nanostructures, operating by scanning a focused beam of electrons across a sample to generate signals such as secondary electrons and backscattered electrons that form high-resolution images. The resolution of SEM typically reaches down to 1-10 nm, enabling detailed examination of nanostructure shapes, sizes, and surface features like pores or aggregations in materials such as nanoparticles or nanowires. In applications, SEM excels at mapping the three-dimensional topography of nanostructures, providing insights into their assembly and defects, which is crucial for understanding structural properties in one-, two-, and three-dimensional forms. Transmission electron microscopy (TEM) offers superior for internal structural analysis of nanostructures, transmitting a high-energy electron beam through ultra-thin samples to produce two-dimensional images based on transmitted electrons, with resolutions approaching 0.1 nm under optimal conditions. This high resolution allows of arrangements, fringes, and defects within nanostructures, such as planes in quantum dots or layer stacking in sheets. Scanning TEM (), a variant of TEM, utilizes a focused probe to scan the sample, where contrast mechanisms like Z-contrast—arising from the (Z) dependence of —enable differentiation of elements based on brightness proportional to Z², facilitating imaging of compositional interfaces in hybrid nanostructures. Atomic force microscopy (AFM) provides nanoscale surface profiling by raster-scanning a sharp probe over the sample, measuring forces between the tip and surface to generate topographic maps with sub-nanometer vertical and lateral resolution down to ~1 nm. Operating in modes such as contact, tapping, or non-contact, AFM is particularly suited for delicate nanostructures, revealing surface roughness, height profiles, and mechanical variations in materials like carbon nanotubes or thin films without requiring vacuum conditions. These techniques support in-situ imaging capabilities to observe dynamic processes in nanostructures, such as , deformation, or transitions under environmental stimuli like temperature or , integrating , TEM, or AFM holders that maintain resolution while allowing real-time monitoring. For instance, in-situ TEM can capture atomic-scale movements during loading, linking to structural evolution. Advancements in aberration-corrected TEM, developed since the early , have dramatically improved by compensating for aberrations, achieving sub-angstrom and enabling direct of light elements and beam-sensitive nanostructures with reduced damage. This correction enhances contrast and signal-to-noise ratios, allowing precise mapping of atomic positions in complex nanostructures like perovskites or heterostructures.

Analytical and Spectroscopic Techniques

Analytical and spectroscopic techniques play a crucial role in elucidating the , bonding environments, and structures of nanostructures, enabling the identification of surface modifications, distributions, and defect states that influence their properties. These methods provide non-destructive or minimally invasive probes at the and molecular levels, complementing approaches by focusing on spectroscopic signatures rather than alone. X-ray photoelectron spectroscopy (XPS) is widely employed for surface-sensitive analysis of nanostructures, typically probing the top 5–10 nm of a sample by measuring the kinetic energies of photoelectrons emitted upon irradiation. The of core-level electrons shifts due to chemical environments, allowing differentiation of oxidation states and coordination in ; for instance, in oxide nanoparticles, shifts in the O 1s peak reveal surface or doping effects. via peak fitting yields elemental ratios and depth profiles when combined with angle-resolved measurements, essential for understanding reactivity in catalytic nanostructures. Raman spectroscopy offers insights into vibrational and rotational modes, facilitating the characterization of lattice dynamics and defects in nanostructures through inelastic light scattering. In two-dimensional materials like , the G band at approximately 1580 cm⁻¹ arises from in-plane sp² carbon stretching vibrations, while the D band at around 1350 cm⁻¹ signals breathing modes of sp³-hybridized six-atom rings associated with defects or edges. The ratio of D to G peak intensities (I_D/I_G) serves as a quantitative metric for defect density, with higher ratios indicating increased disorder from or chemical functionalization. This technique is particularly valuable for non-destructive, in-situ monitoring of nanostructures during or processing. Energy-dispersive X-ray (EDX) spectroscopy, often integrated with scanning electron microscopy (SEM) or transmission electron microscopy (TEM), detects characteristic X-rays generated by electron beam interactions to map compositions spatially. In SEM-EDX, resolutions of 1–5 µm suffice for nanomaterial distributions, whereas TEM-EDX achieves sub-nanometer for core-shell structures, such as segregation in nanoparticles. Quantitative mapping corrects for effects using standards, revealing inhomogeneities like clustering that affect electronic properties. Defect detection is indirect through associated variations, such as oxygen enrichment at imperfections. Synergistic use of (EELS) within TEM enhances analytical capabilities by analyzing energy losses of transmitted electrons, providing both elemental and electronic structure data at atomic scales. Core-loss EELS edges enable fine-structure analysis for bonding states, while low-loss regions detect plasmons or excitons indicative of defects; for example, in carbon nanotubes, shifts in the π* peak near 285 eV signal sp² defect disruptions. Combining EELS with TEM imaging yields correlative maps of composition and defects, such as vacancy sites in materials, with quantification via cross-sections for concentrations down to single atoms. This integration is pivotal for validating nanostructure integrity in device applications.

Applications

In Electronics and Optoelectronics

Nanostructures have revolutionized by enabling devices with enhanced performance at the nanoscale, where quantum effects and high surface-to-volume ratios play crucial roles. In technology, nanowires serve as channels in field-effect transistors (FETs), offering superior electrostatic control and reduced short-channel effects compared to bulk devices. For instance, gate-all-around nanowire FETs achieve sub-10 nm gate lengths while maintaining high on/off current ratios exceeding 10^6, allowing for faster switching speeds up to 100 GHz. These advancements stem from the nanowires' ability to confine carriers effectively, drawing on their electronic properties such as tunable bandgaps. Carbon nanotube (CNT)-based interconnects represent another key application, addressing the limitations of in scaling down integrated circuits by providing lower resistance and higher current densities. Single-walled CNTs exhibit ballistic transport with mean free paths over 1 μm, reducing interconnect resistance by up to 50% compared to at 22 nm nodes, which enhances power efficiency in . In flexible electronics, nanostructures enable bendable transistors and circuits due to their exceptional mechanical flexibility and carrier mobility exceeding 100,000 cm²/V·s in two-dimensional () FETs. These properties allow devices to withstand bending radii as small as 1 mm without performance degradation, paving the way for wearable and foldable . Optoelectronics benefits from nanostructures through improved light-matter interactions, particularly in light-emitting diodes (LEDs) and sensors. LEDs (QD-LEDs) utilize colloidal nanocrystals, such as CdSe or InP, to achieve narrow emission spectra with below 30 nm and external quantum efficiencies surpassing 20%, enabling vibrant displays with lower power consumption. Plasmonic nanostructures, including or silver nanoparticles, enhance sensor sensitivity by localizing electromagnetic fields, achieving detection limits down to single molecules in surface-enhanced (SERS) platforms. In solar cells, nanostructures like nanowires and nanoparticles facilitate light trapping and charge separation, boosting efficiency. Silicon nanowire arrays in photovoltaic devices increase via multiple , yielding power conversion efficiencies up to 15% in thin-film configurations, a significant gain over planar counterparts. nanostructures further exemplify this, with quantum dots enhancing stability and efficiency to over 18% by reducing recombination losses. These optoelectronic applications underscore nanostructures' role in compact, efficient devices for displays, sensing, and in .

In Biomedicine and Energy

Nanostructures have revolutionized biomedicine by enabling targeted drug delivery, enhanced imaging, and precise therapies, leveraging their unique size-dependent properties to interact effectively with biological systems. Liposomes, phospholipid-based nanoparticles, serve as versatile carriers for encapsulating hydrophilic and hydrophobic drugs, improving bioavailability and reducing systemic toxicity through controlled release mechanisms. For instance, liposomal formulations like Doxil have been FDA-approved for cancer treatment, demonstrating prolonged circulation times and site-specific delivery via the enhanced permeability and retention effect. Biocompatibility is a key factor in these applications, influenced by factors such as surface charge, size, and composition, which determine cellular uptake and minimal immune response; neutral or zwitterionic coatings on liposomes enhance stealth properties, mitigating opsonization and extending half-life in vivo. In , nanostructures act as agents to improve and specificity in techniques like MRI and , where superparamagnetic nanoparticles provide T2-weighted for tumor detection, offering higher sensitivity than traditional chelates. nanoparticles and quantum dots further enable multimodal , combining optical and capabilities for real-time monitoring of progression. Targeted therapies exploit nanostructures' optical properties, such as nanorods, which absorb near-infrared light to generate localized heat for photothermal ablation of cancer cells, achieving up to 90% tumor reduction in preclinical models with minimal off-target effects. In energy applications, nanostructures enhance storage and conversion efficiency in devices like batteries and fuel cells. Silicon nanowires as anodes in lithium-ion batteries address volume expansion issues during lithiation, delivering specific capacities exceeding 3000 mAh/g—over ten times that of —while maintaining structural integrity through one-dimensional morphology that facilitates ion diffusion. nanoparticles in fuel cells reduce catalyst loading to below 0.1 mg/cm² by maximizing surface area for the , improving and durability under operational conditions. Dye-sensitized solar cells benefit from TiO₂ nanotube arrays, which provide oriented transport paths, boosting conversion efficiencies to over 7% compared to nanoparticle counterparts, due to reduced recombination and enhanced dye loading. Recent advances in the 2020s have focused on nanostructures for , where nanostructured halide perovskites in tandem cells achieve certified efficiencies above 33%, surpassing single-junction limits through improved light management and defect passivation. These 2D/3D hybrid nanostructures enhance stability against moisture and heat, enabling scalable fabrication via solution processing for cost-effective harvesting.

Challenges and Future Prospects

Safety and Environmental Concerns

Nanostructures, due to their small size and high surface area, can exhibit that varies with particle dimensions and morphology, influencing cellular uptake and subsequent biological effects. For instance, rod-like gold nanoparticles demonstrate higher to cells compared to spherical ones of similar size, primarily because of differences in cellular internalization pathways such as . Similarly, cellular uptake of gold nanoparticles is size- and shape-dependent, with larger spherical particles (e.g., 50 nm) showing enhanced delivery potential but also varying across cell types, including cancer cells. In carbon nanotubes (CNTs), often arises from (ROS) generation, which interacts with cellular components like mitochondria and membranes, leading to and potential DNA damage. This ROS-mediated mechanism contributes to broader cytotoxic effects observed in various cell models exposed to CNTs. Environmentally, nanostructures like silver nanoparticles (AgNPs) pose risks through persistence and in aquatic systems, where they can accumulate in organisms such as , disrupting microbial communities and food chains. Their high reactivity allows AgNPs to release silver ions that exhibit extreme toxicity to non-target in water streams, exacerbating ecological contamination. Regulatory frameworks address these concerns through specific guidelines for . Under the European Union's REACH regulation, nanomaterials are defined as chemical substances or materials comprising very small particles of different shapes and sizes no larger than 100 nm and must be registered with detailed information on nanoforms, including hazard assessments, as updated in 2020 to harmonize reporting across substances. In the United States, the Environmental Protection Agency (EPA) has conducted post-2010 assessments via case studies on nanoscale silver and carbon nanotubes, evaluating environmental fate, exposure, and risk to inform Toxic Substances Control Act (TSCA) reviews for new chemical notices. These efforts highlight the need for tailored risk evaluations due to unique nanomaterial properties. To mitigate toxicity, surface coatings are employed to reduce nanoparticle reactivity and cellular interactions. For example, polymer coatings on silver nanoparticles prevent ion release and associated toxicity in biological systems, while EDTMP coatings on metal oxide nanoparticles passivate surfaces, lowering pulmonary inflammation and cytotoxicity in rodent models. Hydrophobic coatings on zinc oxide nanoparticles similarly decrease environmental toxicity by altering surface charge and limiting ROS production. Such modifications enhance biocompatibility without compromising functionality in applications like biomedicine. Scalability remains a primary hurdle in nanostructure production, particularly in transitioning from laboratory-scale synthesis to industrial volumes. In bottom-up approaches, such as chemical vapor deposition (CVD), achieving uniform deposition across large areas is challenging due to variations in gas flow, temperature gradients, and precursor distribution, leading to inconsistencies in nanostructure size, morphology, and quality. For instance, in the CVD synthesis of vertically aligned carbon nanotubes, geometric nonuniformities arise from catalytic inefficiencies and reactor non-idealities, often resulting in low material utilization. These issues are exacerbated in scaling up, where reactor design limitations hinder repeatability and increase defect rates in 2D materials like graphene or transition metal dichalcogenides. Top-down methods, including lithography and etching, face significant cost barriers stemming from the need for expensive, precision equipment like electron beam systems, which limit throughput and drive up production expenses— for example, the cost of gold nanoparticles can exceed $80,000 per gram due to complex processing, far surpassing raw material prices. Emerging trends are addressing these challenges through innovative and strategies. Post-2020 advancements in AI-optimized have enabled predictive modeling of nanostructure properties, accelerating the of optimal synthesis parameters and reducing trial-and-error in production; for example, algorithms now guide the assembly of carbon nanomaterials and metallic nanoparticles for enhanced uniformity and efficiency. nanostructures, combining materials like 0D quantum dots with 2D sheets or 1D nanowires, are gaining prominence for their synergistic properties, such as improved charge transfer and multifunctionality in and sensing. In quantum computing, 0D nanostructures like quantum dots serve as , leveraging quantum confinement for spin manipulation and coherence times exceeding microseconds, with configurations integrating dots into photonic or superconducting platforms to enable scalable qubit arrays. Looking ahead to 2025 and beyond, neuromorphic devices based on materials like MoS₂ and WSe₂ promise energy-efficient architectures, with 3D-integrated memristor-transistor arrays achieving switching energies as low as 2.9 pJ/bit and footprints of 3–4.5 F², facilitating brain-like for applications. Sustainable synthesis methods are also evolving, utilizing renewable resources such as extracts and solvents to minimize energy use and waste; for instance, bio-mediated approaches yield nanoparticles with controlled sizes while reducing environmental impact compared to traditional chemical routes. Economically, the market is projected to reach approximately $105 billion in 2025, driven by demand in and healthcare, with further growth to over $200 billion by the early 2030s through scalable innovations.

References

  1. [1]
    [PDF] One-Dimensional Nanostructures: Synthesis, Characterization, and ...
    Nanostructures–structures that are defined as having at least one dimension between 1 and 100 nm–have received steadily growing interests as a result of their ...
  2. [2]
    Chapter 1: Introduction to Nanostructures - AIP Publishing
    Nanostructures generally refer to the material systems that are in the range of 1 to 100 nanometers. In a nanostructure, electrons are normally confined in ...
  3. [3]
    [PDF] Introduction to Nanoparticles and Nanostructures - nanoHUB
    • Definition and Examples. • Physical Properties. • Optical Properties ... If the dimensions of a nanostructure are smaller than the electron MFP, then surface ...
  4. [4]
    Nanomaterials: Synthesis and Applications in Theranostics - NIH
    The nanomaterials can be classified as 0D, 1D, 2D, and 3D based on their dimensions. The nanomaterials can be malleable and ductile and they can be drawn into ...2.6. Chemical... · 4. One Dimensional (1d)... · Figure 4<|control11|><|separator|>
  5. [5]
    A Review on Low-Dimensional Nanomaterials: Nanofabrication ...
    Three-dimensional nanomaterials can be constructed based on the arrangement and organization of a group of 0D, 1D, or 2D constituent nanostructures.
  6. [6]
    Nanomaterials: a review of synthesis methods, properties, recent ...
    Feb 24, 2021 · This review discusses a brief history of nanomaterials and their use throughout history to trigger advances in nanotechnology development.
  7. [7]
    Nanostructures: Introduction | Chemical Reviews - ACS Publications
    The term nano does not have a precise definition, and we prefer to leave it that way. We include molecular structures that are in the range of 5 nm, as well ...<|control11|><|separator|>
  8. [8]
    Emerging 0D, 1D, 2D, and 3D nanostructures for efficient point-of ...
    Nanomaterials are categorized as zero-dimensional (0D) (for example, nanoparticles), one-dimensional (1D) (for example, nanotubes & nanorods), two-dimensional ( ...
  9. [9]
    Nanostructured materials - NASA ADS
    Nanostructured materials may be defined as those materials whose structural elements - clusters, crystallites or molecules - have dimensions in the 1 to 100 nm ...
  10. [10]
    Review on nanoparticles and nanostructured materials: history ... - NIH
    The aim of this review is to compare synthetic (engineered) and naturally occurring nanoparticles (NPs) and nanostructured materials (NSMs) to identify their ...
  11. [11]
    Nanostructure - an overview | ScienceDirect Topics
    A nanostructure is a structure of intermediate size between microscopic and molecular structures. In describing nanostructures, it is necessary to differentiate ...
  12. [12]
    Nanoparticle classification, physicochemical properties ...
    Jun 7, 2022 · Larger surface areas and larger surface-to-volume ratios generally increases the reactivity of nanomaterials due to the larger reaction surface ...
  13. [13]
    Nanomaterials: Surface Area to Volume Ratio - OMICS International
    This equation shows that as the radius of a nanoparticle decreases, the surface area increases exponentially relative to its volume [4]. For example, if the ...
  14. [14]
    Types of Nanostructures - ScienceDirect.com
    Nanostructures and nanomaterials can show exceptional properties that are dissimilar to those of equivalent chemical compounds with larger dimensions.
  15. [15]
  16. [16]
  17. [17]
  18. [18]
    Quantum Confinement in Size-Selected, Surface-Oxidized Silicon ...
    Size-selective precipitation and size-exclusion chromatography cleanly separate the silicon nanocrystals from larger crystallites and aggregates.
  19. [19]
  20. [20]
    From Nanotech to Nanoscience - Science History Institute
    Jun 23, 2008 · Medieval artisans discovered through alchemical experimentation that adding gold chloride to molten glass resulted in a red tint, and adding ...
  21. [21]
    Experimental relations of gold (and other metals) to light - Journals
    The Bakerian Lecture. —Experimental relations of gold (and other metals) to light. Michael Faraday.
  22. [22]
    Michael Faraday's recognition of ruby gold: the birth of modern ...
    Michael Faraday's recognition of ruby gold: the birth of modern nanotechnology. His 1857 lecture to the royal society in London.
  23. [23]
    Nanotechnology Through History: Carbon-based Nanoparticles from ...
    Jun 17, 2013 · Since our early ancestors first learned to make fires, humans have been producing carbon-based nanoparticles. The smoke and soot from their ...
  24. [24]
    Electron Microscopy | SpringerLink
    May 23, 2023 · Ruska and Max Knoll built the first transmission electron microscope (TEM). By the end of the 1930s TEMs reached a resolution limit of 10 nm and ...
  25. [25]
    [PDF] There's Plenty of Room at the Bottom - Caltech Magazine
    "There's Plenty of Room at the Bottom" is a transcript of a talk given by Dr. Feynman on December 29 at the annual meeting of the American Physical Society ...
  26. [26]
    C60: Buckminsterfullerene - Nature
    Nov 14, 1985 · Nature volume 318, pages 162–163 (1985)Cite this article. 68k ... H. W. Kroto, J. R. Heath, S. C. O'Brien, R. F. Curl & R. E. Smalley.
  27. [27]
    Press release: The 1996 Nobel Prize in Chemistry - NobelPrize.org
    New forms of the element carbon – called fullerenes – in which the atoms are arranged in closed shells was discovered in 1985 by Robert F. Curl, Harold W. Kroto ...
  28. [28]
    Helical microtubules of graphitic carbon - Nature
    Nov 7, 1991 · Synthesis of carbon nanotubes. On 7 November 1991, Sumio Iijima announced in Nature the preparation of nanometre-size, needle-like tubes of ...Missing: original | Show results with:original
  29. [29]
    [PDF] The Initiative and its Implementation Plan
    Nov 23, 1993 · The President's Committee of Advisors on Science and Technology (PCAST) strongly endorsed the establishment of the NNI, beginning in Fiscal Year ...
  30. [30]
    Electric Field Effect in Atomically Thin Carbon Films - Science
    Oct 22, 2004 · Graphene is the name given to a single layer of carbon atoms densely packed into a benzene-ring structure, and is widely used to describe ...
  31. [31]
    What are Zero-Dimensional Nanomaterials? - AZoNano
    Oct 10, 2025 · 0D nanomaterials are defined by their confinement: in the zeroth dimension, electrons are restricted in all three spatial directions. Common ...Missing: colloidal | Show results with:colloidal
  32. [32]
    Quantum Dot Nanoparticles - an overview | ScienceDirect Topics
    Quantum dots nanoparticles are defined as zero-dimensional semiconductor structures with dimensions smaller than the Bohr radius, primarily composed of ...
  33. [33]
    The Five Methods for Synthesizing Colloidal Quantum Dots - AZoNano
    Jan 23, 2024 · This article from Merck discusses the five different methods used to synthesize monodisperse colloidal quantum dots.
  34. [34]
    [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 ...
  35. [35]
    First-Principles Investigation of the Structure and Properties of Au ...
    Jul 26, 2019 · We present a first-principles investigation of the structure, stability, and reactivity of Au nanoparticles (NPs) supported on ZnO.
  36. [36]
    Quantum Dots and Their Multimodal Applications: A Review - PMC
    Quantum confinement effects are observed when the size is sufficiently small that the energy level spacing of a nanocrystal exceeds kT (where k is Boltzmann's ...Missing: ∝ r²
  37. [37]
    Introduction: 1D Nanomaterials/Nanowires | Chemical Reviews
    Aug 14, 2019 · A review. This article provides a comprehensive review of current research activities that conc. on 1-dimensional (1D) nanostructures-wires ...
  38. [38]
    Structure and Electronic Properties of Carbon Nanotubes
    Thus to first order, zigzag (n,0) or chiral (n,m) SWNTs are metallic when (n − m)/3 is an integer and otherwise semiconducting. Figure 2 (a) Three-dimensional ...Missing: formula | Show results with:formula
  39. [39]
    A comprehensive review of template-synthesized multi-component ...
    In this article, a comprehensive review of template-synthesized multi-component nanowire structures has been given, including multi-segmented, core-shell and ...
  40. [40]
    Hierarchical self-assembly of 3D lattices from polydisperse ... - Nature
    Apr 18, 2019 · We show that our prototypical system of anisometric silver plates with a high polydispersity assemble, unexpectedly, into an ordered, three-dimensional lattice.
  41. [41]
    Three-Dimensional Core-Shell Superstructures - ACS Publications
    Silica aerogels are mesoporous materials (5) made by supercritical fluid (SCF) drying of wet gels. (6-9). The latter are prepared via the so-called sol–gel ...Missing: dendrimers supercrystals
  42. [42]
    Dendrimers: Exploring Their Wide Structural Variety and Applications
    Dendrimers are hyper-branched macromolecules characterized by large numbers of end-group functionalities and a compact molecular structure.
  43. [43]
    Achieving Functionality and Multifunctionality through Bulk and ...
    Feb 9, 2023 · This Feature Article focuses on the different properties of 3DOM materials that provide functionality, including a relatively large surface area, the ...
  44. [44]
    3D Hierarchically ordered porous carbon entrapped Ni ...
    Mar 2, 2021 · A highly active catalyst was prepared by entrapping Ni nanoparticles (NPs) inside three-dimensional hierarchically ordered porous carbon (3D HOPC).3. Results And Discussion · 3.2. Characterization Of... · 3.3. Catalytic Activity
  45. [45]
    Emulsion-confined self-assembly of colloidal nanoparticles into 3D ...
    Self-assembly of faceted binary supercrystals. (F) Schematic illustration of self-assembly of faceted binary nanocrystal supercrystals within emulsion droplets.
  46. [46]
    Advances in lithographic techniques for precision nanostructure ...
    Dec 11, 2023 · Photolithography (PL) is a top-down fabrication technique in which a substance known as a photoresist is exposed to light in certain regions, ...
  47. [47]
    An In-Depth Look at Top Down Nanofabrication - AZoNano
    Dec 21, 2018 · Reactive Ion Etching (RIE). Reactive ion etching is a form of dry etching, i.e., a method that uses a bombardment of ions in the form a ...
  48. [48]
    Synthesis of nanomaterials using various top-down and bottom-up ...
    In this study, the Top-down and Bottom-up methods and techniques of synthesizing nanostructured materials are reviewed, compared, and analyzed.
  49. [49]
    Fabrication of Nanostructures with Bottom-up Approach and Their ...
    The top-down approach involves lateral patterning of bulk materials by either subtractive or additive methods to realize nano-sized structures. Several methods ...
  50. [50]
    Large Scale CVD Synthesis of Single-Walled Carbon Nanotubes
    Chemical vapor deposition of hydrocarbons over metal catalysts has been a classical method to produce carbon materials. Various forms of carbon fibers, ...Missing: seminal | Show results with:seminal
  51. [51]
    The sol-gel process | Chemical Reviews - ACS Publications
    From Gel to Crystal: Mechanism of HfO2 and ZrO2 Nanocrystal Synthesis in ... Nanoparticles in Autofluorescence-Free Biosensing and Bioimaging: A Review.
  52. [52]
    Folding DNA to create nanoscale shapes and patterns - Nature
    Mar 16, 2006 · Here I describe a simple method for folding long, single-stranded DNA molecules into arbitrary two-dimensional shapes.
  53. [53]
    Block Copolymer Based Nanostructures: Materials, Processes, and ...
    We organize this review as follows: a brief introduction to block copolymers, their hybrids, and self-assembly behavior is presented in section . In section , ...
  54. [54]
    Mechanisms of Nucleation and Growth of Nanoparticles in Solution
    If the supersaturation is high such that r̅/r* ≥ 2 then d(Δr)/dt ≤ 0 and the growth of the system is self-sharpening of the size distribution. However, if r̅/r* < ...
  55. [55]
    Layer-by-layer assembly as a versatile bottom-up nanofabrication ...
    Layer-by-layer (LbL) assembly is an easy, inexpensive technique for multilayer formation, allowing various materials to be incorporated, making it a versatile ...Missing: seminal papers
  56. [56]
    Quantum confinement in Si and Ge nanostructures - AIP Publishing
    Jan 6, 2014 · The DOS illustrates that a change in the confinement dimension directly changes the energy occupation level. Thus, a modification in the DOS ...Missing: seminal | Show results with:seminal
  57. [57]
    Quantum Confinement in Si and Ge Nanostructures - arXiv
    Nov 8, 2011 · We apply perturbative effective mass theory as a broadly applicable theoretical model for quantum confinement (QC) in all Si and Ge nanostructures.Missing: seminal | Show results with:seminal<|separator|>
  58. [58]
    [PDF] Density of States: - Georgia Institute of Technology
    Dec 17, 2005 · Derivation of Density of States (1D). For calculating the density of states for a 1D structure (i.e. quantum wire), we can use a similar ...
  59. [59]
    [cond-mat/0103008] Quantum Effects in Coulomb Blockade - arXiv
    Mar 1, 2001 · The review starts with a description of an isolated quantum dot. We discuss the status of the Random Matrix theory (RMT) of the one-electron ...
  60. [60]
    Room Temperature Ballistic Conduction in Carbon Nanotubes
    Multiwalled carbon nanotubes are shown to be ballistic conductors at room temperature, with mean free paths of the order of tens of microns.Introduction · Results · Summary and Conclusion · Appendix
  61. [61]
    Bounds on quantum confinement effects in metal nanoparticles
    Mar 26, 2018 · In this paper, we demonstrate that bounds can be placed on the additional effects due to quantum confinement. The optical behavior of large ...Missing: seminal | Show results with:seminal
  62. [62]
    Tuning the Photoluminescence of Graphene Quantum Dots through ...
    Dec 28, 2012 · In this study, we observe that the photoluminescence (PL) of the GQDs shifts due to charge transfers between functional groups and GQDs.
  63. [63]
    Scanning Electron Microscopy: Principle and Applications in ...
    It passes electrons through an ultra-thin sample to provide two-dimensional imaging of internal structures, such as shape, crystallinity, heterogeneity, ...
  64. [64]
    Microscopic Techniques for Nanomaterials Characterization
    Jan 8, 2025 · “ Scanning Electron Microscopy: Principle and Applications in Nanomaterials Characterization.” In Handbook of Materials Characterization ...Scanning Electron Microscopy · Transmission Electron... · Atomic Force Microscopy
  65. [65]
    Scanning Electron Microscopy: Principle and Applications in ...
    Scanning electron microscopy (SEM) is an important electron microscopy technique that is capable of achieving a detailed visual image of a particle with ...
  66. [66]
    Applications of Transmission Electron Microscopy in Phase ...
    Aug 29, 2023 · We review notable applications of TEM in PEN research, including applications in fields such as metallic nanostructures, carbon allotropes, low-dimensional ...Introduction · Various TEM-Based Techniques · Applications of TEM in PEN
  67. [67]
    Transmission Electron Microscopy of Nanostructures - ScienceDirect
    Finally, an overview of applications of transmission electron microscopy (TEM) for studying nanostructures is given, with emphasis on defect, thin layer, ...
  68. [68]
    Structure determination through Z-contrast microscopy - ScienceDirect
    This review outlined the quantum mechanical basis for regarding Z-contrast imaging and EELS in the STEM as directly interpretable, column-by-column imaging ...
  69. [69]
    Atomic force microscopy for nanoscale mechanical property ...
    Nov 10, 2020 · In this review, we focus on AFM as a nanoscale mechanical property characterization tool and examine various AFM contact and intermittent contact modes.
  70. [70]
    (PDF) Atomic Force Microscopy for Surface Imaging and ...
    Aug 7, 2025 · This chapter presents an overview of Atomic Force Microscopy (AFM) principles followed by details on AFM instrumentation.
  71. [71]
    In situ microscopy techniques for characterizing the mechanical ...
    Three in situ microscopy techniques of AFM, SEM and TEM were used to study the mechanical properties and nanomechanical behavior of two-dimensional (2D) ...
  72. [72]
    Advances on in situ TEM mechanical testing techniques - Frontiers
    In this review, we summarize recent advances on in situ TEM mechanical characterization techniques, including classical tension holders, nanoindentation ...
  73. [73]
    Aberration-corrected scanning transmission electron microscopy for ...
    Mar 13, 2014 · The availability of aberration correctors in HRTEM also led to exciting developments, in particular the so-called NCSI technique for which ...
  74. [74]
    Possibilities and limitations of advanced transmission electron ... - NIH
    Jul 16, 2015 · A major revolution for electron microscopy in the past decade is the introduction of aberration correction, which enables one to increase ...
  75. [75]
    Leveraging Machine Learning for Advanced Nanoscale X-ray Analysis
    Aug 6, 2024 · Energy dispersive X-ray (EDX) spectroscopy in the transmission electron microscope is a key tool for nanomaterials analysis, providing a ...
  76. [76]
    Prospects for detecting individual defect centers using spatially ...
    Oct 14, 2019 · Electron energy loss spectroscopy (EELS) provides a way to detect point defects via their characteristic spectroscopic signals, allowing the ...Missing: nanostructures | Show results with:nanostructures
  77. [77]
    A Review of Liposomes as a Drug Delivery System - PubMed Central
    Feb 17, 2022 · Liposomes have been considered promising and versatile drug vesicles. Compared with traditional drug delivery systems, liposomes exhibit better properties.
  78. [78]
    Engineering precision nanoparticles for drug delivery - Nature
    Dec 4, 2020 · In this Review, we discuss advanced nanoparticle designs utilized in both non-personalized and precision applications that could be applied to improve ...<|separator|>
  79. [79]
    Biocompatibility of nanomaterials and their immunological properties
    Polymers, long chains of repeating monomers, are widely used in nanomedicine due to their biocompatible and biodegradable properties. Polymeric NMs are ...
  80. [80]
    Biocompatibility and nanostructured materials: applications in ...
    Surface charge. Surface charge of the nanoparticles is another important factor which can affect biocompatibility. The zeta potential is commonly used for ...
  81. [81]
    Nanomaterial-based contrast agents - PMC - NIH
    Contrast agents are often needed to accompany anatomical data with functional information or to provide phenotyping of the disease in question.
  82. [82]
    Nanotechnology in medical imaging: probe design and applications
    Nanoparticles have become more and more prevalent in reports of novel contrast agents, especially for molecular imaging, the detection of cellular processes.
  83. [83]
    Gold nanoparticles and gold nanorods in the landscape of cancer ...
    Jun 21, 2023 · The photothermal property of nanoparticles, especially of gold nanorods, causes absorption of the light incident by the light source, and ...Missing: nanostructures | Show results with:nanostructures
  84. [84]
    A critical review of silicon nanowire electrodes and their energy ...
    This review provides a comprehensive evaluation of the Li-storage capacity of various Si nanowire electrodes based on both specific and areal capacity.
  85. [85]
    Nanostructured silicon for high capacity lithium battery anodes
    Nanostructured silicon is promising for high capacity anodes in lithium batteries. The specific capacity of silicon is an order of magnitude higher than that ...
  86. [86]
    Sulfur-anchoring synthesis of platinum intermetallic nanoparticle ...
    Oct 21, 2021 · Yang et al. show that sulfur-doped carbon supports create strong platinum-sulfur bonds that stabilize small platinum alloy nanoparticles.
  87. [87]
    TiO2 nanotubes in dye-sensitized solar cells: Higher efficiencies by ...
    ... solar cell efficiency of 5.2%. This represents in fact the highest efficiency reported up to now for a purely TiO2 nanotube based solar cell. For mixed ...
  88. [88]
    Metal Halide Perovskite for next-generation optoelectronics - eLight
    Jan 4, 2023 · Propelling current interest in the 2020s is that MHPs continue to exhibit new and surprising optoelectronic properties, the certified PCE for ...
  89. [89]
    A comprehensive review on the advancements and challenges in ...
    This review covers perovskite solar cell types, charge transport, tandem cells, and challenges, while also discussing future potential.
  90. [90]
    Size/Shape Dependent Toxicity of Gold Nanomaterials in Skin
    Spherical gold nanoparticles with different sizes are not inherently toxic to human skin cells, but gold nanorods are highly toxic due to the presence of CTAB ...
  91. [91]
    Size- and cell type-dependent cellular uptake, cytotoxicity and ... - NIH
    Aug 28, 2019 · The aim of this study was to systematically investigate the effects of particle size on the uptake and cytotoxicity of AuNPs in normal cells and cancer cells
  92. [92]
    Assessment of Pristine Carbon Nanotubes Toxicity in Rodent Models
    CNTs can induce formation of ROS by interacting with cellular components such as mitochondria and cell membrane and are capable of attenuating intracellular ...
  93. [93]
    Cellular Toxicity and Immunological Effects of Carbon-based ...
    Apr 11, 2019 · Previous studies reported various toxic effects of carbon nanomaterials, including ROS generation, DNA damage, lysosomal damage, mitochondrial ...
  94. [94]
    [PDF] Opinion on Nanosilver: safety, health and environmental effects and ...
    Jul 12, 2013 · Silver nanoparticles: behaviour and effects in the aquatic environment. ... Bioaccumulation of silver nanoparticles into Daphnia magna from ...
  95. [95]
    Results of the public consultation SCENIHR's preliminary opinion on ...
    (2012) have raised concerns regarding the environmental contamination of water streams with nanosilver because of its “extreme toxicity to non-target bacteria.
  96. [96]
    Nanomaterials - ECHA - European Union
    It is used in different European regulations, including REACH and CLP, to harmonise how nanomaterials are defined across legal frameworks. As of 1 January 2020, ...
  97. [97]
    Nanomaterial Case Study: Nanoscale Silver In Disinfectant Spray ...
    Mar 20, 2017 · EPA's Nanotechnology White Paper recommends the development of case studies to identify unique risk assessment considerations and research ...Missing: post- | Show results with:post-
  98. [98]
    Nanomaterial Case Study: Nanoscale Silver in Disinfectant Spray ...
    This draft document presents a case study of engineered nanoscale silver (nano-Ag), focusing on the.Missing: post- | Show results with:post-
  99. [99]
    Effect of Surface Coating on the Toxicity of Silver Nanomaterials on ...
    We have demonstrated that coating of the silver nanoparticles with a biodegradable polymer prevents the toxicity of the powder. Toxicity mechanism has been ...Missing: mitigation | Show results with:mitigation
  100. [100]
    Hydrophobic Surface Coating Can Reduce Toxicity of Zinc Oxide ...
    May 7, 2021 · This study evidenced the influence of surface coatings on the physicochemical properties, toxicity, and toxic mechanisms of ZnO-NPs
  101. [101]
    Reduction of pulmonary toxicity of metal oxide nanoparticles by ...
    Apr 21, 2017 · EDTMP coating could passivate the surface of MOx, reduce their cytotoxicity and pulmonary hazard effects. This coating would be an effective safe design ...