Next Article in Journal / Special Issue
Immune and Inflammatory Networks in Myocardial Infarction: Current Research and Its Potential Implications for the Clinic
Previous Article in Journal
Genome-Wide Identification of Cassava Glyoxalase I Genes and the Potential Function of MeGLYⅠ-13 in Iron Toxicity Tolerance
Previous Article in Special Issue
Phosphatidylserine Supplementation as a Novel Strategy for Reducing Myocardial Infarct Size and Preventing Adverse Left Ventricular Remodeling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Challenges in Metabolomics-Based Tests, Biomarkers Revealed by Metabolomic Analysis, and the Promise of the Application of Metabolomics in Precision Medicine

by
Alessandro Di Minno
1,2,*,
Monica Gelzo
2,3,
Marianna Caterino
2,3,
Michele Costanzo
2,3,
Margherita Ruoppolo
2,3 and
Giuseppe Castaldo
2,3
1
Dipartimento di Farmacia, University of Naples Federico II, 80131 Naples, Italy
2
CEINGE-Biotecnologie Avanzate, 80131 Naples, Italy
3
Department of Molecular Medicine and Medical Biotechnology, School of Medicine, University of Naples Federico II, 80131 Naples, Italy
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(9), 5213; https://doi.org/10.3390/ijms23095213
Submission received: 1 April 2022 / Revised: 29 April 2022 / Accepted: 5 May 2022 / Published: 6 May 2022

Abstract

:
Metabolomics helps identify metabolites to characterize/refine perturbations of biological pathways in living organisms. Pre-analytical, analytical, and post-analytical limitations that have hampered a wide implementation of metabolomics have been addressed. Several potential biomarkers originating from current targeted metabolomics-based approaches have been discovered. Precision medicine argues for algorithms to classify individuals based on susceptibility to disease, and/or by response to specific treatments. It also argues for a prevention-based health system. Because of its ability to explore gene–environment interactions, metabolomics is expected to be critical to personalize diagnosis and treatment. Stringent guidelines have been applied from the very beginning to design studies to acquire the information currently employed in precision medicine and precision prevention approaches. Large, prospective, expensive and time-consuming studies are now mandatory to validate old, and discover new, metabolomics-based biomarkers with high chances of translation into precision medicine. Metabolites from studies on saliva, sweat, breath, semen, feces, amniotic, cerebrospinal, and broncho-alveolar fluid are predicted to be needed to refine information from plasma and serum metabolome. In addition, a multi-omics data analysis system is predicted to be needed for omics-based precision medicine approaches. Omics-based approaches for the progress of precision medicine and prevention are expected to raise ethical issues.

1. Introduction

The term “metabolomics” was first used at the beginning of this millennium to identify the area of functional genomics devoted to the analysis of metabolites [1,2]. Metabolomics defines the comprehensive characterization of small molecules derived from both the genome (i.e., endogenous metabolites) and their interaction with the environment (i.e., exogenous metabolites) [3]. In recent years, methods have advanced for metabolomics and have allowed the for reliable identification, detection, and quantification of new metabolites in food, plant, environmental, animal, and human research. The combined use of untargeted and targeted metabolomics has exhibited many advantages beyond analytical chemistry [4]. In addition to documenting the high hypothesis-generating potential [1], advancements in omics have provided significant information regarding new potential biomarkers [5]. Advanced data processing systems (e.g., informatics) have greatly helped to characterize metabolic pathways in different biological systems [6,7,8,9]. However, the possibility has also emerged that inherent technical limitations in analytical instrumentation and in methods of analyses might have slowed the progress and industrial applications of metabolomics [10,11]. How these shortcomings have been addressed is summarized in Section 2 of this review. Presently, easy and predictable quantification of metabolites is achieved in plasma or serum [6,12]. However, much work needs to be conducted to interpret and explore the overwhelming amount of data to date generated by metabolomics [13]. How to implement the relevance of metabolomics-based tests in biomedical research is discussed in Section 3 of this review. By integrating biomarkers with genetic and phenotypic characteristics that distinguish one patient from another with comparable clinical settings, precision medicine is aimed at systemically evaluating the underlying causes of disease so as to target health interventions to individual needs [14]. Translational opportunities of metabolomics are critical for the progress of precision medicine [13,14,15,16]. The extent to which the criteria applied to gather the information currently employed in precision medicine may help the advancement of metabolomics is discussed in Section 4 of this report.

2. Current Challenges in Targeted Metabolomics

Table 1 reports metabolomics-based biomarkers identified over the last decade in pre-natal and post-natal diagnosis, and in related experimental models by the authors of the present review. During the same time period, a variety of metabolomics-based biomarkers for characterizing environmental contaminants [17] or food derivatives [2], in addition to identifying the risk of diabetes mellitus [18,19], coronary heart disease [20,21,22,23] or cancer [24,25,26,27,28], have been identified. Advantages and disadvantages of different instrumental platforms, whose use is related to the chemical complexity of the biological system analyzed [10], have emerged in all these areas of metabolomics investigation. The fact [29,30] that very sensitive detectors (e.g., MS) that directly reveal very low concentrations of metabolites are not sensitive enough to simultaneously measure high-concentration components arose as a critical disadvantage. The need for different platforms and of different experts (analytical chemists, biologists, statisticians, data scientists and bioinformaticians) to achieve a comprehensive metabolome coverage [2,3] has also been recognized. Indeed, while reliably handling laboratory medicine issues, researchers trained in liquid chromatography–mass spectrometry (LC–MS) often need the help of experts in bioinformatics for the optimal experimental design for individual metabolomics studies and the appropriate statistics to be employed. This argues for large metabolomics groups with expertise and instrumentation sufficient to avoid contract laboratories (to carry out ad hoc experiments). A multifaceted research asset also enables to: (1) set up collaboration platforms with skilled metabolomics groups to increase chances to achieve funding for large program projects and overcome the high costs of analytical instrumentation, and (2) develop specialized training programs to teach beginners the broad spectrum of expertise needed for reliable analyses. How major additional pre-analytical, analytical, and post-analytical hurdles have been (and are being) addressed in metabolomics studies is summarized in the next few paragraphs.
  • Standard operating procedures. The rationale for the wide spectrum of methods used in different metabolomics labs [11,31,32] stems from the following: (1) no single analytical method is sensitive and specific enough to allow for the identification and quantitation of the whole metabolome of even a single biological entity [31]; (2) the metabolome of a cell/organism contains metabolites differing in their concentrations (from g/L to pg/L) [31,33], turnover rates, and stability; and (3) the biology of different living organisms implies diversity in the metabolites to be identified/measured [1,33]. In view of this, the Metabolomics Society has set up the “metabolomics Standards Initiative (MSI)” Committee that has established rules to standardize metabolomics systems [2,34,35]. Quality control and standard operating procedures should be carefully followed to reduce pre-analytical errors [36]. Standardization of steps throughout the study procedure and data analysis (e.g., the analytical platform to be employed, instrumentation performance, the type of analysis to use, and the requirements for the interpretation of the output) prevent the risk of poor-quality control metabolomics protocols, incorrect quantification of metabolites, and deceptive data interpretation [35,37,38]. Few variables should be selected to make metabolomics data reliable [34].
  • Quantification. Most of the data generated by metabolomics rely on normalization of the signal [3,30]. However, semi-quantitative approaches hamper multi-omics integration and translation of metabolomics data into clinical practice [11,39]. Definition of normal concentrations of a metabolite is key for early detection of pre-clinical conditions [40,41,42]. Consistent with the possibility that metabolomics can achieve absolute quantification of the metabolome [43], methods and analytical platforms for absolute quantification of the metabolome using targeted approaches are presently available [3,10]. Examples (Table 2) of technologies, platforms, and protocols for absolute quantification of several metabolites are now available [44,45,46,47,48,49,50,51]. Presently, the possibility is also documented that the relative, or absolute accuracy of quantification of newly discovered metabolites needs newer standardization steps [38,49,52].
  • Choice of separation methods. Reverse phase approaches should be used for the separation of non-polar components (e.g., fatty acids), while normal phase approaches should be preferred to separate polar compounds (e.g., nucleotides and sugars) [11]. Thus, the compounds to be measured and the biochemical pathways to be identified define the separation method to be employed. Robust techniques (e.g., nuclear magnetic resonance, NMR) exhibit rather limited sensitivity of detections [15,53,54]. Advances in analytical instrumentation are overcoming such limitations [55,56]. The use of small-in-size NMR machineries and mass spectrometers provide wide coverage of metabolites [57]. Platforms with high reproducibility and detection consistency are being developed to reveal low concentrations of metabolites [57,58,59]. Two-dimensional chromatographic separations are becoming increasingly widespread [56], and MS based technologies are gradually being employed in a targeted fashion [51,60].
  • Combination of different techniques. “Hyphenation” may be a new frontline in metabolomics [61]. Using standards or library spectra, spectroscopy produces selective information for identification of mixtures of chemical components separated by chromatography. Thus “hyphenation” combines the advantages of both techniques. Combinations of different techniques helps overcome limitations of single techniques and calls for major achievements in metabolomic studies [62]. Hyphenation of liquid chromatography–nuclear magnetic resonance–mass spectrometry liquid chromatography (LC–NMR–MS LC) has been developed for global metabolite profiling and identification of compounds [2]. The setup of such a platform needs to be simplified.
  • Statistical analysis. A robust statistical analysis of the results (e.g., t-tests, ANOVA, principal component analysis [PCA], hierarchical cluster analysis (HCA), partial least square–discriminant analysis [PLSDA], volcano plots, correlation analysis) is critical for the reliability of metabolomics studies. For inherent reasons, the statistical significance for analytes that differ between cohorts is difficult to be determined in untargeted metabolomics. While enhancing the number of false negatives (type II errors), conservative approaches such as the Bonferroni correction limit false positive data (type I error) [4]. False discovery rate (FDR) approaches help address the issue of removing false positives, especially in untargeted metabolomics [63]. For instance, the Q-value calculates the maximally applicable correction to a given dataset [64].
  • Metabolite identification. Both in targeted and untargeted metabolomics, the identification of “true” metabolites pushes upcoming steps of the analysis [65,66,67,68] and informative interpretation of features beyond the standard putative identification based on mass and/or retention time [69]. “True” metabolite identification is also critical for pathway analysis and mapping. The development of the human http://www.hmdb.ca/, accessed on 31 March 2022), food (http://foodb.ca/, accessed on 31 March 2022), DrugBank (https://www.drugbank.ca/, accessed on 31 March 2022) and T3DB metabolome database (http://www.t3db.ca/, accessed on 31 March 2022) helps achieve this goal. Especially for GC–MS methods [31,70] together with commercial metabolite libraries, in-house comprehensive spectral libraries of metabolites help convert putative metabolites/features into positive identifications [68]. Spectral libraries (e.g., Metlin or mzCloud) provide a reliable standard for the identification of the majority of naturally occurring metabolites present in biological materials [71], including those for which kits are not available [72]. Newer bioinformatics tools that employ web-accessed libraries are anticipated to improve automated metabolite identification [7,72].

3. Overpromising but Under-Delivering Translational Results

The high hypothesis-generating potential (and translational skills) of metabolomics is now established [41], and panels of biomarkers have been defined (http://www.mayomedicallaboratories.com, accessed on 31 March 2022). Using advanced analytical, community-based methods [72] and bioinformatics [1,5], (targeted) weaknesses in metabolomics have largely been overcome. Inherent technical limitations that might have delayed clinical and industrial translations of interfaces generated by this strategy have also been minimized. Accordingly, the whole human serum metabolome has been mapped in a UK population [73]. Additional issues are likely to be addressed through community-based approaches [16], and this may expand metabolomics-based opportunities to primary care facilities that have little access to expensive instruments [2], All this progress might be at odds [13] with the perception that metabolomics is overpromising but under-delivering translational results [15]. However: (1) other common biological matrices should be regularly explored, and (2) metabolomics information collected in clinical investigations should make a positive impact on the public [2]. In the present section, examples of how metabolomics-based research (as an emerging discipline) is currently being exploited to expand its role in health and disease are provided, and details on new potential directions to be pursued to improve our understanding of human pathophysiology are summarized.
  • Biomarkers discovery and validation. Biomarkers are defined as objectively measured indicators of normal biological processes, pathogenic processes, or pharmacologic responses to a therapeutic intervention [5]. At variance with other biomarkers [13], metabolites are easily quantified at a low cost [41]. Most currently identified metabolomics-based biomarkers arise from studies that are rather limited in experimental designs [4], statistical robustness and validity [37]. Indeed, to date, biomarker discovery and validation has been often carried out in small uncontrolled trials [74]. Independent validation within the same topic, an attitude that increases confidence in the clinical strength of a potentially metabolomics-based test, is erratic. Because of the lack of a second evaluation in other cohorts, the possibility that any findings these studies have generated might be poorly reproducible should be considered. In keeping with this, a very limited number of the metabolomics-based biomarkers that have been reported to date, are widely employed in clinical practice [75]. Ad hoc, prospective trials are mandatory to validate biomarkers with high chances to impact clinical practice [76,77]. In this respect, numbers of patients to be tested may be limited in studies devoted to rare diseases [78], while they should be large enough in very common clinical settings (e.g., hypercholesterolemia). In the latter case, the possibility of identifying intermediate phenotypes (e.g., subjects with/without high lipoprotein levels in the circulation) should be considered.
  • Newer sources of metabolomic analysis. The roadway of pathophysiology is key to understanding the machinery of diseases and to recognize the cause and the downstream effect of a disease. Investigations of metabolomics-based biomarkers should be carried out accordingly. Advancing towards this direction implies pathophysiology-oriented targeted metabolomics studies, which would be better if conducted in cooperation with other omics communities [79,80]. New sources of metabolomic analyses may be critical in this respect. In addition to plasma [81] and serum metabolome [82,83,84], discoveries in urinary metabolome [85,86,87,88] and in the volatilome, (e.g., breath) [89,90,91] will likely help gather/refine information on the mechanisms of major causes of death. Information-rich metabolomes may be also obtained from cerebrospinal fluid, human saliva, broncho alveolar lavage, sweat, feces, semen, and amniotic fluid. Studies struggle to extend measurements to intact tissues [4]. In clinical pharmacology, models of mammalian cultured primary cells are relevant for adsorption, distribution, metabolism, and excretion–toxicology (ADME–Tox) studies. With appropriate protections, the risk–benefit ratios of these studies may be defined for individual cases/diseases, and biomarkers identified [14,92].
  • Newer directions to be pursued. Together with top causes of death in developed countries (ischemic heart and cerebrovascular disease, and malignancy) [4], the rapid rise of pathogens is acknowledged to increasingly contribute to world-wide mortality [93]. Newer pathogens are emerging [94]. ‘Traffic’ of microbes and the diseases they cause is facilitated in the globalized world of the third millennium. The adaptation into a new human host population may produce ‘new’ mutations in viruses, bacteria, or fungi that allow them to acquire new biological characteristics to adapt to new ecologies and to infect new hosts [95,96,97,98,99,100,101]. Pathogens may also be transmitted by human blood and blood-derived products. Donor selection and blood screening, and methods for their purification/inactivation have reduced the risk of pathogen contamination of blood/plasma-derived products and increased the safety of blood products [102]. However, the poor sensitivity/specificity of current screening methods, and the lack of reliable tests for some pathogens (e.g., prions), should be emphasized [94]. Metabolomics may minimize/eradicate the risk of contaminants in blood, including pathogens. Because of the rise in antimicrobial resistance, many normally harmless opportunistic microorganisms are increasing their pathogenicity, and bacterial infections are predicted to kill more humans than cancer and heart disease in the coming decades [103]. Metabolomics should work to establish ad hoc biomarkers to identify the appropriate strategy and prevent future deaths in the area of antimicrobial resistance.

4. Metabolomics in the Era of Precision Medicine

  • The promise of precision medicine. Current clinical practice focuses on few variables and provides little information on their potential interactions. The identification of variables to classify individuals into sub-populations is critical for precision medicine and precision prevention [104]. Newborn screening for genetic mutations, e.g., phenylketonuria, and progress in dietary intervention to prevent the onset of diseases, are some of the earliest examples of precision medicine and precision prevention. Functional genomics has been the determining factor of an early tailoring approach once key profiles are identified [105]. More recently, information from genomics has been critical for the progress of cancer diagnostics, therapeutics, and prevention [106], and this way of thinking has been extended to the majority of areas in clinical medicine [14]. Cheap genome sequencing [107], powerful methods of functional genomics, large-scale biological databases, and computational tools for analyzing large sets of data have greatly fostered this attitude [108]. However, the genomic-approach based initiatives that have been launched in precision medicine to date, have delivered fewer disease genes than originally expected [3]. Limited information also arose from the approaches based on transcriptomics and everything relating to RNAs [109]. In keeping with this [110], evidence has emerged that: (1) many tumors are not genetically and metabolically homogeneous; (2) metabolic heterogeneity exists also within an individual tumor tissue [3], and (3) obesity-induced changes in adipose tissue microenvironment impact genetics of cardiovascular disease [111]. These limitations have shifted the attention from genomics-centered approaches to the impact of environment determinants in the initiation and progress of malignancy, and vascular disease [112]. At variance with genomics, which only foresees events based on genetic predisposition, metabolomics also reveals events related to gene–environment interactions [3]. As such, metabolomics is a key driver to exploring underlying pathogenic mechanisms of complex polygenic diseases (e.g., cancer, cardiovascular diseases, and diabetes mellitus) for which environmental factors (e.g., diet) substantially impact disease onset and development [113]. Recent improvements in single cell metabolomic analysis [114] lend credence to the possibility of specific treatments for individual metabolic microenvironments within diseased tissues [115].
  • How metabolomic information can potentially benefit the development of precision medicine: an example. Cystathionine β-synthase deficiency (CBSD, EC 4.2.1.22), also known as homocystinuria (OMIM 236200, mean prevalence worldwide 1:335,000, ranges 1:1800–1:900,000), is a recessively inherited disorder of the catabolic pathway (the transsulfuration pathway) for the essential amino acid methionine (Met) [116]. Met is converted to the non-structural amino acid homocysteine (Hcy), via S-adenosylmethionine (SAM) and S-adenosylhomocysteine (SAH), by the release of a methyl group that is used in methylation reactions (e.g., via phosphatidylethanolamine N-methyltransferase, PEMT). CBSD impairs the conversion of homocysteine (Hcy) to cystathionine, leading to Hcy accumulation in plasma (up to 200 µm/L) and urine (homocystinuria) [117,118,119,120]. Severely affected patients with CBSD present ectopia lentis, learning difficulties, connective tissue disturbances including skeletal abnormalities (marfanoid habitus), osteoporosis, propensity to venous and arterial thrombosis, premature atherosclerosis and occasional liver steatosis. Presently, genotype–phenotype correlation in homocystinuria remains obscure [120]. Using an ultra-high-performance liquid chromatography–electrospray ionization–quadrupole time-of-flight–mass spectrometry method, and employing an untargeted lipidomic approach, we have identified a novel biochemical abnormality in plasma from 11 severe CBSD patients (belonging to nine unrelated families and carrying different genetic defects already reported in patients with CBS), consisting of a depletion of phosphatidylcholine (PC; p = 0.02) and lysophosphatidylcholine (LPC; p = 0.003) species containing docosahexaenoic acid (DHA), and a higher than normal medium and long-chain polyunsaturated fatty acids content in phosphatidylethanolamine (PE) and lysophosphatidylethanolamine (LPE) species (p < 0.02). This suggests impaired in vivo PEMT activity. As PEMT needs methyl groups to convert PE into PC, SAM and SAH were measured by LC–MS. Whole blood SAM and SAH concentrations were 1.4-fold (p = 0.015) and 5.3-fold (p = 0.003) higher in CBSD patients than in controls. A positive correlation between SAM/SAH and PC/PE ratios (r = 0.520; p = 0.019) was found. CBSD patients with liver steatosis (5/11) had a significantly lower PC/PE ratio than those without (48.26 ± 18.7 vs. 86.28 ± 14.4, respectively; p = 0.016). After correcting for age and gender, liver steatosis was associated with PE/PC ratio in a multivariate linear regression analysis (β = −0.770; p = 0.009) [78]. Pathophysiological information is that a diminished PEMT expression/activity as reflected by a decrease in hepatic PC/PE ratio, is consistently correlated with hepatic steatosis in mice [121]. SAH accumulation inhibits PEMT, and SAH-mediated impairment of PEMT is linked to hepatic steatosis [121,122]. Additionally, in a transgenic model (HO mice) that expresses very low levels of CBS and high plasma concentrations of Hcy and SAH, a post-translational repression of PEMT that inversely correlates with liver steatosis is present, together with upregulation and down-regulation of phospholipid species and SAM/SAH ratios similar to those found in our CBS patients [123,124]. Together, these findings in CBSD patients highlight the impact of Hcy levels on SAM/SAH levels regardless of the underlying genetic defect, arguing for directions to be pursued to understand the phenotypic heterogeneity of severely affected patients with CBS deficiency, and to provide guidelines to design innovative strategies in this area.
  • Metabolomics towards precision medicine. Tough guidelines have been applied to design studies to attain (and analyze) data to be used in precision medicine and precision prevention approaches. The use of big biobanks and electronic medical records that integrate biological information with clinical data has strengthened and refined information from these studies [125]. Large, prospective, time-consuming and expensive studies are mandatory to validate older, and discover newer, metabolomics-based biomarkers with high translational chances [3]. This is predicted to uncover new pathological pathways and disease biomarkers, to improve disease prognoses, and facilitate treatment selection. To this end, information on metabolomics-based health data collection in families, and new imaging techniques to monitor changes in metabolite levels, are critical for the translation of metabolomics-based results into clinical and industrial application.
    Metabolomics-based health data collection in families. In the second half of the last century, health data collection in “healthy” individuals and their families helped predict disease through the identification of biomarkers suggestive of pre-clinical conditions, and allowed for informative decision making and ad hoc preventive strategies [126]. Such measures had substantial economical and welfare effects (when supported by human validation studies), and provided large-scale biological databases to help predict post-treatment outcomes [108]. Metabolomics-based health data collection is likely to be critical for improved big data analysis and tailored medical decisions. For instance, citrate, an important biomarker of cancer [127], is increased in older healthy individuals [73]. A comprehensive information on citrate levels in healthy individuals of different ages is key for the progress of precision medicine (e.g., preventive screenings and early phases of a malignancy) [128].
    New imaging techniques to monitor changes in metabolite levels. In the analysis of the data on CBSD patients summarized above, multi-dimensional scaling (MDS) analysis, based on lipid abundance, was implemented by the ‘DaMiRseq’ R/Bioconductor package [129] to identify specific clusters or batch effects. Differential analyses (CBSD patients vs. controls) were performed by the ‘limma’ R/Bioconductor package [130], implementing linear models adjusted for the effect of ‘Smoking’ [131,132]. The Benjamini–Hochberg procedure was used to control for the FDR. A lipid was deemed significant if the FDR adjusted p-value was <0.05 and the |log2(Fold Change)| > 1.5. Clustering analysis, performed by MDS showed that, except for smokers, CBSD and control groups were well separated both in positive and negative ion modes. In view of the key role of smoking in the top causes of death, the present example strongly supports the need for newer imaging techniques to strengthen the role of metabolomics in advanced research and avoid false overlapping in lipidomic analysis.
    A high likelihood of translation into a routine clinical test argues for large cohort multi-center studies to validate metabolomics-based biomarkers with high chances to impact clinical practice. Healthy individuals should be seriously considered in new metabolomics studies. Indeed, metabolomics is predicted to be critical for developing medical devices that are unique to a patient (or small groups of patients). However, metabolomics-based data should also help develop devices for the health population (or for field testing of a disease) [4]. To this end, professional and regulatory agencies should provide updated robust guidelines for study design, data acquisition and validation, to be applied from the very beginning of a project [79,80,133]. Conversion of results into products is maximal when ad hoc plans and paths are defined at the start of a project. Upon completion of data acquisition, identification of mechanisms leading to a metabolic pattern increase the chances of successful translation of results into clinical and industrial application [11]. Perhaps together with suppliers involved in developing analytical platforms, the search for cheap and easy miniaturized instruments will be critical for smart modifications of biomarkers. In this respect: (1) methods for absolute quantification of a wide range of metabolites using easy analytical instrumentation should be implemented. Rather than targeted, special attention should be devoted to newer untargeted quantitative metabolomics methods; (2) new better platforms are needed to work together with other omics to progressively increase the number of genes that expose to (or protect from) illnesses; (3) efforts to evaluate the influence of confounding factors (e.g., age, gender, ethnicity, diet) on metabolomics results should be implemented [73], and (4) validation studies in health and disease are urgently needed to remove potential bias. Validation is especially mandatory in view of the: (a) inter-laboratory variation in techniques of different metabolomics institutions [4]; (b) lack of common practices to validate potential biomarkers, (i.e., the absence of generally accepted procedures for metabolic profiling for biomarker discovery); and (c) use of metabolomics data as a source of potential pharmacologically active compounds [134,135].

5. Conclusions

Over the centuries, changes in medical attitudes have dramatically improved the cure, and sometimes prevented the development of diseases (e.g., tuberculosis). The attitude of metabolomics to reliably analyze metabolites, and identify new biological matrices is now established, and technological and computational improvements have greatly enhanced the translational capability of this omics. The existing applications of metabolomics in precision medicine translate to advancements in the diagnosis, prevention, and treatment of disease. Improving instrumentation and implementing standard analytical procedures is predicted to strengthen the impact of metabolomics in future medical care. The roadway of precision medicine and precision prevention is likely to be critical to validate old, and identify new, metabolomics-based biomarkers. Information from sources other than plasma and serum, and advanced pathophysiological analysis will likely refine the picture of a disease based on measurements of the plasma or serum metabolome [136,137].
Considering the information gathered in genomics- and transcriptomics-based initiatives, it is predicted that precise clinical decisions and precision treatments will largely abide by the accuracy of the information available, that is largely omics in nature. Truly integrated multi-omics analyses have not been widely applied. Major effort is now mandatory to develop the analytical infrastructure required to generate, analyze, and annotate multi-omics data and inform decision-making in precision medicine. Broad incorporation of machine learning techniques and systems to provide doctors with fully automated clinical analyzers are likely to be needed to assist in disease diagnosis and treatment and predict prognosis in precision medicine [110]. Major hurdles that omics (first, metabolomics) will face in this new dimension are largely ethical. Firstly, predictive diagnosis will change the relationships among patients and healthcare providers, and increase physician visits, laboratory tests, and patient anxiety. Presently, the poor pathophysiological information about the overwhelming amount of data generated to date hampers translation of metabolomics to clinical practice. A systematic approach to determining (genetic) causality is mandatory. Secondly, using genomic, clinical, personal, and environmental data collected from very large numbers of individuals from various populations, and connecting their health records, “non-responders“ to a treatment, might belong to definite minority populations An effort is needed against discrimination in access to treatments [125].
Table 1. Examples of information collected employing targeted and/or untargeted metabolomics approaches in experimental models of disease and in pre- and post-natal diagnoses in humans.
Table 1. Examples of information collected employing targeted and/or untargeted metabolomics approaches in experimental models of disease and in pre- and post-natal diagnoses in humans.
Models of Disease
Source of MaterialMain FindingsRefs.
Human, adultDysregulation of lipid metabolism and pathological inflammation in patients with COVID-19.[138]
Liver abnormalities involving carbon and nitrogen metabolism in moderate and severe COVID-19 patients [139]
Plasma phospholipid dysregulation in patients with cystathionine-beta synthase deficiency §[78]
Plasma levels of platelet-activating factor and its precursors in patients with familial hypercholesterolemia on treatment with PCSK9 inhibitors §[22]
In vivo thromboxane A2 biosynthesis and endothelial function in patients with familial hypercholesterolemia receiving PCSK-9 inhibitors therapy §[140]
Human, pediatricSerum phospholipid profile allows for the discrimination of infants who develop celiac disease before 8 years of age [141]
AnimalA targeted metabolomic approach to a mouse model of mucopolysaccharidosis IIIB identifies specific amino acid and fatty acid metabolic pathway alterations [142]
Mice model of Glutaric aciduria type I (GA-I, OMIM # 231670), an inborn error of metabolism caused by a deficiency of glutaryl-CoA dehydrogenase. * [143]
Reference Values as Related to Gender Differences
Human, adultSerum metabolomic profiles suggest influence of sex and oral contraceptive use.[144]
Human, pediatricEffect of gender on human premature blood metabolome in neonates.[145]
Effect of gender on urinary excretion of organic acids in children. °[146]
Effect of gender on blood metabolome of female and male human babies.[147]
AnimalEffect of gender on amino acid and carnitine levels in rat tissues (heart, liver, kidney)[148]
* Gaining insights into (brain) pathophysiology, and the development of new therapeutic interventions. ° relevance of analyzing human metabolome. § untargeted metabolomics, combined metabolomic and lipidomic approach.
Table 2. Examples of absolute quantification of metabolites using targeted approaches: source of metabolites, available methods and analytical platforms employed.
Table 2. Examples of absolute quantification of metabolites using targeted approaches: source of metabolites, available methods and analytical platforms employed.
Type, (Numbers), and Source of Metabolites QuantifiedQuantification MethodPlatformRefs.
Amino and non-amino organic acids (67), urine and serum samples. MCF derivatizationGC-MS/MS[46]
Polar primary metabolites (49), chickpea cultivarsBSTFA derivatization of primary metabolitesGC-MS[149]
Amino and non-amino organic acids
(50–100, human biological samples).
Calibration curve-free GC–MS method using MCF GC-MS[150]
Amino metabolites (124), renal cancer tissue, rat urine and plasma. Derivatization assisted sensitivity enhancement with 5-AIQC UPLC-MS/MS[151]
Lipids, lipidomic quantification (222), human serum samples. PRM QTOF LC-MS[152]
Amino acids and metabolites in the urea and tricarboxylic acid
cycles; biogenic amines; acylcarnitines; lipids,
(188, murine tissues).
Absolute IDQ TM p180 Kit (Biocrates)LC-MS/MS and FIA-MS/MS, UPLC MS/MS[153,154]
Essential and non-essential amino acids, phospholipids
(32, human breast cancer).
HR MASNMR[155,156]
Identifying, in one session, different classes of compounds from seeds (amygdalin), flowers (rutin), fruits (isovitexin) leaves
(shikimic acid) and stems (epicatechin) from
Crataegus rhipidophylla Gand (58).
Ratio methodNMR[157]
Legend. 5-AIQC: 5-aminoisoquinolyl-N-hydroxysuccinimidyl carbamate; BSTFA: N, O-bis-(trimethylsilyl)trifluoroacetamide; GC: gas chromatography; LC: liquid chromatography; MCF: methyl chloroformate derivatization; MS: mass spectrometry; UPLC: ultra performance liquid chromatography; PRM: parallel reaction monitoring; HR MAS—high-resolution magic angle spinning; FIA—flow injection analysis; NMR—nuclear magnetic resonance; QTOF—quadrupole time-of-flight.

Author Contributions

Conceptualization, A.D.M., M.G., M.R. and G.C.; methodology, M.C. (Marianna Caterino) and M.C. (Michele Costanzo); validation, M.C. (Marianna Caterino) and M.C. (Michele Costanzo); formal analysis, A.D.M. and M.G.; investigation, A.D.M. and M.G.; writing—original draft preparation, A.D.M. and M.G.; writing—review and editing M.C. (Marianna Caterino) and M.C. (Michele Costanzo); supervision, M.R. and G.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dunn, W.B. Current trends and future requirements for the mass spectrometric investigation of microbial, mammalian and plant metabolomes. Phys. Biol. 2008, 5, 011001. [Google Scholar] [CrossRef] [PubMed]
  2. Pinu, F.R.; Beale, D.J.; Paten, A.M.; Kouremenos, K.; Swarup, S.; Schirra, H.J.; Wishart, D. Systems Biology and Multi-Omics Integration: Viewpoints from the Metabolomics Research Community. Metabolites 2019, 9, 76. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Wishart, D.S. Emerging applications of metabolomics in drug discovery and precision medicine. Nat. Rev. Drug Discov. 2016, 15, 473–484. [Google Scholar] [CrossRef] [PubMed]
  4. Trivedi, D.K.; Hollywood, K.A.; Goodacre, R. Metabolomics for the masses: The future of metabolomics in a personalized world. New Horiz. Transl. Med. 2017, 3, 294–305. [Google Scholar] [CrossRef] [Green Version]
  5. Strimbu, K.; Tavel, J.A. What are biomarkers? Curr. Opin. HIV AIDS 2010, 5, 463–466. [Google Scholar] [CrossRef]
  6. Beger, R.D.; Dunn, W.B.; Bandukwala, A.; Bethan, B.; Broadhurst, D.; Clish, C.B.; Dasari, S.; Derr, L.; Evans, A.; Fischer, S.; et al. Towards quality assurance and quality control in untargeted metabolomics studies. Metabolomics 2019, 15, 4. [Google Scholar] [CrossRef]
  7. Chong, J.; Soufan, O.; Li, C.; Caraus, I.; Li, S.; Bourque, G.; Wishart, D.S.; Xia, J. MetaboAnalyst 4.0: Towards more transparent and integrative metabolomics analysis. Nucleic Acids Res. 2018, 46, W486–W494. [Google Scholar] [CrossRef] [Green Version]
  8. Cui, L.; Lu, H.; Lee, Y.H. Challenges and emergent solutions for LC-MS/MS based untargeted metabolomics in diseases. Mass Spectrom. Rev. 2018, 37, 772–792. [Google Scholar] [CrossRef]
  9. Djoumbou-Feunang, Y.; Pon, A.; Karu, N.; Zheng, J.; Li, C.; Arndt, D.; Gautam, M.; Allen, F.; Wishart, D.S. CFM-ID 3.0: Significantly Improved ESI-MS/MS Prediction and Compound Identification. Metabolites 2019, 9, 72. [Google Scholar] [CrossRef] [Green Version]
  10. Spadarella, G.; Di Minno, A.; Brunetti-Pierri, N.; Mahlangu, J.; Di Minno, G. The evolving landscape of gene therapy for congenital haemophilia: An unprecedented, problematic but promising opportunity for worldwide clinical studies. Blood Rev. 2021, 46, 100737. [Google Scholar] [CrossRef]
  11. Johnson, C.H.; Ivanisevic, J.; Siuzdak, G. Metabolomics: Beyond biomarkers and towards mechanisms. Nat. Rev. Mol. Cell Biol. 2016, 17, 451–459. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Kell, D.B.; Oliver, S.G. The metabolome 18 years on: A concept comes of age. Metabolomics 2016, 12, 148. [Google Scholar] [CrossRef] [PubMed]
  13. Wishart, D.S.; Mandal, R.; Stanislaus, A.; Ramirez-Gaona, M. Cancer Metabolomics and the Human Metabolome Database. Metabolites 2016, 6, 10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Di Minno, G.; Tremoli, E. Tailoring of medical treatment: Hemostasis and thrombosis towards precision medicine. Haematologica 2017, 102, 411–418. [Google Scholar] [CrossRef] [Green Version]
  15. Schrimpe-Rutledge, A.C.; Codreanu, S.G.; Sherrod, S.D.; McLean, J.A. Untargeted Metabolomics Strategies-Challenges and Emerging Directions. J. Am. Soc. Mass Spectrom. 2016, 27, 1897–1905. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Xia, J.; Broadhurst, D.I.; Wilson, M.; Wishart, D.S. Translational biomarker discovery in clinical metabolomics: An introductory tutorial. Metabolomics 2013, 9, 280–299. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Skelton, D.M.; Ekman, D.R.; Martinovic-Weigelt, D.; Ankley, G.T.; Villeneuve, D.L.; Teng, Q.; Collette, T.W. Metabolomics for in situ environmental monitoring of surface waters impacted by contaminants from both point and nonpoint sources. Environ. Sci. Technol. 2014, 48, 2395–2403. [Google Scholar] [CrossRef]
  18. Liu, J.; Semiz, S.; van der Lee, S.J.; van der Spek, A.; Verhoeven, A.; van Klinken, J.B.; Sijbrands, E.; Harms, A.C.; Hankemeier, T.; van Dijk, K.W.; et al. Metabolomics based markers predict type 2 diabetes in a 14-year follow-up study. Metabolomics 2017, 13, 104. [Google Scholar] [CrossRef] [Green Version]
  19. Savolainen, O.; Fagerberg, B.; Vendelbo Lind, M.; Sandberg, A.S.; Ross, A.B.; Bergstrom, G. Biomarkers for predicting type 2 diabetes development-Can metabolomics improve on existing biomarkers? PLoS ONE 2017, 12, e0177738. [Google Scholar] [CrossRef]
  20. Ganna, A.; Salihovic, S.; Sundstrom, J.; Broeckling, C.D.; Hedman, A.K.; Magnusson, P.K.; Pedersen, N.L.; Larsson, A.; Siegbahn, A.; Zilmer, M.; et al. Large-scale metabolomic profiling identifies novel biomarkers for incident coronary heart disease. PLoS Genet. 2014, 10, e1004801. [Google Scholar] [CrossRef]
  21. Bernardo, B.C.; Ooi, J.Y.Y.; Weeks, K.L.; Patterson, N.L.; McMullen, J.R. Understanding Key Mechanisms of Exercise-Induced Cardiac Protection to Mitigate Disease: Current Knowledge and Emerging Concepts. Physiol. Rev. 2018, 98, 419–475. [Google Scholar] [CrossRef] [PubMed]
  22. Di Minno, A.; Orsini, R.C.; Chiesa, M.; Cavalca, V.; Calcaterra, I.; Tripaldella, M.; Anesi, A.; Fiorelli, S.; Eligini, S.; Colombo, G.I.; et al. Treatment with PCSK9 Inhibitors in Patients with Familial Hypercholesterolemia Lowers Plasma Levels of Platelet-Activating Factor and Its Precursors: A Combined Metabolomic and Lipidomic Approach. Biomedicines 2021, 9, 1073. [Google Scholar] [CrossRef] [PubMed]
  23. Marcinkiewicz-Siemion, M.; Ciborowski, M.; Kretowski, A.; Musial, W.J.; Kaminski, K.A. Metabolomics—A wide-open door to personalized treatment in chronic heart failure? Int. J. Cardiol. 2016, 219, 156–163. [Google Scholar] [CrossRef] [PubMed]
  24. Mehta, K.Y.; Wu, H.J.; Menon, S.S.; Fallah, Y.; Zhong, X.; Rizk, N.; Unger, K.; Mapstone, M.; Fiandaca, M.S.; Federoff, H.J.; et al. Metabolomic biomarkers of pancreatic cancer: A meta-analysis study. Oncotarget 2017, 8, 68899–68915. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Nellis, M.; Caperton, C.O.; Liu, K.; Tran, V.; Go, Y.M.; Hallberg, L.M.; Ameredes, B.T.; Jones, D.P.; Boysen, G. Lung metabolome of 1,3-butadiene exposed Collaborative Cross mice reflects metabolic phenotype of human lung cancer. Toxicology 2021, 463, 152987. [Google Scholar] [CrossRef] [PubMed]
  26. Skaripa-Koukelli, I.; Hauton, D.; Walsby-Tickle, J.; Thomas, E.; Owen, J.; Lakshminarayanan, A.; Able, S.; McCullagh, J.; Carlisle, R.C.; Vallis, K.A. 3-Bromopyruvate-mediated MCT1-dependent metabolic perturbation sensitizes triple negative breast cancer cells to ionizing radiation. Cancer Metab. 2021, 9, 37. [Google Scholar] [CrossRef] [PubMed]
  27. Wang, K.X.; Du, G.H.; Qin, X.M.; Gao, L. Compound Kushen Injection intervenes metabolic reprogramming and epithelial-mesenchymal transition of HCC via regulating beta-catenin/c-Myc signaling. Phytomedicine 2021, 93, 153781. [Google Scholar] [CrossRef]
  28. Zhu, Q.; Huang, L.; Yang, Q.; Ao, Z.; Yang, R.; Krzesniak, J.; Lou, D.; Hu, L.; Dai, X.; Guo, F.; et al. Metabolomic analysis of exosomal-markers in esophageal squamous cell carcinoma. Nanoscale 2021, 13, 16457–16464. [Google Scholar] [CrossRef]
  29. Iwamoto, N.; Shimada, T. Recent advances in mass spectrometry-based approaches for proteomics and biologics: Great contribution for developing therapeutic antibodies. Pharmacol. Ther. 2018, 185, 147–154. [Google Scholar] [CrossRef]
  30. Lei, Z.; Huhman, D.V.; Sumner, L.W. Mass spectrometry strategies in metabolomics. J. Biol. Chem. 2011, 286, 25435–25442. [Google Scholar] [CrossRef] [Green Version]
  31. Beale, D.J.; Pinu, F.R.; Kouremenos, K.A.; Poojary, M.M.; Narayana, V.K.; Boughton, B.A.; Kanojia, K.; Dayalan, S.; Jones, O.A.H.; Dias, D.A. Review of recent developments in GC-MS approaches to metabolomics-based research. Metabolomics 2018, 14, 152. [Google Scholar] [CrossRef] [PubMed]
  32. Chetwynd, A.J.; Dunn, W.B.; Rodriguez-Blanco, G. Collection and Preparation of Clinical Samples for Metabolomics. Adv. Exp. Med. Biol. 2017, 965, 19–44. [Google Scholar] [CrossRef] [PubMed]
  33. Pinu, F.R.; Villas-Boas, S.G. Extracellular Microbial Metabolomics: The State of the Art. Metabolites 2017, 7, 43. [Google Scholar] [CrossRef] [PubMed]
  34. Members, M.S.I.B.; Sansone, S.A.; Fan, T.; Goodacre, R.; Griffin, J.L.; Hardy, N.W.; Kaddurah-Daouk, R.; Kristal, B.S.; Lindon, J.; Mendes, P.; et al. The metabolomics standards initiative. Nat. Biotechnol. 2007, 25, 846–848. [Google Scholar] [CrossRef] [PubMed]
  35. Sumner, L.W.; Amberg, A.; Barrett, D.; Beale, M.H.; Beger, R.; Daykin, C.A.; Fan, T.W.; Fiehn, O.; Goodacre, R.; Griffin, J.L.; et al. Proposed minimum reporting standards for chemical analysis Chemical Analysis Working Group (CAWG) Metabolomics Standards Initiative (MSI). Metabolomics 2007, 3, 211–221. [Google Scholar] [CrossRef] [Green Version]
  36. Bernini, P.; Bertini, I.; Luchinat, C.; Nincheri, P.; Staderini, S.; Turano, P. Standard operating procedures for pre-analytical handling of blood and urine for metabolomic studies and biobanks. J. Biomol. NMR 2011, 49, 231–243. [Google Scholar] [CrossRef]
  37. Broadhurst, D.; Goodacre, R.; Reinke, S.N.; Kuligowski, J.; Wilson, I.D.; Lewis, M.R.; Dunn, W.B. Guidelines and considerations for the use of system suitability and quality control samples in mass spectrometry assays applied in untargeted clinical metabolomic studies. Metabolomics 2018, 14, 72. [Google Scholar] [CrossRef] [Green Version]
  38. Fernie, A.R.; Aharoni, A.; Willmitzer, L.; Stitt, M.; Tohge, T.; Kopka, J.; Carroll, A.J.; Saito, K.; Fraser, P.D.; DeLuca, V. Recommendations for reporting metabolite data. Plant Cell 2011, 23, 2477–2482. [Google Scholar] [CrossRef] [Green Version]
  39. Torii, Y.; Kawano, Y.; Sato, H.; Sasaki, K.; Fujimori, T.; Kawada, J.-i.; Takikawa, O.; Lim, C.K.; Guillemin, G.J.; Ohashi, Y.; et al. Quantitative metabolome profiling reveals the involvement of the kynurenine pathway in influenza-associated encephalopathy. Metabolomics 2016, 12, 84. [Google Scholar] [CrossRef]
  40. Ghazi, N.; Arjmand, M.; Akbari, Z.; Mellati, A.O.; Saheb-Kashaf, H.; Zamani, Z. (1)H NMR- based metabolomics approaches as non- invasive tools for diagnosis of endometriosis. Int. J. Reprod. Biomed. 2016, 14, 1–8. [Google Scholar] [CrossRef] [Green Version]
  41. Goldansaz, S.A.; Guo, A.C.; Sajed, T.; Steele, M.A.; Plastow, G.S.; Wishart, D.S. Livestock metabolomics and the livestock metabolome: A systematic review. PLoS ONE 2017, 12, e0177675. [Google Scholar] [CrossRef] [Green Version]
  42. Sun, H.Z.; Wang, D.M.; Wang, B.; Wang, J.K.; Liu, H.Y.; Guan, L.L.; Liu, J.X. Metabolomics of four biofluids from dairy cows: Potential biomarkers for milk production and quality. J. Proteome Res. 2015, 14, 1287–1298. [Google Scholar] [CrossRef] [PubMed]
  43. Wishart, D.S. Computational strategies for metabolite identification in metabolomics. Bioanalysis 2009, 1, 1579–1596. [Google Scholar] [CrossRef]
  44. Bennett, B.D.; Yuan, J.; Kimball, E.H.; Rabinowitz, J.D. Absolute quantitation of intracellular metabolite concentrations by an isotope ratio-based approach. Nat. Protoc. 2008, 3, 1299–1311. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Hu, K.; Westler, W.M.; Markley, J.L. Simultaneous quantification and identification of individual chemicals in metabolite mixtures by two-dimensional extrapolated time-zero (1)H-(13)C HSQC (HSQC(0)). J. Am. Chem. Soc. 2011, 133, 1662–1665. [Google Scholar] [CrossRef] [PubMed]
  46. Kvitvang, H.F.; Andreassen, T.; Adam, T.; Villas-Boas, S.G.; Bruheim, P. Highly sensitive GC/MS/MS method for quantitation of amino and nonamino organic acids. Anal. Chem. 2011, 83, 2705–2711. [Google Scholar] [CrossRef]
  47. Lien, S.K.; Kvitvang, H.F.; Bruheim, P. Utilization of a deuterated derivatization agent to synthesize internal standards for gas chromatography-tandem mass spectrometry quantification of silylated metabolites. J. Chromatogr. A 2012, 1247, 118–124. [Google Scholar] [CrossRef]
  48. Martineau, E.; Tea, I.; Akoka, S.; Giraudeau, P. Absolute quantification of metabolites in breast cancer cell extracts by quantitative 2D (1) H INADEQUATE NMR. NMR Biomed. 2012, 25, 985–992. [Google Scholar] [CrossRef]
  49. Tumanov, S.; Zubenko, Y.; Obolonkin, V.; Greenwood, D.R.; Shmanai, V.; Villas-Bôas, S.G. Calibration curve-free GC–MS method for quantitation of amino and non-amino organic acids in biological samples. Metabolomics 2016, 12, 64. [Google Scholar] [CrossRef]
  50. Vielhauer, O.; Zakhartsev, M.; Horn, T.; Takors, R.; Reuss, M. Simplified absolute metabolite quantification by gas chromatography-isotope dilution mass spectrometry on the basis of commercially available source material. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2011, 879, 3859–3870. [Google Scholar] [CrossRef]
  51. Zhou, J.; Yin, Y. Strategies for large-scale targeted metabolomics quantification by liquid chromatography-mass spectrometry. Analyst 2016, 141, 6362–6373. [Google Scholar] [CrossRef] [PubMed]
  52. Gatalica, Z.; Xiu, J.; Swensen, J.; Vranic, S. Comprehensive analysis of cancers of unknown primary for the biomarkers of response to immune checkpoint blockade therapy. Eur. J. Cancer 2018, 94, 179–186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Godoy-Vitorino, F.; Ortiz-Morales, G.; Romaguera, J.; Sanchez, M.M.; Martinez-Ferrer, M.; Chorna, N. Discriminating high-risk cervical Human Papilloma Virus infections with urinary biomarkers via non-targeted GC-MS-based metabolomics. PLoS ONE 2018, 13, e0209936. [Google Scholar] [CrossRef] [PubMed]
  54. Melvin, S.D.; Lanctot, C.M.; Doriean, N.J.C.; Carroll, A.R.; Bennett, W.W. Untargeted NMR-based metabolomics for field-scale monitoring: Temporal reproducibility and biomarker discovery in mosquitofish (Gambusia holbrooki) from a metal(loid)-contaminated wetland. Environ. Pollut. 2018, 243, 1096–1105. [Google Scholar] [CrossRef] [PubMed]
  55. Sinclair, I.; Bachman, M.; Addison, D.; Rohman, M.; Murray, D.C.; Davies, G.; Mouchet, E.; Tonge, M.E.; Stearns, R.G.; Ghislain, L.; et al. Acoustic Mist Ionization Platform for Direct and Contactless Ultrahigh-Throughput Mass Spectrometry Analysis of Liquid Samples. Anal. Chem. 2019, 91, 3790–3794. [Google Scholar] [CrossRef] [Green Version]
  56. Stoll, D.R.; Carr, P.W. Two-Dimensional Liquid Chromatography: A State of the Art Tutorial. Anal. Chem. 2017, 89, 519–531. [Google Scholar] [CrossRef] [PubMed]
  57. Lu, R.; Zhou, X.; Yin, Q.; Hu, J.; Ni, Z. Miniature nuclear magnetic resonance spectrometer using a partially enclosed permanent magnet. Instrum. Sci. Technol. 2017, 45, 324–337. [Google Scholar] [CrossRef]
  58. Snyder, D.T.; Pulliam, C.J.; Ouyang, Z.; Cooks, R.G. Miniature and Fieldable Mass Spectrometers: Recent Advances. Anal. Chem. 2016, 88, 2–29. [Google Scholar] [CrossRef] [Green Version]
  59. Zhou, X.; Liu, J.; Cooks, R.G.; Ouyang, Z. Development of miniature mass spectrometry systems for bioanalysis outside the conventional laboratories. Bioanalysis 2014, 6, 1497–1508. [Google Scholar] [CrossRef] [Green Version]
  60. Lu, W.; Bennett, B.D.; Rabinowitz, J.D. Analytical strategies for LC-MS-based targeted metabolomics. J. Chromatogr. B Anal. Technol. Biomed. Life Sci 2008, 871, 236–242. [Google Scholar] [CrossRef] [Green Version]
  61. Hirschfeld, T. Instrumentation in the next decade. Science 1985, 230, 286–291. [Google Scholar] [CrossRef] [PubMed]
  62. Patel, K.N.; Patel, J.K.; Patel, M.P.; Rajput, G.C.; Pattel, H.A. Introduction to hyphenated techniques and their applications in pharmacy. Pharm. Methods 2010, 1, 2–13. [Google Scholar] [CrossRef]
  63. Bartroff, J.; Song, J. Sequential Tests of Multiple Hypotheses Controlling False Discovery and Nondiscovery Rates. Seq. Anal. 2020, 39, 65–91. [Google Scholar] [CrossRef] [PubMed]
  64. Storey, J.D.; Tibshirani, R. Statistical significance for genomewide studies. Proc. Natl. Acad. Sci. USA 2003, 100, 9440–9445. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Wishart, D.S. Advances in metabolite identification. Bioanalysis 2011, 3, 1769–1782. [Google Scholar] [CrossRef]
  66. Wishart, D.S. Metabolomics: Applications to Food Science and Nutrition Research. Trends Food Sci. Technol. 2008, 19, 482–493. [Google Scholar] [CrossRef]
  67. Pinu, F.R.; Edwards, P.J.B.; Jouanneau, S.; Kilmartin, P.A.; Gardner, R.C.; Villas-Bôas, S.G.J.M. Sauvignon blanc metabolomics: Grape juice metabolites affecting the development of varietal thiols and other aroma compounds in wines. Metabolomics 2013, 10, 556–573. [Google Scholar] [CrossRef]
  68. Smart, K.F.; Aggio, R.B.; Van Houtte, J.R.; Villas-Boas, S.G. Analytical platform for metabolome analysis of microbial cells using methyl chloroformate derivatization followed by gas chromatography-mass spectrometry. Nat. Protoc. 2010, 5, 1709–1729. [Google Scholar] [CrossRef]
  69. Xia, J.; Wishart, D.S. Using MetaboAnalyst 3.0 for Comprehensive Metabolomics Data Analysis. Curr. Protoc. Bioinform. 2016, 55, 14.10.1–14.10.91. [Google Scholar] [CrossRef]
  70. Pinu, F.R.; Tumanov, S.; Grose, C.; Raw, V.; Albright, A.; Stuart, L.; Villas-Boas, S.G.; Martin, D.; Harker, R.; Greven, M. Juice Index: An integrated Sauvignon blanc grape and wine metabolomics database shows mainly seasonal differences. Metabolomics 2019, 15, 3. [Google Scholar] [CrossRef]
  71. Guijas, C.; Montenegro-Burke, J.R.; Domingo-Almenara, X.; Palermo, A.; Warth, B.; Hermann, G.; Koellensperger, G.; Huan, T.; Uritboonthai, W.; Aisporna, A.E.; et al. METLIN: A Technology Platform for Identifying Knowns and Unknowns. Anal. Chem. 2018, 90, 3156–3164. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Pinu, F.R.; Goldansaz, S.A.; Jaine, J. Translational Metabolomics: Current Challenges and Future Opportunities. Metabolites 2019, 9, 108. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Dunn, W.B.; Lin, W.; Broadhurst, D.; Begley, P.; Brown, M.; Zelena, E.; Vaughan, A.A.; Halsall, A.; Harding, N.; Knowles, J.D.; et al. Molecular phenotyping of a UK population: Defining the human serum metabolome. Metabolomics 2015, 11, 9–26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Munafo, M.R.; Nosek, B.A.; Bishop, D.V.M.; Button, K.S.; Chambers, C.D.; du Sert, N.P.; Simonsohn, U.; Wagenmakers, E.J.; Ware, J.J.; Ioannidis, J.P.A. A manifesto for reproducible science. Nat. Hum. Behav. 2017, 1, 0021. [Google Scholar] [CrossRef] [Green Version]
  75. Poste, G. Bring on the biomarkers. Nature 2011, 469, 156–157. [Google Scholar] [CrossRef]
  76. Harris, R. Overview of screening: Where we are and where we may be headed. Epidemiol. Rev. 2011, 33, 1–6. [Google Scholar] [CrossRef]
  77. Khoury, M.J.; McCabe, L.L.; McCabe, E.R. Population screening in the age of genomic medicine. N. Engl. J. Med. 2003, 348, 50–58. [Google Scholar] [CrossRef] [Green Version]
  78. Di Minno, A.; Anesi, A.; Chiesa, M.; Cirillo, F.; Colombo, G.I.; Orsini, R.C.; Capasso, F.; Morisco, F.; Fiorelli, S.; Eligini, S.; et al. Plasma phospholipid dysregulation in patients with cystathionine-beta synthase deficiency. Nutr. Metab. Cardiovasc. Dis. 2020, 30, 2286–2295. [Google Scholar] [CrossRef]
  79. Pepe, M.S.; Feng, Z.; Janes, H.; Bossuyt, P.M.; Potter, J.D. Pivotal evaluation of the accuracy of a biomarker used for classification or prediction: Standards for study design. J. Natl. Cancer Inst. 2008, 100, 1432–1438. [Google Scholar] [CrossRef] [Green Version]
  80. Pinu, F.R. Metabolomics: Applications to Food Safety and Quality Research. In Microbial Metabolomics: Applications in Clinical, Environmental, and Industrial Microbiology; Beale, D., Kouremenos, K., Palombo, E., Eds.; Springer International Publishing: Berlin/Heidelberg, Germany, 2016; pp. 225–259. [Google Scholar] [CrossRef] [Green Version]
  81. Guo, L.; Milburn, M.V.; Ryals, J.A.; Lonergan, S.C.; Mitchell, M.W.; Wulff, J.E.; Alexander, D.C.; Evans, A.M.; Bridgewater, B.; Miller, L.; et al. Plasma metabolomic profiles enhance precision medicine for volunteers of normal health. Proc. Natl. Acad. Sci. USA 2015, 112, E4901–E4910. [Google Scholar] [CrossRef] [Green Version]
  82. Chung, L.; Moore, K.; Phillips, L.; Boyle, F.M.; Marsh, D.J.; Baxter, R.C. Novel serum protein biomarker panel revealed by mass spectrometry and its prognostic value in breast cancer. Breast Cancer Res. 2014, 16, R63. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Tessitore, A.; Gaggiano, A.; Cicciarelli, G.; Verzella, D.; Capece, D.; Fischietti, M.; Zazzeroni, F.; Alesse, E. Serum biomarkers identification by mass spectrometry in high-mortality tumors. Int. J. Proteom. 2013, 2013, 125858. [Google Scholar] [CrossRef] [PubMed]
  84. Vanmassenhove, J.; Vanholder, R.; Nagler, E.; Van Biesen, W. Urinary and serum biomarkers for the diagnosis of acute kidney injury: An in-depth review of the literature. Nephrol. Dial. Transpl. 2013, 28, 254–273. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Kim, K.; Taylor, S.L.; Ganti, S.; Guo, L.; Osier, M.V.; Weiss, R.H. Urine metabolomic analysis identifies potential biomarkers and pathogenic pathways in kidney cancer. OMICS 2011, 15, 293–303. [Google Scholar] [CrossRef] [Green Version]
  86. Laiakis, E.C.; Morris, G.A.; Fornace, A.J.; Howie, S.R. Metabolomic analysis in severe childhood pneumonia in the Gambia, West Africa: Findings from a pilot study. PLoS ONE 2010, 5, e12655. [Google Scholar] [CrossRef]
  87. Shen, C.; Sun, Z.; Chen, D.; Su, X.; Jiang, J.; Li, G.; Lin, B.; Yan, J. Developing urinary metabolomic signatures as early bladder cancer diagnostic markers. Omics A J. Integr. Biol. 2015, 19, 1–11. [Google Scholar] [CrossRef] [Green Version]
  88. Trivedi, D.K.; Iles, R.K. Shotgun metabolomic profiles in maternal urine identify potential mass spectral markers of abnormal fetal biochemistry—Dihydrouracil and progesterone in the metabolism of Down syndrome. Biomed. Chromatogr. 2015, 29, 1173–1183. [Google Scholar] [CrossRef]
  89. Schnabel, R.; Fijten, R.; Smolinska, A.; Dallinga, J.; Boumans, M.L.; Stobberingh, E.; Boots, A.; Roekaerts, P.; Bergmans, D.; van Schooten, F.J. Analysis of volatile organic compounds in exhaled breath to diagnose ventilator-associated pneumonia. Sci. Rep. 2015, 5, 17179. [Google Scholar] [CrossRef] [Green Version]
  90. Sethi, S.; Nanda, R.; Chakraborty, T. Clinical application of volatile organic compound analysis for detecting infectious diseases. Clin. Microbiol. Rev. 2013, 26, 462–475. [Google Scholar] [CrossRef] [Green Version]
  91. van de Kant, K.D.; van der Sande, L.J.; Jobsis, Q.; van Schayck, O.C.; Dompeling, E. Clinical use of exhaled volatile organic compounds in pulmonary diseases: A systematic review. Respir. Res. 2012, 13, 117. [Google Scholar] [CrossRef] [Green Version]
  92. Pashayan, N.; Pharoah, P. Population-based screening in the era of genomics. Per. Med. 2012, 9, 451–455. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Di Minno, G.; Mannucci, P.M.; Ironside, J.W.; Perno, C.F.; Gurtler, L.; Aledort, L. Convalescent plasma for administration of passive antibodies against viral agents. Haematologica 2020, 105, 2710–2715. [Google Scholar] [CrossRef]
  94. Di Minno, G.; Perno, C.F.; Tiede, A.; Navarro, D.; Canaro, M.; Guertler, L.; Ironside, J.W. Current concepts in the prevention of pathogen transmission via blood/plasma-derived products for bleeding disorders. Blood Rev. 2016, 30, 35–48. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Di Minno, G.; Canaro, M.; Ironside, J.W.; Navarro, D.; Perno, C.F.; Tiede, A.; Guertler, L. Pathogen safety of long-term treatments for bleeding disorders: (Un)predictable risks and evolving threats. Semin. Thromb. Hemost. 2013, 39, 779–793. [Google Scholar] [CrossRef] [PubMed]
  96. Di Minno, G.; Canaro, M.; Ironside, J.W.; Navarro, D.; Perno, C.F.; Tiede, A.; Gurtler, L. Pathogen safety of long-term treatments for bleeding disorders: Still relevant to current practice. Haematologica 2013, 98, 1495–1498. [Google Scholar] [CrossRef] [PubMed]
  97. Jones, K.E.; Patel, N.G.; Levy, M.A.; Storeygard, A.; Balk, D.; Gittleman, J.L.; Daszak, P. Global trends in emerging infectious diseases. Nature 2008, 451, 990–993. [Google Scholar] [CrossRef]
  98. Morens, D.M.; Folkers, G.K.; Fauci, A.S. The challenge of emerging and re-emerging infectious diseases. Nature 2004, 430, 242–249. [Google Scholar] [CrossRef]
  99. Weiss, R.A.; McMichael, A.J. Social and environmental risk factors in the emergence of infectious diseases. Nat. Med. 2004, 10, S70–S76. [Google Scholar] [CrossRef]
  100. Wolfe, N.D.; Dunavan, C.P.; Diamond, J. Origins of major human infectious diseases. Nature 2007, 447, 279–283. [Google Scholar] [CrossRef]
  101. Zappa, A.; Amendola, A.; Romano, L.; Zanetti, A. Emerging and re-emerging viruses in the era of globalisation. Blood Transfus. 2009, 7, 167–171. [Google Scholar] [CrossRef]
  102. Di Minno, G.; Navarro, D.; Perno, C.F.; Canaro, M.; Gurtler, L.; Ironside, J.W.; Eichler, H.; Tiede, A. Pathogen reduction/inactivation of products for the treatment of bleeding disorders: What are the processes and what should we say to patients? Ann. Hematol. 2017, 96, 1253–1270. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. O’Neill, J. Antimicrobial Resistance: Tackling a Crisis for the Health and Wealth of Nations. Available online: https://amr-review.org/sites/default/files/AMR%20Review%20Paper%20-%20Tackling%20a%20crisis%20for%20the%20health%20and%20wealth%20of%20nations_1.pdf (accessed on 31 March 2022).
  104. Beger, R.D.; Dunn, W.; Schmidt, M.A.; Gross, S.S.; Kirwan, J.A.; Cascante, M.; Brennan, L.; Wishart, D.S.; Oresic, M.; Hankemeier, T.; et al. Metabolomics enables precision medicine: “A White Paper, Community Perspective”. Metabolomics 2016, 12, 149. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Collins, F.S.; Varmus, H. A new initiative on precision medicine. N. Engl. J. Med. 2015, 372, 793–795. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Garraway, L.A.; Lander, E.S. Lessons from the cancer genome. Cell 2013, 153, 17–37. [Google Scholar] [CrossRef] [Green Version]
  107. Hayden, E.C. Technology: The $1,000 genome. Nature 2014, 507, 294–295. [Google Scholar] [CrossRef] [Green Version]
  108. Baker, M. Big biology: The ‘omes puzzle. Nature 2013, 494, 416–419. [Google Scholar] [CrossRef] [Green Version]
  109. Olivier, M.; Asmis, R.; Hawkins, G.A.; Howard, T.D.; Cox, L.A. The Need for Multi-Omics Biomarker Signatures in Precision Medicine. Int. J. Mol. Sci. 2019, 20, 4781. [Google Scholar] [CrossRef] [Green Version]
  110. Azad, R.K.; Shulaev, V. Metabolomics technology and bioinformatics for precision medicine. Brief. Bioinform. 2019, 20, 1957–1971. [Google Scholar] [CrossRef]
  111. Fuster, J.J.; Ouchi, N.; Gokce, N.; Walsh, K. Obesity-Induced Changes in Adipose Tissue Microenvironment and Their Impact on Cardiovascular Disease. Circ. Res. 2016, 118, 1786–1807. [Google Scholar] [CrossRef] [Green Version]
  112. Ziegelstein, R.C. Personomics. JAMA Intern. Med. 2015, 175, 888–889. [Google Scholar] [CrossRef]
  113. Lam, S.M.; Wang, Y.; Li, B.; Du, J.; Shui, G. Metabolomics through the lens of precision cardiovascular medicine. J. Genet. Genom. 2017, 44, 127–138. [Google Scholar] [CrossRef] [PubMed]
  114. Zenobi, R. Single-cell metabolomics: Analytical and biological perspectives. Science 2013, 342, 1243259. [Google Scholar] [CrossRef]
  115. Liu, X.; Locasale, J.W. Metabolomics reveals intratumor heterogeneity—Implications for precision medicine. EBioMedicine 2017, 19, 4–5. [Google Scholar] [CrossRef] [Green Version]
  116. Di Minno, G.; Davi, G.; Margaglione, M.; Cirillo, F.; Grandone, E.; Ciabattoni, G.; Catalano, I.; Strisciuglio, P.; Andria, G.; Patrono, C.; et al. Abnormally high thromboxane biosynthesis in homozygous homocystinuria. Evidence for platelet involvement and probucol-sensitive mechanism. J. Clin. Investig. 1993, 92, 1400–1406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Austin, R.C.; Lentz, S.R.; Werstuck, G.H. Role of hyperhomocysteinemia in endothelial dysfunction and atherothrombotic disease. Cell Death Differ. 2004, 11 (Suppl. S1), S56–S64. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Beard, R.S., Jr.; Bearden, S.E. Vascular complications of cystathionine beta-synthase deficiency: Future directions for homocysteine-to-hydrogen sulfide research. Am. J. Physiol. Heart Circ. Physiol. 2011, 300, H13–H26. [Google Scholar] [CrossRef]
  119. Davi, G.; Di Minno, G.; Coppola, A.; Andria, G.; Cerbone, A.M.; Madonna, P.; Tufano, A.; Falco, A.; Marchesani, P.; Ciabattoni, G.; et al. Oxidative stress and platelet activation in homozygous homocystinuria. Circulation 2001, 104, 1124–1128. [Google Scholar] [CrossRef] [Green Version]
  120. Morris, A.A.; Kozich, V.; Santra, S.; Andria, G.; Ben-Omran, T.I.; Chakrapani, A.B.; Crushell, E.; Henderson, M.J.; Hochuli, M.; Huemer, M.; et al. Guidelines for the diagnosis and management of cystathionine beta-synthase deficiency. J. Inherit. Metab. Dis. 2017, 40, 49–74. [Google Scholar] [CrossRef] [Green Version]
  121. Li, Z.; Agellon, L.B.; Allen, T.M.; Umeda, M.; Jewell, L.; Mason, A.; Vance, D.E. The ratio of phosphatidylcholine to phosphatidylethanolamine influences membrane integrity and steatohepatitis. Cell Metab. 2006, 3, 321–331. [Google Scholar] [CrossRef] [Green Version]
  122. Vance, D.E. Physiological roles of phosphatidylethanolamine N-methyltransferase. Biochim. Et Biophys. Acta 2013, 1831, 626–632. [Google Scholar] [CrossRef] [Green Version]
  123. Jacobs, R.L.; Jiang, H.; Kennelly, J.P.; Orlicky, D.J.; Allen, R.H.; Stabler, S.P.; Maclean, K.N. Cystathionine beta-synthase deficiency alters hepatic phospholipid and choline metabolism: Post-translational repression of phosphatidylethanolamine N-methyltransferase is a consequence rather than a cause of liver injury in homocystinuria. Mol. Genet. Metab. 2017, 120, 325–336. [Google Scholar] [CrossRef] [PubMed]
  124. Maclean, K.N.; Sikora, J.; Kozich, V.; Jiang, H.; Greiner, L.S.; Kraus, E.; Krijt, J.; Overdier, K.H.; Collard, R.; Brodsky, G.L.; et al. A novel transgenic mouse model of CBS-deficient homocystinuria does not incur hepatic steatosis or fibrosis and exhibits a hypercoagulative phenotype that is ameliorated by betaine treatment. Mol. Genet. Metab. 2010, 101, 153–162. [Google Scholar] [CrossRef] [PubMed]
  125. Dausset, J. Journal of Biomedicine and Biotechnology. J. Biomed. Biotechnol. 2001, 1, 1–2. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Andersson, C.; Johnson, A.D.; Benjamin, E.J.; Levy, D.; Vasan, R.S. 70-year legacy of the Framingham Heart Study. Nat. Rev. Cardiol. 2019, 16, 687–698. [Google Scholar] [CrossRef]
  127. Mosaoa, R.; Kasprzyk-Pawelec, A.; Fernandez, H.R.; Avantaggiati, M.L. The Mitochondrial Citrate Carrier SLC25A1/CIC and the Fundamental Role of Citrate in Cancer, Inflammation and Beyond. Biomolecules 2021, 11, 141. [Google Scholar] [CrossRef]
  128. Dhanasekaran, S.M.; Barrette, T.R.; Ghosh, D.; Shah, R.; Varambally, S.; Kurachi, K.; Pienta, K.J.; Rubin, M.A.; Chinnaiyan, A.M. Delineation of prognostic biomarkers in prostate cancer. Nature 2001, 412, 822–826. [Google Scholar] [CrossRef]
  129. Chiesa, M.; Colombo, G.I.; Piacentini, L. DaMiRseq-an R/Bioconductor package for data mining of RNA-Seq data: Normalization, feature selection and classification. Bioinformatics 2018, 34, 1416–1418. [Google Scholar] [CrossRef]
  130. Ritchie, M.E.; Phipson, B.; Wu, D.; Hu, Y.; Law, C.W.; Shi, W.; Smyth, G.K. limma powers differential expression analyses for RNA-sequencing and microarray studies. Nucleic Acids Res. 2015, 43, e47. [Google Scholar] [CrossRef]
  131. O’Callaghan, P.; Meleady, R.; Fitzgerald, T.; Graham, I. Smoking and plasma homocysteine. Eur. Heart J. 2002, 23, 1580–1586. [Google Scholar] [CrossRef] [Green Version]
  132. Sobczak, A.J. The effects of tobacco smoke on the homocysteine level--a risk factor of atherosclerosis. Addict. Biol. 2003, 8, 147–158. [Google Scholar] [CrossRef]
  133. Snyder, N.W.; Mesaros, C.; Blair, I.A. Translational metabolomics in cancer research. Biomark. Med. 2015, 9, 821–834. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Albright, J.C.; Henke, M.T.; Soukup, A.A.; McClure, R.A.; Thomson, R.J.; Keller, N.P.; Kelleher, N.L. Large-scale metabolomics reveals a complex response of Aspergillus nidulans to epigenetic perturbation. ACS Chem. Biol. 2015, 10, 1535–1541. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Rattray, N.J.W.; Daouk, R.K. Pharmacometabolomics and Precision Medicine Special Issue Editorial. Metabolomics 2017, 13, 59. [Google Scholar] [CrossRef] [Green Version]
  136. Broadhurst, D.I.; Kell, D.B. Statistical strategies for avoiding false discoveries in metabolomics and related experiments. Metabolomics 2006, 2, 171–196. [Google Scholar] [CrossRef] [Green Version]
  137. Dunn, W.B.; Wilson, I.D.; Nicholls, A.W.; Broadhurst, D. The importance of experimental design and QC samples in large-scale and MS-driven untargeted metabolomic studies of humans. Bioanalysis 2012, 4, 2249–2264. [Google Scholar] [CrossRef] [Green Version]
  138. Caterino, M.; Gelzo, M.; Sol, S.; Fedele, R.; Annunziata, A.; Calabrese, C.; Fiorentino, G.; D’Abbraccio, M.; Dell’Isola, C.; Fusco, F.M.; et al. Dysregulation of lipid metabolism and pathological inflammation in patients with COVID-19. Sci. Rep. 2021, 11, 2941. [Google Scholar] [CrossRef]
  139. Caterino, M.; Costanzo, M.; Fedele, R.; Cevenini, A.; Gelzo, M.; Di Minno, A.; Andolfo, I.; Capasso, M.; Russo, R.; Annunziata, A.; et al. The Serum Metabolome of Moderate and Severe COVID-19 Patients Reflects Possible Liver Alterations Involving Carbon and Nitrogen Metabolism. Int. J. Mol. Sci. 2021, 22, 9548. [Google Scholar] [CrossRef]
  140. Di Minno, A.; Gentile, M.; Iannuzzo, G.; Calcaterra, I.; Tripaldella, M.; Porro, B.; Cavalca, V.; Di Taranto, M.D.; Tremoli, E.; Fortunato, G.; et al. Endothelial function improvement in patients with familial hypercholesterolemia receiving PCSK-9 inhibitors on top of maximally tolerated lipid lowering therapy. Thromb. Res. 2020, 194, 229–236. [Google Scholar] [CrossRef]
  141. Auricchio, R.; Galatola, M.; Cielo, D.; Amoresano, A.; Caterino, M.; De Vita, E.; Illiano, A.; Troncone, R.; Greco, L.; Ruoppolo, M. A Phospholipid Profile at 4 Months Predicts the Onset of Celiac Disease in at-Risk Infants. Sci. Rep. 2019, 9, 14303. [Google Scholar] [CrossRef]
  142. De Pasquale, V.; Caterino, M.; Costanzo, M.; Fedele, R.; Ruoppolo, M.; Pavone, L.M. Targeted Metabolomic Analysis of a Mucopolysaccharidosis IIIB Mouse Model Reveals an Imbalance of Branched-Chain Amino Acid and Fatty Acid Metabolism. Int. J. Mol. Sci. 2020, 21, 4211. [Google Scholar] [CrossRef]
  143. Gonzalez Melo, M.; Remacle, N.; Cudre-Cung, H.P.; Roux, C.; Poms, M.; Cudalbu, C.; Barroso, M.; Gersting, S.W.; Feichtinger, R.G.; Mayr, J.A.; et al. The first knock-in rat model for glutaric aciduria type I allows further insights into pathophysiology in brain and periphery. Mol. Genet. Metab. 2021, 133, 157–181. [Google Scholar] [CrossRef] [PubMed]
  144. Ruoppolo, M.; Campesi, I.; Scolamiero, E.; Pecce, R.; Caterino, M.; Cherchi, S.; Mercuro, G.; Tonolo, G.; Franconi, F. Serum metabolomic profiles suggest influence of sex and oral contraceptive use. Am. J. Transl. Res. 2014, 6, 614–624. [Google Scholar] [PubMed]
  145. Caterino, M.; Ruoppolo, M.; Costanzo, M.; Albano, L.; Crisci, D.; Sotgiu, G.; Saderi, L.; Montella, A.; Franconi, F.; Campesi, I. Sex Affects Human Premature Neonates’ Blood Metabolome According to Gestational Age, Parenteral Nutrition, and Caffeine Treatment. Metabolites 2021, 11, 158. [Google Scholar] [CrossRef] [PubMed]
  146. Caterino, M.; Ruoppolo, M.; Villani, G.R.D.; Marchese, E.; Costanzo, M.; Sotgiu, G.; Dore, S.; Franconi, F.; Campesi, I. Influence of Sex on Urinary Organic Acids: A Cross-Sectional Study in Children. Int. J. Mol. Sci. 2020, 21, 582. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Ruoppolo, M.; Scolamiero, E.; Caterino, M.; Mirisola, V.; Franconi, F.; Campesi, I. Female and male human babies have distinct blood metabolomic patterns. Mol. Biosyst. 2015, 11, 2483–2492. [Google Scholar] [CrossRef]
  148. Ruoppolo, M.; Caterino, M.; Albano, L.; Pecce, R.; Di Girolamo, M.G.; Crisci, D.; Costanzo, M.; Milella, L.; Franconi, F.; Campesi, I. Targeted metabolomic profiling in rat tissues reveals sex differences. Sci. Rep. 2018, 8, 4663. [Google Scholar] [CrossRef]
  149. Dias, D.A.; Hill, C.B.; Jayasinghe, N.S.; Atieno, J.; Sutton, T.; Roessner, U. Quantitative profiling of polar primary metabolites of two chickpea cultivars with contrasting responses to salinity. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2015, 1000, 1–13. [Google Scholar] [CrossRef] [Green Version]
  150. Amabebe, E.; Reynolds, S.; Stern, V.L.; Parker, J.L.; Stafford, G.P.; Paley, M.N.; Anumba, D.O. Identifying metabolite markers for preterm birth in cervicovaginal fluid by magnetic resonance spectroscopy. Metabolomics 2016, 12, 67. [Google Scholar] [CrossRef] [Green Version]
  151. Wang, J.; Zhou, L.; Lei, H.; Hao, F.; Liu, X.; Wang, Y.; Tang, H. Simultaneous Quantification of Amino Metabolites in Multiple Metabolic Pathways Using Ultra-High Performance Liquid Chromatography with Tandem-mass Spectrometry. Sci. Rep. 2017, 7, 1423. [Google Scholar] [CrossRef] [Green Version]
  152. Zhou, J.; Liu, C.; Si, D.; Jia, B.; Zhong, L.; Yin, Y. Workflow development for targeted lipidomic quantification using parallel reaction monitoring on a quadrupole-time of flight mass spectrometry. Anal. Chim. Acta 2017, 972, 62–72. [Google Scholar] [CrossRef]
  153. Jedlicka, L.D.L.; Silva, J.D.C.; Balbino, A.M.; Neto, G.B.; Furtado, D.Z.S.; da Silva, H.D.T.; Cavalcanti, F.B.C.; van der Heijden, K.M.; Penatti, C.A.A.; Bechara, E.J.H.; et al. Effects of Diacetyl Flavoring Exposure in Mice Metabolism. Biomed. Res. Int. 2018, 2018, 9875319. [Google Scholar] [CrossRef] [PubMed]
  154. Zukunft, S.; Prehn, C.; Rohring, C.; Moller, G.; Hrabe de Angelis, M.; Adamski, J.; Tokarz, J. High-throughput extraction and quantification method for targeted metabolomics in murine tissues. Metabolomics 2018, 14, 18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Fan, M.; He, T.; Zhang, P.; Cheng, H.; Zhang, J.; Gao, X.; Li, L. Diffusion-weighted imaging features of breast tumours and the surrounding stroma reflect intrinsic heterogeneous characteristics of molecular subtypes in breast cancer. NMR Biomed. 2018, 31, e3869. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Gogiashvili, M.; Horsch, S.; Marchan, R.; Gianmoena, K.; Cadenas, C.; Tanner, B.; Naumann, S.; Ersova, D.; Lippek, F.; Rahnenfuhrer, J.; et al. Impact of intratumoral heterogeneity of breast cancer tissue on quantitative metabolomics using high-resolution magic angle spinning (1) H NMR spectroscopy. NMR Biomed. 2018, 31, e3862. [Google Scholar] [CrossRef]
  157. Kumar, D.; Thakur, K.; Sharma, S.; Kumar, S. NMR for metabolomics studies of Crataegus rhipidophylla Gand. Anal. Bioanal. Chem. 2019, 411, 2149–2159. [Google Scholar] [CrossRef] [PubMed]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Di Minno, A.; Gelzo, M.; Caterino, M.; Costanzo, M.; Ruoppolo, M.; Castaldo, G. Challenges in Metabolomics-Based Tests, Biomarkers Revealed by Metabolomic Analysis, and the Promise of the Application of Metabolomics in Precision Medicine. Int. J. Mol. Sci. 2022, 23, 5213. https://doi.org/10.3390/ijms23095213

AMA Style

Di Minno A, Gelzo M, Caterino M, Costanzo M, Ruoppolo M, Castaldo G. Challenges in Metabolomics-Based Tests, Biomarkers Revealed by Metabolomic Analysis, and the Promise of the Application of Metabolomics in Precision Medicine. International Journal of Molecular Sciences. 2022; 23(9):5213. https://doi.org/10.3390/ijms23095213

Chicago/Turabian Style

Di Minno, Alessandro, Monica Gelzo, Marianna Caterino, Michele Costanzo, Margherita Ruoppolo, and Giuseppe Castaldo. 2022. "Challenges in Metabolomics-Based Tests, Biomarkers Revealed by Metabolomic Analysis, and the Promise of the Application of Metabolomics in Precision Medicine" International Journal of Molecular Sciences 23, no. 9: 5213. https://doi.org/10.3390/ijms23095213

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop