Introduction
Magnetic nanoparticles (MNPs) have emerged as versatile and highly effective contrast agents in nuclear magnetic resonance imaging (MRI) and have played a pivotal role in advancing this crucial medical imaging tool. Globally, over one hundred million MRI procedures are conducted annually, one-third of which employ magnetic contrast agents to enhance imaging precision and sensitivity [1]. The synthesis, functionalization, and evaluation of MNPs as MRI contrast agents have attracted considerable attention [2, 3]. As researchers have investigated the properties of MNPs, their structural control has played a crucial role in their ongoing evolution and refinement to improve their MRI contrast power. In T1-weighted MRI, under the magnetic field strength range of medical and research scanners, the inorganic structure of MNP contrast agents has a larger influence on their r2 values (transverse relaxivity) than their r1 values (longitudinal relaxivity) [4]. Moreover, reducing the structure-sensitive r2 value can suppress T2 effects (negative contrast) and increase the essential r1/r2 ratio, which will enhance the T1 effect (positive contrast) of the MNPs [5]. In this regard, both computational and experimental methods have been developed for structural control to optimize MNPs and unlock their full capabilities. Computational methods, such as modeling simulations, have the advantage of easily changing the MNP structures, including varying the MNP lattice or size parameters, to predict their T1 or T2 contrast power. After changing the synthesis conditions to produce MNPs of the desired size, an effective structural analysis tool is necessary to confirm that the as-synthesized MNPs have the targeted structures needed to obtain the ideal contrast power. In this regard, high-sensitivity and high-resolution X-ray absorption fine structure (XAFS) spectroscopy can identify both the electronic and geometric structures of materials [6,7], revealing the local atomic environment to shed light on the bond lengths and angles, coordination numbers, and local symmetry [8, 9]. Therefore, it is crucial to combine modeling simulations and XAFS spectroscopy to achieve structural control of MNPs and achieve the best MRI performance.
From a computational perspective, analytical and simulation models have been established to predict the optimal structure of MNPs processing an ideal MRI contrast power depending on their specific applications [10-13]. Using these models, only the production and evaluation of a narrower set of MNPs are required to save a large amount of energy and materials. To obtain statistically reliable results, the time-domain step size of the simulation is set to be proportional to the square of the MNP diameter (Eq. 1 and Eq. 5); in other words, the corresponding simulation time (the reciprocal of the step size) for small MNPs is much higher than that for large MNPs (e.g., 1-nm MNPs take 102 times longer than 10-nm MNPs). Consequently, it is difficult for traditional algorithms and computing resources to simulate MNPs smaller than 5 nm because of their formidable simulation time, which can be on the scale of months. In biomedical applications, small MNPs are particularly important, with a size of <5.5 nm being a critical criterion [14, 15] because MNPs smaller than this can be rapidly cleared through the kidney route in humans and mammals after intravenous injections, diminishing potential post-imaging side effects. Hence, there is a pressing need for a faster simulation method to accurately predict the r1 or r2 values of MNPs smaller than 5 nm. After simulation and synthesis, verifying the structure of these super-small MNPs poses another challenge to characterization tools [16]. Fortunately, the XAFS technique, using an intense and tunable synchrotron radiation X-ray source, is ideal for determining the structures of metallic compounds at a high-resolution Angstrom (Å) scale [17-19]. So far, the XAFS technique has been used to investigate superparamagnetic iron oxide nanoparticles (SPIONs) with a size of >8 nm [20, 21] because their stable magnetization and potentially minimized toxicity make SPIONs the ideal type of MNPs for biomedical MRI applications [22]. However, it has traditionally been challenging to synthesize monodisperse SPIONs smaller than 8 nm while satisfactorily maintaining their crystal structure and superparamagnetism [23]. Therefore, the availability of these smaller SPIONs has limited their use in MRI applications in the past. More recently, small SPIONs have become increasingly popular due to their excellent T1 contrast power [24, 25], whose bright signals are preferred by MRI radiologists over the dark signals of T2 contrast from large SPIONs. Nevertheless, the magnetic and structural properties of these newly synthesized small SPIONs are less characterized, and the structural knowledge of < 8 nm SPIONs revealed by the XAFS technique is lacking. Because the XAFS technique is a powerful tool for accurately determining the local atomic and electronic structures of various materials, a comprehensive investigation of SPIONs smaller than 8 nm using XAFS is highly needed.
Herein, we report a novel two-pronged computational and experimental strategy that combines MATLAB-based simulations and XAFS analysis to develop small SPIONs best suited for MRI. Our MATLAB-based simulation method is empowered by a swift matrix operation algorithm [26] and parallel computing ability [27] to achieve high-speed prediction of the MRI contrast power for SPIONs smaller than 8 nm. We simulated 1-, 2-, 3-, 6-, and 8-nm SPIONs and obtained statistically reliable results within one week, showing that SPIONs with inorganic diameters between 1 and 3 nm outperformed their larger counterparts. To experimentally validate our predictions, we synthesized small SPIONs, followed by the characterization and evaluation of their magnetic properties, MRI contrast powers (r1 and r2 values), and MR imaging performances in vivo in mice. Furthermore, we demonstrated that XAFS using a synchrotron radiation X-ray source could successfully reveal the atomic and electronic structures of these small SPIONs, thereby explaining their magnetic origins. Overall, the as-synthesized SPIONs exhibited attractive structural, magnetic, and imaging characteristics for use in biomedical MRI applications. On the one hand, the computational results could guide and facilitate MNP synthesis through simulation; on the other hand, the experimental XAFS measurements could reveal the MNP structures and underlying mechanisms of their outstanding properties. Thus, our two-pronged computational and experimental strategy provides a convenient and rigorous platform for developing future MRI probes.
MRI simulation theory
Computational MRI studies include analytical and simulation models. Analytical models [10, 11, 28] are derived from theoretical MRI equations by applying reasonable approximations (e.g., strong magnetic field, long blood vessel, and low iron concentration approximations) under specific circumstances. These analytical models can directly calculate the r1 or r2 values of MNPs without any simulation; thus, the calculation processes are fast. However, the r1 or r2 results provided by analytical models are usually a single number without accurate statistical information, such as the standard deviation calculated from numerical simulations. Moreover, some of the above approximations cannot be applied to all types of MNPs in general. In particular, newly structured MNPs will have even higher discrepancies than conventional MNPs. Consequently, simulative models that follow the MRI theory without any major approximation were established. Simulation models [12, 13, 29, 30] utilize modern computers to accurately calculate the r1 or r2 value of MNPs, which often yield statistically reliable results. The established principles of these MNP simulation models [31, 32] are described below.
In biological media, the movement of water molecules follows the pattern of Brownian motions, where they randomly walk in a three-dimensional (3-D) space and generate proton trajectories. The step size of this 3-D random walk of water molecules,
In the real space in which our world resides, only the real part of Eq. (7) is observed for the jth proton (i.e., the projection of the complex vector in Eq. (7) into the real space, which is
To obtain statistically reliable results, we averaged the outcomes from the simulations of many MNP configurations and proton trajectories (i.e., spin trajectories), and obtained the following:
Because MRI scans are completed in real space, the following is true:
Materials and methods
Chemicals and characterizations
Unless otherwise indicated, all of the chemicals and solvents were purchased from Thermo Fisher Scientific and used as received. The air-sensitive materials were handled under a dry nitrogen atmosphere. Transmission electron microscopy (TEM) images of the MNPs were obtained at an electron acceleration voltage of 200 kV. Magnetization measurements were performed on magnetic property measurement system superconducting quantum interference device (SQUID) between –6 and +6 Tesla at 298 K using the direct current scanning mode. An X-ray diffraction (XRD) analysis was performed using a PANalytical X’Pert Pro diffractometer.
Synthesis of SPIONs and ligand exchange
First, 900 mg of Fe(oleate)3 and 190 μL oleic acid were added to a 5-mL mixture of 1-tetradecene and 1-hexadecene, which was degassed in an N2 atmosphere. Next, the reaction mixture was swiftly heated to 250–280 °C (depending on the targeted SPION diameter) and incubated at this temperature for 1 h to form SPIONs. Subsequently, the reaction mixture was cooled to room temperature before the addition of 113 mg of (CH3)3NO. The reaction mixture was again heated to 130 °C and kept at this temperature for 1 h to ensure that the SPIONs became high-quality maghemite crystals. The SPION product was purified through precipitation by adding an acetone and ethanol mixture and centrifuged at 10,000 g for 10 min. Finally, the supernatant containing the byproducts was discarded, and the hydrophobic SNIO pellets were redispersed in 3 mL of hexane. To generate water-soluble SPIONs, further ligand exchange with 2-[2-(2-methoxyethoxy)ethoxy]acetic acid (MEAA) and zwitterionic dopamine sulfonate (ZDS) was performed following our established protocols [34].
MRI simulation setup
The central unit cell, which contained protons and MNPs, had an edge length that was 100 times the MNP diameter. In this case, the volume ratio of MNPs to unit cells (i.e., the metal ion concentration) was kept constant, making the R2 results for MNPs with different sizes comparable. Protons were allowed to randomly walk using the Monte Carlo algorithm inside the central unit cell, and the periodic boundary condition (PBC) was applied. When a proton tried to escape from one side of the central cell, it returned from the opposite side of the central cell. The
In this work, we selected the following parameters: MNP diameter
X-ray absorption spectroscopy and simulation
XAFS data were acquired at the iron’s K-edge around 7120 eV. The storage ring was operated at an energy of 3.0 GeV and a current of 500 mA at beamline 7-BM of the National Synchrotron Light Source II, Brookhaven National Laboratory. A channel-cut (111) Si monochromator was employed to ensure monochromatic synchrotron radiation. Higher harmonics were effectively eliminated by adjusting the second monochromator crystal. XAFS spectra were gathered in the transmission mode, with samples and references prepared as powders spread onto adhesive tape, folded, and affixed to the sample stage. The extended XAFS (EXAFS) analysis proceeded as follows: initially, the amplitude factor (S02) was determined by fitting both the sample and reference data. Critical parameters such as a single correction to the photoelectron energy origin (∆E0), distances (R), and their respective mean square disorder (σ2) values for Fe-O and Fe-Fe bonds were adjusted during the fit. The fitting ranged from 2 to 12 Å-1 for k values and 1 to 4 Å for R values, with a fitting k weight of three. The bond lengths and coordination numbers of the samples were obtained from the EXAFS analysis. The annotations are defined as follows: T = tetrahedral site, O = octahedral site, N = coordination number around the Fe atom, and dAB = the bond length between atom A and atom B. For a 1-nm SPION: Fe-O (T), N = 4, dFe-O = 1.9 Å; Fe-O (O), N = 6, dFe-O = 2.0 Å; Fe-Fe (T-T), N = 4, dFe-Fe = 3.7 Å; Fe (T-O), N = 6.8, dFe-Fe = 3.5 Å; and Fe (O-O), N = 6, dFe-Fe = 3.0 Å. For bulk maghemite: Fe-O (T), N = 4, dFe-O = 2.1 Å; Fe-O (O), N = 6, dFe-O = 1.9 Å; Fe-Fe (T-T), N = 4, dFe-Fe = 3.7 Å; Fe (T-O), N = 8.3, dFe-Fe = 3.5 Å; and Fe (O-O), N = 6, dFe-Fe = 3.0 Å.
The XAFS simulation was performed on the Nano Analytics I using FEFF9 software purchased from the Department of Physics at the University of Washington. The crystal structure data (CIF files) were obtained from the Crystallography Open Database (https://www.crystallography.net/cod/). Nanosphere, nanocube, and nanorod structures are often found in a nanoscale synthesis and were generated by modifying the original CIF files using Avogadro software, an open-source molecular builder, and a visualization tool (http://avogadro.cc/). The XAFS spectra around iron’s K-edge were acquired after ab initio multiple scattering calculations of 1-nm maghemite crystals with different shapes using the FEFF9 software. The simulation parameters were obtained from a maghemite reference file generated by the Demeter 0.9.26 software (written by Dr. Bruce Ravel): amplitude factor S02 = 1.0, the maximum of Rpath = 5.0, space group = Fd3m, cell parameters a = b = c = 8.33 Å, and lattice angles α = β = γ = 90°. These parameters were kept constant for all three nanostructures with different shapes.
Magnetic resonance relaxivity and imaging experiments
Relaxivity measurements were performed using a Bruker® Minispec MQ60 time-domain nuclear magnetic resonance (NMR) spectrometer at 1.5 Tesla, and the results of three replicated measurements were averaged. The iron concentration (CFe) of the SPIONs was determined using a bathophenanthroline assay, following our previously published protocol [35]. Before imaging experiments, mice were first anesthetized by inhalation of a mixture of 98% oxygen and 2% isoflurane. Subsequently, the mice were monitored using a 7 Tesla Bruker® MRI scanner for a baseline period lasting 10 min. Next, a bolus containing 100 µL of 1-nm SPION solution at CFe = 50 mM was administered intravenously via the mouse tail vein under continuous scanning. The pulse sequence and scanning parameters were in accordance with the previously established conditions [36]. All of the mouse experiments were conducted in accordance with the approved institutional protocols of the University Medical Center, Hamburg-Eppendorf.
Results and discussion
The framework of the modeling process is shown in Fig.1a, where the configuration of the iron oxide MNPs and the movement trajectory of a proton (the hydrogen nucleus from a water molecule) were randomly generated by MATLAB using a Monte Carlo algorithm. This simulation process needs to be repeated numerous times, and hundreds or even thousands of randomized MNP configurations and proton trajectories should be considered to produce statistically reliable averaged R2 results [32]. The equation in the inset of Fig. 1a shows how the z-axis component of the magnetic field experienced by the proton (Bz) is calculated in a random scenario (Eqs. (2) and (3)). It can be seen that Bz depends on the root-mean-squared angular frequency shift at the MNP surface (Δωr, a material-specific quantity) and is particularly sensitive to the MNP radius to its third power (
-202410/1001-8042-35-10-005/alternativeImage/1001-8042-35-10-005-F001.jpg)
The high sensitivity of Bz to the MNP radius is illustrated in Fig. 1b, where a smaller MNP generates a smaller magnetic field, depicted by light contour lines, while a larger MNP creates a larger magnetic field, portrayed by dense contour lines. When a proton travels the same movement distance (or for the same time period) around smaller MNPs compared with larger MNPs, the changing speed of the total magnetic field perturbation from MNPs (Bt) caused by small MNPs will be slower. However, because it is the z-axis component, Bz, that affects the MRI signals [37], the changing speed of tilt angle θ must also be considered. Under the same movement distance, the changing speed of tilt angle θ is faster for smaller MNPs than large MNPs, unlike the trend for Bt. Therefore, numerical simulation studies are necessary to determine the behavior of small MNPs. We wrote innovative MATLAB code that utilized the rapid operation of matrices and parallel computing ability, finishing the simulation of several thousand plausible scenarios within a week to obtain Figs. 1c–d. The simulated MRI signals vs. echo times from the two iron oxide MNP diameters (DNP = 1 and 3 nm) were plotted and linearly fitted in Fig. 1c, where R2 was the slope of each fitted line. These R2 values were statistically reliable and satisfied the goodness of fit criterion (>0.99). Because a reduced R2 suppresses the T2 effect of MRI and makes it preferred for T1-w MRI [38], 1-nm iron oxide MNPs (SPIONs) with a lower R2 were expected to perform better than 3-nm SPIONs in T1-w MRI. In addition to the SPION diameter, we could vary other SPION structural parameters, such as their crystal type [39, 40]. Figure 1d shows the complete data set of SPIONs with different sizes and crystal types (maghemite: Fe3O4 and magnetite: γ-Fe2O3). The 1–3-nm SPIONs possessed R2 values that were roughly five–ten times smaller than the R2 values of the 6–8-nm SPIONs. Figure 1d also shows that the maghemite SPIONs generally outperformed their magnetite counterparts in T1-w MRI applications. Overall, the simulation results in Fig. 1 demonstrate that the facile control of SPION structures in modeling simulations can guide the targeted synthesis and characterization of small maghemite SPIONs, saving significant time and materials in experiments.
Because the optimal SPIONs predicted by our Monte Carlo simulations were between 1 and 3 nm in size, we proceeded with the experimental synthesis of 1-, 2-, and 3-nm SPIONs. In the synthetic route shown in Fig. 2a, the Fe(oleate)3 precursors underwent homogeneous, timed, and temperature-controlled thermal decomposition in tailored solvents. Immediately after synthesis, the oleic acid-coated SPIONs were initially hydrophobic. To make them water-soluble and biocompatible, we performed ligand exchanges with 2-[2-(2-methoxyethoxy)ethoxy]acetic acid (MEAA) and zwitterionic dopamine sulfonate (ZDS) ligands in succession following our established protocols [34]. The TEM images shown in Fig. 2b and 2c confirmed the diameters of the 2- and 3-nm SPIONs and their satisfactory dispersity. Because iron is a light-metal element with moderate contrast under TEM, their extremely small diameters made it difficult to clearly visualize the 1-nm SPIONs; instead, their diameters were confirmed using small-angle X-ray scattering (SAXS) [41]. Moreover, our SQUID measurements compared the magnetizations of the 1-, 2-, and 3-nm SPIONs with commercial MRI contrast agents, including Feraheme® (6-nm SPION cores) and Magnevist® (Gd-DTPA). As shown in Figure 2D, the 1-, 2-, and 3-nm SPIONs had lower magnetizations under the same magnetic field strength. Because a low magnetization generally translates to a small r2 value and suppresses T2 effects in MRI [42], the 1–3-nm SPIONs are expected to perform better in T1-w MRI, validating our previous computational predictions. In the enlarged view of Fig. 2d, we quantified the magnetization change (|ΔM4T→6T|%) percentages of all the magnetic substances and found that the 1-nm SPIONs displayed the strongest unsaturated magnetization among the SPIONs. We calculated the surface metal atomic ratios of all the magnetic substances to further explain this phenomenon, which is usually caused by superficial paramagnetic atoms on SPION surfaces [43]. The results are listed in Table 1, which shows that the 1-nm SPIONs possessed the highest surface ratio (80.2%) among the SPIONs. This was consistent with the fact that they had the strongest unsaturated magnetization (44.7%), as shown in Fig. 2d, and would be beneficial for T1-w MRI. Therefore, our synthetic and magnetic characterization data aligned with the computational predictions, rendering it more convincing that 1-nm SPION was the most promising candidate for T1-w MRI applications.
-202410/1001-8042-35-10-005/alternativeImage/1001-8042-35-10-005-F002.jpg)
Species | Core diameter (nm) | Surface metal atoms ratio |
---|---|---|
Feraheme® | 6.0 | 19.4 (Fe%) |
3-nm SPIONs | 3.0 | 36.2 (Fe%) |
2-nm SPIONs | 2.0 | 50.4 (Fe%) |
1-nm SPIONs | 1.0 | 80.2 (Fe%) |
Magnevist® | 0.19 | 100. (Gd%) |
To reveal the relationship between the SPION structures and their magnetic properties, we carefully examined the as-synthesized SPIONs using a combination of XRD and XAFS techniques. In Fig. 3a, it can be observed that all the SPIONs generally matched the three major XRD peaks of the maghemite standard at 2θ = 35.2°, 41.6°, and 63.3°, based on our measurements. Moreover, the SPION XRD peaks were significantly different from the three major XRD peaks for hematite (adopted from the reference [44]) at 2θ = 33.2°, 35.8°, and 49.5°. However, when the SPION diameter decreased from 3 nm to 1 nm, the XRD peaks broadened, and their signal-to-noise ratio also decreased; thus, it became difficult to determine the exact structure of SPIONs smaller than 3 nm. This is because the number of periodic crystal planes diminished with the decrement of nanoparticle diameter [16], leading to a limited number of lattice unit cells within a single SPION and weaker XRD signals. Therefore, we utilized an intense, energy-tunable synchrotron X-ray source [17] to investigate the detailed atomic structure of the 1-nm SPION. The EXAFS full spectra (6900–7800 eV) of the 1-nm SPION and bulk minerals are shown in Figure 3B. Furthermore, based on detailed analyses of the above EXAFS data, the radial distribution functions in Figure 3C showed that the 1-nm SPION was similar to maghemite between 2.0 Å and 3.5 Å but was remarkably different from hematite. Moreover, the XANES results shown in Fig. 3e confirmed the similarity between the 1-nm SPION and maghemite (7135–7150 eV), while validating the difference between the 1-nm SPION and hematite at 7130 eV. Because maghemite is a highly magnetic and stable form of iron oxide minerals [39], the EXAFS and XANES results explained the magnetism of the 1-nm SPIONs. To further unveil the geometry of our 1-nm SPIONs, considering the fact that iron is a light-metal atom and TEM experiments usually struggle to visualize them on a single nanometer scale, we used the FEFF9 software [45] to perform XANES simulations on three representative shapes (Fig. 3d) commonly found in nanoscale synthesis. As shown in Fig. 3f, the overall trend of the nanosphere simulated spectrum was similar to that of the 1-nm SPION experimental XANES in Fig. 3e, and the reasons for this good match were multifold. First, the FEFF9 software uses an ab initio, self-consistent, and real-space multiple scattering approach for rigorous simulation [46], with polarization dependence, core-hole effects, and local field corrections. The FEFF9 calculation also uses an all-electron, real-space, and relativistic Green’s function formalism with no symmetry requirements, leading to a high level of accuracy [47]. Second, multiple sources of experimental evidence, including the results of XRD and XAFS studies, have confirmed that our 1-nm SPION has a maghemite-like structure. Finally, our refined SPION synthesis technique involves the thermal decomposition of the iron precursors at the boiling point of the chosen solvent with ample stirring, ensuring a highly homogeneous reaction temperature throughout the reaction vessel without any temperature gradient. This isotropic and uniform reaction environment naturally leads to a symmetric shape for the SPION products [35, 36], which is spherical. Therefore, the as-proposed maghemite nanosphere represents the geometry of our 1-nm SPIONs, resulting in a sound match between the maghemite nanosphere FEFF9 simulation result and the experimental XANES of the 1-nm SPIONs. Not surprisingly, the simulated spectra of nanocubes and nanorods were generally different from the experimental XANES spectra for the 1-nm SPIONs; in particular, the first white-line peak of the simulated spectrum of the nanocube at 7120 eV (or the simulated spectrum of the nanorod at 7140 eV) significantly differed from the first experimental white-line peak of the 1-nm SPIONs. The XAFS and XRD results of comprehensive structural studies shown in Fig. 3 verified that our SPIONs maintained their maghemite structure and spherical shape even at 1 nm, indicating the capability of our rational synthesis to target a selected SPION conformation and formula.
-202410/1001-8042-35-10-005/alternativeImage/1001-8042-35-10-005-F003.jpg)
To demonstrate the clinical potential of our SPIONs, we evaluated their magnetic resonance (MR) properties using an MRI scanner and compared the results with those of the commercial MRI contrast agents ferumoxytol and Gd-DTPA [36, 48, 49]. In line with our simulation findings, the in vitro MR data in Fig. 4a showed that the T2 relaxation times of the small SPIONs were higher than those of the larger SPIONs, minimizing the corresponding r2 relaxivities [r1,2 = (CM·T1,2)-1 = (T1,2)-1, when metal ion concentration CM = 1 mM in our case]. A lower r2 relaxivity, in turn, suppresses the T2 effects in MRI [38], which can be quantified by further considering the r1 values and calculating the r1/r2 ratios, as displayed in Fig. 4b. It can be seen that the r1/r2 ratio of our 1-nm SPIONs was 0.91, which even approached the r1/r2 ratio of Gd-DTPA, a molecular MRI contrast agent used in clinics [50]. Because the r1/r2 ratio is a crucial criterion and elevated r1/r2 ratios are desirable for T1-w MRI [51], our 1-nm SPIONs would be better than the 3-nm SPIONs, ferumoxytol, and other SPION-based contrast agents reported to date [52] for T1-w MRI applications. Moreover, as illustrated in Fig. 4c, in vivo MRI experiments with mice made it possible to evaluate the in vivo contrast power of the 1-nm SPIONs. The contrast-enhanced MR angiography data in Fig. 4d demonstrated that our 1-nm SPIONs possessed satisfactory T1 contrast power in vivo, with the ability to highlight major blood vessels and capillaries, along with the heart, kidneys, and bladder. This result also confirmed the kidney clearance of 1-nm SPIONs, an essential pharmacokinetic property in clinical applications to clear unbound SPIONs after the imaging process and reduce potential side effects [53]. The in vitro and in vivo MRI study results shown in Fig. 4 demonstrated the outstanding performance of our 1-nm SPIONs and thus justified the use of a two-pronged computational and experimental strategy for efficient and effective MRI contrast agent development.
-202410/1001-8042-35-10-005/alternativeImage/1001-8042-35-10-005-F004.jpg)
Conclusion
Through MRI simulations in MATLAB, the fundamental mechanisms underlying the contrast power of SPIONs with different structures were revealed, showing that SPIONs with sizes in the range of 1–3 nm were optimal for T1-weighted MRI. Our experimental synthesis successfully produced the targeted 1–3-nm SPIONs, as confirmed by TEM and SQUID measurements. In particular, XRD and XAFS experiments and analyses revealed that our SPIONs targeted spherical maghemite cores, explaining their magnetic origin. Under in vitro and in vivo MRI conditions, these SPIONs exhibited attractive whole-body contrast power and pharmacokinetic attributes in mice. Overall, these high-performance qualities make our SPIONs promising candidates for MRI and MR angiography applications. This research lays the groundwork for an efficient and robust strategy for the future development and evaluation of nanoparticulate MRI probes. By providing these insights, our two-pronged computational and experimental method has the potential to offer innovative solutions for enhancing diagnostic precision and patient care.
Bringing MRI to low‐ and middle‐income countries: Directions, challenges and potential solutions
. NMR Biomed.Imaging macrophages with nanoparticles
. Nat. Mater. 13, 125–138 (2014). https://doi.org/10.1038/nmat3780Iron oxide based nanoparticles for multimodal imaging and magnetoresponsive therapy
. Chem. Rev. 115, 10637–10689 (2015). https://doi.org/10.1021/acs.chemrev.5b00112Superparamagnetic colloid suspensions: Water magnetic relaxation and clustering
. J. Magn. Magn. Mater. 293, 532–539 (2005). https://doi.org/10.1016/j.jmmm.2005.01.070Chemical design of nanoprobes for T 1-weighted magnetic resonance imaging
. Mater. Today 19, 336–348 (2016). https://doi.org/10.1016/j.mattod.2015.11.006Synchrotron radiation-based materials characterization techniques shed light on molten salt reactor alloys
. Nucl. Sci. Tech. 31, 6 (2019). https://doi.org/10.1007/s41365-019-0719-7Study of the relationship between the local geometric structure and the stability of La0.6Sr0.4MnO3−δ and La0.6Sr0.4FeO3−δ electrodes
. Nucl. Sci. Tech. 30, 21 (2019). https://doi.org/10.1007/s41365-019-0550-1Application of X-ray absorption fine structure spectroscopy to materials and environmental science
. Appl. Spectrosc. 52, 252A–279A (1998). https://doi.org/10.1366/0003702981944599Modern X-ray spectroscopy: XAS and XES in the laboratory
. Coord. Chem. Rev. 423,NMR relaxation in tissues with weak magnetic inhomogeneities
. Magn. Reson. Med. 44, 144–156 (2000).Strong field behavior of the NMR signal from magnetically heterogeneous tissues
. Magn. Reson. Med. 43, 226–236 (2000)..T(2)-shortening by strongly magnetized spheres: a chemical exchange model
. Magn. Reson. Med. 47, 388–391 (2002). https://doi.org/10.1002/mrm.10064Mapping the microscale origins of magnetic resonance image contrast with subcellular diamond magnetometry
. Nat. Commun. 9, 131 (2018). https://doi.org/10.1038/s41467-017-02471-7Renal clearance of quantum dots
. Nat. Biotechnol. 25, 1165–1170 (2007). https://doi.org/10.1038/nbt1340Design considerations for tumour-targeted nanoparticles
. Nat. Nanotechnol. 5, 42–47 (2010). https://doi.org/10.1038/nnano.2009.314Influence of Nanoparticles Size on XRD Patterns for Small Monodisperse Nanoparticles of Cu 0 and TiO 2 Anatase
. Ind. Eng. Chem. Res. 57, 2526–2536 (2018). https://doi.org/10.1021/acs.iecr.7b04480.EXAFS and XANES analysis of oxides at the nanoscale
. IUCrJ 1, 571–589 (2014). https://doi.org/10.1107/S2052252514021101Local structural evolutions of CuO/ZnO/Al2O3 catalyst for methanol synthesis under operando conditions studied by in situ quick X-ray absorption spectroscopy
. Nucl. Sci. Tech. 28, 21 (2016). https://doi.org/10.1007/s41365-016-0170-yXAFS and SRGI-XRD studies of the local structure of tellurium corrosion of Ni–18%Cr alloy
. Nucl. Sci. Tech. 30, 153 (2019). https://doi.org/10.1007/s41365-019-0673-4The local atomic structure of colloidal superparamagnetic iron oxide nanoparticles for theranostics in oncology
. Biomedicines 6, 78 (2018). https://doi.org/10.3390/biomedicines6030078Nitric acid-stabilized superparamagnetic iron oxide nanoparticles studied with X-rays
. J. Nanopart. Res. 14, 1066 (2012). https://doi.org/10.1007/s11051-012-1066-3SPIO nanoparticles and magnetic erythrocytes as contrast agents for biomedical and diagnostic applications
. J. Magn. Magn. Mater. 541,Magnetic iron oxide nanoparticle (IONP) synthesis to applications: present and future
. Materials (Basel) 13, 4644 (2020). https://doi.org/10.3390/ma13204644A new class of cubic SPIONs as a dual-mode T1 and T2 contrast agent for MRI
. Magn. Reson. Imag. 49, 16–24 (2018). https://doi.org/10.1016/j.mri.2017.09.013Iron oxide nanoparticles as T1 contrast agents for magnetic resonance imaging: fundamentals, challenges, applications, and prospectives
. Adv. Mater. 33,Fast MATLAB evaluation of nonlinear energies using FEM in 2D and 3D: Nodal elements
. Appl. Math. Comput. 424,Parallel sparse matrix-matrix multiplication and indexing: implementation and experiments
. SIAM J. Sci. Comput. 34, C170–C191 (2012). https://doi.org/10.1137/110848244Theory of NMR signal behavior in magnetically inhomogeneous tissues: the static dephasing regime
. Magn. Reson. Med. 32, 749–763 (1994). https://doi.org/10.1002/mrm.1910320610Analytical model of susceptibility-induced MR signal dephasing: effect of diffusion in a microvascular network
. Magn. Reson. Med. 41, 499–509 (1999). https://doi.org/10.1002/(sici)1522-2594(199903)41:3<499:aid-mrm12>3.0.co;2-oOn T2-shortening by weakly magnetized particles: The chemical exchange model
. Magn. Reson. Med. 45, 1014–1020 (2001). https://doi.org/10.1002/mrm.1135On T(2)-shortening by strongly magnetized spheres: a partial refocusing model
. Magn. Reson. Med. 47, 257–263 (2002). https://doi.org/10.1002/mrm.10059T2 relaxation induced by clusters of superparamagnetic nanoparticles: Monte Carlo simulations
. Magn. Reson. Imaging 26, 994–998 (2008). https://doi.org/10.1016/j.mri.2008.01.039.Microscopic susceptibility variation and transverse relaxation: theory and experiment
. Magn. Reson. Med. 31, 601–610 (1994). https://doi.org/10.1002/mrm.1910310605Compact zwitterion-coated iron oxide nanoparticles for biological applications
. Nano Lett. 12, 22–25 (2012). https://doi.org/10.1021/nl202721qCompact zwitterion-coated iron oxide nanoparticles for in vitro and in vivo imaging
. Integr. Biol. 5, 108–114 (2013). https://doi.org/10.1039/C2IB20142AExceedingly small iron oxide nanoparticles as positive MRI contrast agents
. Proc. Natl. Acad. Sci. 114, 2325–2330 (2017). https://doi.org/10.1073/pnas.1620145114.Monte Carlo simulation and theory of proton NMR transverse relaxation induced by aggregation of magnetic particles used as MRI contrast agents
. J. Magn. Reson. 212, 139–148 (2011). https://doi.org/10.1016/j.jmr.2011.06.024Applications of the Dixon technique in the evaluation of the musculoskeletal system
. Radiol. Bras. 54, 33–42 (2021). https://doi.org/10.1590/0100-3984.2019.0086.A review of the magnetic properties, synthesis methods and applications of maghemite
. J. Magn. Magn. Mater. 426, 74–81 (2017). https://doi.org/10.1016/j.jmmm.2016.11.033Synthesis, principles, and properties of magnetite nanoparticles for in vivo imaging applications-a review
. Pharmaceutics 11, 601 (2019). https://doi.org/10.3390/pharmaceutics11110601Single-nanometer iron oxide nanoparticles as tissue-permeable MRI contrast agents
. Proc. Natl. Acad. Sci. USA 118,Structure-relaxivity relationships of magnetic nanoparticles for magnetic resonance imaging
. Adv. Mater. 31,Magnetic nanoparticles: surface effects and properties related to biomedicine applications
. Int. J. Mol. Sci. 14, 21266–21305 (2013). https://doi.org/10.3390/ijms141121266Hematite from Natural Iron Stones as Microwave Absorbing Material on X-Band Frequency Ranges
. IOP Conf. Ser. Mater. Sci. Eng. 196,Calculations of electron energy loss and X-ray absorption spectra in periodic systems without a supercell
. Phys. Rev. B Condens. Matter 81,Ab initio theory and calculations of X-ray spectra
. Comptes Rendus Phys. 10, 548–559 (2009). https://doi.org/10.1016/j.crhy.2008.08.004Parameter-free calculations of X-ray spectra with FEFF9
. Phys. Chem. Chem. Phys. 12, 5503–5513 (2010). https://doi.org/10.1039/B926434E.Superparamagnetic iron oxide nanoparticles: diagnostic magnetic resonance imaging and potential therapeutic applications in neurooncology and central nervous system inflammatory pathologies, a review
. J. Cereb. Blood Flow Metab. 30, 15–35 (2010). https://doi.org/10.1038/jcbfm.2009.192Ferumoxytol: a new, clinically applicable label for stem-cell tracking in arthritic joints with MRI
. Nanomedicine (Lond) 8, 1969–1983 (2013). https://doi.org/10.2217/nnm.12.198A highly effective, nontoxic T1 MR contrast agent based on ultrasmall PEGylated iron oxide nanoparticles
. Nano Lett. 9, 4434–4440 (2009). https://doi.org/10.1021/nl902715vRecent developments in magnetic diagnostic systems
. Chem. Rev. 115, 10690–10724 (2015). https://doi.org/10.1021/cr500698dMechanisms and clinical implications of renal drug excretion
. Drug Metab. Rev. 33, 299–351 (2001). https://doi.org/10.1081/dmr-120000654The authors declare that they have no competing interests.