Next Article in Journal
pH-Responsive Nanophotosensitizer Boosting Antibacterial Photodynamic Therapy by Hydroxyl Radical Generation
Previous Article in Journal
Advancements in Nanotechnology for Spinal Surgery: Innovations in Spinal Fixation Devices for Enhanced Biomechanical Performance and Osteointegration
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Multifunctional Upconversion Nanoparticles Transforming Photoacoustic Imaging: A Review

1
Nepean Hospital, Kingswood, NSW 2747, Australia
2
School of Chemistry and Chemical Engineering, Jiangsu University, Zhenjiang 212013, China
3
School of Materials Science and Engineering, University of New South Wales, Sydney, NSW 2052, Australia
4
School of Natural Sciences, Macquarie University, Sydney, NSW 2109, Australia
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2025, 15(14), 1074; https://doi.org/10.3390/nano15141074
Submission received: 29 May 2025 / Revised: 1 July 2025 / Accepted: 8 July 2025 / Published: 10 July 2025

Abstract

Photoacoustic imaging (PAI) merges the high spatial resolution of optical methods with the deep tissue penetration provided by ultrasound, making it a valuable tool in biomedical imaging. In recent years, a diverse array of photoacoustic contrast agents, spanning both organic and inorganic materials, has been developed. Among them, upconversion nanoparticles (UCNPs) stand out as promising candidates due to their unique optical features, tunable absorption in the near-infrared I (NIR-I, 750–1350 nm) region, and strong potential for both imaging and treatment-related uses. This review discusses the growing significance of UCNPs in the field of PAI, focusing on their structural characteristics, strengths, and existing challenges. Then, we talk about an up-to-date account of the current literature on the use of UCNPs as contrast agents for PAI. Lastly, we discuss the challenges and perspectives of UCNPs as a contrast agent for PAI in preclinical research and clinical diagnosis.

1. Introduction

Numerous non-invasive imaging technologies, for instance, magnetic resonance imaging (MRI), computed tomography (CT), positron emission tomography (PET), and ultrasound (US) [1,2,3,4,5,6], have been extensively applied in the biomedical field. Among these, optical imaging stands out as a convenient method for real-time visualization and quantification of biological processes in living organisms. It avoids the use of ionizing radiation, making it especially attractive for both clinical applications and biological studies [7]. In recent developments, hybrid imaging techniques like X-ray excited optical luminescence (XEOL) combined with PET have been introduced [8], leveraging multiple physical principles to enhance imaging resolution, depth, speed, and three-dimensional reconstruction beyond what single modalities can achieve. Despite these advancements, conventional optical imaging alone has struggled to deliver high-resolution images at centimeter-level depths—until the emergence of photoacoustic imaging (PAI) [9].
PAI, a relatively recent innovation, synergizes the strengths of optical and ultrasound imaging to enable highly sensitive visualization of optical probes [10,11,12]. Over the past decade, PAI has gained prominence as a technique that relies on laser-induced ultrasound generation. It was in 1880 that Alexander Graham Bell first noted the photoacoustic (PA) response, observing sound waves generated by light absorption [13]. PAI has gained traction since the 1990s because of progress in laser technology and ultrasound detectors (Figure 1) [14]. PAI offers impressive spatial (around 100 µm) and temporal (approximately 1 ms) resolution, constrained primarily by the specifications of the US transducer. Concerning traditional imaging, PAI employs the optical wavelengths in the near-infrared (NIR, 700–1500 nm), achieving penetration depths up to several centimeters. Versus conventional imaging modalities [15,16,17,18,19,20,21,22,23,24,25,26,27], PAI utilizes near-infrared (NIR, 700–1500 nm) light to achieve imaging depths reaching several centimeters [28]. One of the main benefits of PAI is its capacity to provide detailed structural, functional, and molecular information about biological tissues using non-ionizing energy sources. Through the integration of optical stimulation and ultrasound signal acquisition, this technique allows for real-time imaging capabilities [29]. PAI emerges as a robust imaging technique capable of offering detailed anatomical and physiological information, particularly valuable for the early diagnosis of certain tumors (Figure 2) [30,31,32,33].
In the realm of PAI, a key area of ongoing research lies in the design and refinement of nanostructured imaging enhancers that offer improved contrast-to-noise ratios, minimal toxicity, and efficient clearance from the bloodstream [28]. To overcome these challenges and improve the sensitivity and depth of imaging, PAI contrast enhancement agents are required [46,47]. The strength and high resolution of PAI have been demonstrated using extrinsic contrast agents such as fluorescent organic dyes [48,49], dye-loaded nanoparticles [50], carbon nanotubes [51,52,53], semiconductor quantum dots (QDs) [54,55,56], metal complexes [57,58,59], and fluorescent proteins [60,61]. These agents have been utilized for multiple types of preclinical PAI applications, including (1) subcellular imaging of organelles like mitochondria and melanosomes in vitro; (2) in vivo tracking of individual cells such as erythrocytes alongside malignant melanocytes; (3) visualizing vascular architectures, neovascularization, and the lymphatic system; (4) evaluating physiological parameters including hemoglobin oxygenation, blood flow dynamics, and metabolic activity within microvasculature; (5) assessing brain activity; and (6) monitoring drug delivery processes, therapeutic outcomes, targeted molecular interactions, and gene expression patterns [62,63,64,65,66,67].
In recent decades, upconversion nanoparticles (UCNPs) have emerged as innovative contrast agents across various biomedical imaging modalities [68,69,70]. These nanoparticles have the unique ability to alter low-energy light, typically in the near-infrared (NIR) region (wavelengths above 700 nm) [71], higher-energy emissions such as ultraviolet (UV, <400 nm), and visible light (400–700 nm) [72,73]. This upconversion (UC) process is based on nonlinear optical phenomena, where multiple photons are absorbed in succession via long-lived intermediate energy states, ultimately producing anti-Stokes emission—light with emission occurring at a shorter wavelength relative to the excitation beam, which was first recognized and formulated in the mid-1960s [74]. With advancements into nanomaterials that display upconversion behavior, their application in biomedical imaging has progressed rapidly. The specific advantages of UCNPs include: (1) the tunable NIR spectral absorbance in the “optical transmission window” of biological tissues (700–1000 nm); (2) no background autofluorescence unlike FL imaging probes; (3) applicability in multimodal imaging such as MRI, CT, PAI, and so on, and ability to enable photo-therapy [44,75,76,77,78,79]; (4) tissue-specific targeting through surface modification; (5) high in vitro and in vivo photostability; and (6) high biocompatibility. All of these merits render UCNPs more appealing than conventional fluorescence nanomaterials for bioimaging [80,81,82]. Therefore, UCNPs have been utilized as a modern iteration of a fluorescent nanoprobe for fluorescent optical imaging [83,84,85], holding great promise to overcome the innate shortcomings associated with dye molecules [86,87,88,89,90], fluorescent proteins, and quantum dots (QDs) that need to be excited with high-energy incident light, such as UV or visible light [91,92,93,94,95,96,97,98,99].
UCNPs as nanoprobe/CA in imaging have been broadly covered in multiple reviews in the past [59,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116]. However, UCNPs as contrast agents for PAI have not yet been dedicatedly summarized and analyzed. With their growing popularity as PAI contrast agents, there is a critical need to summarize the progress of this field in a critical review. Therefore, we systematically outline the recent progress on the use of UCNPs for PAI accounting for their fundamental properties that make UCNPs perfect for PAI agents. We also focus on current breakthroughs in UCNP-based PAI, which is intended to drive innovation in this domain and stimulate further curiosity from fellow researchers. At last, we highlight the remaining challenges and speculate for the prospective real-world use of UCNPs in biomedical imaging technologies.

2. Working Principle of PAI

2.1. Physical Underpinnings of PAI

PAI is a hybrid biomedical imaging technology that can achieve real-time and high-resolution in vivo tissue microanatomy [117]. PAI merges high-contrast optical excitation with strong temporal resolution (<100 µm) and high penetration depth of US imaging [118]. PAI typically uses NIR (750 to 1800 nm) and visible light (400 to 750 nm) to excite acoustic vibrations [119]. The photoacoustic effect involves three main stages: first, the target material absorbs incident photons, leading to a temperature rise; second, this thermal energy causes localized pressure elevation through thermoelastic expansion; and third, the resulting pressure disturbance propagates as acoustic waves via elastic interactions in the surrounding medium [120]. For effective generation of these acoustic signals, the thermal expansion must vary with time. This is commonly accomplished using a nanosecond pulsed laser (1–100 ns) or an intensity-modulated continuous-wave (CW) laser to illuminate the light-absorbing material [121,122]. Pulsed lasers are typically favored since they offer a higher signal-to-noise ratio than continuous-wave lasers at the same optical exposure and beam strength levels [123]. Additionally, the short pulse duration is typically less than the thermal and stress confinement times of the absorber, allowing thermal diffusion and volumetric expansion effects to be disregarded during excitation [124]. In biological tissues, photons interacting with cellular components predominantly undergo elastic scattering. When these scattered photons are absorbed by molecules, rapid thermoelastic expansion occurs, producing broadband ultrasonic waves in the surrounding tissue [125]. These acoustic signals are captured by piezoelectric transducer arrays, and the image is reconstructed by analyzing the amplitude and arrival time of the signals, enabling visualization of the original photoacoustic pressure distribution. A conceptual diagram of this mechanism is shown in Figure 3.
In general, upon absorption of photons by a tissue, the intrinsic UV–Vis chromophore is elevated to an excited state, followed by consequent chains of photon events in the visible and NIR region. In PA, due to NIR absorption capabilities of the chromophore, photon energy is transformed to heat energy via vibrational/collisional relaxation. This localized heating leads to increase in the local pressure. The initial local pressure rises or PA amplitude of waves ( P 0 ) can be calculated using Equation (1).
P 0 = Γ H
In this context, H represents the optical energy or heat absorbed per unit tissue volume. The resulting pressure rise is generally considered to be linearly proportional. Photoacoustic signal efficiency generation largely relies on the Grüneisen parameter Γ , a dimensionless thermodynamic factor that is sensitive to temperature and differs among various absorbing materials. The absorbed energy density is obtained using Equation (2).
H = η t h μ a F
Here, η t h denotes the thermal conversion efficiency, which reflects the proportion of nonradiative energy dissipation after laser excitation (essentially equal to one subtracted by the fluorescence quantum yield) [124,127]. Since Γ is temperature dependent, influenced by temperature, PAI can serve as a tool for thermal monitoring [128]. The optical absorption coefficient μ a (cm−1) reflects the concentration of the absorbing chromophore, and F denotes the local optical fluence (J·m−2). Collectively, these parameters define the total optical energy absorbed or deposited in the tissue [120,127,129]. Thus, combining Equations (1) and (2), PA amplitude of waves ( P 0 ) can be determined as represented in Equation (3)
P 0 = Γ η t h μ a F
The main source of contrast is provided by variations in optical properties, absorption, and scattering by the sample. Among all, absorption tends to be the governing factor for contrast in PA imaging. In tissues where the temperature is high and light is well absorbed, the PA images show an optimal signal-to-noise ratio (SNR). Photoacoustic imaging (PAI) enables visualization of both anatomical and functional features—such as hemoglobin oxygen saturation and metabolic activity—either with or in the absence of external contrast agents [130], thus making this technique suitable for anatomical imaging.

2.2. PAI Implementation: Tomography and Microscopy

According to the mechanism of formation of images, PAI systems are broadly categorized into two types: (i) raster-scanning-based photoacoustic microscopy (PAM) [67,131] and (ii) ultrasound array-based photoacoustic computed tomography (PACT) [132], as depicted in Figure 4. Photoacoustic imaging (PAI) systems can function in either reflection (backward) or transmission (forward) modes [133]. In the reflection mode, the laser source and US transducer are positioned on the identical side of the specimen and oriented perpendicular to it, similar to the arrangement in B-mode ultrasound imaging. In contrast, the transmission mode places the source and detector positioned on opposing sides of the sample, a setup typically used for thin tissue slices or ex vivo samples because of the limited light penetration in thicker tissues.
PAM inherits the features of PAI and is designed for in vivo imaging. PAM offers several benefits surpassing classical optical microscopy approaches, such as deeper imaging depth, highly sensitive structural and functional details, and not requiring optical sectioning to obtain the 3D image. PAM operates in either reflection or transmission mode and generally employs a focused single-element US transducer, paired with mechanical or optical scanning methods for image acquisition. In this configuration, the alignment of optical excitation and acoustic detection into dual focal points enhances overall sensitivity [134]. A single laser pulse produces a one-dimensional ultrasound signal along the axial direction of the laser beam. As the system scans laterally across the sample, these signals are collected point by point to reconstruct a three-dimensional image. The axial resolution is governed by the travel time of the acoustic wave, whereas the transverse resolution is influenced by the spacing between the optical and acoustic foci [134].
In the optical ballistic regime, the lateral resolution of photoacoustic microscopy (PAM) has been significantly improved through tight optical focusing, a technique known as optical-resolution photoacoustic microscopy (OR-PAM). This approach achieves high lateral resolution by relying on precise optical focus, enabling detailed in vivo imaging of individual cells and microvascular structures, like those found in murine auricular vessels. Furthermore, by harnessing nonlinear optical effects, the resolution of PAM has been pushed beyond the conventional optical diffraction limit, resulting in super-resolution imaging—termed photoacoustic nanoscopy (PAN)—which has made it possible to visualize structures as small as single mitochondria in vivo [135]. On the other hand, when acoustic focusing is applied within the optical diffusive regime, a method known as acoustic-resolution photoacoustic macroscopy (AR-PAMac) allows for complete anatomical imaging of small laboratory animals [136]. A notable benefit of this technique is its avoidance of complex reconstruction algorithms, thus minimizing image artifacts. Nonetheless, the reliance on raster scanning in PAM inherently limits the speed at which images can be acquired. To address this challenge, a new waterproof MEMS scanner (weighing 162 g with a diameter of 17 mm) has been created for OR-PAM [137,138]. These miniaturizations include mechanical scanning using a voice-coil motor (~15 Hz frame rate) [139], optical scanning using Galvo mirrors (~2 Hz frame rate) [140], and hybrid scanning through mechanical scanning on one axis and optical scanning on the other axis (~6 Hz frame rate) [141].
PACT is an evolving potent hybrid optical imaging modality utilizing both ultrasonic resolution and optical absorption contrast, while all the essential features are greatly scalable. PAT is extensively applied in research, with several commercial systems already available. In focused-scanning PAT, a focused ultrasound transducer physically scans across the target area. In contrast, array-based PACT typically utilizes transducer arrays arranged in linear, arc-shaped, or circular geometries [142]. A key benefit of this approach is its capability to deliver real-time imaging. Compared to PAM, PACT systems are more readily compatible with standard ultrasound platforms, making them more feasible for clinical integration and commercial development [143]. Typically, PACT offers spatial resolutions ranging from 0.1 to 1.0 mm, while maintaining the capability to image several centimeters deep into soft tissue. Additionally, a practical strength of this tomographic technique is that transducer arrays—such as linear probes—can be used in a handheld format, similar to conventional ultrasound devices.
Figure 4. Multiscale PA imaging. (A) Schematic of OR-PAM, where fast scanning is attained by a MEMS mirror [144]. (B) Schematic of AR-PAM [144]. (C) Schematic of PACT [145]. Abbreviations: PA, photoacoustic; OR-PAM, optical-resolution PAM; MEMS, microelectromechanical scanner; OAC, optical-acoustic combiner; AR-PAM, acoustic-resolution PAM; UT, ultrasonic transducer. Reprinted with permission from Refs. [144,145].
Figure 4. Multiscale PA imaging. (A) Schematic of OR-PAM, where fast scanning is attained by a MEMS mirror [144]. (B) Schematic of AR-PAM [144]. (C) Schematic of PACT [145]. Abbreviations: PA, photoacoustic; OR-PAM, optical-resolution PAM; MEMS, microelectromechanical scanner; OAC, optical-acoustic combiner; AR-PAM, acoustic-resolution PAM; UT, ultrasonic transducer. Reprinted with permission from Refs. [144,145].
Nanomaterials 15 01074 g004

3. Contrast Agent for PAI

3.1. Endogenous PAI—Intrinsic Chromophores

Endogenous photoacoustic imaging (PAI) captures both anatomical and functional data by detecting light absorption from intrinsic tissue components such as hemoglobin, lipids, water, and proteins (Figure 5A). Hemoglobin, responsible for oxygen transport, experiences structural and electronic alterations upon binding oxygen (Figure 5B), leading to distinct spectral shifts. These variations allow for the quantification of total hemoglobin levels and oxygen saturation, which are critical for analyzing tumor vascularization (Figure 5C) [146,147]. PAI can also be utilized to map lipid distribution in vivo [148], utilizing absorption peaks at 930 nm and 1210 nm, attributed to C-H bond overtones, while water content is mapped via its peak absorption near 975 nm [149].

3.2. Exogenous PAI—Key Factors Governing Contrast Agent Design

For effective molecular imaging, a contrast agent should consist of two key components: a signaling molecule responsible for generating the imaging signal, and a targeting unit that guides the agent to specific biological structures or processes for precise localization [150,151].
Nanoparticles used in photoacoustic imaging typically share several key features: (1) strong optical absorption in the near-infrared (NIR) window (usually 650–950 nm), which ensures deeper tissue penetration and efficient photoacoustic signal generation [152]; (2) high photothermal conversion efficiency to generate detectable acoustic waves upon pulsed laser excitation [153]; (3) good biocompatibility and stability under physiological conditions [154]; and (4) appropriate size and surface chemistry for prolonged circulation and targeted delivery [155]. These shared properties enable various types of nanoparticles—such as gold nanostructures, semiconducting polymers, carbon-based materials, and organic dyes—to serve as effective photoacoustic contrast agents.
(i)
Photophysical properties
An optimal signaling molecule for photoacoustic imaging should possess several key features. First, a high molar extinction coefficient is essential for efficient light absorption and strong photoacoustic (PA) signal generation. While a distinct and narrow absorption peak can aid signal isolation under controlled conditions, its effectiveness in vivo is limited due to the broad absorption spectra of endogenous chromophores (e.g., hemoglobin, melanin) and tissue scattering, which compromise spectral specificity and introduce signal cross-talk. As such, relying solely on spectral unmixing is insufficient for accurate in vivo imaging. To overcome these limitations, alternative detection strategies have emerged.
One such approach is the use of radiometric photoacoustic imaging, in which dual-wavelength or dual-emission UCNP systems (e.g., Tm3+/Er3+- or Yb3+/Nd3+-doped nanocrystals) are employed to provide internal calibration [156,157,158,159], thereby reducing variability from probe concentration and environmental effects. Time-gated or lifetime-based detection methods also show promise, leveraging the long luminescence lifetimes of UCNPs to distinguish signal from tissue background and improve temporal resolution [160,161,162].
Furthermore, photo-switchable probes, such as UCNPs coupled with light-activated absorbers (e.g., spiropyrans), can create an on/off contrast mechanism, enhancing signal discrimination [157,163]. Similarly, stimuli-responsive systems—including thermoresponsive or pH-sensitive UCNP composites—enable signal activation only under specific physiological conditions, increasing target specificity and reducing background interference [164,165].
Ideally, the absorption of these constructs should fall within the near-infrared (NIR) window (620–950 nm) to maximize tissue penetration (Figure 5A). In addition, strong photostability is necessary to preserve signal over time, while a low fluorescence quantum yield is preferred to favor nonradiative decay and promote heat generation. Ultimately, the contrast agent must efficiently convert absorbed energy into localized thermal expansion to produce reliable and high-contrast acoustic signals, suitable for deep-tissue, high-resolution biomedical imaging.
Figure 5. Photoacoustic imaging of intrinsic chromophores. (A) Absorption profiles of major endogenous molecules: hemoglobin (blue, 150 g/L), water (black, ~80% tissue volume) [166]; lipid, red line [167,168] (20% by volume in tissue); and protein, green line. Data from http://omlc.ogi.edu/spectra (access on 1 May 2025). (B) The heme moiety, an iron-bound aromatic porphyrin, undergoes structural changes upon oxygenation, leading to distinct spectral shifts. (C) PAI at 640 nm visualizes tumor vasculature before and 24 h after administration of the vascular disrupting agent OXi4503. Repeated vascular features in the skin are marked with green arrows. Scale bars: 1 mm. Figure 5 reprinted with permission from Ref. [169].
Figure 5. Photoacoustic imaging of intrinsic chromophores. (A) Absorption profiles of major endogenous molecules: hemoglobin (blue, 150 g/L), water (black, ~80% tissue volume) [166]; lipid, red line [167,168] (20% by volume in tissue); and protein, green line. Data from http://omlc.ogi.edu/spectra (access on 1 May 2025). (B) The heme moiety, an iron-bound aromatic porphyrin, undergoes structural changes upon oxygenation, leading to distinct spectral shifts. (C) PAI at 640 nm visualizes tumor vasculature before and 24 h after administration of the vascular disrupting agent OXi4503. Repeated vascular features in the skin are marked with green arrows. Scale bars: 1 mm. Figure 5 reprinted with permission from Ref. [169].
Nanomaterials 15 01074 g005
(ii)
Biological properties
Targeted contrast agents need to overcome circulatory and cellular barriers to effectively reach their specific target sites. Upon entering the bloodstream, these agents encounter various biomolecules, which may adsorb onto their surfaces [170,171,172]. This surface adsorption can influence immune recognition and interfere with target specificity, thereby complicating data interpretation and requiring thorough characterization [171,173]. To effectively extravasate from blood vessels into surrounding tissues, contrast agents generally need to be smaller than 10 nm. However, in oncology-focused uses, nanoparticles as large as 100 nm have been shown to traverse the abnormally permeable vasculature typically found in tumors [174]. After reaching the tissue interstitium, the agent may cross cellular barriers either via active targeting—such as binding to cell surface receptors [175], transporters [176], or metabolic enzymes [177]—or through passive mechanisms like diffusion or endocytosis.

3.3. UCNPs as Contrast Agent for PAI

(i)
Synthesis and Functionalization of UCNPs
Long before UCNPs emerged, upconversion (UC) phenomena had already been demonstrated in various bulk materials using diverse host and lanthanide (Ln3+) ion combinations. However, only a limited number of these combinations are viable in colloidal nanocrystals due to synthesis challenges and reduced emission efficiency at the nanoscale. Fluoride-based lattices like NaYF4 are commonly employed as UCNP hosts due to their low phonon energy (~350 cm−1), high optical transparency, and ability to form well-crystallized structures at relatively low synthesis temperatures. Various synthetic methods—including thermal decomposition, hydrothermal synthesis, sol-gel processing, coprecipitation, and ionic liquid-assisted techniques—have been explored to produce high-quality lanthanide-doped NaYF4 UCNPs with precise control over composition, crystallinity, and morphology (Table 1). Among these methods, thermolysis [178] and hydrothermal synthesis [179] are most frequently employed, as they enable fine control over the phase, morphology, size, and stoichiometry of both core and core–shell UCNPs.
To obtain UCNPs with high crystallinity and uniform morphology, the above-mentioned methods are typically conducted in high-boiling-point non-aqueous solvents alongside one or two long-chain ligands. Consequently, the resulting UCNPs are usually coated with hydrophobic ligands (e.g., oleic acid), necessitating subsequent surface modification to render them hydrophilic and enable functionalization (Figure 6) [187]. These approaches can be broadly classified into three main categories: hydrophilic modification, bioconjugation, and functionalization with hybrid materials. Common approaches include the use of (1) acidic ligands [188], (2) polymers [189], or (3) chelating agents [190], all of which typically contain hydrophilic functional groups such as hydroxyls, amines, or carboxylic acids. A comprehensive overview of these modification techniques is provided in the recent review by Chen et al. [191].
(ii)
Dopant/Host Selection Criteria
The distinctive optical behavior of UCNPs is largely determined by their material composition. They typically consist of an inorganic host lattice embedded with light-absorbing sensitizer ions and light-emitting activator ions (Figure 7) [192,193,194]. Inorganic compounds based on rare earth elements—such as NaYF4, NaErF4, NaGdF4, NaLuGdF4, Y2O3, Y2O2S, and GdOCl—serve as excellent host materials due to their low phonon energy, strong chemical and thermal stability, and high optical transparency [191]. Choosing suitable host materials is crucial for synthesizing lanthanide-doped (Ln-doped) nanocrystals with targeted nanoscale features and desirable optical properties, including high upconversion efficiency and tunable emission. An ideal host should be readily synthesized with small, uniform particle size and a narrow size distribution. Additionally, low lattice phonon energy is essential, as it reduces nonradiative losses and enhances radiative emission [101]. Rare earth fluorides, particularly binary REF3 and ternary AREF4 compounds (where RE = rare earth element, A = alkali metal), are the most commonly employed host materials for UCNP fabrication [195,196,197,198,199,200,201]. Among them, hexagonal-phase NaYF4 nanocrystals co-doped with Yb/Er or Yb/Tm are recognized as the most efficient upconversion systems.
Activator ions serve as the primary luminescent centers in UCNPs. Among various options, trivalent lanthanide ions (Ln3+), particularly Er3+, Tm3+, and Ho3+, are most frequently used due to their stepwise energy level structures, which facilitate multiphoton absorption and efficient upconversion [74,101,202,203]. These ions also possess wide energy level gaps, which help suppress nonradiative transitions between excited states, thereby enhancing upconversion efficiency. Moreover, Yb3+, which shows high absorption (~980 nm, 2F7/22F5/2), and other rare earth elements (mainly Er3+, Tm3+, Ho3+, Tb3+, Eu3+, Dy3+, Sm3+, and Gd3+) have been successfully employed as efficient sensitizers and activators in the fabrication of UCNPs, respectively [178,191,204]. The efficiency of energy transfer in UCNPs can be enhanced by carefully tuning the inorganic host matrix along with the concentrations of sensitizer and activator ions, particularly under excitation wavelengths of 808 nm and 980 nm [205]. Importantly, the emission colors of UCNPs can be finely tuned by adjusting the concentrations of sensitizer and activator ions [206], as well as by introducing additional metal ions [111,178,207], with specific details summarized in Table 2.
Importantly, UCNPs exhibit colorful emissions, which make them promising materials for bioapplications [178]. Multiple research teams have investigated UCNPs doped with alternative lanthanide ions such as Gd3+, Yb3+, Ho3+, and Sm3+ as contrast agents for imaging modalities like MRI, CT, and PAI, aimed at enhancing the visualization of specific biological structures [216,217,218]. In recent years, the application of NaYF4-based UCNPs in biomedical applications has expanded rapidly, presenting promising opportunities for the advancement of novel non-invasive bioimaging techniques. Nevertheless, the search for alternative host materials for UCNPs continues in pursuit of improved performance and tailored nanoscale properties. A key limitation of NaYF4-based UCNPs in biological applications is the difficulty in synthesizing ultra-small nanocrystals (below 10 nm) that still exhibit strong upconversion luminescence [191]. To overcome this problem, other fluorides such as NaxScF3+x [219], NaYbF4 [220], KGdF4 [221], CaF2 [222,223], NaLuF4 [224,225,226], and BaLuF5 [227] have recently gained popularity as host materials for producing smaller UCNPs with enhanced brightness.
(iii)
Mechanism of Upconversion—Optical Imaging Agent
The rich energy level structures of lanthanide ions provide upconversion nanoparticles with extensive possibilities for efficient energy transfer processes, which have been thoroughly summarized and analyzed in various review articles [74,228,229]. These upconversion mechanisms can be broadly categorized into five types: (a) excited-state absorption (ESA), (b) energy transfer upconversion (ETU), (c) cooperative sensitization upconversion (CSU), (d) cross-relaxation (CR), and (e) photon avalanche (PAv) [221,230]. It is important to note that the term “upconversion” is occasionally used to refer to the overall result—namely, the conversion of long-wavelength to short-wavelength emission—and at other times to denote a specific underlying mechanism responsible for this process.
All five mechanisms operate through the stepwise absorption of two or more photons by long-lived, metastable energy states. This process results in the population of a higher excited state, from which upconversion emission is generated. Specifically, in ESA, a single ion absorbs multiple photons in succession, with each photon promoting the ion to a higher energy level until it reaches the emissive state (Figure 8A).
ETU resembles ESA in that both involve the sequential absorption of two photons to reach a higher-energy metastable state. However, the key distinction lies in the mechanism of excitation: while ESA occurs within a single ion, ETU involves energy transfer between neighboring ions. In ETU, two adjacent ions each absorb a pump photon of the same energy, with one ion subsequently transferring its energy to the other, promoting it to the metastable level E1 (Figure 8B).
CSU, illustrated in Figure 8C, involves the interaction of three ion centers—typically with ion 1 and ion 3 being of the same type. Compared to ESA and ETU, CSU is significantly less efficient, often by several orders of magnitude, due to its reliance on quasi-virtual pair states that require higher-order quantum mechanical treatment through perturbation theory. Nevertheless, the inherently low efficiency of CSU necessitates highly localized excitation, which could be advantageous for achieving super-resolution imaging not attainable through other upconversion mechanisms.
CR, shown in Figure 8D, is a type of energy transfer arising from interactions between ions, where ion 1 transfers part of its excitation energy to ion 2 via a process involving transitions such as E2 (ion 1) + G (ion 1) + E1 (ion 2). As a fundamental ion–ion interaction mechanism, the efficiency of CR is strongly influenced by dopant concentration. While CR often leads to the well-known phenomenon of concentration quenching, where excessive dopant levels reduce emission efficiency; it can also be strategically utilized to modify emission color or to initiate efficient PAv processes.
PAv, depicted in Figure 8E, is characterized by a unique excitation mechanism that requires the pump power to exceed a specific threshold. The process begins with weak, non-resonant ground-state absorption (GSA) that populates level E1, followed by resonant excited-state absorption (ESA) that promotes electrons to the higher-energy emissive level E2. PAv is readily distinguishable by its threshold-dependent behavior, delayed luminescence onset (often taking seconds), and a steep, nonlinear increase in photoluminescence intensity near the pump threshold.
  • (iv) UCNPs works as contrast agent for PAI
Since Alexander Graham Bell first observed the photoacoustic effect [13], significant research efforts have focused on creating efficient photoacoustic contrast agents. These include both inorganic materials—such as gold-based nanostructures, MXene derivatives, carbon-based nanomaterials, and silicon nanoparticles—and organic agents like small organic molecules, polymers, and DNA-based constructs [231,232,233,234,235]. These contrast agents have significantly broadened the scope of PAI by enhancing contrast and enabling deeper tissue imaging. Among them, upconversion nanoparticles (UCNPs) have drawn significant interest due to their ability to absorb near-infrared (NIR) light and emit visible or higher-energy photons via multi-photon upconversion processes [236]. This anti-Stokes emission reduces background interference, minimizes sample autofluorescence, and lowers the risk of tissue overheating, which is a critical advantage for in vivo imaging applications [237,238,239,240,241]. If luminescence is suppressed, either through surface quenching or hybrid material coupling, this energy is instead dissipated as localized heat, leading to transient thermoelastic expansion of the surrounding medium. This rapid expansion generates an ultrasonic pressure wave, which can be detected by photoacoustic transducers and used to reconstruct high-resolution images. These dual-functional properties position UCNPs as ideal nanoplatforms for both fluorescence and photoacoustic modalities, and as versatile tools for biological sensing and in vivo diagnostics. Representative lanthanide-based nanoprobes and their applications in PAI are summarized in Table 3.
Nevertheless, to fully realize their clinical potential, the in vivo biocompatibility and degradation behavior of UCNPs must be critically addressed.
Although UCNPs—typically composed of rare-earth doped fluoride matrices (e.g., NaYF4:Yb,Er/Tm)—are generally considered chemically stable and photophysically robust, their in vivo biodegradation may release lanthanide ions or fluoride species. These degradation products could potentially induce cytotoxicity, oxidative stress, or organ accumulation, especially upon prolonged exposure or high-dose administration. The cytotoxicity also leads to the rare use of bare UCNPs in medical contrast agents; thus, the application of a suitable surface modification is particularly important when it comes to the health concerns of UCNPs in vivo.
Recent research thus focuses on engineering UCNPs with biocompatible coatings (e.g., silica, PEG, phospholipids) or designing biodegradable core–shell structures to improve clearance and reduce systemic toxicity. Surface modifications not only largely improve the stability of UCNPs, but also enhance the performance of UCNPs as in vivo contrast agents by enhancing the hydrophilicity, biocompatibility, colloidal stability, and even optical intensity. It is well known that the cytotoxicity of UCNPs is a result of toxic ion (mainly fluoride ions and lanthanide ions) leakage during the degradation process. For engineered UCNPs aimed at reducing the potential degradation and increasing safety, the coating acted as a protective layer against the hydroxide ion or phosphate ion in the surrounding environment, which is corrosive to UCNPs and the cause of cytotoxic ion leakage. According to the research of Bastos et al., thick silicon-coated UCNPs have significantly lower fluoride ion leakage (4.42 µM) compare to bare UCNPs (297.37 µM) after 48 h in water [257]. The research of Saleh et al. also reported that the fluoride ion and yttrium ion leakage from thick silicon-coated UCNPs remained at a very low level (≈0 µg·mL−1) after over 70 h, while the fluoride ion leakage from bare UCNPs already exceeded 5000 µg·mL−1 in water, and the yttrium ion leakage exceeded 4000 µg·mL−1 at the same time [258]. As such, the safety and stability increasing effects of UCNP surface modifications were widely proved by experimental data.
Nonetheless, UCNPs with an appropriate surface modification could be excreted via the hepatobiliary excretion pathway in a shorter time period without overt tissue toxicity. This effect will be increased if coordinate with a small particle size, since smaller particles can avoid participation of intracellular catabolism and are rapidly cleared by urine. The research of Liu et al. reported that a hydrophilic polyethylene glycol (PEG)-coated 5.1 nm NaGdF4 UCNP sample exhibited an elimination half-life of 1.4 days after being injected into mice. The excreted particles found in the feces were confirmed by TEM image analysis to be no different in size, size distribution, and shape compared to the original particles. This suggested that PEG-coated UCNPs within living mice were not transformed [259,260]. The safety of such UCNPs can be concluded to be reliable, since the degradation time was obviously longer than the excretion time, and most particles were cleared out of the system before they had enough time to release toxic ions. Additionally, the fastest complete excretion time of NaYF4:Yb,Er@SiO2-PEI UCNPs (50 nm) is 7 days, reported by Zhang et al. [261], which was a noteworthy result in view of the particle size.

4. Photoacoustic Imaging—A Perfect Imaging Modality of UCNPs

Upconversion nanoprobes possess notable features such as excellent photostability, favorable biocompatibility, distinct near-infrared anti-Stokes emission, and improved tissue penetration, making them valuable tools for visualizing biological activities and detecting disease biomarkers [262,263].

4.1. Coated UCNPs for PAI

Lanthanide-doped upconversion nanoparticles (UCNPs) have attracted growing interest as next-generation contrast agents for photoacoustic imaging due to their chemical stability, engineerable photophysical properties, and biocompatibility. Their anti-Stokes emission, wherein low-energy near-infrared (NIR) light is absorbed and converted into higher-energy emission, offers distinct advantages for in vivo applications. This includes minimized background interference—as biological tissues lack strong anti-Stokes signals—and reduced tissue overheating, since NIR light penetrates deeply with lower absorption and scattering, lowering the risk of photothermal damage [264,265].
Although luminescence is not the primary readout in PA imaging, the balance between radiative and nonradiative relaxation pathways is central to UCNP performance. Efficient luminescence implies that the particles’ energy levels can be finely tuned. However, quenching luminescence intentionally through material design (e.g., via surface coatings or dopant optimization) allows more energy to dissipate as heat, which is crucial for generating strong PA signals.
In this context, dopant concentration, shell passivation, and hybrid construct formation are key strategies for modulating energy dissipation and optimizing PA performance.
Doping concentration significantly influences nonradiative relaxation pathways, which are crucial for thermal expansion and, consequently, PA signal generation. Sheng et al. showed that rare-earth-element doped micrometer-sized particles (e.g., NaYF4:Yb3+, Er3+) modified with polyacrylic acid yielded enhanced PA signals, particularly in anisotropic rod-like structures. These enhancements were attributed to optimized doping levels that promote upconversion luminescence quenching and efficient heat generation [242].
Shell passivation, such as inert NaYF4 coatings, plays a crucial role in reducing surface defects and suppressing surface quenching. This not only improves upconversion efficiency, but also increases the proportion of absorbed photon energy converted into heat rather than being lost via radiative emission. As shown in comparative studies, increasing shell thickness enhanced the PA signal more effectively than increasing particle size alone, confirming the pivotal role of surface engineering.
Based on these insights, Zhao and Yu et al. first reported that NaYF4:Yb3+, Er3+ upconversion nanoparticles coated with α-cyclodextrin (UCNP/α-CD) could significantly enhance photoacoustic signals through luminescence quenching-induced nonradiative relaxation. This enhancement arises from the intrinsic heat-generating capability of UCNPs, coupled with improved thermal conductivity due to a phase transition occurring in aqueous environments. Under 980 nm laser excitation, this mechanism enabled strong PA signal output in water, allowing successful in vivo imaging of mouse kidneys (Figure 9A) [254]. These findings support the potential of UCNP/α-CD systems as effective contrast agents for diagnostic imaging applications.
As limited penetration depth is still the main disadvantage for bioimaging diagnostics, hybrid constructs that integrate UCNPs with optically absorbing materials have demonstrated dramatic improvements in PA performance. Hou et al. describe the synthesis of H-TiO2-decorated Nd3+-sensitized UCNPs using a modified Nd: YAG laser method (Nd: UCNPs@H-TiO2), under 808 nm excitation [248]. H-TiO2 not only serves as a visible-light-driven photosensitizer, but can be utilized to generate a PA signal with a high US spatial resolution determined by thermoelastic expansion. This result was confirmed experimentally for mice in comparison to the untreated tumor. Such excellent imaging performance demonstrates that Nd: UCNPs@H-TiO2 serves as an obvious contrast agent in optical absorption between exogenous and endogenous environments to overcome the limitation of fundamental penetration depth (Figure 9B,C) [248]. Meanwhile, the particle size, morphology, and coating are all being explored to optimize UCNPs for PAI applications. Lv et al. designed a novel contrast agent that can work well in their newly designed nanoplatform for PAI. The material consists of a UCNP core–shell structure with an additional mesoporous silica loaded with indocyanine green (ICG) molecule (UCNP@mSiO2-ICG). Compared with pure ICG, sealing the ICG within a mesoporous silica exhibited significantly elevated PA signals due to ICG’s strong NIR absorption and improved stability from silica encapsulation. These constructs not only boosted imaging depth (up to 1.5 cm in vivo), but also mitigated dye leakage and photobleaching (Figure 9D) [246]. Meanwhile, Sheng et al. [253] reported that rare-earth-element-doped micrometer-sized particles can be used for PAI optimized contrast visualization. The authors showed that the morphology, particle size, and proper coating have a major impact on influencing the PA signal intensity, especially emphasizing that coating an inert NaYF4 shell can enhance luminescent emission more significantly than increasing the particle size, indicating that coating an inert shell can effectively reduce the surface quenching effects (Figure 9F,G).
Together, these advances underscore the critical relationship between nanoparticle design and PA imaging performance, offering pathways to optimize contrast agent efficacy for deep-tissue diagnostics and real-time biomedical applications.

4.2. UCNP Nanocomposites for PAI

Whilst the findings discussed above highlight the potential of UCNPs as PAI contrast agents, their effectiveness in deep-tissue diagnostics remains limited due to relatively low photoacoustic conversion efficiency and interference from non-specific background signals [266]. To address these limitations, Zhuang et al. developed ONOO-responsive UCY7 nanoprobes, constructed by coordinating heptamethine cyanine dyes with UCNPs. These nanoprobes demonstrated excellent sensitivity, selectivity, photostability, and biosafety for both in vitro and in vivo detection of ONOO. By integrating the dye with UCNPs, they effectively mitigated the photobleaching issues of cyanine dyes and enhanced tissue penetration by converting visible 660 nm light from P-cy7 into 980 nm NIR light absorbed by UCY7. This strategy led to superior in vitro PAI performance and enabled real-time imaging of liver injury in living subjects [245]. In addition, Xing and Gao et al. developed UCNP–cyanine dye nanocomposites to enable PAI of dynamic redox changes during disease progression (Figure 10) [267].
Du et al. developed an innovative multifunctional nanoplatform by in situ synthesizing ultrasmall Ag2Se nanodots uniformly anchored onto UCNP surfaces via an intermediate CS shell layer (Figure 11). The PA signal gradually increased within 2 h of injection of UCNP@CS@Ag2Se nanocomposites, suggesting that a significant amount of nanocomposites preferentially accumulate at the tumor site. Quantitative analysis revealed that the PA signal within the tumor increased five-fold two hours after injection, which is probably owed to the enhanced permeability and retention (EPR) effect occurring during systemic circulation [250].

4.3. Outcomes of Combining UCNPs for PAI with Other Imaging Modalities

Imaging technologies have greatly advanced our ability to observe biochemical processes and understand the relationship between anatomical structure and biological role at both cellular and anatomical levels. Nonetheless, each imaging modality comes with inherent drawbacks—such as limited sensitivity or resolution—which can hinder the acquisition of precise and dependable data from target tissues. A practical solution to these challenges is the integration of multiple imaging techniques to boost spatial resolution and signal fidelity performance. In particular, combining luminescence imaging with photoacoustic imaging (PAI) offers a cost-effective, intraoperative platform that is well-suited for surgical navigation and enables high-resolution visualization of deep-tissue structures. For instance, the material N-hydroxysuccinimide-labeled indocyanine green (ICG) immobilizing on the surface of pre-fabricated layered UCNPs enabled simultaneous UCL, PAI (Figure 12), and magnetic resonance imaging (MRI) in vivo. Attaching ICG to the UCNP surface significantly boosts UCL intensity under 800 nm excitation, owing to efficient energy transfer. Meanwhile, this unique design increases the imaging depth (from 4.8 mm up to 10 mm) of PAI, and the signal can be increased up to 1.8-fold 6 h post-injection. This effect primarily results from the EPR phenomenon in leaky tumor vasculature combined with cellular uptake via endocytosis. Following this, combination of modalities can provide much more detailed information for clinical diagnosis [243]. In addition, both upconversion luminescence (UCL) and photoacoustic (PA) tumor imaging exhibited excellent biocompatibility and high specificity in sensing, offering promising prospects for deep-tissue multimodal imaging and integrated phototherapeutic applications.
He et al. introduced a method utilizing UCNPs coated with a photo-switchable azobenzene-containing polymer (PAA-Azo), which enhanced PA signals up to six-fold compared to unmodified UCNPs while maintaining stable NIR-II emission. Nevertheless, the relatively long acquisition time (approximately 5 min) limits the real-time applicability of PAI. Integrating PA imaging with NIR-II imaging within a single platform offers a promising approach to enhance diagnostic precision and efficiency (Figure 13) [249]. Additionally, Sheng et al. demonstrated that PAA-coated REDP nanorods show significant potential as dual-modality agents for luminescence and PAI [242].
Although multimodal optical imaging techniques have facilitated the visualization of significant pathological changes in living organisms, most still struggle to dynamically monitor bioindicator responses. This limitation largely stems from challenges in designing stimuli-responsive, multimodal imaging probes. To address this, Xing et al. developed a technology combining deep-tissue-penetrating multispectral optoacoustic tomography (MSOT) with upconversion luminescence (UCL) imaging by integrating reactive oxygen species (ROS)-sensitive chromophores with UCNPs (Figure 14) [244]. The interaction between pathological ROS such as H2O2 and the nanoprobes induces a shift in absorption, detectable through ratiometric MSOT signals. This system offers a self-calibrated cross-referencing approach, maximizing both sensitivity and specificity. Utilizing this combined PA and UCL imaging method, the authors successfully enabled precise monitoring of the intricate link between redox imbalance and hepatic disorders, proving valuable for in vivo evaluation of hepatoprotective therapies. (Figure 14).
In a separate study, Zhang et al. presented a straightforward strategy to develop an upconversion-based nanoagent (UCNPs-DI) by attaching a diketopyrrolopyrrole (DPP) polymer dye to the nanoparticle surface. This dye exhibited strong absorption overlapping with the visible emission of the UCNPs, while also incorporating indocyanine green (ICG) as a photosensitizer responsive to near-infrared (NIR) light. Under pulsed-wave (PW) 980 nm laser excitation, ICG remains inactive, while the visible light emitted by UCNPs is absorbed by DPP to produce strong photoacoustic signals without inducing photothermal or photodynamic effects. In contrast, UCNPs, DPP, free ICG, and UCNPs-I generated only weak signals. These findings demonstrate that UCNP-DI enables enhanced, long-lasting, real-time PA imaging with minimal phototoxicity, facilitating effective PA-guided phototherapy. Additionally, altering the 980 nm laser from PW to continuous-wave (CW) mode significantly boosts its photodynamic therapy effect [268] (Figure 15).
Sun et al. developed a new system using UCNPs in connection with a mixture of thiolated oligonucleotides and heterobifunctional polymer molecules grafted onto nanorod dimers. The synthesized nanorod dimer-UCNP-Ce6 exhibited a notably high efficiency in UCL-PA multimodal imaging systems. Figure 13 confirmed that PAI was effectively attributed to the material’s strong NIR absorbance. Photoacoustic imaging conducted 24 h after contrast agent administration revealed a markedly enhanced signal across the entire tumor area. This enabled clear visualization of tumor depth and structural details, highlighting the potential of DNA-guided multifunctional nanoagent systems for future clinical applications in cancer diagnosis and treatment [252] (Figure 16).
Currently, many contrast agents have been developed to support not only dual-, but also tri- and tetra-modal medical imaging. For instance, He et al. synthesized UCNPs@MS-Au25-PEG materials for PA imaging, demonstrating clear photoacoustic responses in an experiment involving concentration gradients under 808 nm laser excitation (Figure 17A,B). This single NIR light-triggered tri-modal imaging capability enables in vivo bioimaging and imaging-guided malignancy therapy. The presence of nanoparticles can be further validated by PAI, which offers high-resolution, non-invasive visualization of tissue structures, making it well-suited for diagnostics. Notably, the PA signal intensity at neoplasm sites increases over time following injection, showcasing excellent PAI performance. Moreover, PAI provides detailed structural information important for cancer detection and therapy [247]. Additionally, He et al. [251] successfully synthesized NaGdF4:0.3%Nd@-, NaGdF4@-, NaGdF4:10%Yb/1%Er@-, NaGdF4:10%Yb@-, and NaGdF4:10%Yb-(LDNPs-5)-Au25-PEG multi-core–shell structures. This integrated platform enables both energy-blocking and synergistic enhancement through the combination of photoacoustic imaging (PAI), fluorescence imaging (FI), and photothermal imaging (PTI), demonstrating the effectiveness of multimodal imaging for visualizing drug delivery processes (Figure 17C,D).

5. Outlook and Future Developments

Upconversion nanoparticles (UCNPs) have emerged as a promising class of contrast agents for photoacoustic imaging (PAI), offering distinct advantages over conventional dyes and metallic nanoparticles. Their unique optical properties, including narrow emission bands, near-infrared (NIR) excitation capabilities, and deep-tissue penetration, minimize background interference while enabling high-resolution imaging. Additionally, UCNPs possess inherent luminescence that facilitates dual-modal imaging when combined with PAI, allowing cross-validation of signals. The ease of surface functionalization further enhances their utility, as targeting ligands such as antibodies or peptides can be conjugated for precise subcellular imaging and theranostic applications. These features make UCNPs particularly valuable in preclinical research, where they have demonstrated excellent performance in tumor imaging, vascular mapping, and drug delivery monitoring in animal models [269,270].
Despite these advantages, significant challenges must be addressed before UCNPs can transition from laboratory research to widespread clinical use in diagnostics. One major limitation is biocompatibility, as certain UCNP compositions may trigger inflammatory responses or exhibit long-term accumulation in organs such as the liver and spleen. Furthermore, the relatively high doses required to generate detectable photoacoustic signals raise concerns about potential toxicity in human applications. Targeting efficiency in vivo remains another hurdle, with studies showing less than 5% tumor uptake in some cases due to nonspecific binding and clearance by the reticuloendothelial system. These biological challenges are compounded by technical and regulatory barriers, including the lack of standardized protocols for toxicity assessment and the high costs associated with scalable synthesis and purification compared to clinically approved contrast agents like indocyanine green (ICG).
To bridge this gap between preclinical success and clinical adoption, future research should prioritize three key areas. First, optimizing biocompatibility through innovative surface coatings—such as biodegradable silica or phospholipid layers—could mitigate immune responses while maintaining optical performance. Second, enhancing functional performance by engineering UCNPs with higher absorption coefficients in the NIR-II window (1000–1700 nm) would improve signal sensitivity at lower doses. Active targeting strategies, such as pH-responsive ligands or tumor microenvironment-specific probes, could further boost specificity. Finally, concerted efforts toward clinical integration are essential. This includes validating UCNP-based PAI in large-animal models to assess scalability and combining UCNP-enhanced molecular imaging with existing clinical ultrasound systems to leverage established infrastructure.
While UCNPs may not replace conventional imaging modalities like MRI or CT, their unique capabilities position them as powerful complementary tools. By addressing current limitations in biocompatibility, targeting efficiency, and regulatory readiness, UCNPs have the potential to transition from research curiosities to clinically viable contrast agents. As photoacoustic imaging technology advances and our understanding of nanomaterial design deepens, UCNPs could play a pivotal role in unlocking new diagnostic and theranostic applications—ultimately enhancing the precision and functionality of medical imaging.

Author Contributions

Y.Z. and Z.L. contribute equally. Y.Z. and Z.L. wrote the manuscript. Y.H., Z.D. and J.P. revised the review. Y.H. and J.P. conceived the idea and supervised the project. All authors have read and agreed to the published version of the manuscript.

Funding

Y.H. acknowledges the support from the National Natural Science Foundation of China (22308127). The authors would like to sincerely thank Xiaoxue Xu, from the University of Technology Sydney (UTS), and Tushar Kumeria, from the University of New South Wales (UNSW), for their valuable suggestions and assistance in reviewing and correcting this manuscript. Their insightful feedback greatly contributed to improving the clarity and quality of this work.

Conflicts of Interest

The authors declare no competing financial interests.

References

  1. Miao, Q.; Xie, C.; Zhen, X.; Lyu, Y.; Duan, H.; Liu, X.; Jokerst, J.V.; Pu, K. Molecular afterglow imaging with bright, biodegradable polymer nanoparticles. Nat. Biotechnol. 2017, 35, 1102–1110. [Google Scholar] [CrossRef] [PubMed]
  2. Wu, L.; Ishigaki, Y.; Hu, Y.; Sugimoto, K.; Zeng, W.; Harimoto, T.; Sun, Y.; He, J.; Suzuki, T.; Jiang, X.; et al. H2S-activatable near-infrared afterglow luminescent probes for sensitive molecular imaging in vivo. Nat. Commun. 2020, 11, 446. [Google Scholar] [CrossRef]
  3. Zhou, H.; Yang, H.; Tang, L.; Wang, Y.; Li, Y.; Liu, N.; Zeng, X.; Yan, Y.; Wu, J.; Chen, S.; et al. Mn-Loaded apolactoferrin dots for in vivo MRI and NIR-II cancer imaging. J. Mater. Chem. C 2019, 7, 9448–9454. [Google Scholar] [CrossRef]
  4. Tu, L.; Xu, Y.; Ouyang, Q.; Li, X.; Sun, Y. Recent advances on small-molecule fluorophores with emission beyond 1000 nm for better molecular imaging in vivo. Chin. Chem. Lett. 2019, 30, 1731–1737. [Google Scholar] [CrossRef]
  5. Guo, Z.; Wang, M.; Agyekum, A.A.; Wu, J.; Chen, Q.; Zuo, M.; El-Seedi, H.R.; Tao, F.; Shi, J.; Ouyang, Q.; et al. Quantitative detection of apple watercore and soluble solids content by near infrared transmittance spectroscopy. J. Food Eng. 2020, 279, 109955. [Google Scholar] [CrossRef]
  6. Li, C.; Niu, Z.; Zuo, M.; Wang, T.; Zou, X.; Sun, Z. Detection and identification of foreign bodies in conditioned steak based on ultrasound imaging. Food Sci. Technol. Res. 2024, 30, 269–280. [Google Scholar] [CrossRef]
  7. Ong, S.Y.; Zhang, C.; Dong, X.; Yao, S.Q. Recent Advances in Polymeric Nanoparticles for Enhanced Fluorescence and Photoacoustic Imaging. Angew. Chem. Int. Ed. 2021, 60, 17797–17809. [Google Scholar] [CrossRef]
  8. Chen, X.; Song, J.; Chen, X.; Yang, H. X-ray-activated nanosystems for theranostic applications. Chem. Soc. Rev. 2019, 48, 3073–3101. [Google Scholar] [CrossRef]
  9. Upputuri, P.K.; Pramanik, M. Photoacoustic imaging in the second near-infrared window: A review. J. Biomed. Opt. 2019, 24, 040901. [Google Scholar] [CrossRef]
  10. Nie, L.; Chen, X. Structural and functional photoacoustic molecular tomography aided by emerging contrast agents. Chem. Soc. Rev. 2014, 43, 7132–7170. [Google Scholar] [CrossRef]
  11. Li, W.; Chen, X. Gold nanoparticles for photoacoustic imaging. Nanomedicine 2015, 10, 299–320. [Google Scholar] [CrossRef] [PubMed]
  12. Huang, J.; Pu, K. Activatable Molecular Probes for Second Near-Infrared Fluorescence, Chemiluminescence, and Photoacoustic Imaging. Angew. Chem. Int. Ed. 2020, 59, 11717–11731. [Google Scholar] [CrossRef]
  13. Bell, A.G. On the production and reproduction of sound by light. Am. J. Sci. 1880, 3, 305–324. [Google Scholar] [CrossRef]
  14. Wang, L.V. Tutorial on photoacoustic microscopy and computed tomography. IEEE J. Sel. Top. Quantum Electron. 2008, 14, 171–179. [Google Scholar] [CrossRef]
  15. Khulal, U.; Zhao, J.; Hu, W.; Chen, Q. Nondestructive quantifying total volatile basic nitrogen (TVB-N) content in chicken using hyperspectral imaging (HSI) technique combined with different data dimension reduction algorithms. Food Chem. 2016, 197, 1191–1199. [Google Scholar] [CrossRef]
  16. Li, H.; Kutsanedzie, F.; Zhao, J.; Chen, Q. Quantifying Total Viable Count in Pork Meat Using Combined Hyperspectral Imaging and Artificial Olfaction Techniques. Food Anal. Methods 2016, 9, 3015–3024. [Google Scholar] [CrossRef]
  17. Chitra, S.; Muhammad Mudassir Arif, C.; Jitendra, P. Classification of pulse flours using near-infrared hyperspectral imaging. LWT 2022, 154, 112799. [Google Scholar]
  18. Zhu, Y.; Zou, X.; Shen, T.; Shi, J.; Zhao, J.; Holmes, M.; Li, G. Determination of total acid content and moisture content during solid-state fermentation processes using hyperspectral imaging. J. Food Eng. 2016, 174, 75–84. [Google Scholar] [CrossRef]
  19. Yuan, L.-M.; Cai, J.-R.; Sun, L.; Ye, C. A Preliminary Discrimination of Cluster Disqualified Shape for Table Grape by Mono-Camera Multi-Perspective Simultaneously Imaging Approach. Food Anal. Methods 2016, 9, 758–767. [Google Scholar] [CrossRef]
  20. Weng, S.; Tang, P.; Yuan, H.; Guo, B.; Yu, S.; Huang, L.; Xu, C. Hyperspectral imaging for accurate determination of rice variety using a deep learning network with multi-feature fusion. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2020, 234, 118237. [Google Scholar] [CrossRef]
  21. Sun, J.; Lu, X.; Mao, H.; Wu, X.; Gao, H. Quantitative Determination of Rice Moisture Based on Hyperspectral Imaging Technology and BCC-LS-SVR Algorithm. J. Food Process Eng. 2017, 40, e12446. [Google Scholar] [CrossRef]
  22. Jiang, S.; Sun, J.; Xin, Z.; Mao, H.; Wu, X.; Li, Q. Visualizing distribution of pesticide residues in mulberry leaves using NIR hyperspectral imaging. J. Food Process Eng. 2017, 40, e12510. [Google Scholar] [CrossRef]
  23. Zhou, X.; Sun, J.; Mao, H.; Wu, X.; Zhang, X.; Yang, N. Visualization research of moisture content in leaf lettuce leaves based on WT-PLSR and hyperspectral imaging technology. J. Food Process Eng. 2018, 41, e12647. [Google Scholar] [CrossRef]
  24. Xu, Y.; Chen, Q.; Liu, Y.; Sun, X.; Huang, Q.; Ouyang, Q.; Zhao, J. A Novel Hyperspectral Microscopic Imaging System for Evaluating Fresh Degree of Pork. Korean J. Food Sci. Anim. Resour. 2018, 38, 362–375. [Google Scholar]
  25. Chen, X.; Ding, H.; Yuan, L.-M.; Cai, J.-R.; Chen, X.; Lin, Y. New approach of simultaneous, multi-perspective imaging for quantitative assessment of the compactness of grape bunches. Aust. J. Grape Wine Res. 2018, 24, 413–420. [Google Scholar] [CrossRef]
  26. Li, Y.; Sun, J.; Wu, X.; Chen, Q.; Lu, B.; Dai, C. Detection of viability of soybean seed based on fluorescence hyperspectra and CARS-SVM-AdaBoost model. J. Food Process. Preserv. 2019, 43, e14238. [Google Scholar] [CrossRef]
  27. Ge, X.; Sun, J.; Lu, B.; Chen, Q.; Xun, W.; Jin, Y. Classification of oolong tea varieties based on hyperspectral imaging technology and BOSS-LightGBM model. J. Food Process Eng. 2019, 42, e13289. [Google Scholar] [CrossRef]
  28. Maturi, M.; Locatelli, E.; Monaco, I.; Comes Franchini, M. Current concepts in nanostructured contrast media development for in vivo photoacoustic imaging. Biomater. Sci. 2019, 7, 1746–1775. [Google Scholar] [CrossRef]
  29. Zhu, Y.; Feng, T.; Cheng, Q.; Wang, X.; Du, S.; Sato, N.; Yuan, J.; Kuniyil Ajith Singh, M. Towards Clinical Translation of LED-Based Photoacoustic Imaging: A Review. Sensors 2020, 20, 2484. [Google Scholar] [CrossRef]
  30. Mao, W.; Tang, J.; Dai, L.; He, X.; Li, J.; Cai, L.; Liao, P.; Jiang, R.; Zhou, J.; Wu, H. A General Strategy to Design Highly Fluorogenic Far-Red and Near-Infrared Tetrazine Bioorthogonal Probes. Angew. Chem. Int. Ed. 2021, 60, 2393–2397. [Google Scholar] [CrossRef]
  31. Gargiulo, S.; Albanese, S.; Mancini, M. State-of-the-Art Preclinical Photoacoustic Imaging in Oncology: Recent Advances in Cancer Theranostics. Contrast Media Mol. Imaging 2019, 2019, 5080267. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, Z.; Yang, F.; Zhang, W.; Xiong, K.; Yang, S. Towards in vivo photoacoustic human imaging: Shining a new light on clinical diagnostics. Fundam. Res. 2024, 4, 1314–1330. [Google Scholar] [CrossRef] [PubMed]
  33. Rich, L.J.; Seshadri, M. Photoacoustic monitoring of tumor and normal tissue response to radiation. Sci. Rep. 2016, 6, 21237. [Google Scholar] [CrossRef]
  34. Lovelock, J. A photoionization detector for gases and vapours. Nature 1960, 188, 401. [Google Scholar] [CrossRef] [PubMed]
  35. Rosencwaig, A. Photoacoustic spectroscopy of solids. Opt. Commun. 1973, 7, 305–308. [Google Scholar] [CrossRef]
  36. Vinson, F.S.; Eggleton, R.; Meiss, R. Variations in acoustic velocity in skeletal muscle determined by acoustic microscopy. In Ultrasound in Medicine; Springer: Berlin/Heidelberg, Germany, 1978; Volume 4, pp. 519–534. [Google Scholar]
  37. Helander, P.; Lundström, I.; McQueen, D. Photoacoustic study of layered samples. J. Appl. Phys. 1981, 52, 1146–1151. [Google Scholar] [CrossRef]
  38. Masujima, T.; Munekane, Y.; Kawai, C.; Yoshida, H.; Imai, H.; Juing-Yi, L.; Sato, Y. Photoacoustic imaging immunoassay for biological component microanalysis. In Proceedings of the Photoacoustic and Photothermal Phenomena: Proceedings of the 5th International Topical Meeting, Heidelberg, Germany, 27–30 July 1987; pp. 558–560. [Google Scholar]
  39. Masujima, T. X-ray photoacoustics for characterization and non-destructive evaluation. In Photoacoustic and Photothermal Phenomena II, Proceedings of the 6th International Topical Meeting, Baltimore, MD, USA, 31 July –3 August 1989; Springer: Berlin/Heidelberg, Germany, 1990; pp. 222–234. [Google Scholar]
  40. Kruger, R.A. Photoacoustic ultrasound. Med. Phys. 1994, 21, 127–131. [Google Scholar] [CrossRef]
  41. Kruger, R.A.; Liu, P.; Fang, Y.R.; Appledorn, C.R. Photoacoustic ultrasound (PAUS)—Reconstruction tomography. Med. Phys. 1995, 22, 1605–1609. [Google Scholar] [CrossRef]
  42. Oraevsky, A.A.; Karabutov, A.A.; Solomatin, S.V.; Savateeva, E.V.; Andreev, V.A.; Gatalica, Z.; Singh, H.; Fleming, R.D. Laser optoacoustic imaging of breast cancer in vivo. In Proceedings of the Biomedical Optoacoustics II, San Jose, CA, USA, 20–26 January 2001; pp. 6–15. [Google Scholar]
  43. Consalvey, M.; Perkins, R.G.; Paterson, D.M.; Underwood, G.J. PAM fluorescence: A beginners guide for benthic diatomists. Diatom Res. 2005, 20, 1–22. [Google Scholar] [CrossRef]
  44. Sun, Y.; Zhu, X.; Peng, J.; Li, F. Core–shell lanthanide upconversion nanophosphors as four-modal probes for tumor angiogenesis imaging. Acs. Nano 2013, 7, 11290–11300. [Google Scholar] [CrossRef]
  45. Gulzar, A.; Xu, J.; Yang, P.; He, F.; Xu, L. Upconversion processes: Versatile biological applications and biosafety. Nanoscale 2017, 9, 12248–12282. [Google Scholar] [CrossRef] [PubMed]
  46. Chen, P.-J.; Hu, S.-H.; Fan, C.-T.; Li, M.-L.; Chen, Y.-Y.; Chen, S.-Y.; Liu, D.-M. A novel multifunctional nano-platform with enhanced anti-cancer and photoacoustic imaging modalities using gold-nanorod-filled silica nanobeads. Chem. Commun. 2013, 49, 892–894. [Google Scholar] [CrossRef]
  47. Li, C.; Liu, C.; Fan, Y.; Ma, X.; Zhan, Y.; Lu, X.; Sun, Y. Recent development of near-infrared photoacoustic probes based on small-molecule organic dye. RSC Chem. Biol. 2021, 2, 743–758. [Google Scholar] [CrossRef] [PubMed]
  48. Zhang, J.; Campbell, R.E.; Ting, A.Y.; Tsien, R.Y. Creating new fluorescent probes for cell biology. Nat. Rev. Mol. Cell Biol. 2002, 3, 906–918. [Google Scholar] [CrossRef] [PubMed]
  49. Schröck, E.; Du Manoir, S.; Veldman, T.; Schoell, B.; Wienberg, J.; Ferguson-Smith, M.; Ning, Y.; Ledbetter, D.; Bar-Am, I.; Soenksen, D. Multicolor spectral karyotyping of human chromosomes. Science 1996, 273, 494–497. [Google Scholar] [CrossRef]
  50. Chaudhary, Z.; Khan, G.M.; Abeer, M.M.; Pujara, N.; Wan-Chi Tse, B.; McGuckin, M.A.; Popat, A.; Kumeria, T. Efficient photoacoustic imaging using indocyanine green (ICG) loaded functionalized mesoporous silica nanoparticles. Biomater. Sci. 2019, 7, 5002–5015. [Google Scholar] [CrossRef]
  51. de la Zerda, A.; Liu, Z.; Bodapati, S.; Teed, R.; Vaithilingam, S.; Khuri-Yakub, B.T.; Chen, X.; Dai, H.; Gambhir, S.S. Ultrahigh sensitivity carbon nanotube agents for photoacoustic molecular imaging in living mice. Nano Lett. 2010, 10, 2168–2172. [Google Scholar] [CrossRef]
  52. de la Zerda, A.; Bodapati, S.; Teed, R.; May, S.Y.; Tabakman, S.M.; Liu, Z.; Khuri-Yakub, B.T.; Chen, X.; Dai, H.; Gambhir, S.S. Family of enhanced photoacoustic imaging agents for high-sensitivity and multiplexing studies in living mice. ACS Nano 2012, 6, 4694–4701. [Google Scholar] [CrossRef]
  53. Pramanik, M.; Song, K.H.; Swierczewska, M.; Green, D.; Sitharaman, B.; Wang, L.V. In vivo carbon nanotube-enhanced non-invasive photoacoustic mapping of the sentinel lymph node. Phys. Med. Biol. 2009, 54, 3291. [Google Scholar] [CrossRef]
  54. Wu, X.; Liu, H.; Liu, J.; Haley, K.N.; Treadway, J.A.; Larson, J.P.; Ge, N.; Peale, F.; Bruchez, M.P. Immunofluorescent labeling of cancer marker Her2 and other cellular targets with semiconductor quantum dots. Nat. Biotechnol. 2003, 21, 41–46. [Google Scholar] [CrossRef]
  55. Kim, S.; Lim, Y.T.; Soltesz, E.G.; De Grand, A.M.; Lee, J.; Nakayama, A.; Parker, J.A.; Mihaljevic, T.; Laurence, R.G.; Dor, D.M. Near-infrared fluorescent type II quantum dots for sentinel lymph node mapping. Nat. Biotechnol. 2004, 22, 93–97. [Google Scholar] [CrossRef]
  56. Jaiswal, J.K.; Mattoussi, H.; Mauro, J.M.; Simon, S.M. Long-term multiple color imaging of live cells using quantum dot bioconjugates. Nat. Biotechnol. 2003, 21, 47–51. [Google Scholar] [CrossRef] [PubMed]
  57. Zhao, Q.; Li, F.; Huang, C. Phosphorescent chemosensors based on heavy-metal complexes. Chem. Soc. Rev. 2010, 39, 3007–3030. [Google Scholar] [CrossRef] [PubMed]
  58. Zhao, Q.; Huang, C.; Li, F. Phosphorescent heavy-metal complexes for bioimaging. Chem. Soc. Rev. 2011, 40, 2508–2524. [Google Scholar] [CrossRef] [PubMed]
  59. 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]
  60. Shaner, N.C.; Steinbach, P.A.; Tsien, R.Y. A guide to choosing fluorescent proteins. Nat. Methods 2005, 2, 905–909. [Google Scholar] [CrossRef]
  61. Nagai, T.; Ibata, K.; Park, E.S.; Kubota, M.; Mikoshiba, K.; Miyawaki, A. A variant of yellow fluorescent protein with fast and efficient maturation for cell-biological applications. Nat. Biotechnol. 2002, 20, 87–90. [Google Scholar] [CrossRef]
  62. Zhang, Y.; Jeon, M.; Rich, L.J.; Hong, H.; Geng, J.; Zhang, Y.; Shi, S.; Barnhart, T.E.; Alexandridis, P.; Huizinga, J.D. Non-invasive multimodal functional imaging of the intestine with frozen micellar naphthalocyanines. Nat. Nanotechnol. 2014, 9, 631–638. [Google Scholar] [CrossRef]
  63. Lee, C.; Jeon, M.; Jeon, M.Y.; Kim, J.; Kim, C. In vitro photoacoustic measurement of hemoglobin oxygen saturation using a single pulsed broadband supercontinuum laser source. Appl. Opt. 2014, 53, 3884–3889. [Google Scholar] [CrossRef]
  64. Koo, J.; Jeon, M.; Oh, Y.; Kang, H.W.; Kim, J.; Kim, C.; Oh, J. In vivo non-ionizing photoacoustic mapping of sentinel lymph nodes and bladders with ICG-enhanced carbon nanotubes. Phys. Med. Biol. 2012, 57, 7853. [Google Scholar] [CrossRef]
  65. Akers, W.J.; Kim, C.; Berezin, M.; Guo, K.; Fuhrhop, R.; Lanza, G.M.; Fischer, G.M.; Daltrozzo, E.; Zumbusch, A.; Cai, X. Noninvasive photoacoustic and fluorescence sentinel lymph node identification using dye-loaded perfluorocarbon nanoparticles. ACS Nano 2011, 5, 173–182. [Google Scholar] [CrossRef] [PubMed]
  66. Kim, C.; Song, H.-M.; Cai, X.; Yao, J.; Wei, A.; Wang, L.V. In vivo photoacoustic mapping of lymphatic systems with plasmon-resonant nanostars. J. Mater. Chem. 2011, 21, 2841–2844. [Google Scholar] [CrossRef]
  67. Kim, C.; Jeon, M.; Wang, L.V. Nonionizing photoacoustic cystography in vivo. Opt. Lett. 2011, 36, 3599–3601. [Google Scholar] [CrossRef] [PubMed]
  68. Ouyang, Q.; Zhang, M.; Yang, Y.; Din, Z.U.; Chen, Q. Mesoporous silica-modified upconversion biosensor coupled with real-time ion release properties for ultrasensitive detection of Staphylococcus aureus in meat. Food Control 2023, 145, 109444. [Google Scholar] [CrossRef]
  69. Yin, L.; Hu, X.; Hao, M.; Shi, J.; Zou, X.; Dusabe, K.D. Upconversion nanoparticles-based background-free selective fluorescence sensor developed for immunoassay of fipronil pesticide. J. Food Meas. Charact. 2023, 17, 3125–3133. [Google Scholar] [CrossRef]
  70. Zhang, B.; Li, H.; Pan, W.; Chen, Q.; Ouyang, Q.; Zhao, J. Dual-Color Upconversion Nanoparticles (UCNPs)-Based Fluorescent Immunoassay Probes for Sensitive Sensing Foodborne Pathogens. Food Anal. Methods 2017, 10, 2036–2045. [Google Scholar] [CrossRef]
  71. Zhang, Y.; Hassan, M.M.; Rong, Y.; Liu, R.; Li, H.; Ouyang, Q.; Chen, Q. An upconversion nanosensor for rapid and sensitive detection of tetracycline in food based on magnetic-field-assisted separation. Food Chem. 2022, 373, 131497. [Google Scholar] [CrossRef]
  72. Yahui, L.; Yanxiao, L.; Di, Z.; Weilong, T.; Jiyong, S.; Zhihua, L.; Hanyu, L.; Yinyin, Y.; Liu, Y.; Xin, W.; et al. A fluorescence resonance energy transfer probe based on functionalized graphene oxide and upconversion nanoparticles for sensitive and rapid detection of zearalenone. LWT 2021, 147, 111541. [Google Scholar]
  73. Huang, Y.; Du, Z.; Bao, G.; Fang, G.; Cappadona, M.; McClements, L.; Tuch, B.E.; Lu, H.; Xu, X. Smart Drug-Delivery System of Upconversion Nanoparticles Coated with Mesoporous Silica for Controlled Release. Pharmaceutics 2023, 15, 89. [Google Scholar] [CrossRef]
  74. Auzel, F. Upconversion and anti-stokes processes with f and d ions in solids. Chem. Rev. 2004, 104, 139–174. [Google Scholar] [CrossRef]
  75. Hilderbrand, S.A.; Weissleder, R. Near-infrared fluorescence: Application to in vivo molecular imaging. Curr. Opin. Chem. Biol. 2010, 14, 71–79. [Google Scholar] [CrossRef]
  76. Zhou, J.; Yu, M.; Sun, Y.; Zhang, X.; Zhu, X.; Wu, Z.; Wu, D.; Li, F. Fluorine-18-labeled Gd3+/Yb3+/Er3+ co-doped NaYF4 nanophosphors for multimodality PET/MR/UCL imaging. Biomaterials 2011, 32, 1148–1156. [Google Scholar] [CrossRef] [PubMed]
  77. Liu, Z.; Pu, F.; Huang, S.; Yuan, Q.; Ren, J.; Qu, X. Long-circulating Gd2O3: Yb3+, Er3+ up-conversion nanoprobes as high-performance contrast agents for multi-modality imaging. Biomaterials 2013, 34, 1712–1721. [Google Scholar] [CrossRef]
  78. Fan, W.; Shen, B.; Bu, W.; Chen, F.; Zhao, K.; Zhang, S.; Zhou, L.; Peng, W.; Xiao, Q.; Xing, H. Rattle-structured multifunctional nanotheranostics for synergetic chemo-/radiotherapy and simultaneous magnetic/luminescent dual-mode imaging. J. Am. Chem. Soc. 2013, 135, 6494–6503. [Google Scholar] [CrossRef] [PubMed]
  79. Cheng, L.; Wang, C.; Liu, Z. Upconversion nanoparticles and their composite nanostructures for biomedical imaging and cancer therapy. Nanoscale 2013, 5, 23–37. [Google Scholar] [CrossRef] [PubMed]
  80. Gu, B.; Zhang, Q. Recent advances on functionalized upconversion nanoparticles for detection of small molecules and ions in biosystems. Adv. Sci. 2018, 5, 1700609. [Google Scholar] [CrossRef]
  81. Rong, Y.; Ali, S.; Ouyang, Q.; Wang, L.; Li, H.; Chen, Q. Development of a bimodal sensor based on upconversion nanoparticles and surface-enhanced Raman for the sensitive determination of dibutyl phthalate in food. J. Food Compos. Anal. 2021, 100, 103929. [Google Scholar] [CrossRef]
  82. Rennie, C.; Huang, Y.; Siwakoti, P.; Du, Z.; Padula, M.; Bao, G.; Tuch, B.E.; Xu, X.; McClements, L. In Vitro Evaluation of A Hybrid Drug-Delivery Nanosystem for Fibrosis Prevention in Cell Therapy for Type 1 Diabetes. Nanomedicine 2023, 18, 53–66. [Google Scholar] [CrossRef]
  83. Vetrone, F.; Naccache, R.; de la Fuente, A.J.; Sanz-Rodríguez, F.; Blazquez-Castro, A.; Rodriguez, E.M.; Jaque, D.; Solé, J.G.; Capobianco, J.A. Intracellular imaging of HeLa cells by non-functionalized NaYF4: Er3+, Yb3+ upconverting nanoparticles. Nanoscale 2010, 2, 495–498. [Google Scholar] [CrossRef]
  84. Krämer, K.W.; Biner, D.; Frei, G.; Güdel, H.U.; Hehlen, M.P.; Lüthi, S.R. Hexagonal sodium yttrium fluoride based green and blue emitting upconversion phosphors. Chem. Mater. 2004, 16, 1244–1251. [Google Scholar] [CrossRef]
  85. Li, Z.; Zhang, Y.; Jiang, S. Multicolor core/shell-structured upconversion fluorescent nanoparticles. Adv. Mater. 2008, 20, 4765–4769. [Google Scholar] [CrossRef]
  86. Lin, H.; Duan, Y.; Man, Z.; Zareef, M.; Wang, Z.; Chen, Q. Quantitation of volatile aldehydes using chemoselective response dyes combined with multivariable data analysis. Food Chem. 2021, 353, 129485. [Google Scholar] [CrossRef] [PubMed]
  87. Huang, X.-W.; Zou, X.-B.; Shi, J.-Y.; Li, Z.-H.; Zhao, J.-W. Colorimetric sensor arrays based on chemo-responsive dyes for food odor visualization. Trends Food Sci. Technol. 2018, 81, 90–107. [Google Scholar]
  88. Lin, H.; Man, Z.X.; Kang, W.C.; Guan, B.B.; Chen, Q.S.; Xue, Z.L. A novel colorimetric sensor array based on boron-dipyrromethene dyes for monitoring the storage time of rice. Food Chem. 2018, 268, 300–306. [Google Scholar] [CrossRef]
  89. Lin, H.; Kang, W.; Kutsanedzie, F.Y.H.; Chen, Q. A Novel Nanoscaled Chemo Dye–Based Sensor for the Identification of Volatile Organic Compounds During the Mildewing Process of Stored Wheat. Food Anal. Methods 2019, 12, 2895–2907. [Google Scholar] [CrossRef]
  90. Ningqiu, T.; Jun, S.; Min, X.; Kunshan, Y.; Yan, C.; Dengjie, L. Identification of fumigated and dyed Lycium barbarum by hyperspectral imaging technology. J. Food Process Eng. 2022, 45, e13950. [Google Scholar] [CrossRef]
  91. Wang, F.; Banerjee, D.; Liu, Y.; Chen, X.; Liu, X. Upconversion nanoparticles in biological labeling, imaging, and therapy. Analyst 2010, 135, 1839–1854. [Google Scholar] [CrossRef]
  92. Zou, Y.; Shi, Y.; Wang, T.; Ji, S.; Zhang, X.; Shen, T.; Huang, X.; Xiao, J.; Farag, M.A.; Shi, J.; et al. Quantum dots as advanced nanomaterials for food quality and safety applications: A comprehensive review and future perspectives. Compr. Rev. Food Sci. Food Saf. 2024, 23, e13339. [Google Scholar] [CrossRef]
  93. Yosri, N.; Gao, S.; Zhou, R.; Wang, C.; Zou, X.; El-Seedi, H.R.; Guo, Z. Innovative quantum dots-based SERS for ultrasensitive reporting of contaminants in food: Fundamental concepts and practical implementations. Food Chem. 2025, 467, 142395. [Google Scholar] [CrossRef]
  94. Zhang, C.; Han, Y.; Lin, L.; Deng, N.; Chen, B.; Liu, Y. Development of Quantum Dots-Labeled Antibody Fluorescence Immunoassays for the Detection of Morphine. J. Agric. Food Chem. 2017, 65, 1290–1295. [Google Scholar] [CrossRef]
  95. Zhang, X.; Yu, X.; Wang, J.; Wang, Q.; Meng, H.; Wang, Z. One-Step Core/Multishell Quantum Dots-Based Fluoroimmunoassay for Screening of Deoxynivalenol in Maize. Food Anal. Methods 2018, 11, 2569–2578. [Google Scholar] [CrossRef]
  96. Li, Y.; Luo, S.; Sun, L.; Kong, D.; Sheng, J.; Wang, K.; Dong, C. A Green, Simple, and Rapid Detection for Amaranth in Candy Samples Based on the Fluorescence Quenching of Nitrogen-Doped Graphene Quantum Dots. Food Anal. Methods 2019, 12, 1658–1665. [Google Scholar] [CrossRef]
  97. Bi, X.; Li, L.; Luo, L.; Liu, X.; Li, J.; You, T. A ratiometric fluorescence aptasensor based on photoinduced electron transfer from CdTe QDs to WS2 NTs for the sensitive detection of zearalenone in cereal crops. Food Chem. 2022, 385, 132657. [Google Scholar] [CrossRef]
  98. Liang, N.; Hu, X.; Li, W.; Wang, Y.; Guo, Z.; Huang, X.; Li, Z.; Zhang, X.; Zhang, J.; Xiao, J.; et al. A dual-signal fluorescent sensor based on MoS2 and CdTe quantum dots for tetracycline detection in milk. Food Chem. 2022, 378, 132076. [Google Scholar] [CrossRef] [PubMed]
  99. Hu, X.; Shi, J.; Shi, Y.; Zou, X.; Arslan, M.; Zhang, W.; Huang, X.; Li, Z.; Xu, Y. Use of a smartphone for visual detection of melamine in milk based on Au@Carbon quantum dots nanocomposites. Food Chem. 2019, 272, 58–65. [Google Scholar] [CrossRef]
  100. Zhou, J.; Liu, Z.; Li, F. Upconversion nanophosphors for small-animal imaging. Chem. Soc. Rev. 2012, 41, 1323–1349. [Google Scholar] [CrossRef]
  101. Wang, F.; Liu, X. Recent advances in the chemistry of lanthanide-doped upconversion nanocrystals. Chem. Soc. Rev. 2009, 38, 976–989. [Google Scholar] [CrossRef]
  102. Zhou, J.; Liu, Q.; Feng, W.; Sun, Y.; Li, F. Upconversion luminescent materials: Advances and applications. Chem. Rev. 2015, 115, 395–465. [Google Scholar] [CrossRef]
  103. Zheng, W.; Huang, P.; Tu, D.; Ma, E.; Zhu, H.; Chen, X. Lanthanide-doped upconversion nano-bioprobes: Electronic structures, optical properties, and biodetection. Chem. Soc. Rev. 2015, 44, 1379–1415. [Google Scholar] [CrossRef]
  104. Sun, L.-D.; Wang, Y.-F.; Yan, C.-H. Paradigms and challenges for bioapplication of rare earth upconversion luminescent nanoparticles: Small size and tunable emission/excitation spectra. Acc. Chem. Res. 2014, 47, 1001–1009. [Google Scholar] [CrossRef]
  105. Sedlmeier, A.; Gorris, H.H. Surface modification and characterization of photon-upconverting nanoparticles for bioanalytical applications. Chem. Soc. Rev. 2015, 44, 1526–1560. [Google Scholar] [CrossRef] [PubMed]
  106. Prodi, L.; Rampazzo, E.; Rastrelli, F.; Speghini, A.; Zaccheroni, N. Imaging agents based on lanthanide doped nanoparticles. Chem. Soc. Rev. 2015, 44, 4922–4952. [Google Scholar] [CrossRef]
  107. Park, Y.I.; Lee, K.T.; Suh, Y.D.; Hyeon, T. Upconverting nanoparticles: A versatile platform for wide-field two-photon microscopy and multi-modal in vivo imaging. Chem. Soc. Rev. 2015, 44, 1302–1317. [Google Scholar] [CrossRef]
  108. Muhr, V.; Wilhelm, S.; Hirsch, T.; Wolfbeis, O.S. Upconversion nanoparticles: From hydrophobic to hydrophilic surfaces. Acc. Chem. Res. 2014, 47, 3481–3493. [Google Scholar] [CrossRef] [PubMed]
  109. Liu, X.; Yan, C.-H.; Capobianco, J.A. Photon upconversion nanomaterials. Chem. Soc. Rev. 2015, 44, 1299–1301. [Google Scholar] [CrossRef]
  110. Li, X.; Zhang, F.; Zhao, D. Lab on upconversion nanoparticles: Optical properties and applications engineering via designed nanostructure. Chem. Soc. Rev. 2015, 44, 1346–1378. [Google Scholar] [CrossRef] [PubMed]
  111. Wang, Y.; Song, S.; Zhang, S.; Zhang, H. Stimuli-responsive nanotheranostics based on lanthanide-doped upconversion nanoparticles for cancer imaging and therapy: Current advances and future challenges. Nano Today 2019, 25, 38–67. [Google Scholar] [CrossRef]
  112. Dong, H.; Du, S.-R.; Zheng, X.-Y.; Lyu, G.-M.; Sun, L.-D.; Li, L.-D.; Zhang, P.-Z.; Zhang, C.; Yan, C.-H. Lanthanide nanoparticles: From design toward bioimaging and therapy. Chem. Rev. 2015, 115, 10725–10815. [Google Scholar] [CrossRef]
  113. Chen, X.; Peng, D.; Ju, Q.; Wang, F. Photon upconversion in core–shell nanoparticles. Chem. Soc. Rev. 2015, 44, 1318–1330. [Google Scholar] [CrossRef]
  114. Chen, G.; Ågren, H.; Ohulchanskyy, T.Y.; Prasad, P.N. Light upconverting core–shell nanostructures: Nanophotonic control for emerging applications. Chem. Soc. Rev. 2015, 44, 1680–1713. [Google Scholar] [CrossRef]
  115. Li, X.; Zhang, F.; Zhao, D. Highly efficient lanthanide upconverting nanomaterials: Progresses and challenges. Nano Today 2013, 8, 643–676. [Google Scholar] [CrossRef]
  116. Xu, W.; Chen, X.; Song, H. Upconversion manipulation by local electromagnetic field. Nano Today 2017, 17, 54–78. [Google Scholar] [CrossRef]
  117. Lee, C.; Jeon, M.; Kim, C. Photoacoustic imaging in nanomedicine. In Applications of Nanoscience in Photomedicine; Elsevier: Amsterdam, The Netherlands, 2015; pp. 31–47. [Google Scholar]
  118. Kim, C.; Favazza, C.; Wang, L.V. In vivo photoacoustic tomography of chemicals: High-resolution functional and molecular optical imaging at new depths. Chem. Rev. 2010, 110, 2756–2782. [Google Scholar] [CrossRef] [PubMed]
  119. Sordillo, L.A.; Pu, Y.; Pratavieira, S.; Budansky, Y.; Alfano, R.R. Deep optical imaging of tissue using the second and third near-infrared spectral windows. J. Biomed. Opt. 2014, 19, 056004. [Google Scholar] [CrossRef]
  120. Cox, B.T.; Laufer, J.G.; Beard, P.C.; Arridge, S.R. Quantitative spectroscopic photoacoustic imaging: A review. J. Biomed. Opt. 2012, 17, 061202. [Google Scholar] [CrossRef]
  121. Maslov, K.I.; Wang, L.V. Photoacoustic imaging of biological tissue with intensity-modulated continuous-wave laser. J. Biomed. Opt. 2008, 13, 024006. [Google Scholar] [CrossRef]
  122. Wang, X.; Pang, Y.; Ku, G.; Xie, X.; Stoica, G.; Wang, L.V. Noninvasive laser-induced photoacoustic tomography for structural and functional in vivo imaging of the brain. Nat. Biotechnol. 2003, 21, 803–806. [Google Scholar] [CrossRef]
  123. Lashkari, B.; Mandelis, A. Comparison between pulsed laser and frequency-domain photoacoustic modalities: Signal-to-noise ratio, contrast, resolution, and maximum depth detectivity. Rev. Sci. Instrum. 2011, 82, 094903. [Google Scholar] [CrossRef]
  124. Xia, J.; Yao, J.; Wang, L.V. Photoacoustic tomography: Principles and advances. Electromagn. Waves 2014, 147, 1–22. [Google Scholar] [CrossRef]
  125. Xu, M.; Wang, L.V. Photoacoustic imaging in biomedicine. Rev. Sci. Instrum. 2006, 77, 041101. [Google Scholar] [CrossRef]
  126. Han Seung, H. Review of Photoacoustic Imaging for Imaging-Guided Spinal Surgery. Neurospine 2018, 15, 306–322. [Google Scholar] [CrossRef] [PubMed]
  127. Xia, J.; Kim, C.; Lovell, J.F. Opportunities for photoacoustic-guided drug delivery. Curr. Drug Targets 2015, 16, 571–581. [Google Scholar] [CrossRef] [PubMed]
  128. Pramanik, M.; Wang, L.V. Thermoacoustic and photoacoustic sensing of temperature. J. Biomed. Opt. 2009, 14, 054024. [Google Scholar] [CrossRef]
  129. Beard, P. Biomedical photoacoustic imaging. Interface Focus 2011, 1, 602–631. [Google Scholar] [CrossRef]
  130. Kim, J.; Park, S.; Jung, Y.; Chang, S.; Park, J.; Zhang, Y.; Lovell, J.F.; Kim, C. Programmable real-time clinical photoacoustic and ultrasound imaging system. Sci. Rep. 2016, 6, 35137. [Google Scholar] [CrossRef]
  131. Yao, J.; Wang, L.V. Photoacoustic microscopy. Laser Photonics Rev. 2013, 7, 758–778. [Google Scholar] [CrossRef]
  132. Wang, L.V.; Xia, J.; Yao, J. Photoacoustic Neuroimaging. In Neurophotonics and Brain Mapping; CRC Press: Boca Raton, FL, USA, 2017; pp. 235–256. [Google Scholar]
  133. Zhang, C.; Wang, Y. Comparison of various imaging modes for photoacoustic tomography. In Proceedings of the 13th International Conference on Biomedical Engineering, Valletta, Malta, 24–26 February 2020; pp. 121–124. [Google Scholar]
  134. Wang, L.V.; Hu, S. Photoacoustic tomography: In vivo imaging from organelles to organs. Science 2012, 335, 1458–1462. [Google Scholar] [CrossRef] [PubMed]
  135. Danielli, A.; Maslov, K.I.; Garcia-Uribe, A.; Winkler, A.M.; Li, C.; Wang, L.; Chen, Y.; Dorn, G.W., II; Wang, L.V. Label-free photoacoustic nanoscopy. J. Biomed. Opt. 2014, 19, 086006. [Google Scholar] [CrossRef]
  136. Jeon, M.; Kim, J.; Kim, C. Multiplane spectroscopic whole-body photoacoustic imaging of small animals in vivo. Med. Biol. Eng. Comput. 2016, 54, 283–294. [Google Scholar] [CrossRef]
  137. Kim, J.Y.; Lee, C.; Park, K.; Lim, G.; Kim, C. Fast optical-resolution photoacoustic microscopy using a 2-axis water-proofing MEMS scanner. Sci. Rep. 2015, 5, 7932. [Google Scholar] [CrossRef]
  138. Kim, J.Y.; Lee, C.; Park, K.; Lim, G.; Kim, C. A PDMS-based 2-axis waterproof scanner for photoacoustic microscopy. Sensors 2015, 15, 9815–9826. [Google Scholar] [CrossRef]
  139. Harrison, T.; Ranasinghesagara, J.; Lu, H.; Zemp, R.J. Fast-scanning ultrasonic-photoacoustic biomicroscope: In vivo performance. In Proceedings of the Photons Plus Ultrasound: Imaging and Sensing 2010, San Francisco, CA, USA, 24–26 January 2010; p. 75641X. [Google Scholar]
  140. Xie, Z.; Jiao, S.; Zhang, H.F.; Puliafito, C.A. Laser-scanning optical-resolution photoacoustic microscopy. Opt. Lett. 2009, 34, 1771–1773. [Google Scholar] [CrossRef]
  141. Rao, B.; Li, L.; Maslov, K.; Wang, L. Hybrid-scanning optical-resolution photoacoustic microscopy for in vivo vasculature imaging. Opt. Lett. 2010, 35, 1521–1523. [Google Scholar] [CrossRef]
  142. Heijblom, M.; Steenbergen, W.; Manohar, S. Clinical photoacoustic breast imaging: The twente experience. IEEE Pulse 2015, 6, 42–46. [Google Scholar] [CrossRef] [PubMed]
  143. Cha, M.G.; Lee, S.; Park, S.; Kang, H.; Lee, S.G.; Jeong, C.; Lee, Y.-S.; Kim, C.; Jeong, D.H. A dual modal silver bumpy nanoprobe for photoacoustic imaging and SERS multiplexed identification of in vivo lymph nodes. Nanoscale 2017, 9, 12556–12564. [Google Scholar] [CrossRef] [PubMed]
  144. Yajing, L.; Liming, N.; Xiaoyuan, C. Photoacoustic Molecular Imaging: From Multiscale Biomedical Applications Towards Early-Stage Theranostics. Trends Biotechnol. 2016, 34, 420–433. [Google Scholar]
  145. Jung, D.; Park, S.; Lee, C.; Kim, H. Recent Progress on Near-Infrared Photoacoustic Imaging: Imaging Modality and Organic Semiconducting Agents. Polymers 2019, 11, 1693. [Google Scholar] [CrossRef]
  146. Laufer, J.G.; Zhang, E.Z.; Treeby, B.E.; Cox, B.T.; Beard, P.C.; Johnson, P.; Pedley, B. In vivo preclinical photoacoustic imaging of tumor vasculature development and therapy. J. Biomed. Opt. 2012, 17, 056016. [Google Scholar] [CrossRef]
  147. Bohndiek, S.E.; Sasportas, L.S.; Machtaler, S.; Jokerst, J.V.; Hori, S.; Gambhir, S.S. Photoacoustic tomography detects early vessel regression and normalization during ovarian tumor response to the antiangiogenic therapy trebananib. J. Nucl. Med. 2015, 56, 1942–1947. [Google Scholar] [CrossRef]
  148. Guggenheim, J.A.; Allen, T.J.; Plumb, A.; Zhang, E.Z.; Rodriguez-Justo, M.; Punwani, S.; Beard, P.C. Photoacoustic imaging of human lymph nodes with endogenous lipid and hemoglobin contrast. J. Biomed. Opt. 2015, 20, 050504. [Google Scholar] [CrossRef]
  149. Xu, Z.; Wang, L.V.; Zhu, Q. In vivo photoacoustic tomography of mouse cerebral edema induced by cold injury. J. Biomed. Opt. 2011, 16, 066020. [Google Scholar] [CrossRef]
  150. Sarpong, F.; Yu, X.; Zhou, C.; Hongpeng, Y.; Uzoejinwa, B.B.; Bai, J.; Wu, B.; Ma, H. Influence of anti-browning agent pretreatment on drying kinetics, enzymes inactivation and other qualities of dried banana (Musa ssp.) under relative humidity-convective air dryer. J. Food Meas. Charact. 2018, 12, 1229–1241. [Google Scholar] [CrossRef]
  151. Zhang, H.; Mahunu, G.K.; Castoria, R.; Apaliya, M.T.; Yang, Q. Augmentation of biocontrol agents with physical methods against postharvest diseases of fruits and vegetables. Trends Food Sci. Technol. 2017, 69, 36–45. [Google Scholar] [CrossRef]
  152. Zeng, X.; Tang, L.; Zhang, W.; Hong, X.; Xiao, Y. Shape and size effects of gold nanoparticles for tumor photoacoustic imaging and photothermal therapy within the NIR-I and NIR-II Biowindows. Small 2025, 21, 2412296. [Google Scholar] [CrossRef] [PubMed]
  153. Zhang, L.; Wang, M.; Wu, F.; Liu, L.; Ren, X.; Hai, Z. Intracellular Formation of Hemicyanine Nanoparticle Enhances Tumor-Targeting Photoacoustic Imaging and Photothermal Therapy. Adv. Healthc. Mater. 2023, 12, 2202676. [Google Scholar] [CrossRef]
  154. Wang, F.; Men, X.; Chen, H.; Mi, F.; Xu, M.; Men, X.; Yuan, Z.; Lo, P.K. Second near-infrared photoactivatable biocompatible polymer nanoparticles for effective in vitro and in vivo cancer theranostics. Nanoscale 2021, 13, 13410–13420. [Google Scholar] [CrossRef]
  155. Sridharan, B.; Lim, H.G. Advances in photoacoustic imaging aided by nano contrast agents: Special focus on role of lymphatic system imaging for cancer theranostics. J. Nanobiotechnol. 2023, 21, 437. [Google Scholar] [CrossRef] [PubMed]
  156. Li, F.; Tu, L.; Zhang, Y.; Huang, D.; Liu, X.; Zhang, X.; Du, J.; Fan, R.; Yang, C.; Krämer, K.W. Size-dependent lanthanide energy transfer amplifies upconversion luminescence quantum yields. Nat. Photonics 2024, 18, 440–449. [Google Scholar] [CrossRef]
  157. Fardian-Melamed, N.; Skripka, A.; Ursprung, B.; Lee, C.; Darlington, T.P.; Teitelboim, A.; Qi, X.; Wang, M.; Gerton, J.M.; Cohen, B.E. Infrared nanosensors of piconewton to micronewton forces. Nature 2025, 637, 70–75. [Google Scholar] [CrossRef]
  158. Casar, J.R.; McLellan, C.A.; Shi, C.; Stiber, A.; Lay, A.; Siefe, C.; Parakh, A.; Gaerlan, M.; Gu, X.W.; Goodman, M.B. Upconverting microgauges reveal intraluminal force dynamics in vivo. Nature 2025, 637, 76–83. [Google Scholar] [CrossRef]
  159. Chen, T.; Su, L.; Lin, L.; Ge, X.; Bai, F.; Niu, M.; Wang, C.; Song, J.; Guo, S.; Yang, H. Mesoporous radiosensitized nanoprobe for enhanced NIR-II photoacoustic imaging-guided accurate radio-chemotherapy. Nano Res. 2022, 15, 4154–4163. [Google Scholar] [CrossRef]
  160. Al-Salihi, M.; Ghellab, S.E.; Li, Y.; Luo, C.; Kalsoom, U.-E.; Liu, L. Effective Rapid Fluorescence Lifetime Imaging of the Brain: A Novel Approach Using Upconversion Photoluminescence Lifetime Based on Gate-Width Acquisition. Nano Lett. 2024, 24, 14973–14982. [Google Scholar] [CrossRef] [PubMed]
  161. Dukhno, O.; Ghosh, S.; Greiner, V.; Bou, S.; Godet, J.; Muhr, V.; Buchner, M.; Hirsch, T.; Mély, Y.; Przybilla, F. Targeted single particle tracking with upconverting nanoparticles. ACS Appl. Mater. Interfaces 2024, 16, 11217–11227. [Google Scholar] [CrossRef] [PubMed]
  162. Ding, L.; Chen, C.; Shan, X.; Liu, B.; Wang, D.; Du, Z.; Zhao, G.; Su, Q.P.; Yang, Y.; Halkon, B. Optical nonlinearity enabled super-resolved multiplexing microscopy. Adv. Mater. 2024, 36, 2308844. [Google Scholar] [CrossRef] [PubMed]
  163. Du, P.; Wei, Y.; Liang, Y.; An, R.; Liu, S.; Lei, P.; Zhang, H. Near-Infrared-Responsive Rare Earth Nanoparticles for Optical Imaging and Wireless Phototherapy. Adv. Sci. 2024, 11, 2305308. [Google Scholar] [CrossRef]
  164. Liu, X.-M.; Zhu, Z.-Z.; He, X.-R.; Zou, Y.-H.; Chen, Q.; Wang, X.-Y.; Liu, H.-M.; Qiao, X.; Wang, X.; Xu, J.-Y. NIR light and GSH dual-responsive upconversion nanoparticles loaded with multifunctional platinum (IV) prodrug and RGD peptide for precise cancer therapy. ACS Appl. Mater. Interfaces 2024, 16, 40753–40766. [Google Scholar] [CrossRef]
  165. Das, B. Transition Metal Complex-Loaded Nanosystems: Advances in Stimuli-Responsive Cancer Therapies. Small 2025, 21, 2410338. [Google Scholar] [CrossRef]
  166. Hale, G.M.; Querry, M.R. Optical constants of water in the 200-nm to 200-μm wavelength region. Appl. Opt. 1973, 12, 555–563. [Google Scholar] [CrossRef]
  167. Van Veen, R.; Sterenborg, H.J.; Pifferi, A.; Torricelli, A.; Chikoidze, E.; Cubeddu, R. Determination of visible near-IR absorption coefficients of mammalian fat using time-and spatially resolved diffuse reflectance and transmission spectroscopy. J. Biomed. Opt. 2005, 10, 054004. [Google Scholar] [CrossRef]
  168. Tsai, C.-L.; Chen, J.-C.; Wang, W.-J. Near-infrared absorption property of biological soft tissue constituents. J. Med. Biol. Eng. 2001, 21, 7–14. [Google Scholar]
  169. Weber, J.; Beard, P.C.; Bohndiek, S.E. Contrast agents for molecular photoacoustic imaging. Nat. Methods 2016, 13, 639–650. [Google Scholar] [CrossRef] [PubMed]
  170. Tenzer, S.; Docter, D.; Kuharev, J.; Musyanovych, A.; Fetz, V.; Hecht, R.; Schlenk, F.; Fischer, D.; Kiouptsi, K.; Reinhardt, C. Rapid formation of plasma protein corona critically affects nanoparticle pathophysiology. Nat. Nanotechnol. 2013, 8, 772–781. [Google Scholar] [CrossRef]
  171. Lynch, I.; Dawson, K.A. Protein-nanoparticle interactions. Nano Today 2008, 3, 40–47. [Google Scholar] [CrossRef]
  172. Zhou, M.; Huang, H.; Wang, D.; Lu, H.; Chen, J.; Chai, Z.; Yao, S.Q.; Hu, Y. Light-Triggered PEGylation/dePEGylation of the Nanocarriers for Enhanced Tumor Penetration. Nano Lett. 2019, 19, 3671–3675. [Google Scholar] [CrossRef]
  173. Lundqvist, M. Nanoparticles: Tracking protein corona over time. Nat. Nanotechnol. 2013, 8, 701–702. [Google Scholar] [CrossRef] [PubMed]
  174. Petros, R.A.; DeSimone, J.M. Strategies in the design of nanoparticles for therapeutic applications. Nat. Rev. Drug Discov. 2010, 9, 615–627. [Google Scholar] [CrossRef] [PubMed]
  175. Nie, L.; Wang, S.; Wang, X.; Rong, P.; Ma, Y.; Liu, G.; Huang, P.; Lu, G.; Chen, X. In vivo volumetric photoacoustic molecular angiography and therapeutic monitoring with targeted plasmonic nanostars. Small 2014, 10, 1585–1593. [Google Scholar] [CrossRef]
  176. Chatni, M.R.; Xia, J.; Maslov, K.I.; Guo, Z.; Wang, K.; Anastasio, M.A.; Wang, L.V.; Sohn, R.; Arbeit, J.M.; Zhang, Y. Tumor glucose metabolism imaged in vivo in small animals with whole-body photoacoustic computed tomography. J. Biomed. Opt. 2012, 17, 076012. [Google Scholar] [CrossRef]
  177. Dragulescu-Andrasi, A.; Kothapalli, S.-R.; Tikhomirov, G.A.; Rao, J.; Gambhir, S.S. Activatable oligomerizable imaging agents for photoacoustic imaging of furin-like activity in living subjects. J. Am. Chem. Soc. 2013, 135, 11015–11022. [Google Scholar] [CrossRef]
  178. Wen, S.; Zhou, J.; Zheng, K.; Bednarkiewicz, A.; Liu, X.; Jin, D. Advances in highly doped upconversion nanoparticles. Nat. Commun. 2018, 9, 2415. [Google Scholar] [CrossRef]
  179. Zhang, Y.; Zhang, L.; Deng, R.; Tian, J.; Zong, Y.; Jin, D.; Liu, X. Multicolor barcoding in a single upconversion crystal. J. Am. Chem. Soc. 2014, 136, 4893–4896. [Google Scholar] [CrossRef] [PubMed]
  180. Xie, X.; Gao, N.; Deng, R.; Sun, Q.; Xu, Q.-H.; Liu, X. Mechanistic investigation of photon upconversion in Nd3+-sensitized core–shell nanoparticles. J. Am. Chem. Soc. 2013, 135, 12608–12611. [Google Scholar] [CrossRef] [PubMed]
  181. He, S.; Johnson, N.J.; Nguyen Huu, V.A.; Huang, Y.; Almutairi, A. Leveraging spectral matching between photosensitizers and upconversion nanoparticles for 808 nm-activated photodynamic therapy. Chem. Mater. 2018, 30, 3991–4000. [Google Scholar] [CrossRef]
  182. Campos-Goncalves, I.; Costa, B.F.; Santos, R.F.; Duraes, L. Superparamagnetic core-shell nanocomplexes doped with Yb3+: Er3+/Ho3+ rare-earths for upconversion fluorescence. Mater. Des. 2017, 130, 263–274. [Google Scholar] [CrossRef]
  183. Rinkel, T.; Raj, A.N.; Dühnen, S.; Haase, M. Synthesis of 10 nm β-NaYF4: Yb, Er/NaYF4 core/shell upconversion nanocrystals with 5 nm particle cores. Angew. Chem. Int. Ed. 2016, 55, 1164–1167. [Google Scholar] [CrossRef]
  184. Xu, Z.; Kang, X.; Li, C.; Hou, Z.; Zhang, C.; Yang, D.; Li, G.; Lin, J. Ln3+ (Ln = Eu, Dy, Sm, and Er) ion-doped YVO4 nano/microcrystals with multiform morphologies: Hydrothermal synthesis, growing mechanism, and luminescent properties. Inorg. Chem. 2010, 49, 6706–6715. [Google Scholar] [CrossRef]
  185. Wang, M.; Liu, J.-L.; Zhang, Y.-X.; Hou, W.; Wu, X.-L.; Xu, S.-K. Two-phase solvothermal synthesis of rare-earth doped NaYF4 upconversion fluorescent nanocrystals. Mater. Lett. 2009, 63, 325–327. [Google Scholar] [CrossRef]
  186. Tang, S.; Huang, M.; Wang, J.; Yu, F.; Shang, G.; Wu, J. Hydrothermal synthesis and luminescence properties of GdVO4: Ln3+ (Ln = Eu, Sm, Dy) phosphors. J. Alloys Compd. 2012, 513, 474–480. [Google Scholar] [CrossRef]
  187. Min, Y.; Li, J.; Liu, F.; Padmanabhan, P.; Yeow, E.K.; Xing, B. Recent advance of biological molecular imaging based on lanthanide-doped upconversion-luminescent nanomaterials. Nanomaterials 2014, 4, 129–154. [Google Scholar] [CrossRef]
  188. Mao, L.; Lu, Z.; He, N.; Zhang, L.; Deng, Y.; Duan, D. A new method for improving the accuracy of miRNA detection with NaYF 4: Yb, Er upconversion nanoparticles. Sci. China Chem. 2017, 60, 157–162. [Google Scholar] [CrossRef]
  189. Liu, B.; Li, C.; Xing, B.; Yang, P.; Lin, J. Multifunctional UCNPs@ PDA-ICG nanocomposites for upconversion imaging and combined photothermal/photodynamic therapy with enhanced antitumor efficacy. J. Mater. Chem. B 2016, 4, 4884–4894. [Google Scholar] [CrossRef]
  190. Zhang, L.; Ling, B.; Wang, L.; Chen, H. A near-infrared luminescent Mn2+-doped NaYF4: Yb, Tm/Fe3+ upconversion nanoparticles redox reaction system for the detection of GSH/Cys/AA. Talanta 2017, 172, 95–101. [Google Scholar] [CrossRef]
  191. Chen, G.; Qiu, H.; Prasad, P.N.; Chen, X. Upconversion nanoparticles: Design, nanochemistry, and applications in theranostics. Chem. Rev. 2014, 114, 5161–5214. [Google Scholar] [CrossRef]
  192. Tian, Y.; Tian, Y.; Huang, P.; Wang, L.; Shi, Q.; Cui, C.E. Effect of Yb3+ concentration on upconversion luminescence and temperature sensing behavior in Yb3+/Er3+ co-doped YNbO4 nanoparticles prepared via molten salt route. Chem. Eng. J. 2016, 297, 26–34. [Google Scholar] [CrossRef]
  193. Cheng, X.; Pan, Y.; Yuan, Z.; Wang, X.; Su, W.; Yin, L.; Xie, X.; Huang, L. Er3+ sensitized photon upconversion nanocrystals. Adv. Funct. Mater. 2018, 28, 1800208. [Google Scholar] [CrossRef]
  194. Liang, G.; Wang, H.; Shi, H.; Wang, H.; Zhu, M.; Jing, A.; Li, J.; Li, G. Recent progress in the development of upconversion nanomaterials in bioimaging and disease treatment. J. Nanobiotechnol. 2020, 18, 1–22. [Google Scholar] [CrossRef] [PubMed]
  195. Passuello, T.; Pedroni, M.; Piccinelli, F.; Polizzi, S.; Marzola, P.; Tambalo, S.; Conti, G.; Benati, D.; Vetrone, F.; Bettinelli, M. PEG-capped, lanthanide doped GdF3 nanoparticles: Luminescent and T2 contrast agents for optical and MRI multimodal imaging. Nanoscale 2012, 4, 7682–7689. [Google Scholar] [CrossRef]
  196. Schäfer, H.; Ptacek, P.; Zerzouf, O.; Haase, M. Synthesis and optical properties of KYF4/Yb, Er nanocrystals, and their surface modification with undoped KYF4. Adv. Funct. Mater. 2008, 18, 2913–2918. [Google Scholar] [CrossRef]
  197. Chen, C.; Sun, L.-D.; Li, Z.-X.; Li, L.-L.; Zhang, J.; Zhang, Y.-W.; Yan, C.-H. Ionic liquid-based route to spherical NaYF4 nanoclusters with the assistance of microwave radiation and their multicolor upconversion luminescence. Langmuir 2010, 26, 8797–8803. [Google Scholar] [CrossRef]
  198. Schäfer, H.; Ptacek, P.; Eickmeier, H.; Haase, M. Synthesis of Hexagonal Yb3+, Er3+-Doped NaYF4 Nanocrystals at Low Temperature. Adv. Funct. Mater. 2009, 19, 3091–3097. [Google Scholar] [CrossRef]
  199. Chen, D.; Yu, Y.; Huang, F.; Wang, Y. Phase transition from hexagonal LnF3 (Ln = La, Ce, Pr) to cubic Ln0.8M0.2F2.8 (M = Ca, Sr, Ba) nanocrystals with enhanced upconversion induced by alkaline-earth doping. Chem. Commun. 2011, 47, 2601–2603. [Google Scholar] [CrossRef]
  200. Yin, W.; Zhao, L.; Zhou, L.; Gu, Z.; Liu, X.; Tian, G.; Jin, S.; Yan, L.; Ren, W.; Xing, G. Enhanced red emission from GdF3: Yb3+, Er3+ upconversion nanocrystals by Li+ doping and their application for bioimaging. Chem.–A Eur. J. 2012, 18, 9239–9245. [Google Scholar] [CrossRef]
  201. Li, F.; Li, C.; Liu, X.; Bai, T.; Dong, W.; Zhang, X.; Shi, Z.; Feng, S. Microwave-assisted synthesis and up–down conversion luminescent properties of multicolor hydrophilic LaF3: Ln3+ nanocrystals. Dalton Trans. 2013, 42, 2015–2022. [Google Scholar] [CrossRef]
  202. Wang, F.; Liu, X. Upconversion multicolor fine-tuning: Visible to near-infrared emission from lanthanide-doped NaYF4 nanoparticles. J. Am. Chem. Soc. 2008, 130, 5642–5643. [Google Scholar] [CrossRef] [PubMed]
  203. Gnach, A.; Bednarkiewicz, A. Lanthanide-doped up-converting nanoparticles: Merits and challenges. Nano Today 2012, 7, 532–563. [Google Scholar] [CrossRef]
  204. Rafique, R.; Baek, S.H.; Chang, S.-J.; Gul, A.R.; Park, T.J. A facile hydrothermal synthesis of highly luminescent NaYF4: Yb3+/Er3+ upconversion nanoparticles and their biomonitoring capability. Mater. Sci. Eng. C 2019, 99, 1067–1074. [Google Scholar] [CrossRef] [PubMed]
  205. Chien, Y.H.; Chan, K.K.; Yap, S.H.K.; Yong, K.T. NIR-responsive nanomaterials and their applications; upconversion nanoparticles and carbon dots: A perspective. J. Chem. Technol. Biotechnol. 2018, 93, 1519–1528. [Google Scholar] [CrossRef]
  206. Li, Z.; Yuan, H.; Yuan, W.; Su, Q.; Li, F. Upconversion nanoprobes for biodetections. Coord. Chem. Rev. 2018, 354, 155–168. [Google Scholar] [CrossRef]
  207. Tong, R.; Lin, H.; Chen, Y.; An, N.; Wang, G.; Pan, X.; Qu, F. Near-infrared mediated chemo/photodynamic synergistic therapy with DOX-UCNPs@ mSiO2/TiO2-TC nanocomposite. Mater. Sci. Eng. C 2017, 78, 998–1005. [Google Scholar] [CrossRef] [PubMed]
  208. Wang, M.; Abbineni, G.; Clevenger, A.; Mao, C.; Xu, S. Upconversion nanoparticles: Synthesis, surface modification and biological applications. Nanomed. Nanotechnol. Biol. Med. 2011, 7, 710–729. [Google Scholar] [CrossRef]
  209. Shangguan, M.; Xia, H.; Wang, C.; Qiu, J.; Shentu, G.; Zhang, Q.; Dou, X.; Pan, J.-W. All-fiber upconversion high spectral resolution wind lidar using a Fabry-Perot interferometer. Opt. Express 2016, 24, 19322–19336. [Google Scholar] [CrossRef] [PubMed]
  210. Chen, D.; Yu, Y.; Huang, P.; Wang, Y. Nanocrystallization of lanthanide trifluoride in an aluminosilicate glass matrix: Dimorphism and rare earth partition. CrystEngComm 2009, 11, 1686–1690. [Google Scholar] [CrossRef]
  211. He, M.; Pang, X.; Liu, X.; Jiang, B.; He, Y.; Snaith, H.; Lin, Z. Monodisperse dual-functional upconversion nanoparticles enabled near-infrared organolead halide perovskite solar cells. Angew. Chem. Int. Ed. 2016, 55, 4280–4284. [Google Scholar] [CrossRef]
  212. Wang, M.; Mi, C.; Zhang, Y.; Liu, J.; Li, F.; Mao, C.; Xu, S. NIR-responsive silica-coated NaYbF4: Er/Tm/Ho upconversion fluorescent nanoparticles with tunable emission colors and their applications in immunolabeling and fluorescent imaging of cancer cells. J. Phys. Chem. C 2009, 113, 19021–19027. [Google Scholar] [CrossRef]
  213. Gu, Z.; Yan, L.; Tian, G.; Li, S.; Chai, Z.; Zhao, Y. Recent advances in design and fabrication of upconversion nanoparticles and their safe theranostic applications. Adv. Mater. 2013, 25, 3758–3779. [Google Scholar] [CrossRef]
  214. Idris, N.M.; Gnanasammandhan, M.K.; Zhang, J.; Ho, P.C.; Mahendran, R.; Zhang, Y. In vivo photodynamic therapy using upconversion nanoparticles as remote-controlled nanotransducers. Nat. Med. 2012, 18, 1580–1585. [Google Scholar] [CrossRef] [PubMed]
  215. Jin, B.; Wang, S.; Lin, M.; Jin, Y.; Zhang, S.; Cui, X.; Gong, Y.; Li, A.; Xu, F.; Lu, T.J. Upconversion nanoparticles based FRET aptasensor for rapid and ultrasenstive bacteria detection. Biosens. Bioelectron. 2017, 90, 525–533. [Google Scholar] [CrossRef]
  216. Kumar, R.; Nyk, M.; Ohulchanskyy, T.Y.; Flask, C.A.; Prasad, P.N. Combined optical and MR bioimaging using rare earth ion doped NaYF4 nanocrystals. Adv. Funct. Mater. 2009, 19, 853–859. [Google Scholar] [CrossRef]
  217. Xing, H.; Bu, W.; Ren, Q.; Zheng, X.; Li, M.; Zhang, S.; Qu, H.; Wang, Z.; Hua, Y.; Zhao, K. A NaYbF4: Tm3+ nanoprobe for CT and NIR-to-NIR fluorescent bimodal imaging. Biomaterials 2012, 33, 5384–5393. [Google Scholar] [CrossRef]
  218. Ni, D.; Bu, W.; Zhang, S.; Zheng, X.; Li, M.; Xing, H.; Xiao, Q.; Liu, Y.; Hua, Y.; Zhou, L. Single Ho3+-doped upconversion nanoparticles for high-performance T2-weighted brain tumor diagnosis and MR/UCL/CT multimodal imaging. Adv. Funct. Mater. 2014, 24, 6613–6620. [Google Scholar] [CrossRef]
  219. Teng, X.; Zhu, Y.; Wei, W.; Wang, S.; Huang, J.; Naccache, R.; Hu, W.; Tok, A.I.Y.; Han, Y.; Zhang, Q. Lanthanide-doped NaxScF3+x nanocrystals: Crystal structure evolution and multicolor tuning. J. Am. Chem. Soc. 2012, 134, 8340–8343. [Google Scholar] [CrossRef] [PubMed]
  220. Chen, G.; Ohulchanskyy, T.Y.; Kumar, R.; Ågren, H.; Prasad, P.N. Ultrasmall monodisperse NaYF4: Yb3+/Tm3+ nanocrystals with enhanced near-infrared to near-infrared upconversion photoluminescence. ACS Nano 2010, 4, 3163–3168. [Google Scholar] [CrossRef] [PubMed]
  221. Wong, H.-T.; Vetrone, F.; Naccache, R.; Chan, H.L.W.; Hao, J.; Capobianco, J.A. Water dispersible ultra-small multifunctional KGdF4: Tm3+, Yb3+ nanoparticles with near-infrared to near-infrared upconversion. J. Mater. Chem. 2011, 21, 16589–16596. [Google Scholar] [CrossRef]
  222. Wang, G.; Peng, Q.; Li, Y. Upconversion luminescence of monodisperse CaF2: Yb3+/Er3+ nanocrystals. J. Am. Chem. Soc. 2009, 131, 14200–14201. [Google Scholar] [CrossRef]
  223. Dong, N.-N.; Pedroni, M.; Piccinelli, F.; Conti, G.; Sbarbati, A.; Ramírez-Hernández, J.E.; Maestro, L.M.; Iglesias-de la Cruz, M.C.; Sanz-Rodriguez, F.; Juarranz, A. NIR-to-NIR two-photon excited CaF2:Tm3+, Yb3+ nanoparticles: Multifunctional nanoprobes for highly penetrating fluorescence bio-imaging. ACS Nano 2011, 5, 8665–8671. [Google Scholar] [CrossRef]
  224. Liu, Q.; Sun, Y.; Yang, T.; Feng, W.; Li, C.; Li, F. Sub-10 nm hexagonal lanthanide-doped NaLuF4 upconversion nanocrystals for sensitive bioimaging in vivo. J. Am. Chem. Soc. 2011, 133, 17122–17125. [Google Scholar] [CrossRef]
  225. Shi, F.; Wang, J.; Zhai, X.; Zhao, D.; Qin, W. Facile synthesis of β-NaLuF4: Yb/Tm hexagonal nanoplates with intense ultraviolet upconversion luminescence. CrystEngComm 2011, 13, 3782–3787. [Google Scholar] [CrossRef]
  226. Yang, D.; Dai, Y.; Ma, P.; Kang, X.; Cheng, Z.; Li, C.; Lin, J. One-step synthesis of small-sized and water-solule NaREF4 upconversion nanoparticles for in vitro cell imaging and drug delivery. Chem.–A Eur. J. 2013, 19, 2685–2694. [Google Scholar] [CrossRef]
  227. Sarkar, S.; Meesaragandla, B.; Hazra, C.; Mahalingam, V. Sub-5 nm Ln3+-doped BaLuF5 Nanocrystals: A Platform to Realize Upconversion via Interparticle Energy Transfer (IPET). Adv. Mater. 2013, 25, 856–860. [Google Scholar] [CrossRef]
  228. Yersin, H.; Donges, D. Low-lying electronic states and photophysical properties of organometallic Pd (II) and Pt (II) compounds. Modern research trends presented in detailed case studies. In Transition Metal and Rare Earth Compounds: Excited States, Transitions, Interactions II; Springer: Berlin/Heidelberg, Germany, 2001; pp. 81–186. [Google Scholar]
  229. Gamelin, D.R.; Güdel, H.U. Design of luminescent inorganic materials: New photophysical processes studied by optical spectroscopy. Acc. Chem. Res. 2000, 33, 235–242. [Google Scholar] [CrossRef]
  230. Qiu, H.; Chen, G.; Sun, L.; Hao, S.; Han, G.; Yang, C. Ethylenediaminetetraacetic acid (EDTA)-controlled synthesis of multicolor lanthanide doped BaYF5 upconversion nanocrystals. J. Mater. Chem. 2011, 21, 17202–17208. [Google Scholar] [CrossRef]
  231. Yang, L.; Li, J.; Pan, W.; Wang, H.; Li, N.; Tang, B. Fluorescence and photoacoustic dual-mode imaging of tumor-related mRNA with a covalent linkage-based DNA nanoprobe. Chem. Commun. 2018, 54, 3656–3659. [Google Scholar] [CrossRef]
  232. Ye, Z.; Srivastava, P.K.; Xu, Y.; Wang, W.; Jing, L.; Chen, S.-L.; Tu, C.-C. Surface-Functionalized Silicon Nanoparticles as Contrast Agents for Photoacoustic Microscopy Imaging. ACS Appl. Nano Mater. 2019, 2, 7577–7584. [Google Scholar] [CrossRef]
  233. Mantri, Y.; Jokerst, J.V. Engineering Plasmonic Nanoparticles for Enhanced Photoacoustic Imaging. ACS Nano 2020, 14, 9408–9422. [Google Scholar] [CrossRef] [PubMed]
  234. Jana, N.R.; Gearheart, L.; Murphy, C.J. Seeding growth for size control of 5–40 nm diameter gold nanoparticles. Langmuir 2001, 17, 6782–6786. [Google Scholar] [CrossRef]
  235. Hirsch, L.R.; Gobin, A.M.; Lowery, A.R.; Tam, F.; Drezek, R.A.; Halas, N.J.; West, J.L. Metal nanoshells. Ann. Biomed. Eng. 2006, 34, 15–22. [Google Scholar] [CrossRef] [PubMed]
  236. Xu, Y.; Kutsanedzie, F.Y.H.; Ali, S.; Wang, P.; Li, C.; Ouyang, Q.; Li, H.; Chen, Q. Cysteamine-mediated upconversion sensor for lead ion detection in food. J. Food Meas. Charact. 2021, 15, 4849–4857. [Google Scholar] [CrossRef]
  237. Wu, W.; Ahmad, W.; Hassan, M.M.; Wu, J.; Ouyang, Q.; Chen, Q. An upconversion biosensor based on inner filter effect for dual-role recognition of sulfadimethoxine in aquatic samples. Food Chem. 2024, 437, 137832. [Google Scholar] [CrossRef]
  238. Ouyang, Q.; Rong, Y.; Wang, B.; Ahmad, W.; Liu, S.; Chen, Q. An innovative solid-phase biosensor for rapid on-site detection of N-nitrosodimethylamine incorporating zein film and upconversion nanoparticles. Food Chem. 2024, 430, 136981. [Google Scholar] [CrossRef]
  239. Li, S.; Wu, J.; Zhang, S.; Jiao, T.; Wei, J.; Chen, X.; Chen, Q.; Chen, Q. Inner filter effect-based upconversion nanosensor for rapid detection of thiram pesticides using upconversion nanoparticles and dithizone–cadmium complexes. Food Chem. 2024, 434, 137438. [Google Scholar] [CrossRef]
  240. Li, H.; Bei, Q.; Hassan, M.M.; Marimuthu, M.; Adade, S.Y.S.S.; Chen, Q.; Zareef, M. A Cu2+-modulated UCNPs@RBD sensor for sensitive detection of tetracyclines in food based on the spirolactam open-loop reaction. J. Food Compos. Anal. 2024, 133, 106374. [Google Scholar] [CrossRef]
  241. Liu, Y.; Ouyang, Q.; Li, H.; Chen, M.; Zhang, Z.; Chen, Q. Turn-On Fluoresence Sensor for Hg2+ in Food Based on FRET between Aptamers-Functionalized Upconversion Nanoparticles and Gold Nanoparticles. J. Agric. Food Chem. 2018, 66, 6188–6195. [Google Scholar] [CrossRef] [PubMed]
  242. Sheng, Y.; Liao, L.-D.; Thakor, N.; Tan, M.C. Rare-earth doped particles as dual-modality contrast agent for minimally-invasive luminescence and dual-wavelength photoacoustic imaging. Sci. Rep. 2014, 4, 6562. [Google Scholar] [CrossRef]
  243. Liu, Y.; Kang, N.; Lv, J.; Zhou, Z.; Zhao, Q.; Ma, L.; Chen, Z.; Ren, L.; Nie, L. Deep photoacoustic/luminescence/magnetic resonance multimodal imaging in living subjects using high-efficiency upconversion nanocomposites. Adv. Mater. 2016, 28, 6411–6419. [Google Scholar] [CrossRef] [PubMed]
  244. Wang, Z.; Ai, X.; Zhang, Z.; Wang, Y.; Wu, X.; Haindl, R.; Yeow, E.K.; Drexler, W.; Gao, M.; Xing, B. NIR nanoprobe-facilitated cross-referencing manifestation of local disease biology for dynamic therapeutic response assessment. Chem. Sci. 2020, 11, 803–811. [Google Scholar] [CrossRef]
  245. Zhuang, H.; Li, B.; Zhao, M.; Wei, P.; Yuan, W.; Zhang, M.; Han, X.; Chen, Y.; Yi, T. Real-time monitoring and accurate diagnosis of drug-induced hepatotoxicity in vivo by ratio-fluorescence and photoacoustic imaging of peroxynitrite. Nanoscale 2020, 12, 10216–10225. [Google Scholar] [CrossRef]
  246. Lv, R.; Wang, D.; Xiao, L.; Chen, G.; Xia, J.; Prasad, P.N. Stable ICG-loaded upconversion nanoparticles: Silica core/shell theranostic nanoplatform for dual-modal upconversion and photoacoustic imaging together with photothermal therapy. Sci. Rep. 2017, 7, 15753. [Google Scholar] [CrossRef]
  247. He, F.; Yang, G.; Yang, P.; Yu, Y.; Lv, R.; Li, C.; Dai, Y.; Gai, S.; Lin, J. A new single 808 nm NIR light-induced imaging-guided multifunctional cancer therapy platform. Adv. Funct. Mater. 2015, 25, 3966–3976. [Google Scholar] [CrossRef]
  248. Hou, Z.; Deng, K.; Wang, M.; Liu, Y.; Chang, M.; Huang, S.; Li, C.; Wei, Y.; Cheng, Z.; Han, G. Hydrogenated titanium oxide decorated upconversion nanoparticles: Facile laser modified synthesis and 808 nm near-infrared light triggered phototherapy. Chem. Mater. 2019, 31, 774–784. [Google Scholar] [CrossRef]
  249. He, S.; Song, J.; Liu, J.; Liu, L.; Qu, J.; Cheng, Z. Enhancing Photoacoustic Intensity of Upconversion Nanoparticles by Photoswitchable Azobenzene-Containing Polymers for Dual NIR-II and Photoacoustic Imaging In Vivo. Adv. Opt. Mater. 2019, 7, 1900045. [Google Scholar] [CrossRef]
  250. Du, K.; Lei, P.; Dong, L.; Zhang, M.; Gao, X.; Yao, S.; Feng, J.; Zhang, H. In situ decorating of ultrasmall Ag2Se on upconversion nanoparticles as novel nanotheranostic agent for multimodal imaging-guided cancer photothermal therapy. Appl. Mater. Today 2020, 18, 100497. [Google Scholar] [CrossRef]
  251. He, F.; Feng, L.; Yang, P.; Liu, B.; Gai, S.; Yang, G.; Dai, Y.; Lin, J. Enhanced up/down-conversion luminescence and heat: Simultaneously achieving in one single core-shell structure for multimodal imaging guided therapy. Biomaterials 2016, 105, 77–88. [Google Scholar] [CrossRef]
  252. Sun, M.; Xu, L.; Ma, W.; Wu, X.; Kuang, H.; Wang, L.; Xu, C. Hierarchical plasmonic nanorods and upconversion core–satellite nanoassemblies for multimodal imaging-guided combination phototherapy. Adv. Mater. 2016, 28, 898–904. [Google Scholar] [CrossRef]
  253. Sheng, Y.; Liao, L.-D.; Bandla, A.; Liu, Y.-H.; Thakor, N.; Tan, M.C. Size and shell effects on the photoacoustic and luminescence properties of dual modal rare-earth-doped nanoparticles for infrared photoacoustic imaging. ACS Biomater. Sci. Eng. 2016, 2, 809–817. [Google Scholar] [CrossRef] [PubMed]
  254. Maji, S.K.; Sreejith, S.; Joseph, J.; Lin, M.; He, T.; Tong, Y.; Sun, H.; Yu, S.W.-K.; Zhao, Y. Upconversion Nanoparticles as a Contrast Agent for Photoacoustic Imaging in Live Mice. Adv. Mater. 2014, 26, 5633–5638. [Google Scholar] [CrossRef] [PubMed]
  255. Rieffel, J.; Chen, F.; Kim, J.; Chen, G.; Shao, W.; Shao, S.; Chitgupi, U.; Hernandez, R.; Graves, S.A.; Nickles, R.J. Hexamodal imaging with porphyrin-phospholipid-coated upconversion nanoparticles. Adv. Mater. 2015, 27, 1785–1790. [Google Scholar] [CrossRef]
  256. Xu, Y.; Liang, H.; Zeng, Q.; He, F.; Liu, C.; Gai, S.; Ding, H.; Yang, P. A bubble-enhanced lanthanide-doped up/down-conversion platform with tumor microenvironment response for dual-modal photoacoustic and near-infrared-II fluorescence imaging. J. Colloid Interface Sci. 2024, 659, 149–159. [Google Scholar] [CrossRef] [PubMed]
  257. Bastos, V.; Oskoei, P.; Andresen, E.; Saleh, M.I.; Rühle, B.; Resch-Genger, U.; Oliveira, H. Stability, dissolution, and cytotoxicity of NaYF4-upconversion nanoparticles with different coatings. Sci. Rep. 2022, 12, 3770. [Google Scholar] [CrossRef]
  258. Saleh, M.I.; Rühle, B.; Wang, S.; Radnik, J.; You, Y.; Resch-Genger, U. Assessing the protective effects of different surface coatings on NaYF4:Yb3+, Er3+ upconverting nanoparticles in buffer and DMEM. Sci. Rep. 2020, 10, 19318. [Google Scholar] [CrossRef]
  259. Liu, C.; Gao, Z.; Zeng, J.; Hou, Y.; Fang, F.; Li, Y.; Qiao, R.; Shen, L.; Lei, H.; Yang, W.; et al. Magnetic/Upconversion Fluorescent NaGdF4:Yb,Er Nanoparticle-Based Dual-Modal Molecular Probes for Imaging Tiny Tumors In Vivo. ACS Nano 2013, 7, 7227–7240. [Google Scholar] [CrossRef]
  260. Sun, Y.; Feng, W.; Yang, P.; Huang, C.; Li, F. The biosafety of lanthanide upconversion nanomaterials. Chem. Soc. Rev. 2015, 44, 1509–1525. [Google Scholar] [CrossRef] [PubMed]
  261. Abdul Jalil, R.; Zhang, Y. Biocompatibility of silica coated NaYF4 upconversion fluorescent nanocrystals. Biomaterials 2008, 29, 4122–4128. [Google Scholar] [CrossRef] [PubMed]
  262. Wang, S.; Lin, J.; Wang, T.; Chen, X.; Huang, P. Recent Advances in Photoacoustic Imaging for Deep-Tissue Biomedical Applications. Theranostics 2016, 6, 2394–2413. [Google Scholar] [CrossRef]
  263. Li, H.; Liu, H.; Wong, K.-L.; All, A.H. Lanthanide-doped upconversion nanoparticles as nanoprobes for bioimaging. Biomater. Sci. 2024, 12, 4650–4663. [Google Scholar] [CrossRef] [PubMed]
  264. Wu, S.; Butt, H.J. Near-infrared-sensitive materials based on upconverting nanoparticles. Adv. Mater. 2016, 28, 1208–1226. [Google Scholar] [CrossRef]
  265. Wang, P.; Li, H.; Hassan, M.M.; Guo, Z.; Zhang, Z.-Z.; Chen, Q. Fabricating an Acetylcholinesterase Modulated UCNPs-Cu2+ Fluorescence Biosensor for Ultrasensitive Detection of Organophosphorus Pesticides-Diazinon in Food. J. Agric. Food Chem. 2019, 67, 4071–4079. [Google Scholar] [CrossRef]
  266. Lee, T.; Baac, H.W.; Li, Q.; Guo, L.J. Efficient photoacoustic conversion in optical nanomaterials and composites. Adv. Opt. Mater. 2018, 6, 1800491. [Google Scholar] [CrossRef]
  267. Ai, X.; Wang, Z.; Cheong, H.; Wang, Y.; Zhang, R.; Lin, J.; Zheng, Y.; Gao, M.; Xing, B. Multispectral optoacoustic imaging of dynamic redox correlation and pathophysiological progression utilizing upconversion nanoprobes. Nat. Commun. 2019, 10, 1087. [Google Scholar] [CrossRef]
  268. Yang, Y.; Huang, J.; Wei, W.; Zeng, Q.; Li, X.; Xing, D.; Zhou, B.; Zhang, T. Switching the NIR upconversion of nanoparticles for the orthogonal activation of photoacoustic imaging and phototherapy. Nat. Commun. 2022, 13, 3149. [Google Scholar] [CrossRef]
  269. Kim, D.W.; Wrede, P.; Rodríguez-Camargo, A.; Chen, Y.; Dogan, N.O.; Glück, C.; Lotsch, B.V.; Razansky, D.; Sitti, M. Upconversion Nanoparticle-Covalent Organic Framework Core–shell Particles as Therapeutic Microrobots Trackable with Optoacoustic Imaging. Adv. Mater. 2025, 2418425. [Google Scholar] [CrossRef]
  270. Rizvi, S.F.A.; Zhang, H.; Fang, Q. Engineering peptide drug therapeutics through chemical conjugation and implication in clinics. Med. Res. Rev. 2024, 44, 2420–2471. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Timeline of photoacoustic imaging development. Tracing back to Alexander Graham Bell’s early discovery of photoacoustic principles to its eventual application in clinical settings, the technology has evolved over several decades [13,34,35,36,37,38,39,40,41,42,43,44,45].
Figure 1. Timeline of photoacoustic imaging development. Tracing back to Alexander Graham Bell’s early discovery of photoacoustic principles to its eventual application in clinical settings, the technology has evolved over several decades [13,34,35,36,37,38,39,40,41,42,43,44,45].
Nanomaterials 15 01074 g001
Figure 2. Scheme of photoacoustic imaging technology in precise disease detection and diagnosis.
Figure 2. Scheme of photoacoustic imaging technology in precise disease detection and diagnosis.
Nanomaterials 15 01074 g002
Figure 3. Fundamental mechanism of photoacoustic imaging. Absorbed light elevates local temperature, triggering thermoelastic expansion that generates US pressure waves through adjacent tissue. These waves are captured by US transducers, and image reconstruction is performed using their intensity and arrival time. Adapted with permission from Ref. [126].
Figure 3. Fundamental mechanism of photoacoustic imaging. Absorbed light elevates local temperature, triggering thermoelastic expansion that generates US pressure waves through adjacent tissue. These waves are captured by US transducers, and image reconstruction is performed using their intensity and arrival time. Adapted with permission from Ref. [126].
Nanomaterials 15 01074 g003
Figure 6. Commonly used methods of functionalizing UCNPs. Reproduced with permission from Ref. [187].
Figure 6. Commonly used methods of functionalizing UCNPs. Reproduced with permission from Ref. [187].
Nanomaterials 15 01074 g006
Figure 7. (A) The basic composition of UCNPs and (B) schematic illustration of the mechanism of organic dye-sensitized UCNPs. Reproduced with permission from Ref. [194].
Figure 7. (A) The basic composition of UCNPs and (B) schematic illustration of the mechanism of organic dye-sensitized UCNPs. Reproduced with permission from Ref. [194].
Nanomaterials 15 01074 g007
Figure 8. Principle of UC processes for lanthanide-doped UCNPs: (A) excited-state absorption (ESA), (B) energy transfer upconversion (ETU), (C) cooperative sensitization upconversion (SCU), (D) cross-relaxation (CR), and (E) photon avalanche (PAv). The red, violet, and green lines represent photon excitation, energy transfer, and emission processes, respectively.
Figure 8. Principle of UC processes for lanthanide-doped UCNPs: (A) excited-state absorption (ESA), (B) energy transfer upconversion (ETU), (C) cooperative sensitization upconversion (SCU), (D) cross-relaxation (CR), and (E) photon avalanche (PAv). The red, violet, and green lines represent photon excitation, energy transfer, and emission processes, respectively.
Nanomaterials 15 01074 g008
Figure 9. Strategies for designing lanthanide nanoparticles as photoacoustic diagnostic markers. (A) Diagram showing fluorescence damping followed by photoacoustic signal amplification from UCNP/α-CD dispersed in H2O. High-definition acoustic responses and phantom images were captured using nanosecond pulsed laser excitation at 980 nm. Reprinted with permission from Ref. [254]. PA images at the neoplasm region for mice before (B) and after (C) intratumoral injection of Nd: UCNPs@H-TiO2 nanocomposites, respectively. Reproduced with permission from Ref. [248]. (D) Mouse brain imaging was conducted following injection of UCNP@SiO2-ICG using 800 nm excitation with both pulsed- and continuous-wave lasers (n = 3). SSS refers to the superior sagittal sinus. (E) The in vivo photoacoustic intensity versus the intravenous injection time. Reproduced with permission from Ref. [246]. (F) Impact of nanoparticle shell thickness and size on UCNPs’ PA signals during NIR imaging. (G) Photoacoustic assessments were conducted with a dark-field PAM instrument with 50 MHz central frequency and confocal detection with approximately 4 ns laser pulses (975 nm) at a 10 Hz pulsing frequency. Reused with permission from Ref. [253].
Figure 9. Strategies for designing lanthanide nanoparticles as photoacoustic diagnostic markers. (A) Diagram showing fluorescence damping followed by photoacoustic signal amplification from UCNP/α-CD dispersed in H2O. High-definition acoustic responses and phantom images were captured using nanosecond pulsed laser excitation at 980 nm. Reprinted with permission from Ref. [254]. PA images at the neoplasm region for mice before (B) and after (C) intratumoral injection of Nd: UCNPs@H-TiO2 nanocomposites, respectively. Reproduced with permission from Ref. [248]. (D) Mouse brain imaging was conducted following injection of UCNP@SiO2-ICG using 800 nm excitation with both pulsed- and continuous-wave lasers (n = 3). SSS refers to the superior sagittal sinus. (E) The in vivo photoacoustic intensity versus the intravenous injection time. Reproduced with permission from Ref. [246]. (F) Impact of nanoparticle shell thickness and size on UCNPs’ PA signals during NIR imaging. (G) Photoacoustic assessments were conducted with a dark-field PAM instrument with 50 MHz central frequency and confocal detection with approximately 4 ns laser pulses (975 nm) at a 10 Hz pulsing frequency. Reused with permission from Ref. [253].
Nanomaterials 15 01074 g009
Figure 10. Schematic representation of UCNP-based responsive nanoprobes tailored for photoacoustic bioimaging applications. (A) UCNP–dye composite nanoprobes designed for monitoring dynamic correlations of reactive oxygen species (ROS) and reactive nitrogen species (RNS) in liver pathology. (B) Multispectral optoacoustic tomography (MSOT) images at 680 nm and 800 nm showing regions of interest (ROIs) in liver cross-sections with pseudo-color mapping after nanoprobe injection in models treated with lipopolysaccharide (LPS) and acetaminophen (APA). Reproduced with permission from Ref. [267].
Figure 10. Schematic representation of UCNP-based responsive nanoprobes tailored for photoacoustic bioimaging applications. (A) UCNP–dye composite nanoprobes designed for monitoring dynamic correlations of reactive oxygen species (ROS) and reactive nitrogen species (RNS) in liver pathology. (B) Multispectral optoacoustic tomography (MSOT) images at 680 nm and 800 nm showing regions of interest (ROIs) in liver cross-sections with pseudo-color mapping after nanoprobe injection in models treated with lipopolysaccharide (LPS) and acetaminophen (APA). Reproduced with permission from Ref. [267].
Nanomaterials 15 01074 g010
Figure 11. Multimodal imaging performance of UCNP@CS@Ag2Se nanocomposites in vitro and in vivo. (A) PA signal intensity of UCNP@CS@Ag2Se nanocomposites at varying concentrations in vitro, (B) PA signal response showing concentration dependence, and (C) time-course PA imaging of pre- and post-injection assessment in mice with tumors of UCNP@CS@Ag2Se nanocomposites. Reproduced with permission from Ref. [250].
Figure 11. Multimodal imaging performance of UCNP@CS@Ag2Se nanocomposites in vitro and in vivo. (A) PA signal intensity of UCNP@CS@Ag2Se nanocomposites at varying concentrations in vitro, (B) PA signal response showing concentration dependence, and (C) time-course PA imaging of pre- and post-injection assessment in mice with tumors of UCNP@CS@Ag2Se nanocomposites. Reproduced with permission from Ref. [250].
Nanomaterials 15 01074 g011
Figure 12. (A) Schematic of deep-tissue imaging utilizing high-performance multi-shell UCNPs for in vivo PAI, UCL, and MRI. The layered structure enhances PAI penetration depth, supporting applications such as early tumor detection, drug tracking, and surgical navigation. (B) Photoacoustic images showing tumor and vasculature visualization in mice post-intravenous nanoprobe injection. Reproduced with permission from Ref. [243].
Figure 12. (A) Schematic of deep-tissue imaging utilizing high-performance multi-shell UCNPs for in vivo PAI, UCL, and MRI. The layered structure enhances PAI penetration depth, supporting applications such as early tumor detection, drug tracking, and surgical navigation. (B) Photoacoustic images showing tumor and vasculature visualization in mice post-intravenous nanoprobe injection. Reproduced with permission from Ref. [243].
Nanomaterials 15 01074 g012
Figure 13. (A) PA imaging and related signal intensities of UCNPs and UCNPs@PAA-Azo across a series of concentrations (0.04, 0.2, 1.0, and 5.0 mg/mL), demonstrating a concentration-dependent enhancement. (B) NIR-II fluorescence (1350 nm, 300 ms exposure) and signal comparison between UCNPs and UCNPs@PAA-Azo in vials. (C) Dual-modality imaging of lymph nodes (LNs) in mice using UCNPs@PAA-Azo: photograph showing NIR-II (red) and PA (blue) imaging zones; dynamic PAI and NIR-II imaging of sacral (SC) and popliteal (PO) LNs post-injection (0, 10, 30, 60, 120 min); guided surgical removal of LNs under NIR-II imaging. Reproduced with permission from Ref. [249].
Figure 13. (A) PA imaging and related signal intensities of UCNPs and UCNPs@PAA-Azo across a series of concentrations (0.04, 0.2, 1.0, and 5.0 mg/mL), demonstrating a concentration-dependent enhancement. (B) NIR-II fluorescence (1350 nm, 300 ms exposure) and signal comparison between UCNPs and UCNPs@PAA-Azo in vials. (C) Dual-modality imaging of lymph nodes (LNs) in mice using UCNPs@PAA-Azo: photograph showing NIR-II (red) and PA (blue) imaging zones; dynamic PAI and NIR-II imaging of sacral (SC) and popliteal (PO) LNs post-injection (0, 10, 30, 60, 120 min); guided surgical removal of LNs under NIR-II imaging. Reproduced with permission from Ref. [249].
Nanomaterials 15 01074 g013
Figure 14. Multi-platform lanthanide nanoprobes for combined integrated luminescence and photoacoustic approaches to liver pathology visualization. (A) Diagram of UCL/PA dual-modality probes developed for concurrent tracking of liver disease progression and therapy evaluation. (B,C) In vivo ratiometric imaging via UCL and PA techniques to monitor liver injury caused by medication and antioxidant treatment outcomes. Reproduced with permission from Ref. [244].
Figure 14. Multi-platform lanthanide nanoprobes for combined integrated luminescence and photoacoustic approaches to liver pathology visualization. (A) Diagram of UCL/PA dual-modality probes developed for concurrent tracking of liver disease progression and therapy evaluation. (B,C) In vivo ratiometric imaging via UCL and PA techniques to monitor liver injury caused by medication and antioxidant treatment outcomes. Reproduced with permission from Ref. [244].
Nanomaterials 15 01074 g014
Figure 15. Design and evaluation of orthogonal UCNPs-DI nanoagent. (A) Illustration of UCNPs-DI enabling photoacoustic imaging-guided, on-demand therapy. (B) Intracellular 1O2 generation in SOS G-stained MCF7 cells under various treatments. Scale bar: 50 μm. (C) Confocal fluorescence images of MCF7 cells co-stained with calcein AM (green, live) and propidium iodide (PI, red, dead) following different treatments. Scale bar: 100 μm. (D) Flow cytometric evaluation of MCF7 cell viability under different treatment conditions. Reprinted with permission from Ref. [268].
Figure 15. Design and evaluation of orthogonal UCNPs-DI nanoagent. (A) Illustration of UCNPs-DI enabling photoacoustic imaging-guided, on-demand therapy. (B) Intracellular 1O2 generation in SOS G-stained MCF7 cells under various treatments. Scale bar: 50 μm. (C) Confocal fluorescence images of MCF7 cells co-stained with calcein AM (green, live) and propidium iodide (PI, red, dead) following different treatments. Scale bar: 100 μm. (D) Flow cytometric evaluation of MCF7 cell viability under different treatment conditions. Reprinted with permission from Ref. [268].
Nanomaterials 15 01074 g015
Figure 16. In vivo multimodal imaging. (A) UCL imaging, (B) photoacoustic imaging of HeLa tumor-bearing mice captured at various time intervals after i.v. injection with nanorod dimer-UCNP-Ce6 (200 µL 2 mg·mL−1, in terms of the nanorod amount). The colors in the diagram represent the strength of the signal. Reprinted with permission from Ref. [252].
Figure 16. In vivo multimodal imaging. (A) UCL imaging, (B) photoacoustic imaging of HeLa tumor-bearing mice captured at various time intervals after i.v. injection with nanorod dimer-UCNP-Ce6 (200 µL 2 mg·mL−1, in terms of the nanorod amount). The colors in the diagram represent the strength of the signal. Reprinted with permission from Ref. [252].
Nanomaterials 15 01074 g016
Figure 17. (A,B) PA signal intensity of UCNPs@MS-Au25-PEG decreases with dilution in distilled water. (C) Diagram showing the structure of the designed LDNPs-5. (D) In vivo PA imaging of tumor sites at different times after intravenous injection of LDNPs-5-Au25-PEG. Reproduced with permission from Ref. [247] and Ref. [251].
Figure 17. (A,B) PA signal intensity of UCNPs@MS-Au25-PEG decreases with dilution in distilled water. (C) Diagram showing the structure of the designed LDNPs-5. (D) In vivo PA imaging of tumor sites at different times after intravenous injection of LDNPs-5-Au25-PEG. Reproduced with permission from Ref. [247] and Ref. [251].
Nanomaterials 15 01074 g017
Table 1. Main synthetic methods for UCNP preparation and their advantages and disadvantages.
Table 1. Main synthetic methods for UCNP preparation and their advantages and disadvantages.
Main Synthetic MethodMaterial, RefsSize Range (nm)AdvantagesDisadvantages
Thermal decompositionNaNdF4 [180]
β-NaErF4 [181]
50–500High-quality, uniform sizeIntermediate toxicity, high cost
MicroemulsionNaYF4 [182]
LaF3 [183]
4–500Easy to operate, narrow size distribution, high stabilityCalcination or annealing usually required
Phase transfer hydrothermal synthesis (La-Dy)VO4 [184]
YVO4 [184]
NaYF4 [185]
10–1000Good dispersion, simple procedures, tunable sizeSpecialized reaction vessels are needed
Sol-gel processingGdVO4 [186]30–600Cheap raw materials, simple proceduresBroad particle size and unsuitable for bioapplication
Table 2. The doping ratio of different ions of UCNPs as contrast agents working in PAI and their major emission wavelengths.
Table 2. The doping ratio of different ions of UCNPs as contrast agents working in PAI and their major emission wavelengths.
Dopant Ions and CompositionMajor λ Emissions (nm) Reference
Host LatticeSensitizerActivatorShellBlueGreenRed
β-NaYF420%Yb3+2%Er3+ 450, 476520654Ref. [208]
20%Yb3+0.2%Tm3+ 540 Ref. [209]
20%Yb3+2%Ho3+ 541 Ref. [209]
β-NaYF420%Yb3+0.3%Tm3+20%Yb3+, 2%Er3+450520653Ref. [210]
475540
Li+ doped β-NaYF4 20%Yb3+0.5%Tm3+ 452, 479 650Ref. [211]
Mn+ doped β-NaYF4 20%Yb3+2%Er3+ 657Ref. [212]
β-NaYF418.6%Yb3+2.2%Er3+TRITC-SiO2407 Ref. [213]
25%Yb3+0.3%Tm3+SiO2450, 479521,
539
651Ref. [85]
25%Yb3+0.3%Tm3+FITC-SiO2450, 479521,
539,
580
651Ref. [214]
α-NaYF425%Yb3+0.3%Tm3+QD-SiO2450, 479540651Ref. [214]
20%Yb3+2%Er3+ 411 660Ref. [202]
20%Yb3+0.2%Tm3+ 450540644Ref. [215]
20%Yb3+0.2%Er3+ 475525693Ref. [215]
Table 3. Summary of representative UCNPs applied for photoacoustic imaging.
Table 3. Summary of representative UCNPs applied for photoacoustic imaging.
Photoacoustic Contrast AgentExcitation λ (nm)PA SensitivitySize (nm)Multimodal Imaging CapabilityIn Vivo/In VitroReference
PAA-NaYF4:Yb,Er520, 975-40–60PAI In vivoRef. [242]
NaYF4:Yb:Er@NaYF4:Yb@NaNdF4:Yb@NaYF4 @NaGdF4-HAD-G2 ICG (CS2-ICG)808/540, 650ICG54.3 ± 4.1 MRI, PA, UCLbothRef. [243]
CyB/NaGdF4:Yb/Tm/Er@NaGdF4980/650, 800Isoniazid (INH)~78PAI, UCLIn vivoRef. [244]
UCY7680ONOO~35PAI, RFLIIn vivoRef. [245]
UCNP@mSiO2-ICG800ICG~60PA, UCLbothRef. [246]
UCNPs@MS-Au25
-PEG
808 28–46PA, MRI, CT, PDT, PTTbothRef. [247]
Nd:UCNPs@H-TiO2808H-TiO2~60PA, PDT, PTT, UCLbothRef. [248]
UCNP@PAA-AzoNaYF4:Yb30%, Tm0.5%, Nd5%808Azobenzene-containing poly~21PA, NIR-IIbothRef. [249]
UCNPs@CS#Ag2Se808Ag+141.8UCL, CT, PAI, PTTbothRef. [250]
LDNPs-Au25-PEG808 ~40PTT, PA, PTI, PDT, MRIbothRef. [251]
Nanorod (NR) dimer-UCNP-Ce6980, 808NR-dimer~70PA, UCL, CT, MRIIn vivoRef. [252]
Core/Shell β-NaYF4:Yb,Er975 50.3 ± 5.6 PAT bothRef. [253]
NaYF4:Yb/Er@
α-cyclodextrin
980α-CD PAIIn vivoRef. [254]
PoP-UCNP980PoP combined with detergent74 ± 3.6PA, FL, UC, PET, CL, CTIn vivoRef. [255]
LDNP@DMSN-Au@CaCO3980Au-NPs~174PA, NIR-II, FLbothRef. [256]
FL, fluorescence; UC, NIR-to-NIR upconversion (UC) luminescence; PET, positron emission tomography; CL, Cerenkov luminescence; CT, computed tomography; RFLI, ratio-fluorescent imaging; NIR-II, second near-infrared; MRI, magnetic resonance imaging; PDT, photodynamic therapy; PTT, photothermal therapy; UCL, upconversion luminescence; PAT, photoacoustic tomography; PTI, photothermal imaging.
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

Zhang, Y.; Li, Z.; Du, Z.; Pan, J.; Huang, Y. Multifunctional Upconversion Nanoparticles Transforming Photoacoustic Imaging: A Review. Nanomaterials 2025, 15, 1074. https://doi.org/10.3390/nano15141074

AMA Style

Zhang Y, Li Z, Du Z, Pan J, Huang Y. Multifunctional Upconversion Nanoparticles Transforming Photoacoustic Imaging: A Review. Nanomaterials. 2025; 15(14):1074. https://doi.org/10.3390/nano15141074

Chicago/Turabian Style

Zhang, Yuqian, Zerui Li, Ziqing Du, Jianming Pan, and Yanan Huang. 2025. "Multifunctional Upconversion Nanoparticles Transforming Photoacoustic Imaging: A Review" Nanomaterials 15, no. 14: 1074. https://doi.org/10.3390/nano15141074

APA Style

Zhang, Y., Li, Z., Du, Z., Pan, J., & Huang, Y. (2025). Multifunctional Upconversion Nanoparticles Transforming Photoacoustic Imaging: A Review. Nanomaterials, 15(14), 1074. https://doi.org/10.3390/nano15141074

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