Skip to main content
  • Critical Review
  • Open access
  • Published:

Multi-nuclear magnetic resonance spectroscopy: state of the art and future directions

Abstract

With the development of heteronuclear fluorine, sodium, phosphorus, and other probes and imaging technologies as well as the optimization of magnetic resonance imaging (MRI) equipment and sequences, multi-nuclear magnetic resonance (multi-NMR) has enabled localize molecular activities in vivo that are central to a variety of diseases, including cardiovascular disease, neurodegenerative pathologies, metabolic diseases, kidney, and tumor, to shift from the traditional morphological imaging to the molecular imaging, precision diagnosis, and treatment mode. However, due to the low natural abundance and low gyromagnetic ratios, the clinical application of multi-NMR has been hampered. Several techniques have been developed to amplify the NMR sensitivity such as the dynamic nuclear polarization, spin-exchange optical pumping, and brute-force polarization. Meanwhile, a wide range of nuclei can be hyperpolarized, such as 2H, 3He, 13C, 15 N, 31P, and 129Xe. The signal can be increased and allows real-time observation of biological perfusion, metabolite transport, and metabolic reactions in vivo, overcoming the disadvantages of conventional magnetic resonance of low sensitivity. HP-NMR imaging of different nuclear substrates provides a unique opportunity and invention to map the metabolic changes in various organs without invasive procedures. This review aims to focus on the recent applications of multi-NMR technology not only in a range of preliminary animal experiments but also in various disease spectrum in human. Furthermore, we will discuss the future challenges and opportunities of this multi-NMR from a clinical perspective, in the hope of truly bridging the gap between cutting-edge molecular biology and clinical applications.

Key points

  • The multi-NMR by HP has been applied in many diseases. It has allowed for real-time assessment of human metabolism in vivo and can promote a shift from morphological to molecular level.

  • Work is ongoing in clinical translation of multi-NMR, with numerous spectrum of human diseases, and bridging the gap between the molecular level imaging and clinical applications will foster and introduce future opportunities in this field.

Background

Since 1938, the phenomenon of NMR discovered by Rabi when sent a beam of molecules through magnetic field and found that they could be made to emit radio waves at specific frequencies, great advances in the research and clinical applications of magnetic resonance (MR) imaging have been obtained and leaved remarkable contributions in human health [1]. At present, the development of medical imaging is not only confined to observation of lesions from morphological changes, but also toward a more in-depth comprehensive assessment of biological behavior, pathophysiological, and metabolic changes, and thus to achieve a more precise personalized medicine. With the development of MR hardware (magnets, gradients, radiofrequency coils) and software (pulse sequences, data reconstruction algorithms), the clinical application of multi-nuclear magnetic resonance spectroscopy provides the possibility of real-time dynamic visualization of the metabolic changes and specific substrate to metabolic conversion of the lesion in vivo, which also pushes the MR imaging into a more in-depth molecular level [2].

In theory, any nucleus that can be detected by NMR can be imaged with MR imaging. 1H is the mostly used atom because of the high concentration in human body and large value of the gyromagnetic ratio which enables to achieve high-resolution T1- or T2-weighted MR images. However, the complicated background signal, signal overlaps, and the ability to only reflect a few molecules such as such as choline, creatine, NAA, glycine, myo-inositol, lactate, alanine, and acetate for 1H also limit its application in the real-time metabolic activity. Using of higher magnetic fields, developing of new detectors, increasing the signal with hyperpolarization, and obtaining the signal with ultra-fast pulse sequences make the other nucleus imaging spectroscopy like 2H, 13C, 15 N, 18F, 23Na, 31P, and 129Xe come true [3,4,5]. These multi-NMR spectroscopies provide the ability to insight the multiple metabolic activities and reflect the real metabolism in vivo and have the potential to really bridge the gap between the cutting-edge molecular biology, biochemistry, and metabolism research.

In this review, we will provide an overview of the clinical application of multi-nuclear magnetic resonance spectroscopy in the various biologic processes with different lesions. In addition, we will also discuss the future challenges and opportunities of this novel imaging technique from a clinical perspective, hoping to real bridge the gap between the cutting-edge molecular biology and clinical applications.

Nervous system

Metabolic abnormalities are key factors in many neurological disorders. The ability to accurately measure this metabolic damage could improve the detection and diagnosis of disease progression and facilitate the development of new treatments accordingly. Hyperpolarized (HP)-carbon-13 magnetic resonance spectroscopic imaging (MRSI) is an emerging imaging technique that allows the noninvasive, real-time measurement of enzyme activity in living organisms (Fig. 1). To date, this metabolic imaging method has been used mainly in the fields of cancer and cardiology and has recently begun to be used to study neurological diseases [6, 7]. HP-13C MRSI has promising applications in neurodegenerative diseases, traumatic brain injury, stroke, and brain tumors [8].

Fig. 1
figure 1

The workflow of hyperpolarized carbon-13 MRI in biomedical imaging

Neurodegenerative diseases

Mononuclear phagocytes (MPs) play a crucial role in the progression of multiple sclerosis (MS) and other neurodegenerative diseases [9,10,11]. Following activation and subsequent differentiation toward a proinflammatory phenotype, abnormalities in MPs metabolism occur, leading to increased glycolysis and lactate production. HP-13C MRSI is a clinically translatable imaging modality that allows for the noninvasive detection of metabolic pathways in real time. 13C MRSI not only tracks the Warburg effect within the tumor but also allows for a realistic neurological inflammatory response. The significant increase in HP-[1-13C]-pyruvate to HP-[1-13C]-lactate conversion, on the one hand, indicated the presence of a high density of proinflammatory MPs. On the other hand, it suggested that the increase in HP-[1-13C]-lactate may be mediated by upregulation of pyruvate dehydrogenase (PDH) kinase-1 in activated MPs leading to regional PDH inhibition [12]. Overall, the preclinical results of the cuprizone model demonstrated the potential of 13C MRSI of HP-[1-13C]-pyruvate as a neuroimaging method for the assessment of inflammatory lesions. This method could prove useful not only for multiple sclerosis, but also for other neurological disorders with an inflammatory component in the future.

Traumatic disease

Traumatic brain injury (TBI) can lead to disturbances in energetic metabolism in the brain but monitoring of metabolic activity is currently achieved mainly by microdialysis and 18F-fluorodeoxyglucose (18F-FDG) positron emission tomography (PET) [13]. Proton magnetic resonance spectroscopy (1H MRS) also detected focal or systemic elevated lactate and related PDH caused by primary or secondary injured inflammation [14, 15]. HP-13C MRSI could be used as a direct, rapid, and noninvasive method to explore the effects of TBI on energetic metabolism in the brain [16, 17]. In rats with moderate TBI induced by control cortical impingement on one cerebral hemisphere, measured by injection of HP-[1-13C]-pyruvate, the injured side of the brain was found to produce a 13C-bicarbonate signal 24 ± 6% lower than the injured side, while the HP-bicarbonate-to-HP-lactate ratio was 33 ± 8% lower than that of the injured side [16]. In controls, there were no significant differences in signals between the two cerebral sides. The results suggest that mitochondrial pyruvate metabolism is impaired, resulting in reduced aerobic respiration at the injury site after TBI. Clinically, HP-[1-13C]-pyruvate MRSI is used to image patients with acute mild TBI several days after head trauma, but with no obvious anatomic changes. One patient showed high levels of lactic acid production at the injured site, and both patients showed a notable reduction in bicarbonate production [17] (Fig. 2). This study indicates using HP-pyruvate feasibly to image the metabolic changes in TBI patients and proves the transformability and sensitivity of this technology to the changes in cerebral metabolism after mild TBI.

Fig. 2
figure 2

Reproduced with permission from Elsevier Ltd (License Number: 5324100264979). iScience. 2020 Nov 30;23(12):101,885

Metabolism of acute traumatic brain injury (TBI) imaged by hyperpolarized 13C-pyruvate. The patient (35 years old, male), injured on the left frontal scalp, was imaged 28 h after the injury. a An axial slice that includes the injured region was prescribed for 13C imaging. b T1-weighted and T2-weighted 1 H FLAIR and (c) ASL images showed swelling and increased perfusion in the injured site outside the skull (yellow circle), whereas no cerebral contusion, hemorrhage or hypoperfusion were detected. d Region with increased total hyperpolarized 13C signal image matched to the scalp edema. e Increased [1-13C]lactate conversion and (f) decreased [13C]HCO3 – production were detected in the brain tissues of the injured hemisphere. g Averaged spectra over the injured brain region and the contralateral normal-appearing brain region.

Stroke

Ischemic stroke is a progressive process characterized by impaired but reversible damage to the metabolic activity of the penumbra tissue so that if blood flow is restored, full recovery of function is possible. Currently, ischemic cases are identified by a mismatched area between perfusion and diffusion on MRI. Conventional perfusion and diffusion proton MR images have a large background signal, and the commonly used contrast injection is an invasive method, but many patients with the renal injury cannot use contrast agents and results have shown that gadolinium contrast agents can cause nephrogenic systemic fibrosclerosis. Due to the high rate of glucose extraction in the penumbra, characterized by anaerobic glycolysis and subsequent increased lactate production, this can be detected by the altered metabolism of HP-13C-pyruvate. HP-13C MRI may therefore have an important role in understanding the development and response of the ischemic putative difficult zone in stroke and is also translatable to concurrent studies of metabolism and perfusion in humans [18] (Fig. 3). In preclinical trials, an endothelin-1-induced ischemic stroke model was used to assess metabolic alterations after acute ischemic stroke in rats [19]. The results showed increased total lactate production in the ischemic hemispheric dark zone compared to the contralateral brain. This was the result of a combination of increased pyruvate in the blood and lactate dehydrogenase (LDH)-mediated lactate production. When estimating penumbra viability, measures of penumbra metabolic activity may be more sensitive than conventional imaging techniques. Likewise, since the 129Xe signal is proportional to cerebral blood flow as a perfusion missing agent for cerebral blood flow, HP-129Xe MRI can also be used for the study of ischemic semi-dark areas as a perfusion missing agent of cerebral blood flow. The high lipid solubility, lack of biological tissue background signals, and high sensitivity of HP-129Xe allow the detection of HP-129Xe signals even at low concentrations, which is advantageous for the identification of under-perfused brain tissue. Using HP-129Xe MRI, Zhou et al. [20] successfully detected a local decrease in cerebral blood flow and studied ischemic lesions in the core of stroke by performing a permanent middle cerebral artery occlusion on the right brain of rats. This demonstrated that HP-129Xe MRI can be used as a complementary tool to conventional MRI to study the pathophysiology of cerebral under perfusion.

Fig. 3
figure 3

Iterative Decomposition with Echo Asymmetry and Least squares estimation (IDEAL) spiral 13C imaging demonstrating metabolite distribution in the healthy human brain. NeuroImage Volume 189, 1 April 2019, Pages 171–179

Brain tumor

The current “open questions” in neuro-oncology imaging are to evaluate tumor response during and after treatment on clinically relevant timescales, to differentiate between recurrent or residual tumors and treatment-related changes with high specificity, to classify tumor pathological aberrations at the molecular level, and to predict tumor response to treatment. Metabolic biomarkers could potentially address these issues and could in principle monitor response before genetic alterations or more delayed morphological changes or recurrence.

Hydrogen is the basic building block of biological organic molecule, and therefore, deuterium (hydrogen 2 (2H)) is characterized by nondestructive labeling, labeling flexibility, and potential adaptability. At present, in the early stages of research, animal studies have shown that 2H MRSI and MRI can be used to quantify 2H-labeled glucose metabolism in tumors [21,22,23].

HP-[1-13C]-pyruvate MRI is a novel method for demonstrating energetic metabolism in the human brain and brain tumors, in contrast to the clinical needs that cannot be met by current conventional imaging techniques [24,25,26,27,28] (Fig. 4). HP-pyruvate MRSI was a valuable tool for monitoring the phosphatidylinositol 3-kinase inhibitor LY294002 in mouse models of malignant glioma and breast cancer, where inhibition of the phosphatidylinositol 3-kinase pathway led to reduced LDH activity, thereby reducing the signal for HP conversion of pyruvate to lactate [29]. Park et al. also used HP-13C-pyruvate MRSI to assess the prognosis of a rat glioma model [24]. The signal-to-noise ratio of 13C-pyruvate and its metabolic13C-lactate were both significantly higher than that of normal brain tissue, where the blood–brain barrier (BBB) was able to restrict the entry of pyruvate into brain cells, whereas the disruption of the BBB in gliomas resulted in the smooth entry of 13C-pyruvate into brain tissue, which had lagged its research in oncology and cardiology. Nevertheless, Hurd et al. found that injections of [1-13C]-ethyl pyruvate (a lipophilic pyruvate derivative) were able to cross the BBB faster than pyruvate and had a significantly higher polarization ratio than total carbon in brain tissue [25], opening a new direction for HP MRSI in the brain. In preliminary studies performed in patients with primary brain tumors, HP-[1-13C]-pyruvate rapidly crossed the BBB and was converted to the metabolites lactate and bicarbonate. Non-tumor areas showed a high bicarbonate signal, while tumor areas had a low bicarbonate signal, suggesting that this technique may be valuable for exploring mitochondrial oxidative metabolism and its alterations in brain tumors [26]. Another recent study similarly formalized the feasibility of obtaining high-quality metabolite maps in patients with primary or metastatic brain tumors [27]. These initial clinical studies demonstrate the feasibility of HP-[1-13C]-pyruvate MRI for cerebral metabolism assessment.

Fig. 4
figure 4

Pyruvate metabolism in a recurrent glioblastoma and treatment-related changes (Patient 2, a and b) and anaplastic oligodendroglioma (Patient 1, c and d). An enhancing mass is located in the right cingulate gyrus (green oval) (a). The hyperpolarized pyruvate map shows high signal corresponding to the cortex/juxtacortical regions and superior sagittal sinus. The volume normalized pyruvate signal in the lesion is 1.3-fold higher than the entire brain. The HP lactate map shows mildly elevated signal corresponding to the lesion, similar to background lactate production by the brain. The volume normalized lactate signal in the lesion is 1.0-fold of the lactate over the entire brain (cortex and white matter) (b). The T2-FLAIR image shows an expansile left frontal mass and contracting right frontal hematoma. The left frontal mass has minimal, patchy enhancement and a small component of elevated plasma volume (green oval) visualized by DCE MRI (c). The HP pyruvate map shows high signal corresponding to the cortex/juxtacortical regions and superior sagittal sinus, but relatively lower signal corresponding to the small hyperperfused component. The HP lactate map shows mildly elevated signal corresponding to the hyperperfused component, similar to background lactate production by the brain (d). Cancer Res. 2018;78(14):3755–3760

Genitourinary system disease

Tumors

In prostate cancer, antisense oligonucleotides targeting monocarboxylate transporter (MCT)-4 that inhibit lactate secretion may be useful in treating castration-resistant prostate cancer and other cancers because they can interfere with the reprogrammed energetic metabolism of cancer, an emerging hallmark of cancer [30]. Recent preclinical experiments demonstrated that MCT1 regulated pyruvate transport into the cell, MCT4 regulated lactate transport outside the cell body and MCT1 expression had a strong positive correlation with Gleason grade of prostate cancer [31]. Similarly, Keshari et al. used the HP-13C-pyruvate co-reactor to study renal cancer cells and found that metabolic pathways could be observed in real time and that changes in the microenvironment caused by oxidation, flux, pH, and substrate could be controlled, revealing the importance of MCT4 transport in disease onset and progression and that MCT4-mediated regulation of lactate transport could be detected by HP-13C MRSI [32]. Sriram et al. found that in malignant cells, MCT4 transport of lactate to the extracellular compartment was increased, so the ratio of intracellular to extracellular lactate predicted the aggressiveness of tumor cells [33]. Another research confirmed by HP-13C-pyruvate metabolism in in vivo renal tumor tissue that, compared with benign renal tumors, approximately 70–80% of renal cell carcinomas (RCC) had increased lactate production and extracellular translocation was increased, and its aberrant metabolism was mainly caused by increased expression of LDHA and MCT4 [34]. Therefore, benign renal tumors can be differentiated from RCC by assessing the rate of HP-[1-13C]-pyruvate-to-lactate conversion and extracellular transport of lactate. In addition, HP-[1-13C]-pyruvate MRI can noninvasively assess the aggressiveness of RCC. Sriram et al. [35] found that the expression of LDHA-catalyzed conversion of pyruvate to lactate varied among tumor lines, and HP- [1-13C]-pyruvate MRI and diffusion-weighted imaging were used to differentiate between benign renal tumors and renal cell carcinomas by tumor [1-13C]-pyruvate conversion rate, LDHA expression rate, and [1-13C]-lactate apparent diffusion coefficient values to assess the aggressiveness of renal tumors and to guide the corresponding clinical treatment.

The first clinical trials of HP-[1-13C]-pyruvate MRI and MRSI were successfully used in prostate cancer patients [36]. Recently, Sushentsev N et al. [37] also found that HP 13C-MRI can noninvasively differentiate indolent from aggressive prostate tumors based on their characteristic metabolic features (Fig. 5). Preclinical studies have shown that elevated [1-13C]-lactate is observed in animal models of prostate cancer and that the [1-13C]-lactate-to-[1-13C]-pyruvate ratio at the cancer site increased with increasing cancer stage; this ratio decreased with effective treatment [38,39,40]. This clinical trial study firstly demonstrated that HP-[1-13C]-pyruvate MRSI could be used safely and effectively in humans and secondly observed a significant [1-13C]-lactate peak and an increase in the [1-13C]-lactate-to-[1-13C]-pyruvate ratio in the lesion, while healthy prostate tissue and surrounding vascular tissue showed no or very low [1-13C]-lactate peaks [36]. This result is consistent with the results of preclinical experimental studies [38]. Albers et al. performed HP-[1-13C]-pyruvate MRSI in a transgenic mouse model of prostate cancer and found significant differences in the lactate content of highly differentiated prostate cancer, poorly differentiated prostate cancer, normal tissue surrounding the prostate, and metastatic lymph nodes [38]. Another noteworthy finding is that in a patient with biopsy-confirmed bilateral prostate cancer foci, conventional T2-weighted image, ADC image, and 1H MRSI on staging showed lesions only on the right side of the gland, while [1-13C]-pyruvate MRSI monitored areas of high [1-13C]-lactate-to-[1-13C]-pyruvate signal ratio on both the left and right sides of the gland, which was later confirmed by MR guided re-biopsy that confirmed a bilateral cancer focus [36]. This heralds the value of [1-13C]-pyruvate metabolic imaging for noninvasive tumor diagnosis, demonstrating that metabolic changes precede changes in MRI signal. In addition, the surprising finding that HP-[1-13C]-pyruvate MRSI can detect bilateral cancers may become a particularly important monitoring tool for patients in the slow-growing cancer category. Furthermore, for assessing tumor response to treatment, HP-[1-13C]-pyruvate MRSI could detect reduced pyruvate–lactate conversion early after treatment of prostate cancer [41], though low uptake and high background uptake in primary prostate cancer tissue when detected by 18F-FDG PET uptake imaging modalities observe the metabolic response of cancer cells more difficult [42]. Therefore, HP-[1-13C]-pyruvate MRSI has the potential to provide an early assessment of tumor treatment efficacy.

Fig. 5
figure 5

Representative example of an MR-occult transition zone tumor detected on hyperpolarized 13C-MRI. a Post-surgical histopathological assessment confirmed the diagnosis of multifocal adenocarcinoma of the prostate. An International Society of Urological Pathology (ISUP) grade 3 target lesion was visible on 1H-MRI in the left peripheral zone: Prostate Imaging-Reporting and Data System (PI-RADS) 5 target lesion; black region of interest (ROI). b Standard-of-care T2-weighted MRI demonstrating a marked area of low signal intensity corresponding to the target lesion in the left peripheral zone. c ADC map demonstrating a corresponding focus of markedly restricted diffusion in the left peripheral zone. d Dynamic contrast-enhanced (DCE) MRI demonstrating the area of early enhancement in the left peripheral zone. e Pyruvate signal-to-noise ratio (SNR) map with two areas of high pyruvate signal, both corresponding to histopathology-confirmed tumor foci. f Lactate SNR map demonstrating high [1-13C]lactate signal in the grade 3 left peripheral zone lesion. g Total carbon SNR map showing higher signal in the left peripheral zone tumor. h Kpl map (presented as s − 1) showing a higher rate of pyruvate-to-lactate conversion in the more aggressive left peripheral zone lesion. Nat Commun. 2022 Jan 24;13(1):466

Other kidney lesions

Alport syndrome (AS) is a genetic disorder characterized by impaired renal function. The development of noninvasive tools for early diagnosis and monitoring of renal function during disease progression is of clinical importance. HP-13C MRI is an emerging technology that allows noninvasive, real-time measurement of metabolism in vivo. A study investigated the feasibility of using this technique to assess changes in renal metabolism in a mouse model of AS. Mice with AS had significantly lower lactate levels from 4 to 7 weeks of age, whereas control mice had no change in lactate levels over time. The results of this study suggest that HP-13C MRI may provide a potential noninvasive tool for characterizing disease progression in AS. Furthermore, HP-13C MRSI was used for noninvasive assessment of oxidative stress and mitochondrial PDH activity after ischemia–reperfusion injury in the kidney. The changes in renal redox capacity and mitochondrial PDH activity on day 7 after injury in unilateral rat kidneys were observed by HP-[1-13C]-dehydroascorbate and13C-pyruvate MRSI, respectively, and it was found that on day 7 of ischemia–reperfusion injury, renal 13C-vitamin C-to-vitamin C plus dehydroascorbate ratio and 13C-bicarbonate-to-13C-pyruvate ratio were reduced, consistent with their low redox capacity, and PDH activity was impaired [43].

Liver diseases

Metabolic liver disease

A growing body of studies suggests that HP 13C MRI is promising for assessing liver metabolism, diagnosing liver injury by mapping the dynamic conversion of HP 13C substrate [44,45,46,47]. Evidences indicates that increased [1-13C]pyruvate-to-alanine-and-lactate conversion was observed in a model of chemically induced inflammatory liver injury, and the conversion of [1-13C]DHA-to-vitamin C was noted reduction in a diet-induced steatohepatitis model [2, 46, 48]. Jeong et al. found that the increased levels of alanine and lactate are useful biomarkers for fatty liver diseases. Merritt et al. studied the metabolism of HP [1-13C]pyruvate in tricarboxylic acid (TCA) cycle in rat livers and indicated that the major pathway for entry of HP [1-13C]pyruvate into the hepatic TCA cycle is via pyruvate carboxylase, and that cataplerotic flux through phosphoenolpyruvate carboxykinase (PERCK) is the primary source of [13C]bicarbonate [44, 49]. Another study conducted by Emine et al. [50] also showed that [13C]bicarbonate labeled from hyperpolarized [1- 13 C]pyruvate is an in vivo marker of hepatic gluconeogenesis in fasted state. Furthermore, the metabolic changes can be also quantified by HP 13C [45, 51,52,53]. Smith et al. [54]. measured the time to peak (TTP) of HP 13C-pyruvate and its metabolites in non-alcoholic fatty liver disease (NAFLD) and normal pigs, and their results revealed that the decreased liver lactate TTP in 13C spectra in NAFLD pigs, indicating an increased rate of lactate production and a disturbance in liver lipid synthesis. These results suggested that HP 13C can be used as a method for assessing in vivo hepatic metabolism and may have the potential to address a currently unmet clinical need for investigating mechanisms of fatty liver diseases (FLD), noninvasive evaluation of these pathways during the detection and treatment of FLD in general.

Hepatocellular carcinoma

Cancer development is a multi-step process involving genetic and cellular alterations, and the multi-step hemodynamic changes for HCC during hepatocarcinogenesis have been well documented [55]. The routine used diagnostic criteria mainly refers on the vascular criteria due to the angiogenesis or neovascularization and the portal flow diminishes. However, since HCC occurs mostly within the liver parenchyma in cirrhosis, its structure is heterogeneous and manifests as multiple mass-like nodules [56]. Thus, the detection of small or very early HCC can be extremely challenging with routine imaging modalities using the vascular criteria. Using HP 13C MRI, Hu et al. [57] examined a model of Myc gene-induced liver cancer and found that changes in tumor metabolism precede all observable morphological and histological changes. In particularly, their results also showed that the conversion of pyruvate to alanine was significantly increased in precancerous tissue and which has become the highest alanine content region, whereas alanine was absent in healthy or established tumors. In addition to the application of early HCC detection, HP 13C MRI is also useful for the assessment of the biological characteristics [58, 59]. Bliemsrieder et al. [60] presented an in vivo imaging technique for metabolic tumor phenotyping in rat models of HCC, and higher pyruvate-to-lactate conversion rates and lactate signal in subcutaneous tumors were derived from high lactate-to-alanine tumor cells and which may represent a higher biological aggressiveness. Thus, metabolic alterations detected by HP 13C MRI provide more insights into tumor heterogeneous characteristics and therefore may act as a biological and prognostic biomarker for HCC.

Pulmonary diseases

Currently, imaging techniques commonly used in clinical practice for the diagnosis of lung diseases include chest X-ray and computed tomography, which can obtain structural information about the lung, but they are both radioactive and cannot obtain imaging information about the gas exchange function of the lung, while usually the occurrence and development of diseases undergo a process from functional to structural lesions. Therefore, there is an urgent need to develop a non-radioactive imaging technique to detect the gas exchange function of the lung, which can be used for the early study and diagnosis of major lung diseases. Noninvasive, non-radioactive MRI has been used to obtain images of most living organs and tissues, such as the brain, heart, and abdomen, and is widely used in in vivo imaging studies and clinical disease diagnosis. However, the conventional proton (1H)-based MRI technique is not applicable to the lung because the lung tissue is mostly at the gas–solid interface, and thus, the magnetization rate varies greatly, resulting in a very short transverse relaxation time (T2) of protons, and even with ultrashort TE imaging sequences [61, 62], only partial information of the airways of the lung can be obtained. Moreover, the lung is composed of mostly cavernous structures, and the density of gas in the alveoli is about 1,000 times lower than that of ordinary tissues, and it is difficult to image the gas in the alveoli with conventional MRI techniques. To achieve MRI of lung gases, it is necessary to increase the polarization of the gas and thus enhance the sensitivity of the MRI signal.

The noble gases (3He or 129Xe) nuclear spins are hyperpolarized, using the spin-exchange optical pumping (SEOP) technique, by several orders of magnitude compared to the Boltzmann population [63]. This leads to amplification of the MRI sensitivity by orders of magnitude compared to that of thermal equilibrium, enabling sensitive diagnosis of lung diseases by measuring associated changes in the lung's structure and function. 3He and 129Xe are presently the most used gases for HP-gas lung MRI (Fig. 6), and in recent years, HP-83Kr and HP-19F have also been reported in the literature for lung gas MRI studies [64,65,66,67,68]. HP-gas MRI could obtain lung images reflecting lung morphology and ventilation status, enabling visual detection of lung diseases [69,70,71,72,73], and HP-gas MRI could differentiate the severity of asthma [73] (Fig. 7). In combination with proton-based lung contour imaging, HP-gas MRI of lung structures can quantitatively study pulmonary ventilation disorders. Lange et al. used lung contour and ventilation information from HP-3He MRI to calculate ventilation deficits in healthy volunteers and patients with different degrees of asthma, and their results were not only consistent with conventional respiratory volume measurements but also allowed differentiation of asthma severity [70]. It was also possible to detect ventilation deficits in smokers with normal lung function test [71]. Since then, researchers have developed a series of semi-automated and automated algorithms and have gradually moved toward HP-129Xe MRI of the lungs [72] and have demonstrated that the use of HP-129Xe MRI and semi-automated calculations can also differentiate between healthy volunteers and patients with chronic obstructive pulmonary disease patients (Fig. 8).

Fig. 6
figure 6

The workflow of hyperpolarized Xe 129 MRI in biomedical imaging. Preparation of the hyperpolarized Xe 129 (a), and the magnetic resonance imaging of the lung using Xe 129 (b)

Fig. 7
figure 7

Reproduced with permission from Elsevier Ltd (License Number: 5324281420540). J Allergy Clin Immunol. 2003 Jun;111(6):1205–11

Coronal MR images obtained immediately after inhalation of hyperpolarized Helium-3 gas in a healthy normal volunteer (a) and in patients with mild (FEV1 of 132% of predicted value. b Moderate (FEV1 of 83% of predicted value. c Severe (FEV1 of 34% of predicted value. d Asthma. The distribution of the gas is homogenous in the normal volunteer, and ventilation defects are seen with increasing numbers in the asthmatic patients with increasing severity (arrows pointing at several defects).

Fig. 8
figure 8

Reproduced with permission from John Wiley and Sons (License Number: 5324291003853). NMR Biomed. 2013 April; 26(4): 424–435

Representative 129Xe ventilation (left) and 1H SSFP anatomical image (right) slices of three COPD subjects. Note, a variety of ventilation defects score percentage (VDS%) are observed increasingly from 33.3% (a), 50.0% (b) to 66.7% (c) throughout the lungs. SSFP; Steady-state Free Precession.

Using ultra-fast imaging sequences, HP-gas lung imaging allows not only static imaging of the lung lobes but also dynamic imaging of the trachea and bronchi of the lung [72] and obtain images of the tracheal distribution with sub-millimeter resolution [74]. Albert et al. evaluated HP-gas tracheal imaging and demonstrated that the bronchial diameters calculated using two different HP-gas imaging modalities for human class 0–5 bronchi were consistent with the classical Weibel model [75] (Fig. 9). Driehuys et al. examined healthy and partially fibrotic rats using 129Xe imaging and found the significant restriction of ventilation in the lesioned portion [76].

Fig. 9
figure 9

Reproduced with permission from John Wiley and Sons (License Number: 532430065345). Reproduced with permission from John Wiley and Sons (License Number: 5324291003853). NMR Biomed. 2013 April; 26(4): 424–435

Representative dynamic coronal MR images using hyperpolarized 3He MRI. Thresholding a dynamic projection image. a Before thresholding; b After thresholding.

The longitudinal relaxation time T1 of HP-gas decreased sharply with increasing paramagnetic oxygen concentration in the gas [77]. After entering the lung, HP-gas mixed with oxygen remaining in the alveoli, and in areas of high oxygen concentration, the T1 of HP-gas became shorter and weaker and appeared as a low signal on ventilation imaging, so that HP-gas lung ventilation MRI can reflect the distribution of oxygen concentration. The changes in oxygen concentration in the alveoli were caused by gas exchange and ventilation in the alveoli, which can reflect lesions caused by the gas exchange in the lungs [78,79,80,81], such as chronic obstructive pulmonary disease and COVID-19 [82, 83].

The apparent diffusion coefficient (ADC) distribution of HP-gas in the lung can reflect the local structural information of the lung, which in turn can reflect the gas diffusion function in the alveoli and enable the detection of lung diseases. For example, in patients with emphysema, the alveoli became larger, and the restriction of gas movement was reduced, resulting in an increase in the ADC value of the gas in the emphysematous region [70, 84,85,86]. A three-channel system (1H, 3He, and 129Xe) could be used to simultaneously measure the diffusion coefficients of 3He and 129Xe, giving information on lung structure at different scales in a single sample [87]. In addition to the detection of lung structures, HP-gas ventilation imaging can also contribute to the diagnosis of lung cancer. Branca et al. achieved targeted detection of lung tumors by adding a superparamagnetic nanoparticle contrast agent that targets the lung tumor, affecting the homogeneity of the magnetic field in the targeted region through the superparamagnetic contrast agent, thereby altering T2* signal [88].

HP-gas MRI is non-radioactive and noninvasive and has great potential for the diagnosis and treatment of lung diseases, attracting great interest from many researchers. Although the spinning ratio of 129Xe is only one-third of that of 3He, resulting in a low MR sensitivity, the wide availability, and high natural abundance of 129Xe compared with the extremely scarce resources of 3He bring opportunities for the application of 129Xe MRI, especially the good lipid solubility and chemical shift sensitivity of 129Xe, which makes it unique for the detection of lung gas exchange function and has great potential for the early detection of lung diseases. It has great potential for the early detection of lung diseases. However, the HP-129Xe MRI signal sensitivity needs to be further improved to obtain higher resolution lung images and more accurate lung function information, which requires innovative methods, techniques, and instruments for generating HP-gas, optimization of HP-gas delivery systems to reduce the loss of polarization, design of new imaging coils to utilize polarization more effectively, and development of new imaging pulse sequences to obtain more functional lung information. The development of new imaging pulse sequences to obtain more functional information about the lung. Cross-innovation with clinical medicine and biochemistry is also needed for better development and application of this technology.

Heart disease and diabetes

Phosphorus-31 is primarily used to study cell membrane phospholipid metabolism concerning ATP energy metabolism. By analyzing the concentration of phosphorus-containing metabolites in the tissue, the pH of the tissue can be obtained as well as determining the concentration of magnesium ions in the tissue [89]. 31P MRS can be used to measure the myocardial high-energy phosphates phosphocreatine and ATP and to determine the ratio of phosphocreatine to ATP. Calculation of the ratio has been useful in identifying ischemia in animals [90, 91] and humans with coronary stenosis [89, 92,93,94,95].

Up to 40% of myocardial oxidative energy is derived from glucose and lactate in the blood, through which pyruvate is produced and enters the tricarboxylic acid cycle. HP-[1-13C]-pyruvate has been shown to assess cardiac PDH levels and fluxes in vitro and in vivo to provide important information on myocardial activity [96, 97]. Currently, HP MRS is mostly used to assess metabolic changes in diseases such as myocardial infarction, heart failure, cardiac hypertrophy, and cardiac tumors [98, 99] (Fig. 10).

Fig. 10
figure 10

Reproduced with permission from Elsevier Ltd (License Number: 5324651196851). Magn Reson Imaging. 2020 May;68:9–17

13C images of pyruvate (a), lactate (b), lactate/pyruvate (L/P) area under the curve (AUC) ratio (c), and bicarbonate (d) in the LV measured using dynamic HP [1-13C]pyruvate MRI. Metabolite images (color) are overlaid on T1-weighted axial images (grayscale) for anatomical reference.

Dodd et al. applied HP-[1-13C]-pyruvate and HP-[2-13C]-pyruvate to assess changes in mitochondrial metabolism after myocardial infarction in rats [97]. The level of acetyl coenzyme A produced via PDH at week 6 post-infarction was normal, but its oxidation was reduced (slowing of the tricarboxylic acid cycle). At week 22 post-infarction, PDH levels were found to correlate significantly with cardiac ejection fraction, and both acetyl coenzyme A production and its oxidation were reduced. This predicts that HP MRSI has important research value in the assessment of cardiac metabolism. The literature reported that HP-[1-13C]-pyruvate detected a significant decrease in [1-13C]-alanine and [1-13C]-lactate signals in the ischemia–reperfusion region of the left anterior descending coronary artery after 30 min of the blockade in a model of myocardial infarction, which was associated with cellular damage due to decreased cellular metabolism after severe myocardial hypoxia [100]. The HP-[1-13C]-pyruvate MRSI was used to determine the changes of pyruvate metabolism in the heart and liver of rats treated with metformin, and the ratio of [1-13C]-lactate to [1-13C]-pyruvate in the heart and liver increased from 0.10 to 0.27 (p = 0.02) and from 0.36 to 0.87 (p = 0.02), respectively, demonstrating that HP-[1-13C]-pyruvate MRSI can detect metformin-induced changes in cellular redox biology [101].

Diabetic nephropathy is one of the late complications of diabetes mellitus and one of the main causes of advanced renal failure. However, its pathogenesis remains not fully understood, and hypoxia in intrarenal tissues due to disturbances in oxygen metabolism is thought to be an important cause [102]. Laustsen C et al. used HP-[1-13C]-pyruvate MRI to observe alterations in renal metabolism in early-type I diabetes and found that the [1-13C]-lactate-to-[1-13C]-pyruvate ratio was significantly elevated in diabetic kidneys, whereas the [1-13C]-bicarbonate to [1-13C]-pyruvate ratio was unchanged, renal oxygen consumption was increased, and the effective oxygen consumption rate was reduced [103]. Under different oxygen concentration conditions, the metabolism of [1-13C]-pyruvate-to-lactate and alanine conversion in diabetic kidneys increased during acute hypoxia but did not alter bicarbonate flux, whereas lactate levels were normal at high oxygen concentrations [104]. This may explain why there is more nephropathy in diabetic patients living at high altitude [105]. Irregular insulin therapy increased the burden on the kidneys, which increased the uptake of [1-13C]-pyruvate and simultaneously increase the metabolic fluxes of anaerobic and aerobic metabolism (elevated lactate, alanine, and bicarbonate signals), leading to increased consumption of energy substances [106]. This suggests that strict glycemic control is important in insulin-dependent diabetes and that disturbances in renal metabolism in poor glycemic control are dependent on the consumption of energy substrates, which may herald new therapeutic targets for diabetic nephropathy.

In insulin-resistant type II diabetes, excessive hepatic gluconeogenesis and glycogenolysis can cause a rise in blood glucose, and HP-[1-13C]-pyruvate MRSI allowed noninvasive dynamic real-time measurement of hepatic metabolic alterations in vivo [52]. The livers of diabetic mice exhibit higher signals of 13C-oxaloacetate, 13C-aspartate, and 13C-malate, compared to normal mice injected with glucagon gluconeogenesis was increased similarly. Two weeks after metformin treatment, hepatic gluconeogenesis was reduced, 13C-aspartate and 13C-malate and the amount of 13C-labeled aspartate and malate decreased, the conversion of pyruvate to aspartate and malate also decreased, and blood glucose levels decreased, which is consistent with a downregulation of gluconeogenesis. The [1-13C]-pyruvate can be used as a biomarker to diagnose liver dysfunction in diabetes and to help assess treatment. The successful use of HP-13C metabolic imaging in diabetic disease models has a positive impact on the mechanism, treatment, and management of clinical diabetes itself and its complications.

Challenges and future opportunities

NMR spectroscopy is the only non-radiation tool that provides the opportunity to insight the molecular metabolically changes in vivo and provides detailed structural information for molecules. NMR is thus important in real-time imaging of metabolic activities not only for biochemists but also for the clinical scientists. The exploration of the detected metabolic signals with high sensitivity, efficiency, and stability is the external pursuit for the realization of practically, clinical usefully NMR spectroscopy imaging, and this always remains the challenges. To improve the sensitivity of the NMR spectroscopy and overcome the challenges, four major directions are pushed to increase the detection sensitivity of NMR spectroscopy.

Firstly, higher magnetic fields. The sensitivity of NMR spectroscopy increases with the higher magnetic fields, and the signal of the routine used 1H as well as the less sensitivity nucleus such as 2H, 13C, 15 N, 31P, 18F, and 23Na will be greatly boosted with the increased magnetic fields, which provides a potential direction despite many practical difficulties. At present, many higher magnetic fields strength have been used clinically or scientifically, the cutting edge of the clinically available higher magnetic fields is 7.0 T and 9.4 T in human [107, 108], and even the much more higher 11.7 T is potentially available [109]. However, the expensive cost, massive use of liquid helium, and huge site requirements limit its transition from dedicated laboratories to widespread clinical use, and innovative design and application of newly superconducting materials may breakthrough the dilemma.

Secondly, development of high sensitivity NMR coils. Another very important strategy is to increase the signal sensitivity by developing higher sensitivity NMR coils. With the development of new materials and innovative design, more elements can be synthesized into the coil with higher density. In addition, the coil can be also designed more closely fitted with the surface of human body that means shorter coil-to-sample distance and which can also help further improve the signal sensitivity. The air coil may represent this new direction [110]. Besides, commercially available dual-tuned (1H and 13C) multichannel coils that allow high signal-to-noise ratio imaging for all body areas are essential [2].

Thirdly, using hyperpolarization to improve the signal sensitivity. Dynamic nuclear polarization technique has been emerged as a powerful molecular strategy which allows safe, non-radiation injury, real-time, and metabolic-specific investigation of the dynamic metabolic process. Hyperpolarized MRI of 13C-labeled compounds have been shown to increase the signal sensitivity more than 10,000-fold [2]; this offers great potential to trace specific metabolic process in various lesions. However, its main limitation is that the current hyperpolarization technique can be only applicable to few molecular. Generation and co-hyperpolarization of multiple molecules provide the potential to make large number of molecules to be hyperpolarized and simultaneously insight the multiple metabolic pathways. Additionally, the other limitation of the HP techniques is that the produced HP substances or HP contrast agents generally cannot be re-hyperpolarized after their administration, and the hard-won HP state will exponentially decay back to the equilibrium. This decay ostensibly poses a fundamental limit on the time scale of biochemical processes that can be probed by HP contrast agents and requires HP lifetimes that are sufficiently log (> ten of seconds) for agent manipulation after its production, administration (e.g., inhalation or injection), in vivo delivery, and observation of metabolic and functional event [111]. Thus, innovative polarization approaches should be developed to overcome the limitations of shorter hyperpolarized signal lasts time and relatively sophisticated hyperpolarization technique in the produce and probe delivery.

Last but not least, ultra-fast MR imaging sequences and innovative parsing algorithm [112]. To develop ultra-fast MR imaging sequences is another promising approach to increase the signal sensitivity. As mentioned above, the hyperpolarized signal only lasts for few minutes; thus, maximizing signal acquisition in a limited time can further increase the signal sensitivity. In addition, using innovative parsing algorithm to separate the acquired signal and to further strength the target signal may also contribute to the signal improvement.

Conclusions

Multi-nuclear magnetic resonance spectroscopy is an emerging technique. With the development of magnetic resonance software and hardware, multi-nuclear magnetic resonance imaging has been applied to the basic and clinical transformation of various systems of human. Its unique advantages for displaying the real-time, dynamic metabolic process in different pathological processes provide the possibility for early diagnosis, evaluation of therapeutic efficacy, decision-making, drug development, and even exploration of new disease mechanisms. Despite these achievements, there is much progress yet to be made. Innovation of the polarization technique, exploration of new probes, quantification, and standardization of the results and construction of more prospective multicenter trials can further promote the clinical transformation of the advanced NMR technique.

Availability of data and materials

The data and material are included in this manuscript.

Abbreviations

18F-FDG:

18F-fluorodeoxyglucose

AS:

Alport syndrome

BBB:

Blood–brain barrier

HP:

Hyperpolarization

LDH:

Lactate dehydrogenase

MCT:

Monocarboxylate transporter

MP:

Mononuclear phagocyte

MRI:

Magnetic resonance imaging

MRSI:

Magnetic resonance spectroscopic imaging

MS:

Multiple sclerosis

NMR:

Nuclear magnetic resonance

PDH:

Pyruvate dehydrogenase

PET:

Positron emission tomography

RCC:

Renal cell carcinoma

TBI:

Traumatic brain injury

References

  1. Edelman RR (2014) The history of MR imaging as seen through the pages of radiology. Radiology 273:S181-200

    Article  PubMed  Google Scholar 

  2. Wang ZJ, Ohliger MA, Larson PEZ et al (2019) Hyperpolarized (13)C MRI: state of the art and future directions. Radiology 291:273–284

    Article  PubMed  Google Scholar 

  3. Stewart NJ, Matsumoto S (2021) Biomedical applications of the dynamic nuclear polarization and parahydrogen induced polarization techniques for hyperpolarized (13)C MR imaging. Magn Reson Med Sci 20:1–17

    Article  CAS  PubMed  Google Scholar 

  4. Adamson EB, Ludwig KD, Mummy DG, Fain SB (2017) Magnetic resonance imaging with hyperpolarized agents: methods and applications. Phys Med Biol 62:R81-r123

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Jacobs MA, Stearns V, Wolff AC et al (2010) Multiparametric magnetic resonance imaging, spectroscopy and multinuclear (23Na) imaging monitoring of preoperative chemotherapy for locally advanced breast cancer. Acad Radiol 17:1477–1485

    Article  PubMed  PubMed Central  Google Scholar 

  6. Le Page LM, Guglielmetti C, Taglang C, Chaumeil MM (2020) Imaging brain metabolism using hyperpolarized (13)C magnetic resonance spectroscopy. Trends Neurosci 43:343–354

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  7. Grist JT, Miller JJ, Zaccagna F et al (2020) Hyperpolarized (13)C MRI: a novel approach for probing cerebral metabolism in health and neurological disease. J Cereb Blood Flow Metab 40:1137–1147

    Article  PubMed  PubMed Central  Google Scholar 

  8. Hurd RE, Yen YF, Chen A, Ardenkjaer-Larsen JH (2012) Hyperpolarized 13C metabolic imaging using dissolution dynamic nuclear polarization. J Magn Reson Imaging 36:1314–1328

    Article  PubMed  Google Scholar 

  9. Kettenmann H, Hanisch UK, Noda M, Verkhratsky A (2011) Physiology of microglia. Physiol Rev 91:461–553

    Article  CAS  PubMed  Google Scholar 

  10. Tannahill GM, Iraci N, Gaude E, Frezza C, Pluchino S (2015) Metabolic reprograming of mononuclear phagocytes in progressive multiple sclerosis. Front Immunol 6:106

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  11. Kelly B, O’Neill LA (2015) Metabolic reprogramming in macrophages and dendritic cells in innate immunity. Cell Res 25:771–784

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  12. Guglielmetti C, Najac C, Didonna A, Van der Linden A, Ronen SM, Chaumeil MM (2017) Hyperpolarized (13)C MR metabolic imaging can detect neuroinflammation in vivo in a multiple sclerosis murine model. Proc Natl Acad Sci U S A 114:E6982-e6991

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Carpenter KL, Jalloh I, Gallagher CN et al (2014) (13)C-labelled microdialysis studies of cerebral metabolism in TBI patients. Eur J Pharm Sci 57:87–97

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Robertson CL, Saraswati M, Fiskum G (2007) Mitochondrial dysfunction early after traumatic brain injury in immature rats. J Neurochem 101:1248–1257

    Article  CAS  PubMed  Google Scholar 

  15. Guglielmetti C, Chou A, Krukowski K et al (2017) In vivo metabolic imaging of traumatic brain injury. Sci Rep 7:17525

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  16. DeVience SJ, Lu X, Proctor J et al (2017) Metabolic imaging of energy metabolism in traumatic brain injury using hyperpolarized [1-(13)C]pyruvate. Sci Rep 7:1907

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  17. Hackett EP, Pinho MC, Harrison CE et al (2020) Imaging acute metabolic changes in patients with mild traumatic brain injury using hyperpolarized [1-(13)C]pyruvate. iScience 23:101885

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Grist JT, McLean MA, Riemer F et al (2019) Quantifying normal human brain metabolism using hyperpolarized [1-(13)C]pyruvate and magnetic resonance imaging. Neuroimage 189:171–179

    Article  CAS  PubMed  Google Scholar 

  19. Xu Y, Ringgaard S, Mariager C et al (2017) Hyperpolarized (13)C magnetic resonance imaging can detect metabolic changes characteristic of penumbra in ischemic stroke. Tomography 3:67–73

    Article  PubMed  PubMed Central  Google Scholar 

  20. Zhou X, Sun Y, Mazzanti M et al (2011) MRI of stroke using hyperpolarized 129Xe. NMR Biomed 24:170–175

    Article  PubMed  CAS  Google Scholar 

  21. De Feyter HM, Behar KL, Corbin ZA et al (2018) Deuterium metabolic imaging (DMI) for MRI-based 3D mapping of metabolism in vivo. Sci Adv 4:eaat7314

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  22. Hesse F, Somai V, Kreis F, Bulat F, Wright AJ, Brindle KM (2021) Monitoring tumor cell death in murine tumor models using deuterium magnetic resonance spectroscopy and spectroscopic imaging. Proc Natl Acad Sci U S A. https://doi.org/10.1073/pnas.2014631118

    Article  PubMed  PubMed Central  Google Scholar 

  23. Kreis F, Wright AJ, Hesse F, Fala M, Hu DE, Brindle KM (2020) Measuring tumor glycolytic flux in vivo by using fast deuterium MRI. Radiology 294:289–296

    Article  PubMed  Google Scholar 

  24. Park I, Larson PEZ, Zierhut ML et al (2010) Hyperpolarized C-13 magnetic resonance metabolic imaging: application to brain tumors. Neuro Oncol 12:133–144

    Article  PubMed  PubMed Central  Google Scholar 

  25. Hurd RE, Yen YF, Mayer D et al (2010) Metabolic imaging in the anesthetized rat brain using hyperpolarized 1-C-13 pyruvate and 1-C-13 ethyl pyruvate. Magn Reson Med 63:1137–1143

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Park I, Larson PEZ, Gordon JW et al (2018) Development of methods and feasibility of using hyperpolarized carbon-13 imaging data for evaluating brain metabolism in patient studies. Magn Reson Med 80:864–873

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Miloushev VZ, Granlund KL, Boltyanskiy R et al (2018) Metabolic imaging of the human brain with hyperpolarized (13)C pyruvate demonstrates (13)C lactate production in brain tumor patients. Cancer Res 78:3755–3760

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Day SE, Kettunen MI, Cherukuri MK et al (2011) Detecting response of rat C6 glioma tumors to radiotherapy using hyperpolarized [1- 13C]pyruvate and 13C magnetic resonance spectroscopic imaging. Magn Reson Med 65:557–563

    Article  CAS  PubMed  Google Scholar 

  29. Ward CS, Venkatesh HS, Chaumeil MM et al (2010) Noninvasive detection of target modulation following phosphatidylinositol 3-kinase inhibition using hyperpolarized 13C magnetic resonance spectroscopy. Cancer Res 70:1296–1305

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Choi SY, Xue H, Wu R et al (2016) The MCT4 gene: a novel, potential target for therapy of advanced prostate cancer. Clin Cancer Res 22:2721–2733

    Article  CAS  PubMed  Google Scholar 

  31. Granlund KL, Tee SS, Vargas HA et al (2020) Hyperpolarized MRI of human prostate cancer reveals increased lactate with tumor grade driven by monocarboxylate transporter 1. Cell Metab 31:105-114.e103

    Article  CAS  PubMed  Google Scholar 

  32. Keshari KR, Sriram R, Koelsch BL et al (2013) Hyperpolarized 13C-pyruvate magnetic resonance reveals rapid lactate export in metastatic renal cell carcinomas. Cancer Res 73:529–538

    Article  CAS  PubMed  Google Scholar 

  33. Sriram R, Van Criekinge M, Hansen A et al (2015) Real-time measurement of hyperpolarized lactate production and efflux as a biomarker of tumor aggressiveness in an MR compatible 3D cell culture bioreactor. NMR Biomed 28:1141–1149

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Sriram R, Van Criekinge M, DeLos SJ et al (2016) Non-invasive differentiation of benign renal tumors from clear cell renal cell carcinomas using clinically translatable hyperpolarized (13)C pyruvate magnetic resonance. Tomography 2:35–42

    Article  PubMed  PubMed Central  Google Scholar 

  35. Sriram R, Gordon J, Baligand C et al (2018) Non-invasive assessment of lactate production and compartmentalization in renal cell carcinomas using hyperpolarized (13)C pyruvate MRI. Cancers (Basel) 10

  36. Nelson SJ, Kurhanewicz J, Vigneron DB et al (2013) Metabolic imaging of patients with prostate cancer using hyperpolarized [1–13C]pyruvate. Sci Transl Med 5:198ra108

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  37. Sushentsev N, McLean MA, Warren AY et al (2022) Hyperpolarised (13)C-MRI identifies the emergence of a glycolytic cell population within intermediate-risk human prostate cancer. Nat Commun 13:466

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Albers MJ, Bok R, Chen AP et al (2008) Hyperpolarized 13C lactate, pyruvate, and alanine: noninvasive biomarkers for prostate cancer detection and grading. Cancer Res 68:8607–8615

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Day SE, Kettunen MI, Gallagher FA et al (2007) Detecting tumor response to treatment using hyperpolarized 13C magnetic resonance imaging and spectroscopy. Nat Med 13:1382–1387

    Article  CAS  PubMed  Google Scholar 

  40. Dafni H, Larson PE, Hu S et al (2010) Hyperpolarized 13C spectroscopic imaging informs on hypoxia-inducible factor-1 and myc activity downstream of platelet-derived growth factor receptor. Cancer Res 70:7400–7410

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Zierhut ML, Yen YF, Chen AP et al (2010) Kinetic modeling of hyperpolarized 13C1-pyruvate metabolism in normal rats and TRAMP mice. J Magn Reson 202:85–92

    Article  CAS  PubMed  Google Scholar 

  42. Kanamaru H, Oyama N, Akino H, Okada K (2000) Evaluation of prostate cancer using FDG-PET. Hinyokika Kiyo 46:851–853

    CAS  PubMed  Google Scholar 

  43. Baligand C, Qin H, True-Yasaki A et al (2017) Hyperpolarized (13) C magnetic resonance evaluation of renal ischemia reperfusion injury in a murine model. NMR Biomed 30

  44. Moon CM, Oh CH, Ahn KY et al (2017) Metabolic biomarkers for non-alcoholic fatty liver disease induced by high-fat diet: In vivo magnetic resonance spectroscopy of hyperpolarized [1-(13)C] pyruvate. Biochem Biophys Res Commun 482:112–119

    Article  CAS  PubMed  Google Scholar 

  45. Moon CM, Shin SS, Lim NY et al (2018) Metabolic alterations in a rat model of hepatic ischaemia reperfusion injury: in vivo hyperpolarized (13) C MRS and metabolic imaging. Liver Int 38:1117–1127

    Article  CAS  PubMed  Google Scholar 

  46. Josan S, Billingsley K, Orduna J et al (2015) Assessing inflammatory liver injury in an acute CCl4 model using dynamic 3D metabolic imaging of hyperpolarized [1-(13)C]pyruvate. NMR Biomed 28:1671–1677

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Moreno KX, Satapati S, DeBerardinis RJ, Burgess SC, Malloy CR, Merritt ME (2014) Real-time detection of hepatic gluconeogenic and glycogenolytic states using hyperpolarized [2-13C]dihydroxyacetone. J Biol Chem 289:35859–35867

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Wilson DM, Di Gialleonardo V, Wang ZJ et al (2017) Hyperpolarized (13)C spectroscopic evaluation of oxidative stress in a rodent model of steatohepatitis. Sci Rep 7:46014

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Kim GW, Ahn KY, Kim YH, Jeong GW (2016) Time-course metabolic changes in high-fat diet-induced obesity rats: a pilot study using hyperpolarized (13)C dynamic MRS. Magn Reson Imaging 34:1199–1205

    Article  PubMed  CAS  Google Scholar 

  50. Can E, Bastiaansen JAM, Couturier DL, Gruetter R, Yoshihara HAI, Comment A (2022) [(13)C]bicarbonate labelled from hyperpolarized [1-(13)C]pyruvate is an in vivo marker of hepatic gluconeogenesis in fasted state. Commun Biol 5:10

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Chen J, Hackett EP, Kovacs Z, Malloy CR, Park JM (2021) Assessment of hepatic pyruvate carboxylase activity using hyperpolarized [1-(13) C]-l-lactate. Magn Reson Med 85:1175–1182

    Article  CAS  PubMed  Google Scholar 

  52. Lee P, Leong W, Tan T, Lim M, Han W, Radda GK (2013) In vivo hyperpolarized carbon-13 magnetic resonance spectroscopy reveals increased pyruvate carboxylase flux in an insulin-resistant mouse model. Hepatology 57:515–524

    Article  CAS  PubMed  Google Scholar 

  53. Wang JX, Merritt ME, Sherry D, Malloy CR (2016) A general chemical shift decomposition method for hyperpolarized (13) C metabolite magnetic resonance imaging. Magn Reson Chem 54:665–673

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Smith LM, Pitts CB, Friesen-Waldner LJ et al (2021) In vivo magnetic resonance spectroscopy of hyperpolarized [1-(13) C]pyruvate and proton density fat fraction in a guinea pig model of non-alcoholic fatty liver disease development after life-long western diet consumption. J Magn Reson Imaging 54:1404–1414

    Article  PubMed  Google Scholar 

  55. Yoshimitsu K (2014) Transarterial chemoembolization using iodized oil for unresectable hepatocellular carcinoma: perspective from multistep hepatocarcinogenesis. Hepat Med 6:89–94

    PubMed  PubMed Central  CAS  Google Scholar 

  56. Park YN, Kim MJ (2011) Hepatocarcinogenesis: imaging-pathologic correlation. Abdom Imaging 36:232–243

    Article  PubMed  Google Scholar 

  57. Hu S, Balakrishnan A, Bok RA et al (2011) 13C-pyruvate imaging reveals alterations in glycolysis that precede c-Myc-induced tumor formation and regression. Cell Metab 14:131–142

    Article  CAS  PubMed  Google Scholar 

  58. Gallagher FA, Kettunen MI, Day SE, Lerche M, Brindle KM (2008) 13C MR spectroscopy measurements of glutaminase activity in human hepatocellular carcinoma cells using hyperpolarized 13C-labeled glutamine. Magn Reson Med 60:253–257

    Article  CAS  PubMed  Google Scholar 

  59. von Morze C, Larson PE, Hu S et al (2011) Imaging of blood flow using hyperpolarized [(13)C]urea in preclinical cancer models. J Magn Reson Imaging 33:692–697

    Article  Google Scholar 

  60. Bliemsrieder E, Kaissis G, Grashei M et al (2021) Hyperpolarized (13)C pyruvate magnetic resonance spectroscopy for in vivo metabolic phenotyping of rat HCC. Sci Rep 11:1191

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Kveder M, Zupancic I, Lahajnar G et al (1988) Water proton NMR relaxation mechanisms in lung tissue. Magn Reson Med 7:432–441

    Article  CAS  PubMed  Google Scholar 

  62. Togao O, Tsuji R, Ohno Y, Dimitrov I, Takahashi M (2010) Ultrashort echo time (UTE) MRI of the lung: assessment of tissue density in the lung parenchyma. Magn Reson Med 64:1491–1498

    Article  PubMed  Google Scholar 

  63. Walker TG, Happer W (1997) Spin-exchange optical pumping of noble-gas nuclei. Rev Mod Phys 69:629–642. https://doi.org/10.1103/revmodphys.69.629

  64. Hughes-Riley T, Six JS, Lilburn DML et al (2013) Cryogenics free production of hyperpolarized 129Xe and 83Kr for biomedical MRI applications. J Magn Reson 237:23–33

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Pavlovskaya GE, Cleveland ZI, Stupic KF, Basaraba RJ, Meersmann T (2005) Hyperpolarized krypton-83 as a contrast agent for magnetic resonance imaging. Proc Natl Acad Sci U S A 102:18275–18279

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Six JS, Hughes-Riley T, Lilburn DM et al (2014) Pulmonary MRI contrast using surface quadrupolar relaxation (SQUARE) of hyperpolarized (83)Kr. Magn Reson Imaging 32:48–53

    Article  PubMed  PubMed Central  Google Scholar 

  67. Couch MJ, Fox MS, Viel C et al (2016) Fractional ventilation mapping using inert fluorinated gas MRI in rat models of inflammation and fibrosis. NMR Biomed 29:545–552

    Article  CAS  PubMed  Google Scholar 

  68. Gutberlet M, Kaireit TF, Voskrebenzev A et al (2018) Free-breathing dynamic (19)F gas MR imaging for mapping of regional lung ventilation in patients with COPD. Radiology 286:1040–1051

    Article  PubMed  Google Scholar 

  69. Mathew L, Evans A, Ouriadov A et al (2008) Hyperpolarized 3He magnetic resonance imaging of chronic obstructive pulmonary disease: reproducibility at 3.0 tesla. Acad Radiol 15:1298–1311

    Article  PubMed  Google Scholar 

  70. de Lange EE, Altes TA, Patrie JT et al (2006) Evaluation of asthma with hyperpolarized helium-3 MRI: correlation with clinical severity and spirometry. Chest 130:1055–1062

    Article  PubMed  Google Scholar 

  71. Woodhouse N, Wild JM, Paley MN et al (2005) Combined helium-3/proton magnetic resonance imaging measurement of ventilated lung volumes in smokers compared to never-smokers. J Magn Reson Imaging 21:365–369

    Article  PubMed  Google Scholar 

  72. Virgincar RS, Cleveland ZI, Kaushik SS et al (2013) Quantitative analysis of hyperpolarized 129Xe ventilation imaging in healthy volunteers and subjects with chronic obstructive pulmonary disease. NMR Biomed 26:424–435

    Article  CAS  PubMed  Google Scholar 

  73. Samee S, Altes T, Powers P et al (2003) Imaging the lungs in asthmatic patients by using hyperpolarized helium-3 magnetic resonance: assessment of response to methacholine and exercise challenge. J Allergy Clin Immunol 111:1205–1211

    Article  CAS  PubMed  Google Scholar 

  74. Johnson GA, Cofer GP, Hedlund LW, Maronpot RR, Suddarth SA (2001) Registered (1)H and (3)He magnetic resonance microscopy of the lung. Magn Reson Med 45:365–370

    Article  CAS  PubMed  Google Scholar 

  75. Lewis TA, Tzeng YS, McKinstry EL et al (2005) Quantification of airway diameters and 3D airway tree rendering from dynamic hyperpolarized 3He magnetic resonance imaging. Magn Reson Med 53:474–478

    Article  PubMed  PubMed Central  Google Scholar 

  76. Driehuys B, Pollaro J, Cofer GP (2008) In vivo MRI using real-time production of hyperpolarized 129Xe. Magn Reson Med 60:14–20

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Saam B, Happer W, Middleton H (1995) Nuclear relaxation of 3He in the presence of O2. Phys Rev A 52:862–865

    Article  CAS  PubMed  Google Scholar 

  78. Cieślar K, Stupar V, Canet-Soulas E, Gaillard S, Crémillieux Y (2007) Alveolar oxygen partial pressure and oxygen depletion rate mapping in rats using 3He ventilation imaging. Magn Reson Med 57:423–430

    Article  PubMed  Google Scholar 

  79. Cieślar K, Alsaid H, Stupar V et al (2007) Measurement of nonlinear pO2 decay in mouse lungs using 3He-MRI. NMR Biomed 20:383–391

    Article  PubMed  Google Scholar 

  80. Kadlecek S, Mongkolwisetwara P, Xin Y et al (2011) Regional determination of oxygen uptake in rodent lungs using hyperpolarized gas and an analytical treatment of intrapulmonary gas redistribution. NMR Biomed 24:1253–1263

    Article  CAS  PubMed  Google Scholar 

  81. Wild JM, Fichele S, Woodhouse N, Paley MN, Kasuboski L, van Beek EJ (2005) 3D volume-localized pO2 measurement in the human lung with 3He MRI. Magn Reson Med 53:1055–1064

    Article  PubMed  Google Scholar 

  82. Marshall H, Parra-Robles J, Deppe MH, Lipson DA, Lawson R, Wild JM (2014) (3)He pO2 mapping is limited by delayed-ventilation and diffusion in chronic obstructive pulmonary disease. Magn Reson Med 71:1172–1178

    Article  CAS  PubMed  Google Scholar 

  83. Li H, Zhao X, Wang Y et al (2021) Damaged lung gas exchange function of discharged COVID-19 patients detected by hyperpolarized (129)Xe MRI. Sci Adv. https://doi.org/10.1126/sciadv.abc8180

    Article  PubMed  PubMed Central  Google Scholar 

  84. Kaushik SS, Cleveland ZI, Cofer GP et al (2011) Diffusion-weighted hyperpolarized 129Xe MRI in healthy volunteers and subjects with chronic obstructive pulmonary disease. Magn Reson Med 65:1154–1165

    Article  PubMed  Google Scholar 

  85. Saam BT, Yablonskiy DA, Kodibagkar VD et al (2000) MR imaging of diffusion of (3)He gas in healthy and diseased lungs. Magn Reson Med 44:174–179

    Article  CAS  PubMed  Google Scholar 

  86. Salerno M, Altes TA, Mugler JP 3rd, Nakatsu M, Hatabu H, de Lange EE (2001) Hyperpolarized noble gas MR imaging of the lung: potential clinical applications. Eur J Radiol 40:33–44

    Article  CAS  PubMed  Google Scholar 

  87. Wild JM, Marshall H, Xu X et al (2013) Simultaneous imaging of lung structure and function with triple-nuclear hybrid MR imaging. Radiology 267:251–255

    Article  PubMed  Google Scholar 

  88. Driehuys B, Cofer GP, Pollaro J, Mackel JB, Hedlund LW, Johnson GA (2006) Imaging alveolar-capillary gas transfer using hyperpolarized 129Xe MRI. Proc Natl Acad Sci U S A 103:18278–18283

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  89. Buchthal SD, den Hollander JA, Merz CN et al (2000) Abnormal myocardial phosphorus-31 nuclear magnetic resonance spectroscopy in women with chest pain but normal coronary angiograms. N Engl J Med 342:829–835

    Article  CAS  PubMed  Google Scholar 

  90. Schaefer S, Camacho SA, Gober J et al (1989) Response of myocardial metabolites to graded regional ischemia: 31P NMR spectroscopy of porcine myocardium in vivo. Circ Res 64:968–976

    Article  CAS  PubMed  Google Scholar 

  91. Schaefer S, Schwartz GG, Gober JR et al (1990) Relationship between myocardial metabolites and contractile abnormalities during graded regional ischemia. Phosphorus-31 nuclear magnetic resonance studies of porcine myocardium in vivo. J Clin Invest 85:706–713

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Nunnally RL, Bottomley PA (1981) Assessment of pharmacological treatment of myocardial infarction by phosphorus-31 NMR with surface coils. Science 211:177–180

    Article  CAS  PubMed  Google Scholar 

  93. Flaherty JT, Weisfeldt ML, Bulkley BH, Gardner TJ, Gott VL, Jacobus WE (1982) Mechanisms of ischemic myocardial cell damage assessed by phosphorus-31 nuclear magnetic resonance. Circulation 65:561–570

    Article  CAS  PubMed  Google Scholar 

  94. Bottomley PA, Herfkens RJ, Smith LS, Bashore TM (1987) Altered phosphate metabolism in myocardial infarction: P-31 MR spectroscopy. Radiology 165:703–707

    Article  CAS  PubMed  Google Scholar 

  95. Neubauer S, Krahe T, Schindler R et al (1992) 31P magnetic resonance spectroscopy in dilated cardiomyopathy and coronary artery disease. altered cardiac high-energy phosphate metabolism in heart failure. Circulation 86:1810–1818

    Article  CAS  PubMed  Google Scholar 

  96. Malloy CR, Merritt ME, Sherry AD (2011) Could 13C MRI assist clinical decision-making for patients with heart disease? NMR Biomed 24:973–979

    Article  PubMed  PubMed Central  Google Scholar 

  97. Dodd MS, Atherton HJ, Carr CA et al (2014) Impaired in vivo mitochondrial Krebs cycle activity after myocardial infarction assessed using hyperpolarized magnetic resonance spectroscopy. Circ Cardiovasc Imaging 7:895–904

    Article  PubMed  PubMed Central  Google Scholar 

  98. Rider OJ, Tyler DJ (2013) Clinical implications of cardiac hyperpolarized magnetic resonance imaging. J Cardiovasc Magn Reson 15:93

    Article  PubMed  PubMed Central  Google Scholar 

  99. Barton GP, Macdonald EB, Goss KN, Eldridge MW, Fain SB (2020) Measuring the link between cardiac mechanical function and metabolism during hyperpolarized (13)C-pyruvate magnetic resonance experiments. Magn Reson Imaging 68:9–17

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Merritt ME, Harrison C, Storey C, Jeffrey FM, Sherry AD, Malloy CR (2007) Hyperpolarized 13C allows a direct measure of flux through a single enzyme-catalyzed step by NMR. Proc Natl Acad Sci U S A 104:19773–19777

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Lewis AJ, Miller JJ, McCallum C et al (2016) Assessment of metformin-induced changes in cardiac and hepatic redox state using hyperpolarized[1-13C]pyruvate. Diabetes 65:3544–3551

    Article  CAS  PubMed  Google Scholar 

  102. Hansell P, Welch WJ, Blantz RC, Palm F (2013) Determinants of kidney oxygen consumption and their relationship to tissue oxygen tension in diabetes and hypertension. Clin Exp Pharmacol Physiol 40:123–137

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Laustsen C, Ostergaard JA, Lauritzen MH et al (2013) Assessment of early diabetic renal changes with hyperpolarized 1–13C pyruvate. Diabetes Metab Res Rev 29:125–129

    Article  CAS  PubMed  Google Scholar 

  104. Laustsen C, Lycke S, Palm F et al (2014) High altitude may alter oxygen availability and renal metabolism in diabetics as measured by hyperpolarized 1-C-13 pyruvate magnetic resonance imaging. Kidney Int 86:67–74

    Article  CAS  PubMed  Google Scholar 

  105. Hochman ME, Watt JP, Reid R, O’Brien KL (2007) The prevalence and incidence of end-stage renal disease in native American adults on the Navajo reservation. Kidney Int 71:931–937

    Article  CAS  PubMed  Google Scholar 

  106. Laustsen C, Lipso K, Ostergaard JA et al (2014) Insufficient insulin administration to diabetic rats increases substrate utilization and maintains lactate production in the kidney. Physiol Rep. https://doi.org/10.14814/phy2.12233

    Article  PubMed  PubMed Central  Google Scholar 

  107. Paech D, Nagel AM, Schultheiss MN et al (2020) Quantitative dynamic oxygen 17 MRI at 7.0 T for the cerebral oxygen metabolism in glioma. Radiology 295:181–189

    Article  PubMed  Google Scholar 

  108. Murali-Manohar S, Borbath T, Wright AM, Soher B, Mekle R, Henning A (2020) T(2) relaxation times of macromolecules and metabolites in the human brain at 9.4 T. Magn Reson Med 84:542–558

    Article  CAS  PubMed  Google Scholar 

  109. Tanoue M, Saito S, Takahashi Y et al (2019) Amide proton transfer imaging of glioblastoma, neuroblastoma, and breast cancer cells on a 11.7 T magnetic resonance imaging system. Magn Reson Imaging 62:181–190

    Article  CAS  PubMed  Google Scholar 

  110. Cogswell PM, Trzasko JD, Gray EM et al (2021) Application of adaptive image receive coil technology for whole-brain imaging. AJR Am J Roentgenol 216:552–559

    Article  PubMed  Google Scholar 

  111. Nikolaou P, Goodson BM, Chekmenev EY (2015) NMR hyperpolarization techniques for biomedicine. Chemistry 21:3156–3166

    Article  CAS  PubMed  Google Scholar 

  112. Feinberg DA, Setsompop K (2013) Ultra-fast MRI of the human brain with simultaneous multi-slice imaging. J Magn Reson 229:90–100

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Funding

This work was supported by Science and Technology Support Program of Sichuan Province (Grant number 2021YFS0144, Grant number 2021YFS0021), China Postdoctoral Science Foundation (2021M692289), Post-Doctor Research Project, West China Hospital, Sichuan University (Grant number 2020HXBH130).

Author information

Authors and Affiliations

Authors

Contributions

YW, QL, FC, and C-WY collected materials and the literature. F-F G, TZ, SW, and C-WY organized the literature and extracted data. C-WY, YW, and J-Z W wrote the manuscript. YW, H-Y J, and BS revised the manuscript. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Bin Song.

Ethics declarations

Ethics approval and consent to participate

The authors are accountable for all aspects of the work in ensuring that questions related to the accuracy or integrity of any part of the work are appropriately investigated and resolved.

Consent for publication

All authors and all patients involved are agreed to the publication of this manuscript.

Competing interests

The authors have no conflicts of interest to declare.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, 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 changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's 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/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wei, Y., Yang, C., Jiang, H. et al. Multi-nuclear magnetic resonance spectroscopy: state of the art and future directions. Insights Imaging 13, 135 (2022). https://doi.org/10.1186/s13244-022-01262-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13244-022-01262-z

Keywords