Next Article in Journal
Enhancing the Resistance to Shear Instability in Cu/Zr Nanolaminates Through Amorphous Interfacial Layer
Previous Article in Journal
Hybrid Platinum(IV)-Naproxen Nanostructured Drugs Reprogram Melanoma Cells and Overpower Cisplatin
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Chiral Rare Earth Nanomaterials: Synthesis, Optical Properties, and Potential Applications

1
School of Arts and Sciences, Shanghai Dianji University, Shanghai 200240, China
2
College of Sciences, Shanghai Institute of Technology, Shanghai 201418, China
3
State Key Laboratory of Precision Spectroscopy, East China Normal University, Shanghai 200241, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2025, 15(17), 1321; https://doi.org/10.3390/nano15171321
Submission received: 14 July 2025 / Revised: 17 August 2025 / Accepted: 27 August 2025 / Published: 28 August 2025
(This article belongs to the Special Issue Advances in Luminescent Rare-Earth Nanomaterials)

Abstract

Chiral rare earth nanomaterials, which impart optical activity through chiral structures or ligands, possess promising application potential in optoelectronic displays, biosensing, information encryption, and catalysis due to unique properties like circularly polarized luminescence, exceptional photostability, and tunable optics. This review systematically summarizes recent advancements, focusing on synthetic strategies, distinctive optical properties, and experimental demonstrations of applications across various fields. Finally, prospects and challenges for future development are discussed. These studies advance the understanding of circularly polarized luminescence and enable the flexible design and fabrication of chiral rare earth nanomaterials with engineered functionalities, addressing practical challenges in optoelectronic displays and biomedicine.

1. Introduction

Chirality is a property of an object or molecule that cannot be superimposed onto its mirror image [1,2]. Chirality widely exists in nature and can be observed across all scales, from smallest subatomic particles to vast galactic structures. The study of this property is of great significance in fields such as pharmaceutical, bio-pharmaceutical, food, and cosmetic industries [3]. In materials science, chirality influences not only the optical activity of chiral nanomaterials, but also their physical and chemical properties, including electronic spin and catalytic selectivity [4,5,6]. At the nanoscale, the effects of chirality are often significantly amplified, leading to novel phenomena that are absent in bulk materials. For instance, self-assembly is a unifying characteristic of nanoscale particles, enabling the translation of chirality from molecular and nanometer scales to submicron and micrometer scales. This multiscale range allows for the observation of various electronic, chemical, optical, and biological effects, which is essential to engineering sensors as well as photonic, optoelectronic, and information technologies [7]. With the scientific advancement in the fields of nanoscience and chirality, a wide variety of chiral nanomaterials are synthesized and self-assembled, such as chiral quantum dots [8,9,10,11], chiral perovskites [12,13,14,15], and chiral nanocomposites [16]. In recent years, chiral nanomaterials have garnered significant attention in fields such as biology, chemistry, life sciences, chiral nano photonics, and metamaterials [17,18,19,20]. This growing interest can be largely attributed to the introduction of chiral degrees of freedom, which enhance the properties of materials at the nanoscale. The optical properties of chiral nanomaterials can be precisely controlled by manipulating their sizes and shapes.
Rare earth elements consist of 15 lanthanide elements, and the elements Scandium (Sc) and Yttrium (Y). Lanthanide elements have unique electron configurations of [Xe] 4fn−15d0−16s2 (n = 1–15), and the trivalent lanthanide ions (Ln3+) are the most common and stable despite having several chemical valences [21]. The unpaired 4f electrons in rare earth elements are shielded by outer 5s and 5p orbitals, making them relatively insensitive to crystal field effects. Compared to that, transition metals like Cu and Co have partially filled d orbitals [22], which have broader and less defined electronic transitions, rare earth elements have sharp, well-defined electronic transitions. Due to the unique 4f energy levels, rare earth nanomaterials have the advantages of excellent photostability, a large anti-Stokes shift, long luminescence lifetime, and sharp-band emission [23]. Owing to their unique optical properties, Ln3+ doped luminescent nanocrystals are promising for applications ranging from biosensor, lasing, photovoltaic devices, information security, and electrocatalysis [23,24,25].
Chiral rare earth nanomaterials not only exhibit the inherent advantages of rare earth elements but also introduce optical activity through their chiral structures or chiral ligands. This combination imparts unique application potential in optical sensors, biological labeling, and next-generation display technology [26,27,28,29]. Chiral nano-sized rare earth materials demonstrate significant circular dichroism (CD) and circularly polarized luminescence (CPL) properties. CD and CPL spectroscopies are the typical technical means for characterizing the optical activity of chiral materials. CD spectroscopy can precisely analyze the ground-state optical properties of materials by detecting the absorption difference in left- and right- circularly polarized light, providing key evidence for processes such as chiral ligand modification, nanoparticle self-assembly, or lattice symmetry breaking. CPL spectroscopy directly reflects the chiral information of the excited state in chiral luminophores or phosphors by recording the intensity difference in left- and right- circularly polarized light during the luminescence process, and it is particularly suitable for studying the f-f transition characteristics of rare earth ions (such as Eu3+, Tb3+, etc.). The generation mechanism of CPL involves the emission of a photon from a chiral excited state, which gains polarization due to a nonzero dot product between the electric dipole transition moment and the imaginary magnetic dipole transition moment, characterizing the electronic transition responsible for luminescence [30].
The research on chiral rare earth materials is expected to expand the application range of CPL from the visible light region to the ultraviolet (UV) and near-infrared regions [31], possessing high stability [32,33]. For example, cerium (Ce)-based rare earth metal halides, such as (R/S-MBA)CeCl4·2CH3OH (MBA = α-methylbenzylammounium), exhibit pronounced CD in the ultraviolet region (340–400 nm) and display distinct CPL characteristics in the visible light region (300–700 nm), expanding the application potential of chiral optoelectronics (such as UV-CPL light sources) [31]. Furthermore, by modifying the luminescent ions incorporated into the lattice, color-tunable CPL can be achieved, and this type of CPL exhibits excellent thermal stability and can maintain luminescence at temperatures exceeding 300 °C [32]. Chiral rare earth fluoride nanoparticles induced by helical silica can still maintain their CPL activity even after calcination at 400 °C [33]. This resistance to high temperatures enables encryption and display applications in high-temperature environments. By doping Er3+/Yb3+ ions into bismuth oxychloride-based inorganic nanostructured phosphor, modified with chiral sugar alcohols, exhibiting enhanced upconversion luminescence compared to non-chiral phosphors [34], and thus expanding its application in biological imaging and photodynamic therapy. These advancements establish a solid foundation for practical applications of chiral rare earth nanomaterials.
Recent reviews have summarized the advances in chiral rare earth nanomaterials [29,35,36], but this work sets itself apart by proposing a unified framework for understanding the origins of chirality and synthesis methods and incorporating groundbreaking applications such as stimulus-responsive circularly polarized luminescence and high-security encryption that have been reported in 2024–2025. This article provides a comprehensive overview of recent advancements in chiral rare earth nanomaterials, focusing on three critical aspects: synthesis strategies, optical properties, and potential applications (Figure 1). We begin by exploring the origins of chirality in chiral rare earth nanomaterials and categorizing synthesis strategies into three primary methods: hydrothermal/solvothermal methods, chiral template-assisted synthesis, and self-assembly. Next, we present the optical properties, including CD and CPL. Given the unique optical characteristics of chiral rare earth nanomaterials, we then discuss their potential applications in circularly polarized light-emitting devices, optical anti-counterfeiting and information encryption, as well as drug delivery, therapy, and bioimaging. Finally, we highlight current challenges and future perspectives in this rapidly evolving field.

2. The Origin of Chirality

The crystalline structure, shape, and interaction of chiral ligands are the factors that result in chirality in inorganic nanomaterials [37]. For semiconductor nanocrystals, the origins of chirality are well known and can be classified into three types: (1) semiconductor nanocrystal core containing dislocations and defects, (2) incorporation and arrangement of nanocrystals into chiral superstructures, and (3) interactions between chiral ligands and the nanocrystal [38]. In strict analogy with the classification in semiconductor nanocrystals, chirality in chiral rare earth nanomaterials originate from (1) intrinsic chirality arising from chiral crystal, (2) chiral assemblies of achiral rare earth nanomaterials, and (3) ligand-induced chirality into nanomaterials. According to their chiral sources, there are intrinsic chiral rare earth nanomaterials, chiral co-assembled rare earth nanomaterials, and ligand-induced chiral rare earth nanomaterials, respectively, as shown in Figure 2.
In intrinsic chiral rare earth nanomaterials, chirality is determined by the inherent structure at the atomic scale, such as the inherent chiral space group of the crystal structure. Chiral objects or molecules exist in two non-overlapping forms in space, known as enantiomers [39]. The enantiomer with molecular chirality is used to guide the symmetry breaking of the chiral nanocrystals, resulting in a preferential chiral space group [40]. The resultant nanocrystals are categorized as L- or D-form based on the molecular enantiomer, and X-ray diffraction can verify that the sample’s chirality originates from the chiral lattice. LaPO4:RE3+ (RE = Eu, Tb, Tm) nanowires are grown in a direction controlled by chiral molecular inducers, forming nanoscale structures with atomic-scale chirality [32,41]. These nanomaterials, regardless of their external shapes, have chiral lattice arrangements internally and can interact with polarized light [37]. These intrinsically chiral rare earth nanomaterials are more stable than other assemblies, showing clear CPL signals even at temperatures over 300 °C [32].
Interestingly, nanoscale chirality can also be achieved by co-assembly of achiral nanomaterials with chiral templates such as DNA [42,43,44], liquid crystal [35,45,46,47], chiral polymer [48,49], and organogelators [50,51,52]. In chiral co-assembled rare earth nanomaterials, chiral templates (such as liquid crystals, helical silica, etc.) are utilized as hosts to guide the formation of chiral structures, where rare earth materials are used as guests. For example, HS@CeF3:RE (HS = helical silicas) nanocomposites synthesized using helical silica templates [33], and chiral luminescent nanomaterials prepared by liquid crystal templates [35]. These chiral assemblies of achiral rare earth nanomaterials are expected to be useful for studying their interparticle coupling and responsive dynamic properties, which could lead to a wide range of potential applications [35,53,54].
Interaction with chiral ligands or molecules is another strategy to induce chirality in nanomaterials. Ligand exchange between chiral ligands and initial achiral ligands is a commonly used method to induce chirality in chalcogenide metal semiconductors, such as CdS and CdSe quantum dots [10,55]. This process does not destroy the core structure of the quantum dots. In ligand-induced chiral rare earth nanomaterials, chirality can also arise from interactions with chiral ligands such as amino acids or short peptides. Unlike the ligand-exchange method in quantum dots, Dawn E. Barry et al. designed near-infrared lanthanide-emissive Langmuir−Blodgett monolayers using Nd (III) directed self-assembly synthesis of chiral amphiphilic ligands [56]. These ligands impart both amphiphilicity and chirality to the complexes.

3. Synthesis Methods

According to the origin of chirality in chiral rare earth nanomaterials, we classify the synthetic methods into three categories: solvothermal method, template-assisted co-assembly, and self-assembly method. The first method synthesizes intrinsic chiral rare earth nanomaterials, while the latter two methods synthesize non-intrinsic chiral rare earth nanomaterials. The intrinsic chiral rare earth nanomaterials usually have higher stability than the non-intrinsic chiral rare earth nanomaterials due to the atomic chirality. In the following, we will elaborate on the synthetic methods of chiral rare earth nanomaterials in detail.

3.1. Hydrothermal/Solvothermal Method

Hydrothermal/solvothermal method has garnered significant attention for its innovative use of chiral molecules as inducers [32,34,41,57]. This method leverages the unique properties of chiral molecules to direct the preferential growth of materials along specific chiral axes. Enantiomers play a critical role in the synthesis and development of chiral nanocrystals, as they precisely guide the process of symmetry breaking. This guidance ensures that the formation of these nanocrystals favors a specific, preferred chiral space group. Consequently, the final nanocrystalline structures, which reflect this directed symmetry breaking, are accurately classified and designated as either L-form or D-form, corresponding directly to the identity of the molecular enantiomer used in their fabrication. This systematic approach not only underscores the significance of enantiomer selection but also illustrates the meticulous control exercised over the chirality of the resulting nanomaterials [40,41].
Researchers have developed a novel family of intrinsically chiral rare earth nanowires via this facile solution method [32,41], as shown in Figure 3. By incorporating different kinds of rare earth ions (such as europium (Eu3+), terbium (Tb3+), and thulium (Tm3+)), these chiral nanowires emit multicolor CPL (e.g., red, green, or blue). Chiral organic and organic−inorganic hybrid system also possesses color-tunable CPL but suffers from chiral instability employing unstable chiral supramolecules or their dependence on chiral surface ligands [58,59,60]. However, the chiral nanowires synthesized by this method have high thermal stability at high temperatures of up to 300 °C due to the atomic chirality, which could be well coupled with other functional materials. What is more, by injecting this chiral nanowire dispersion into poly (vinyl alcohol) solution and after drying naturally, a substrate-free flexible chiral film is obtained. Thus, this method not only realizes atomic-scale chiral crystals with robust, color-tunable CPL properties but also enables innovative CPL application prospects.

3.2. Chiral Template-Assisted Synthesis Method

Compared to intrinsic chiral rare earth nanomaterials, chiral co-assembled rare earth nanomaterials exhibit a broader variety of types. Currently, various systems have been reported that achieve chirality through the co-assembly of non-chiral luminescent guests (rare earth luminescent nanoparticles) and chiral hosts. Typically, the chiral host provides a chiral confined space for the non-chiral luminescent guest. When the luminescent guest assembles within this chiral confined space, chirality can be effectively transferred from the chiral host to the luminescent guest. Due to the diversity in the selection of chiral hosts and luminescent guests, this strategy can achieve tunable optical properties by regulating different luminescent centers and chiral hosts, thereby demonstrating significant versatility. The following sections introduce several common chiral templates, including chiral metal–organic frameworks (MOFs), chiral liquid crystals, chiral organic gels, chiral helical SiO2 structures, and chiral layered structures.

3.2.1. Co-Assemble with Chiral MOFs Template

MOFs are a subclass of coordination polymers known for their highly ordered crystalline porous structure self-assembled by metal ions and organic ligands [61], which have attracted considerable attention from researchers. Introducing chirality into these MOF materials is expected to confer several advantageous properties, such as asymmetric catalysis and enantioselective recognition or separation [62]. Nanomaterials such as quantum dots and upconversion nanoparticles are easily loaded into the chiral MOFs during the synthetic process. ZIF-8 is a representative zeolitic imidazolate framework (a subclass of MOFs) material with a sodalite topology, formed by zinc ions and 2-methylimidazole (Hmim) [63,64]. The chiral ZIF can be synthesized through the method of mixed-ligand co-assembly [65]. For example, L-histidine (L-His) and Hmim were dissolved in a mixed solvent of methanol and water. A small amount of triethylamine was then added to the solution under stirring. After stirring for 10 min, the resulting mixed-ligand solution was gradually added to a methanol solution of Zn(NO3)2·6H2O. The mixture was stirred at room temperature for 24 h. The colorless product was then washed with large amounts of water and methanol, collected by centrifugation, and dried under vacuum. Various types of launchers, including lanthanide-doped upconversion nanoparticles (UCNPs), can be effectively incorporated into chiral MOFs through a straightforward in situ synthesis approach as shown in Figure 4a [65]. In addition, through the self-assembly of chiral MOFs and perovskite nanocrystals (NCs), Zhang et al. prepared a pair of crystalline enantiomeric adducts, (P)-(+)/(M)-(−)-EuMOF⊃MAPbX3 (where MA = CH3NH3+, and X = Cl, Br, I). This co-assembly strategy aims to embed achiral MAPbBr3 perovskite NCs within chiral MOFs by inheriting the chirality of the host MOFs through host–guest Eu–Br and Pb–O coordination bonds [58]. These works open a new avenue for the general fabrication of solid-state CPL composite materials [60,66,67].

3.2.2. Chiral Liquid-Crystalline Template-Assisted Synthesis Method

In recent, chiral liquid-crystalline templates have also been used to synthesize chiral nanomaterials [68,69,70,71]. Cholesteric or chiral nematic liquid crystals with helical nanostructures are the most commonly used method to achieve CPL enhancement by using chiral liquid crystal templates. As a chiral template, one of attractive features of chiral nematic liquid crystal is the photonic bandgap effect, which selectively reflects circularly polarized light with the same handedness and transmits light with the opposite direction. It provides a way to control the chiral direction. Cellulose nanocrystals (CNCs), as renewable nanomaterials, can be produced in large quantities through acid hydrolysis of natural fibers [29]. When CNCs are dispersed in water solvents at a certain concentration, they can self-assemble to form left-handed nematic liquid crystals. Self-assembled photonic films made from CNCs can selectively reflect CPL, which is similar to some crustaceans that reflect light through a spiral structure [72,73]. The intriguing properties of CNCs have inspired research into generating CPL by doping fluorescent chromophores into photonic cellulose films, for example, Li et al. obtained CNC-based chiral photonic films fabricated via the co-assembly of CNCs liquid-crystalline template, glycerol, and lanthanide-doped UCNPs as shown in Figure 4b [74]. During the synthesis process, (NaYF4:Tm/Yb) UCNPs were first prepared. Polyvinyl alcohol (PVA) was added to an aqueous dispersion of UCNPs and then the mixture was stirred for 1 h at ambient temperature to give the UCNP@PVA dispersion. UCNP@PVA dispersion and different amounts of glycerol were added to the CNCs suspension with vigorous stirring for 5 h, the suspension was cast in a rectangular plastic container and left to dry naturally for 2 to 3 days, thereby obtaining chiral photonic films. Li et al. indicated that with the incorporation of multiple-emissive chromophores into a photonic cellulose film with an adjustable photonic bandgap (PBG), tunable CPL might be achieved simply by adjusting the PBG of chiral host. Thus, by combining a variety of luminescent nanomaterials—such as quantum dots, perovskite nanocrystals, and rare earth nanoparticles—this method provides a convenient and environmentally friendly way to obtain CPL materials.

3.2.3. Chiral Gel-Assisted Synthesis Method

Chiral gels can also function as chiral hosts, which create a chiral microenvironment for the system. Achiral nanomaterials can emit circularly polarized luminescence when encapsulated within chiral helical nanotubes via a co-gelation process [26,50,51,52]. For instance, the co-assembly of carbon dots and chiral gelators can successfully construct composite materials with tunable chiral luminescence, where both the luminescence intensity and direction depend on the excitation wavelength [50]. Liu et al. used chiral gelator lipid N,N′-bis(octadecyl)-l-glutamic diamide (LGAm) or its enantiomer DGAm to co-assemble with achiral perovskite nanocrystals, inducing chiral transfer and causing the nanocrystals to exhibit chiral luminescence. The direction of luminescence depends on the chirality of the lipids [51]. These studies indicate that the handedness direction of the CPL in co-gel systems can be modulated by the excitation wavelength and molecular chirality. For rare earth materials, Zhou et al. designed an organic gel (BzLG), where BzLG is an abbreviation of an organogelator containing benzimidazole moiety, N,N2032-bisoctadecyl-2-(1Hbenzimidazole-2-carbonyl)-L-glutamic amide [75]. They developed various chiral nanostructures by introducing metal ions, which can also coordinate with rare earth metal ions. Glutamine served as the gelation agent, while the benzimidazole group acted as the coordination unit for rare earth ions. Upon the introduction of Eu(NO3)3 or Tb(NO3)3, the material preferentially assembled into a flower-like morphology consisting of aggregated nanotubes. Conversely, when EuCl3 or TbCl3 were added, nanofibers were produced, highlighting the influence of different anions on the interaction between metal ions and the gel matrix. This interaction, in turn, enables control over the gel’s structural morphology. Additionally, achiral lanthanide-doped UCNPs (NaYF4:Yb/Er or NaYF4:Yb/Tm) can be encapsulated into chiral helical nanotubes through the procedure of co-gelation (as shown in Figure 4c) [26]. These nanomaterials display upconverted CPL with wavelength ranging from 300 nm to 850 nm.

3.2.4. Chiral Silica-Assisted Synthesis Method

Chiral silicas (nanospheres, twisted nanorods, and helical ribbons, etc.) have been found to exhibit chirality at either the molecular or nanoscale level [76,77,78]. These materials are considered promising chiral templates for the development of CPL-active compounds due to their stable architecture and ease of surface functionalization [79,80,81]. Lu. et al. constructed a novel type of pure inorganic chiral nanocomposites HS@CeF3:RE by in situ assembly strategy [33]. Left-/right-handed helical silicas (L/R-HS) are selected as chiral hosts and RE-doped CeF3 nanoparticles (NPs) are selected as luminophores. As shown in Figure 4d, the left- or right-handed helical ribbon structure was used as a template for inorganic transcription to obtain HS. R-HS was successively surface-functionalized with 3-aminopropyltriethoxysilane and succinic anhydride. Rare earth ions were attracted to the HS surface through coordination and electrostatic interactions. Subsequently, NaHF2 was added as a fluorine source. With the aid of ultrasonic treatment, CeF3:TbNP was rapidly in situ assembled within 2 min, thereby preparing L/R-HS@CeF3:Tb nanocomposites. In addition, Duan et al. also reported an impressive upconversion CPL (UC-CPL) example from a co-assembled system composed of helical platform and UCNPs [82]. This work has suggested that an achiral molecule C3-symmetric benzene-1,3,5-tricarboxamid with three identical benzoic acid arms (BTABA) can form chiral nano helices through symmetry breaking. This method provides a universal approach for designing tunable UC-CPL materials, greatly expanding the research scope of CPL materials. Importantly, the co-assemblies exhibit enhanced CPL, which has been confirmed in many co-assembly systems [83].
Chiral SiO2 (non-helical) can also be formed through the induction of chiral organic templates. Jin et al. [79] successfully constructed a novel solid-state inorganic CPL material system containing luminescent sub-10-nanometer rare earth oxide nanoparticles (guests) and chiral silica nanofibers (hosts) by a simple two-step adsorption–calcination method. Jin et al. [84] first mixed polyethyleneimine (PEI) with chiral tartaric acid (tart) to form chiral crystal complexes (CCCs). Subsequently, they dispersed the CCC in a mixed solution of water and tetramethyl orthosilicate (TMOS), and deposited silica through hydrolysis and condensation reactions. Finally, they removed the organic template via high-temperature calcination, resulting in high-temperature-resistant chiral silica (PEI/tart@SiO2 nanofibers) as shown in Figure 4e. Jin’s team [79] incorporated rare earth ions into PEI/tart@SiO2 nanofibers. The research found that this inorganic system exhibits CD optical activity in the ultraviolet wavelength region. This work opens up new avenues for the development of CPL systems based on inorganic materials and indicates that chiral silica hosts can effectively transfer chiral information to rare earth oxides guests.

3.2.5. Chiral Layered Structure-Assisted Synthesis Method

Under the influence of chiral molecules, BiOCl tends to form chiral layered nanostructures [85]. Inorganic nanostructures with a chiral layered morphology can also function as inorganic chiral hosts. Qiu et al. [34] synthesized BiOCl: 5%Yb3+/4%Er3+ (BYE) UC-CPL materials with D-sorbitol (D-Sor) (Figure 4f) using a hydrothermal method followed by calcination in air. Compared to the BiOCl material without chiral ligands (A-BYE), D-BYE exhibits a significantly stronger upconversion luminescence intensity. D-BYE demonstrates pronounced CPL properties under 980 nm laser excitation, whereas A-BYE does not. This indicates that the chiral layered structure is not only essential for the emission intensity of CPL but also enhances the photoluminescence quantum yield (PLQY). The utilization of rare earth-doped chiral layered nanomaterials offers a novel strategy for achieving UC-CPL.
Figure 4. (a) Schematic synthesis of chiral MOFs and CPL-active MOFs [65]. (b) Illustration showing CNC-based chiral photonic films fabricated via the co-assembly of CNCs, glycerol, and lanthanide-doped UCNPs [74]. (c) Schematic representation of upconverted circularly polarized luminescence based on co-gel with UCNPs [26]. (d) Schematic illustration for the synthesis of R-HS@CeF3: rare earth nanocomposites [33]. (e) Illustration showing geometrically organized chiral segments dispersed throughout the silica wall, along with confined chromophores [84]. (f) SEM images of D-BYE [34].
Figure 4. (a) Schematic synthesis of chiral MOFs and CPL-active MOFs [65]. (b) Illustration showing CNC-based chiral photonic films fabricated via the co-assembly of CNCs, glycerol, and lanthanide-doped UCNPs [74]. (c) Schematic representation of upconverted circularly polarized luminescence based on co-gel with UCNPs [26]. (d) Schematic illustration for the synthesis of R-HS@CeF3: rare earth nanocomposites [33]. (e) Illustration showing geometrically organized chiral segments dispersed throughout the silica wall, along with confined chromophores [84]. (f) SEM images of D-BYE [34].
Nanomaterials 15 01321 g004

3.3. Self-Assembly Method

The development of functional nanostructures through lanthanide-directed self-assembly has become an increasingly important area of research, given their potential applications in molecular recognition, sensing, imaging, and optical devices [56,86,87,88]. Two-dimensional heterostructure films (D-/L-Se/DPNUCNPs films) were fabricated by self-assembly at the liquid–liquid interface using D-/L-Se NPs and NaYF4:Yb/Tm upconversion nanoparticles (DPNUCNPs) [89]. As shown in Figure 5a, first, a monolayer D-/L-Se film (1L D-/L-Se film) was formed on a quartz substrate using the liquid–liquid interfacial assembly technique. Following this, a layer of DPNUCNPs was applied to the surface of the chiral Se film by the same method, resulting in the formation of a two-dimensional heterogeneous structure composed of 1L D-/L-Se and 1L DPNUCNPs films. Furthermore, the CPL performance of the films was optimized by controlling the assembly sequence and the number of layers as shown in Figure 5b. The basic mechanism of chirality transfer within these 2D self-assembled films is that chirality is transferred from the excited state of the chiral Se NPs to the excited state of the DPNUCNPs during the energy transfer process from DPNUCNPs to Se NPs. In addition, chiral amphiphilic Nd (III) complexes form monolayer films via the Langmuir–Blodgett technique [56].

4. Chiroptical Properties

Chiroptical (chiral–optical) properties refer to the optical characteristics of molecules or objects [90]. These properties arise due to the strong interaction between chiral nanomaterials and electromagnetic waves, leading to phenomena such as CD and CPL [35,91,92]. Generally, natural light from sunlight or other common sources is considered unpolarized, with its electromagnetic waves emitted and propagated randomly at various polarization angles. The polarized light can be simply classified into linearly, circularly, elliptically, and partially polarized light [93,94,95]. Experimentally, circularly polarized light can be obtained from non-polarized light by using a linear polarizer and a quarter-wave plate. However, using polarizers causes the loss of about 50% of light due to the conversion of the unpolarized light into linearly polarized light. Chiral materials exhibit unique optical activities, including CD and CPL. In the following, we will first introduce the basic concepts related to CD and CPL and then delve into the recent advancements of chiroptical properties in chiral rare earth nanomaterials.

4.1. Basic Concepts Related to CD and CPL

CD and CPL involve the differential absorption and emission rates of left- and right-circularly polarized light, respectively. CD and CPL spectra are essential techniques for characterizing the optical activity of chiral materials. Figure 6 illustrates the mechanisms of CD and CPL measurements. When right- and left-handed circularly polarized light alternately pass through a chiral material (for example, an L-form material), right-handed circularly polarized light is absorbed less than left-handed circularly polarized light, as shown in Figure 6a. Conversely, when right- and left-handed circularly polarized light alternately pass through its enantiomer (D-form material), right-handed circularly polarized light is absorbed preferentially. The CD intensity is proportional to ALAR, where AL and AR represent the absorbance of left- and right-handed circularly polarized light, respectively. Consequently, the CD spectrum exhibits a mirror image with positive and negative signals. On the other hand, unpolarized light is used to excite a chiral material. As shown in Figure 6b, L- and D-form materials preferentially emit right- and left-handed CPL, respectively. The CPL intensity is proportional to ILIR, where IL and IR represent the intensities of left- and right-handed CPL, respectively. Positive and negative signals correspond to L- and D-form materials, respectively. CD and CPL spectra can precisely analyze the ground and excited state optical properties of materials. The complementary combination of these two techniques not only enables the quantitative assessment of the anisotropy factor (g factor) of materials but also reveals the transfer and amplification mechanism of chirality, laying a theoretical foundation for the design of high-performance optoelectronic devices [82,96].
Both the CD and CPL spectra are ultimately based on the difference in intensity of circularly polarized light. Therefore, the magnitude method of CD and CPL are similar [83]. Circular dichroism refers to the differential absorption rates of L-CPL and R-CPL by a sample. The definition of the asymmetry factor g in circular dichroism spectroscopy is g CD = 2 ( A L A R ) / ( A L + A R ) , where AL and AR denote absorption coefficients of L-CPL and R-CPL, respectively. Experimentally, CD data are typically measured in terms of ellipticity (θ) (unit: deg or mdeg) [97]. This relationship of ellipticity (unit: mdeg) and gCD is expressed in Equation (1):
g CD = ellipticity × absorbance 32980  
where the “ellipticity” and the “absorbance” can be directly obtained from CD spectra. When a chiral luminescent system is excited, the emission intensities of the left and right CPL that it produces are not equal [30,98,99]. Due to the significant challenges associated with directly measuring absolute emission intensities, the luminescent dissymmetry factor ( g lum ) is employed to evaluate the extent of CPL. The mathematical expression for the g lum factor is provided in Equation (2), as shown below:
g lum = 2 I L I R I L + I R
where IR and IL represent the emission intensities of R-CPL and L-CPL, respectively. Experimentally, glum = [ellipticity/(32,980/ln 10)]/total fluorescence intensity at the CPL extremum [26]. The value of g lum ranges from +2 to −2, with ±2 indicating that the system emits ideal circularly polarized light, while a g lum value of 0 corresponds to unpolarized light [100]. Notably, when using g CD values and g lum values as evaluation metrics, these parameters effectively eliminate the influence of differences in sample concentration, variations in luminescence intensity, and the use of different instruments during experimentation, resulting in more accurate and reliable outcomes [29].

4.2. Chiroptical of Chiral Nanomaterials

Lu et al. developed a novel rare earth fluoride nanocomposite material and investigated its optical properties and potential applications [33]. Figure 7a,b show the XRD patterns of L/R-HS@CeF3: Tb and SAXS patterns of R-HS and R-HS@CeF3: Tb nanocomposites. As shown in Figure 7c, the broad excitation band (220 nm~300 nm) is attributed to the allowed 4f→5d transition of Ce3+ ions. Under excitation at 254 nm, R-HS@CeF3: Tb exhibited strong emission peaks of Tb3+ at 489, 542, 584, and 620 nm, corresponding to the 5D47FJ (J = 6, 5, 4, 3) transitions, respectively (Figure 7d). The CD spectra of L/R-HS@CeF3: Tb exhibited mirror-image symmetry, indicating that the nanoparticles were oriented in opposite directions on the surfaces of L/RHS (Figure 7e). Compared to other chiral nanomaterials, such as chiral quantum dots, it is difficult to observe the CD signals of chiral rare earth nanomaterials. This challenge may be attributed to the low concentration of rare earth nanomaterials in the complexes and the strong scattering effects [26].
Research on the CPL of chiral rare earth nanomaterials has made significant progress [101,102]. As shown in Table 1, the glum values range from 4.7 × 10−3 to 1.1. The g factor of intrinsic chiral rare earth nanomaterials is generally larger than that of non-intrinsic nanomaterials, which may be related to the degree of chiral assembly. The magnitude of glum can be controlled by adjusting the glycerol content [74], the number of film layers [89], and the incident light [66]. For example, the CPL activity of self-assembled films of chiral selenium nanoparticles can be effectively enhanced by controlling the assembly sequence and the number of layers, reaching a maximum value of 0.68 [89]. Additionally, by introducing stimulus-responsive groups or materials, the CPL properties can be made responsive to external conditions such as temperature and humidity [74]. Furthermore, composite materials composed of perovskite and chiral lanthanide MOFs can respond to various stimuli, including chemical substances, temperature, and light, enabling reversible switching of CPL [58].
The color of CPL in chiral rare earth nanomaterials can be tuned by varying the doped ions or by applying an external stimulus, such as glycerol [32,33,74]. For example, phosphate nanowires of LaPO4, GdPO4, and TbPO4 doped with Eu3+, Tb3+, or Tm3+, as well as rare earth-doped CeF3 nanoparticles assembled within helical silica, exhibit CPL emissions in different colors [32,33]. This tunability is primarily achieved by adjusting the type and concentration of rare earth ions, which controls the CPL emission wavelength. Additionally, chiral europium halide (R/S-3BrMBA)3EuCl6 demonstrates magnetic field-tunable red CPL at room temperature, with its degree of polarization modulated by an external magnetic field [105]. Furthermore, the CPL emission intensity of chiral nanomaterials is generally higher than that of their non-chiral counterparts [34].
Some chiral rare earth nanomaterials exhibit excellent thermal stability, maintaining CPL activity even under high-temperature conditions [32,33]. These studies indicate that chiral inorganic nanomaterials hold significant potential for applications in the CPL field. By integrating chiral assembly strategies, rare earth-doped luminescent materials, and photonic crystal structures, it is possible to fabricate efficient, tunable, and stimulus-responsive CPL materials.

5. Potential Applications

Chiral substances play a crucial role in biological systems. Chiral inorganic nanomaterials have similar size, charge, surface properties, and morphology to natural chiral nanomaterials, but usually exhibit extraordinary properties, such as high g factor values and enantiomeric configuration [28]. Chiral nanomaterials with intrinsic chirality or spatial asymmetry at the nanoscale are currently in the limelight of both fundamental research and diverse important potential applications, such as biosensing, drug delivery, early diagnosis, bioimaging, and disease treatment. In the following, we will elaborate the application prospects of chiral rare earth nanomaterials in circularly polarized luminescence devices, information encryption, biosensing, bioimaging, and disease diagnosis and treatment.

5.1. Circularly Polarized Light-Emitting Devices

CPL materials have broad application prospects in fields such as information security, 3D displays, and optoelectronic devices. Currently, most reported CPL materials are organic molecular systems, which are sensitive to environmental factors and exhibit poor stability. Inorganic nanomaterials, on the other hand, offer better control over optical performance and higher chemical stability, making them a hotspot in CPL material research. By doping different kinds of rare earth ions (such as Eu3+, Tb3+, Tm3+) into the crystal structure, the emission colors of CPL can be tuned. Wang et al. successfully prepared chiral LaPO4:RE3+ (RE = Eu, Tb, Tm) nanowires, achieving CPL emissions in red, green, and blue colors [32]. The presence of atomic-scale chirality ensures that CPL emissions are unaffected by molecular ligands and exhibit excellent thermal stability, maintaining stable luminescence even at high temperatures of 300 °C. Furthermore, they dispersed the nanowires into a polymer matrix to create transparent, flexible CPL films, laying the foundation for practical applications of CPL materials.
Li et al. constructed chiral photonic films using cellulose nanocrystals (CNCs) and discovered that glycerol-based composite films exhibit humidity-responsive UC-CPL at blue light wavelengths, with glum values varying with relative humidity as shown in Figure 8a,b [74]. This humidity-responsive UC-CPL material offers new insights for the development of intelligent CPL devices. This opens new avenues for designing high-stability, tunable CPL materials, facilitates a deeper understanding of inorganic chiral information, and provides fresh insights for the development of novel optoelectronic devices with outstanding CPL properties.

5.2. Optical Anti-Counterfeiting and Information Encryption

CPL can be a type of luminescence generated by the intrinsic chirality of luminescent molecules or complexes. It encodes the “fingerprints” of chiral molecules, which cannot be replicated. Conventional CPL spectrometers are slow and costly, limiting their application in security inks. In recent years, the emergence of next-generation fast CPL spectrometers and CPL microscopes has made rapid verification and imaging of CPL security inks possible. Lanthanide metal complexes hold great potential in security inks, particularly in the realm of CPL applications. With the continuous advancement of CPL analysis and reading technologies, CPL-active lanthanide metal security inks are expected to become an important component of the next generation of security inks [106]. Additionally, Chiral lanthanide lumino-glass (such as europium complexes) exhibit switchable CPL patterns under UV light, leading to new applications of CPL materials as security inks [107].
Lu et al. reported a novel inorganic chiral nanocomposite made of helical silica HS and rare earth fluoride nanoparticles, exhibiting multicolor CPL and time-resolved photoluminescence (TRPL) properties for advanced anti-counterfeiting applications as shown in Figure 9 [33]. They mixed different nanocomposites with PVA to create films and fabricated plant patterns. Under daylight, the patterns are transparent; however, when exposed to UV light, the patterns emit different colors of light and exhibit dynamic and chiral signals, enabling advanced anti-counterfeiting features. Additionally, they delved into the application potential of these nanocomposites for information encryption. By incorporating different nanocomposites into a 96-well plate, they created multicolor codes that can be decrypted step-by-step through various decryption methods (UV light, TRPL, CPL), achieving multilayer optical encryption. Additionally, Hao et al. reported that the chiral Se/DPNUCNPs films were also patterned for encryption applications [89]. Notably, this upconversion CPL of chiral Se/DPNUCNPs films were excited by near-infrared irradiation (980 nm), which holds significant potential for future applications in encryption inks due to its hard replication.

5.3. Drug Delivery, Therapy, and Bioimaging

Chiral rare earth nanomaterials have good biocompatibility and can be used in biomedical applications such as bioimaging and photodynamic therapy. In recent years, a large number of studies have demonstrated that rare earth-doped nanomaterials have shown significant application value in the biomedical field, especially in bioimaging and photodynamic therapy [108,109,110,111]. These studies have confirmed their advantages from different perspectives, for example, the unique advantages of upconversion nanoparticles in bioimaging and antibacterial treatment [108,109], diagnosis, and treatment [111]. Wang et al. introduced some applications of chiral inorganic nanomaterials, including the treatment of Alzheimer’s disease, cancer, and viral infections, demonstrating their significant potential in biomedicine and bioengineering [112]. Chiral rare earth complexes can selectively bind to specific biomolecules, such as proteins and nucleic acids, through their CPL properties [113]. Additionally, by leveraging the CPL characteristics of chiral rare earth materials, image contrast can be enhanced, thereby improving the resolution of biomedical imaging. For example, chiral BiOCl:Er3+/Yb3+ nanostructures show promise for biological imaging due to their enhanced upconversion luminescence [34].
Kuang et al. developed a gold nanorod (NR)-UCNP tetramer assembly featuring tunable optical properties and dual-mode biosensing capabilities [114]. When the oligonucleotide targets are present, the hairpin-like DNA strands within the tetramers extend as shown in Figure 10a. This extension increases the gap length, resulting in a decrease in both upconversion luminescence and CD signal intensities as shown in Figure 10b–e. Owing to the strong optical activity and enhanced luminescence of the tetramer, it can be employed for dual-mode biosensing, simultaneously detecting upconversion luminescence, and circular dichroism signals. This method can detect DNA at concentrations as low as the attomolar level and demonstrates high specificity and reliability.

6. Summary and Outlook

This paper reviews the research progress on chiral rare earth nanomaterials. Firstly, it introduces the chirality origins and synthesis methods for these materials in detail. Researchers have achieved controllable design of the morphology, size, and chiral features of rare earth nanomaterials by carefully adjusting the synthesis processes. For intrinsic chiral rare earth nanomaterials, hydrothermal/solvothermal methods are commonly used, utilizing rare earth ions with excellent luminescent properties (such as Eu3+, Tb3+, and Tm3+), rare earth oxides, and upconversion nanoparticles. These intrinsic chiral rare earth nanomaterials not only exhibit tunable luminescent colors but also demonstrate good thermal stability. In terms of non-intrinsic chiral hosts, chiral MOFs, chiral liquid-crystalline structures, chiral helical structures, chiral organic gels, and chiral SiO2 provide favorable chiral environments for luminescent guest species. When luminescent guests are assembled along the spatial structure of the chiral hosts, chiral composites can be constructed. Next, the optical properties of chiral rare earth nanomaterials are introduced, including CD and CPL spectra. The gCD-value and glum-value eliminate the influences of concentration and luminescence intensity, making them significant for the study of the optical properties of chiral materials. Finally, based on the excellent optical performance of chiral rare earth nanomaterials, we introduce their potential applications in CPL devices, optical anti-counterfeiting and information encryption, drug delivery, and bioimaging.
Despite significant progress in the research of chiral rare earth nanomaterials, several challenges remain. Key issues include achieving large-scale, cost-effective synthesis of high-quality materials, further enhancing chiroptical properties, and addressing their stability and biosafety in complex environments. Further research is essential to improve the luminescence efficiency [115,116] and CPL intensity of chiral rare earth materials while expanding their range of applications. For instance, optimizing chiral templates and controlling the coordination environment of rare earth ions could significantly enhance CPL signal strength. Additionally, to facilitate their use in biomedical applications, it is crucial to improve water solubility and develop more water-soluble chiral groups suitable for aqueous solutions. The development and application of CPL in the biomedical field remain in their early stages. To achieve clinical translation, technical bottlenecks must be further addressed. For example, future efforts could focus on preparing chiral rare earth nanomaterials with highly responsive CPL to biological small molecules to overcome current limitations. Such advancements will contribute to building a medical platform that integrates diagnosis and real-time treatment monitoring, thereby enabling visualization of the treatment process.

Author Contributions

Conceptualization, L.Z., P.L. and F.C.; methodology, L.Z. and H.Z.; validation, X.W. and Y.Y.; formal analysis, Y.X. and F.C.; investigation, L.Z., P.L., H.Z., R.H., Y.X., F.C., X.W. and Y.Y.; writing—original draft preparation, L.Z., P.L. and F.C.; writing—review and editing, P.L. and Y.Y.; visualization, L.Z., P.L. and Y.Y.; supervision, X.W.; project administration, L.Z., P.L., H.Z., R.H., Y.X., F.C., X.W. and Y.Y.; funding acquisition, P.L. and R.H. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the National Natural Science Foundation of China (Grants No 12004239 and 12104311) and the Shanghai Chenguang Program (Grant No. 22CGA74).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Deng, Y.; Wang, M.; Zhuang, Y.; Liu, S.; Huang, W.; Zhao, Q. Circularly polarized luminescence from organic micro-/nano-structures. Light Sci. Appl. 2021, 10, 76. [Google Scholar] [CrossRef]
  2. Zhuang, T.T.; Li, Y.; Gao, X.; Wei, M.; de Arquer, F.P.G.; Todorović, P.; Tian, J.; Li, G.; Zhang, C.; Li, X.; et al. Regioselective magnetization in semiconducting nanorods. Nat. Nanotechnol. 2020, 15, 192–197. [Google Scholar] [CrossRef]
  3. Kandula, J.S.; Rayala, V.P.K.; Pullapanthula, R. Chirality: An inescapable concept for the pharmaceutical, bio-pharmaceutical, food, and cosmetic industries. Sep. Sci. Plus 2023, 6, 2200131. [Google Scholar] [CrossRef]
  4. Abbas, S.U.; Li, J.J.; Liu, X.; Siddique, A.; Shi, Y.X.; Hou, M.; Yang, K.; Nosheen, F.; Cui, X.Y.; Zheng, G.C.; et al. Chiral metal nanostructures: Synthesis, properties and applications. Rare Met. 2023, 42, 2489–2515. [Google Scholar] [CrossRef]
  5. Liu, J.; Yang, L.; Qin, P.; Zhang, S.; Yung, K.K.L.; Huang, Z. Recent advances in inorganic chiral nanomaterials. Adv. Mater. 2021, 33, 2005506. [Google Scholar] [CrossRef] [PubMed]
  6. Wei, H.; Wang, J.; Han, L.; Duan, Y. Chiral mesostructured Yb3+, Er3+ codoped CeO2 with upconverted circularly polarized luminescence. Chem. Eur. J. 2025, 31, e202403836. [Google Scholar] [CrossRef] [PubMed]
  7. Kotov, N.A.; Liz-Marzan, L.M.; Weiss, P.S. Chiral nanostructures: New twists. ACS Nano 2021, 15, 12457–12460. [Google Scholar] [CrossRef]
  8. Hou, X.; Song, J.; Wu, Q.; Lv, H. Chiral carbon quantum dots as fluorescent probe for rapid chiral recognition of isoleucine enantiomers. Anal. Chim. Acta 2021, 1184, 339012. [Google Scholar] [CrossRef] [PubMed]
  9. Mukhina, M.V.; Maslov, V.G.; Baranov, A.V.; Fedorov, A.V.; Orlova, A.O.; Purcell-Milton, F.; Govan, J.; Gun’ko, Y.K. Intrinsic chirality of CdSe/ZnS quantum dots and quantum rods. Nano Lett. 2015, 15, 2844–2851. [Google Scholar] [CrossRef]
  10. Tohgha, U.; Deol, K.K.; Porter, A.G.; Bartko, S.G.; Choi, J.K.; Leonard, B.M.; Varga, K.; Kubelka, J.; Muller, G.; Balaz, M. Ligand induced circular dichroism and circularly polarized luminescence in CdSe quantum dots. ACS Nano 2013, 7, 11094–11102. [Google Scholar] [CrossRef]
  11. Choi, J.K.; Haynie, B.E.; Tohgha, U.; Pap, L.; Elliott, K.W.; Leonard, B.M.; Dzyuba, S.V.; Varga, K.; Kubelka, J.; Balaz, M. Chirality inversion of CdSe and CdS quantum dots without changing the stereochemistry of the capping ligand. ACS Nano 2016, 10, 3809–3815. [Google Scholar] [CrossRef]
  12. Ma, S.; Ahn, J.; Moon, J. Chiral perovskites for next-generation photonics: From chirality transfer to chiroptical activity. Adv. Mater. 2021, 33, 2005760. [Google Scholar] [CrossRef]
  13. Ding, J.; Wang, H.; Tang, J.; Zhang, L.; Zhang, W.; Wang, Q. Chiral metal halide perovskites toward room temperature spin light-emitting diodes: Insights and perspectives. Chem. Phys. Rev. 2025, 6, 021303. [Google Scholar] [CrossRef]
  14. Wang, J.; Mao, B.; Vardeny, Z.V. Chirality induced spin selectivity in chiral hybrid organic–inorganic perovskites. J. Chem. Phys. 2023, 159, 091002. [Google Scholar] [CrossRef]
  15. He, T.; Cui, Y.; Li, J.; Gao, Y. Chiral lead halide perovskite nanocrystals: Construction strategies and photophysical properties. Adv. Quantum Technol. 2023, 6, 2300178. [Google Scholar] [CrossRef]
  16. Tan, H.; Huang, Y.; Dong, S.; Bai, Z.; Chen, C.; Wu, X.; Chao, M.; Yan, H.; Wang, S.; Geng, D.; et al. A chiral nanocomplex for multitarget therapy to alleviate neuropathology and rescue alzheimer’s cognitive deficits. Small 2023, 19, 2303530. [Google Scholar] [CrossRef]
  17. Xiao, K.; Xue, Y.; Yang, B.; Zhao, L. Ion-pairing chirality transfer in atropisomeric biaryl-centered gold clusters. CCS Chem. 2021, 2, 555–565. [Google Scholar] [CrossRef]
  18. Wang, Y.; Xu, J.; Wang, Y.; Chen, H. Emerging chirality in nanoscience. Chem. Soc. Rev. 2013, 42, 2930–2962. [Google Scholar] [CrossRef] [PubMed]
  19. Ben-Moshe, A.; Maoz, B.M.; Govorov, A.O.; Markovich, G. Chirality and chiroptical effects in inorganic nanocrystal systems with plasmon and exciton resonances. Chem. Soc. Rev. 2013, 42, 7028–7041. [Google Scholar] [CrossRef]
  20. Crassous, J. Chiral transfer in coordination complexes: Towards molecular materials. Chem. Soc. Rev. 2009, 38, 830–845. [Google Scholar] [CrossRef] [PubMed]
  21. Xu, J.; Chen, X.; Xu, Y.; Du, Y.; Yan, C. Ultrathin 2D rare-earth nanomaterials: Compositions, syntheses, and applications. Adv. Mater. 2020, 32, 1806461. [Google Scholar] [CrossRef]
  22. Power, P.P. Stable two-coordinate, open-shell (d1–d9) transition metal complexes. Chem. Rev. 2012, 112, 3482–3507. [Google Scholar] [CrossRef]
  23. Sun, L.D.; Dong, H.; Zhang, P.Z.; Yan, C.H. Upconversion of rare earth nanomaterials. Annu. Rev. Phys. Chem. 2015, 66, 619–642. [Google Scholar] [CrossRef] [PubMed]
  24. Zhang, H.; Chen, Z.H.; Liu, X.; Zhang, F. A mini-review on recent progress of new sensitizers for luminescence of lanthanide doped nanomaterials. Nano Res. 2020, 13, 1795–1809. [Google Scholar] [CrossRef]
  25. Li, C.; Wang, P.; He, M.; Yuan, X.; Fang, Z.; Li, Z. Rare earth-based nanomaterials in electrocatalysis. Coord. Chem. Rev. 2023, 489, 215204. [Google Scholar] [CrossRef]
  26. Jin, X.; Sang, Y.; Shi, Y.; Li, Y.; Zhu, X.; Duan, P.; Liu, M. Optically active upconverting nanoparticles with induced circularly polarized luminescence and enantioselectively triggered photopolymerization. ACS Nano 2019, 13, 2804–2811. [Google Scholar] [CrossRef] [PubMed]
  27. Kwon, J.; Choi, W.J.; Jeong, U.; Jung, W.; Hwang, I.; Park, K.H.; Ko, S.G.; Park, S.M.; Kotov, N.A.; Yeom, J. Recent advances in chiral nanomaterials with unique electric and magnetic properties. Nano Converg. 2022, 9, 32. [Google Scholar] [CrossRef]
  28. Wang, F.; Yue, X.; Ding, Q.; Lin, H.; Xu, C.; Li, S. Chiral inorganic nanomaterials for biological applications. Nanoscale 2023, 15, 2541–2552. [Google Scholar] [CrossRef]
  29. Chen, Y.; Dong, S.; Wang, P. Recent advances in circularly polarized luminescence of chiral rare earth nanomaterials. Sci. Sin. Chim. 2024, 54, 1178–1193. [Google Scholar] [CrossRef]
  30. Jiang, S.; Kotov, N.A. Circular polarized light emission in chiral inorganic nanomaterials. Adv. Mater. 2023, 35, 2108431. [Google Scholar] [CrossRef]
  31. Niu, X.; Zeng, Z.; Wang, Z.; Lu, H.; Sun, B.; Zhang, H.L.; Chen, Y.; Du, Y.; Long, G. The first chiral cerium halide towards circularly-polarized luminescence in the UV region. Sci. China Chem. 2024, 67, 1961–1968. [Google Scholar] [CrossRef]
  32. Tan, L.; Li, J.; Jin, Y.; Wen, Z.; Cheng, Y.; Fu, W.; Wang, P.P. Multicolor circularly polarized luminescence from inorganic crystalline nanostructures induced by atomic chirality. Nano Lett. 2023, 23, 4384–4389. [Google Scholar] [CrossRef] [PubMed]
  33. Yuan, W.; Lu, S.; Li, X.; Li, Z.; Gu, X.; Liu, X.; Ren, Z.; Wang, F.; Chen, X. Helically assembled rare earth fluoride nanoparticles with multicolor circularly polarized luminescence for high-security anti-counterfeiting. Aggregate 2025, 6, e70042. [Google Scholar] [CrossRef]
  34. Zhang, H.; Wei, H.; Xu, L.; Li, Y.; Song, Z.; Zhou, D.; Wang, Q.; Long, Z.; Yang, Y.; Wen, Y.; et al. Chiral inorganic nanostructured BiOCl co-doped with Er3+/Yb3+ exhibits circularly polarized luminescence and enhanced upconversion luminescence. Ceram. Int. 2023, 49, 30436–30442. [Google Scholar] [CrossRef]
  35. Zhang, X.; Xu, Y.; Valenzuela, C.; Zhang, X.; Wang, L.; Feng, W.; Li, Q. Liquid crystal-templated chiral nanomaterials: From chiral plasmonics to circularly polarized luminescence. Light-Sci. Appl. 2022, 11, 223. [Google Scholar] [CrossRef]
  36. Francés-Soriano, L.; Zaballos-García, E.; Pérez-Prieto, J. Up-and down-shifting nanomaterials with chiroptical responses: Origin of their chiroptical activity and applications. Adv. Opt. Mater. 2023, 11, 2300337. [Google Scholar] [CrossRef]
  37. Fu, W.; Tan, L.; Wang, P. Chiral inorganic nanomaterials for photo (electro) catalytic conversion. ACS Nano 2023, 17, 16326–16347. [Google Scholar] [CrossRef]
  38. Puri, M.; Ferry, V.E. Circular dichroism of CdSe nanocrystals bound by chiral carboxylic acids. ACS Nano 2017, 11, 12240–12246. [Google Scholar] [CrossRef]
  39. Bentley, R. The nose as a stereochemist. Enantiomers and odor. Chem. Rev. 2006, 106, 4099–4112. [Google Scholar] [CrossRef]
  40. Ben-Moshe, A.; Govorov, A.O.; Markovich, G. Enantioselective synthesis of intrinsically chiral mercury sulfide nanocrystals. Angew. Chem. Int. Ed. 2013, 52, 1275–1279. [Google Scholar] [CrossRef]
  41. Hananel, U.; Ben-Moshe, A.; Diamant, H.; Markovich, G. Spontaneous and directed symmetry breaking in the formation of chiral nanocrystals. Proc. Natl. Acad. Sci. USA 2019, 116, 11159–11164. [Google Scholar] [CrossRef]
  42. Lan, X.; Lu, X.; Shen, C.; Ke, Y.; Ni, W.; Wang, Q. Au nanorod helical superstructures with designed chirality. J. Am. Chem. Soc. 2015, 137, 457–462. [Google Scholar] [CrossRef]
  43. Zhu, J.; Wu, F.; Han, Z.; Shang, Y.; Liu, F.; Yu, H.; Yu, L.; Li, N.; Ding, B. Strong light–matter interactions in chiral plasmonic–excitonic systems assembled on DNA origami. Nano Lett. 2021, 21, 3573–3580. [Google Scholar] [CrossRef] [PubMed]
  44. Kuzyk, A.; Schreiber, R.; Zhang, H.; Govorov, A.O.; Liedl, T.; Liu, N. Reconfigurable 3D plasmonic metamolecules. Nat. Mater. 2014, 13, 862–866. [Google Scholar] [CrossRef] [PubMed]
  45. Wang, L.; Urbas, A.M.; Li, Q. Nature-inspired emerging chiral liquid crystal nanostructures: From molecular self-assembly to DNA mesophase and nanocolloids. Adv. Mater. 2020, 32, 1801335. [Google Scholar] [CrossRef]
  46. He, Y.; Lin, S.; Guo, J.; Li, Q. Circularly polarized luminescent self-organized helical superstructures: From materials and stimulus-responsiveness to applications. Aggregate 2021, 2, 141. [Google Scholar] [CrossRef]
  47. Gonçalves, D.P.; Prévôt, M.E.; Üstünel, Ş.; Ogolla, T.; Nemati, A.; Shadpour, S.; Hegmann, T. Recent progress at the interface between nanomaterial chirality and liquid crystals. Liq. Cryst. Rev. 2021, 9, 1–34. [Google Scholar] [CrossRef]
  48. Cheng, G.; Xu, D.; Lu, Z.; Liu, K. Chiral self-assembly of nanoparticles induced by polymers synthesized via reversible addition–fragmentation chain transfer polymerization. ACS Nano 2019, 13, 1479–1489. [Google Scholar] [CrossRef]
  49. Chiu, P.T.; Yang, C.Y.; Xie, Z.H.; Chang, M.Y.; Hung, Y.C.; Ho, R.M. Gold nanohelices for chiral plasmonic films by templated electroless plating. Adv. Opt. Mater. 2021, 9, 2002133. [Google Scholar] [CrossRef]
  50. Li, A.; Zheng, D.; Zhang, M.; Wu, B.; Zhu, L. Chirality transfer in carbon dot-composited sol–gel systems for excitation-dependent circularly polarized luminescence. Langmuir 2020, 36, 8965–8970. [Google Scholar] [CrossRef]
  51. Shi, Y.; Duan, P.; Huo, S.; Li, Y.; Liu, M. Endowing perovskite nanocrystals with circularly polarized luminescence. Adv. Mater. 2018, 30, 1705011. [Google Scholar] [CrossRef] [PubMed]
  52. Jung, S.H.; Jeon, J.; Kim, H.; Jaworski, J.; Jung, J.H. Chiral arrangement of achiral Au nanoparticles by supramolecular assembly of helical nanofiber templates. J. Am. Chem. Soc. 2014, 136, 6446–6452. [Google Scholar] [CrossRef]
  53. Albano, G.; Pescitelli, G.; Di Bari, L. Chiroptical properties in thin films of π-conjugated systems. Chem. Rev. 2020, 120, 10145–10243. [Google Scholar] [CrossRef] [PubMed]
  54. Maniappan, S.; Jadhav, A.B.; Kumar, J. Template assisted generation of chiral luminescence in organic fluorophores. Front. Chem. 2021, 8, 557650. [Google Scholar] [CrossRef]
  55. Tohgha, U.; Varga, K.; Balaz, M. Achiral CdSe quantum dots exhibit optical activity in the visible region upon post-synthetic ligand exchange with D-or L-cysteine. Chem. Commun. 2013, 49, 1844–1846. [Google Scholar] [CrossRef]
  56. Barry, D.E.; Kitchen, J.A.; Albrecht, M.; Faulkner, S.; Gunnlaugsson, T. Near infrared (NIR) lanthanide emissive langmuir–blodgett monolayers formed using Nd (III) directed self-assembly synthesis of chiral amphiphilic ligands. Langmuir 2013, 29, 11506–11515. [Google Scholar] [CrossRef] [PubMed]
  57. Zhao, W.; Huang, X.; Hu, S.; Yang, F.; Zhong, J. Rare-earth nanocrystalline scintillators for biomedical application: A review. Ceram. Int. 2025, 51, 16936–16950. [Google Scholar] [CrossRef]
  58. Zhang, C.; Li, Z.S.; Dong, X.Y.; Niu, Y.Y.; Zang, S.Q. Multiple responsive CPL switches in an enantiomeric pair of perovskite confined in lanthanide MOFs. Adv. Mater. 2022, 34, 2109496. [Google Scholar] [CrossRef]
  59. Chen, H.; Gu, Z.G.; Zhang, J. Chiral-induced ultrathin covalent organic frameworks nanosheets with tunable circularly polarized luminescence. J. Am. Chem. Soc. 2022, 144, 7245–7252. [Google Scholar] [CrossRef]
  60. Zhai, R.; Xiao, Y.; Gu, Z.; Zhang, J. Tunable chiroptical application by encapsulating achiral lanthanide complexes into chiral MOF thin films. Nano Res. 2022, 15, 1102–1108. [Google Scholar] [CrossRef]
  61. Katsoulidis, A.P.; Antypov, D.; Whitehead, G.F.; Carrington, E.J.; Adams, D.J.; Berry, N.G.; Darling, G.R.; Dyer, M.S.; Rosseinsky, M.J. Chemical control of structure and guest uptake by a conformationally mobile porous material. Nature 2019, 565, 213–217. [Google Scholar] [CrossRef]
  62. Zhao, T.; Han, J.; Jin, X.; Liu, Y.; Liu, M.; Duan, P. Enhanced circularly polarized luminescence from reorganized chiral emitters on the skeleton of a zeolitic imidazolate framework. Angew. Chem. 2019, 131, 5032–5036. [Google Scholar] [CrossRef]
  63. Huang, X.C.; Lin, Y.Y.; Zhang, J.P.; Chen, X.M. Ligand-directed strategy for zeolite-type metal–organic frameworks: Zinc (II) imidazolates with unusual zeolitic topologies. Angew. Chem. Int. Ed. 2006, 45, 1557–1559. [Google Scholar] [CrossRef]
  64. Park, K.S.; Ni, Z.; Côté, A.P.; Choi, J.Y.; Huang, R.; Uribe-Romo, F.J.; Chae, H.K.; O’Keeffe, M.; Yaghi, O.M. Exceptional chemical and thermal stability of zeolitic imidazolate frameworks. Proc. Natl. Acad. Sci. USA 2006, 103, 10186–10191. [Google Scholar] [CrossRef] [PubMed]
  65. Zhao, T.; Han, J.; Jin, X.; Zhou, M.; Liu, Y.; Duan, P.; Liu, M. Dual-mode induction of tunable circularly polarized luminescence from chiral metal-organic frameworks. Research 2020, 2020, 6452123. [Google Scholar] [CrossRef] [PubMed]
  66. Zhao, T.; Han, J.; Shi, Y.; Zhou, J.; Duan, P. Multi-light-responsive upconversion-and-downshifting-based circularly polarized luminescent switches in chiral metal–organic frameworks. Adv. Mater. 2021, 33, 2101797. [Google Scholar] [CrossRef] [PubMed]
  67. Hao, T.; Xu, B.; Wang, X.; Zheng, H.; Li, S.; Wang, F.; Zhang, J. Enhanced circularly polarized luminescence in rare-earth MOFs via framework chirality and host–guest energy transfer. Polyoxometalates 2025, 4, 9140095. [Google Scholar] [CrossRef]
  68. Shi, Y.; Zhou, Z.; Miao, X.; Liu, Y.J.; Fan, Q.; Wang, K.; Luo, D.; Sun, X.W. Circularly polarized luminescence from semiconductor quantum rods templated by self-assembled cellulose nanocrystals. J. Mater. Chem. C 2020, 8, 1048–1053. [Google Scholar] [CrossRef]
  69. Lu, D.; Li, M.; Gao, X.; Yu, X.; Wei, L.; Zhu, S.; Xu, Y. Cellulose nanocrystal films with NIR-II circularly polarized light for cancer detection applications. ACS Nano 2022, 17, 461–471. [Google Scholar] [CrossRef]
  70. Nguyen, T.D.; Hamad, W.Y.; MacLachlan, M.J. Near-IR-sensitive upconverting nanostructured photonic cellulose films. Adv. Opt. Mater. 2017, 5, 1600514. [Google Scholar] [CrossRef]
  71. Kang, S.; Biesold, G.M.; Lee, H.; Bukharina, D.; Lin, Z.; Tsukruk, V.V. Dynamic chiro-optics of bio-inorganic nanomaterials via seamless co-assembly of semiconducting nanorods and polysaccharide nanocrystals. Adv. Funct. Mater. 2021, 31, 2104596. [Google Scholar] [CrossRef]
  72. Sharma, V.; Crne, M.; Park, J.O.; Srinivasarao, M. Structural origin of circularly polarized iridescence in jeweled beetles. Science 2009, 325, 449–451. [Google Scholar] [CrossRef]
  73. Chiou, T.H.; Kleinlogel, S.; Cronin, T.; Caldwell, R.; Loeffler, B.; Siddiqi, A.; Goldizen, A.; Marshall, J. Circular polarization vision in a stomatopod crustacean. Curr. Biol. 2008, 18, 429–434. [Google Scholar] [CrossRef]
  74. Li, W.; Xu, M.; Ma, C.; Liu, Y.; Zhou, J.; Chen, Z.; Wang, Y.; Yu, H.; Li, J.; Liu, S. Tunable upconverted circularly polarized luminescence in cellulose nanocrystal based chiral photonic films. ACS Appl. Mater. Interfaces 2019, 11, 23512–23519. [Google Scholar] [CrossRef]
  75. Zhou, X.; Jin, Q.; Zhang, L.; Shen, Z.; Jiang, L.; Liu, M. Self-Assembly of hierarchical chiral nanostructures based on metal–benzimidazole interactions: Chiral nanofibers, nanotubes, and microtubular flowers. Small 2016, 34, 4743–4752. [Google Scholar] [CrossRef]
  76. Xu, L.; Guo, M.; Hung, C.T.; Shi, X.L.; Yuan, Y.; Zhang, X.; Jin, R.H.; Li, W.; Dong, Q.; Zhao, D. Chiral skeletons of mesoporous silica nanospheres to mitigate alzheimer’s β-amyloid aggregation. J. Am. Chem. Soc. 2023, 145, 7810–7819. [Google Scholar] [CrossRef]
  77. Wang, Y.; Li, W.; He, Z.; Yin, W.; Chen, X.; Zhang, J.; Li, H. Multichiral mesoporous silica screws with chiral differential mucus penetration and mucosal adhesion for oral drug delivery. ACS Nano 2024, 18, 16166–16183. [Google Scholar] [CrossRef] [PubMed]
  78. Delclos, T.; Aimé, C.; Pouget, E.; Brizard, A.; Huc, I.; Delville, M.H.; Oda, R. Individualized silica nanohelices and nanotubes: Tuning inorganic nanostructures using lipidic self-assemblies. Nano Lett. 2008, 8, 1929–1935. [Google Scholar] [CrossRef] [PubMed]
  79. Sugimoto, M.; Liu, X.L.; Tsunega, S.; Nakajima, E.; Abe, S.; Nakashima, T.; Kawai, T.; Jin, R.H. Circularly polarized luminescence from inorganic materials: Encapsulating guest lanthanide oxides in chiral silica hosts. Chem. Eur. J. 2018, 24, 6519–6524. [Google Scholar] [CrossRef] [PubMed]
  80. Liu, P.; Battie, Y.; Kimura, T.; Okazaki, Y.; Pranee, P.; Wang, H.; Pouget, P.; Nlate, S.; Sagawa, T.; Oda, R. Chiral perovskite nanocrystal growth inside helical hollow silica nanoribbons. Nano Lett. 2023, 23, 3174–3180. [Google Scholar] [CrossRef]
  81. Liu, P.; Chen, W.; Okazaki, Y.; Battie, Y.; Brocard, L.; Decossas, M.; Pouget, E.; Müller-Buschbaum, P.; Kauffmann, B.; Pathan, S.; et al. Optically active perovskite CsPbBr3 nanocrystals helically arranged on inorganic silica nanohelices. Nano Lett. 2020, 20, 8453–8460. [Google Scholar] [CrossRef] [PubMed]
  82. Zhou, M.; Sang, Y.; Jin, X.; Chen, S.; Guo, J.; Duan, P.; Liu, M. Steering nanohelix and upconverted circularly polarized luminescence by using completely achiral components. ACS Nano 2021, 15, 2753–2761. [Google Scholar] [CrossRef] [PubMed]
  83. Sang, Y.; Han, J.; Zhao, T.; Duan, P.; Liu, M. Circularly polarized luminescence in nanoassemblies: Generation, amplification, and application. Adv. Mater. 2020, 32, 1900110. [Google Scholar] [CrossRef]
  84. Hiroyuki, M.; Ren-Hua, J. High-temperature-resistant chiral silica generated on chiral crystalline templates at neutral pH and ambient conditions. Angew. Chem. Int. Edit. 2012, 51, 5862–5865. [Google Scholar]
  85. Zhao, K.; Zhang, L.; Wang, J.; Li, Q.; He, W.; Yin, J.J. Surface structure-dependent molecular oxygen activation of BiOCl single-crystalline nanosheets. J. Am. Chem. Soc. 2013, 135, 15750–15753. [Google Scholar] [CrossRef]
  86. Binnemans, K. Lanthanide-based luminescent hybrid materials. Chem. Rev. 2009, 109, 4283–4374. [Google Scholar] [CrossRef]
  87. Pandya, S.; Yu, J.; Parker, D. Engineering emissive europium and terbium complexes for molecular imaging and sensing. Dalton Trans. 2006, 23, 2757–2766. [Google Scholar] [CrossRef]
  88. Eliseeva, S.V.; Bünzli, J.C.G. Lanthanide luminescence for functional materials and bio-sciences. Chem. Soc. Rev. 2010, 39, 189–227. [Google Scholar] [CrossRef]
  89. Zhangsun, H.; Xu, L.; Qu, A.; Xu, C.; Kuang, H.; Hao, C. Generating strong circularly polarized luminescence from self-assembled films of chiral selenium nanoparticles and upconversion nanoparticles. Chem. Int. Ed. 2025, 64, e202419884. [Google Scholar] [CrossRef] [PubMed]
  90. Collins, J.T.; Kuppe, C.; Hooper, D.C.; Sibilia, C.; Centini, M.; Valev, V.K. Chirality and chiroptical effects in metal nanostructures: Fundamentals and current trends. Adv. Opt. Mater. 2017, 5, 1700182. [Google Scholar] [CrossRef]
  91. Nizar, N.S.; Sujith, M.; Swathi, K.; Sissa, C.; Painelli, A.; Thomas, K.G. Emergent chiroptical properties in supramolecular and plasmonic assemblies. Chem. Soc. Rev. 2021, 50, 11208–11226. [Google Scholar] [CrossRef]
  92. Mun, J.; Kim, M.; Yang, Y.; Badloe, T.; Ni, J.; Chen, Y.; Qiu, C.W.; Rho, J. Electromagnetic chirality: From fundamentals to nontraditional chiroptical phenomena. Light-Sci. Appl. 2020, 9, 139. [Google Scholar] [CrossRef]
  93. Stalder, M.; Schadt, M. Linearly polarized light with axial symmetry generated by liquid-crystal polarization converters. Opt. Lett. 1996, 21, 1948–1950. [Google Scholar] [CrossRef] [PubMed]
  94. Nikolova, L.; Todorov, T.; Ivanov, M.T.; Andruzzi, F.; Hvilsted, S.; Ramanujam, P.S. Polarization holographic gratings in side-chain azobenzene polyesters with linear and circular photoanisotropy. Appl. Opt. 1996, 35, 3835–3840. [Google Scholar] [CrossRef] [PubMed]
  95. Trippe, S. Polarization and polarimetry: A review. J. Korean Astron. Soc. 2014, 1401, 1911. [Google Scholar] [CrossRef]
  96. Han, J.; Guo, S.; Lu, H.; Liu, S.; Zhao, Q.; Huang, W. Recent progress on circularly polarized luminescent materials for organic optoelectronic devices. Adv. Opt. Mater. 2018, 6, 1800538. [Google Scholar] [CrossRef]
  97. Warning, L.A.; Miandashti, A.R.; McCarthy, L.A.; Zhang, Q.; Landes, C.F.; Link, S. Nanophotonic approaches for chirality sensing. ACS Nano 2021, 15, 15538–15566. [Google Scholar] [CrossRef]
  98. Moroni, M.; Coccia, C.; Malavasi, L. Chiral 2D and quasi-2D hybrid organic inorganic perovskites: From fundamentals to applications. Chem. Commun. 2024, 60, 9310. [Google Scholar] [CrossRef] [PubMed]
  99. Sánchez-Carnerero, E.M.; Agarrabeitia, A.R.; Moreno, F.; Maroto, B.L.; Muller, G.; Ortiz, M.J.; de la Moya, S. Circularly polarized luminescence from simple organic molecules. Chem. Eur. J. 2015, 21, 13488–13500. [Google Scholar] [CrossRef] [PubMed]
  100. Carr, R.; Evans, N.H.; Parker, D. Lanthanide complexes as chiral probes exploiting circularly polarized luminescence. Chem. Soc. Rev. 2012, 41, 7673–7686. [Google Scholar] [CrossRef]
  101. Ma, B.; Wu, Y.; Zhang, S.; Wang, S.; Qiu, J.; Zhao, L.; Guo, D.; Duan, J.; Sang, Y.; Li, L.; et al. Terbium–aspartic acid nanocrystals with chirality-dependent tunable fluorescent properties. ACS Nano 2017, 11, 1973–1981. [Google Scholar] [CrossRef]
  102. Muller, G. Luminescent chiral lanthanide (III) complexes as potential molecular probes. Dalton Trans. 2009, 44, 9692–9707. [Google Scholar] [CrossRef] [PubMed]
  103. Yang, X.; Zhou, M.; Wang, Y.; Duan, P. Electric-field-regulated energy transfer in chiral liquid crystals for enhancing upconverted circularly polarized luminescence through steering the photonic bandgap. Adv. Mater. 2020, 32, 2000820. [Google Scholar] [CrossRef]
  104. Jin, X.; Zhou, M.; Han, J.; Li, B.; Zhang, T.; Jiang, S.; Duan, P. A new strategy to achieve enhanced upconverted circularly polarized luminescence in chiral perovskite nanocrystals. Nano Res. 2022, 15, 1047–1053. [Google Scholar] [CrossRef]
  105. Niu, X.; Li, Y.; Lu, H.; Wang, Z.; Zhang, Y.; Shao, T.; Wang, H.; Gull, S.; Sun, B.; Zhang, H.L.; et al. Chiral europium halides with high-performance magnetic field tunable red circularly polarized luminescence at room temperature. Nat. Commun. 2025, 16, 2525. [Google Scholar] [CrossRef]
  106. MacKenzie, L.E.; Robert, P. Circularly polarized lanthanide luminescence for advanced security inks. Nat. Rev. Chem. 2021, 5, 109–124. [Google Scholar] [CrossRef]
  107. Kitagawa, Y.; Wada, S.; Islam, M.D.J.; Saita, K.; Gon, M.; Fushimi, K.; Tanaka, K.; Maeda, S.; Hasegawa, Y. Chiral lanthanide lumino-glass for a circularly polarized light security device. Commun. Chem. 2020, 3, 119. [Google Scholar] [CrossRef] [PubMed]
  108. Rezende, T.K.; Barbosa, H.P.; Dos Santos, L.F.; Lima, K.d.O.; de Matos, P.A.; Tsubone, T.M.; Gonçalves, R.R.; Ferrari, J.L. Upconversion rare earths nanomaterials applied to photodynamic therapy and bioimaging. Front. Chem. 2022, 10, 1035449. [Google Scholar] [CrossRef]
  109. Meng, J.; Cui, Y.; Wang, Y. Rare earth-doped nanocrystals for bioimaging in the near-infrared region. J. Mater. Chem. B 2022, 10, 8596–8615. [Google Scholar] [CrossRef]
  110. Chen, Z.; Wang, X.; Yang, M.; Ming, J.; Yun, B.; Zhang, L.; Wang, X.; Yu, P.; Xu, J.; Zhang, H.; et al. An extended NIR-II superior imaging window from 1500 to 1900 nm for high-resolution in vivo multiplexed imaging based on lanthanide nanocrystals. Angew. Chem. Int. Ed. 2023, 62, e202311883. [Google Scholar] [CrossRef] [PubMed]
  111. Meng, J.; Cui, Y.; Wang, Y. Rare earth-doped nanoparticles for near-infrared II image-guided photodynamic therapy. ACS Appl. Nano Mater. 2024, 7, 11008–11018. [Google Scholar] [CrossRef]
  112. Gao, Q.; Tan, L.; Wen, Z.; Fan, D.; Hui, J.; Wang, P.P. Chiral inorganic nanomaterials: Harnessing chirality-dependent interactions with living entities for biomedical applications. Nano Res. 2023, 16, 11107–11124. [Google Scholar] [CrossRef]
  113. Zhao, C.; Qu, X. Recent progress on molecular recognition and modulation of nucleic acids using chiral rare-earth complexes. Prog. Chem. 2013, 25, 539–544. [Google Scholar]
  114. Wu, X.; Xu, L.; Ma, W.; Liu, L.; Kuang, H.; Kotov, N.A.; Xu, C. Propeller-like nanorod-upconversion nanoparticle assemblies with intense chiroptical activity and luminescence enhancement in aqueous phase. Adv. Mater. 2016, 28, 5907–5915. [Google Scholar] [CrossRef]
  115. Liang, P.; Wang, Y.; Cheng, L.; Hu, R.; Men, Y.; Li, J.; Jia, T.; Sun, Z.; Feng, D. Impact of photo-irradiation on the optical and spin properties of chiral CdS quantum dot films. Appl. Phys. Lett. 2025, 126, 151902. [Google Scholar] [CrossRef]
  116. Wang, Y.; Liang, P.; Men, Y.; Jiang, M.; Cheng, L.; Li, J.; Jia, T.; Sun, Z.; Feng, D. Light-induced photoluminescence enhancement in chiral CdSe quantum dot films. J. Chem. Phys. 2024, 160, 161102. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Synthesis methods, optical properties, and application of chiral rare earth nanomaterials.
Figure 1. Synthesis methods, optical properties, and application of chiral rare earth nanomaterials.
Nanomaterials 15 01321 g001
Figure 2. Schematic diagram of chirality origin of chiral rare earth nanomaterials. (a) Intrinsically chiral crystal. (b) Co-assembly between nanomaterials and a chiral template. (c) Chiral transfer between inorganic nanomaterials and chiral ligands.
Figure 2. Schematic diagram of chirality origin of chiral rare earth nanomaterials. (a) Intrinsically chiral crystal. (b) Co-assembly between nanomaterials and a chiral template. (c) Chiral transfer between inorganic nanomaterials and chiral ligands.
Nanomaterials 15 01321 g002
Figure 3. Schematic illustration of the rational design of chiral nanocrystals with atomic-scale chirality for tunable color circularly polarized luminescence emission [32].
Figure 3. Schematic illustration of the rational design of chiral nanocrystals with atomic-scale chirality for tunable color circularly polarized luminescence emission [32].
Nanomaterials 15 01321 g003
Figure 5. Schemes of (a) the preparation procedure for the 2D heterostructure films through interfacial self-assembly of D-/L-Se NPs and DPNUCNPs; (b) enhanced-CPL by controlling the layer number of D-/L-Se/DPNUCNPs films under 980 nm excitation [89].
Figure 5. Schemes of (a) the preparation procedure for the 2D heterostructure films through interfacial self-assembly of D-/L-Se NPs and DPNUCNPs; (b) enhanced-CPL by controlling the layer number of D-/L-Se/DPNUCNPs films under 980 nm excitation [89].
Nanomaterials 15 01321 g005
Figure 6. Schematic illustration of circular dichroism (CD) (a) and circularly polarized luminescence (CPL) (b).
Figure 6. Schematic illustration of circular dichroism (CD) (a) and circularly polarized luminescence (CPL) (b).
Nanomaterials 15 01321 g006
Figure 7. (a) XRD patterns of L/R-HS@CeF3: Tb. (b) SAXS patterns of R-HS and R-HS@CeF3: Tb nanocomposites (the insets represent two-dimensional SAXS images). PL excitation spectrum under 542 nm emission (c) and PL emission spectrum under 254 nm excitation (d) of R-HS@CeF3: Tb. (e) CD spectra of L/R-HS@CeF3: Tb. (f) CPL spectra of CeF3: Tb and L/R-HS@CeF3: Tb under 254 nm excitation (the inset represents the corresponding image of R-HS@CeF3: Tb upon 254 nm UV irradiation) [33].
Figure 7. (a) XRD patterns of L/R-HS@CeF3: Tb. (b) SAXS patterns of R-HS and R-HS@CeF3: Tb nanocomposites (the insets represent two-dimensional SAXS images). PL excitation spectrum under 542 nm emission (c) and PL emission spectrum under 254 nm excitation (d) of R-HS@CeF3: Tb. (e) CD spectra of L/R-HS@CeF3: Tb. (f) CPL spectra of CeF3: Tb and L/R-HS@CeF3: Tb under 254 nm excitation (the inset represents the corresponding image of R-HS@CeF3: Tb upon 254 nm UV irradiation) [33].
Nanomaterials 15 01321 g007
Figure 8. Humidity-dependent UC-CPL observed in the glycerol-integrated photonic film. CPL spectra (a) and the corresponding glum distribution curves (b) of film under varying relative humidity (RH) conditions [74].
Figure 8. Humidity-dependent UC-CPL observed in the glycerol-integrated photonic film. CPL spectra (a) and the corresponding glum distribution curves (b) of film under varying relative humidity (RH) conditions [74].
Nanomaterials 15 01321 g008
Figure 9. The illustration shows (ae) anti-counterfeiting patterns and (f) multilayer optical information encryption codes based on L/RHS@CeF3:RE nanocomposites. (a) Preparation process of luminescent plant pattern. (b) Photographs of patterns under daylight and UV light. (c) Time-dependent CIE chromaticity diagram for petal (left) and stamen (right). (d) Dynamic PL changes in flower stamen and petal. (e) CPL spectra of pattern under 254 nm UV excitation. (f) Demonstration of the step-by-step decryption of multicolor codes through the modes of UV light, TRPL, and CPL. Insets show the arrayed nanocomposites in a 96-well plate, the corresponding photograph under 254 nm irradiation, and the CIE chromaticity diagram at a delay time of 1.5 ms [33].
Figure 9. The illustration shows (ae) anti-counterfeiting patterns and (f) multilayer optical information encryption codes based on L/RHS@CeF3:RE nanocomposites. (a) Preparation process of luminescent plant pattern. (b) Photographs of patterns under daylight and UV light. (c) Time-dependent CIE chromaticity diagram for petal (left) and stamen (right). (d) Dynamic PL changes in flower stamen and petal. (e) CPL spectra of pattern under 254 nm UV excitation. (f) Demonstration of the step-by-step decryption of multicolor codes through the modes of UV light, TRPL, and CPL. Insets show the arrayed nanocomposites in a 96-well plate, the corresponding photograph under 254 nm irradiation, and the CIE chromaticity diagram at a delay time of 1.5 ms [33].
Nanomaterials 15 01321 g009
Figure 10. (a) Schematic illustration for DNA biosensing. The CD (b) and upconversion luminescence (d) curves with increasing concentrations of DNA solution. The CD (c) and upconversion luminescent (e) calibration curves for DNA detection [114].
Figure 10. (a) Schematic illustration for DNA biosensing. The CD (b) and upconversion luminescence (d) curves with increasing concentrations of DNA solution. The CD (c) and upconversion luminescent (e) calibration curves for DNA detection [114].
Nanomaterials 15 01321 g010
Table 1. Summary of the maximum representative glum in this review. (BTABA = benzene-1,3,5-tricarboxamid with three identical benzoic acid arms, NCs = nanocrystals, DAEC = diarylethene derivative).
Table 1. Summary of the maximum representative glum in this review. (BTABA = benzene-1,3,5-tricarboxamid with three identical benzoic acid arms, NCs = nanocrystals, DAEC = diarylethene derivative).
ChiralityChirality OriginCPL-Active SystemMaximum glumReference
IntrinsicChiral latticeGdPO4:Eu3+ phosphate1.1 × 10−1[32]
Chiral latticeNaYF4:Eu3+ phosphate4 × 10−1[41]
Non-intrinsicChiral helical nanotubesNaYF4:Yb/Tm UCNPs5.48 × 10−3[26]
Helical silicaCeF3:Tb3+ nanoparticles4.7 × 10−3[33]
Chiral MOFNaYF4:Yb/Er UCNPs1.2 × 10−2[65]
Cellulose nanocrystalNaYF4:Tm/Yb UCNPs1.56 × 10−1[74]
Chiral nematic liquid crystalUCNPs and CsPbBr3 NCs1.1[103]
Chiral MOFDAEC and UCNP-Tm7.8 × 10−2[66]
Chiral CsPbBr3 NCsNaYF4:Yb/Tm UCNPs5.0 × 10−3[104]
BTABANaYF4:Yb/Er nanoparticles1.2 × 10−2[82]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhao, L.; Liang, P.; Zhao, H.; Hu, R.; Xu, Y.; Chen, F.; Wang, X.; Yao, Y. Chiral Rare Earth Nanomaterials: Synthesis, Optical Properties, and Potential Applications. Nanomaterials 2025, 15, 1321. https://doi.org/10.3390/nano15171321

AMA Style

Zhao L, Liang P, Zhao H, Hu R, Xu Y, Chen F, Wang X, Yao Y. Chiral Rare Earth Nanomaterials: Synthesis, Optical Properties, and Potential Applications. Nanomaterials. 2025; 15(17):1321. https://doi.org/10.3390/nano15171321

Chicago/Turabian Style

Zhao, Lei, Pan Liang, Hua Zhao, Rongrong Hu, Yangyang Xu, Fangfang Chen, Xianghu Wang, and Yunhua Yao. 2025. "Chiral Rare Earth Nanomaterials: Synthesis, Optical Properties, and Potential Applications" Nanomaterials 15, no. 17: 1321. https://doi.org/10.3390/nano15171321

APA Style

Zhao, L., Liang, P., Zhao, H., Hu, R., Xu, Y., Chen, F., Wang, X., & Yao, Y. (2025). Chiral Rare Earth Nanomaterials: Synthesis, Optical Properties, and Potential Applications. Nanomaterials, 15(17), 1321. https://doi.org/10.3390/nano15171321

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop