Abstract
Fluorescent molecules with specific target moieties are essential for histopathological analysis, but their limited tissue penetration depth makes in vivo, in situ color encoding analysis challenging. Magnetic resonance imaging (MRI) offers deep tissue penetration. When combined with chemical shift-encoded MRI reporters, it enables in vivo chemical shift encoding for biotarget imaging and analysis. These reporters require both strong signal intensity and large chemical shift window. However, conventional proton MRI reporters, with low sensitivity and a small chemical shift window, limit their in vivo applications. Here, we describe a chemical shift-encoded hyperpolarized 129Xe MRI reporter based on the multivariate metal-organic framework, NiZn-ZIF-8, to overcome these challenges. The proposed NiZn-ZIF-8 gives distinct chemical shifts for dissolved and entrapped 129Xe without signal interference, enhancing the 129Xe NMR signal by 210 times compared to dissolved 129Xe in water and biological media. This enables detection threshold at â 4 fM concentrations, setting a record for the lowest concentration of xenon hosts detected in nanomaterials. Additionally, NiZn-ZIF-8 exhibits good in vivo MRI performance, allowing xenon encoding and distinction in rat lungs. NiZn-ZIF-8 represents a versatile and powerful platform for advanced molecular imaging and in vivo biomedical diagnostics.
Similar content being viewed by others
Introduction
Fluorescence imaging offers numerous advantages, such as high sensitivity and real-time imaging, making it an invaluable tool for studying molecular mechanisms and early disease diagnosis1,2. Many fluorescent molecules have been developed, which can color-encode different targets for histopathological analysis and cellular imaging based on their emission wavelengths after being labeled with specific targeting moieties. However, challenges remain in vivo color-encoded imaging, including background interference, fluorescence signal decay, and limited tissue penetration depth.
Magnetic resonance imaging (MRI) is a modality in clinical diagnostics and biomedical research that allows for the non-invasive visualization of internal structures and biological processes without limitations in tissue penetration depth. Chemical shift is one of the most important parameters in nuclear magnetic resonance (NMR), closely related to the chemical environment of the atomic nucleus. It plays a crucial role in substance identification and molecular structure analysis. Different substances can be encoded by their chemical shifts for analysis. When combined with functionalized chemical shift-encoded MRI reporters, it enables in vivo chemical shift-encoded imaging for biotargets. To achieve effective chemical shift-encoded MRI, the reporters must provide both strong signal intensity and a large chemical shift window. However, conventional proton chemical shift-encoded MRI reporters suffer from fundamental limitations, such as low sensitivity and a small chemical shift window, which limit their application in complex biological systems.
Because of the associated high sensitivity, large chemical shift window, and absence of biological background signals, hyperpolarized 129Xe gas, produced by the spin-exchange optical pumping (SEOP) method, has established itself as a promising agent3,4,5,6,7,8,9. Dissolution of this gas in aqueous media or in blood results in a single peak at â 193 ppm, far off from the gaseous 129Xe signal in â 0 ppm. Trapping xenon atoms with molecular cages, such as cryptophanes and cucurbit[6]uril, generates a distinct NMR signal for entrapped xenon, which differs from that of dissolved xenon and xenon gas, allowing for encoding by chemical shift for analysis. These chemical shift-encoded molecular cages, similar to fluorescent molecules, can be functionalized with various target moieties, enabling the detection of different biomolecules or cells10,11,12,13,14. However, these molecular cages typically exhibit weak 129Xe NMR signals, particularly in aqueous and biological environments, due to low xenon loading and signal intensity. Most of such drawbacks hampered their practical applications in molecular imaging, much more so in vivo experiments, where the signal is required to be strong and stable, and specificity often becomes an issue. So far, enhancement efforts toward 129Xe NMR signals through physical and chemical modifications have been minor15,16; thus, there is an urgent need for the strategy of amplifying 129Xe signal intensity with its biocompatibility preserved.
Metal-organic frameworks (MOFs) offer a versatile platform to address these challenges. MOFs are crystalline materials that consist of metal clusters or ions bridged by organic ligands, forming highly ordered, porous networks with tunable pore sizes, high surface areas, and robust chemical stability. Such characteristics make MOFs attractive for applications involving gas storage, separation, and sensing17,18,19,20,21,22,23. Among them, multivariate MOFs (MTV-MOFs) are of particular interest because they integrate multiple building blocks into a single lattice without changing their topology. This MTV methodology allows for fine-tuning of the MOFâs internal environment24,25,26, resulting in significantly higher affinities for guest molecules, enhanced signal intensities, and improved selectivity compared to single-component MOFs27,28,29,30,31. In this paper, we introduce a multivariate MOF, NiZn-ZIF-8, purposefully engineered to enhance the entrapped 129Xe NMR signal intensity, to develop a chemical shift-encoded reporter for hyperpolarized 129Xe MRI in vivo. Incorporation of nickel into the ZIF-8 framework provides some subtle adjustments of its pore environment in such a way that hyperpolarized xenon entrapment and enhanced signal amplification can take place. NiZn-ZIF-8 overcomes critical limitations of traditional xenon-based MRI reporters with fM sensitivity and stability in aqueous and biological media. This work introduces a versatile platform for advanced 129Xe molecular imaging, with far-reaching possibilities toward transformative in vivo applications, and extends the utility of MOFs in biomedical diagnostics.
Previous studies have shown that zeolites and MOFs can adsorb xenon atoms32,33,34,35,36. Especially, MOFs such as ZIF-8, IRMOF-1, IRMOF-8, and IRMOF-10 have a strong affinity for xenon atoms, making them effective hosts for hyperpolarized 129Xe MRI probes37,38. Among these, ZIF-8 stands out because of its high surface area and porosity, which allow it to trap xenon atoms efficiently37. Unlike traditional hosts like cryptophanes and cucurbit[6]uril39,40, ZIF-8 can effectively concentrate xenon within its porous structure while keeping it separate from the surrounding aqueous phase. This confinement creates a distinct chemical shift in 129Xe NMR signals, providing a reliable platform for enhancing signal strength.
Building on ZIF-8âs natural advantages and the flexibility of MTV-MOFs, we developed NiZn-ZIF-8, to amplify 129Xe NMR signals in aqueous environments. By incorporating nickel into the ZIF-8 structure, we created a chemical shift-encoded reporter that not only boosts the intensity of the 129Xe signal but also distinguishes between dissolved and entrapped xenon through distinct chemical shifts.
To implement this strategy, we synthesized a series of nickel-substituted multivariate ZIF-8 materials, called NiZn-ZIF-8, carefully fine-tuning the nickel content to optimize the 129Xe NMR signal (Fig. 1a, b). The optimized NiZn-ZIF-8 showed a 33% increase in the entrapped 129Xe NMR signal intensity compared to standard ZIF-8, along with an impressive 210-fold enhancement over dissolved 129Xe in water and biological media under the same conditions. Additionally, NiZn-ZIF-8 achieves a detection threshold as low as â 4 fM. Acting as a highly sensitive, chemical shift-encoded reporter for hyperpolarized 129Xe MRI, NiZn-ZIF-8 demonstrates good in vivo MRI performance. The entrapped xenon can be clearly imaged and distinguished from both dissolved xenon and xenon gas in rat lungs. Its propertiesâsuch as good dispersibility in water, a hydrophobic pore environment, pore sizes that closely match xenon atoms, and high stability in aqueous mediaâmake it an ideal platform for concentrating xenon and enhancing signal strength. By addressing key challenges in sensitivity and signal limitations, NiZn-ZIF-8 marks an advancement in molecular imaging technologies, opening possibilities for biomedical applications.
a Schematic illustration showing the increase in hyperpolarized 129Xe NMR signal intensity upon addition of NiZn-ZIF-8 to water. b Structural representation of NiZn-ZIF-8, highlighting nickel substitution within the ZIF-8 framework.c Hyperpolarized 129Xe NMR spectra of NiZn-ZIF-8 (25âmgâmL-1), pristine ZIF-8 (25âmgâmL-1), and pure water, demonstrating a significant signal enhancement for NiZn-ZIF-8. d Hyperpolarized 129Xe NMR spectra of NiZn-ZIF-8 (25âmgâmL-1), pristine ZIF-8 (25âmgâmL-1), and culture medium (containing 10% FBS), showing signal enhancement in a biologically relevant environment. e Quantified 129Xe NMR signal enhancement fold from (c), showing aâ>â210-fold enhancement for NiZn-ZIF-8 compared to pure water (nâ=â3 independent experimental replicates). f Quantified 129Xe NMR signal enhancement fold from (d), confirming similar signal enhancements in biological media (nâ=â3 independent experimental replicates). All NMR spectra were acquired using a zg pulse sequence (rectangular pulse, p1â=â31.8 μs), with a single scan (NSâ=â1) and line broadening (LBâ=â10âHz). Data in (e, f) were presented as mean valuesâ±âSD. Source data are provided as a Source Data file.
Results
Preparation and characterization of MOF samples
A series of NiZn-ZIF-8 nanoparticles were synthesized by adjusting the molar ratio of nickel to zinc salts in a methanol solution containing 2-methylimidazolate. The final nickel content in the nanoparticles increased as the initial nickel-to-zinc ratio increased (Supplementary Fig. 1, Supplementary Table 1), and all of these samples exhibited a positive Zeta potential (Supplementary Table 2).
Scanning electron microscopy (SEM) showed that incorporating nickel influenced both the size and shape of the particles. At lower nickel levels (below 0.13%), the nanoparticles maintained a uniform size, while higher nickel content (ââ¥â0.33%) resulted in more irregular particle shapes (Fig. 2b, Supplementary Figs. 2â8). Despite this, all NiZn-ZIF-8 nanoparticles retained their characteristic dodecahedral shape, as confirmed through SEM and transmission electron microscopy (TEM) (Fig. 2b, c, Supplementary Figs. 2â17).
a Schematic representation of the NiZn-ZIF-8 crystal structure. b SEM images of NiZn-ZIF-8 nanoparticles showing consistent dodecahedral morphology across samples (scale bars: 1 μm). c TEM images of NiZn-ZIF-8 (0.08% nickel), illustrating uniform particle size and structure (scale bar: 200ânm). d High-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) and elemental mapping of NiZn-ZIF-8 (0.08% nickel), confirming the homogeneous distribution of Ni and Zn within the nanoparticles (scale bars: 200ânm). e Relative 129Xe NMR signal intensities for NiZn-ZIF-8, pristine ZIF-8, and other MOF samples, demonstrating the superior signal enhancement achieved by NiZn-ZIF-8 (nâ=â3 independent experimental replicates). f Hyperpolarized 129Xe NMR spectra of NiZn-ZIF-8 nanoparticles (0.08% nickel) at various concentrations (5.0, 10, 15, 20, and 25âmgâmL-1) in aqueous solutions, showing a concentration-dependent increase in signal intensity. g Normalized 129Xe NMR signal intensities for NiZn-ZIF-8 (red) and pristine ZIF-8 (purple) as a function of MOF concentration, indicating that NiZn-ZIF-8 exhibits aâââ33% greater xenon entrapment capacity compared to pristine ZIF-8 (nâ=â3 independent experimental replicates). All NMR spectra were acquired using a zg pulse sequence (rectangular pulse, p1â=â31.8 μs), with a single scan (NSâ=â1) and line broadening (LBâ=â10âHz). Data in (e, g) were presented as mean valuesâ±âSD. Source data are provided as a Source Data file.
Elemental mapping revealed that nickel and zinc atoms were evenly distributed throughout individual NiZn-ZIF-8 crystals (Fig. 2d, Supplementary Figs. 18â26). For comparison, we also synthesized and characterized Cu2+- and Co2+-doped ZIF-8 materials (CuZn-ZIF-8 and CoZn-ZIF-8), which exhibited similar dodecahedral shapes (Supplementary Figs. 27â30). Thermogravimetric analysis (TGA) demonstrated the thermal stability of all MOFs, with no significant weight loss observed below 400â°C (Supplementary Fig. 31).
Nitrogen adsorption measurements at 77âK confirmed the permanent porosity of these materials (Supplementary Figs. 32â56). Notably, NiZn-ZIF-8 showed an increased BET surface area of â 2000 m² g-1, surpassing that of pristine ZIF-8. In contrast, CuZn-ZIF-8 and CoZn-ZIF-8 maintained surface area values similar to pristine ZIF-8 (Supplementary Table 3). Pore size analysis using the nonlocal density functional theory (NLDFT) model revealed an average pore size of about 11âà for NiZn-ZIF-8, which is consistent with that of pristine ZIF-8 (Supplementary Table 4).
Powder X-ray diffraction (PXRD) analysis confirmed that NiZn-ZIF-8 retained its crystalline integrity and phase purity, showing patterns identical to those of pristine ZIF-8 without any additional peaks (Supplementary Figs. 57â59). This indicates that incorporating Ni2+, Cu2+, or Co2+ does not disrupt the original ZIF-8 lattice structure. Further validation through single-crystal X-ray diffraction revealed the partial substitution of Zn2+ with Ni2+ in the framework backbone (Supplementary Table 5). The stability of NiZn-ZIF-8 was thoroughly tested. Its PXRD pattern remained unchanged even after being immersed in water for one week, demonstrating high stability in aqueous environments (Supplementary Fig. 60).
Further evidence of nickel incorporation was provided by diffuse reflectance visible spectroscopy, which showed two characteristic absorption peaks at 580ânm and 790ânm, corresponding to the [NiN4] clusters (Supplementary Fig. 61). These findings align with previous studies41, confirming the stable integration of nickel into the MOF framework.
NiZn-ZIF-8 nanoparticles demonstrate good thermal stability and water stability, and improved porosity compared to pristine ZIF-8. The partial replacement of Zn2+ with Ni2+ introduces valuable properties, creating a robust and adaptable platform for efficient xenon entrapment and significant enhancement of the 129Xe NMR signal.
Investigation of the xenon entrapment capability of MOF
When dispersed in water, solvent-free NiZn-ZIF-8 nanoparticles formed a uniform and stable colloidal solution (Supplementary Fig. 62). Hyperpolarized 129Xe NMR spectra of this dispersion (5âmgâmL-1) revealed two distinct peaks: one at 193.4 ppm, corresponding to dissolved 129Xe in the aqueous phase, and another at 82.2 ppm, representing 129Xe trapped within the pores of NiZn-ZIF-8 (Supplementary Fig. 63). Additionally, a third peak at 84.1 ppm, observed in the solid state, indicates that NiZn-ZIF-8 retains its dry micropores even when dispersed in a liquid colloidal environment.
Incorporating nickel into the ZIF-8 framework had minimal impact on the chemical shift of entrapped 129Xe, but it significantly affected signal intensity (integrated peak area) (Supplementary Fig. 64). The signal intensity increased with rising nickel content but decreased beyond a critical threshold (Fig. 2e). This may be due to two key factors: 1) Paramagnetic effects of nickel, which shorten 129Xe T1 relaxation times (Supplementary Fig. 65, Supplementary Table 6). 2) Reduced dispersity and non-uniform particle morphology at higher nickel concentrations (Fig. 2b, Supplementary Figs. 2-8).
Through optimization, the ideal range for nickel substitution was identified as 0.08%â0.13%. At an optimal 0.08% nickel content, NiZn-ZIF-8 achieved a 129Xe NMR signal intensity that was 212-fold higher than dissolved 129Xe in water (Fig. 1c, e). In biologically relevant culture media containing 10% fetal bovine serum (FBS), the entrapped 129Xe signal was similarly impressive, showing a 216-fold enhancement compared to dissolved 129Xe (Fig. 1d, f; for the calculation, see Methods). More importantly, although NiZn-ZIF-8 is in blood plasma solutions, the entrapped 129Xe NMR signal remains strong (Supplementary Fig. 66), making it promising for potential biological applications.
Further analysis revealed that the 129Xe NMR signal intensity was directly proportional to the concentration of NiZn-ZIF-8 and pristine ZIF-8 nanoparticles, displaying a clear linear relationship over a concentration range of 1â25âmgâmL-1, without altering the entrapped 129Xe chemical shift (Fig. 2f, g, Supplementary Fig. 67). Importantly, NiZn-ZIF-8 consistently demonstrated stronger xenon entrapment compared to pristine ZIF-8, with the slope of the linear fit (0.0418âmLâmg-1) being â 33% higher than that of ZIF-8 (0.0313âmLâmg-1) (Fig. 2g). We calculated that the number of entrapped xenon atoms per nanoparticle is 3.03 Ã 107 for NiZn-ZIF-8 and 2.29 Ã 107 for ZIF-8, respectively. In contrast, ZIF-8 doped with Cu2+ or Co2+ did not improve 129Xe NMR signal intensity (Supplementary Fig. 68).
These findings collectively highlight the superior xenon entrapment capabilities of NiZn-ZIF-8, resulting in a significant boost in 129Xe NMR signal intensity in solution. The optimized nickel doping enhances xenon concentration properties, positioning NiZn-ZIF-8 as a highly promising and customizable platform for hyperpolarized 129Xe molecular imaging applications.
NiZn-ZIF-8 with good hyperpolarized 129Xe MRI performance
To assess the performance of NiZn-ZIF-8 in hyperpolarized 129Xe MRI, experiments were carried out using high nanoparticle concentrations in a culture medium. A dual-compartment system was employed, with a 5âmm NMR tube containing a NiZn-ZIF-8 dispersion (40âmgâmL-1) placed inside a 10âmm NMR tube filled with pure culture medium (Fig. 3a). Proton MRI showed strong signal intensity in the 5âmm tube, attributed to proton relaxation enhancement caused by NiZn-ZIF-8, while the outer tube displayed much weaker signal intensity (Fig. 3b). In hyperpolarized 129Xe MRI, the culture medium alone produced a signal corresponding to dissolved 129Xe. In contrast, the NiZn-ZIF-8 dispersion showed no detectable signal for dissolved 129Xe, indicating strong xenon entrapment and relaxation effects. Instead, a distinct and intense signal was observed for entrapped 129Xe (Fig. 3c). Notably, the signals for dissolved and entrapped xenon were clearly separated, eliminating any signal interference.
a Schematic illustration of the NMR tube setup used in the experiment, consisting of a 5âmm inner tube containing NiZn-ZIF-8 dispersion (40âmgâmL-1) and a 10âmm outer tube containing pure culture medium. b Proton MRI images showing weak signal intensity in the 10âmm tube (culture medium) and strong signal intensity in the 5âmm tube, attributed to proton relaxation enhancement by NiZn-ZIF-8. c Chemical shift-encoded hyperpolarized 129Xe MRI results for NiZn-ZIF-8, demonstrating the separation of dissolved 129Xe (193.4 ppm) and entrapped 129Xe (82.2 ppm) signals into distinct colors without interference. d Schematic illustration of the intratracheal instillation of NiZn-ZIF-8 solution in rats (20âmgâmL-1, 0.2âmL). e In vivo hyperpolarized 129Xe NMR spectrum exhibits a strong 129Xe NMR signal entrapped by NiZn-ZIF-8, demonstrating its xenon entrapment capability in biological systems. f Hyperpolarized 129Xe images of a rat after NiZn-ZIF-8 instillation clearly show the signal of xenon entrapped by NiZn-ZIF-8 (purple), xenon dissolved in TP (green), and xenon gas (gray). g 3D images reconstructed from the simultaneous imaging of xenon gas (gray), xenon entrapped within NiZn-ZIF-8 (purple), and xenon dissolved in TP (green). h Hyperpolarized 129Xe images of a rat after NiZn-ZIF-8 instillation clearly show the signal of xenon entrapped by NiZn-ZIF-8 (purple), xenon dissolved in RBC (cyan), and xenon gas (gray). i 3D images reconstructed from the simultaneous imaging of xenon gas, xenon entrapped within NiZn-ZIF-8, and xenon dissolved in RBC.
To investigate the imaging performance of NiZn-ZIF-8 in vivo, hyperpolarized 129Xe NMR and MRI were conducted following the intratracheal instillation of NiZn-ZIF-8 solution (20âmgâmL-1, 0.2âmL) in rats (Fig. 3d-g). Hyperpolarized 129Xe NMR spectra revealed four characteristic signal peaks: the peak at 0 ppm corresponded to xenon gas, the peak at 90 ppm to the entrapped 129Xe within NiZn-ZIF-8, and the peaks at 195 and 212 ppm to the dissolved 129Xe in tissue and plasma (TP), and red blood cells (RBC), respectively (Fig. 3e). Because of the strong entrapped xenon signal and its chemical shift being far from that of xenon gas and dissolved xenon, simultaneous imaging of xenon gas, dissolved xenon, and entrapped xenon in the lungs was achieved. The signals were clearly observed and distinguished (Fig. 3f-i). Notably, even with just 4âmg of NiZn-ZIF-8 nanoparticles, the entrapped xenon signal was sufficiently strong for analysis, demonstrating the good chemical shift-encoded 129Xe MRI performance of NiZn-ZIF-8 in vivo. This highlights the great potential of NiZn-ZIF-8 for future biological applications.
The xenon entrapment ability of NiZn-ZIF-8 and ZIF-8 was further analyzed using the hyperpolarized xenon chemical exchange saturation transfer (Hyper-CEST) method16. Both NiZn-ZIF-8 and pristine ZIF-8 exhibited a saturation response centered around 82 ppm. However, at an equivalent concentration of 40âμgâmL-1, NiZn-ZIF-8 demonstrated a significantly stronger Hyper-CEST effect compared to pristine ZIF-8, aligning with the results observed in direct 129Xe NMR measurements (Fig. 4a).
a Hyper-CEST spectra comparing NiZn-ZIF-8 (purple circle) and pristine ZIF-8 (blue circle) at the same concentration (40âμgâmL-1), showing a stronger Hyper-CEST effect for NiZn-ZIF-8. b Hyper-CEST spectra of NiZn-ZIF-8 at varying concentrations (20, 30, 40, 50, and 60âμgâmL-1) in aqueous solutions, highlighting the concentration-dependent increase in signal intensity. c Plot of the Hyper-CEST effect at 82 ppm as a function of NiZn-ZIF-8 concentration, demonstrating a linear relationship over the range of 20â60âμgâmL-1. d Hyper-CEST MRI phantom images of NiZn-ZIF-8 at different concentrations, illustrating the dose-dependent enhancement in MRI signal intensity. Source data are provided as a Source Data file.
Further Hyper-CEST experiments conducted across a range of NiZn-ZIF-8 concentrations (20â60âμgâmL-1) revealed a clear linear relationship between the Hyper-CEST effect and nanoparticle concentration (Fig. 4b, c). Remarkably, even at a low concentration of 20âμgâmL-1, NiZn-ZIF-8 achieved a Hyper-CEST effect of â 30%, highlighting its capability to trap xenon. Corresponding Hyper-CEST MRI also demonstrated strong signal intensity even at low concentrations of NiZn-ZIF-8 nanoparticles, with intensity measurements showed a strong linear relationship with NiZn-ZIF-8 concentration (Fig. 4d, Supplementary Figs. 69, 70).
To determine the detection threshold, time-dependent saturation transfer spectra were collected for serial dilutions of NiZn-ZIF-8 at 100, 10, 2, and 1âμgâmL-1 (Supplementary Fig. 71). The detection threshold was identified as 2âμgâmL-1, which corresponds to a low concentration of â 4 fM, calculated using nanoparticle tracking analysis. The detection threshold of pristine ZIF-8 was identified as 3âμgâmL-1 by the same method (Supplementary Fig. 72). The detection threshold of NiZn-ZIF-8 is about 25 times lower than that of nanoemulsion droplets (100 fM)42, and about 6250 times lower than that of the gas vesicles (25 pM)43. This represents the lowest concentration of detected xenon hosts of nanomaterials to date, marking a significant advancement in hyperpolarized 129Xe NMR field (Supplementary Table 7)42,43,44,45,46.
NiZn-ZIF-8 proves to be an effective xenon host, capable of concentrating xenon atoms with unmatched sensitivity and significantly boosting 129Xe NMR signals in aqueous solutions. These results underscore its potential for hyperpolarized 129Xe MRI at low concentrations, positioning it as a powerful tool for advanced molecular imaging and diagnostic applications.
Investigation of xenon interactions within the MOF pores
The role of nickel ions in influencing xenon interactions within the NiZn-ZIF-8 pores was studied using variable-temperature hyperpolarized 129Xe NMR and two-dimensional 129Xe exchange spectroscopy (2D EXSY). In aqueous solutions containing NiZn-ZIF-8, the chemical shift of dissolved 129Xe showed a nonlinear temperature dependence: it increased with temperature up to 300âK and then decreased at higher temperatures (Fig. 5a, Supplementary Fig. 73). A similar trend was observed for pure water and pristine ZIF-8 (Fig. 5a). In contrast, the chemical shift of entrapped 129Xe within NiZn-ZIF-8 exhibited a linear relationship with temperature over the range of 278â320âK. The slope and intercept for NiZn-ZIF-8 were â0.0731â±â0.0004 ppm K-1 and 103.98â±â0.12 ppm, respectively, compared to â0.0759â±â0.0012 ppm K-1 and 104.51â±â0.36 ppm for pristine ZIF-8. These slight differences suggest that the incorporation of nickel into the framework alters the interactions between xenon and the MOF pore environment, fine-tuning its properties. Interestingly, the temperature-dependent behavior of entrapped 129Xe in NiZn-ZIF-8 differs from molecular cages like cryptophane and metal-organic capsules47,48,49,50,51, which typically trap only a single xenon atom per cavity. In contrast, MOFs can hold multiple xenon atoms within each pore, allowing xenon-xenon interactions to influence the NMR signal. At higher temperatures, the increased thermal motion of xenon atoms reduces their residence time within the pores and weakens xenon-xenon interactions, leading to upfield shifts in the 129Xe NMR signal.
a Temperature dependence of hyperpolarized 129Xe chemical shift for dissolved 129Xe in water, pristine ZIF-8, and NiZn-ZIF-8. Dissolved 129Xe exhibits nonlinear behavior, while entrapped 129Xe in NiZn-ZIF-8 shows a linear trend (nâ=â3 independent experimental replicates). b Schematic illustration of the chemical exchange process between dissolved 129Xe and entrapped 129Xe within NiZn-ZIF-8 pores. c 2D 129Xe EXSY spectrum of NiZn-ZIF-8 at a mixing time of 1.5âms, showing cross peaks indicative of xenon exchange between dissolved and entrapped states. d Exchange rate determination: Cross peak (purple circle), diagonal peak of dissolved 129Xe (red circle), and diagonal peak of entrapped 129Xe (blue circle) intensities from 2D 129Xe EXSY as a function of mixing time, fitted to a two-site exchange model, yielding a chemical exchange rate of 108â±â27âHz for NiZn-ZIF-8. e MD simulation setup: Snapshot of the simulation box (34 à 34 à 102 à 3) after equilibration, containing NiZn-ZIF-8 surrounded by 600 H2O molecules and 60 xenon atoms. Xenon atoms are shown in red; carbon, light gray; hydrogen, white; nitrogen, blue; and metal, light blue. H2O molecules are represented as a semi-transparent gray volume. f, g Xenon entrapment dynamics: Change in the number of xenon atoms entrapped by NiZn-ZIF-8 in (f) pure xenon and (g) xenon-water environments over the simulation timescale, demonstrating higher xenon entrapment in the xenon-water mixture. h Average xenon entrapment: MD simulation results showing the mean number of xenon atoms entrapped in pristine ZIF-8 (red box), NiZn-ZIF-8 (purple box), and CuZn-ZIF-8 (blue box) during the final 5âns of the simulation in the xenon-water environment. NiZn-ZIF-8 exhibits significantly higher xenon entrapment. nââ=ââ5001 independent data points. The box plots show the interquartile range, with the centre lines indicating the medians and the whiskers the minimum and maximum values. Data in (a) were presented as mean valuesâ±âSD. Statistical significance was calculated using unpaired two-tailed Studentâs t-test. Source data are provided as a Source Data file.
The dynamics of xenon exchange between the entrapped and dissolved states were examined using 2D EXSY (Supplementary Figs. 74-81). For the NiZn-ZIF-8, at a 0âms mixing time, no cross peaks were observed, indicating that no exchange occurred between the two states (Supplementary Fig. 76a). However, at a 1.5âms mixing time, cross peaks appeared, confirming chemical exchange between the entrapped and dissolved xenon (Fig. 5c). By fitting the peak intensities over time to a two-site exchange model52, the exchange rate for NiZn-ZIF-8 was determined to be 108â±â27âHz (Fig. 5d). In comparison, pristine ZIF-8 and CuZn-ZIF-8 showed faster exchange rates of 124â±â27âHz and 214â±â48âHz, respectively (Supplementary Figs. 77â81). The slower exchange rate in NiZn-ZIF-8 indicates that nickel incorporation enhances the MOFâs affinity for xenon, reducing the rate of exchange between the entrapped and dissolved states. This increased affinity aligns with the observed improvements in xenon entrapment and the enhanced 129Xe NMR signal intensity.
Xenon adsorption isotherms
Xenon adsorption isotherms were performed at 298 and 273âK to investigate the interaction strength between the xenon and the framework (Supplementary Fig. 82). We found that the xenon uptake amounts of NiZn-ZIF-8 is consistent with the pristine ZIF-8, but the isosteric heats of adsorption (Qst) of xenon in the framework is different. The Qst of xenon at nearly zero loadings for NiZn-ZIF-8 and ZIF-8 are estimated to be 16.93 and 16.83âkJâmol-1, respectively (Supplementary Figs. 83, 84). The Qst of xenon in the NiZn-ZIF-8 is slightly higher than that of pristine ZIF-8, demonstrating that the xenon atoms have a higher affinity with the NiZn-ZIF-8. Therefore, the nickel in the framework may play a key role in enhancing the affinity between the xenon and the framework.
Molecular dynamics simulations
To gain deeper insight into xenon entrapment, molecular dynamics (MD) simulations were conducted for NiZn-ZIF-8, CuZn-ZIF-8, and pristine ZIF-8 (Fig. 5eâh, Supplementary Movie 1, and Supplementary Figs. 85â90). Snapshots taken every 1.0âps along the simulation trajectory quantified the number of xenon atoms trapped within the MOF pores. The simulations revealed that xenon was rapidly entrapped in both pure xenon and xenon-water environments (Fig. 5f, g). In the pure xenon environment, the average number of xenon atoms trapped by NiZn-ZIF-8 has no significant difference compared with pristine ZIF-8 and CuZn-ZIF-8 (Fig. 5f, Supplementary Figs. 87, 88). This consistent with the xenon adsorption isotherms (Supplementary Fig. 82). While in the xenon-water mixture (Fig. 5g, Supplementary Fig. 89), NiZn-ZIF-8 showed a significantly higher average number of trapped xenon atoms (16.6) compared to pristine ZIF-8 (13.4) and CuZn-ZIF-8 (14.4) (Fig. 5h). Notably, the number of xenon atoms trapped in the xenon-water mixture was significantly higher than in the pure xenon environment. This result indicates that the MOF pores remain hydrophobic even when dispersed in water, allowing xenon atoms to preferentially occupy the pore interiors rather than dissolving into the surrounding water.
From MD simulations, we calculated the distances between xenon and metal atoms for each snapshot. The free-energy profiles were determined by analyzing the probability distributions of these distances (Supplementary Fig. 90). The results indicate that the free-energy minima for pristine ZIF-8, CuZn-ZIF-8, and NiZn-ZIF-8 occur at a distance of 5.8âà . However, the depth of the free-energy profiles varies: â24.78âkJâmol-1 for pristine ZIF-8, â24.87âkJâmol-1 for CuZn-ZIF-8, and â25.16âkJâmol-1 for NiZn-ZIF-8. These findings suggest that NiZn-ZIF-8 exhibits a stronger ability to trap the xenon atoms compared to both pristine ZIF-8 and CuZn-ZIF-8, as indicated by the deeper free-energy minimum.
These results confirm that incorporating nickel into the framework enhances the materialâs affinity for xenon, resulting in increased xenon entrapment. This improved trapping capability directly explains the enhanced 129Xe NMR signal intensity observed for NiZn-ZIF-8.In summary, incorporating nickel into the ZIF-8 framework significantly improves its affinity for xenon and increases the number of trapped xenon atoms, providing a dual advantage for molecular imaging. The enhanced BET surface area and superior xenon retention make NiZn-ZIF-8 a suitable material for concentrating xenon and amplifying NMR signals in aqueous solutions.
Discussion
This study demonstrates that NiZn-ZIF-8, developed through a MTV-MOF strategy, serves as an effective platform for enhancing hyperpolarized 129Xe NMR signal intensity in aqueous environments. By incorporating nickel into the ZIF-8 framework, we created a material with greater xenon affinity, improved xenon trapping capacity, and extended xenon residence time within its pores. These advancements led to aââ 33% increase in 129Xe NMR signal intensity compared to pristine ZIF-8 and a 210-fold enhancement relative to dissolved xenon in water and biological media. NiZn-ZIF-8 also excels as a chemical shift-encoded MRI reporter, precisely distinguishing between dissolved and entrapped xenon signals, and demonstrates good chemical shift-encoded MRI performance in rat lungs. With a detection threshold of â 4 fM, enabling the detection of trace molecules using Hyper-CEST MRI. These findings highlight NiZn-ZIF-8âs immense potential for advanced molecular imaging in complex biological systems, addressing critical challenges in signal intensity and sensitivity faced by existing probes.
This MTV strategy opens up broader opportunities to design next-generation MOFs with customizable properties. This expands their potential applications beyond 129Xe MRI into fields such as biomedical imaging, optics, and electronics. By showcasing how strategic compositional tuning can unlock new functionalities, this work marks a significant advancement in the development of functional MOFs for transformative applications in molecular imaging and diagnostics.In the future, the chemical shift-encoded reporter, NiZn-ZIF-8, will be functionalized with targeting moieties such as hyaluronic acid, RGD, or other relevant targeting moieties. When targeting-functionalized NiZn-ZIF-8 is combined with hyperpolarized 129Xe, it could enable targeted molecular imaging in vivo, particularly for the detection of diseases. For example, in the context of pulmonary diseases like lung cancer, hyperpolarized 129Xe combined with NiZn-ZIF-8 could provide imaging of the lungs, overcoming the low sensitivity challenges typically associated with conventional imaging agents.
Methods
Materials
Zinc acetate dehydrate (Analytical reagent, 99.7%), nickel acetate tetrahydrate, zinc nitrate hexahydrate (Analytical reagent, 99.7%), nickel nitrate hexahydrate (Analytical reagent, 99.7%), copper acetylacetonate (Analytical reagent, 99.7%), cobalt acetylacetonate (Analytical reagent, 99.7%), anhydrous methanol (Analytical reagent, 99.7%) were purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). 2-methyl-imidazole (mIM, 98%) was purchased from Aladdin. All chemicals were used as received without further purification unless otherwise stated.
Characterization
Characterization of the synthesized materials was conducted using a variety of techniques to assess their structural and thermal properties, as well as their morphology and elemental composition. BrunauerâEmmettâTeller (BET) surface area and pore size measurements were carried out through nitrogen adsorption and desorption isotherms at 77âK using a Micromeritics 3Flex 5.0 instrument. Powder X-ray diffraction (PXRD) patterns were recorded on a Bruker D8 (3âkW) diffractometer to confirm the crystallinity and phase purity of the samples. Thermal stability was evaluated using thermogravimetric analysis and differential scanning calorimetry (TGA-DSC) on a Mettler TG-DSC 3+ instrument in an air atmosphere over a temperature range of 30â800â°C, with a controlled air flow rate of 60âmLâmin-1.
The size and morphology of the MOF nanoparticles were examined using high-resolution SEM equipped with energy-dispersive X-ray spectroscopy (EDX) and TEM performed on an FEI Titan transmission electron microscope operating at 200âkV. Elemental mapping was conducted using scanning transmission electron microscopy (STEM) with a high-angle annular dark-field (HAADF) detector to confirm the homogeneous distribution of elements within the MOF frameworks. Additionally, elemental analysis was performed using inductively coupled plasma optical emission spectroscopy (ICP-OES) with an ICPOES760 instrument (Agilent) to quantify the elemental composition of the samples.
Synthesis of pristine ZIF-8
To synthesize pristine ZIF-8, zinc acetate dihydrate (Zn(OAc)2â·â2H2O, 542.2âmg, 2.47âmmol) and mIM (810.6âmg, 9.87âmmol) were each dissolved in 50âmL of MeOH in separate 250âmL conical flasks. The dissolution was aided by ultrasound to ensure complete solubilization of the reagents. The mIM solution was then poured into the zinc acetate solution under vigorous stirring, and mixing was continued until the solutions were fully combined. Stirring was stopped, and the resulting mixture was allowed to stand undisturbed for 24âh. A white precipitate formed during this time was recovered by centrifugation. The collected product was washed three times with MeOH to remove any unreacted metal salts or ligands, yielding the final ZIF-8 material.
Synthesis of NiZn-ZIF-8
To synthesize NiZn-ZIF-8, nickel acetate tetrahydrate (Ni(OAc)2â·â4H2O, 49.7âmg, 0.2âmmol) and zinc acetate dihydrate (Zn(OAc)2â·â2H2O, 175.6âmg, 0.8âmmol) were dissolved separately in 50âmL of MeOH with the help of ultrasound in a 250âmL conical flask. Simultaneously, mIM (656âmg, 8âmmol) was dissolved in 50âmL of MeOH in a separate 250âmL conical flask using ultrasound. The mIM solution was then added to the metal acetate solution under vigorous stirring. Stirring was continued for 3âh, after which the mixture was allowed to stand undisturbed for 24âh. The precipitate formed was collected by centrifugation and washed three times with MeOH to remove any unreacted metal salts or ligands. Samples of NiZn-ZIF-8 with varying nickel percentages were synthesized using the same procedure by altering the molar ratio of n(Ni(OAc)2â·â4H2O)/n(Zn(OAc)2â·â2H2O), while maintaining the total moles of the two metal salts at 1âmmol. The resulting nickel percentages in the MOF samples were as follows: for n(Ni(OAc)2â·â4H2O)/n(Zn(OAc)2â·â2H2O)â=â1/3 â 0.11% nickel; 1/2 â 0.13% nickel; 1/1 â 0.33% nickel; 2/1 â 0.64% nickel; 3/1 â 0.84% nickel; and 4/1 â 1.76% nickel.
Synthesis of CuZn-ZIF-8
Copper-substituted ZIF-8 (CuZn-ZIF-8) was synthesized using a procedure similar to that for NiZn-ZIF-8. Copper acetylacetonate (Cu(AcAc)2) (52.4âmg, 0.2âmmol) and zinc acetate dihydrate (Zn(OAc)2â·â2H2O) (175.6âmg, 0.8âmmol) were dissolved in 50âmL of MeOH using ultrasound. Separately, mIM (656âmg, 8âmmol) was dissolved in 50âmL of MeOH with ultrasound. The mIM solution was then poured into the copper and zinc solution under vigorous stirring, which was continued for 3âh. The mixture was left to stand undisturbed for 24âh, and the precipitate was collected by centrifugation. The product was washed three times with MeOH to remove unreacted salts and ligands. ICP-OES analysis indicated that the copper content in the CuZn-ZIF-8 sample was 0.06%.
Synthesis of CoZn-ZIF-8
Cobalt-substituted ZIF-8 (CoZn-ZIF-8) was synthesized by dissolving cobalt acetylacetonate (Co(AcAc)2) (4.97âmg, 0.02âmmol) and zinc acetate dihydrate (Zn(OAc)2â·â2H2O, 215.1âmg, 0.98âmmol) in 50âmL of MeOH using ultrasound in a 250âmL conical flask. Separately, mIM (656âmg, 8âmmol) was dissolved in 50âmL of MeOH in another 250âmL conical flask with the aid of ultrasound. The mIM solution was then poured into the cobalt-zinc solution under vigorous stirring, which was continued for 3âh. The mixture was allowed to stand undisturbed for 24âh, and the resulting precipitate was recovered by centrifugation. The product was washed three times with MeOH to remove any unreacted metal salts and ligands. ICP-OES analysis indicated that the cobalt percentage in the CoZn-ZIF-8 sample was 0.33%.
Sample activation procedures for NiZn-ZIF-8
The as-synthesized NiZn-ZIF-8 samples were activated to remove residual solvents. The samples were first centrifuged and washed three times with anhydrous MeOH to exchange the pore solvent. The MeOH-exchanged samples were transferred as a suspension to a quartz cell, and the solvent was decanted. The remaining MeOH within the pores was removed by evacuating the samples under vacuum (10-2âTorr) at room temperature for 12âh. This was followed by heating the samples to 120â°C at a controlled rate of 1â°C min-1 for 12âh. The same controlled rate was applied during the subsequent cooling process.
ICP-OES analysis
The elemental content of zinc, nickel, copper, and cobalt in the MOF samples was quantified using ICP-OES. For a typical analysis, 5âmg of the MOF sample was added to 10âmL of a 4% nitric acid (HNO3) aqueous solution. The mixture was sonicated for 10âmin and then heated to 100â°C for 2âh to ensure complete digestion. After cooling to room temperature, the resulting solution was diluted to a total volume of 10âmL and filtered using a 0.22 μm filter to remove particulates. Standard solutions containing 1, 5, 10, 20, 50, 100, and 200 ppm of nickel, copper, cobalt, or zinc in 3% nitric acid were prepared. A calibration curve was generated by plotting chromatographic peak intensities against the known concentrations of the standards, enabling precise quantification of the metal content in the MOF samples.
Thermal gravimetric analysis
TGA measurements were performed using a TGA analyzer, with samples placed in aluminum oxide pans under a continuous airflow atmosphere. The balance gas was nitrogen (N2) at a flow rate of 40.0âmLâmin-1, and the sample gas was air at a flow rate of 60.0âmLâmin-1. Samples were heated at a constant rate of 10â°C min-1 during all TGA experiments.
N2 adsorption analysis
All N2 adsorption experiments were carried out on a Micromeritics 3 Flex automatic volumetric instrument. Measurements were performed at 77âK using a liquid nitrogen bath. Ultra-high purity N2 was used as the adsorbate. Prior to isotherm measurements, ZIF-8 and the series of NiZn-ZIF-8 samples were degassed on the 3 Flex instrument for 10âh at 120â°C. Pore size distributions for the MOFs were analyzed using quenched NLDFT with a carbon slit-pore model.
PXRD crystallography
PXRD data for MOF samples were collected using a Bruker D8 3âkW diffractometer with Cu-Kα1 X-ray radiation (λâ=â1.5406âà ) in transmission geometry.
MOF nanoparticle concentration analysis
NiZn-ZIF-8 nanoparticles (2âmg) were dispersed in 40âmL of pure water and sonicated for 5âmin. A 1âmL aliquot of this stock solution was then diluted to 3âmL with pure water and sonicated for an additional 5âmin. The nanoparticle concentration in the diluted solution was analyzed using a nanoparticle tracking analysis (NTA) system (ZetaView PMX 110). The measurements were performed at a controlled temperature of 23 to 30â°C. The particle concentration was converted to molar concentration using the formula (molar concentration = particle concentration/NA, NA is Avogadroâs constant).
Animal preparation
All procedures were approved by the Review Board of the Innovation Academy for Precision Measurement Science and Technology, Chinese Academy of Sciences, and were performed in accordance with the national Regulations for the Administration of Affairs Concerning Experimental Animals. Healthy male and female SpragueâDawley rats (body mass 240â260âg) were provided by the Hubei Provincial Center for Disease Control and Prevention (Wuhan, Hubei, China). Prior to surgery, the rats were anesthetized with 5% isoflurane (RWD, Shenzhen, Guangdong, China) and maintained under 2% isoflurane anesthesia during the procedure.
Hyperpolarized 129Xe experiments
Hyperpolarized 129Xe gas was generated using a commercial 129Xe hyperpolarizer (âââ40% polarization, verImagin Healthcare, Wuhan, China) in continuous flow mode. All 129Xe NMR and MRI experiments were performed on a 9.4âT Bruker AV400 wide-bore NMR spectrometer (Bruker Biospin, Ettlingen, Germany) or a 7âT animal MRI scanner (Bruker Biospec 70/20 USR, Billerica, MA, USA) equipped with a home-built dual-tuned birdcage coil.
For hyperpolarized 129Xe experiments in solution, a gas mixture of 10% N2, 88% He, and 2% Xe (26.4% natural abundance of 129Xe) from Spectra Gases was used. The gas mixture was maintained at a pressure of 3.5âbar and flowed at 0.1 SLPM. During experiments, after the gas mixture was polarized, it was bubbled directly into a 10âmm NMR tube containing the MOF sample for 20 or 60âs (single scan). A 3âs delay was allowed to collapse the bubbles before acquiring the spectrum using a zg sequence (rectangular pulse, p1â=â31.8 μs). For variable-temperature experiments, temperatures were varied from 278 to 320âK in 2âK increments. A stabilization time of 10âmin was allowed at each temperature before acquiring the 129Xe NMR spectrum. The sample temperature was controlled using a variable temperature (VT) unit integrated into the NMR spectrometer. 129Xe NMR chemical shifts were referenced to the signal of gaseous xenon at 0 ppm.
For hyperpolarized 129Xe MRI experiments in solution, 16 scans were acquired and averaged. Images were collected using a Rapid Acquisition with Relaxation Enhancement (RARE) sequence with the following parameters: slice thickness = 30âmm, matrix size = 32 Ã 32, FOVâ=â30 Ã 30 mm, in-plane resolution = 0.9375 Ã 0.9375 mm, bandwidth = 5400âHz, echo time = 8.5âms, centric k-space encoding, no partial Fourier transform acceleration, rare factor = 4. For each excitation, the gas mixture was bubbled into the solution for 20âs, followed by a 3âs delay to allow bubble collapse, after which the image was acquired.
For in vivo hyperpolarized 129Xe experiments, a gas mixture of 10% N2, 88% He, and 2% Xe (86% enriched abundance of 129Xe) from Spectra Gases was used. Polarized xenon gas was accumulated and frozen in a magnetic field-protected cold finger using liquid nitrogen, then thawed into a 1.0âL Tedlar bag for delivery to the rats by mechanical ventilation. A 3D Ultra-short Echo Time (UTE) simultaneous imaging sequence was used to obtain images of the dissolved xenon (xenon dissolved in TP or RBC), entrapped xenon, and xenon gas at the same time, with the following parameters: TRâ=â250âms, matrix size = 192 à 192 à 192, FOVâ=â15 à 15 à 15âcm, 1000 projections with Fibonacci trajectory, and the reconstructed spatial resolution = 0.47 à 0.47 à 0.47 mm. In Hyper-CEST experiments, the hyperpolarized gas mixture was bubbled into a 10âmm NMR tube for 20âs, followed by a 3âs delay for bubble collapse. A continuous wave (cw) pulse (6.5 µT, 10âs) was applied to saturate xenon within the NiZn-ZIF-8 pores, and the 129Xe NMR spectrum was subsequently acquired in a single scan. Spectral processing included Lorentzian broadening (LBâ=â5âHz).
For detection limit tests, the polarized gas mixture was bubbled into a 10âmm NMR tube for 20âs, followed by a 3âs delay for bubble collapse. A selective saturation pulse was applied, and the spectrum was acquired. The cw-saturation pulses were tuned to either the resonant frequency of xenon within the NiZn-ZIF-8 pores or off-resonance. Saturation times ranged from 0 to 20âs in 1âs increments under a 13 µT field. Spectra were acquired in single scans and processed using Lorentzian broadening (LBâ=â5âHz).
For Hyper-CEST MRI experiments, 8 on-resonant and 8 off-resonant scans were acquired and averaged. Images were collected using a RARE sequence with the following parameters: slice thickness = 30âmm, matrix size = 32 à 32, FOVâ=â30 à 30 mm, in-plane resolution = 0.9375 à 0.9375 mm, bandwidth = 5400âHz, echo time = 4.65âms, centric k-space encoding, no partial Fourier transform acceleration, rare factor = 8. For each excitation, the gas mixture was bubbled into the solution for 20âs, followed by a 3âs delay to allow bubble collapse. A saturation pulse (6âs, 13 µT) was applied to saturate xenon within the NiZn-ZIF-8 pores or off-resonance, after which the image was acquired. MR images were processed in MATLAB (R2014a, MathWorks, Natick, MA). The raw 32 à 32 image matrix was interpolated to a 64 à 64 matrix. Hyper-CEST effects for on-resonant saturation were calculated relative to off-resonant saturation using the formula:
Point-by-point analysis was performed, and a mask was applied to exclude non-sample areas and regions with normalized signal intensities below 0.2.
Hyperpolarized 129Xe two-dimensional exchange spectroscopy (2D EXSY)
Hyperpolarized 129Xe 2D EXSY experiments were conducted at 298âK using a 90°-t1-90°-tm-90°-t2 pulse sequence and the States-TPPI method for phase-sensitive detection in the indirect dimension. The gas mixture was bubbled into the solution for 20âs, followed by a 3âs delay for bubble collapse. The mixing time (tm) was varied from 0 to 20âms, during which transverse magnetization was eliminated using phase cycling in the pulse sequence. The 2D NMR data were acquired with 32 and 256 points in the t1 and t2 dimensions, respectively, with 4 scans per point.
T 1 relaxation measurements for hyperpolarized 129Xe
A small flip angle pulse sequence was used to measure the T1 relaxation time of hyperpolarized 129Xe. After hyperpolarized gas mixture was bubbled into the solution for 20âs, a 3âs delay was allowed for bubble collapse. A small flip angle non-selective pulse was then applied to the sample, and a free induction decay (FID) was collected immediately after the pulse. The signal was allowed to decay before applying another small flip angle pulse to the remaining magnetization, followed by the collection of another FID. This process was repeated for the desired number of points along the decay curve. The collected signal was fitted to a monoexponential decay model described by the equation:
where Mt is the magnetization at time t, M0 is the initial magnetization, and T1 is the spin-lattice relaxation time.
However, the T1 measured with this pulse sequence will be shorter than the T1 measured with a conventional pulse sequence. Every pulse decreases the amount of magnetization, shortening the time needed for the signal to decay to zero. However, the effect of the small flip angle pulses can be corrected by this equation:
In the above equation, T1 is the corrected T1, while T1* is the measured T1. The angle θ is the small flip angle used to sample the magnetization over time and Ï is the time between samplings.
Calculation for signal enhancement fold
The enhancement fold is derived as follows: the mass concentration of the MOF be denoted as c1, and the signal intensity produced by the entrapped xenon in the MOF at this concentration be M1. The signal per unit mass concentration of MOF is therefore M1/c1. For pure water, let the mass concentration be c2, and the signal intensity of the dissolved xenon in pure water be M2. The signal per unit mass concentration of water is M2/c2. Thus, the enhancement fold, which compares the signal intensity per unit mass concentration of entrapped xenon in the MOF to the signal intensity per unit mass concentration of xenon in pure water, is calculated as:
Xenon adsorption analysis and isosteric heat of adsorption
Xenon adsorption/desorption isotherms were performed on a Micromeritics 3 Flex automatic volumetric instrument. Before the isotherm measurement, NiZn-ZIF-8 and pristine ZIF-8 powders were degassed on 3 Flex for 10âh at 120â°C. The analysis temperature was kept at 273 and 298âK, respectively. Isosteric heat of adsorption (Qst) is used to evaluate the adsorption affinity of the gas molecules to the adsorbent material. Higher Qst stands for stronger interactions between the adsorbate molecules and the adsorbent. The Clausius-Clapeyron equation was employed to calculate the Qst:
Where na (mmol g-1) is the amount of adsorbed gas, T (K) is the temperature, P (kPa) is the pressure, Qst (kJ mol-1) is the isosteric heat of adsorption.
Molecular dynamics simulations
The initial structures for the MD simulations were constructed by three regions, namely as bath, MOF, and bath regions, which are segregated equally along the Z-axis. In the MOF region, the structures were adopted from the Cambridge Crystallographic Data Centre (CCDC No. 602542, http://www.ccdc.cam.ac.uk). The structure with a volume of 34 Ã 34 Ã 34 Ã 3 is composed of 2 Ã 2 Ã 2 unit cells with the periodic structures in X-, Y-, and Z-axes. Terminal groups along in Z-axis were saturated by adding methyl (âââCH3) groups. Each constructed structure was positioned at the center of the simulation box with dimensions 34 Ã 34 Ã 102 Ã 3, surrounded by 60 xenon atoms (Supplementary Fig. 85), using the Universal Force Field (UFF).
Statistical analysis
Data are expressed as meanâ±âstandard deviation (SD) from at least three independent replicates. Statistical analysis was conducted using one-way analysis of variance (ANOVA), and two tailed Studentâs t-test was used for two-group comparisons, all performed within OriginPro 2021. Significance levels are indicated by asterisks: a p-value of less than 0.05 is the benchmark for statistical significance, ** for pâ<â0.01, and *** for pâ<â0.001.
Data availability
The data that support the findings of this study, including the full image dataset, are available from the corresponding authors upon request. Source data are provided with this paper.
References
Zeng, Z., Liew, S. S., Wei, X. & Pu, K. Hemicyanine-based near-infrared activatable probes for imaging and diagnosis of diseases. Angew. Chem. Int. Ed. 60, 26454 (2021).
Wang, X. et al. Fluorescent probes for disease diagnosis. Chem. Rev. 124, 7106â7164 (2024).
Mugler, J. P. & Altes, T. A. Hyperpolarized 129Xe MRI of the human lung. J. Magn. Reson. Imaging 37, 313â331 (2013).
Li, H. et al. Damaged lung gas exchange function of discharged COVID-19 patients detected by hyperpolarized 129Xe MRI. Sci. Adv. 7, eabc8180 (2021).
Fang, Y. et al. Rapid pulmonary 129Xe ventilation MRI of discharged COVID-19 patients with Zigzag sampling. Magn. Reson. Med. 92, 956â966 (2024).
Rao, Q. et al. Assessment of pulmonary physiological changes caused by aging, cigarette smoking, and COPD with hyperpolarized 129Xe magnetic resonance. Eur. Radiol. 34, 7450â7459 (2024).
Zhou, Q. et al. Evaluation of injuries caused by coronavirus disease 2019 using multi-nuclei magnetic resonance imaging. Magn. Reson. Lett. 1, 2â10 (2021).
Doganay, O. et al. Time-series hyperpolarized xenon-129 MRI of lobar lung ventilation of COPD in comparison to V/Q-SPECT/CT and CT. Eur. Radiol. 29, 4058â4067 (2019).
Wang, J. M. et al. Using hyperpolarized 129Xe MRI to quantify regional gas transfer in idiopathic pulmonary fibrosis. Thorax 73, 21â28 (2018).
Rose, H. M. et al. Development of an antibody-based, modular biosensor for 129Xe NMR molecular imaging of cells at nanomolar concentrations. Proc. Natl. Acad. Sci. USA 111, 11697â11702 (2014).
Witte, C. et al. Live-cell MRI with xenon Hyper-CEST biosensors targeted to metabolically labeled cell-surface glycans. Angew. Chem. Int. Ed. 54, 2806â2810 (2015).
Klippel, S. et al. Cell tracking with caged xenon: using cryptophanes as MRI reporters upon cellular internalization. Angew. Chem. Int. Ed. 53, 493â496 (2014).
Wang, Y., Roose, B. W., Philbin, J. P., Doman, J. L. & Dmochowski, I. J. Programming a molecular relay for ultrasensitive biodetection through 129Xe NMR. Angew. Chem. Int. Ed. 55, 1733â1736 (2016).
Zhang, B. et al. An intracellular diamine oxidase triggered hyperpolarized 129Xe magnetic resonance biosensor. Chem. Commun. 54, 13654â13657 (2018).
Zhou, X., Graziani, D. & Pines, A. Hyperpolarized xenon NMR and MRI signal amplification by gas extraction. Proc. Natl. Acad. Sci. USA 106, 16903â16906 (2009).
Schröder, L., Lowery, T. J., Hilty, C., Wemmer, D. E. & Pines, A. Molecular imaging using a targeted magnetic resonance hyperpolarized biosensor. Science 314, 446â449 (2006).
Chen, Z. et al. Fine-tuning a robust metalâorganic framework toward enhanced clean energy gas storage. J. Am. Chem. Soc. 143, 18838â18843 (2021).
Smoljan, C. S. et al. Engineering metalâorganic frameworks for selective separation of hexane isomers using 3-dimensional linkers. J. Am. Chem. Soc. 145, 6434â6441 (2023).
Liu, Y. et al. Remarkably enhanced gas separation by partial self-conversion of a laminated membrane to metalâorganic frameworks. Angew. Chem. Int. Ed. 54, 3028â3032 (2015).
Huang, J. et al. A microporous hydrogen-bonded organic framework based on hydrogen-bonding tetramers for efficient Xe/Kr separation. Angew. Chem. Int. Ed 62, e202315987 (2023).
Jiang, J., Furukawa, H., Zhang, Y.-B. & Yaghi, O. M. High methane storage working capacity in metal-organic frameworks with acrylate links. J. Am. Chem. Soc. 138, 10244â10251 (2016).
Wang, H. et al. Docking of Cu(I) and Ag(I) in metalâorganic frameworks for adsorption and separation of xenon. Angew. Chem. Int. Ed 60, 3417â3421 (2021).
Banerjee, D. et al. Potential of Metalâorganic frameworks for separation of xenon and krypton. Acc. Chem. Res. 48, 211â219 (2015).
Deng, H. et al. Multiple functional groups of varying ratios in metal-organic frameworks. Science 327, 846â850 (2010).
Dong, Z., Sun, Y., Chu, J., Zhang, X. & Deng, H. Multivariate metalâorganic frameworks for dialing-in the binding and programming the release of drug molecules. J. Am. Chem. Soc. 139, 14209â14216 (2017).
Zhang, Z., Hanikel, N., Lyu, H. & Yaghi, O. M. Broadly tunable atmospheric water harvesting in multivariate metalâorganic frameworks. J. Am. Chem. Soc. 144, 22669â22675 (2022).
Fan, W. et al. Optimizing multivariate metal-organic frameworks for efficient C2H2/CO2 separation. J. Am. Chem. Soc. 142, 8728â8737 (2020).
Zhang, Y. et al. Introduction of functionality, selection of topology, and enhancement of gas adsorption in multivariate metal-organic framework-177. J. Am. Chem. Soc. 137, 2641â2650 (2015).
Fan, W. et al. Multivariate polycrystalline metal-organic framework membranes for CO2/CH4 separation. J. Am. Chem. Soc. 143, 17716â17723 (2021).
Chen, C.-X. et al. Enhancing photocatalytic hydrogen production via the construction of robust multivariate Ti-MOF/COFÂ composites. Angew. Chem. Int. Ed. 61, e202114071 (2022).
Wang, Y. et al. A tunable multivariate metal-organic framework as a platform for designing photocatalysts. J. Am. Chem. Soc. 143, 6333â6338 (2021).
Deliere, L. et al. Role of silver nanoparticles in enhanced xenon adsorption using silver-loaded zeolites. J. Phys. Chem. C 118, 25032â25040 (2014).
Daniel, C. et al. Xenon capture on silver-loaded zeolites: characterization of very strong adsorption sites. J. Phys. Chem. C 117, 15122â15129 (2013).
Bachetzky, C. et al. Adsorption of xenon, krypton, and their mixture on the flexible metalâorganic framework DUT-8(Ni). J. Phys. Chem. C 128, 6997â7006 (2024).
Xiong, S. et al. A microporous metalâorganic framework with commensurate adsorption and highly selective separation of xenon. J. Mater. Chem. A 6, 4752â4758 (2018).
Zeng, D. et al. Metalâorganic frameworks possessing suitable pores for Xe/Kr separation. Inorg. Chem. 63, 5151â5157 (2024).
Zeng, Q. et al. Hyperpolarized Xe NMR signal advancement by metal-organic framework entrapment in aqueous solution. Proc. Natl. Acad. Sci. USA 117, 17558â17563 (2020).
Yang, Y. et al. Coloring ultrasensitive MRI with tunable metal-organic frameworks. Chem. Sci. 12, 4300â4308 (2021).
Spence, M. M. et al. Functionalized xenon as a biosensor. Proc. Natl. Acad. Sci. USA 98, 10654â10657 (2001).
Kunth, M., Witte, C., Hennig, A. & Schröder, L. Identification, classification, and signal amplification capabilities of high-turnover gas binding hosts in ultra-sensitive NMR. Chem. Sci. 6, 6069â6075 (2015).
Li, R. et al. Nickel-substituted zeolitic imidazolate frameworks for time-resolved alcohol sensing and photocatalysis under visible light. J. Mater. Chem. A 2, 5724â5729 (2014).
Stevens, T. K., Ramirez, R. M. & Pines, A. Nanoemulsion contrast agents with sub-picomolar sensitivity for xenon NMR. J. Am. Chem. Soc. 135, 9576â9579 (2013).
Shapiro, M. G. et al. Genetically encoded reporters for hyperpolarized xenon magnetic resonance imaging. Nat. Chem. 6, 630â635 (2014).
Zeng, Q. et al. Mitochondria targeted and intracellular biothiol triggered hyperpolarized 129Xe magnetofluorescent biosensor. Anal. Chem. 89, 2288â2295 (2017).
Wang, Y. & Dmochowski, I. J. Cucurbit[6]uril is an ultrasensitive Xe-129 NMR contrast agent. Chem. Commun. 51, 8982â8985 (2015).
Zeng, Q. et al. Ultrasensitive molecular building block for biothiol NMR detection at picomolar concentrations. iScience 24, 103515 (2021).
Schilling, F. et al. MRI thermometry based on encapsulated hyperpolarized xenon. ChemPhysChem 11, 3529â3533 (2010).
Roukala, J. et al. Encapsulation of xenon by a self-assembled Fe4L6 metallosupramolecular cage. J. Am. Chem. Soc. 137, 2464â2467 (2015).
Du, K., Zemerov, S. D., Parra, S. H., Kikkawa, J. M. & Dmochowski, I. J. Paramagnetic organocobalt capsule revealing xenon hostâguest chemistry. Inorg. Chem. 59, 13831â13844 (2020).
Du, K., Zemerov, S. D., Carroll, P. J. & Dmochowski, I. J. Paramagnetic shifts and guest exchange kinetics in ConFe4ân metalâorganic capsules. Inorg. Chem. 59, 12758â12767 (2020).
Jayapaul, J. et al. Hyper-CEST NMR of metal organic polyhedral cages reveals hidden diastereomers with diverse guest exchange kinetics. Nat. Commun. 13, 1708 (2022).
Tallavaara, P. & Jokisaari, J. 2D 129Xe EXSY of xenon atoms in a thermotropic liquid crystal confined to a controlled-pore glass. Phys. Chem. Chem. Phys. 8, 4902â4907 (2006).
Acknowledgements
This work is supported by the Strategic Priority Research Program, CAS (XDB0540000), National Natural Science Foundation of China (82127802, 82441015), Key Research Program of Frontier Sciences, CAS (ZDBS-LY-JSC004), Hubei Provincial Key Technology Foundation of China (2021ACA013), Major Program (JD) of Hubei Province (2023BAA021), National Key Research and Development Program of China (2022YFC2410000), Hubei Provincial Natural Science Foundation of China (2022CFA050). Q.G. and L.S.B. acknowledge the support from the CAS Youth Interdisciplinary Team (JCTD-2022-13). Q.Z. acknowledges the support from the Youth Innovation Promotion Association, CAS (2023347), WUHAN TALENT, and the Hubei Provincial Natural Science Foundation of China (2023AFB790), Y. Yang acknowledges the support from the National Natural Science Foundation of China (22274162).
Author information
Authors and Affiliations
Contributions
X. Zhou, Q.G. and Q.Z. conceived the research. Q.G. and Q.Z. designed the experiments. Q.Z., Q.Y., M.Z., R.Z., Z.W., Z. X., Y. Yuan, and X. Zhao performed all the experiments. Q.Z., W.J., Y. Yang, and H.L. analyzed the data. Q.Z. wrote the original manuscript. Q.G., L.Z., X. Zhou, M.Y., and L.-S.B. edited the paper. Q.Z. and Q.Y. contributed equally to this work. All authors discussed the results and commented on the paper.
Corresponding authors
Ethics declarations
Competing interests
The authors declare no competing interests.
Peer review
Peer review information
Nature Communications thanks Ivan Dmochowski who co-reviewed with Yannan Lin; and the other anonymous reviewer(s) for their contribution to the peer review of this work. A peer review file is available.
Additional information
Publisherâs note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Source data
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the articleâs Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the articleâs Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
About this article
Cite this article
Zeng, Q., Yue, Q., Zhang, M. et al. Multivariate metal-organic frameworks enable chemical shift-encoded MRI with femtomolar sensitivity for biological systems. Nat Commun 16, 6832 (2025). https://doi.org/10.1038/s41467-025-62110-4
Received:
Accepted:
Published:
DOI: https://doi.org/10.1038/s41467-025-62110-4







