Next Article in Journal
In Situ Synthesis of RMB6-TMB2 Composite Nanopowders via One-Step Solid-State Reduction
Next Article in Special Issue
Review of Current Achievements in Dendrimers and Nanomaterials for Potential Detection and Remediation of Chemical, Biological, Radiological and Nuclear Contamination—Integration with Artificial Intelligence and Remote Sensing Technologies
Previous Article in Journal
A First-Principles Investigation of the Structural, Electronic, Optical, and Mechanical Properties of Hydrogen Storage Ordered Vacancy Double Perovskite X2MH6 Materials
Previous Article in Special Issue
Special Issue “Advanced Nanomaterials for Bioimaging”
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Current Status and Future Aspects of Gadolinium Oxide Nanoparticles as Positive Magnetic Resonance Imaging Contrast Agents

1
Department of Chemistry, College of Natural Sciences, Kyungpook National University, Taegu 41566, Republic of Korea
2
Institute of Biomedical Engineering Research, Kyungpook National University, Taegu 41944, Republic of Korea
3
Department of Chemistry Education, Teachers’ College, Kyungpook National University, Taegu 41566, Republic of Korea
4
Department of Molecular Medicine, School of Medicine, Kyungpook National University, Taegu 41944, Republic of Korea
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2025, 15(17), 1340; https://doi.org/10.3390/nano15171340
Submission received: 30 July 2025 / Revised: 25 August 2025 / Accepted: 29 August 2025 / Published: 1 September 2025
(This article belongs to the Special Issue Advanced Nanomaterials for Bioimaging: 2nd Edition)

Abstract

Although numerous studies have investigated gadolinium oxide (Gd2O3) nanoparticles (NPs) as positive (T1) magnetic resonance imaging (MRI) contrast agents (CAs), comprehensive reviews on this topic remain scarce. Therefore, it is essential to evaluate their current status and outline prospects. Despite promising physicochemical properties such as considerably higher relaxivities compared to 3–5 s−1mM−1 of clinically approved Gd(III)-chelate contrast agents and encouraging results from in vivo animal studies such as highly improved contrast enhancements, drug loading, and tumor targeting, extensive in vivo toxicity assessments including long-term toxicity and formulation advancements suitable for renal excretion (d < ~3 nm) are still required for clinical translation. This review summarizes the synthesis, characterization, in vitro and in vivo toxicity, and in vivo MRI applications of surface-modified Gd2O3 NPs as T1 MRI CAs. Finally, future perspectives on the development of surface-modified Gd2O3 NPs as potential next-generation T1 MRI CAs are discussed.

Graphical Abstract

1. Introduction

Molecular imaging plays a critical role in medical treatments by providing detailed anatomical information through high-resolution imaging [1,2]. Various imaging modalities have been developed, including fluorescence imaging (FI), magnetic resonance imaging (MRI), X-ray computed tomography (CT), positron emission tomography (PET), ultrasound imaging (USI), and single-photon emission computed tomography (SPECT) [3,4]. These techniques enable the detection, diagnosis, and monitoring of diseases. Among them, MRI is particularly valued for its non-invasive nature as it employs non-ionizing radiofrequency radiation and allows for whole-body imaging with unlimited penetration depth in soft tissues, making it a widely adopted modality [5]. However, MRI suffers from relatively low sensitivity due to the small population difference between the two proton spin states [6,7], which limits its ability to detect small or early-stage lesions. This limitation can be mitigated by the use of contrast agents (CAs), which enhance signal intensity by accelerating proton spin relaxation processes, thereby improving detection sensitivity [8,9].
Various types of MRI CAs have been investigated, including Gd(III)- and Mn(II)-chelates [10,11], Gd- [12] and Mn-containing nanoparticles (NPs) [13] for positive (T1) MRI, and superparamagnetic iron oxide nanoparticles (SPIONs) [14,15] for negative (T2) MRI. Currently, only Gd(III)-chelates are approved for clinical use as T1 MRI CAs [16,17]. Although SPIONs were previously commercialized, their use is now minimal [18]. Gd3+, with seven unpaired 4f-electrons and a high spin magnetic moment (s = 7/2), is ideal for use in T1 MRI CAs [19]. Consequently, Gd2O3 NPs can induce stronger T1 proton spin relaxations than Mn-and Fe-based NPs owing to lower electron spin magnetic moments of Mn2+ (s = 5/2) and Fe3+ (s = 5/2) than Gd3+ [16,20,21,22,23,24,25]. For example, this can be noticed from the r1 value (33.4 s−1mM−1) of free Gd3+ ions-grafted NDs [26], which is higher than that (13.5 s−1mM−1) of free Mn2+ ions-grafted NDs [27]. However, after PVP coating, r1 value of free Gd3+ ions-grafted NDs dropped to 15.9 s−1mM−1 [28]. T1 MRI CAs are generally preferred over T2 MRI CAs because signal enhancement is easier to detect than signal loss [29]. Moreover, SPIONs can cause signal darkening and susceptibility-induced artifacts, which may complicate image interpretation [30,31,32,33]. In contrast, Gd2O3 NPs produce the high-quality positive contrast images with less artificial signal voids, thereby offering clearer diagnostic outcomes [34].
However, Gd3+ ions are inherently toxic and non-biodegradable through metabolic processes [35]. To mitigate this toxicity, they are commonly administered as Gd(III)-chelates [35,36,37,38,39]. Although Gd(III)-chelates effectively enhance positive contrasts in MRI, they exhibit low sensitivity due to their low longitudinal proton spin relaxivity (r1) values and lack of specificity. Moreover, they have been associated with serious adverse effects, including nephrogenic systemic fibrosis (NSF) in patients with renal impairment, a condition characterized by skin and organ thickening and darkening, ultimately impairing organ function [40,41,42]. Recent studies have also indicated that repeated administration of Gd(III)-chelates can lead to gadolinium deposition in the brain, potentially causing neurotoxicity [43,44].
With advancements in nanotechnology, Gd2O3 NPs have emerged as promising alternatives to Gd(III)-chelates, offering several advantages. These NPs exhibit a high density of Gd3+ ions per NP, resulting in higher r1 values [45,46,47,48], compared to 3–5 s−1mM−1 [17,38] of Gd(III)-chelates. They also demonstrate prolonged blood circulation times due to the enhanced permeability and retention (EPR) effect [46], enabling extended imaging and diagnosing windows [47]. Furthermore, their high surface-to-volume ratios facilitate drug loading and conjugation of targeting ligands, making them suitable for theranostic applications [48,49,50]. However, their drawback is potential toxicity, which needs to be clearly and thoroughly investigated for clinical translations. For biomedical applications, Gd2O3 NPs must be surface-modified with hydrophilic and biocompatible ligands to ensure colloidal stability, biocompatibility, and functionalization potential. Various synthesis methods have been reported, including polyol [51,52], dimethyl sulfoxide (DMSO) [53,54], thermal decomposition [55], and hydrothermal methods [56]. Surface modifications have employed a range of biocompatible and hydrophilic ligands such as polyacrylic acid (PAA) [51,57], poly(methyl vinyl ether-alt-maleic acid) (PMVEMA) [58,59,60], poly(acrylic acid-co-maleic acid) (PAAMA) [61,62], polyaspartic acid (PASA) [63], D-glucuronic acid [64], citric acid (CA) [65], dextran [66], polyethylene glycol (PEG) [67], or polyvinylpyrrolidone (PVP) [68]. These ligand coatings are critical not only for stability and functionality but also for reducing toxicity by preventing the release of free Gd3+ ions from the NPs.
To date, few comprehensive reviews have focused specifically on surface-coated Gd2O3 NPs as T1 MRI CAs. However, as illustrated in Figure 1, a substantial number of research articles have been reported on surface-coated Gd2O3 NPs as T1 MRI CAs and their biomedical applications. Therefore, it is both timely and necessary to evaluate the current status of this research and to discuss future directions. As shown in Figure 1, the annual growth rate of publications increased up to 2019 and then decreased afterwards. This trend reflects the increase in the development of various types of gadolinium NP formulations, in vivo MRI, and biotoxicity studies up to 2019, and then the decrease afterwards. This review covers key aspects, including synthesis methods, physicochemical characterization, in vitro and in vivo toxicity assessments, and in vivo MRI applications. Finally, future perspectives for clinical translation of surface-coated Gd2O3 NPs are outlined.

2. Synthesis of Surface-Modified Gd2O3 NPs

To date, various synthetic approaches have been employed to prepare surface-modified Gd2O3 NPs for biomedical applications. These include polyol [51,52], DMSO [53], thermal decomposition [55], and hydrothermal methods [56]. As summarized in Table 1, each method presents distinct advantages and limitations in terms of synthesis complexity, scalability, and environmental impact.

2.1. Polyol Method

The polyol method is effective for synthesizing ultrasmall Gd2O3 NPs, with average particle diameters of approximately 2.0 nm, via a one-pot process that simultaneously enables surface modification with hydrophilic and biocompatible ligands [34,51,58,61,63,64,69,70,71,72,73,74,75]. High boiling point polyols, such as ethylene glycol (EG), diethylene glycol (DEG), triethylene glycol (TEG), and polyethylene glycol (PEG), are commonly used as solvents. These polyols also function as stabilizing agents, preventing particle aggregation. A general scheme for the one-pot polyol synthesis of surface-coated Gd2O3 NPs is displayed in Figure 2a. The process involves a Gd3+ precursor, surface-coating ligands, NaOH, and a polyol solvent, all reacted under atmospheric conditions with magnetic stirring. Initially, the Gd3+ precursor solution is prepared by dissolving Gd3+ salt, such as GdCl3·xH2O, and ligands such as PAA in a polyol such as TEG within a three-necked round-bottom flask. This mixture is stirred magnetically at room temperature under atmospheric conditions until complete dissolution. Separately, a NaOH solution is prepared by dissolving NaOH in TEG in a beaker and is then slowly added to the precursor solution until the pH reaches 8–10. The resulting mixture is stirred at ~110 °C for ~12 h to yield ligand-coated Gd2O3 NPs.
For example, Lee’s group synthesized ultrasmall Gd2O3 NPs using various hydrophilic ligands, such as PAA, PMVEMA, PAAMA, PASA, CA, and D-glucuronic acid, via the one-pot polyol method [51,58,61,63,64]. In another study, Guleria et al. synthesized Gd2O3 NPs using polyols of varying chain lengths, including DEG, TEG, tetraethylene glycol (TeEG), and PEG-200, and observed a linear correlation between the glycol chain length and the resulting NP size [70].

2.2. DMSO Method

The DMSO method is suitable for low-temperature synthesis (<60 °C), enabling the production of ultrasmall Gd2O3 NPs with uniform particle size and simultaneous one-pot surface modification using hydrophilic and biocompatible ligands [53,76,77]. As shown in Figure 2b, Gd3+ precursor such as Gd(acetate)3·xH2O is dissolved in DMSO under atmospheric conditions with magnetic stirring until a transparent solution is obtained. Tetramethylammonium hydroxide (TMAH) solution is then added dropwise to adjust the pH to ~8, resulting in a cloudy solution. This mixture is stirred magnetically for 24 h to allow the formation of ultrasmall Gd2O3 NPs. For surface coating, a hydrophilic ligand such as PAA is added to the reaction mixture, followed by the addition of TMAH to maintain pH at ~8. The solution is stirred for another 24 h to ensure uniform ligand surface coating. The resulting ligand-coated Gd2O3 NPs are purified by repeated ethanol washing and centrifugation to remove unreacted reagents and solvents. Finally, the surface-coated NPs are dispersed in triple-distilled water and subjected to dialysis against water to eliminate residual impurities.
For example, Uvdal and co-workers synthesized Gd2O3 NPs (d = 4–5 nm) by dissolving Gd(acetate)3·xH2O in DMSO and adding TMAH in ethanol under ambient conditions with magnetic stirring [53]. In another example, Cui et al. synthesized Gd2O3 NPs using Gd(acetate)3·xH2O and TMAH in DMSO at room temperature and subsequently conjugated the NPs with carboxyfluorescein [FI]-polyethylene glycol [PEG]-bombesin (BBN), resulting in NPs with a particle diameter of 52.3 nm [77].

2.3. Thermal Decomposition

The thermal decomposition method yields monodispersed Gd2O3 NPs with controlled particle size; however, it requires high temperatures, organic solvents such as OA and OM, and an inert gas atmosphere, rendering the synthesis costly and environmentally unfriendly [55,78,79,80,81,82]. Additionally, this method does not permit one-pot synthesis for surface modification. As shown in Figure 2(cI), Gd3+ precursor is added to a mixture of OA and OM in a three-neck round-bottom flask and heated to ~110 °C under a flow of argon or nitrogen to remove moisture and dissolved gases. The reaction mixture is then refluxed at 290–340 °C for 1–6 h. After cooling to room temperature, the Gd2O3 NPs are precipitated by adding ethanol, collected by centrifugation, and redispersed in toluene: this washing procedure is repeated three times. The purified NPs are subsequently dispersed and stored in hexane or chloroform. For surface coating, the hydrophobic ligands on the NP surfaces, primarily OA, with a minor contribution from OM due to the stronger binding affinity of the -COOH group than the -NH2 group, are replaced with hydrophilic ligands such as polyvinyl pyrrolidone (PVP) through ligand exchange (Figure 2(cII)). The resulting hydrophilic ligand-coated Gd2O3 NPs are dispersed in water and further purified through dialysis against water.
For example, Fang et al. synthesized monodispersed ultrasmall Gd2O3 NPs (d = 2.9 nm) capped with hydrophobic OA, via thermal decomposition of a Gd(OA)3 precursor in a mixture of OA and OM, which acted as both solvent and surfactant [78]. The OA-capped Gd2O3 NPs were subsequently converted to PVP-Gd2O3 NPs through ligand exchange with PVP for use as T1 MRI CAs. Similarly, Cai et al. prepared OA/OM-capped Gd2O3 nanoplates (length × thickness = 10 × 1 nm) via thermal decomposition of Gd(OA)3 at 340 °C using OA and OM as solvent and capping agents. These nanoplates were encapsulated with N-dodecyl-polyethylene imine (PEI)-PEG polymers to form hydrophilic nanoplate clusters (d = 95 nm) for application as T1 MRI CAs [55].

2.4. Hydrothermal Method

The hydrothermal method is considered environmentally friendly as it utilizes water as the solvent. The synthesis is controlled by adjusting temperature and pressure [46,56,83,84,85,86]. As illustrated in Figure 2d, the Gd3+ precursor is dissolved in deionized water. Separately, a hydrophilic ligand is dissolved in deionized water at room temperature and added to the Gd3+ solution under magnetic stirring. An aqueous solution of NaOH or urea is then added dropwise under magnetic stirring to adjust the pH to 9–10. The resulting solution is transferred into an autoclave and subjected to hydrothermal treatment at elevated temperatures (typically above 150 °C) for 6–24 h. The final precipitates are washed three times with ethanol, dispersed in triple-distilled water, and further purified via dialysis against water.
For example, Wu et al. synthesized hyaluronic acid (HA)-coated Gd2O3 NPs (d = 105 nm) using the hydrothermal method [46]. In this procedure, HA was dissolved in water under vigorous magnetic stirring at ambient temperature overnight. GdCl3·6H2O and NaOH were then added, followed by magnetic stirring for 5 min to form a homogeneous clear solution. The mixture was transferred to an autoclave, sealed, and heated at 120 °C for 6 h. The resulting HA-coated Gd2O3 NPs were purified through dialysis against water to remove impurities.

3. Various Characterizations

Following synthesis, various characterization techniques can be employed to determine the physicochemical properties of surface-modified Gd2O3 NPs, as summarized in Table 2 [51,53,61,87]. These include high-resolution transmission electron microscopy (HRTEM) to determine particle size and morphology; dynamic light scattering (DLS) to measure hydrodynamic diameter; X-ray diffraction (XRD) to determine the crystal structure; Fourier transform-infrared (FT-IR) absorption spectroscopy to analyze surface coating; and thermogravimetric analysis (TGA) to estimate the surface-coating content. Additional methods include zeta potential measurements to assess the surface charge; inductively coupled plasma-atomic emission spectroscopy (ICP-AES) to quantify Gd concentration in solution; and vibrating sample magnetometry (VSM) or magnetic property measurement system (MPMS) to measure magnetic properties. MRI is used to determine the r1 and r2 relaxivities. In vitro cellular toxicity is assessed using standard assays such as methyl thiazolyl tetrazolium (MTT), cell counting kit (CCK)-8, and water-soluble tetrazolium (WST)-1 to evaluate cell viability. Collectively, these methods constitute a standard characterization protocol for surface-coated Gd2O3 NPs.

3.1. Particle Size and Morphology

Particle size and morphology are critical parameters for biomedical applications. These features are typically characterized using HRTEM operated at acceleration voltages above 200 kV. For TEM imaging, NPs are dispersed onto a copper grid coated with a carbon film by drop-casting a solution sample prepared in water or a water/ethanol mixture. High-resolution imaging also enables measurement of lattice fringes, which can aid in NP identification. In addition, elemental mapping provides the spatial distribution of constituent elements, allowing identification of both the NPs and their surface-coating ligands.
For example, Miao et al. synthesized PAA-coated ultrasmall Gd2O3 NPs (davg = 2.0 ± 0.1 nm) using a one-pot polyol method; the NPs remained colloidally stable in aqueous media without precipitation (Figure 3a) [51]. Ahrén et al. prepared Gd2O3 NPs with particle diameters of 4–5 nm via the DMSO method at room temperature (Figure 3b) [53]. These NPs exhibited good aqueous stability, with a zeta potential of +34.7 mV due to the presence of ammonium acetate on their surfaces. Cai et al. prepared OA/OM-capped Gd2O3 square nanoplates (thickness × length = 1 × 10 nm) via thermal decomposition of Gd(OA)3 at 340 °C, using OA and OM as both solvent and capping agent (Figure 3c) [55]. Yang et al. synthesized Gd(OH)3 nanorods by hydrothermal treatment of GdCl3 and triethylamine in water at 160 °C for 8 h in an autoclave [56]. Figure 3d displays the TEM image of the as-synthesized Gd(OH)3 nanorods (length × width = 200 × 10 nm).

3.2. Crystal Structure

XRD analysis provides critical information on the crystal structure of NPs and allows identification of the synthesized products. Powder samples are typically used, and the 2θ scan range is set between 10 and 100°. Diffraction peaks are indexed using Miller indices (hkl), and the unit cell parameters (a, b, c, α, β, γ) are calculated using Bragg’s law, in combination with interplanar spacing equations derived from the assigned crystal system. In addition, the average particle diameter can be roughly estimated using the Scherrer equation, which employs the full width at half maximum (FWHM) of selected XRD peaks [88]. Ultrasmall NPs (d < 3 nm) often exhibit poor crystallinity and may appear amorphous in XRD due to their limited long-range order. The pair distribution function analysis is recommended for ultrasmall NPs with amorphous XRD patterns because it offers the short-range local atomic structure of amorphous NPs [89]. However, heat treatment can induce particle growth and crystallization, allowing crystal structure determination [90].
For example, the XRD patterns corresponding to the gadolinium-based nanomaterials displayed in the TEM images in Figure 3a–d are provided in Figure 3e–h, respectively. The ultrasmall Gd2O3 NPs synthesized via the polyol and DMSO methods (TEM images in Figure 3a and Figure 3b, respectively) displayed broad, amorphous XRD patterns, indicating poor crystallinity (bottom spectra in Figure 3e,f) [51,53]. After thermal treatment, the NPs exhibited diffraction peaks corresponding to the cubic phase of Gd2O3 (top spectrum in Figure 3e) [51]. The top spectrum in Figure 3f represents the reference pattern of commercial bulk cubic Gd2O3 [53]. The Gd2O3 nanoplates synthesized via the thermal decomposition method (TEM image in Figure 3c), displayed broad peaks in the XRD pattern (top spectrum in Figure 3g) [55], which were indexed to a poorly developed monoclinic phase of Gd2O3 (JCPDS No 43-1015) [80,82,91], as shown in the reference pattern (bottom spectrum in Figure 3g). The XRD pattern of the Gd(OH)3 nanorods synthesized via the hydrothermal method (TEM image in Figure 3d) is presented in the top spectrum of Figure 3h and matches well with the reference pattern for the hexagonal phase of Gd(OH)3 (JCPDS No. 83-2037) (bottom spectrum in Figure 3h) [56].

3.3. Hydrodynamic Diameter

The hydrodynamic diameters (aavg) of NPs are measured using samples dispersed in water. The aavg values are typically larger than the corresponding davg values due to the presence of surface-coating ligands and the hydration layer formed by adsorbed water molecules. The aavg is influenced by particle size, morphology, and hydrophilicity of the surface ligands. Larger and more hydrophilic ligands attract more water molecules, thereby increasing the aavg and enhancing colloidal stability. In particular, polymer ligands tend to attract more water molecules than small molecules, resulting in larger aavg values and improved colloidal stability [92].
For example, Miao et al. synthesized ultrasmall Gd2O3 NPs (d = 2.0 nm) coated with PAA of varying molecular weight by the polyol method. They observed an increase in aavg values (10.3 → 11.1 → 11.3 nm) with increasing PAA size (1200 → 5100 → 15,000 amu) (Figure 4a) [75]. Cui et al. synthesized Gd2O3 NPs (d = 52.3 nm) grafted with carboxyfluorescein (FI)-polyethylene glycol (PEG)-bombesin (BBN) via the DMSO method [77], and reported a large aavg of 90.6 nm (Figure 4b), attributed to the combination of large particle size and bulky polymeric and biomolecular coating. Similarly, Cai et al. synthesized Gd2O3@N-dodecyl-PEI-PEG nanoplates (thickness × length = 1 × 10 nm), which exhibited a large aavg value of 95 nm (Figure 4c) owing to the thick polymer coating [55]. Siribbal et al. synthesized CA-coated hollow Gd2O3 NPs (d = 100 nm) using a hydrothermal method [93], reporting a very large aavg value of 233.7 nm, due to the large particle size and surface modification (Figure 4d).

3.4. Surface-Coating Analysis

The surface coating of surface-modified Gd2O3 NPs can be analyzed using FT-IR absorption spectroscopy and TGA [94,95]. Dried powder samples are used for both measurements to minimize the influence of water. Characteristic absorption peaks of surface-coating ligands in the FT-IR absorption spectra provide direct evidence of their presence on the Gd2O3 NP surfaces. Additionally, their coating amount can be estimated from TGA data, as most organic ligands undergo thermal decomposition below 450 °C via oxidation in a flow of hot air during TGA [96,97,98]. The coating amount is typically expressed in weight percent (wt.%), calculated from the mass loss observed in the TGA curve, excluding the initial mass loss between room temperature and ~100 °C due to desorption of water and air.
For example, Ho et al. synthesized arginylglycylaspartic acid (RGD)-PAA-coated ultrasmall Gd2O3 NPs (denoted as RGD-PAA-UGNPs) for cancer imaging using in vivo T1 MRI in tumor-model mice [95]. They confirmed the successful coating of RGD-PAA onto UGNPs by recording FT-IR absorption spectra of PAA-UGNPs, RGD-PAA-UGNPs, free PAA, and free RGD (Figure 5a). The C=O stretching peak at 1553 cm−1 in both the PAA-UGNP and RGD–PAA–UGNP spectra confirmed the successful coordination of PAA ligands to the UGNP surfaces. This peak exhibited a red shift from that of free PAA at ~1700 cm−1 due to electrostatic interactions between the COO groups of PAA and surface Gd3+ ions of UGNP. Furthermore, the FT–IR absorption spectrum of RGD–PAA–UGNPs showed characteristic RGD peaks, including N–H bending at 1544 cm−1 and C–N stretching at 1390 cm−1, confirming successful conjugation of RGD to PAA through amide bond formation.
For surface-coating amount estimation, Park et al. measured TGA curves of ultrasmall Gd2O3 NPs coated with five different kinds of ligands: PEI, pimelic acid, azelaic acid, suberic acid, and sebacic acid (Figure 5b) [99]. The water and air desorption from the NPs occurred between room temperature and ~105 °C. Mass drops between ~105 and ~450 °C were attributed to the thermal decomposition or oxidative combustion of the coating ligands. The residual mass in TGA curves corresponded to Gd2O3 NPs. The measured surface-coating amounts were 52.3% for pimelic acid, 60.9% for suberic acid, 55.7% for azelaic acid, 63.1% for sebacic acid, and 30.5% for PEI-70000.

3.5. Zeta Potentials

The zeta potential (ζ) of surface-coated Gd2O3 NPs dispersed in water provides insight into their surface charge in aqueous media. Negative zeta potentials indicate the presence of negatively charged functional groups such as -OH and -COOH on NP surfaces, whereas positive zeta potentials reflect the presence of positively charged functional groups such as -NH2. Zwitterionic ligands such as amino acids typically result in low zeta potential values [100]. High absolute zeta potential values are needed to ensure good colloidal stability [101,102].
For example, Marasini et al. reported polyaspartic acid (PSA)-coated Gd2O3 NPs as a dual-modal T1 and T2 MRI CA [63]. Figure 5c shows a highly negative zeta potential of −28.0 mV in aqueous media, consistent with their observed good colloidal stability, evidenced by the absence of NP precipitation following synthesis.

3.6. Magnetic Properties

The magnetic properties of surface-coated Gd2O3 NPs are typically assessed by recording magnetization (M) versus applied magnetic field (H) (M–H curves) and M versus temperature (T) (M–T curves) using either VSM or MPMS. The M–H curves provide saturation magnetization, coercivity, and remanence, while the M–T curves yield the blocking temperature. These curves are measured using 10–20 mg of powdered samples. Notably, M values of surface-coated Gd2O3 NPs arises predominantly from the spin magnetic moment of Gd3+ (s = 7/2), and is largely independent of both ligand type and particle size [21].
For example, Guleria et al. measured M–H and M–T curves of various polyol-coated Gd2O3 NPs [70]. Figure 6a and Figure 6b show the M–H curves of DEG-, TEG-, TeEG-, and PEG-coated Gd2O3 NPs at 5 and 300 K, respectively. The higher M values at 5 K are attributed to enhanced spin alignment of 4f-electron spins (s = 7/2) at lower temperatures, due to reduced thermal fluctuation. Consistent with the absence of hysteresis in the M–H curves, no magnetic transitions were observed in the M–T curves as depicted in Figure 6c,d, confirming the paramagnetic nature of the Gd2O3 NPs down to T = 2 K. Moreover, the negligible variation in magnetization among the DEG-, TEG-, TeEG-, and PEG-coated Gd2O3 NPs confirms that the magnetic moment originates primarily from the paramagnetic Gd2O3 core, as the surface-coating ligands are nonmagnetic and have minimal influence on the magnetic behavior.

3.7. Water Proton Spin Relaxivities (r1 and r2 Values)

In MRI, T1 (longitudinal or spin-lattice) and T2 (transverse or spin-spin) relaxation times indicate how quickly the excited proton spins return to their equilibrium state. T1 describes the recovery time of the longitudinal (i.e., z-component) magnetization along the main magnetic field. T2 describes the decay time of the transverse (i.e., x-y component) magnetization perpendicular to the main magnetic field [103,104].
The positive (T1) contrast efficacy increases with increasing r1 value while keeping r2/r1 ratio close to one, whereas negative (T2) contrast efficacy increases with increasing r2 value with r2/r1 ratio as large as possible. The r1 and r2 values are determined by measuring T1 and T2 water proton spin relaxation times at various Gd concentrations in aqueous solution samples using an MRI scanner. The inverse 1/T1 and 1/T2 relaxation times are plotted as a function of Gd concentration, and r1 and r2 values are obtained from the corresponding slopes. The r2 value is always greater than r1 because T1 relaxation always accompanies T2 relaxation whereas the opposite does not happen, making r2/r1 ratio always greater one. Therefore, positive (T1) contrast efficacy increases with increasing r1 value with r2/r1 ratio close to one, which are preferred to T1 MRI contrast agents. As shown in Figure 7, T1 relaxation is primarily induced by the interaction between the Gd3+ ion on Gd2O3 NP surface and water molecule (called inner sphere interaction model), whereas T2 relaxation is induced by the interaction between NP magnetic moment and water molecule (called outer sphere interaction model) [104].
For example, Guleria et al. measured the r1 and r2 values of Gd2O3 NPs coated with various ligands [70]. Figure 8a and Figure 8b depict the plots of 1/T1 and 1/T2 relaxation times, respectively, for DEG-, TEG-, TeEG-, and PEG-coated Gd2O3 NPs dispersed in water as function of Gd concentration. Both r1 and r2 values increased with increasing glycol chain length, as displayed in Figure 8c and Figure 8d, respectively, suggesting that proton relaxation rates are proportional to the thickness of the hydrophilic coating layer, probably due to enhanced access of outer-sphere water molecules facilitated by the thicker coating.

3.8. Toxicity of Surface-Modified Gd2O3 NPs

Gd2O3 NPs have demonstrated significant potential as T1 MRI CAs due to their considerably higher r1 and r2 values compared to Gd(III)-chelates [47,97,105,106,107,108,109,110,111,112]. However, their clinical translation is hindered by concerns over toxicity, primarily because free Gd3+ ions are toxic and cannot be metabolized by the body [113,114]. Therefore, complete excretion of injected Gd2O3 NPs, preferably via the renal system, without the release of free Gd3+ ions is essential. For clinically approved Gd(III)-chelates, clinical acute and chronic toxicities are well reported [113,114]: acute toxicity includes immediate hypersensitivity allergic-like reactions and physiologic reactions such as coldness, warmth, or pain at the injection site, nausea with or without vomiting, headache, paresthesia, and dizziness and chronic toxicity includes NSF, kidney disease, neurotoxicity, and gadolinium deposition disease such as bone pain and skin and subcutaneous tissue burning pain. However, there exist no clinical studies on acute and chronic toxicity for surface-modified Gd2O3 NPs. Therefore, we discussed in vitro cellular and in vivo animal toxicity studies reported so far in this section. Notably, free Gd3+ ion has an ionic radius comparable to that of Ca2+, enabling them to interfere with Ca2+-dependent metabolic process and potentially cause adverse effects [115,116]. To mitigate this risk, Gd2O3 NPs must be coated with hydrophilic and biocompatible ligands such as PAA, PMVEMA, PAAMA, PASA, D-glucuronic acid, CA, dextran, PEG, and PVP [51,58,61,63,64,66,68,73,75,93]. This review addresses both in vitro cellular toxicity and in vivo toxicity (i.e., histological analysis, body weight variation, biodistribution, clearance, and immune response).
Although there is a dose conversion formula for drugs between animals and human [117], the FDA-approved dose for gadolinium-based MRI contrast agents is 0.1 mmol/kg for human as well as animals, except for the liver-specific gadoxetic acid dose (0.025 mmol/kg) [118].

3.8.1. In Vitro Cellular Toxicity

In vitro cellular toxicity studies should be conducted prior to in vivo animal experiments to ensure that the surface-coated Gd2O3 NPs are nontoxic at the cellular level. Experimental evidence indicates that bare Gd2O3 NPs are inherently toxic; thus, surface coating with hydrophilic and biocompatible ligands is necessary to reduce cytotoxicity [69,93,119].
Common techniques for assessing cell viability include MTT, WST-1, and CCK-8 assays [120,121]. The MTT assay is a colorimetric method based on the metabolic activity of the cells, whereby mitochondrial dehydrogenases enzymes reduce the tetrazolium salt MTT into water-insoluble formazan crystals. After incubating cells with surface-coated Gd2O3 NPs (typically for 48 h at 37 °C), the resulting formazan crystals are solubilized using an organic solvent or detergent, and absorbance is measured at 570 nm. The WST-1 assay operates on a similar principle, but uses the tetrazolium salt WST-1, which is converted into a highly water-soluble formazan product by mitochondrial dehydrogenase enzymes; absorbance is measured at 450 nm following incubation with NP samples. The CCK-8 assay employs the water-soluble tetrazolium salt WST-8, which is reduced by viable cells into a water-soluble formazan product that also absorbs at 450 nm. Among these assays, WST-8 has been reported as the most sensitive.
For example, Ahmad et al. demonstrated the nontoxicity of PAA-coated Gd2O3 NPs in normal human embryonic kidney (HEK293) and human liver cancer (HepG2) cell lines [120]. As shown in Figure 9a, PAA-coated Gd2O3 NPs exhibited high cell viability in both cell lines up to a Gd concentration of 500 μM. Figure 9b presents the in vitro cytotoxicity results of PAAMA-coated ultrasmall Gd2O3 NPs in human prostate cancer (DU145) and normal mouse hepatocyte (NCTC1469) cell lines, showing no toxicity up to 500 µM Gd [61]. Similarly, Cui et al. evaluated the in vitro cytotoxicity of Gd2O3-FI-PEG and Gd2O3-FI-PEG-BBN NPs in the prostate cancer (PC)-3 cell line [77]. As shown in Figure 9c, cells treated with various Gd concentrations for both 24 and 48 h, demonstrated good viability for both types of surface-coated NPs. These findings consistently confirm the nontoxicity of surface-coated Gd2O3 NPs and underscore the importance of hydrophilic and biocompatible ligand surface coatings in mitigating cytotoxicity.

3.8.2. In Vivo Toxicity

Histology
Histological analysis enables the investigation of the in vivo toxicity of ligand-coated Gd2O3 NPs in organs and muscles at the cellular levels. To perform histological studies, mice are intravenously (IV) injected with ligand-coated Gd2O3 NPs and sacrificed at designated time points after injection to extract relevant organs and muscles. The harvested tissues are sectioned and imaged using an optical microscope, and the results are compared with those of control animals that did not receive NP injections.
For example, Sun et al. conducted a histological assessment of major organs such as the liver, kidney, spleen, and heart 30 days after IV administration of Gd2O3@BSA NPs [102]. As shown in Figure 10, no significant alterations or signs of toxicity were observed compared to control tissues, indicating the nontoxicity of Gd2O3@BSA NPs.
Body Weight Change
Body weight is another important indicator of in vivo toxicity and is measured using a balance before and at regular intervals after IV injection of surface-coated Gd2O3 NPs into mice tails. In the absence of NP-induced toxicity, body weight should remain stable over time. As shown in Figure 11a, no significant weight loss or gain was observed in mice at any of the tested injection doses (25, 50, and 100 mg/kg) compared to the control group without NP injection, indicating that the Gd2O3@BSA NPs did not induce toxicity [102].
Biodistribution
Biodistribution studies enable the tracking of injected NPs within the body. A solution of Gd-containing NPs is IV injected into mice tails, which are then sacrificed at specific time points after injection to extract organs and muscle for quantification of accumulated Gd concentrations. The biodistribution profile of injected NPs depends on particle size, morphology, and the nature of surface-coating ligands. Ideally, the NPs should be completely excreted from the body via the renal system, similar to Gd(III)-chelates, as free Gd3+ ions are toxic and cannot be eliminated through normal metabolic processes [123].
For example, Dai et al. investigated the biodistribution of PEG-Gd2O3 NPs (hydrodynamic diameter = 30–40 nm) in comparison with commercial Magnetist (Gd-DTPA) as control. They measured Gd accumulation in the heart, lung, liver, spleen, and kidneys at 1 and 12 h after IV injection into mice tails (injection dose = 15 μmol Gd/kg) [124]. As shown in Figure 11b, Magnevist exhibited considerably higher accumulation in the kidneys compared to other organs at 1 h post-injection, indicating rapid renal clearance. Additionally, the Gd concentration in all organs was higher for Magnevist than for PEG-Gd2O3 NPs, suggesting faster systemic circulation. At 12 h post-injection, Gd levels in all organs for Magnevist were very low, confirming efficient renal excretion. In contrast, PEG-Gd2O3 NPs showed considerable hepatic accumulation at 12 h, with Gd levels in the liver approximately 7.53 times higher than at 1 h. While the Gd concentration in the heart decreased over time, it increased in the lung, spleen and kidneys, indicating prolonged circulation and predominant hepatic clearance of PEG-Gd2O3 NPs. As another example, Ashouri et al. investigated the biodistribution of β-cyclodextrin-coated Gd2O3 (Gd2O3@PCD) NPs (d < 100 nm, a = 967.6 nm), using Omniscan (Gd-DTPA-BMA) and Dotarem (Gd-DOTA) as commercial controls in renally impaired rats [125]. As shown in Figure 11c, Gd2O3@PCD NPs exhibited only slightly higher Gd accumulation in all organs compared to Omniscan but considerably higher Gd accumulation than Dotarem, 7 days after tail vein injection (total Gd dose: 1.2 mmol Gd/kg for Gd2O3@PCD NPs; 30 mmol Gd/kg for Omniscan and Dotarem, administered as 0.2 or 5.0 mmol/kg per week for 6 weeks). These findings indicate that Dotarem was efficiently excreted through the renal system even in renally impaired rats, whereas both Gd2O3@PCD NPs and Omniscan showed notable organ accumulation, particularly in the skin. Yang et al. investigated the long-term biodistribution of the Gd(OH)3 nanorods in mice up to 150 days post-injection [56]. The Gd(OH)3 nanorods predominantly accumulated in the liver, spleen, and lungs. They showed that most Gd(OH)3 nanorods were rapidly cleared from the circulating blood, with accumulation in these organs increasing initially and then decreasing over time.
Clearance
Surface-modified Gd2O3 NPs should be completely cleared from the body within appropriate time frames after injection owing to their inherent toxicity. Clearance occurs via two primary pathways: renal excretion and hepatobiliary excretion, with the renal excretion being the preferred route. The feasibility of renal excretion primarily depends on particle size and hydrodynamic diameter. Gd2O3 NPs with diameters smaller than 5 nm and hydrodynamic diameter below 10 nm can be efficiently excreted through the renal system [123,126,127], whereas larger NPs are typically sequestered by macrophages in the liver and spleen and cleared via the hepatobiliary pathway [128,129,130].
For instance, for ultrasmall NPs (d < 5 nm), Wu et al. synthesized polymaleic acid (PMA)-coated extremely small gadolinium oxide NPs (ES-GON-PMA) (d = 2.1 nm) and evaluated their clearance following IV tail injection at a dosage of 5.0 mg/kg [128]. Gd levels were measured in urine and feces over a 24 h period. As shown in Figure 11d, urinary Gd content decreased rapidly over time, while fecal Gd levels remained negligible, indicating predominant renal excretion due to the ultrasmall particle size of ES-GdON-PMA. In contrast, for large NPs (d > 5 nm), Tian et al. investigated the excretion profile of larger Gd2O3@SiO2 NPs (d = 37.29 ± 1.03 nm) after IV administration into mice tails (15 µmol/kg) [129]. As shown in Figure 11e, excretion was monitored weekly in both feces and urine over a period of 12 weeks. The results exhibited considerably higher Gd content in feces than in urine, indicating that the large NPs were primarily cleared via the hepatobiliary pathway due to their insufficient renal clearance.
Figure 11. (a) Body weight variation over time in mice (N = 4) following injection with different doses of Gd2O3@BSA NPs (25, 50, and 100 mg/kg) compared to a control group (no injection), monitored for 30 days [102]. (b) Biodistribution of Magnevist and PEG-Gd2O3 NPs in the heart, lung, liver, spleen, and kidney at 1 and 12 h after IV injection (15 μmol Gd/kg) (N = 3; statistical significance p ** < 0.001 compared with the other group in the same organ; ID = percentage of injection dose per gram of organ) [124]. (c) Biodistribution of Gd2O3@PCD NPs, Omniscan, and Dotarem 7 days after tail vein injection (0.2 mmol/kg per week × 6 weeks = total 1.2 mmol Gd/kg for Gd2O3@PCD NPs; 5.0 mmol/kg per week × 6 weeks = total 30 mmol Gd/kg for Omniscan and Dotarem) (N = 5) [125]. (d) Time-dependent excretion of ES-GdON-PMA-9 (d = 2.1 nm) via urine and feces after IV injection (5.0 mg Gd/kg) (N = 3; I.D.%/g = percentage of injected dose per gram of urine or feces) [128]. (e) Excretion profile of Gd2O3@SiO2 NPs (d = 37.29 ± 1.03 nm) through feces and urine up to 12 weeks after mice tail vein injection (15 µmol/kg) (N = 3) [129].
Figure 11. (a) Body weight variation over time in mice (N = 4) following injection with different doses of Gd2O3@BSA NPs (25, 50, and 100 mg/kg) compared to a control group (no injection), monitored for 30 days [102]. (b) Biodistribution of Magnevist and PEG-Gd2O3 NPs in the heart, lung, liver, spleen, and kidney at 1 and 12 h after IV injection (15 μmol Gd/kg) (N = 3; statistical significance p ** < 0.001 compared with the other group in the same organ; ID = percentage of injection dose per gram of organ) [124]. (c) Biodistribution of Gd2O3@PCD NPs, Omniscan, and Dotarem 7 days after tail vein injection (0.2 mmol/kg per week × 6 weeks = total 1.2 mmol Gd/kg for Gd2O3@PCD NPs; 5.0 mmol/kg per week × 6 weeks = total 30 mmol Gd/kg for Omniscan and Dotarem) (N = 5) [125]. (d) Time-dependent excretion of ES-GdON-PMA-9 (d = 2.1 nm) via urine and feces after IV injection (5.0 mg Gd/kg) (N = 3; I.D.%/g = percentage of injected dose per gram of urine or feces) [128]. (e) Excretion profile of Gd2O3@SiO2 NPs (d = 37.29 ± 1.03 nm) through feces and urine up to 12 weeks after mice tail vein injection (15 µmol/kg) (N = 3) [129].
Nanomaterials 15 01340 g011
Immunotoxicity
The immunotoxicity assay involves the measurements of (1) ROS in neutrophils from peripheral blood, (2) innate immunity cluster of differentiation (CD) markers such as CD40, CD206, CD11b, and CD80/CD86 in monocytes/macrophages in peripheral blood, (3) adaptive immunity CD markers such as CD69 in lymphocytes from peripheral blood and lymph node, and (4) cytokines including IL-1β, IL-2, and IL-4 in serum [129,131].
For example, Tian et al. evaluated the immunotoxicity of Gd2O3@SiO2 NPs (Gd-NPs; d = 37.29 ± 1.03 nm) by quantifying ROS production, CD maker expression, and cytokine levels in BALB/c mice divided into three groups: PBS (100 µL, negative control), Gd-DTPA (15 µmol/kg), and Gd-NP (15 µmol Gd/kg), with N = 5 for each group [129]. As shown in Figure 12a–c, a statistically significant difference was observed between the Gd-NP and PBS groups in the expression levels of CD206, CD40, IL-1β, IL-2, and IL-4. However, no significant difference were found between the Gd-NP and Gd-DTPA groups, except ROS generation. As shown in Figure 12c, the Gd-DTPA and Gd-NP groups exhibited increased ROS production of 29% and 75%, respectively, relative to the PBS group. Although the Gd-NP group showed the highest ROS generation, this result is not critical, as ROS is a nonspecific mediator of immune response, capable of promoting either pathogen clearance or immunosuppression during tissue repair. Therefore, the data suggest that Gd-NPs induce minimal immunotoxicity. Similarly, Zheng et al. investigated the immunotoxicity of Gd2O3:Eu3+ NPs (d = 7.4 nm) in BALB/c mice using PBS and Gd-DTPA as controls [131] and obtained comparable results, further supporting that Gd2O3:Eu3+ NPs exhibit minimal immunotoxicity.

4. Factors Affecting r1 and r2 Values

The r1 and r2 values of surface-coated Gd2O3 NPs are key factors determining their performance as MRI CAs [5]. These values are influenced by several factors, such as particle diameter, surface-coating ligands, solution pH, temperature, and applied MR field strength [64,97,132,133,134,135]. Table 3 lists r1 and r2 values of various surface-coated Gd2O3 NPs prepared by different methods and measured under diverse conditions, illustrating the dependence of r1 and r2 values on these factors. This dependence highlights the potential to optimize the efficacy of surface-coated Gd2O3 NPs as T1 MRI CAs through controlled tuning of these variables. Notably, as provided in Table 3, r1 values of surface-modified Gd2O3 NPs are considerably higher than 3–5 s−1mM−1 [17,38] of clinically approved T1 MRI contrast agents, Gd(III)-chelates.

4.1. Particle Diameter

Monodispersed ultrasmall Gd2O3 NPs are preferred for T1 MRI applications because a higher proportion of surface Gd3+ ions is available to interact with surrounding water molecules, thereby enhancing the r1 value. Previous studies have consistently shown that the r1 value increases with decreasing particle diameter [80,109,143] and that ultrasmall Gd2O3 NPs have higher r1 values than Gd(III)-chelates, suggesting the existence of an optimal particle size for maximizing r1 value. This was demonstrated by Park et al. [64] and Rahman et al. [143]. According to Solomon-Bloembergen-Morgan theory [38,134], the T1 relaxation is due to interaction between Gd3+ ion and proton spins. As the number of surface Gd3+ ions of NPs increases, T1 relaxation (or r1 value) increases for ultrasmall d < ~2.5 nm [64,143] because majority of Gd3+ ions in NPs exist on NP surface due to high surface-to-volume (S/V) ratios, whereas T1 relaxation (or r1 value) decreases as d increases for d > ~2.5 nm because of low S/V ratios at which majority of Gd3+ ions in NPs are internal inactive ones for T1 relaxations.
Rahman et al. reported that the r1 value of CA-coated Gd2O3 NPs increased as particle size decreased below 3.0 nm at 7 T, reaching a maximum value at approximately 2.3 nm; further size reduction led to a decline in r1 value (Figure 13a) [143]. Park et al. roughly estimated the optimal particle size to lie between 1.1 and 2.5 nm [64]. In another study, the r1 values of PAA-octylamine-coated Gd2O3 NPs (d = 2, 5, 8, 11, and 22 nm) [80] and PEG-polysiloxane-coated Gd2O3 NPs (d = 2.2, 3.8, and 4.6 nm) [109] decreased with increasing particle diameter at 1.41 and 7 T, respectively (Figure 13b). This trend likely results from particle diameters exceeding the optimal range. Conversely, the r2 value, which is influenced by the NP magnetic moment, tends to increase with particle size due to the corresponding increase in magnetic moment per NP [64].

4.2. Surface-Coating Ligands

Surface coating is another critical factor influencing the r1 and r2 values of Gd2O3 NPs [71,75,97]. Appropriate selection of coating ligands can enhance the accessibility of water molecules near the Gd2O3 NP surface, thereby increasing the r1 and r2 values.
For example, Tegafaw et al. investigated the relaxometric properties of Gd2O3 NPs coated with various ligands, including small diacids with hydrophobic chains (succinic acid, glutaric acid, and terephthalic acid) and large PEIs (PEI-1300 and PEI-10,000) with hydrophilic chains [71]. As shown in Figure 13c, both r1 and r2 values decreased with increasing ligand size. This indicates that small, hydrophilic ligands promote closer interaction between water molecules and the NP surface, thereby enhancing r1 and r2 values. Similarly, Miao et al. reported that the r1 and r2 values of ultrasmall Gd2O3 NPs decreased as the molecular weight of PAA (Mw  =  1200, 5100, 15,000 Da) increased (Figure 13c) [75]. Kim et al. also observed a reduction in r1 and r2 values as the molecular weight of PEGD (Mw  =  250 and 600 Da) increased (Figure 13c) [97]. These results indicate that smaller, hydrophilic ligands are preferable for achieving higher r1 and r2 values.

4.3. Solution pH

The solution pH influences the surface charge and colloidal stability of surface-coated Gd2O3 NPs, thereby affecting their r1 and r2 values. pH can be adjusted by adding acid or base. For example, in the case of PAA-coated Gd2O3 NPs, the carboxyl groups on PAA ligands become protonated under acidic conditions, potentially leading to NP aggregation. Aggregation increases particle size and reduces the surface-to-volume ratios, which may limit water exchange between bulk water and the Gd2O3 NP surface, thereby decreasing the r1 value. Conversely, under basic conditions, the carboxyl groups are deprotonated, resulting in negatively charged surfaces that promote electrostatic repulsion between NPs. This enhanced repulsion improves colloidal stability and increases water accessibility to the NP surface, potentially enhancing r1 values. In contrast, the r2 value, which increases with the NP magnetic moment [64], may rise with NP aggregation due to the increase in the magnetic moment per NP. However, excessive aggregation can still compromise colloidal stability, which may in turn limit this effect.

4.4. Temperature

The r1 and r2 values are also influenced by temperature, as key parameters such as the diffusion coefficient (D) and rotational correlation time (τR) of water molecules and CAs are temperature-dependent [133,134,144,145]. As temperature increases, D increases and τR decreases, thereby shortening the interaction time between water proton spins and the NPs. This reduction in interaction time leads to decreased r1 and r2 values. For example, Goussuin et al. demonstrated that the r1 and r2 values of Gd(OH)3 NPs decreased with increasing temperature at 1.41 T, as shown in Figure 13d [144].

4.5. Applied Magnetic Field Strength (H)

Roch et al. theoretically investigated the r1 and r2 values of superparamagnetic NPs (d = 5 nm) as a function of the applied magnetic field strength (H) or proton Larmor frequency [133]. As shown in Figure 13e, three distinct regions were identified: Region 1, where r1 and r2 values remain nearly constant at H < 1 MHz; Region 2, where r1 and r2 values increase with H between 1 and 10 MHz; and Region 3, where r1 and r2 values decrease at H > 10 MHz (the clinical region).
For example, Carniato et al. synthesized CA-coated GdF3 NPs (d = 2.2–2.3 nm) and observed three regions in the r1 nuclear magnetic resonance dispersion (NMRD) profile, as shown in Figure 13f [135], in agreement with theoretical prediction [133]. Additional experimental r1 value plots as a function of H in Region 3 are displayed in Figure 13g, demonstrating that the r1 values of various ligand-coated Gd2O3 NPs and Gd(III)-chelates decrease with increasing H in this region [73,107,140,146]. Therefore, the proton signal enhancement by T1 MRI CAs diminishes with increasing H in the clinical region. However, due to enhanced proton spin alignment along H at higher field strengths, resulting in a larger proton spin population difference, the net proton signal enhancement at high H remains considerably greater than that predicted solely from the T1 MRI CA contribution.
Figure 13. Plot of r1 values of (a) CA-coated Gd2O3 NPs as a function of particle diameter (1.5–194.0 nm) at 7 T [143] and (b) PAA-octylamine- [80] and PEG-polysiloxane- [109] coated Gd2O3 NPs as a function of particle diameter at 1.41 and 7 T, respectively. (c) Plot of r1 values of various ligand-coated Gd2O3 NPs as a function of ligand size at 1.5 T [71,75,97]. (d) Plot of r1 values of Gd(OH)3 NPs as a function of temperature at 1.41 T [144]. (e) Theoretical NMRD profile of superparamagnetic NPs (d = 5 nm) [133]. (f) NMRD profile of CA-coated GdF3 NPs (d = 2.2–2.3 nm) [135]. (g) Plots of r1 values of various ligand-coated Gd2O3 NPs and Gd(III)-DTPA as a function of H [73,107,140,146].
Figure 13. Plot of r1 values of (a) CA-coated Gd2O3 NPs as a function of particle diameter (1.5–194.0 nm) at 7 T [143] and (b) PAA-octylamine- [80] and PEG-polysiloxane- [109] coated Gd2O3 NPs as a function of particle diameter at 1.41 and 7 T, respectively. (c) Plot of r1 values of various ligand-coated Gd2O3 NPs as a function of ligand size at 1.5 T [71,75,97]. (d) Plot of r1 values of Gd(OH)3 NPs as a function of temperature at 1.41 T [144]. (e) Theoretical NMRD profile of superparamagnetic NPs (d = 5 nm) [133]. (f) NMRD profile of CA-coated GdF3 NPs (d = 2.2–2.3 nm) [135]. (g) Plots of r1 values of various ligand-coated Gd2O3 NPs and Gd(III)-DTPA as a function of H [73,107,140,146].
Nanomaterials 15 01340 g013

5. In Vivo MRI Applications

5.1. Imaging in Normal Mice

The FDA-approved dose for gadolinium-based MRI contrast agents is 0.1 mmol/kg for human as well as animals, except for the liver-specific gadoxetic acid dose (0.025 mmol/kg) [118]. Numerous in vivo studies have reported the use of surface-coated Gd2O3 NPs as T1 MRI CAs, demonstrating their strong potential as T1 MRI CAs. For example, Ahmad et al. investigated PMVEMA-coated ultrasmall Gd2O3 NPs (d = 1.9 nm) as a T1 MRI CA [58]. Following IV administration of the aqueous NP solution into mice tails, positive (i.e., bright) contrast enhancement was observed in the liver and kidneys, as shown in Figure 14a. Signal-to-noise ratios (SNRs) of regions of interest (ROIs) in these organs initially increased, reached maxima, and then decreased over time, indicating accumulation followed by excretion of the NPs, predominantly through the renal pathway, thus confirming their role as T1 MRI CAs (Figure 14b). In another example, Yue et al. reported carbon-coated Gd2O3 (Gd2O3@C) NPs (d = 3.1 nm) as a T1 MRI CA [69]. Upon IV administration into mice tails, the NPs produced positive contrast enhancement in the liver, kidneys, and bladder (Figure 14c). The SNR-ROIs in these organs, plotted as a function of time (Figure 14d), showed that the contrast enhancements peaked between the point of administration and 30 min post-injection, and subsequently declined. These results further confirm the function of the NPs as a T1 MRI CA.

5.2. Imaging in Cancer Model Mice

Owing to the EPR effects, NPs can accumulate more readily in cancer cells than in normal cells (called passive targeting) [147]. Furthermore, the accumulation of NPs in cancer cells can be considerably enhanced by conjugating cancer-targeting ligands onto the NP surface (called active targeting).
Dai et al. showed that PEG-Gd2O3 NPs (r1 = 29.0 s−1 mM−1) displayed stronger contrast enhancement of the tumor in in vivo MR images than Magnevist (r1 = 4.2 s−1 mM−1) at the same amount of Gd injection, demonstrating the superiority of PEG-Gd2O3 NPs to commercial molecular contrast agent Magnevist [124]. Shen et al. evaluated T1 MR images of U87MG tumor model nude mice using exceedingly small (ES), RGD2-conjugated PAA-coated Gd2O3 (simply referred to as ES-GON5-PAA@RGD2) NPs [107]. As shown in Figure 15a, the tumor region exhibited a very weak MRI signal prior to IV injection. Following injection, signal intensity increased markedly, peaking at 2 h post-administration. Quantitative analysis in Figure 15b revealed a maximum ΔSNR [=100% × (SNRpost − SNRpre)/SNRpre] of 372 ± 56%, which is considerably higher than the ΔSNR of <80% of typically observed value with conventional Gd(III)-chelates. This remarkably enhanced tumor ΔSNR can be attributed to the combined effects of the high relaxivity value of the NPs (r1 = 68.7 ± 2.3 mM–1s–1, r2/r1 = 1.03 ± 0.03 at 1.5 T), the active targeting capability of RGD2 toward integrin αvβ3-overexpressed on tumor cells, and the EPR effect.

6. Conclusions and Future Outlook

Considerable research progress has been made on surface-coated Gd2O3 NPs as promising T1 MRI CAs. However, no comprehensive review on this subject has been reported to data. Therefore, a thorough overview addressing current status and future perspectives of surface-coated Gd2O3 NPs as T1 MRI CAs is highly warranted. Such a review would substantially advance research in this field and assist researchers in understanding the current state of the art and identifying future research directions.
This review has covered synthesis methods (i.e., polyol, DMSO, thermal decomposition, and hydrothermal techniques); characterization techniques (i.e., TEM, DLS, XRD, FTIR, TGA, zeta potential, VSM, and MRI); in vitro cellular and in vivo toxicity assessments (i.e., histology, biodistribution, body weight change, clearance, and immunotoxicity); and in vivo MRI applications of surface-modified Gd2O3 NPs. In addition, the influence of various factors (i.e., particle diameter, surface-coating ligands, solution pH, temperature, and applied MR field) on performance parameters of MRI CAs (i.e., r1 and r2 values) was discussed, as this information is critical for the rational design of optimized Gd2O3 NPs for specific applications. By comparing the strengths and limitations of each synthesis method, researchers can select the most appropriate approach for their intended application. The characterization protocols described herein may also serve as standard procedures for studies. Although previous studies have demonstrated the strong potential of surface-modified Gd2O3 NPs as T1 MRI CAs, further development of advanced formulations that fully meet in vivo toxicity criteria is essential for clinical translation. For instance, nontoxic, renally excretable formulations that prevent the release of free Gd3+ ions should be pursued.
While remarkable progress has been achieved in surface-modified Gd2O3 NPs, successful translation to clinical trial phases demands standardized methodologies. Future research should focus on conducting comprehensive toxicity assessments, including acute, sub-chronic, and chronic studies aligned with Food and Drug Administration (FDA) and European Medicines Agency (EMA) guidelines [148,149]. Standardized protocols for evaluating renal clearance and nephrotoxicity are also needed for clinical trial phases. Additionally, scalable manufacturing processes compliant with Good Manufacturing Practices (GMPs) are necessary for commercialization. GMP-compliant scale-up strategies involve a systematic approach to increasing production volume while maintaining product quality and regulatory compliance. Key aspects include facility design, process validation, quality control, supply chain management, and workforce training; all tailored to meet GMP requirements.
The surfaced-modified Gd2O3 NPs exhibited higher r1 values compared to surfaced-modified Mn and Fe-based NPs owing to the higher electron spin magnetic moment of Gd3+ (s = 7/2) compared to Mn2+ (s = 5/2) and Fe3+ (s = 5/2). The only disadvantage of surfaced-modified Gd2O3 NPs compared to surfaced-modified Mn- and Fe-based NPs is their higher potential toxicity. Therefore, advancement in biocompatible ligand coatings is critical to ensure the safety profile of surfaced-modified Gd2O3 NPs. Several studies assessed colloidal stability of surface-modified lanthanide oxide NPs in water, phosphate-buffered saline (PBS), or serum-containing media [51,69,150], demonstrating good colloidal stability overtime and thus, good coating stability.
Over the past decade, extensive efforts have been devoted to developing new T1 MRI CAs as alternatives to Gd(III)-chelates, which suffer from low sensitivity due to their low r1 values, as well as non-specificity and short imaging windows resulting from rapid blood clearance. These limitations often necessitate the administration of large doses. Surface-modified Gd2O3 NPs have emerged as promising alternatives, offering improved performance by overcoming the shortcomings of Gd(III)-chelates. Moreover, surface-modified Gd2O3 NPs can be easily conjugated with cancer-targeting ligands and therapeutic drugs, enabling targeted theranostic applications. They can be further functionalized via integration with other imaging agents such as FI, CT, PET, and SPECT agents to promote their application as multimodal imaging agents. While many of these advantages have been demonstrated in in vivo animal studies, considerable challenges remain in translating these systems into clinical practice. As discussed above, the development of advanced formulations and comprehensive in vivo toxicity evaluations are critical next steps. Achieving clinical translation will require sustained interdisciplinary research efforts.

Author Contributions

Conceptualization, E.M. and T.T.; methodology, T.T. and E.M.; validation, Y.L., D.Z., A.B., J.K. and X.C.; writing—original draft preparation, E.M. and T.T.; writing—review and editing, Y.C. and G.H.L.; funding acquisition, J.K., Y.C. and G.H.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Basic Science Research Program of the National Research Foundation (NRF) funded by the Korea government (Ministry of Science, and Information and Communications Technology: MSIT) (Basic Research Laboratory, No. RS-2024-00406209) and the NRF funded by the Ministry of Education (Post-Doc. Growth Type Cooperational Research, No. RS-2024-00459895).

Data Availability Statement

The data presented in this study are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Wu, M.; Shu, J. Multimodal Molecular Imaging: Current Status and Future Directions. Contrast Media Mol. Imaging 2018, 2018, 1382183. [Google Scholar] [CrossRef]
  2. Chen, Z.-Y.; Wang, Y.-X.; Lin, Y.; Zhang, J.-S.; Yang, F.; Zhou, Q.-L.; Liao, Y.-Y. Advance of Molecular Imaging Technology and Targeted Imaging Agent in Imaging and Therapy. BioMed Res. Int. 2014, 2014, 819324. [Google Scholar] [CrossRef]
  3. Han, X.; Xu, K.; Taratula, O.; Farsad, K. Applications of Nanoparticles in Biomedical Imaging. Nanoscale 2019, 11, 799–819. [Google Scholar] [CrossRef]
  4. Kim, S.J.; Lee, H.-Y. In Vivo Molecular Imaging in Preclinical Research. Lab. Anim. Res. 2022, 38, 31. [Google Scholar] [CrossRef]
  5. Ni, D.; Bu, W.; Ehlerding, E.B.; Cai, W.; Shi, J. Engineering of Inorganic Nanoparticles as Magnetic Resonance Imaging Contrast Agents. Chem. Soc. Rev. 2017, 46, 7438–7468. [Google Scholar] [CrossRef]
  6. Chouliaras, L.; O’Brien, J.T. The Use of Neuroimaging Techniques in the Early and Differential Diagnosis of Dementia. Mol. Psychiatry 2023, 28, 4084–4097. [Google Scholar] [CrossRef]
  7. McCarthy, J.; Collins, D.L.; Ducharme, S. Morphometric MRI as a Diagnostic Biomarker of Frontotemporal Dementia: A Systematic Review to Determine Clinical Applicability. NeuroImage Clin. 2018, 20, 685–696. [Google Scholar] [CrossRef]
  8. Iyad, N.; Ahmad, M.S.; Alkhatib, S.G.; Hjouj, M. Gadolinium contrast agents- challenges and opportunities of a multidisciplinary approach: Literature review. Eur. J. Radiol. Open 2023, 11, 100503. [Google Scholar] [CrossRef] [PubMed]
  9. Zhou, Z.; Lu, Z.-R. Gadolinium-Based Contrast Agents for MR Cancer Imaging. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 2013, 5, 1–18. [Google Scholar] [CrossRef] [PubMed]
  10. Palagi, L.; Longo, D.L.; Toth, E.; Quattrocchi, C.C.; van der Molen, A.J.; Aime, S.; Gianolio, E. Molecular and supramolecular routes to enhance Gadolinium-based contrast agents relaxivity: How far are we from the theoretical optimal value? Eur. J. Med. Chem. 2025, 292, 117668. [Google Scholar] [CrossRef] [PubMed]
  11. Wan, F.; Wu, L.; Chen, X.; Zhang, Y.; Jiang, L. Research progress on manganese complexes as contrast agents for magnetic resonance imaging. Polyhedron 2023, 242, 116489. [Google Scholar] [CrossRef]
  12. Zhao, D.; Baek, A.; Tegafaw, T.; Liu, T.; Liu, T.; Yue, H.; Mulugeta, E.; Yang, J.; Park, J.A.; Lee, H.; et al. Poly(acrylic acid)-Grafted BixGd2–xO3 Nanoparticles as Multimodal T1/T2 MRI and X-ray CT Contrast Agents. ACS Appl. Nano Mater. 2025, 8, 13663–13675. [Google Scholar] [CrossRef]
  13. Marasini, S.; Yue, H.; Ho, S.-L.; Park, J.-A.; Kim, S.; Yang, J.-U.; Cha, H.; Liu, S.; Tegafaw, T.; Ahmad, M.Y.; et al. In vivo positive magnetic resonance imaging of brain cancer (U87MG) using folic acid-conjugated polyacrylic acid-coated ultrasmall manganese oxide nanoparticles. Appl. Sci. 2021, 11, 2596. [Google Scholar] [CrossRef]
  14. Yue, H.; Zhao, D.; Tegafaw, T.; Ahmad, M.Y.; Saidi, A.K.A.A.; Ying, L.; Lee, S.; Cha, H.; Yang, B.W.; Chae, K.S.; et al. Core-Shell Fe3O4@C Nanoparticles as Highly Effective T2 Magnetic Resonance Imaging Contrast Agents: In Vitro and In Vivo Studies. Nanomaterials 2024, 14, 177. [Google Scholar] [CrossRef]
  15. Liu, Y.; Ho, S.L.; Tegafaw, T.; Zhao, D.; Ahmad, M.Y.; Saidi, A.K.A.A.; Cha, H.; Lee, S.; Lee, H.; Kim, S.; et al. Clustered ultra-small iron oxide nanoparticles as potential T1/T2 dual–modal magnetic resonance imaging contrast agents and application to tumor model. Nanotechnology 2024, 35, 505101. [Google Scholar] [CrossRef]
  16. Geraldes, C.F.G.C. Rational Design of Magnetic Nanoparticles as T1–T2 Dual-Mode MRI Contrast Agents. Molecules 2024, 29, 1352. [Google Scholar] [CrossRef]
  17. Wahsner, J.; Gale, E.M.; Rodríguez-Rodríguez, A.; Caravan, P. Chemistry of MRI Contrast Agents: Current Challenges and New Frontiers. Chem. Rev. 2019, 119, 957–1057. [Google Scholar] [CrossRef]
  18. Li, L.; Jiang, W.; Luo, K.; Song, H.; Lan, F.; Wu, Y.; Gu, Z. Superparamagnetic Iron Oxide Nanoparticles as MRI contrast agents for Non-invasive Stem Cell Labeling and Tracking. Theranostics 2013, 3, 595–615. [Google Scholar] [CrossRef]
  19. Scarciglia, A.; Papi, C.; Romiti, C.; Leone, A.; Di Gregorio, E.; Ferrauto, G. Gadolinium-Based Contrast Agents (GBCAs) for MRI: A Benefit–Risk Balance Analysis from a Chemical, Biomedical, and Environmental Point of View. Glob. Chall. 2025, 9, 2400269. [Google Scholar] [CrossRef] [PubMed]
  20. Antwi-Baah, R.; Wang, Y.; Chen, X.; Yu, K. Metal-Based Nanoparticle Magnetic Resonance Imaging Contrast Agents: Classifications, Issues, and Countermeasures toward Their Clinical Translation. Adv. Mater. Interfaces 2022, 9, 2101710. [Google Scholar] [CrossRef]
  21. Tegafaw, T.; Liu, S.; Ahmad, M.Y.; Saidi, A.K.A.A.; Zhao, D.; Liu, Y.; Nam, S.-W.; Chang, Y.; Lee, G.H. Magnetic Nanoparticle-Based High-Performance Positive and Negative Magnetic Resonance Imaging Contrast Agents. Pharmaceutics 2023, 15, 1745. [Google Scholar] [CrossRef]
  22. Caspani, S.; Magalhães, R.; Araújo, J.P.; Sousa, C.T. Magnetic Nanomaterials as Contrast Agents for MRI. Materials 2020, 13, 2586. [Google Scholar] [CrossRef]
  23. Lee, S.H.; Kim, B.H.; Na, H.B.; Hyeon, T. Paramagnetic inorganic nanoparticles as T1 MRI contrast agents. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 2014, 6, 196–209. [Google Scholar] [CrossRef]
  24. Hu, F.; Zhao, Y.S. Inorganic nanoparticle-based T1 and T1/T2 magnetic resonance contrast probes. Nanoscale 2012, 4, 6235–6243. [Google Scholar] [CrossRef] [PubMed]
  25. Pellico, J.; Ellis, C.M.; Davis, J.J. Nanoparticle-Based Paramagnetic Contrast Agents for Magnetic Resonance Imaging. Contrast Media Mol. Imaging 2019, 2019, 1845637. [Google Scholar] [CrossRef] [PubMed]
  26. Panich, A.M.; Shames, A.I.; Goren, S.D.; Yudina, E.B.; Aleksenskii, A.E.; Vul, A.Y. Examining relaxivities in suspensions of nanodiamonds grafted by magnetic entities: Comparison of two approaches. Magn. Reson. Mater. Phys. Biol. Med. 2020, 33, 885–888. [Google Scholar] [CrossRef] [PubMed]
  27. Panich, A.M.; Salti, M.; Aleksenskii, A.E.; Kulvelis, Y.V.; Chizhikova, A.; Vul, A.Y.; Shames, A.I. Suspensions of manganese-grafted nanodiamonds: Preparation, NMR, and MRI study. Diam. Relat. Mater. 2023, 131, 109591. [Google Scholar] [CrossRef]
  28. Panich, A.M.; Salti, M.; Prager, O.; Swissa, E.; Kulvelis, Y.V.; Yudina, E.B.; Aleksenskii, A.E.; Goren, S.D.; Vul, A.Y.; Shames, A.I. PVP-coated Gd-grafted nanodiamonds as a novel and potentially safer contrast agent for in vivo MRI. Magn. Reson. Med. 2021, 86, 935–942. [Google Scholar] [CrossRef]
  29. Jeon, M.; Halbert, M.V.; Stephen, Z.R.; Zhang, M. Iron Oxide Nanoparticles as T1 Contrast Agents for Magnetic Resonance Imaging: Fundamentals, Challenges, Applications, and Prospectives. Adv. Mater. 2021, 33, 1906539. [Google Scholar] [CrossRef] [PubMed]
  30. Liu, T.; Spincemaille, P.; de Rochefort, L.; Wong, R.; Prince, M.; Wang, Y. Unambiguous Identification of Superparamagnetic Iron Oxide Particles through Quantitative Susceptibility Mapping of the Nonlinear Response to Magnetic Fields. Magn. Reson. Imaging 2010, 28, 1383–1389. [Google Scholar] [CrossRef]
  31. van den Bos, E.J.; Baks, T.; Moelker, A.D.; Kerver, W.; van Geuns, R.-J.; van der Giessen, W.J.; Duncker, D.J.; Wielopolski, P.A. Magnetic Resonance Imaging of Haemorrhage within Reperfused Myocardial Infarcts: Possible Interference with Iron Oxide-Labelled Cell Tracking? Eur. Heart J. 2006, 27, 1620–1626. [Google Scholar] [CrossRef]
  32. Girard, O.M.; Du, J.; Agemy, L.; Sugahara, K.N.; Kotamraju, V.R.; Ruoslahti, E.; Bydder, G.M.; Mattrey, R.F. Optimization of Iron Oxide Nanoparticle Detection Using Ultrashort Echo Time Pulse Sequences: Comparison of T1, T2*, and Synergistic T1−T2* Contrast Mechanisms. Magn. Reson. Med. 2011, 65, 1649–1660. [Google Scholar] [CrossRef]
  33. Eibofner, F.; Steidle, G.; Kehlbach, R.; Bantleon, R.; Schick, F. Positive Contrast Imaging of Iron Oxide Nanoparticles with Susceptibility-Weighted Imaging. Magn. Reson. Med. 2010, 64, 1027–1038. [Google Scholar] [CrossRef]
  34. Hedlund, A.; Ahrén, M.; Gustafsson, H.; Abrikossova, N.; Warntjes, M.; Jönsson, J.I.; Uvdal, K.; Engström, M. Gd2O3 nanoparticles in hematopoietic cells for MRI contrast enhancement. Int. J. Nanomed. 2011, 6, 3233–3240. [Google Scholar] [CrossRef]
  35. Blomqvist, L.; Nordberg, G.F.; Nurchi, V.M.; Aaseth, J.O. Gadolinium in Medical Imaging—Usefulness, Toxic Reactions and Possible Countermeasures—A Review. Biomolecules 2022, 12, 742. [Google Scholar] [CrossRef]
  36. Pichler, V.; Martinho, R.P.; Temming, L.; Segers, T.; Wurm, F.R.; Koshkina, O. The Environmental Impact of Medical Imaging Agents and the Roadmap to Sustainable Medical Imaging. Adv. Sci. 2025, 12, 2404411. [Google Scholar] [CrossRef] [PubMed]
  37. Idée, J.-M.; Port, M.; Robic, C.; Medina, C.; Sabatou, M.; Corot, C. Role of Thermodynamic and Kinetic Parameters in Gadolinium Chelate Stability. J. Magn. Reson. Imaging 2009, 30, 1249–1258. [Google Scholar] [CrossRef] [PubMed]
  38. Caravan, P.; Ellison, J.J.; McMurry, T.J.; Lauffer, R.B. Gadolinium(III) Chelates as MRI Contrast Agents:  Structure, Dynamics, and Applications. Chem. Rev. 1999, 99, 2293–2352. [Google Scholar] [CrossRef]
  39. Oliveira, I.S.; Hedgire, S.S.; Li, W.; Ganguli, S.; Prabhakar, A.M. Blood pool contrast agents for venous magnetic resonance imaging. Cardiovasc. Diagn. Ther. 2016, 6, 508–518. [Google Scholar] [CrossRef]
  40. Thomsen, H.S.; Morcos, S.K.; Almén, T.; Bellin, M.-F.; Bertolotto, M.; Bongartz, G.; Clement, O.; Leander, P.; Heinz-Peer, G.; Reimer, P.; et al. Nephrogenic systemic fibrosis and gadolinium-based contrast media: Updated ESUR Contrast Medium Safety Committee guidelines. Eur. Radiol. 2013, 23, 307–318. [Google Scholar] [CrossRef] [PubMed]
  41. Thomsen, H.S. Nephrogenic Systemic Fibrosis: History and Epidemiology. Radiol. Clin. N. Am. 2009, 47, 827–831. [Google Scholar] [CrossRef]
  42. Bernstein, E.J.; Schmidt-Lauber, C.; Kay, J. Nephrogenic Systemic Fibrosis: A Systemic Fibrosing Disease Resulting from Gadolinium Exposure. Best Pract. Res. Clin. Rheumatol. 2012, 26, 489–503. [Google Scholar] [CrossRef]
  43. Murata, N.; Murata, K.; Gonzalez-Cuyar, L.F.; Maravilla, K.R. Gadolinium tissue deposition in brain and bone. Magn. Reson. Imaging 2016, 34, 1359–1365. [Google Scholar] [CrossRef]
  44. Khairinisa, M.A.; Ariyani, W.; Tsushima, Y.; Koibuchi, N. Effects of Gadolinium Deposits in the Cerebellum: Reviewing the Literature from In Vitro Laboratory Studies to In Vivo Human Investigations. Int. J. Environ. Res. Public Health 2021, 18, 7214. [Google Scholar] [CrossRef] [PubMed]
  45. Roux, S.; Faure, A.-C.; Mandon, C.; Dufort, S.; Rivière, C.; Bridot, J.-L.; Mutelet, B.; Marquette, C.A.; Josserand, V.; Le Duc, G.; et al. Multifunctional Gadolinium Oxide Nanoparticles: Towards Image-Guided Therapy. Imaging Med. 2010, 2, 211–223. [Google Scholar] [CrossRef]
  46. Wu, C.; Cai, R.; Zhao, T.; Wu, L.; Zhang, L.; Jin, J.; Xu, L.; Li, P.; Li, T.; Zhang, M.; et al. Hyaluronic Acid-Functionalized Gadolinium Oxide Nanoparticles for Magnetic Resonance Imaging-Guided Radiotherapy of Tumors. Nanoscale Res. Lett. 2020, 15, 94. [Google Scholar] [CrossRef]
  47. Ahmad, M.W.; Xu, W.; Kim, S.J.; Baeck, J.S.; Chang, Y.; Bae, J.E.; Chae, K.S.; Park, J.A.; Kim, T.J.; Lee, G.H. Potential dual imaging nanoparticle: Gd2O3 nanoparticle. Sci. Rep. 2015, 5, 8549. [Google Scholar] [CrossRef]
  48. Ho, S.L.; Yue, H.; Tegafaw, T.; Ahmad, M.Y.; Liu, S.; Nam, S.-W.; Chang, Y.; Lee, G.H. Gadolinium Neutron Capture Therapy (GdNCT) Agents from Molecular to Nano: Current Status and Perspectives. ACS Omega 2022, 7, 2533–2553. [Google Scholar] [CrossRef]
  49. Zhang, Y.; Tang, N.; Zhou, H.; Zhu, Y. Surface engineered multifunctional nano-systems for localised drug delivery against thyroid cancer: A review of current practices. Biomed. Pharmacother. 2024, 176, 116840. [Google Scholar] [CrossRef] [PubMed]
  50. Al-Thani, A.N.; Jan, A.G.; Abbas, M.; Geetha, M.; Sadasivuni, K.K. Nanoparticles in cancer theragnostic and drug delivery: A comprehensive review. Life Sci. 2024, 352, 122899. [Google Scholar] [CrossRef] [PubMed]
  51. Miao, X.; Ho, S.L.; Tegafaw, T.; Cha, H.; Chang, Y.; Oh, I.T.; Yaseen, A.M.; Marasini, S.; Ghazanfari, A.; Yue, H.; et al. Stable and Non-Toxic Ultrasmall Gadolinium Oxide Nanoparticle Colloids (Coating Material = Polyacrylic Acid) as High-Performance T1 Magnetic Resonance Imaging Contrast Agents. RSC Adv. 2018, 8, 3189–3197. [Google Scholar] [CrossRef] [PubMed]
  52. Liu, S.; Yue, H.; Ho, S.L.; Kim, S.; Park, J.A.; Tegafaw, T.; Ahmad, M.Y.; Kim, S.; Saidi, A.K.A.A.; Zhao, D.; et al. Enhanced Tumor Imaging Using Glucosamine-Conjugated Polyacrylic Acid-Coated Ultrasmall Gadolinium Oxide Nanoparticles in Magnetic Resonance Imaging. Int. J. Mol. Sci. 2022, 23, 1792. [Google Scholar] [CrossRef]
  53. Ahrén, M.; Selegård, L.; Söderlind, F.; Linares, M.; Kauczor, J.; Norman, P.; Käll, P.-O.; Uvdal, K. A Simple Polyol-Free Synthesis Route to Gd2O3 Nanoparticles for MRI Applications: An Experimental and Theoretical Study. J. Nanopart. Res. 2012, 14, 1006. [Google Scholar] [CrossRef]
  54. Zhao, D.; Liu, Y.; Ho, S.L.; Tegafaw, T.; Saidi, A.K.A.A.; Lee, H.; Ahn, D.; Nam, H.; Park, J.A.; Yang, J.-U.; et al. Multi-functional GdEuxTb1−xO3 (x = 0 to 1) nanoparticles: Colour tuning optical properties, water proton spin relaxivities, and X-ray attenuation properties. Nanoscale 2024, 16, 16998–17008. [Google Scholar] [CrossRef] [PubMed]
  55. Cai, W.; Zhang, Y.; Wang, J.; Wang, Z.; Tian, Y.; Liu, H.; Pan, H.; Fu, L.; Chen, W.; Wu, C.; et al. Engineering the Surface of Gd2O3 Nanoplates for Improved T1-Weighted Magnetic Resonance Imaging. Chem. Eng. J. 2020, 380, 122473. [Google Scholar] [CrossRef]
  56. Yang, Y.; Sun, Y.; Liu, Y.; Peng, J.; Wu, Y.; Zhang, Y.; Feng, W.; Li, F. Long-term in vivo biodistribution and toxicity of Gd(OH)3 nanorods. Biomaterials 2013, 34, 508–515. [Google Scholar] [CrossRef]
  57. Tegafaw, T.; Liu, Y.; Ho, S.L.; Liu, S.; Ahmad, M.Y.; Saidi, A.K.A.A.; Zhao, D.; Ahn, D.; Nam, H.; Chae, W.-S.; et al. High-Quantum-Yield Ultrasmall Ln2O3 (Ln = Eu, Tb, or Dy) Nanoparticle Colloids in Aqueous Media Obtained via Photosensitization. Langmuir 2023, 39, 15338–15342. [Google Scholar] [CrossRef]
  58. Ahmad, M.Y.; Ahmad, M.W.; Yue, H.; Ho, S.L.; Park, J.A.; Jung, K.-H.; Cha, H.; Marasini, S.; Ghazanfari, A.; Liu, S.; et al. In Vivo Positive Magnetic Resonance Imaging Applications of Poly(Methyl Vinyl Ether-Alt-Maleic Acid)-Coated Ultra-Small Paramagnetic Gadolinium Oxide Nanoparticles. Molecules 2020, 25, 1159. [Google Scholar] [CrossRef]
  59. Saidi, A.K.A.A.; Baek, A.; Liu, S.; Mengesha, T.T.; Ahmad, M.Y.; Zhao, D.; Liu, Y.; Yang, J.U.; Park, J.A.; Hwang, D.W.; et al. Lead Oxide Nanoparticles as X-Ray Contrast Agents for In Vitro and In Vivo Imaging. ACS Appl. Nano Mater. 2023, 6, 20129–20138. [Google Scholar] [CrossRef]
  60. Saidi, A.K.A.A.; Ghazanfari, A.; Liu, S.; Tegafaw, T.; Ahmad, M.Y.; Zhao, D.; Liu, Y.; Yang, S.H.; Hwang, D.W.; Yang, J.-U.; et al. Facile Synthesis and X-ray Attenuation Properties of Ultrasmall Platinum Nanoparticles Grafted with Three Types of Hydrophilic Polymers. Nanomaterials 2023, 13, 806. [Google Scholar] [CrossRef] [PubMed]
  61. Jang, Y.-J.; Liu, S.; Yue, H.; Park, J.A.; Cha, H.; Ho, S.L.; Marasini, S.; Ghazanfari, A.; Ahmad, M.Y.; Miao, X.; et al. Hydrophilic Biocompatible Poly(Acrylic Acid-Co-Maleic Acid) Polymer as a Surface-Coating Ligand of Ultrasmall Gd2O3 Nanoparticles to Obtain a High r1 Value and T1 MR Images. Diagnostics 2021, 11, 2. [Google Scholar] [CrossRef] [PubMed]
  62. Saidi, A.K.A.A.; Ghazanfari, A.; Baek, A.; Tegafaw, T.; Ahmad, M.Y.; Zhao, D.; Liu, Y.; Yang, J.-u.; Park, J.A.; Yang, B.W.; et al. Ultrasmall Cerium Oxide Nanoparticles as Highly Sensitive X-Ray Contrast Agents and Their Antioxidant Effect. RSC Adv. 2024, 14, 3647–3658. [Google Scholar] [CrossRef]
  63. Marasini, S.; Yue, H.; Ghazanfari, A.; Ho, S.L.; Park, J.A.; Kim, S.; Cha, H.; Liu, S.; Tegafaw, T.; Ahmad, M.Y.; et al. Polyaspartic Acid-Coated Paramagnetic Gadolinium Oxide Nanoparticles as a Dual-Modal T1 and T2 Magnetic Resonance Imaging Contrast Agent. Appl. Sci. 2021, 11, 8222. [Google Scholar] [CrossRef]
  64. Park, J.Y.; Baek, M.J.; Choi, E.S.; Woo, S.; Kim, J.H.; Kim, T.J.; Jung, J.C.; Chae, K.S.; Chang, Y.; Lee, G.H. Paramagnetic Ultrasmall Gadolinium Oxide Nanoparticles as Advanced T1 MRI Contrast Agent: Account for Large Longitudinal Relaxivity, Optimal Particle Diameter, and In Vivo T1 MR Images. ACS Nano 2009, 3, 3663–3669. [Google Scholar] [CrossRef]
  65. Liu, S.; Tegafaw, T.; Ho, S.L.; Yue, H.; Zhao, D.; Liu, Y.; Mulugeta, E.; Chen, X.; Lee, H.; Ahn, D.; et al. Magnetic Resonance Imaging and X-Ray Imaging Properties of Ultrasmall Lanthanide Oxide (Ln = Eu, Gd, and Tb) Nanoparticles Synthesized via Thermal Decomposition. Molecules 2025, 30, 2519. [Google Scholar] [CrossRef]
  66. Xu, W.; Miao, X.; Oh, I.; Chae, K.S.; Cha, H.; Chang, Y.; Lee, G.H. Dextran-Coated Ultrasmall Gd2O3 Nanoparticles as Potential T1 MRI Contrast Agent. ChemistrySelect 2016, 1, 6086–6091. [Google Scholar] [CrossRef]
  67. Ahab, A.; Rohman, F.; Iskandar, F.; Haryanto, F.; Arif, I. A simple straightforward thermal decomposition synthesis of PEG-covered Gd2O3 (Gd2O3@PEG) nanoparticles. Adv. Powder Technol. 2016, 27, 1800–1805. [Google Scholar] [CrossRef]
  68. Vahdatkhah, P.; Madaah Hosseini, H.R.; Khodaei, A.; Montazerabadi, A.R.; Irajirad, R.; Oghabian, M.A.; Delavari, H.H. Rapid Microwave-Assisted Synthesis of PVP-Coated Ultrasmall Gadolinium Oxide Nanoparticles for Magnetic Resonance Imaging. Chem. Phys. 2015, 453–454, 35–41. [Google Scholar] [CrossRef]
  69. Yue, H.; Marasini, S.; Ahmad, M.Y.; Ho, S.L.; Cha, H.; Liu, S.; Jang, Y.J.; Tegafaw, T.; Ghazanfari, A.; Miao, X.; et al. Carbon-coated ultrasmall gadolinium oxide (Gd2O3@C) nanoparticles: Application to magnetic resonance imaging and fluorescence properties. Colloids Surf. A Physicochem. Eng. Asp. 2020, 586, 124261. [Google Scholar] [CrossRef]
  70. Guleria, A.; Pranjali, P.; Meher, M.K.; Chaturvedi, A.; Chakraborti, S.; Raj, R.; Poluri, K.M.; Kumar, D. Effect of Polyol Chain Length on Proton Relaxivity of Gadolinium Oxide Nanoparticles for Enhanced Magnetic Resonance Imaging Contrast. J. Phys. Chem. C 2019, 123, 18061–18070. [Google Scholar] [CrossRef]
  71. Tegafaw, T.; Xu, W.; Lee, S.H.; Chae, K.S.; Cha, H.; Chang, Y.; Lee, G.H. Ligand-size and ligand-chain hydrophilicity effects on the relaxometric properties of ultrasmall Gd2O3 nanoparticles. AIP Adv. 2016, 6, 065114. [Google Scholar] [CrossRef]
  72. Fortin, M.A.; Petoral, R.M.; Söderlind, F.; Klasson, A.; Engström, M.; Veres, T.; Käll, P.O.; Uvdal, K. Polyethylene glycol-covered ultra-small Gd2O3 nanoparticles for positive contrast at 1.5 T magnetic resonance clinical scanning. Nanotechnology 2007, 18, 395501. [Google Scholar] [CrossRef]
  73. Faucher, L.; Tremblay, M.; Lagueux, J.; Gossuin, Y.; Fortin, M.-A. Rapid Synthesis of PEGylated Ultrasmall Gadolinium Oxide Nanoparticles for Cell Labeling and Tracking with MRI. ACS Appl. Mater. Interfaces 2012, 4, 4506–4515. [Google Scholar] [CrossRef] [PubMed]
  74. Rahman, A.T.M.A.; Vasilev, K.; Majewski, P. Ultra small Gd2O3 nanoparticles: Absorption and emission properties. J. Colloid. Interface Sci. 2011, 354, 592–596. [Google Scholar] [CrossRef]
  75. Miao, X.; Xu, W.; Cha, H.; Chang, Y.; Oh, I.T.; Chae, K.S.; Tegafaw, T.; Ho, S.L.; Kim, S.J.; Lee, G.H. Ultrasmall Gd2O3 nano-particles surface-coated by polyacrylic acid (PAA) and their PAA-size dependent relaxometric properties. Appl. Surf. Sci. 2019, 477, 111–115. [Google Scholar] [CrossRef]
  76. Hu, Z.; Ahrén, M.; Selegård, L.; Skoglund, C.; Söderlind, F.; Engström, M.; Zhang, X.; Uvdal, K. Highly Water-Dispersible Surface-Modified Gd2O3 Nanoparticles for Potential Dual-Modal Bioimaging. Chem. Eur. J. 2013, 19, 12658–12667. [Google Scholar] [CrossRef]
  77. Cui, D.; Lu, X.; Yan, C.; Liu, X.; Hou, M.; Xia, Q.; Xu, Y.; Liu, R. Gastrin-releasing peptide receptor-targeted gadolinium oxide-based multifunctional nanoparticles for dual magnetic resonance/fluorescent molecular imaging of prostate cancer. Int. J. Nanomed. 2017, 12, 6787–6797. [Google Scholar] [CrossRef] [PubMed]
  78. Fang, J.; Chandrasekharan, P.; Liu, X.-L.; Yang, Y.; Lv, Y.-B.; Yang, C.-T.; Ding, J. Manipulating the surface coating of ultra-small Gd2O3 nanoparticles for improved T1-weighted MR imaging. Biomaterials 2014, 35, 1636–1642. [Google Scholar] [CrossRef]
  79. Gu, W.; Song, G.; Li, S.; Shao, C.; Yan, C.; Ye, L. Chlorotoxin-conjugated, PEGylated Gd2O3 nanoparticles as a glioma-specific magnetic resonance imaging contrast agent. RSC Adv. 2014, 4, 50254–50260. [Google Scholar] [CrossRef]
  80. Cho, M.; Sethi, R.; Narayanan, J.S.A.; Lee, S.S.; Benoit, D.N.; Taheri, N.; Decuzzi, P.; Colvin, V.L. Gadolinium oxide nanoplates with high longitudinal relaxivity for magnetic resonance imaging. Nanoscale 2014, 6, 13637–13645. [Google Scholar] [CrossRef]
  81. Khue, N.T.T.; Tam, L.T.T.; Dung, N.T.; Tam, L.T.; Chung, N.X.; Linh, N.T.N.; Vinh, N.D.; Quy, B.M.; Lu, L. Water-dispersible Gadolinium Oxide Nanoplates as an Effective Positive Magnetic Resonance Imaging Contrast Agent. ChemistrySelect 2022, 7, e202202062. [Google Scholar] [CrossRef]
  82. Cao, Y.C. Synthesis of Square Gadolinium-Oxide Nanoplates. J. Am. Chem. Soc. 2004, 126, 7456–7457. [Google Scholar] [CrossRef]
  83. Liu, Z.; Liu, X.; Yuan, Q.; Dong, K.; Jiang, L.; Li, Z.; Ren, J.; Qu, X. Hybrid mesoporous gadolinium oxide nanorods: A platform for multimodal Imaging and Enhanced Insoluble Anticancer Drug Delivery with Low Systemic toxicity. J. Mater. Chem. 2012, 22, 14982–14990. [Google Scholar] [CrossRef]
  84. Jegadeesan, P.; Sen, S.; Padmaprabu, C.; Srivastava, S.K.; Das, A.; Amirthapandian, S. Morphological and optical investigations on Gd2O3 nanostructures. Inorg. Chem. Commun. 2023, 150, 110493. [Google Scholar] [CrossRef]
  85. Wyantuti, S.; Fadhilatunnisa, B.; Fauzia, R.P.; Jia, Q.; Rahmani, A.A.; Irkham; Bahti, H.H. Response surface methodology box-behnken design to optimise the hydrothermal synthesis of gadolinium nanoparticles. Chin. J. Anal. Chem. 2023, 51, 100316. [Google Scholar] [CrossRef]
  86. Kang, J.-G.; Min, B.-K.; Sohn, Y. Synthesis and characterization of Gd(OH)3 and Gd2O3 nanorods. Ceram. Int. 2015, 41, 1243–1248. [Google Scholar] [CrossRef]
  87. Mulugeta, E.; Tegafaw, T.; Liu, Y.; Zhao, D.; Baek, A.; Kim, J.; Chang, Y.; Lee, G.H. Synthesis, Characterization, Magnetic Properties, and Applications of Carbon Dots as Diamagnetic Chemical Exchange Saturation Transfer Magnetic Resonance Imaging Contrast Agents: A Review. Nanomaterials 2025, 15, 542. [Google Scholar] [CrossRef]
  88. Uvarov, V.; Popov, I. Metrological characterization of X-ray diffraction methods at different acquisition geometries for determination of crystallite size in nano-scale materials. Mater. Charact. 2013, 85, 111–123. [Google Scholar] [CrossRef]
  89. Billinge, S.J.L. The rise of the X-ray atomic pair distribution function method: A series of fortunate events. Philos. Trans. R. Soc. A 2019, 377, 20180413. [Google Scholar] [CrossRef] [PubMed]
  90. Whba, F.; Mohamed, F.; Rosli, N.R.A.M.; Rahman, I.A.; Idris, M.I. The crystalline structure of gadolinium oxide nanoparticles (Gd2O3-NPs) synthesized at different temperatures via X-ray diffraction (XRD) technique. Radiat. Phys. Chem. 2021, 179, 109212. [Google Scholar] [CrossRef]
  91. Mahajan, S.V.; Dickerson, J.H. Synthesis of monodisperse Sub-3 nm RE2O3 and Gd2O3:RE3+ nanocrystals. Nanotechnology 2007, 18, 325605. [Google Scholar] [CrossRef]
  92. Rahmani, A.A.; Jia, Q.; Bahti, H.H.; Fauzia, R.P.; Wyantuti, S. Recent advances in lanthanide-based nanoparticle contrast agents for magnetic resonance imaging: Synthesis, characterization, and applications. OpenNano 2025, 21, 100226. [Google Scholar] [CrossRef]
  93. Siribbal, M.S.; Jurewicz, A.; Hassan, M.; Iqbal, S.; Hu, Z.; Uvdal, K.; Hussain, M.S.; Mathur, S.; Ilyas, S. Biocompatible Hollow Gadolinium Oxide Nanocarriers for the Transport of Bioactive Molecules to Cells. ACS Appl. Nano Mater. 2024, 7, 12408–12419. [Google Scholar] [CrossRef]
  94. Ahmad, M.Y.; Yue, H.; Tegafaw, T.; Liu, S.W.; Ho, S.L.; Lee, G.H.; Nam, S.W.; Chang, Y.M. Functionalized Lanthanide Oxide Nanoparticles for Tumor Targeting, Medical Imaging, and Therapy. Pharmaceutics 2021, 13, 1890. [Google Scholar] [CrossRef] [PubMed]
  95. Ho, S.L.; Choi, G.; Yue, H.; Kim, H.-K.; Jung, K.-H.; Park, J.A.; Kim, M.H.; Lee, Y.J.; Kim, J.Y.; Miao, X.; et al. In Vivo neutron capture therapy of cancer using ultrasmall gadolinium oxide nanoparticles with cancer-targeting ability. RSC Adv. 2020, 10, 865–874. [Google Scholar] [CrossRef]
  96. Kunc, F.; Gallerneault, M.; Kodra, O.; Brinkmann, A.; Lopinski, G.P.; Johnston, L.J. Surface chemistry of metal oxide nanoparticles: NMR and TGA quantification. Anal. Bioanal. Chem. 2022, 414, 4409–4425. [Google Scholar] [CrossRef]
  97. Kim, C.R.; Baeck, J.S.; Chang, Y.; Bae, J.E.; Chae, K.S.; Lee, G.H. Ligand-size dependent water proton relaxivities in ultrasmall gadolinium oxide nanoparticles and in vivo T1 MR images in a 1.5 T MR field. Phys. Chem. Chem. Phys. 2014, 16, 19866–19873. [Google Scholar] [CrossRef]
  98. Kattel, K.; Park, J.Y.; Xu, W.; Kim, H.G.; Lee, E.J.; Bony, B.A.; Heo, W.C.; Jin, S.; Baeck, J.S.; Chang, Y.; et al. Paramagnetic dysprosium oxide nanoparticles and dysprosium hydroxide nanorods as T2 MRI contrast agents. Biomaterials 2012, 33, 3254–3261. [Google Scholar] [CrossRef]
  99. Park, J.Y.; Kim, S.J.; Lee, G.H.; Jin, S.; Chang, Y.; Bae, J.E.; Chae, K.S. Various ligand-coated ultrasmall gadolinium-oxide nanoparticles: Water proton relaxivity and in-vivo T1 MR image. J. Korean Phys. Soc. 2015, 66, 1295–1302. [Google Scholar] [CrossRef]
  100. Schlenoff, J.B. Zwitteration: Coating surfaces with zwitterionic functionality to reduce nonspecific adsorption. Langmuir 2014, 30, 9625–9636. [Google Scholar] [CrossRef] [PubMed]
  101. Honary, S.; Zahir, F. Effect of Zeta Potential on the Properties of Nano-Drug Delivery Systems—A Review (Part 2). Trop. J. Pharm. Res. 2013, 12, 265–273. [Google Scholar]
  102. Sun, X.; Kou, B. Biocompatibility and potential anticancer activity of gadolinium oxide (Gd2O3) nanoparticles against nasal squamous cell carcinoma. BMC Biotechnol. 2024, 24, 53. [Google Scholar] [CrossRef]
  103. Frencken, A.L.; Blasiak, B.; Tomanek, B.; Kruk, D.; van Veggel, F.C.J.M. Utilizing T1- and T2-Specific Contrast Agents as “Two Colors” MRI Correlation. Materials 2025, 18, 2290. [Google Scholar] [CrossRef]
  104. Dash, A.; Blasiak, B.; Tomanek, B.; van Veggel, F.C.J.M. Validation of Inner, Second, and Outer Sphere Contributions to T1 and T2 Relaxation in Gd3+-Based Nanoparticles Using Eu3+ Lifetime Decay as a Probe. J. Phys. Chem. C 2018, 122, 11557–11569. [Google Scholar] [CrossRef]
  105. Faucher, L.; Guay-Bégin, A.-A.; Lagueux, J.; Côté, M.-F.; Petitclerc, É.; Fortin, M.-A. Ultra-small gadolinium oxide nanoparticles to image brain cancer cells in vivo with MRI. Contrast Media Mol. Imaging 2011, 6, 209–218. [Google Scholar] [CrossRef]
  106. Yin, J.; Wang, X.; Zheng, H.; Zhang, J.; Qu, H.; Tian, L.; Zhao, F.; Shao, Y. Silica Nanoparticles Decorated with Gadolinium Oxide Nanoparticles for Magnetic Resonance and Optical Imaging of Tumors. ACS Appl. Nano Mater. 2021, 4, 3767–3779. [Google Scholar] [CrossRef]
  107. Shen, Z.; Fan, W.; Yang, Z.; Liu, Y.; Bregadze, V.I.; Mandal, S.K.; Yung, B.C.; Lin, L.; Liu, T.; Tang, W.; et al. Exceedingly Small Gadolinium Oxide Nanoparticles with Remarkable Relaxivities for Magnetic Resonance Imaging of Tumors. Small 2019, 15, 1903422. [Google Scholar] [CrossRef] [PubMed]
  108. Mekuria, S.L.; Debele, T.A.; Tsai, H.-C. Encapsulation of Gadolinium Oxide Nanoparticle (Gd2O3) Contrasting Agents in PAMAM Dendrimer Templates for Enhanced Magnetic Resonance Imaging in Vivo. ACS Appl. Mater. Interfaces 2017, 9, 6782–6795. [Google Scholar] [CrossRef] [PubMed]
  109. Bridot, J.-L.; Faure, A.-C.; Laurent, S.; Rivière, C.; Billotey, C.; Hiba, B.; Janier, M.; Josserand, V.; Coll, J.-L.; Vander Elst, L.; et al. Hybrid Gadolinium Oxide Nanoparticles:  Multimodal Contrast Agents for in Vivo Imaging. J. Am. Chem. Soc. 2007, 129, 5076–5084. [Google Scholar] [CrossRef]
  110. Wang, F.; Peng, E.; Zheng, B.; Li, S.F.Y.; Xue, J.M. Synthesis of Water-Dispersible Gd2O3/GO Nanocomposites with Enhanced MRI T1 Relaxivity. J. Phys. Chem. C 2015, 119, 23735–23742. [Google Scholar] [CrossRef]
  111. Kryza, D.; Taleb, J.; Janier, M.; Marmuse, L.; Miladi, I.; Bonazza, P.; Louis, C.; Perriat, P.; Roux, S.; Tillement, O.; et al. Biodistribution Study of Nanometric Hybrid Gadolinium Oxide Particles as a Multimodal SPECT/MR/Optical Imaging and Theragnostic Agent. Bioconjug. Chem. 2011, 22, 1145–1152. [Google Scholar] [CrossRef]
  112. Ahrén, M.; Selegård, L.; Klasson, A.; Söderlind, F.; Abrikossova, N.; Skoglund, C.; Bengtsson, T.; Engström, M.; Käll, P.-O.; Uvdal, K. Synthesis and Characterization of PEGylated Gd2O3 Nanoparticles for MRI Contrast Enhancement. Langmuir 2010, 26, 5753–5762. [Google Scholar] [CrossRef]
  113. Ramalho, J.; Ramalho, M.; Jay, M.; Burke, L.M.; Semelka, R.C. Gadolinium toxicity and treatment. Magn. Reson. Imaging 2016, 34, 1394–1398. [Google Scholar] [CrossRef]
  114. Grobner, T.; Prischl, F.C. Gadolinium and nephrogenic systemic fibrosis. Kidney Int. 2007, 72, 260–264. [Google Scholar] [CrossRef]
  115. Garcia, J.; Liu, S.Z.; Louie, A.Y. Biological effects of MRI contrast agents: Gadolinium retention, potential mechanisms and a role for phosphorus. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2017, 375, 20170180. [Google Scholar] [CrossRef] [PubMed]
  116. Kloc, M.; Halasa, M.; Wosik, J.; Ghobrial, R.M. Gadolinium-based MRI contrast agent effects on calcium signaling and actin-dependent cell functions. Magn. Med. 2025, 1, 100004. [Google Scholar] [CrossRef]
  117. Saadh, M.J.; Haddad, M.; Dababneh, M.F.; Bayan, M.F.; Al-Jaidi, B.A. A Guide for Estimating the Maximum Safe Starting Dose and Conversion it between Animals and Humans. Syst. Rev. Pharm. 2020, 11, 98–101. [Google Scholar]
  118. Cheong, B.Y.C.; Wilson, J.M.; Preventza, O.A.; Muthupillai, R. Gadolinium-Based Contrast Agents: Updates and Answers to Typical Questions Regarding Gadolinium Use. Tex. Heart Inst. J. 2022, 49, e217680. [Google Scholar] [CrossRef] [PubMed]
  119. Ho, S.L.; Cha, H.; Oh, I.T.; Jung, K.-H.; Kim, M.H.; Lee, Y.J.; Miao, X.; Tegafaw, T.; Ahmad, M.Y.; Chae, K.S.; et al. Magnetic resonance imaging, gadolinium neutron capture therapy, and tumor cell detection using ultrasmall Gd2O3 nanoparticles coated with polyacrylic acid-rhodamine B as a multifunctional tumor theragnostic agent. RSC Adv. 2018, 8, 12653–12665. [Google Scholar] [CrossRef]
  120. Stoddart, M.J. Mammalian Cell Viability: Methods and Protocols; Humana Press: New York, NY, USA, 2011. [Google Scholar]
  121. Kamiloglu, S.; Sari, G.; Ozdal, T.; Capanoglu, E. Guidelines for cell viability assays. Food Front. 2020, 1, 332–349. [Google Scholar] [CrossRef]
  122. Ahmad, M.Y.; Liu, S.; Tegafaw, T.; Saidi, A.K.A.A.; Zhao, D.; Liu, Y.; Lee, S.; Kim, S.; Yang, S.H.; Hwang, D.W.; et al. Biotin-Conjugated Poly(Acrylic Acid)-Grafted Ultrasmall Gadolinium Oxide Nanoparticles for Enhanced Tumor Imaging. Eur. J. Inorg. Chem. 2023, 26, e202300430. [Google Scholar] [CrossRef]
  123. Li, X.D.; Sun, Y.H.; Ma, L.N.; Liu, G.F.; Wang, Z.X. The Renal Clearable Magnetic Resonance Imaging Contrast Agents: State of the Art and Recent Advances. Molecules 2020, 25, 5072. [Google Scholar] [CrossRef]
  124. Dai, Y.; Wu, C.; Wang, S.; Li, Q.; Zhang, M.; Li, J.; Xu, K. Comparative study on in vivo behavior of PEGylated gadolinium oxide nanoparticles and Magnevist as MRI contrast agent. Nanomed. Nanotechnol. Biol. Med. 2018, 14, 547–555. [Google Scholar] [CrossRef]
  125. Ashouri, H.; Riyahi Alam, N.; Khoobi, M.; Haghgoo, S.; Rasouli, Z.; Gholami, M. NSF evaluation of gadolinium biodistribution in renally impaired rats: Using novel metabolic Gd2O3 nanoparticles coated with β-cyclodextrin (Gd2O3@PCD) in MR molecular Imaging. Magn. Reson. Imaging 2024, 107, 120–129. [Google Scholar] [CrossRef]
  126. Sarikhani, A.; Alamzadeh, Z.; Beik, J.; Irajirad, R.; Mirrahimi, M.; Mahabadi, V.P.; Kamrava, S.K.; Ghaznavi, H.; Khoei, S. Ultrasmall Fe3O4 and Gd2O3 hybrid nanoparticles for T1-weighted MR imaging of cancer. Cancer Nanotechnol. 2022, 13, 43. [Google Scholar] [CrossRef]
  127. Xie, M.; Xu, Y.; Huang, J.; Li, Y.; Wang, L.; Yang, L.; Mao, H. Going Even Smaller: Engineering Sub-5 nm Nanoparticles for Improved Delivery, Biocompatibility and Functionality. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 2020, 12, e1644. [Google Scholar] [CrossRef] [PubMed]
  128. Wu, L.; Lu, X.; Lu, Y.; Shi, M.; Guo, S.; Feng, J.; Yang, S.; Xiong, W.; Xu, Y.; Yan, C.; et al. Kilogram-Scale Synthesis of Extremely Small Gadolinium Oxide Nanoparticles as a T1-Weighted Contrast Agent for Magnetic Resonance Imaging. Small 2024, 20, 2308547. [Google Scholar] [CrossRef]
  129. Tian, X.; Yang, F.; Yang, C.; Peng, Y.; Chen, D.; Zhu, J.; He, F.; Li, L.; Chen, X. Toxicity evaluation of Gd2O3@SiO2 nanoparticles prepared by laser ablation in liquid as MRI contrast agents in vivo. Int. J. Nanomed. 2014, 9, 4043–4053. [Google Scholar] [CrossRef]
  130. Longmire, M.; Choyke, P.L.; and Kobayashi, H. Clearance Properties of Nano-Sized Particles and Molecules as Imaging Agents: Considerations and Caveats. Nanomedicine 2008, 3, 703–717. [Google Scholar] [CrossRef]
  131. Zheng, C.; Tian, X.; Cai, J.; Huang, L.; Wang, S.; Yang, F.; Ma, Y.; Xie, F.; Li, L. In vivo immunotoxicity of Gd2O3: Eu3+ nanoparticles and the associated molecular mechanism. J. Biochem. Mol. Toxicol. 2020, 34, e22562. [Google Scholar] [CrossRef] [PubMed]
  132. Roch, A.; Gossuin, Y.; Muller, R.N.; Gillis, P. Superparamagnetic colloid suspensions: Water magnetic relaxation and clustering. J. Magn. Magn. Mater. 2005, 293, 532–539. [Google Scholar] [CrossRef]
  133. Roch, A.; Muller, R.N.; Gillis, P. Theory of proton relaxation induced by superparamagnetic particles. J. Chem. Phys. 1999, 110, 5403–5411. [Google Scholar] [CrossRef]
  134. Lauffer, R.B. Paramagnetic Metal Complexes as Water Proton Relaxation Agents for NMR Imaging: Theory and Design. Chem. Rev. 1987, 87, 901–927. [Google Scholar] [CrossRef]
  135. Carniato, F.; Thangavel, K.; Tei, L.; Botta, M. Structure and dynamics of the hydration shells of citrate-coated GdF3 nanoparticles. J. Mater. Chem. B 2013, 1, 2442–2446. [Google Scholar] [CrossRef] [PubMed]
  136. Park, J.Y.; Choi, E.S.; Baek, M.J.; Lee, G.H.; Woo, S.; Chang, Y. Water-Soluble Ultra Small Paramagnetic or Superparamagnetic Metal Oxide Nanoparticles for Molecular MR Imaging. Eur. J. Inorg. Chem. 2009, 2009, 2477–2481. [Google Scholar] [CrossRef]
  137. Guay-Bégin, A.-A.; Chevallier, P.; Faucher, L.; Turgeon, S.; Fortin, M.-A. Surface Modification of Gadolinium Oxide Thin Films and Nanoparticles Using Poly(ethylene glycol)-Phosphate. Langmuir 2012, 28, 774–782. [Google Scholar] [CrossRef]
  138. Kattel, K.; Park, J.Y.; Xu, W.; Kim, H.G.; Lee, E.J.; Bony, B.A.; Heo, W.C.; Lee, J.J.; Jin, S.; Baeck, J.S.; et al. A Facile Synthesis, In Vitro and In Vivo MR Studies of d-Glucuronic Acid-Coated Ultrasmall Ln2O3 (Ln = Eu, Gd, Dy, Ho, and Er) Nanoparticles as a New Potential MRI Contrast Agent. ACS Appl. Mater. Interfaces 2011, 3, 3325–3334. [Google Scholar] [CrossRef]
  139. Xu, W.; Park, J.Y.; Kattel, K.; Ahmad, M.W.; Bony, B.A.; Heo, W.C.; Jin, S.; Park, J.W.; Chang, Y.; Kim, T.J.; et al. Fluorescein-polyethyleneimine coated gadolinium oxide nanoparticles as T1 magnetic resonance imaging (MRI)–cell labeling (CL) dual agents. RSC Adv. 2012, 2, 10907–10915. [Google Scholar] [CrossRef]
  140. Ni, K.; Zhao, Z.; Zhang, Z.; Zhou, Z.; Yang, L.; Wang, L.; Ai, H.; Gao, J. Geometrically Confined Ultrasmall Gadolinium Oxide Nanoparticles Boost the T1 Contrast Ability. Nanoscale 2016, 8, 3768–3774. [Google Scholar] [CrossRef]
  141. Ahmad, M.W.; Ahmad, M.Y.; Ul-Islam, M.; Xu, W.; Tegafaw, T.; Wazwaz, A.A.; Ibrehem, A.S.; Khan, M.S.; Oh, I.-T.; Chae, K.S.; et al. Synthesis, MR Relaxivities, and In Vitro Cytotoxicity of 3,5-Diiodo-L-Tyrosine-Coated Gd2O3 Nanoparticles. Bionanoscience 2019, 9, 179–185. [Google Scholar] [CrossRef]
  142. Miao, X.; Yue, H.; Ho, S.L.; Cha, H.; Marasini, S.; Ghazanfari, A.; Ahmad, M.Y.; Liu, S.; Tegafaw, T.; Chae, K.-S.; et al. Synthesis, Biocompatibility, and Relaxometric Properties of Heavily Loaded Apoferritin with D-Glucuronic Acid-Coated Ultrasmall Gd2O3 Nanoparticles. Bionanoscience 2021, 11, 380–389. [Google Scholar] [CrossRef]
  143. Rahman, A.T.M.A.; Majewski, P.; Vasilev, K. Gd2O3 nanoparticles: Size-dependent nuclear magnetic resonance. Contrast Media Mol. Imaging 2013, 8, 92–95. [Google Scholar] [CrossRef]
  144. Gossuin, Y.; Hocq, A.; Vuong, Q.L.; Disch, S.; Hermann, R.P.; Gillis, P. Physico-chemical and NMR relaxometric characterization of gadolinium hydroxide and dysprosium oxide nanoparticles. Nanotechnology 2008, 19, 475102. [Google Scholar] [CrossRef] [PubMed]
  145. Marangoni, V.S.; Neumann, O.; Henderson, L.; Kaffes, C.C.; Zhang, H.; Zhang, R.; Bishnoi, S.; Ayala-Orozco, C.; Zucolotto, V.; Bankson, J.A.; et al. Enhancing T1 magnetic resonance imaging contrast with internalized gadolinium(III) in a multilayer nanoparticle. Proc. Natl. Acad. Sci. USA 2017, 114, 6960–6965. [Google Scholar] [CrossRef] [PubMed]
  146. Rinck, P.A.; Muller, R.N. Field strength and dose dependence of contrast enhancement by gadolinium-based MR contrast agents. Eur. Radiol. 1999, 9, 998–1004. [Google Scholar] [CrossRef]
  147. Wu, J. The enhanced permeability and retention (EPR) effect: The significance of the concept and methods to enhance its application. J. Pers. Med. 2021, 11, 771. [Google Scholar] [CrossRef]
  148. Hua, S.; de Matos, M.B.C.; Metselaar, J.M.; Storm, G. Current Trends and Challenges in the Clinical Translation of Nanoparticulate Nanomedicines: Pathways for Translational Development and Commercialization. Front. Pharmacol. 2018, 9, 790. [Google Scholar] [CrossRef]
  149. Đorđević, S.; Gonzalez, M.M.; Conejos-Sánchez, I.; Carreira, B.; Pozzi, S.; Acúrcio, R.C.; Satchi-Fainaro, R.; Florindo, H.F.; Vicent, M.J. Current hurdles to the translation of nanomedicines from bench to the clinic. Drug Deliv. Transl. Res. 2021, 12, 500–525. [Google Scholar] [CrossRef] [PubMed]
  150. Marasini, S.; Yue, H.; Ho, S.L.; Park, J.A.; Kim, S.; Jung, K.-H.; Cha, H.; Liu, S.; Tegafaw, T.; Ahmad, M.Y.; et al. Synthesis, Characterizations, and 9.4 Tesla T2 MR Images of Polyacrylic Acid-Coated Terbium(III) and Holmium(III) Oxide Nanoparticles. Nanomaterials 2021, 11, 1355. [Google Scholar] [CrossRef]
Figure 1. Number of publications on Gd2O3 NPs related to biomedical applications and T1 MRI CAs, based on data from (a) Scopus and (b) Web of Science, up to 20 June 2025.
Figure 1. Number of publications on Gd2O3 NPs related to biomedical applications and T1 MRI CAs, based on data from (a) Scopus and (b) Web of Science, up to 20 June 2025.
Nanomaterials 15 01340 g001
Figure 2. Synthesis schemes for hydrophilic and biocompatible ligand-coated Gd2O3 NPs: (a) one-pot polyol method, (b) one-pot DMSO method, (c) (I) thermal decomposition method yielding OA/OM-coated Gd2O3 NPs and (II) subsequent ligand exchange with hydrophilic and biocompatible ligands, and (d) one-pot hydrothermal method.
Figure 2. Synthesis schemes for hydrophilic and biocompatible ligand-coated Gd2O3 NPs: (a) one-pot polyol method, (b) one-pot DMSO method, (c) (I) thermal decomposition method yielding OA/OM-coated Gd2O3 NPs and (II) subsequent ligand exchange with hydrophilic and biocompatible ligands, and (d) one-pot hydrothermal method.
Nanomaterials 15 01340 g002
Figure 3. TEM images: (a) PAA-coated Gd2O3 NPs synthesized via the polyol method [51], (b) Gd2O3 NPs synthesized via the DMSO method [d222 indicates the lattice spacing between the (222) planes] [53], (c) OA/OM-capped Gd2O3 nanoplates synthesized via the thermal decomposition method [55], and (d) Gd(OH)3 nanorods synthesized via the hydrothermal method [56]. XRD patterns: (e) PAA-coated Gd2O3 NPs synthesized via the polyol method before (bottom spectrum) and after (top spectrum) TGA [51]. (f) Gd2O3 NPs synthesized via the DMSO method (bottom spectrum) and bulk Gd2O3 reference (top spectrum) [53]. (g) OA/OM-capped Gd2O3 nanoplates synthesized via the thermal decomposition method (top spectrum) and reference pattern with monoclinic structure (JCPDS No 43-1015) (bottom spectrum) [55]. (h) Gd(OH)3 nanorods synthesized via the hydrothermal method (top spectrum) and reference pattern with hexagonal structure (JCPDS No. 83-2037) (bottom spectrum) [56].
Figure 3. TEM images: (a) PAA-coated Gd2O3 NPs synthesized via the polyol method [51], (b) Gd2O3 NPs synthesized via the DMSO method [d222 indicates the lattice spacing between the (222) planes] [53], (c) OA/OM-capped Gd2O3 nanoplates synthesized via the thermal decomposition method [55], and (d) Gd(OH)3 nanorods synthesized via the hydrothermal method [56]. XRD patterns: (e) PAA-coated Gd2O3 NPs synthesized via the polyol method before (bottom spectrum) and after (top spectrum) TGA [51]. (f) Gd2O3 NPs synthesized via the DMSO method (bottom spectrum) and bulk Gd2O3 reference (top spectrum) [53]. (g) OA/OM-capped Gd2O3 nanoplates synthesized via the thermal decomposition method (top spectrum) and reference pattern with monoclinic structure (JCPDS No 43-1015) (bottom spectrum) [55]. (h) Gd(OH)3 nanorods synthesized via the hydrothermal method (top spectrum) and reference pattern with hexagonal structure (JCPDS No. 83-2037) (bottom spectrum) [56].
Nanomaterials 15 01340 g003
Figure 4. DLS patterns: (a) PAA-coated Gd2O3 NPs (PAA = 1200, 5100, and 15,000 Da) [75], (b) FI-PEG-BBN-coated Gd2O3 NPs [77], (c) Gd2O3@N-dodecyl-PEI-PEG nanoplate clusters [55], and (d) CA-coated hollow Gd2O3 NPs [93]. a = hydrodynamic diameter.
Figure 4. DLS patterns: (a) PAA-coated Gd2O3 NPs (PAA = 1200, 5100, and 15,000 Da) [75], (b) FI-PEG-BBN-coated Gd2O3 NPs [77], (c) Gd2O3@N-dodecyl-PEI-PEG nanoplate clusters [55], and (d) CA-coated hollow Gd2O3 NPs [93]. a = hydrodynamic diameter.
Nanomaterials 15 01340 g004
Figure 5. (a) FT-IR absorption spectra of PAA, PAA-UGNP, RGD, and RGD-PAA-UGNP (subscript labels “s” and “b” denote stretching and bending vibrations, respectively) [95]. (b) TGA curves of PEI, pimelic acid, azelaic acid, suberic acid, and sebacic acid coated-Gd2O3 NPs [99]. (c) Zeta potential (ζ) curve of PSA-coated Gd2O3 NPs [63].
Figure 5. (a) FT-IR absorption spectra of PAA, PAA-UGNP, RGD, and RGD-PAA-UGNP (subscript labels “s” and “b” denote stretching and bending vibrations, respectively) [95]. (b) TGA curves of PEI, pimelic acid, azelaic acid, suberic acid, and sebacic acid coated-Gd2O3 NPs [99]. (c) Zeta potential (ζ) curve of PSA-coated Gd2O3 NPs [63].
Nanomaterials 15 01340 g005
Figure 6. M–H curves of DEG-, TEG-, TeEG-, and PEG-coated Gd2O3 NPs at (a) T = 5 and (b) 300 K. M–T curves of (c) DEG- and TEG-coated Gd2O3 NPs, and (d) TeEG- and PEG-coated Gd2O3 NPs at H = 100 Oe (T = 2–340 K) [70].
Figure 6. M–H curves of DEG-, TEG-, TeEG-, and PEG-coated Gd2O3 NPs at (a) T = 5 and (b) 300 K. M–T curves of (c) DEG- and TEG-coated Gd2O3 NPs, and (d) TeEG- and PEG-coated Gd2O3 NPs at H = 100 Oe (T = 2–340 K) [70].
Nanomaterials 15 01340 g006
Figure 7. Inner and outer sphere water proton spin relaxation mechanisms. The dotted lines indicate Gd3+–water and NP magnetic moment–water interactions. O2− ions were omitted in Gd2O3 NP.
Figure 7. Inner and outer sphere water proton spin relaxation mechanisms. The dotted lines indicate Gd3+–water and NP magnetic moment–water interactions. O2− ions were omitted in Gd2O3 NP.
Nanomaterials 15 01340 g007
Figure 8. Plots of (a) inverse 1/T1 and (b) 1/T2 relaxation times of DEG-, TEG-, TeEG-, and PEG-coated Gd2O3 NPs dispersed in water as functions of Gd concentration at 3 T. The slopes correspond to r1 and r2 values, respectively. Effect of glycol chain length on (c) r1 and (d) r2 values [70].
Figure 8. Plots of (a) inverse 1/T1 and (b) 1/T2 relaxation times of DEG-, TEG-, TeEG-, and PEG-coated Gd2O3 NPs dispersed in water as functions of Gd concentration at 3 T. The slopes correspond to r1 and r2 values, respectively. Effect of glycol chain length on (c) r1 and (d) r2 values [70].
Nanomaterials 15 01340 g008
Figure 9. In vitro cell viabilities of (a) PAA-Gd2O3 NPs in HEK293 and HepG2 cell lines after 48 h of incubation [122], (b) PAAMA-coated Gd2O3 NPs in DU145 and NCTC1469 cell lines after 48 h of incubation [61], and (c) Gd2O3-FI-PEG-BBN and Gd2O3-FI-PEG in PC-3 cell line after 24 h (left) and 48 h (right) of incubation [77].
Figure 9. In vitro cell viabilities of (a) PAA-Gd2O3 NPs in HEK293 and HepG2 cell lines after 48 h of incubation [122], (b) PAAMA-coated Gd2O3 NPs in DU145 and NCTC1469 cell lines after 48 h of incubation [61], and (c) Gd2O3-FI-PEG-BBN and Gd2O3-FI-PEG in PC-3 cell line after 24 h (left) and 48 h (right) of incubation [77].
Nanomaterials 15 01340 g009
Figure 10. Histological analysis of the liver, kidney, spleen and heart 30 days after IV administration of Gd2O3@BSA NPs into mice tails (scale bar: 50 μm; number of mice, N = 4) [102].
Figure 10. Histological analysis of the liver, kidney, spleen and heart 30 days after IV administration of Gd2O3@BSA NPs into mice tails (scale bar: 50 μm; number of mice, N = 4) [102].
Nanomaterials 15 01340 g010
Figure 12. (a) Mean fluorescence intensity (MFI) obtained via flow cytometry to evaluate toxicity on innate immune markers (i.e., CD11b, CD206, CD40, CD80, and CD86) in monocytes/macrophages from peripheral blood 48 h after injection of PBS (100 µL, negative control), Gd-DTPA (15 μmol/kg), and Gd-NPs (15 μmol/kg) in BALB/c mice (N = 5 per group). (b) Serum concentrations of cytokines such as IL-1β, IL-2, and IL-4. (c) MFI of the adaptive immune marker CD69 in lymphocytes from peripheral blood and lymph nodes, and ROS levels in neutrophils from peripheral blood. Statistical significance determined by t-test (* p < 0.05) [129].
Figure 12. (a) Mean fluorescence intensity (MFI) obtained via flow cytometry to evaluate toxicity on innate immune markers (i.e., CD11b, CD206, CD40, CD80, and CD86) in monocytes/macrophages from peripheral blood 48 h after injection of PBS (100 µL, negative control), Gd-DTPA (15 μmol/kg), and Gd-NPs (15 μmol/kg) in BALB/c mice (N = 5 per group). (b) Serum concentrations of cytokines such as IL-1β, IL-2, and IL-4. (c) MFI of the adaptive immune marker CD69 in lymphocytes from peripheral blood and lymph nodes, and ROS levels in neutrophils from peripheral blood. Statistical significance determined by t-test (* p < 0.05) [129].
Nanomaterials 15 01340 g012
Figure 14. (a) In vivo T1 MR images of a mouse at 3 T before (“Pre”) and after IV administration of the solution sample via the mouse tail. Small circles indicate ROIs, and dotted circles indicate the liver and kidneys (administration dose = ~0.1 mmol Gd/kg) and (b) SNR plots of the ROIs in the liver and kidneys as a function of time [58]. (c) In vivo coronal and axial T1 MR images of the (I) liver, (II) kidneys, and (III) bladder of mice before (labeled as pre) and after IV injection of Gd2O3@C NPs (administration dose = ~0.1 mmol Gd/kg) and (d) SNR-ROI plots as a function of time [69]. The red circles in MR images indicate ROIs.
Figure 14. (a) In vivo T1 MR images of a mouse at 3 T before (“Pre”) and after IV administration of the solution sample via the mouse tail. Small circles indicate ROIs, and dotted circles indicate the liver and kidneys (administration dose = ~0.1 mmol Gd/kg) and (b) SNR plots of the ROIs in the liver and kidneys as a function of time [58]. (c) In vivo coronal and axial T1 MR images of the (I) liver, (II) kidneys, and (III) bladder of mice before (labeled as pre) and after IV injection of Gd2O3@C NPs (administration dose = ~0.1 mmol Gd/kg) and (d) SNR-ROI plots as a function of time [69]. The red circles in MR images indicate ROIs.
Nanomaterials 15 01340 g014
Figure 15. (a) In vivo axial T1 MR images of U87MG tumor-bearing nude mice acquired at 1.5 T after IV injection of ES-GON5-PAA@RGD2 NPs (arrows indicate the tumor). (b) ΔSNR in the tumor at various time points. The injection Gd dosage was 5.0 mgkg−1 [107].
Figure 15. (a) In vivo axial T1 MR images of U87MG tumor-bearing nude mice acquired at 1.5 T after IV injection of ES-GON5-PAA@RGD2 NPs (arrows indicate the tumor). (b) ΔSNR in the tumor at various time points. The injection Gd dosage was 5.0 mgkg−1 [107].
Nanomaterials 15 01340 g015
Table 1. Summary of synthesis methods with associated advantages and disadvantages.
Table 1. Summary of synthesis methods with associated advantages and disadvantages.
MethodSolventAdvantageDisadvantageRef.
PolyolPolyol such as ethylene glycol, diethylene glycol, triethylene glycol, polyethylene glycol
surface modification in one-pot synthesis
ultrasmall NPs
small-scale synthesis
poor crystallinity
[51,52]
DMSODMSO
low temperature synthesis (<60 °C)
surface modification in one-pot synthesis
ultrasmall NPs
small-scale synthesis
poor crystallinity
[53]
Thermal decompositionHigh boiling point organic solvent such as oleic acid, oleylamine
high crystallinity.
monodisperse NPs with particle size control
expensive (organic solvent, inert gas)
post synthesis surface modification
solvent waste
small-scale synthesis
[55]
HydrothermalWater
environmentally friendly
large-scale synthesis
requires autoclave
[56]
Table 2. Various characterization techniques and information.
Table 2. Various characterization techniques and information.
Characterization TechniqueInformation
TEMParticle size, morphology
ICP-AESGd concentration in aqueous solution sample
XRDCrystal structure
DLSHydrodynamic diameter
FT-IRSurface coating
TGASurface-coating amount
Zeta potentialSurface charge
VSM, MPMSMagnetic properties
MRIr1, r2 values
MTT, CCK-8, WST-1in vitro cellular toxicity
Table 3. r1 and r2 values of surface-coated Gd2O3 NPs under various conditions, including coating ligands, particle size, hydrodynamic diameter, applied magnetic field, and temperature.
Table 3. r1 and r2 values of surface-coated Gd2O3 NPs under various conditions, including coating ligands, particle size, hydrodynamic diameter, applied magnetic field, and temperature.
NPSynthesis MethodLigandSize (nm)r1
(s−1mM−1)
r2
(s−1mM−1)
H (tesla)T (°C)Ref.
TEMDLS
Gd2O3polyolSuccinic acid1.34.1112.515.41.522[71]
Glutaric acid1.34.151313.2
Terephthalic acid1.34.1911.614.4
PEI-13001.312.7189.1
PEI-10,0001.313.875.17.6
D-glucuronic acid2.4 4.2527.11
PAA-1200210.35.636.61.522[75]
PAA-5100211.14.731
PAA-15,000211.3429.2
PEGD-6000.9 14.1819.151.522[136]
D-glucuronic acid1 12.5612.95
Lactobionic acid0.9 11.5713.38
D-glucuronic acid149.910.51.522[64]
DEG 17 ± 22.12.81.525[137]
PEG-phosphate (reaction in water) 152 ± 201014.2
PEG-phosphate (reaction in ethanol) 68 ± 1111.414.5
PEG-polysiloxane2.23.38.811.4725[109]
3.85.28.828.8
4.68.94.428.9
D-glucuronic acid2.4 4.2527.111.522[138]
Fluorescein-polyethyleneimine3.927.56.7620.271.522[139]
PEG1.39.116.217.70.47 [73]
14.217.21.41
10.915.97
10.417.211.7
MSN2.362.2545.0848.810.5 [140]
16.95 7
D-Glucuronic acid1.36.224 ± 260 ± 51.522[97]
PEGD-2509.75 ± 155 ± 2
PEGD-60012.10.1 ± 0.110 ± 1
3,5-Diiodo-L-tyrosine2 9.2438.271.5 [141]
Carbon3.118.916.2624.121.522[69]
PAA510026.33137.41.522[51]
Polyaspartic acid212.719.153.7322[63]
Apoferritin-D-Glucuronic acid1.929.18.715.71.522[142]
PVP2.5 10.2814.473 [68]
DEG13 1.1413.53 [70]
TEG16 2.616.4
TeEG19 3.9922.6
PEG21 5.7528.7
Dextran1.512.412.229.31.522[66]
Poly(acrylic acid-co-maleic acid) (PAAMA)1.8940.663.4322[61]
Poly(methyl vinyl ether-alt-maleic acid) (PMVEMA)1.919.836.274322[58]
PEG<32.89.413.41.519.5[72]
DEG<3 6.415.2
PEG<40 0.17.6
Gd2O3DMSO 4.75–106.97.91.41 [53]
5(6)-carboxyfluorescein-polyethylene glycol-bombesin52.390.64.23 3 [77]
TEG based carboxylate ligand4136.49.11.41 [76]
Gd2O3Thermal decompositionPVP2.915.712.12333.184736[78]
N-Dodecyl-PEI-PEG9.19514.13 9.4 [55]
PAA-OA227.847.282.41.41 [80]
528.145.675.1
831.233.770.0
1133.032.467.3
2247.121.132.4
poly (maleic anhydride-alt-1-octadecene) polymer10 × 1.13216.95 1.5 [81]
CTX-PEG-TETT9.09 8.41 7 [79]
Gd2O3HydrothermalHyaluronic acid105 6 3 [46]
CA100 ± 20233.7 ± 101.85.31.4140[93]
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

Mulugeta, E.; Tegafaw, T.; Liu, Y.; Zhao, D.; Chen, X.; Baek, A.; Kim, J.; Chang, Y.; Lee, G.H. Current Status and Future Aspects of Gadolinium Oxide Nanoparticles as Positive Magnetic Resonance Imaging Contrast Agents. Nanomaterials 2025, 15, 1340. https://doi.org/10.3390/nano15171340

AMA Style

Mulugeta E, Tegafaw T, Liu Y, Zhao D, Chen X, Baek A, Kim J, Chang Y, Lee GH. Current Status and Future Aspects of Gadolinium Oxide Nanoparticles as Positive Magnetic Resonance Imaging Contrast Agents. Nanomaterials. 2025; 15(17):1340. https://doi.org/10.3390/nano15171340

Chicago/Turabian Style

Mulugeta, Endale, Tirusew Tegafaw, Ying Liu, Dejun Zhao, Xiaoran Chen, Ahrum Baek, Jihyun Kim, Yongmin Chang, and Gang Ho Lee. 2025. "Current Status and Future Aspects of Gadolinium Oxide Nanoparticles as Positive Magnetic Resonance Imaging Contrast Agents" Nanomaterials 15, no. 17: 1340. https://doi.org/10.3390/nano15171340

APA Style

Mulugeta, E., Tegafaw, T., Liu, Y., Zhao, D., Chen, X., Baek, A., Kim, J., Chang, Y., & Lee, G. H. (2025). Current Status and Future Aspects of Gadolinium Oxide Nanoparticles as Positive Magnetic Resonance Imaging Contrast Agents. Nanomaterials, 15(17), 1340. https://doi.org/10.3390/nano15171340

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