Next Article in Journal
Definitions of Normal Liver Fat and the Association of Insulin Sensitivity with Acquired and Genetic NAFLD—A Systematic Review
Next Article in Special Issue
Metabolomic Studies of Oral Biofilm, Oral Cancer, and Beyond
Previous Article in Journal
Plasma Lipoprotein-Associated Phospholipase A2 Levels Correlated with the Cardio-Ankle Vascular Index in Long-Term Type 2 Diabetes Mellitus Patients
Previous Article in Special Issue
Gut Microbiota and Nonalcoholic Fatty Liver Disease: Insights on Mechanism and Application of Metabolomics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advantages and Pitfalls of Mass Spectrometry Based Metabolome Profiling in Systems Biology

Max Planck Institute for Molecular Genetics, Ihnestraße 63-73, 14195 Berlin, Germany
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2016, 17(5), 632; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms17050632
Submission received: 22 March 2016 / Revised: 19 April 2016 / Accepted: 21 April 2016 / Published: 27 April 2016
(This article belongs to the Special Issue Metabolomic Technologies in Medicine)

Abstract

:
Mass spectrometry-based metabolome profiling became the method of choice in systems biology approaches and aims to enhance biological understanding of complex biological systems. Genomics, transcriptomics, and proteomics are well established technologies and are commonly used by many scientists. In comparison, metabolomics is an emerging field and has not reached such high-throughput, routine and coverage than other omics technologies. Nevertheless, substantial improvements were achieved during the last years. Integrated data derived from multi-omics approaches will provide a deeper understanding of entire biological systems. Metabolome profiling is mainly hampered by its diversity, variation of metabolite concentration by several orders of magnitude and biological data interpretation. Thus, multiple approaches are required to cover most of the metabolites. No software tool is capable of comprehensively translating all the data into a biologically meaningful context yet. In this review, we discuss the advantages of metabolome profiling and main obstacles limiting progress in systems biology.

1. Introduction

Metabolomics offers a framework for analyzing individuals with a specific phenotype at a molecular level in cell biology, personalized medicine, and systems biology [1,2,3]. Metabolites are small molecules participating in metabolic reactions, which are necessary for cellular function, maintenance and growth [4]. Typically, metabolites range from 50 to 1500 Da, while their concentrations span several orders of magnitude. The metabolome is highly dynamic, time-dependent, and metabolites are sensitive to many environmental conditions. It is still unknown how many different metabolite species exist within a cell or organism; predicted are at least 2000 in mammals [5], the Human Metabolome Database (HMDB) [6] currently has more than 8000 verified entries. Plants and bacteria outnumber the diversity of compounds by two orders of magnitude; more than 200,000 metabolites are estimated, for example, to exist in the plant kingdom [7,8]. Metabolites are extensively exchanged with the environment, e.g., food intake, excretion, inhalation, secondary metabolites, such as medications, flavorings, and recreational drugs, which can be further processed by the gut microbiome or organs. Furthermore, metabolites are, chemically, very diverse (polarity, charge, pKa, solubility, volatility, stability, and reactivity), consequently no single method can capture and analyze the entire metabolome at once. Hence, many extraction methods were developed to identify and quantify specific classes of metabolites. The most commonly used approaches to explore the metabolome are gas chromatography (GC), liquid chromatography (LC) or, to a lesser extent, capillary electrophoresis (CE), online coupled to a mass spectrometer (MS), as well as nuclear magnetic resonance (NMR) spectroscopy [9]. Nevertheless, the contextual and chemical diversity makes the detection of a whole set of metabolites present in a biological sample an ambitious goal [10,11].
Systems biology is a novel subject in the field of life science, which aims to understand biological systems as a whole entity, under a variety of conditions [12]. The availability of huge amounts of quantitative data and the development of computational methodologies are the main reasons for the emergence of systems biology [12]. The complete biological model can only be discovered if different levels of gene expression, proteins, and metabolites are considered in an analysis. Multi-omics data can be combined as predictor variables in order to identify models that predict phenotypes and elucidate biomarkers [13].
Metabolites are regarded as the final response of a biological system to environmental changes or gene regulations and, thus, an aberrant metabolism can be linked to disease severity and phenotypes. The metabolome is, thereby, the most predictive of phenotype [1,4]. Metabolic profiling refers to the comprehensive identification and quantification of metabolites in a biological system [14,15]. Systems biology-based metabolome profiling is a new and promising field, which is still puzzled by several aspects, such as sample preparation, identification, standardization, and translating results into biological meaningful data.
Metabolomics and systems biology-driven studies have a great potential for hypotheses generation and dissection of signaling networks in an unbiased or untargeted fashion [16]. The main advantage of systems biology-based metabolomics is the connection of metabolic networks to the underlying reaction pathway structure [7]. Thus, systems biology based metabolomics has a lot of benefits, such as elucidating connectivity maps of pathways, but is still far from covering the entire metabolome.
In this review, we will focus on the advantages and challenges in metabolome profiling for systems biological approaches and cover various techniques and their limitations with a special spotlight on real-time metabolome profiling as a very promising new strategy.

2. Sample Preparation

Typically, a sample preparation protocol contains a solvent extraction step, ultrafiltration, solid-phase extraction and, optionally, a chemical derivatization step [1]. The optimization of sample pretreatment is an important point in method development, since it ensures reproducibility. This review is not focusing on pretreatment strategies and we refer to a variety of other review articles [17,18,19,20].

2.1. Derivatization

Derivatization methods apply reagents, which chemically modify and transform a specific target structure. Thus, derivatization is one of the most effective method to improve the detection characteristics of metabolites in GC- or LC-MS, binding to LC columns, or to stabilize compounds. For example, formaldehyde can be used to label the amine groups through reductive amination. Silylation is a common derivatization reagent in GC-MS, as it transfers a wide range of functional groups, even though some products are not very stable [21].
By derivatization, a sensitivity gain by several orders of magnitude can be achieved. The drawbacks are that an additional derivatization step is introduced and not all metabolites can be derivatized by only one reagent. Furthermore, the mass spectra are different in terms of parent and fragment masses from their endogenous counterparts. Thus, databases for endogenous compounds cannot be applied for reference matching and derivatization-specific databases are rare or have to be built from scratch.
A combination of derivatization and internal standards (ISs), named isotope-coded derivatization (ICD), is a good alternative for relative quantification, in order to improve analytical precision, (reviewed in [22,23]). Thereby, one sample is derivatized with a naturally-labeled reagent, while the other sample is separately derivatized with the isotopically-labeled reagent. After differential labelling, the samples are pooled and processed jointly to avoid any method biases. Thus, this method offers new perspectives for quantitative metabolite profiling, where current protocols can easily comprise over a hundred metabolites [22]. ICD is very similar to the classic stable isotope labeling by amino acids in cell culture (SILAC) approach used in proteomics, where specific amino acids are incorporated into newlysynthesized proteins [24]. Thus, ICD works well for selected compounds or compound groups, such as amino acids, glycans, phosphometabolites, amines, nucleotides, carboxylic- and fatty acids [22].

2.2. Internal Standards

In mass spectrometry, ISs are the method of choice for absolute or relative quantification, calibration curves, recovery rate determination, or to correct for matrix effects. Typically, ISs are isotopically (13C, 15N or deuterium)-labeled counterparts of compounds of interest. The ISs have the same chemical behavior as their endogenous counterparts, do not show drastic chromatographic isotope effects, and have the same retention time. The application of ISs is widespread, but one drawback is that they are not 100% labeled. Thus, an additional and indistinguishable unlabeled peak of typically 1%–2% intensity, depending on the impurity ratio may occur in the mass spectrum. Therefore, the amount of introduced ISs should be reduced to a minimum in order to avoid these interferences. Furthermore, isotopically-labeled ISs are not practical for large profiling approaches, since they are expensive and not available for the majority of compounds or it is impossible like for de novo identifications. To circumvent this bottleneck, only a few ISs, matching to similar compound groups, are selected. Ionization suppression or enhancement is matrix-specific and shows fluctuations over the elution time. This is not always properly reflected by only using a few ISs.
Alternatives for isotopically-labeled ISs are synthetic compounds, which are very similar to the target compound, such as a 17-carbon chain sphingosine instead of the natural 18-carbon chain sphingosine [25]. Label-free approaches, which aim to determine the relative amount of proteins based on precursor signal intensities or on spectral counting, are well established in proteomics [26]. Several methods have been developed to align chromatographic profiles of non-targeted data without the need of ISs for normalization, like applied in XCMS Online [27,28].

3. Data Acquisition and Analysis

3.1. GC-, LC-MS, and NMR in Metabolomics

Most metabolic profiling studies are performed using chromatographic separation online coupled to mass spectrometers, usually gas chromatography (GC)-MS and liquid chromatography (LC)-MS, as well as nuclear magnetic resonance (NMR) spectroscopy [29].
NMR is generally accepted as the gold standard in metabolite structural elucidation, due to its high selectivity, analytical reproducibility, non-destructive nature, and simplicity of sample preparation. However, it suffers from relatively low sensitivity compared to MS [30].
Mass spectrometry-based approaches for metabolic profiling have the advantages of a high sensitivity and selectivity, as well as high throughput and depth of coverage [11]. GC-MS has advantages of a greater chromatographic resolution [31,32,33] when compared to LC-MS-based methods, a good retention of small compounds that elute early with the solvent front in reverse-phase LC-MS methods [34], and large spectral libraries [35]. Unfortunately, the thermal stability of the stationary phase, metabolites and their derivatives limit the metabolome coverage derived by GC-MS [11]. Furthermore, several metabolites can just be analyzed by GC-MS after derivatization and this might introduce variability and produce derivatization artifacts [30].
Due to its high sensitivity and wide range of molecules that can be analyzed, the usage of LC-MS has expanded rapidly over the past ten years [1,29].
Bingol and colleagues introduced a strategy, which combines direct infusion MS with NMR (SUMMIT MS/NMR) for the identification of unknown metabolites in complex mixtures [9]. First, the chemical formulas of compounds in the mixture are identified from accurate masses by MS and, subsequently, all possible structures are generated [9]. Second, NMR spectra of each member of the structural manifold are predicted and compared with the experimental spectra to identify the molecular structure that match the information obtained from MS/NMR [9]. With this method, different types of metabolites, such as amino acids, polyamines, nucleic acids, nucleosides, and carbohydrate conjugates could be identified in an E. coli extract [9]. This approach opens new possibilities in biomedicine, synthetic chemistry, or food sciences for high-throughput identification of unknown metabolites and to overcome limitations of databases [9]. Thus, it is possible to combine different approaches, but due to instrumentation limitations usually just one analytical method is used [36,37].

3.2. Chromatographic Dimension

The metabolome is a collection of compounds with diverse physicochemical properties; no single retention mechanism is adequate to resolve complex sample mixtures that vary widely in their polarity, charge, and stability [1]. Usually, polar and nonpolar metabolites are separated and analyzed in parallel by according analytical columns.
Reversed-phase liquid chromatography (RPLC) is extensively used, because of its reproducible and predictable retention times, wide applicability to various classes of metabolites, and mobile phase compatibility for direct coupling to electrospray ionization mass spectrometry (ESI-MS) [1]. The major disadvantage of RPLC is the poor recovery of hydrophilic compounds. This problem can be solved by the use of hydrophilic interaction chromatography (HILIC), which requires longer re-equilibration and can show retention time drifts [38]. Nevertheless, HILIC is the most commonly used separation mode in LC-MS for polar metabolites that is used together with RPLC for comprehensive metabolomic studies.
The combined use of HILIC and RPLC is an optimal method to increase metabolome coverage, because it is ideal for retaining extremely polar and lipophilic compounds [39]. The main disadvantage is that the applied mobile phases of HILIC and RPLC are incompatible and, thus, two separate runs are necessary, which reduces sample throughput. Furthermore, different chromatographic conditions can cause redundant data analysis and sample preparation is more complicated due to the requirement of different sample buffers [39,40].
To overcome these problems, Haggarty and colleagues applied a method where a sample is injected onto both, RP and HILIC columns in tandem in a single injection [41]. To solve the problem with the solvent strength incompatibility between the columns, a high-aqueous mobile phase at low flow rate for the RPLC column and a high-acetonitrile mobile phase at high flow rate for the HILIC column were used [39].
This tandem column approach was already successfully applied with quadrupole, time of flight (TOF), and orbitrap mass analyzers [39,41,42,43]. This method could significantly increase the amount of identified metabolites in a single analytical run, but one disadvantage is that isomers cannot be separated [39].

3.3. Non-Targeted, Targeted, and Real-Time Metabolome Profiling

Experimental approaches can be classified into non-targeted and targeted metabolomics. Non-targeted metabolomics approaches are widely used for de novo analyte identification, where no prior knowledge about compounds is necessary. In contrast, targeted metabolomics aims to detect a priori selected analytes, on the basis of known parent and fragment masses, with a method termed multiple reaction monitoring (MRM). Once optimized for every transition, the MRM approach has the advantage of selective and sensitive measurement of the selected analytes in complex samples. The specificity is achieved by fragmenting the analyte and by monitoring both parent and one [44,45,46,47,48] or more product ions simultaneously [49,50,51,52,53].
Savolainen and colleagues developed a GC-MS method using a high scan speed (20,000 Da/s), which combines targeted and non-targeted metabolomics into one single accurate, reproducible, and reliable method [34]. This method has the advantage of measuring predefined metabolites relevant to the study question and, additionally, identifies unexpected metabolites [34].
Direct sample injection on high-resolution mass spectrometers is an effective way to maximize analytical throughput, since sampling, sample preparation, and measurements are time-intensive and difficult to automate [54,55]. The applicability of direct injection in metabolomics is extended by advanced instrumentation capable of high-resolution, accurate mass measurements and tandem MS (such as FT-ICR-MS and orbitrap MS) [30]. Flow injection analysis/mass spectrometry (FIA/MS), a method using a HPLC without a column, was applied for metabolite fingerprinting [56], screening in drug discovery [57] and detection of pesticides in food [58]. Matrix effects resulting in inaccurate identification and quantification are the main disadvantages of direct sample injection.
Fuhrer and colleagues established a robust platform for high-throughput, accurate mass, and non-targeted metabolome profiling of E. coli extracts [54]. This method is ideal for initial untargeted metabolome screens where a variety of samples need to be analyzed rapidly [54].
Link and colleagues went a step further and performed real-time metabolome profiling by direct injection of living bacteria, yeast, and mammalian cells into a Q-TOF MS in order to reveal metabolic switches between starvation and growth [55]. They detected 10,466 ions with distinct mass-to-charge ratios, which could be assigned to around 300 metabolites in 15–30 second cycles. The real-time data were confirmed by comparison to a manual sampling and extraction method by flow-injection TOF-MS. By applying the real-time method, more time points were analyzed, more metabolites were identified, and detected ion intensities were higher than those derived from manually-obtained samples [55]. Thus, metabolome profiling of living cells can be monitored in real-time over extended periods to follow the dynamics of metabolic processes. Furthermore, the consequences of any kind of cell manipulation can be studied in real-time, leading to new experimental approaches that could not be performed before.

3.4. Metabolic Flux Analysis

The original method of metabolic flux analysis (MFA) studies systems at a metabolic steady state, by determining metabolite transport rates within and out of tissues, cells, or their sub-compartments, which are then balanced in reaction networks to provide estimates of intracellular fluxes [59]. The limitation of this classical MFA approach is that it cannot resolve fluxes through parallel pathways, such as glycolysis and pentose phosphate pathway, circular pathways or reversible reactions [60]. To overcome the limitation in MFA, isotope 13C tracers, such as the main substrates for mammalian cell culture, glucose and glutamine, can be used [60]. Labeled nutrients in biological systems spread through the network as a function of metabolic activity and produces labeling patterns in the backbone of metabolic intermediates over time [59]. In vivo fluxes are calculated by mathematical models that describe the label propagation from isotopic patterns and extracellular transport rates [59].
Isotope labeled-MFA of co-cultures can be performed, in order to determine inter-species metabolite exchange and population dynamics. This method simplifies experimental and analytical procedures for co-culture flux analysis, improves accuracy, and allows flux elucidation in systems where physical separation of cells or proteins is difficult [61].
Recently, we applied an isotope based flux analysis in order to determine the enzymatic activity of mutated pyrroline-5-carboxylate synthase (P5CS), a protein of the mitochondrial proline cycle. The mutation causes an autosomal-dominant form of cutis laxa with progeroid features. The flux analysis revealed that mutated cells had a reduced P5CS enzymatic activity leading to a delayed proline accumulation [62].
The advantage of fluxomics is that the turnover of virtually any metabolite can be studied by isotope labeling. New pathways or disease-causing genes can be elucidated, a perfect method for specific tasks. Systems biological approaches are mainly hampered by the complexity of the entire system, as most pathways are not straightly leading from A to B, introducing a bias and are, hence, difficult to compute.

3.5. Mass Spectrometry Imaging

Mass spectrometry imaging (MSI) is a spatially-resolved label-free technique, which can directly identify and map the spatial distribution and abundance of known or unknown molecular species in tissues [63]. Metabolites are the direct measure of metabolic activities and can therefore be correlated to phenotypes. Many critical parameters, such as the discrepancy between the low number of detected metabolites versus the real number of possible metabolites, the bias against compound classes and the overlap of compartmentalized metabolic processes in tissue samples, complicate the interpretation of metabolite profiles [7]. As the two or three dimensional distribution of metabolites in specific tissue parts or even subcellular compartments are crucial for many scientific questions, one option to resolve their precise location is MSI. Conventional imaging techniques are not suitable for simultaneous visualization of endogenous molecules. MSI, using matrix-assisted laser desorption/ionization quadrupole ion trap timeofflight (MALDI-QIT-TOF), was used to visualize the spatial distribution of metabolites in tissue sections [64], or plant surface metabolites [65]. Nanospray desorption electrospray ionization (nano-DESI), combined with MS/MS, enabled simultaneous imaging and identification of a large number of metabolites and lipids with a spatial resolution [66]. Three-dimensional secondary ion mass spectrometry (SIMS) imaging was used to investigate the cellular metabolite uptake on the single cell level [67]. Even MS free systems, such as an electrochemical camera chip for simultaneous imaging of multiple metabolites in biofilms was applied [68].
Limited by the absence of a comprehensive MSI method, normalization, and sample preparation techniques suitable for all metabolite classes, we can barely talk about a systems-wide approach, which covers the metabolome. Derivatization enables increased detection of metabolites that are hard to ionize, but may cause analyte suppression or delocalization [69]. Nevertheless, MSI in all its variants is a valuable diagnostic and systems biology tool, capable of providing comprehensive 2D–3D spatial distribution of selected metabolites, like amino acids, lipids, carnitines, or carbohydrates [63,70] in a wide variety of cells and tissues.

3.6. Metabolite Databases

Metabolite databases are repositories of metabolite information, based on experimental confirmation or in silico prediction, and usually contain MS/MS data in order to facilitate metabolite identification. The most established repositories for biological compounds are METLIN, featuring 242,000 entries, 14,000 with high-resolution MS/MS data [71], the Human Metabolome Database (HMDB), with more than 40,000 annotated metabolite entries, including “detected” metabolites (those with measured concentrations or experimental confirmation of their existence) and “expected” metabolites (those for which biochemical pathways are known or human intake/exposure is frequent, but the compound has yet to be detected in the body) [6], and MassBank [72].
Several in silico-generated compound databases are available as well, such as MINE, containing over 571,000 compounds [73], MyCompoundID, using an evidence-based metabolome library (EML) with 375,809 predicted metabolites [74]. Matching metabolite spectra against such huge in silico databases can give rise to too many false-positive identifications just by chance. Based on our experience, databases contain several erroneous entries. Especially, the MS spectra are sometimes of poor quality, they are recorded by different groups with different MS instrumentations, and the provided instrument settings are sometimes very scarce.
All data derived from metabolome profiling studies should be made publically available in according depository databases. Thus, full transparency is guaranteed and datasets can be reused for future identification, quantification, secondary analysis, or systems biology approaches. Data storage, maintenance, and harmonization are critical points, since abandoned or broken depositories would cause a significant loss of knowledge. Currently, just a few metabolomics-specific depositories, such as MetaboLights [75] and PMR [76], are available.

3.7. Metabolite Identification

Due to the complexity and chemical diversity of the metabolome with metabolite concentrations spread over a wide range of magnitudes, identification is still a challenge. In addition to the variety of instruments and their different operation and acquisition methods, even more programs exist for data processing, typically for a specific class of data. For example, different instrument settings used in LC-MS, such as collision energy, resolution, targeted, non-targeted, and fragmentation methods, such as CID (collision induced dissociation), HCD (higher energy collision dissociation), ETD (electron-transfer dissociation), and PQD (pulsed-Q dissociation) [77,78,79,80], can give rise to different fragments species and intensities. Hence, it is a tough task to use or build metabolite-specific spectral databases for compound identification.
One of the best, but time consuming, options is to create an instrument specific MRM method for a set of preselected metabolites. The advantage of using up to three MRMs and their MRM ion ratios, which are stable properties between transitions [81] for robust and unambiguous analysis of complex samples, was demonstrated in a screening approach of rotenone-treated HeLa cells [53]. In this study, a MRM method was developed using pure substances and the transitions had to co-elute at the same retention time with identical peak shapes. MRM ion ratios have been shown to be essential to avoid false-positive identifications in an example for inosine 5′-monophosphate [53]. Schürmann and colleagues revealed the need of a third transition as well [52]. In their study, product ions from a co-eluting interfering matrix compound were consistent with MRMs of two sebuthylazine transitions [52]. Following the European Union directive 2002/657/EC, which regulates the confirmation of suspected positive identifications, this would have resulted in a false-positive finding. To date there is an insufficient discussion about the relative importance of using certain transitions for compound identification [52]. Fragment mis-assignments can originate from impure Q1 isolation [39], co-eluting isobaric compounds, wrong database entries, or incorrect peak picking.
Due to its complex nature, non-targeted metabolomics has to be linked to advanced chemometric techniques, to reduce the data complexity into a smaller set of manageable signals [82]. Analytical methods and required improvements in non-targeted metabolomics are extensively described in a review of Alonso and colleagues [83]. A METLIN search for the exemplary parent mass 136 Da reveals 131 isobaric, but unique metabolites, many of them with very similar structures and, thus, almost similar fragment spectra. As a result, a ranked list based on similarity scores is provided and cutoff values have to be used to verify the identification. It is debatable whether this is sufficient enough and how many false-positive identifications are actually included. A big effort is being done to improve spectral databases, but the development of accurate automatic identification algorithms is still subject to the availability of an exhaustive set of reference metabolite spectra [83].
As target identification is one of the most critical steps, in silico target identification methods, including chemical similarity database searches, are used, such as CSNAP (Chemical Similarity Network Analysis Pulldown) [84]. Several strategies exist on how unknown peaks can be deciphered and interpreted, but validation guidelines were missing for a long time. A collection of guidelines/minimum requirements for the validation of metabolite identification, were finally conceived by the Metabolomics Standards Initiative (MSI in 2005, in order to allow data to be efficiently applied, shared and reused [85]. The initiative “COordination of Standards in MetabOlomicS” (COSMOS) is generating robust data infrastructures and exchange standards for metabolomics data and metadata [86].

4. Biological Interpretation of Results

At this point, a large knowledge gap exists in the translation from changes in metabolite concentration in body fluids to organ biochemistry and (molecular) physiological interpretation [20]. Existing scarce information are usually only available for specific species, organs, or body fluids and cannot be transferred from one another easily. Software tools, such as pathway or enrichment analysis are not taking that into account.
What does it actually mean, if metabolite X from a distinct pathway is found to be decreased? On one hand, downstream enzymes could have an increased activity or, on the other hand, the enzyme producing metabolite X could have a decreased activity; both can lead to a decreased level of metabolite X. Enzyme activities can be influenced by their specific product by steric inhibition/feedback-inhibition or activation, and thereby guarantee a balanced homeostasis in the cell. This strategy avoids large metabolic changes with a potential negative effect for the cell.
Many metabolites play a role in several pathways and are the product or substrate of many different enzymes or processes. Thus, it is a challenge to pinpoint an altered metabolite to a specific pathway or enzyme. Nevertheless, the pathway information may already give the correct answer, or at least a hint, for a biological question. Changes in metabolite abundances can be mapped to specific pathways, thereby providing mechanistic information of the process under study. Data derived from metabolic profiling can be complemented by genome, proteome, clinical, and environmental data, which supports the discovery of potential biomarkers that would not have been identified with targeted studies alone [87].
Metabolome profiles have the advantage to detect unknown compounds (shotgun) and alterations on a global scale (shotgun and targeted). As many metabolites usually originate from several different processes, specific methods have to be used for validation, like a knockdown of the according enzyme or isotope-labeled MFA. Many bioinformatics tools are continuously created to address several important questions in the metabolomic field, like interpreting profiling data. MetaboAnalyst (www.metaboanalyst.ca), permits a comprehensive metabolomic data analysis, visualization and interpretation, including complex statistical calculations [88]. Metabolite pathway enrichment analysis (MPEA) was designed for the visualization and biological interpretation of metabolic profiling data at the system level. The tool tests whether metabolites involved in some predefined pathways occur towards the top or bottom of a ranked query compound list [89]. Integrated Molecular Pathway Level Analysis (IMPaLA) [90] is a tool for pathway over-representation and enrichment analysis with expression- and metabolite data. This proves the importance of metabolomics, but on the downside, many of these programs get funding for only a few years and are abandoned thereafter and egress, or even become useless, as file formats and necessary accompanying software changes permanently. Furthermore, most of the programs require specific raw or input data, which are frequently not interchangeable between programs.

Biomarker Discovery

Biomarker discovery is driven by applying new instrumentation, protocols, and software tools, in order to find novel and specific key metabolic features, which are characteristic for specific pathological conditions, diseases, or cancer. Surprisingly, the clinical breakthrough is still out of sight. Nevertheless, metabolomic key features indicative for diseases such as depression [91], schizophrenia [92,93], cardiovascular and coronary artery disease [94], diabetes [95,96], and cancers, such as liver [97], ovarian [98], and breast cancers have been reported [99]. The metabolic level variation between people, as well as within tissues and time points, is huge and dynamic. The genome, epigenome, transcriptome, and proteome states are much more stable compared to the high fluctuating metabolites. The aim in the biomarker field is to find biomarkers, which can precisely detect an early malignancy in order to achieve the best treatment effects and finally the highest survival rates for patients. So far, there were no new approved biomarkers in recent years [100,101]. The current strategy to find single biomarkers for a disease is hampered by high and dynamic fluctuations of metabolites. As every disease not only changes one metabolite, but entire metabolic pathways, we probably should search for differentially regulated pathways or metabolite classes to be the more robust biomarkers in the future.
This can be accomplished the best by an integrative approach taking many omics subdisciplines into account [102]. Thus, accurate multi-data analyses will be the key to reveal, assess, and track molecular patterns, which reflect disease-perturbed networks [102,103].

5. Conclusions

The metabolome, lying closest to the phenotype and the most predictive of phenotype, is of emerging interest for systems biologists. Metabolomics has a long lasting importance in the field of biomarker detection and, finally, for drug treatments. No universal instrument or method exists yet, which is capable of measuring the entire metabolome at once. Currently, we are still in the status of metabolomic profiling, intensive parallelization of different approaches and methods are transforming metabolomic profiling steadily into metabolome profiling. For a holistic view, integration of genome, transcriptome, proteome, and/or metabolome datasets can significantly enhance the insight into biological questions and molecular interactions between these disciplines. The integration of different molecular profiling data into a comprehensive entity is still challenging, especially in the bioinformatics area. Knowledge about the function of metabolites and, specifically, their multiple biological interactions are still missing. Thus, interpretation of results has to be handled with care. Our review highlights the advantages of metabolome profiles for systems biology approaches, as well as the current pitfalls, such as data interpretation and translation of results into a biologically-useful meaning.

Acknowledgments

Our work is supported by the Max Planck Society.

Author Contributions

The manuscript was written through contributions of Ina Aretz and David Meierhofer. Both authors have given approval to the final version of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kuehnbaum, N.L.; Britz-McKibbin, P. New advances in separation science for metabolomics: Resolving chemical diversity in a post-genomic era. Chem. Rev. 2013, 113, 2437–2468. [Google Scholar] [CrossRef] [PubMed]
  2. Johnson, C.H.; Gonzalez, F.J. Challenges and opportunities of metabolomics. J. Cell. Physiol. 2012, 227, 2975–2981. [Google Scholar] [CrossRef] [PubMed]
  3. Patti, G.J.; Yanes, O.; Siuzdak, G. Innovation: Metabolomics: The apogee of the omics trilogy. Nat. Rev. Mol. Cell Biol. 2012, 13, 263–269. [Google Scholar] [CrossRef] [PubMed]
  4. Dettmer, K.; Aronov, P.A.; Hammock, B.D. Mass spectrometry-based metabolomics. Mass Spectrom. Rev. 2007, 26, 51–78. [Google Scholar] [CrossRef] [PubMed]
  5. Harrigan, G.G.; Goodacre, R. (Eds.) Metabolic Profiling: Its Role in Biomarker Discovery and Gene Function Analysis; Springer US: Boston, MA, USA, 2003.
  6. Wishart, D.S.; Jewison, T.; Guo, A.C.; Wilson, M.; Knox, C.; Liu, Y.; Djoumbou, Y.; Mandal, R.; Aziat, F.; Dong, E.; et al. HMDB 3.0—The human metabolome database in 2013. Nucleic Acids Res. 2013, 41, D801–D807. [Google Scholar] [CrossRef] [PubMed]
  7. Weckwerth, W. Metabolomics in systems biology. Annu. Rev. Plant Biol. 2003, 54, 669–689. [Google Scholar] [CrossRef] [PubMed]
  8. Saito, K.; Matsuda, F. Metabolomics for functional genomics, systems biology, and biotechnology. Annu. Rev. Plant Biol. 2010, 61, 463–489. [Google Scholar] [CrossRef] [PubMed]
  9. Bingol, K.; Bruschweiler-Li, L.; Yu, C.; Somogyi, A.; Zhang, F.; Brüschweiler, R. Metabolomics beyond spectroscopic databases: A combined MS/NMR strategy for the rapid identification of new metabolites in complex mixtures. Anal. Chem. 2015, 87, 3864–3870. [Google Scholar] [CrossRef] [PubMed]
  10. Veuthey, J.-L.; Guillarme, D.; Schoenmakers, P.J.; Forcisi, S.; Moritz, F.; Kanawati, B.; Tziotis, D.; Lehmann, R.; Schmitt-Kopplin, P. Liquid chromatography–mass spectrometry in metabolomics research: Mass analyzers in ultra high pressure liquid chromatography coupling. J. Chromatogr. A 2013, 1292, 51–65. [Google Scholar]
  11. Wang, Y.; Liu, S.; Hu, Y.; Li, P.; Wan, J.-B. Current state of the art of mass spectrometry-based metabolomics studies—A review focusing on wide coverage, high throughput and easy identification. RSC Adv. 2015, 5, 78728–78737. [Google Scholar] [CrossRef]
  12. Lee, S. Systems biology—A pivotal research methodology for understanding the mechanisms of traditional medicine. J. Pharmacopunct. 2015, 18, 11–18. [Google Scholar] [CrossRef] [PubMed]
  13. Ritchie, M.D.; Holzinger, E.R.; Li, R.; Pendergrass, S.A.; Kim, D. Methods of integrating data to uncover genotype-phenotype interactions. Nat. Rev. Genet. 2015, 16, 85–97. [Google Scholar] [CrossRef] [PubMed]
  14. Clarke, C.J.; Haselden, J.N. Metabolic profiling as a tool for understanding mechanisms of toxicity. Toxicol. Pathol. 2008, 36, 140–147. [Google Scholar] [CrossRef] [PubMed]
  15. Yu, T.; Bai, Y. Analyzing LC/MS metabolic profiling data in the context of existing metabolic networks. Curr. Metab. 2013, 1, 83–91. [Google Scholar]
  16. Caldana, C.; Fernie, A.R.; Willmitzer, L.; Steinhauser, D. Unraveling retrograde signaling pathways: Finding candidate signaling molecules via metabolomics and systems biology driven approaches. Front. Plant Sci. 2012, 3, 267. [Google Scholar] [CrossRef] [PubMed]
  17. Hounoum, B.M.; Blasco, H.; Emond, P.; Mavel, S. Liquid chromatography-high resolution mass spectrometry-based cell metabolomics: Experimental design, recommendations and applications. TrAC Trends Anal. Chem. 2015, 75, 118–128. [Google Scholar] [CrossRef]
  18. Fumes, B.H.; Silva, M.R.; Andrade, F.N.; Nazario, C.E.D.; Lanças, F.M. Recent advances and future trends in new materials for sample preparation. TrAC Trends Anal. Chem. 2015, 71, 9–25. [Google Scholar] [CrossRef]
  19. Raterink, R.-J.; Lindenburg, P.W.; Vreeken, R.J.; Ramautar, R.; Hankemeier, T. Recent developments in sample-pretreatment techniques for mass spectrometry-based metabolomics. TrAC Trends Anal. Chem. 2014, 61, 157–167. [Google Scholar] [CrossRef]
  20. Scalbert, A.; Brennan, L.; Fiehn, O.; Hankemeier, T.; Kristal, B.S.; van Ommen, B.; Pujos-Guillot, E.; Verheij, E.; Wishart, D.; Wopereis, S. Mass-spectrometry-based metabolomics: Limitations and recommendations for future progress with particular focus on nutrition research. Metabolomics 2009, 5, 435–458. [Google Scholar] [CrossRef] [PubMed]
  21. Tallent, W.H.; Kleiman, R.; Cope, D.G. Gas—Liquid chromatography of silylated mixtures from lipolysis of triglycerides containing unusual fatty acyl groups. J. Lipid Res. 1966, 7, 531–535. [Google Scholar] [PubMed]
  22. Bruheim, P.; Kvitvang, H.F.N.; Villas-Boas, S.G. Stable isotope coded derivatizing reagents as internal standards in metabolite profiling. J. Chromatogr. A 2013, 1296, 196–203. [Google Scholar] [CrossRef] [PubMed]
  23. Toyo’oka, T. LC-MS determination of bioactive molecules based upon stable isotope-coded derivatization method. J. Pharm. Biomed. Anal. 2012, 69, 174–184. [Google Scholar] [CrossRef] [PubMed]
  24. Ong, S.-E.; Blagoev, B.; Kratchmarova, I.; Kristensen, D.B.; Steen, H.; Pandey, A.; Mann, M. Stable isotope labeling by amino acids in cell culture, SILAC, as a simple and accurate approach to expression proteomics. Mol. Cell. Proteom. 2002, 1, 376–386. [Google Scholar] [CrossRef]
  25. Lieser, B.; Liebisch, G.; Drobnik, W.; Schmitz, G. Quantification of sphingosine and sphinganine from crude lipid extracts by HPLC electrospray ionization tandem mass spectrometry. J. Lipid Res. 2003, 44, 2209–2216. [Google Scholar] [CrossRef] [PubMed]
  26. Cox, J.; Hein, M.Y.; Luber, C.A.; Paron, I.; Nagaraj, N.; Mann, M. MaxLFQ allows accurate proteome-wide label-free quantification by delayed normalization and maximal peptide ratio extraction. Mol. Cell. Proteom. 2014, 13, 2513–2526. [Google Scholar] [CrossRef] [PubMed]
  27. Smith, C.A.; Want, E.J.; O’Maille, G.; Abagyan, R.; Siuzdak, G. XCMS: Processing mass spectrometry data for metabolite profiling using nonlinear peak alignment, matching, and identification. Anal. Chem. 2006, 78, 779–787. [Google Scholar] [CrossRef] [PubMed]
  28. Tautenhahn, R.; Patti, G.J.; Rinehart, D.; Siuzdak, G. XCMS Online: A web-based platform to process untargeted metabolomic data. Anal. Chem. 2012, 84, 5035–5039. [Google Scholar] [CrossRef]
  29. Contrepois, K.; Jiang, L.; Snyder, M. optimized analytical procedures for the untargeted metabolomic profiling of human urine and plasma by combining hydrophilic interaction (hilic) and reverse-phase liquid chromatography (RPLC)–Mass spectrometry. Mol. Cell. Proteom. 2015, 14, 1684–1695. [Google Scholar] [CrossRef] [PubMed]
  30. Lei, Z.; Huhman, D.V.; Sumner, L.W. Mass spectrometry strategies in metabolomics. J. Biol. Chem. 2011, 286, 25435–25442. [Google Scholar] [CrossRef] [PubMed]
  31. Chan, E.C.Y.; Pasikanti, K.K.; Hong, Y.; Ho, P.C.; Mahendran, R.; Raman Nee Mani, L.; Chiong, E.; Esuvaranathan, K. Metabonomic profiling of bladder cancer. J. Proteome Res. 2015, 14, 587–602. [Google Scholar] [CrossRef] [PubMed]
  32. Pasikanti, K.K.; Ho, P.C.; Chan, E.C.Y. Gas chromatography/mass spectrometry in metabolic profiling of biological fluids. J. Chromatogr. B. 2008, 871, 202–211. [Google Scholar] [CrossRef] [PubMed]
  33. Koek, M.M.; Muilwijk, B.; van der Werf, M.J.; Hankemeier, T. Microbial metabolomics with gas chromatography/mass spectrometry. Anal. Chem. 2006, 78, 1272–1281. [Google Scholar] [CrossRef] [PubMed]
  34. Savolainen, O.I.; Sandberg, A.-S.; Ross, A.B. A Simultaneous metabolic profiling and quantitative multimetabolite metabolomic method for human plasma using gas-chromatography tandem mass spectrometry. J. Proteome Res. 2016, 15, 259–265. [Google Scholar] [CrossRef] [PubMed]
  35. Want, E.J.; Nordström, A.; Morita, H.; Siuzdak, G. From exogenous to endogenous: The inevitable imprint of mass spectrometry in metabolomics. J. Proteome Res. 2007, 6, 459–468. [Google Scholar] [CrossRef] [PubMed]
  36. Psychogios, N.; Hau, D.D.; Peng, J.; Guo, A.C.; Mandal, R.; Bouatra, S.; Sinelnikov, I.; Krishnamurthy, R.; Eisner, R.; Gautam, B.; et al. The human serum metabolome. PLoS ONE 2011, 6, e16957. [Google Scholar] [CrossRef] [PubMed]
  37. Büscher, J.M.; Czernik, D.; Ewald, J.C.; Sauer, U.; Zamboni, N. Cross-platform comparison of methods for quantitative metabolomics of primary metabolism. Anal. Chem. 2009, 81, 2135–2143. [Google Scholar] [CrossRef] [PubMed]
  38. Gray, N.; Heaton, J.; Musenga, A.; Cowan, D.A.; Plumb, R.S.; Smith, N.W. Comparison of reversed-phase and hydrophilic interaction liquid chromatography for the quantification of ephedrines using medium-resolution accurate mass spectrometry. J. Chromatogr. A 2013, 1289, 37–46. [Google Scholar] [CrossRef] [PubMed]
  39. Yan, Z.; Yan, R. Increase the accessibility and scale of targeted metabolomics: Construction of a human urinary metabolome-wide multiple reaction monitoring library using directly-coupled reversed-phase and hydrophilic interaction chromatography. Anal. Chim. Acta 2015, 894, 65–75. [Google Scholar] [CrossRef] [PubMed]
  40. Hemström, P.; Irgum, K. Hydrophilic interaction chromatography. J. Sep. Sci. 2006, 29, 1784–1821. [Google Scholar] [CrossRef] [PubMed]
  41. Haggarty, J.; Oppermann, M.; Dalby, M.J.; Burchmore, R.J.; Cook, K.; Weidt, S.; Burgess, K.E.V. Serially coupling hydrophobic interaction and reversed-phase chromatography with simultaneous gradients provides greater coverage of the metabolome. Metabolomics 2015, 11, 1465–1470. [Google Scholar] [CrossRef] [PubMed]
  42. Falasca, S.; Petruzziello, F.; Kretz, R.; Rainer, G.; Zhang, X. Analysis of multiple quaternary ammonium compounds in the brain using tandem capillary column separation and high resolution mass spectrometric detection. J. Chromatogr. A 2012, 1241, 46–51. [Google Scholar] [CrossRef] [PubMed]
  43. Greco, G.; Grosse, S.; Letzel, T. Serial coupling of reversed-phase and zwitterionic hydrophilic interaction LC/MS for the analysis of polar and nonpolar phenols in wine. J. Sep. Sci. 2013, 36, 1379–1388. [Google Scholar] [CrossRef] [PubMed]
  44. Wei, R.; Li, G.; Seymour, A.B. High-throughput and multiplexed LC/MS/MRM method for targeted metabolomics. Anal. Chem. 2010, 82, 5527–5533. [Google Scholar] [CrossRef] [PubMed]
  45. Lu, W.; Clasquin, M.F.; Melamud, E.; Amador-Noguez, D.; Caudy, A.A.; Rabinowitz, J.D. Metabolomic analysis via reversed-phase ion-pairing liquid chromatography coupled to a stand alone orbitrap mass spectrometer. Anal. Chem. 2010, 82, 3212–3221. [Google Scholar] [CrossRef] [PubMed]
  46. Buré, C.; Ayciriex, S.; Testet, E.; Schmitter, J.-M. A single run LC-MS/MS method for phospholipidomics. Anal. Bioanal. Chem. 2013, 405, 203–213. [Google Scholar] [CrossRef] [PubMed]
  47. Bielawski, J.; Pierce, J.S.; Snider, J.; Rembiesa, B.; Szulc, Z.M.; Bielawska, A. Comprehensive quantitative analysis of bioactive sphingolipids by high-performance liquid chromatography-tandem mass spectrometry. Methods Mol. Biol. 2009, 579, 443–467. [Google Scholar] [PubMed]
  48. Preinerstorfer, B.; Schiesel, S.; Lämmerhofer, M.; Lindner, W. Metabolic profiling of intracellular metabolites in fermentation broths from β-lactam antibiotics production by liquid chromatography-tandem mass spectrometry methods. J. Chromatogr. A 2010, 1217, 312–328. [Google Scholar] [CrossRef] [PubMed]
  49. Kitteringham, N.R.; Jenkins, R.E.; Lane, C.S.; Elliott, V.L.; Park, B.K. Multiple reaction monitoring for quantitative biomarker analysis in proteomics and metabolomics. J. Chromatogr. B 2009, 877, 1229–1239. [Google Scholar] [CrossRef] [PubMed]
  50. Bueno, M.J.M.; Agüera, A.; Gómez, M.J.; Hernando, M.D.; García-Reyes, J.F.; Fernandez-Alba, A.R. Application of liquid chromatography/quadrupole-linear Ion trap mass spectrometry and time-of-flight mass spectrometry to the determination of pharmaceuticals and related contaminants in wastewater. Anal. Chem. 2007, 79, 9372–9384. [Google Scholar] [CrossRef] [PubMed]
  51. Gros, M.; Petrović, M.; Barceló, D. Tracing pharmaceutical residues of different therapeutic classes in environmental waters by using liquid chromatography/quadrupole-linear ion trap mass spectrometry and automated library searching. Anal. Chem. 2009, 81, 898–912. [Google Scholar] [CrossRef] [PubMed]
  52. Schürmann, A.; Dvorak, V.; Crüzer, C.; Butcher, P.; Kaufmann, A. False-positive liquid chromatography/tandem mass spectrometric confirmation of sebuthylazine residues using the identification points system according to EU directive 2002/657/EC due to a biogenic insecticide in tarragon. Rapid Commun. Mass Spectrom. 2009, 23, 1196–1200. [Google Scholar] [CrossRef] [PubMed]
  53. Gielisch, I.; Meierhofer, D. Metabolome and proteome profiling of complex I deficiency induced by rotenone. J. Proteome Res. 2015, 14, 224–235. [Google Scholar] [CrossRef] [PubMed]
  54. Fuhrer, T.; Heer, D.; Begemann, B.; Zamboni, N. High-throughput, accurate mass metabolome profiling of cellular extracts by flow injection-time-of-flight mass spectrometry. Anal. Chem. 2011, 83, 7074–7080. [Google Scholar] [CrossRef] [PubMed]
  55. Link, H.; Fuhrer, T.; Gerosa, L.; Zamboni, N.; Sauer, U. Real-time metabolome profiling of the metabolic switch between starvation and growth. Nat. Methods 2015, 12, 1091–1097. [Google Scholar] [CrossRef] [PubMed]
  56. Beckmann, M.; Parker, D.; Enot, D.P.; Duval, E.; Draper, J. High-throughput, nontargeted metabolite fingerprinting using nominal mass flow injection electrospray mass spectrometry. Nat. Protoc. 2008, 3, 486–504. [Google Scholar] [CrossRef] [PubMed]
  57. Roddy, T.P.; Horvath, C.R.; Stout, S.J.; Kenney, K.L.; Ho, P.-I.; Zhang, J.-H.; Vickers, C.; Kaushik, V.; Hubbard, B.; Wang, Y.K. Mass spectrometric techniques for label-free high-throughput screening in drug discovery. Anal. Chem. 2007, 79, 8207–8213. [Google Scholar] [CrossRef] [PubMed]
  58. Nanita, S.C.; Stry, J.J.; Pentz, A.M.; McClory, J.P.; May, J.H. Fast extraction and dilution flow injection mass spectrometry method for quantitative chemical residue screening in food. J. Agric. Food Chem. 2011, 59, 7557–7568. [Google Scholar] [CrossRef] [PubMed]
  59. Sá, J.; Duarte, T.; Carrondo, M.; Alves, P.; Teixeira, A. Metabolic Flux Analysis: A Powerful Tool in Animal Cell Culture; Al-Rubeai, M., Ed.; Springer International Publishing: Cham, Switzerland, 2015; Volume 9. [Google Scholar]
  60. Niklas, J.; Heinzle, E. Metabolic flux analysis in systems biology of mammalian cells. Adv. Biochem. Eng. Biotechnol. 2012, 127, 109–132. [Google Scholar] [PubMed]
  61. Gebreselassie, N.A.; Antoniewicz, M.R. 13C-metabolic flux analysis of co-cultures: A novel approach. Metab. Eng. 2015, 31, 132–139. [Google Scholar] [CrossRef] [PubMed]
  62. Fischer-Zirnsak, B.; Escande-Beillard, N.; Ganesh, J.; Tan, Y.X.; Al Bughaili, M.; Lin, A.E.; Sahai, I.; Bahena, P.; Reichert, S.L.; Loh, A.; et al. Recurrent de novo mutations affecting residue arg138 of pyrroline-5-carboxylate synthase cause a progeroid form of autosomal-dominant cutis laxa. Am. J. Hum. Genet. 2015, 97, 483–492. [Google Scholar] [CrossRef] [PubMed]
  63. Li, T.; He, J.; Mao, X.; Bi, Y.; Luo, Z.; Guo, C.; Tang, F.; Xu, X.; Wang, X.; Wang, M.; et al. In situ biomarker discovery and label-free molecular histopathological diagnosis of lung cancer by ambient mass spectrometry imaging. Sci. Rep. 2015, 5, 14089. [Google Scholar] [CrossRef] [PubMed]
  64. Zaima, N.; Hayasaka, T.; Goto-Inoue, N.; Setou, M. Matrix-assisted laser desorption/ionization imaging mass spectrometry. Int. J. Mol. Sci. 2010, 11, 5040–5055. [Google Scholar] [CrossRef] [PubMed]
  65. Shroff, R.; Schramm, K.; Jeschke, V.; Nemes, P.; Vertes, A.; Gershenzon, J.; Svatoš, A. Quantification of plant surface metabolites by matrix-assisted laser desorption-ionization mass spectrometry imaging: Glucosinolates on Arabidopsis thaliana leaves. Plant J. 2015, 81, 961–972. [Google Scholar] [CrossRef] [PubMed]
  66. Lanekoff, I.; Burnum-Johnson, K.; Thomas, M.; Short, J.; Carson, J.P.; Cha, J.; Dey, S.K.; Yang, P.; Prieto Conaway, M.C.; Laskin, J. High-speed tandem mass spectrometric in situ imaging by nanospray desorption electrospray ionization mass spectrometry. Anal. Chem. 2013, 85, 9596–9603. [Google Scholar] [CrossRef] [PubMed]
  67. Passarelli, M.K.; Newman, C.F.; Marshall, P.S.; West, A.; Gilmore, I.S.; Bunch, J.; Alexander, M.R.; Dollery, C.T. Single-cell analysis: Visualizing pharmaceutical and metabolite uptake in cells with label-free 3D mass spectrometry imaging. Anal. Chem. 2015, 87, 6696–6702. [Google Scholar] [CrossRef] [PubMed]
  68. Bellin, D.L.; Sakhtah, H.; Zhang, Y.; Price-Whelan, A.; Dietrich, L.E.P.; Shepard, K.L. Electrochemical camera chip for simultaneous imaging of multiple metabolites in biofilms. Nat. Commun. 2016, 7, 10535. [Google Scholar] [CrossRef] [PubMed]
  69. Goodwin, R.J.A. Sample preparation for mass spectrometry imaging: Small mistakes can lead to big consequences. J. Proteom. 2012, 75, 4893–4911. [Google Scholar] [CrossRef] [PubMed]
  70. Harvey, D.J. Analysis of carbohydrates and glycoconjugates by matrix-assisted laser desorption/ionization mass spectrometry: An update for 2007–2008. Mass Spectrom. Rev. 2012, 31, 183–311. [Google Scholar] [CrossRef] [PubMed]
  71. Smith, C.A.; O’Maille, G.; Want, E.J.; Qin, C.; Trauger, S.A.; Brandon, T.R.; Custodio, D.E.; Abagyan, R.; Siuzdak, G. METLIN: A metabolite mass spectral database. Ther. Drug Monit. 2005, 27, 747–751. [Google Scholar] [CrossRef] [PubMed]
  72. Horai, H.; Arita, M.; Kanaya, S.; Nihei, Y.; Ikeda, T.; Suwa, K.; Ojima, Y.; Tanaka, K.; Tanaka, S.; Aoshima, K.; et al. MassBank: A public repository for sharing mass spectral data for life sciences. J. Mass Spectrom. 2010, 45, 703–714. [Google Scholar] [CrossRef] [PubMed]
  73. Jeffryes, J.G.; Colastani, R.L.; Elbadawi-Sidhu, M.; Kind, T.; Niehaus, T.D.; Broadbelt, L.J.; Hanson, A.D.; Fiehn, O.; Tyo, K.E.J.; Henry, C.S. MINEs: Open access databases of computationally predicted enzyme promiscuity products for untargeted metabolomics. J. Cheminform. 2015, 7, 44. [Google Scholar] [CrossRef] [PubMed]
  74. Huan, T.; Tang, C.; Li, R.; Shi, Y.; Lin, G.; Li, L. MyCompoundID MS/MS Search: Metabolite identification using a library of predicted fragment-ion-spectra of 383,830 possible human metabolites. Anal. Chem. 2015, 87, 10619–10626. [Google Scholar] [CrossRef] [PubMed]
  75. Haug, K.; Salek, R.M.; Conesa, P.; Hastings, J.; de Matos, P.; Rijnbeek, M.; Mahendraker, T.; Williams, M.; Neumann, S.; Rocca-Serra, P.; et al. MetaboLights—An open-access general-purpose repository for metabolomics studies and associated meta-data. Nucleic Acids Res. 2013, 41, D781–D786. [Google Scholar] [CrossRef] [PubMed]
  76. Hur, M.; Campbell, A.A.; Almeida-de-Macedo, M.; Li, L.; Ransom, N.; Jose, A.; Crispin, M.; Nikolau, B.J.; Wurtele, E.S. A global approach to analysis and interpretation of metabolic data for plant natural product discovery. Nat. Prod. Rep. 2013, 30, 565–583. [Google Scholar] [CrossRef] [PubMed]
  77. Ichou, F.; Schwarzenberg, A.; Lesage, D.; Alves, S.; Junot, C.; Machuron-Mandard, X.; Tabet, J.-C. Comparison of the activation time effects and the internal energy distributions for the CID, PQD and HCD excitation modes. J. Mass Spectrom. 2014, 49, 498–508. [Google Scholar] [CrossRef] [PubMed]
  78. Sleno, L.; Volmer, D.A. Ion activation methods for tandem mass spectrometry. J. Mass Spectrom. 2004, 39, 1091–1112. [Google Scholar] [CrossRef] [PubMed]
  79. Oppermann, M.; Damoc, N.E.; Crone, C.; Moehring, T.; Muenster, H.; Hornshaw, M. High precision measurement and fragmentation analysis for metabolite identification. Methods Mol. Biol. 2012, 860, 145–156. [Google Scholar] [PubMed]
  80. Donohoe, G.C.; Maleki, H.; Arndt, J.R.; Khakinejad, M.; Yi, J.; McBride, C.; Nurkiewicz, T.R.; Valentine, S.J. A new ion mobility-linear ion trap instrument for complex mixture analysis. Anal. Chem. 2014, 86, 8121–8128. [Google Scholar] [CrossRef] [PubMed]
  81. Kushnir, M.M.; Rockwood, A.L.; Nelson, G.J.; Yue, B.; Urry, F.M. Assessing analytical specificity in quantitative analysis using tandem mass spectrometry. Clin. Biochem. 2005, 38, 319–327. [Google Scholar] [CrossRef] [PubMed]
  82. Roberts, L.D.; Souza, A.L.; Gerszten, R.E.; Clish, C.B. Targeted metabolomics. Curr. Protoc. Mol. Biol. 2012. [Google Scholar] [CrossRef]
  83. Alonso, A.; Marsal, S.; Julià, A. Analytical methods in untargeted metabolomics: State of the art in 2015. Front. Bioeng. Biotechnol. 2015, 3, 23. [Google Scholar] [CrossRef] [PubMed]
  84. Lo, Y.-C.; Senese, S.; Li, C.-M.; Hu, Q.; Huang, Y.; Damoiseaux, R.; Torres, J.Z. Large-scale chemical similarity networks for target profiling of compounds identified in cell-based chemical screens. PLoS Comput. Biol. 2015, 11, e1004153. [Google Scholar] [CrossRef] [PubMed]
  85. Sumner, L.W.; Amberg, A.; Barrett, D.; Beale, M.H.; Beger, R.; Daykin, C.A.; Fan, T.W.-M.; 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] [PubMed]
  86. Salek, R.M.; Neumann, S.; Schober, D.; Hummel, J.; Billiau, K.; Kopka, J.; Correa, E.; Reijmers, T.; Rosato, A.; Tenori, L.; et al. COordination of Standards in MetabOlomicS (COSMOS): Facilitating integrated metabolomics data access. Metabolomics 2015, 11, 1587–1597. [Google Scholar] [CrossRef] [PubMed]
  87. Weckwerth, W.; Morgenthal, K. Metabolomics: From pattern recognition to biological interpretation. Drug Discov. Today 2005, 10, 1551–1558. [Google Scholar] [CrossRef]
  88. Xia, J.; Sinelnikov, I.V.; Han, B.; Wishart, D.S. MetaboAnalyst 3.0—Making metabolomics more meaningful. Nucleic Acids Res. 2015, 43, W251–W257. [Google Scholar] [CrossRef] [PubMed]
  89. Kankainen, M.; Gopalacharyulu, P.; Holm, L.; Oresic, M. MPEA—Metabolite pathway enrichment analysis. Bioinformatics 2011, 27, 1878–1879. [Google Scholar] [CrossRef] [PubMed]
  90. Kamburov, A.; Cavill, R.; Ebbels, T.M.D.; Herwig, R.; Keun, H.C. Integrated pathway-level analysis of transcriptomics and metabolomics data with IMPaLA. Bioinformatics 2011, 27, 2917–2918. [Google Scholar] [CrossRef] [PubMed]
  91. Paige, L.A.; Mitchell, M.W.; Krishnan, K.R.R.; Kaddurah-Daouk, R.; Steffens, D.C. A preliminary metabolomic analysis of older adults with and without depression. Int. J. Geriatr. Psychiatry 2007, 22, 418–423. [Google Scholar] [CrossRef] [PubMed]
  92. Holmes, E.; Tsang, T.M.; Huang, J.T.-J.; Leweke, F.M.; Koethe, D.; Gerth, C.W.; Nolden, B.M.; Gross, S.; Schreiber, D.; Nicholson, J.K.; et al. Metabolic profiling of CSF: Evidence that early intervention may impact on disease progression and outcome in schizophrenia. PLoS Med. 2006, 3, e327. [Google Scholar] [CrossRef] [PubMed]
  93. Kaddurah-Daouk, R. Metabolic profiling of patients with schizophrenia. PLoS Med. 2006, 3, e363. [Google Scholar] [CrossRef] [PubMed]
  94. Brindle, J.T.; Antti, H.; Holmes, E.; Tranter, G.; Nicholson, J.K.; Bethell, H.W.L.; Clarke, S.; Schofield, P.M.; McKilligin, E.; Mosedale, D.E.; et al. Rapid and noninvasive diagnosis of the presence and severity of coronary heart disease using 1H–NMR-based metabonomics. Nat. Med. 2002, 8, 1439–1444. [Google Scholar] [CrossRef] [PubMed]
  95. Yi, L.-Z.; He, J.; Liang, Y.-Z.; Yuan, D.-L.; Chau, F.-T. Plasma fatty acid metabolic profiling and biomarkers of type 2 diabetes mellitus based on GC/MS and PLS-LDA. FEBS Lett. 2006, 580, 6837–6845. [Google Scholar] [CrossRef] [PubMed]
  96. Wang, C.; Kong, H.; Guan, Y.; Yang, J.; Gu, J.; Yang, S.; Xu, G. Plasma phospholipid metabolic profiling and biomarkers of type 2 diabetes mellitus based on high-performance liquid chromatography/electrospray mass spectrometry and multivariate statistical analysis. Anal. Chem. 2005, 77, 4108–4116. [Google Scholar] [CrossRef] [PubMed]
  97. Yang, J.; Xu, G.; Zheng, Y.; Kong, H.; Pang, T.; Lv, S.; Yang, Q. Diagnosis of liver cancer using HPLC-based metabonomics avoiding false-positive result from hepatitis and hepatocirrhosis diseases. J. Chromatogr. B. Anal. Technol. Biomed. Life Sci. 2004, 813, 59–65. [Google Scholar] [CrossRef] [PubMed]
  98. Odunsi, K.; Wollman, R.M.; Ambrosone, C.B.; Hutson, A.; McCann, S.E.; Tammela, J.; Geisler, J.P.; Miller, G.; Sellers, T.; Cliby, W.; et al. Detection of epithelial ovarian cancer using 1H–NMR-based metabonomics. Int. J. Cancer 2005, 113, 782–788. [Google Scholar] [CrossRef] [PubMed]
  99. Monteiro, M.S.; Carvalho, M.; Bastos, M.L.; Guedes de Pinho, P. Metabolomics analysis for biomarker discovery: Advances and challenges. Curr. Med. Chem. 2013, 20, 257–271. [Google Scholar] [CrossRef] [PubMed]
  100. Diamandis, E.P. Cancer biomarkers: Can we turn recent failures into success? J. Natl. Cancer Inst. 2010, 102, 1462–1467. [Google Scholar] [CrossRef] [PubMed]
  101. Konforte, D.; Diamandis, E.P. Is early detection of cancer with circulating biomarkers feasible? Clin. Chem. 2013, 59, 35–37. [Google Scholar] [CrossRef] [PubMed]
  102. Meierhofer, D.; Weidner, C.; Sauer, S. Integrative analysis of transcriptomics, proteomics, and metabolomics data of white adipose and liver tissue of high-fat diet and rosiglitazone-treated insulin-resistant mice identified pathway alterations and molecular hubs. J. Proteome Res. 2014, 13, 5592–5602. [Google Scholar] [CrossRef] [PubMed]
  103. Wang, K.; Lee, I.; Carlson, G.; Hood, L.; Galas, D. Systems biology and the discovery of diagnostic biomarkers. Dis. Markers 2010, 28, 199–207. [Google Scholar] [CrossRef] [PubMed]

Share and Cite

MDPI and ACS Style

Aretz, I.; Meierhofer, D. Advantages and Pitfalls of Mass Spectrometry Based Metabolome Profiling in Systems Biology. Int. J. Mol. Sci. 2016, 17, 632. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms17050632

AMA Style

Aretz I, Meierhofer D. Advantages and Pitfalls of Mass Spectrometry Based Metabolome Profiling in Systems Biology. International Journal of Molecular Sciences. 2016; 17(5):632. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms17050632

Chicago/Turabian Style

Aretz, Ina, and David Meierhofer. 2016. "Advantages and Pitfalls of Mass Spectrometry Based Metabolome Profiling in Systems Biology" International Journal of Molecular Sciences 17, no. 5: 632. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms17050632

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