Next Article in Journal
Drug-Resistance Biomarkers in Patient-Derived Colorectal Cancer Organoid and Fibroblast Co-Culture System
Next Article in Special Issue
Cadmium Stress Signaling Pathways in Plants: Molecular Responses and Mechanisms
Previous Article in Journal
Genomic Evolution Strategy in SARS-CoV-2 Lineage B: Coevolution of Cis Elements
Previous Article in Special Issue
Positive Effects of Physical Activity on Insulin Signaling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Multi Omics Applications in Biological Systems

by
Cristian D. Gutierrez Reyes
1,*,
Gerardo Alejo-Jacuinde
2,†,
Benjamin Perez Sanchez
2,†,
Jesus Chavez Reyes
3,
Sherifdeen Onigbinde
1,
Damir Mogut
4,
Irma Hernández-Jasso
3,
Denisse Calderón-Vallejo
3,
J. Luis Quintanar
3 and
Yehia Mechref
1,*
1
Department of Chemistry and Biochemistry, Texas Tech University, Lubbock, TX 79409, USA
2
Department of Plant and Soil Science, Institute of Genomics for Crop Abiotic Stress Tolerance (IGCAST), Texas Tech University, Lubbock, TX 79409, USA
3
Center of Basic Sciences, Department of Physiology and Pharmacology, Autonomous University of Aguascalientes, Aguascalientes 20392, Mexico
4
Department of Food Biochemistry, Faculty of Food Science, University of Warmia and Mazury in Olsztyn, 10-719 Olsztyn, Poland
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Curr. Issues Mol. Biol. 2024, 46(6), 5777-5793; https://doi.org/10.3390/cimb46060345
Submission received: 19 April 2024 / Revised: 31 May 2024 / Accepted: 5 June 2024 / Published: 11 June 2024
(This article belongs to the Special Issue Latest Review Papers in Molecular Biology 2024)

Abstract

:
Traditional methodologies often fall short in addressing the complexity of biological systems. In this regard, system biology omics have brought invaluable tools for conducting comprehensive analysis. Current sequencing capabilities have revolutionized genetics and genomics studies, as well as the characterization of transcriptional profiling and dynamics of several species and sample types. Biological systems experience complex biochemical processes involving thousands of molecules. These processes occur at different levels that can be studied using mass spectrometry-based (MS-based) analysis, enabling high-throughput proteomics, glycoproteomics, glycomics, metabolomics, and lipidomics analysis. Here, we present the most up-to-date techniques utilized in the completion of omics analysis. Additionally, we include some interesting examples of the applicability of multi omics to a variety of biological systems.

1. Introduction

Omics is an interdisciplinary research field focused on understanding the interactions within biological systems from a holistic perspective, considering the entire system rather than individual parts [1,2]. This field emphasizes the use of high-throughput technologies and large-scale data generated at the molecular level. Computational tools are then employed to integrate and analyze the data, enabling an understanding of biological processes. The “omics” concept encloses the subfields of genomics [3], transcriptomics [4], proteomics [5], glycoproteomics [6], glycomics [7], metabolomics [8], lipidomics [9], and several others, providing a comprehensive analysis of the corresponding biomolecules.
In recent years, omics science has emerged as a pivotal area of research in understanding cellular and molecular systems, significantly enhancing our comprehension of human health, development, diagnosis, prognosis, and aging [10]. Omics technologies offer detailed insights by cataloging variations linked with different stages of diseases or a variety of biological processes, thereby serving as a critical tool for monitoring disease progression and developing biomarkers. Stand-alone omics approaches offer a restricted point of view of the studied system, offering insights into specific molecules and their potential significance in metabolic pathways. Additionally, these approaches lack the necessary information for a comprehensive understanding of dynamic biological processes. In this regard, multi omics plays a central role in unraveling the relationships between different biomolecules and their interactions by simultaneously examining different single omics approaches (Figure 1), to provide a holistic view of the biological system [11]. This allows for the validation of individual findings and reduces the risk of false positives. As a fundamental component of system biology, multi omics allows researchers to explore biological systems as interconnected networks. However, the integration of omics data is challenging due to factors such as large data sets, data heterogeneity, sample sizes, and the need for advanced statistical methods [1]. Multi omics approaches could shed light on the fundamental causes of diseases, their functional repercussions, and pertinent interactions [12,13]. Together, omics technologies offer a comprehensive view of the biological status and changes occurring within an organism, especially in the context of disease. The integration of data from these diverse sources allows researchers to construct a more comprehensive model of disease mechanisms. This integrated approach helps identify potential diagnostic markers and therapeutic targets and enhances our understanding of the complex network of biological pathways involved in disease etiology and progression. By enabling a systems-level analysis, omics sciences facilitate the identification of key regulatory nodes and pathways that could be targeted for intervention, paving the way for personalized medicine and improved healthcare outcomes [14,15,16].
This review provides a comprehensive and up-to-date description of the most widely used omics techniques and describes their applicability for understanding biological systems. Additionally, our review focuses on holistic approaches that have demonstrated the benefits of multi omics applications.

2. Genomics and Transcriptomics

A few decades ago, scientists studied single genes at a time. Current technological innovations allow fast and cost-effective sequencing of entire genomes [17]. This tremendous change in sequencing capability is reflected in the first human genome, which was produced by an international consortium that cost billions of dollars; today, the sequencing of individual genomes can be completed for a reasonable price in a relatively short time [18]. Such technological progress encourages research efforts focused on genomic studies rather than single genes. Genomics encompasses a wide range of approaches to study and understand organisms’ DNA. This includes the analysis of genome structure and composition (genes as well as non-coding regions), their functions and interaction with each other, variation, evolution, and more. The significant improvement in the throughput and accuracy of sequencing capabilities has had an important impact on the identification of novel coding and regulatory regions [19], as well as on our understanding of the spatial organization of the DNA within a cell nucleus [20]. Genomics has also revolutionized the pharmaceutical industry by allowing the identification of protein targets and accelerating drug discovery [21,22,23]. The field has seen rapid expansion, with models and methods developed for analyzing genome rearrangements [24], gene duplication, phylogenetic networks [25], and gene clustering [26].
While genomics provides an overview of all the genetic information of organisms, transcriptomics refers to the study of gene expression patterns of the entire set of RNA molecules in an individual [27]. High-throughput sequencing technologies (e.g., Illumina) allow quantifying the number of times a given sequence is present in a sample. In global transcriptomics analyses, RNA is isolated from tissue samples and further prepared to enrich messenger RNAs (mRNAs), small RNAs (sRNAs), long non-coding RNAs (lncRNA), etc. The resulting data represent a snapshot of the enriched RNA abundance from the bulk of cells within the sample. These analyses provide an average representation of gene expression across a population of cells [28]. Although transcriptomics offers an important source of information, it does not tell the whole story about transcripts. It does not consider the rate at which transcripts are produced or degraded, nor how much of the transcripts are translated into proteins. Additional techniques, such as polysome profiling, allow researchers to estimate the translation rate of a transcript by measuring the number of ribosomes attached to individual mRNAs [29,30], and ribosome profiling can tell the ribosome footprint by isolating the section of the transcript that is bound to the ribosome [31,32]. More sophisticated library preparations involving reversible inhibition of transcription elongation combined with tagging nascent RNA can also provide information about the specific transcription rates of mRNAs and their degradation, as well as providing clues about the splicing process [33,34,35]. More recently, long-read sequencing techniques, such as PacBio Iso-Seq or Oxford Nanopore direct RNA-seq, allow the sequencing of complete transcripts to detect multiple isoforms from the same gene and their dynamics between distinct treatments [36].
Multicellular organisms show a great diversity in cellular composition. For instance, even though a human being originates from a single diploid cell as tissue develops, different cells fall into a heterogeneity of cellular fates influenced by internal and external factors [37,38]. Moreover, once fully developed, cells from the same tissues undergo distinct tasks. Currently, platforms such as Fluidigm C1, Clontech iCell8, BD Rhapsody, and 10× Genomics Chromium allow the isolation of individual cells or nuclei to prepare single-cell RNA-seq (scRNA-seq) libraries [39]. Every platform has its strengths and limitations. A recent study compared the droplet-based 10× Chromium and the microwell-based BD Rhapsody [40]. The results of this study indicate that the microwell-based platform exhibits better mRNA capture efficiency, but it can be under-represented or even lose larger cell types compared to the droplet-based system. The scRNA-seq technology has been applied to understand the dynamic activities of transcription regulation over time and the trajectories of cell states at the individual cell level, for instance, in response to neuronal activation [41]. Moreover, generating multiple single-cell data has allowed scientists to produce reference single-cell atlases to compare cellular conditions across various research investigations [42]. Although scRNA-seq has significantly increased our understanding of the complex cellular heterogeneity of tissues and organs, it has some disadvantages, such as the effect of cell dissociation and the loss of positional information [43]. The spatial context of the cells has a significant role in developmental biology, cellular communication, and disease study, among others. Emerging technologies to determine cell gene expression preserving in situ spatial locations within a tissue are known as spatial transcriptomics (ST). These technologies can be broadly divided into imaging-based and sequencing-based methods [44]. Imaging-based technologies rely on fluorescence in situ hybridization of probes that target specific genes. Conversely, sequencing-based technologies depend on spatial barcodes on an array to capture transcripts unbiasedly. Recent reviews have discussed currently available ST technologies’ characteristics, applications, advantages, and limitations [45,46]. Commercially available ST technologies for whole transcriptome profiling include the 10× Visium, which resolves several to dozens of cells [47]. New methods such as Stereo-seq can provide single-cell resolution, and it has been successfully used to study the spatiotemporal transcriptomic dynamics during organogenesis [48].

3. Proteomics, Glycoproteomics, and Glycomics

Proteomics and its associated glycoproteomics and glycomics techniques are dynamic research areas working to elucidate the associations between protein structure and function [49,50,51]. Protein abundance and post-translational modifications are the fundamental parameters utilized to evaluate the dynamic behavior of proteins. The exponential growth of these omics was possible due to their integration with mass spectrometry-based (MS-based) analysis, which enormously contributed to the creation of large and reliable data sets from a variety of sample types such as plants [52], animal models, human bodily fluids, tissues, etc. [53,54,55,56].
In MS-based proteomics, the proteins are identified using two common approaches: bottom up and top down. In the bottom-up approach, the intact proteins are digested into peptides prior to introduction to the mass spectrometer, where they are detected and fragmented to facilitate identification [57]. In top-down proteomics, the proteins are ionized and fragmented in the intact form in the mass spectrometer [58]. This approach allows an important characterization of protein isoforms and provides a general picture of the protein’s post-translational modifications, such as glycosylation, acetylation, and phosphorylation [59,60]. The protein purification stages are crucial for the obtention of sensitive and reproducible results. In the past decade, proteomics analysis was commonly performed on bodily fluids such as serum, plasma, and cerebrospinal fluid (CSF) and different types of tissues [61,62,63]. These investigations mainly searched for abundance changes in the proteome of disease samples compared to healthy controls. At present, proteomics analysis has extended to more complicated samples such as extra vesicular exosomes (EVs) [64], nucleosomes [65], or single-cell approaches [66]. For instance, Van den Ackerveken et al. [65] developed a fast and robust enrichment method able to isolate nucleosomes from plasma. Further top-down proteomics analysis of colorectal cancer (CRC) samples showed alterations in the histone PTMs after alterations of epigenetic modification [65,67]. The approach was able to identify and quantify 13 histone PTMs with significant changes in abundance between the CRC and the control samples. Recent advancements in MS instruments, such as the development of the Orbitrap Astral (Thermo Fisher Sci.), allowed researchers in the field to address single-cell proteomics using label-free quantitative (LFQ) analysis. Ye et al. [68] developed a lossless LFQ single-cell proteomics (SCP) method used for the identification of over 40,000 peptides and 5000 proteins in Hela cells. The approach integrated sample preparation utilizing the cellenONE, the proteoCHIP EVO96, and the direct transference of the sample to Evotip disposal trap columns. The samples were subsequently analyzed using the Evosep One LC, with whisper flow gradients linked to a narrow-window data independent acquisition (DIA) system.
Glycoproteomics and glycomics MS-based analysis is dedicated to the comprehensive study of carbohydrate moieties present in cells and organisms. These branches of omics science delve into understanding the structure, function, and dynamics of glycans attached to proteins and lipids across different biological systems and their role in health and disease. By analyzing these structures, glycomics and glycoproteomics analysis provides insights into cellular communication [69,70], immune response [71], pathogen interaction [72,73], and disease progression [74], making it a crucial field in biomedical research and therapeutic development. Advanced technologies such as high-performance liquid chromatography–mass spectrometry (LC-MS) are employed to decipher the complex and diverse structures of glycans and the protein O- and N-site heterogeneity. Derivatization techniques are essential due to the low ionization efficiency and poor stability of native glycan and glycopeptide structures that hinder their accurate quantitation and detection by MS systems [75,76,77]. Through derivatization, researchers can achieve more accurate glycan and glycopeptide profiling, facilitating a deeper understanding of glycan functions in biological processes and disease mechanisms. Common glycan derivatization techniques include hydrazide chemistry [78], carbamate derivatization [79], chemical labeling by reductive amination [80], and permethylation [81]. In the case of glycopeptides, the glycan moiety can be derivatized through selective reductive amination reactions directed to the sialic acid [76,77]. The glycopeptides can also be derivatized, targeting the peptide N-terminus using tandem mass tags (TMTs) [82]. Alterations in glycan expression are linked to various diseases, including cancer [83,84,85], neurodegenerative diseases [86,87,88], autoimmune disorders [89,90], and infectious diseases [73,91,92]. These changes can impact cell signaling [93], immune recognition [94], and molecular stability [95], leading to altered cellular behaviors and disease progression. For example, in cancer, altered glycosylation can promote tumor growth, facilitate metastasis, and enable evasion of immune surveillance [96]. In autoimmune diseases, aberrant glycosylation can lead to improper immune responses against self-antigens [90]. Understanding these changes is crucial for developing diagnostic markers and therapeutic targets. Integrating glycomics and proteomics analyses enhances our understanding of disease mechanisms by providing a comprehensive view of the glycome and its interaction with proteins [97,98,99]. Mechref and co-workers [53] have shown that integrating proteomics and glycomics analysis can shed light on the glycoprotein source of glycans and provide mechanistic insight into the clinicopathological changes observed in different disease conditions [53]. This strategy was applied to samples derived from healthy and mild cognitive impairment (MCI) patients. The results pointed to the isomeric N-glycans GlcNAc5,Hex6,Neu5Ac3 and GlcNAc4,Hex5,Fuc,Neu5Ac as biomarker candidates to differentiate healthy and MCI patients. Moreover, the proteomics analysis showed a correlation of the identified glycoproteins with neuroinflammation, an important process in the progression of brain-related diseases. The same research group also utilized proteomics-glycomics approaches to describe the protein glycosylation differences in 11 SARS-CoV2 variants. A significantly large abundance of sialofucosylated N-glycans was observed in the variants of concern when compared to the variants of interest. The proteomics analyses were used for the validation of variant amino acid sequences [92].

4. Metabolomics and Lipidomics

The metabolome includes all the small-molecule metabolite parts of a biological sample, including structural functions, stimulatory and inhibitory effects, fuel storage molecules, and others [100]. Lipids are structurally diverse molecules that perform multiple important functions within the cell. They are characterized by their hydrophobic nature and fundamental components of the cellular membranes [101]. Either the metabolome or the lipidome can be studied at the single-cell level or in a more complex matrix such as a bodily fluid or an organ tissue [100,102,103]. The analyses of these biological systems using high-throughput targeted or untargeted metabolomics or lipidomics MS-based analysis provide an invaluable opportunity to study and quantify differential abundance [104].
Recently, a few techniques based on single-cell metabolomics (SCM) have been developed, including Patch clamp-based nano ESI-MS, CE-MS-based SCMS, MALDI-MSI-based SCMS, LALDI-MSI-based SCMS, SIMS-MSI-based SCMS, and mass cytometry-based SCMS [105]. Metabolites are crucial components of biochemical pathways and cellular functions; therefore, quantitative metabolomics has become a powerful tool in many research fields, including those investigating cancer and neurodegenerative disorders. In this sense, there is a challenge to explore and identify metabolic hallmarks that can track disease development and progression [106]. Cancer cells can reprogram their metabolic pathways to adapt tumor microenvironments and survive pharmacological treatments [107]. This phenomenon, known as metabolic plasticity, results in a heterogenous tumor with different metabolic phenotypes that are hard to characterize and consequently difficult to treat [108]. In this regard, SCM has been applied in the identification of sub-types of circulating tumor cells with distinct metastatic potential [109], and provided useful information for drug therapy [110,111]. In addition, SCM has been used to study the dynamic cancer cell–cell interaction during the drug resistance process [112] and the potential role of microbiota in the development and progression of different cancer types [113]. As metabolites directly reflect the cell behavior, the integration of SCM with other single-cell omics could suggest mechanisms underlying tumor development and drug resistance. Metabolomics have also been used to investigate neurodegenerative disorders [114]. Despite the expanding prevalence of SCM, scarce evidence of studying neurodegenerative disorders at the single-cell level has been reported. Metabolomics was applied to a preclinical model of Alzheimer’s disease (AD); seven AD-related neurotransmitters and sixteen biomarkers were identified by SCM in PC12 cells [115]. It is only a matter of time before SCM will be employed on neurodegenerative disorders to explore the individual role of each neuron and glia cell in the progress of neurodegeneration.
Other most affordable LC–MS/MS strategies employed for metabolomics analysis of biological samples have served to identify potential biomarkers in human serum from patients with diabetes [116], schizophrenia [117], ischemic stroke [118], and other health conditions [119,120,121]. Similar technique also have been used to assess food quality [122,123,124,125,126] and the toxic effects of xenobiotics [127,128]. In order to enhance accuracy, precision, and throughput, common LC-MS techniques have employed 13C or 18O isotopic approaches [129,130,131,132]. MALDI imagining is an MS-based technique widely used for the characterization of tissues derived from several health conditions facilitating their diagnosis and therapy [133,134,135,136,137].
Lipids possess significant nutritional value due to their high caloric density and their capacity to transport essential substances that the human body cannot produce independently. Lipids are also involved in intra- and extracellular signaling processes [138]. Lipidomics focuses on the study and characterization of cellular and bodily fluid lipids, along with their interactions and functions within the body. This field is a crucial tool in healthcare for identifying biomarkers useful in diagnosing and developing therapies for various diseases [139,140]. For example, the nervous system contains a large amount of lipids. It has been suggested that neurological disorders may be associated with lipid homeostasis, among other factors [103,141]. It has also been described that there are changes in the lipidomic profiles of bioactive lipids in plasma and CSF, as well as in specific anatomical regions of the brain associated with the pathological characteristics of AD and Parkinson’s disease [142]. Significant alterations in plasma lipids have also been described in rats and mice with spinal cord and sciatic nerve injury, which indicates that lipid metabolism could be related to the recovery and/or damage processes following nerve injury [143,144]. The application of lipidomics to disorders associated with metabolic syndrome has been widely described. Lipidomics plays a key role in risk prediction studies and therapeutic monitoring of diseases related to metabolic syndrome, given the close association of lipids with these diseases [145]. Furthermore, this discipline has been valuable for determining population profiles, conducting research on pathogenesis, and for identifying biomarkers and monitoring therapeutic responses [146]. It has been described that the lipidome is altered in many neoplastic diseases. The main applications of lipidomics in these types of disorders have focused on the detection and classification of neoplastic cells or tissues, on differentiation between neoplastic and normal environments, in the evaluation of cancer treatments, and for the discovery of new tumor biomarkers [147].

5. Multi Omics Integration

The study of organisms as complex systems requires the integration of distinct approaches (Figure 1). For instance, although genomic and transcriptomic analyses offer powerful tools in our understanding of molecular processes, downstream gaps cannot be filled without integrating other approaches. None of the above methods provide information about what happens to a gene’s translated product or the effect of metabolite accumulation. Consideration of these questions highlights the relevance of multi omics analysis, as they provide a more complete and holistic landscape of cell molecular interactions. Table 1 summarizes additional omics approaches, focusing on the multi-integration of molecular data sets.
Omics technologies are high-throughput molecular assays that produce large data sets of information. Public repositories facilitate the accessibility of this information to the scientific community. Some examples are the following: for high-throughput sequencing SRA [173]; for functional genomics and DNA sequence information NCBI GEO [174], GenBank [175]; for spatial transcriptomics CROST [176]; for proteomics and glycoproteomics PRIDE [177], Proteome Xchange [178]; for glycomics GlycoPOST [179]; for metabolomics and lipidomics MetaboLights [180], HMDB [181]; and many others for multi omics as listed in the references [182,183,184]. However, single omics relay in the application of common statistics mainly focusing on the differentiation of the studied groups, for example, principal component analysis (PCA) plots; heat maps; receiver operating characteristic (ROC) curves; dot plots; statistical tests such as Anova, t-test, Mann–Whitney U Test, and several others in accordance with the data size and distribution; and the appropriated false discovery rate (FDR) corrections such as Benjamini-Hochberg, Bonferroni, etc. [185,186,187]. Regarding the multi omics integration, the application of computational approaches is necessary due to the data complexity. Ingenuity pathway analysis (IPA) is an extensive database from QIAGEN Co. able to integrate transcriptomics, proteomics, and metabolomics data in diverse biological pathways [188]. A similar free alternative is iPathwayGuide developed from ADVAITA Co. [189]; many other software developments and its applications in diverse multi omics investigations are described in the following references: IOAT [190], multi omics of primary immunodeficiencies [191], Omics TIDE [192], IBAG [193], cancer omics [194], and TiMEG [195].

Foodomics

Foodomics represents the integration of advanced omics technologies in food and nutrition studies, aimed at enhancing consumer welfare, health, and understanding [196]. This interdisciplinary field merges food chemistry, biological sciences, and data analysis techniques to comprehend food and its constituents at a molecular level, with the overarching goal of improving food quality, safety, and nutrition by exploring the intricate interactions between food components and the human body. In foodomics research, a primary focus lies in understanding the interplay between food constituents and biological systems, striving to identify biomarkers that objectively measure food intake and shed light on its physiological effects [197]. Foodomics methodologies find application in various studies aimed at linking dietary patterns with health outcomes. Research outcomes can utilize metabolomics techniques to identify specific compounds in diets and analyze their impact on health. For instance, ongoing studies apply metabolome analysis to unveil food-specific compounds (FSCs) that can bridge dietary patterns, such as the Mediterranean-style (MED) diet, with overall health outcomes. Drawing from a controlled feeding MED intervention, the study validates the identification of FSCs from eight distinct foods, detectable in biospecimens post-consumption [198]. Another example is the use of nuclear magnetic resonance (NMR) to reveal alterations of protein and carbohydrate metabolisms in different diets [199]. This study provides new insights into the effects of a healthy diet on glycemia, reduction of inflammation, and weight loss among obese individuals, as well as alteration of the gut microbiota metabolism.
Dilmore et al. [200] evidenced that the modified Mediterranean ketogenic diet (MMKD) could offer potential benefits in combating memory decline associated with AD. The study implemented MMKD dietary interventions for participants with mild cognitive impairment (MCI) to assess their cognitive function. The study utilized shotgun metagenomics to analyze the gut microbiota composition of participants, providing insights into how these dietary interventions influenced the microbiome. Additionally, untargeted metabolomics was employed to analyze metabolite profiles in samples collected at various intervals during the dietary interventions, to identify changes in metabolites related to cognitive function and gut health. However, the untargeted metabolomics platform has limitations in characterizing lipid metabolism and mitochondrial function compared to targeted platforms. Through metagenomic and metabolomic analyses, potential biomarkers associated with cognitive function and gut health were identified. For example, changes in the levels of GABA-producing microbes and GABA-regulating bacteria were observed in individuals with MCI following the MMKD, suggesting a link between gut microbiota composition and cognitive status. Overall, the study integrated a multi omics approach to investigate the relationship between dietary interventions, gut microbiota composition, metabolite profiles, and cognitive function in individuals at risk for AD.
Foodomics represents a dynamic and interdisciplinary field that holds promise for enhancing our understanding of food and its effects on health and well-being. This field has implications for food industry practices, public health policies, and personalized nutrition strategies, especially with the rise of alternative sources of nutrients.

6. Concluding Remarks

Omics sciences stand at the forefront of biomedical research, offering unprecedented insights into the molecular underpinnings of diseases. Their application in disease study promises to revolutionize the diagnosis, treatment, and prevention of human diseases, highlighting their indispensable role in advancing medical science and improving patient care. In this review, we have discussed the current technological innovations that allow fast and cost-effective sequencing of entire genomes. Additionally, we revised a set of novel and accurate transcriptomics techniques focused on showing how the transcript patterns are affected by the development or disease progression. The development of modern MS techniques has allowed the acquisition of high-throughput data sets in the proteomics, metabolomics, and lipidomics fields. Additionally, we described examples of recent studies using single cells or derived from diverse bodily fluids, as well as plant materials. This review offers an extensive and up-to-date overview of the most widely used omics techniques, detailing the relevance of the omics field in understanding different types of biological systems using a holistic view.

Author Contributions

Conceptualization, Y.M., C.D.G.R., and G.A.-J.; investigation, C.D.G.R., G.A.-J., B.P.S., J.C.R., S.O., D.M., I.H.-J., and D.C.-V.; resources, Y.M.; writing—original draft preparation, C.D.G.R., G.A.-J., B.P.S., J.C.R., S.O., D.M., I.H.-J., and D.C.-V.; writing—review and editing, C.D.G.R., G.A.-J., and J.L.Q.; visualization, C.D.G.R., G.A.-J., and B.P.S.; supervision, Y.M. and C.D.G.R.; project administration, Y.M., C.D.G.R., and G.A.-J.; funding acquisition, Y.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by grants from the National Institutes of Health, NIH, including 1R01GM112490-08 (YM), 1R01GM130091-05 (YM), and 1U01CA225753-05 (YM). The work is also supported by grants from the Robert A. Welch Foundation no. D-0005 and The CH Foundation (YM). Project financially supported by Minister of Education and Science in the range of the program entitled “Regional Initiative of Excellence” for the years 2019-2023, project no. 010/RID/2018/19 (DM).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Karahalil, B. Overview of Systems Biology and Omics Technologies. Curr. Med. Chem. 2016, 23, 4221–4230. [Google Scholar] [CrossRef] [PubMed]
  2. Westerhoff, H.V.; Palsson, B.O. The evolution of molecular biology into systems biology. Nat. Biotechnol. 2004, 22, 1249–1252. [Google Scholar] [CrossRef] [PubMed]
  3. Munung, N.S.; Ambele, M.A.; Moela, P. Advancing global equity in cancer genomics—Challenges and opportunities in Sub-Saharan Africa. Curr. Opin. Genet. Dev. 2021, 66, 20–24. [Google Scholar] [CrossRef] [PubMed]
  4. Heidecker, B.; Hare, J.M. The use of transcriptomic biomarkers for personalized medicine. Heart Fail. Rev. 2007, 12, 1–11. [Google Scholar] [CrossRef]
  5. He, T. Implementation of Proteomics in Clinical Trials. Proteom. Clin. Appl. 2019, 13, e1800198. [Google Scholar] [CrossRef] [PubMed]
  6. Mechref, Y.; Peng, W.; Gautam, S.; Ahmadi, P.; Lin, Y.; Zhu, J.; Zhang, J.; Liu, S.; Singal, A.G.; Parikh, N.D.; et al. Mass spectrometry based biomarkers for early detection of HCC using a glycoproteomic approach. Adv. Cancer Res. 2023, 157, 23–56. [Google Scholar] [CrossRef]
  7. Peng, W.; Zhao, J.; Dong, X.; Banazadeh, A.; Huang, Y.; Hussien, A.; Mechref, Y. Clinical application of quantitative glycomics. Expert Rev. Proteom. 2018, 15, 1007–1031. [Google Scholar] [CrossRef]
  8. Dona, A.C.; Coffey, S.; Figtree, G. Translational and emerging clinical applications of metabolomics in cardiovascular disease diagnosis and treatment. Eur. J. Prev. Cardiol. 2016, 23, 1578–1589. [Google Scholar] [CrossRef]
  9. Astarita, G.; Stocchero, M.; Paglia, G. Unbiased Lipidomics and Metabolomics of Human Brain Samples. Methods Mol. Biol. 2018, 1750, 255–269. [Google Scholar] [CrossRef]
  10. Aizat, W.M.; Ismail, I.; Noor, N.M. Recent Development in Omics Studies. Adv. Exp. Med. Biol. 2018, 1102, 1–9. [Google Scholar] [CrossRef]
  11. Hasin, Y.; Seldin, M.; Lusis, A. Multi-omics approaches to disease. Genome Biol. 2017, 18, 83. [Google Scholar] [CrossRef] [PubMed]
  12. 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]
  13. Dihazi, H.; Asif, A.R.; Beißbarth, T.; Bohrer, R.; Feussner, K.; Feussner, I.; Jahn, O.; Lenz, C.; Majcherczyk, A.; Schmidt, B.; et al. Integrative omics—From data to biology. Expert Rev. Proteom. 2018, 15, 463–466. [Google Scholar] [CrossRef] [PubMed]
  14. Raufaste-Cazavieille, V.; Santiago, R.; Droit, A. Multi-omics analysis: Paving the path toward achieving precision medicine in cancer treatment and immuno-oncology. Front. Mol. Biosci. 2022, 9, 962743. [Google Scholar] [CrossRef] [PubMed]
  15. Sen, P.; Orešič, M. Integrating Omics Data in Genome-Scale Metabolic Modeling: A Methodological Perspective for Precision Medicine. Metabolites 2023, 13, 855. [Google Scholar] [CrossRef]
  16. Misra, B.B.; Langefeld, C.D.; Olivier, M.; Cox, L.A. Integrated Omics: Tools, Advances, and Future Approaches. J. Mol. Endocrinol. 2018, 62, R21–R45. [Google Scholar] [CrossRef]
  17. Katara, A.; Chand, S.; Chaudhary, H.; Chaudhry, V.; Chandra, H.; Dubey, R.C. Evolution and applications of Next Generation Sequencing and its intricate relations with chromatographic and spectrometric techniques in modern day sciences. J. Chromatogr. Open 2024, 5, 100121. [Google Scholar] [CrossRef]
  18. McGuire, A.L.; Gabriel, S.; Tishkoff, S.A.; Wonkam, A.; Chakravarti, A.; Furlong, E.E.M.; Treutlein, B.; Meissner, A.; Chang, H.Y.; López-Bigas, N.; et al. The road ahead in genetics and genomics. Nat. Rev. Genet. 2020, 21, 581–596. [Google Scholar] [CrossRef]
  19. Cano-Gamez, E.; Trynka, G. From GWAS to Function: Using Functional Genomics to Identify the Mechanisms Underlying Complex Diseases. Front. Genet. 2020, 11, 505357. [Google Scholar] [CrossRef]
  20. Jerkovic’, I.; Cavalli, G. Understanding 3D genome organization by multidisciplinary methods. Nat. Rev. Mol. Cell Biol. 2021, 22, 511–528. [Google Scholar] [CrossRef]
  21. Schneider, M.; Radoux, C.J.; Hercules, A.; Ochoa, D.; Dunham, I.; Zalmas, L.-P.; Hessler, G.; Ruf, S.; Shanmugasundaram, V.; Hann, M.M.; et al. The PROTACtable genome. Nat. Rev. Drug Discov. 2021, 20, 789–797. [Google Scholar] [CrossRef] [PubMed]
  22. Zhang, L.; Song, J.; Kong, L.; Yuan, T.; Li, W.; Zhang, W.; Hou, B.; Lu, Y.; Du, G. The strategies and techniques of drug discovery from natural products. Pharmacol. Ther. 2020, 216, 107686. [Google Scholar] [CrossRef] [PubMed]
  23. Kiriiri, G.K.; Njogu, P.M.; Mwangi, A.N. Exploring different approaches to improve the success of drug discovery and development projects: A review. Future J. Pharm. Sci. 2020, 6, 27. [Google Scholar] [CrossRef]
  24. Goel, M.; Sun, H.; Jiao, W.-B.; Schneeberger, K. SyRI: Finding genomic rearrangements and local sequence differences from whole-genome assemblies. Genome Biol. 2019, 20, 277. [Google Scholar] [CrossRef]
  25. Blair, C.; Ané, C. Phylogenetic Trees and Networks Can Serve as Powerful and Complementary Approaches for Analysis of Genomic Data. Syst. Biol. 2019, 69, 593–601. [Google Scholar] [CrossRef] [PubMed]
  26. Gilchrist, C.L.M.; Booth, T.J.; van Wersch, B.; van Grieken, L.; Medema, M.H.; Chooi, Y.-H. Cblaster: A remote search tool for rapid identification and visualization of homologous gene clusters. Bioinform. Adv. 2021, 1, vbab016. [Google Scholar] [CrossRef]
  27. Vishwanath, P.P.; Bidaramali, V.; Lata, S.; Yadav, R.K. Transcriptomics: Illuminating the molecular landscape of vegetable crops: A review. J. Plant Biochem. Biotechnol. 2024. [Google Scholar] [CrossRef]
  28. Lee, J.-Y. The Principles and Applications of High-Throughput Sequencing Technologies. Dev. Reprod. 2023, 27, 9–24. [Google Scholar] [CrossRef]
  29. Nguyen, H.L.; Duviau, M.-P.; Cocaign-Bousquet, M.; Nouaille, S.; Girbal, L. Multiplexing polysome profiling experiments to study translation in Escherichia coli. PLoS ONE 2019, 14, e0212297. [Google Scholar] [CrossRef]
  30. Pringle, E.S.; McCormick, C.; Cheng, Z. Polysome Profiling Analysis of mRNA and Associated Proteins Engaged in Translation. Curr. Protoc. Mol. Biol. 2019, 125, e79. [Google Scholar] [CrossRef]
  31. Brar, G.A.; Weissman, J.S. Ribosome profiling reveals the what, when, where and how of protein synthesis. Nat. Rev. Mol. Cell Biol. 2015, 16, 651–664. [Google Scholar] [CrossRef] [PubMed]
  32. Liang, S.; Bellato, H.M.; Lorent, J.; Lupinacci, F.C.S.; Oertlin, C.; van Hoef, V.; Andrade, V.P.; Roffé, M.; Masvidal, L.; Hajj, G.N.M.; et al. Polysome-profiling in small tissue samples. Nucleic Acids Res. 2018, 46, e3. [Google Scholar] [CrossRef] [PubMed]
  33. Fuchs, G.; Voichek, Y.; Rabani, M.; Benjamin, S.; Gilad, S.; Amit, I.; Oren, M. Simultaneous measurement of genome-wide transcription elongation speeds and rates of RNA polymerase II transition into active elongation with 4sUDRB-seq. Nat. Protoc. 2015, 10, 605–618. [Google Scholar] [CrossRef] [PubMed]
  34. Fuchs, G.; Voichek, Y.; Benjamin, S.; Gilad, S.; Amit, I.; Oren, M. 4sUDRB-seq: Measuring genomewide transcriptional elongation rates and initiation frequencies within cells. Genome Biol. 2014, 15, R69. [Google Scholar] [CrossRef] [PubMed]
  35. Zhang, X.-O.; Fu, Y.; Mou, H.; Xue, W.; Weng, Z. The temporal landscape of recursive splicing during Pol II transcription elongation in human cells. PLoS Genet. 2018, 14, e1007579. [Google Scholar] [CrossRef] [PubMed]
  36. Zhao, L.; Zhang, H.; Kohnen, M.V.; Prasad, K.V.S.K.; Gu, L.; Reddy, A.S.N. Analysis of Transcriptome and Epitranscriptome in Plants Using PacBio Iso-Seq and Nanopore-Based Direct RNA Sequencing. Front. Genet. 2019, 10, 430951. [Google Scholar] [CrossRef] [PubMed]
  37. Song, Y.; Xu, X.; Wang, W.; Tian, T.; Zhu, Z.; Yang, C. Single cell transcriptomics: Moving towards multi-omics. Analyst 2019, 144, 3172–3189. [Google Scholar] [CrossRef] [PubMed]
  38. Adil, A.; Kumar, V.; Jan, A.T.; Asger, M. Single-Cell Transcriptomics: Current Methods and Challenges in Data Acquisition and Analysis. Front. Neurosci. 2021, 15, 591122. [Google Scholar] [CrossRef] [PubMed]
  39. Wen, L.; Li, G.; Huang, T.; Geng, W.; Pei, H.; Yang, J.; Zhu, M.; Zhang, P.; Hou, R.; Tian, G.; et al. Single-cell technologies: From research to application. Innovation 2022, 3, 100342. [Google Scholar] [CrossRef]
  40. Salcher, S.; Heidegger, I.; Untergasser, G.; Fotakis, G.; Scheiber, A.; Martowicz, A.; Noureen, A.; Krogsdam, A.; Schatz, C.; Schäfer, G.; et al. Comparative analysis of 10X Chromium vs. BD Rhapsody whole transcriptome single-cell sequencing technologies in complex human tissues. Heliyon 2024, 10, e28358. [Google Scholar] [CrossRef]
  41. Qiu, Q.; Hu, P.; Qiu, X.; Govek, K.W.; Cámara, P.G.; Wu, H. Massively parallel and time-resolved RNA sequencing in single cells with scNT-seq. Nat. Methods 2020, 17, 991–1001. [Google Scholar] [CrossRef]
  42. Andreatta, M.; Corria-Osorio, J.; Müller, S.; Cubas, R.; Coukos, G.; Carmona, S.J. Interpretation of T cell states from single-cell transcriptomics data using reference atlases. Nat. Commun. 2021, 12, 2965. [Google Scholar] [CrossRef]
  43. Adema, K.; Schon, M.A.; Nodine, M.D.; Kohlen, W. Lost in space: What single-cell RNA sequencing cannot tell you. Trends Plant Sci. 2024. [Google Scholar] [CrossRef]
  44. Wang, Y.; Liu, B.; Zhao, G.; Lee, Y.; Buzdin, A.; Mu, X.; Zhao, J.; Chen, H.; Li, X. Spatial transcriptomics: Technologies, applications and experimental considerations. Genomics 2023, 115, 110671. [Google Scholar] [CrossRef]
  45. Chen, T.-Y.; You, L.; Hardillo, J.A.U.; Chien, M.-P. Spatial Transcriptomic Technologies. Cells 2023, 12, 2042. [Google Scholar] [CrossRef]
  46. Piñeiro, A.J.; Houser, A.E.; Ji, A.L. Research Techniques Made Simple: Spatial Transcriptomics. J. Investig. Dermatol. 2022, 142, 993–1001.e1001. [Google Scholar] [CrossRef]
  47. Shi, W.; Zhang, J.; Huang, S.; Fan, Q.; Cao, J.; Zeng, J.; Wu, L.; Yang, C. Next-Generation Sequencing-Based Spatial Transcriptomics: A Perspective from Barcoding Chemistry. JACS Au 2024, 4, 1723–1743. [Google Scholar] [CrossRef]
  48. Chen, A.; Liao, S.; Cheng, M.; Ma, K.; Wu, L.; Lai, Y.; Qiu, X.; Yang, J.; Xu, J.; Hao, S.; et al. Spatiotemporal transcriptomic atlas of mouse organogenesis using DNA nanoball-patterned arrays. Cell 2022, 185, 1777–1792.e1721. [Google Scholar] [CrossRef]
  49. Al-Amrani, S.; Al-Jabri, Z.; Al-Zaabi, A.; Alshekaili, J.; Al-Khabori, M. Proteomics: Concepts and applications in human medicine. World J. Biol. Chem. 2021, 12, 57–69. [Google Scholar] [CrossRef]
  50. Gutierrez Reyes, C.D.; Jiang, P.; Donohoo, K.; Atashi, M.; Mechref, Y.S. Glycomics and glycoproteomics: Approaches to address isomeric separation of glycans and glycopeptides. J. Sep. Sci. 2021, 44, 403–425. [Google Scholar] [CrossRef]
  51. Peng, W.; Gutierrez Reyes, C.D.; Gautam, S.; Yu, A.; Cho, B.G.; Goli, M.; Donohoo, K.; Mondello, S.; Kobeissy, F.; Mechref, Y. MS-based glycomics and glycoproteomics methods enabling isomeric characterization. Mass. Spectrom. Rev. 2023, 42, 577–616. [Google Scholar] [CrossRef] [PubMed]
  52. Alejo-Jacuinde, G.; Nájera-González, H.-R.; Chávez Montes, R.A.; Gutierrez Reyes, C.D.; Barragán-Rosillo, A.C.; Perez Sanchez, B.; Mechref, Y.; López-Arredondo, D.; Yong-Villalobos, L.; Herrera-Estrella, L. Multi-omic analyses reveal the unique properties of chia (Salvia hispanica) seed metabolism. Commun. Biol. 2023, 6, 820. [Google Scholar] [CrossRef]
  53. Reyes, C.D.G.; Hakim, M.A.; Atashi, M.; Goli, M.; Gautam, S.; Wang, J.; Bennett, A.I.; Zhu, J.; Lubman, D.M.; Mechref, Y. LC-MS/MS Isomeric Profiling of N-Glycans Derived from Low-Abundant Serum Glycoproteins in Mild Cognitive Impairment Patients. Biomolecules 2022, 12, 1657. [Google Scholar] [CrossRef] [PubMed]
  54. Cho, B.G.; Gutierrez Reyes, C.D.; Mechref, Y. N-Glycomics of Cerebrospinal Fluid: Method Comparison. Molecules 2021, 26, 1712. [Google Scholar] [CrossRef]
  55. Gonçalves, E.; Poulos, R.C.; Cai, Z.; Barthorpe, S.; Manda, S.S.; Lucas, N.; Beck, A.; Bucio-Noble, D.; Dausmann, M.; Hall, C.; et al. Pan-cancer proteomic map of 949 human cell lines. Cancer Cell 2022, 40, 835–849.e838. [Google Scholar] [CrossRef]
  56. Rontogianni, S.; Synadaki, E.; Li, B.; Liefaard, M.C.; Lips, E.H.; Wesseling, J.; Wu, W.; Altelaar, M. Proteomic profiling of extracellular vesicles allows for human breast cancer subtyping. Commun. Biol. 2019, 2, 325. [Google Scholar] [CrossRef]
  57. Dupree, E.J.; Jayathirtha, M.; Yorkey, H.; Mihasan, M.; Petre, B.A.; Darie, C.C. A Critical Review of Bottom-Up Proteomics: The Good, the Bad, and the Future of this Field. Proteomes 2020, 8, 14. [Google Scholar] [CrossRef]
  58. Toby, T.K.; Fornelli, L.; Srzentić, K.; DeHart, C.J.; Levitsky, J.; Friedewald, J.; Kelleher, N.L. A comprehensive pipeline for translational top-down proteomics from a single blood draw. Nat. Protoc. 2019, 14, 119–152. [Google Scholar] [CrossRef] [PubMed]
  59. Chen, L.; Kashina, A. Post-translational Modifications of the Protein Termini. Front. Cell Dev. Biol. 2021, 9, 719590. [Google Scholar] [CrossRef]
  60. Lee, J.M.; Hammarén, H.M.; Savitski, M.M.; Baek, S.H. Control of protein stability by post-translational modifications. Nat. Commun. 2023, 14, 201. [Google Scholar] [CrossRef]
  61. Johnson, E.C.B.; Bian, S.; Haque, R.U.; Carter, E.K.; Watson, C.M.; Gordon, B.A.; Ping, L.; Duong, D.M.; Epstein, M.P.; McDade, E.; et al. Cerebrospinal fluid proteomics define the natural history of autosomal dominant Alzheimer’s disease. Nat. Med. 2023, 29, 1979–1988. [Google Scholar] [CrossRef] [PubMed]
  62. Palstrøm, N.B.; Matthiesen, R.; Rasmussen, L.M.; Beck, H.C. Recent Developments in Clinical Plasma Proteomics-Applied to Cardiovascular Research. Biomedicines 2022, 10, 162. [Google Scholar] [CrossRef] [PubMed]
  63. Ma, M.; Huo, S.; Zhang, M.; Qian, S.; Zhu, X.; Pu, J.; Rasam, S.; Xue, C.; Shen, S.; An, B.; et al. In-depth mapping of protein localizations in whole tissue by micro-scaffold assisted spatial proteomics (MASP). Nat. Commun. 2022, 13, 7736. [Google Scholar] [CrossRef] [PubMed]
  64. Kangas, P.; Nyman, T.A.; Metsähonkala, L.; Burns, C.; Tempest, R.; Williams, T.; Karttunen, J.; Jokinen, T.S. Towards optimised extracellular vesicle proteomics from cerebrospinal fluid. Sci. Rep. 2023, 13, 9564. [Google Scholar] [CrossRef] [PubMed]
  65. Van den Ackerveken, P.; Lobbens, A.; Turatsinze, J.-V.; Solis-Mezarino, V.; Völker-Albert, M.; Imhof, A.; Herzog, M. A novel proteomics approach to epigenetic profiling of circulating nucleosomes. Sci. Rep. 2021, 11, 7256. [Google Scholar] [CrossRef] [PubMed]
  66. Petrosius, V.; Schoof, E.M. Recent advances in the field of single-cell proteomics. Transl. Oncol. 2023, 27, 101556. [Google Scholar] [CrossRef]
  67. Liu, R.; Wu, J.; Guo, H.; Yao, W.; Li, S.; Lu, Y.; Jia, Y.; Liang, X.; Tang, J.; Zhang, H. Post-translational modifications of histones: Mechanisms, biological functions, and therapeutic targets. MedComm 2023, 4, e292. [Google Scholar] [CrossRef] [PubMed]
  68. Ye, Z.; Sabatier, P.; Hoeven, L.V.D.; Phlairaharn, T.; Hartlmayr, D.; Izaguirre, F.; Seth, A.; Joshi, H.J.; Bekker-Jensen, D.B.; Bache, N.; et al. High-throughput and scalable single cell proteomics identifies over 5000 proteins per cell. bioRxiv 2023. [Google Scholar] [CrossRef]
  69. Chandler, K.B.; Leon, D.R.; Kuang, J.; Meyer, R.D.; Rahimi, N.; Costello, C.E. N-Glycosylation regulates ligand-dependent activation and signaling of vascular endothelial growth factor receptor 2 (VEGFR2). J. Biol. Chem. 2019, 294, 13117–13130. [Google Scholar] [CrossRef]
  70. Perkey, E.; Maurice De Sousa, D.; Carrington, L.; Chung, J.; Dils, A.; Granadier, D.; Koch, U.; Radtke, F.; Ludewig, B.; Blazar, B.R.; et al. GCNT1-Mediated O-Glycosylation of the Sialomucin CD43 Is a Sensitive Indicator of Notch Signaling in Activated T Cells. J. Immunol. 2020, 204, 1674–1688. [Google Scholar] [CrossRef]
  71. Van Coillie, J.; Schulz, M.A.; Bentlage, A.E.H.; de Haan, N.; Ye, Z.; Geerdes, D.M.; van Esch, W.J.E.; Hafkenscheid, L.; Miller, R.L.; Narimatsu, Y.; et al. Role of N-Glycosylation in FcγRIIIa interaction with IgG. Front. Immunol. 2022, 13, 987151. [Google Scholar] [CrossRef]
  72. Ozdilek, A.; Paschall, A.V.; Dookwah, M.; Tiemeyer, M.; Avci, F.Y. Host protein glycosylation in nucleic acid vaccines as a potential hurdle in vaccine design for nonviral pathogens. Proc. Natl. Acad. Sci. USA 2020, 117, 1280–1282. [Google Scholar] [CrossRef]
  73. Onigbinde, S.; Reyes, C.D.G.; Fowowe, M.; Daramola, O.; Atashi, M.; Bennett, A.I.; Mechref, Y. Variations in O-Glycosylation Patterns Influence Viral Pathogenicity, Infectivity, and Transmissibility in SARS-CoV-2 Variants. Biomolecules 2023, 13, 1467. [Google Scholar] [CrossRef]
  74. Munkley, J.; Mills, I.G.; Elliott, D.J. The role of glycans in the development and progression of prostate cancer. Nat. Rev. Urol. 2016, 13, 324–333. [Google Scholar] [CrossRef]
  75. Zhou, S.; Veillon, L.; Dong, X.; Huang, Y.; Mechref, Y. Direct comparison of derivatization strategies for LC-MS/MS analysis of N-glycans. Analyst 2017, 142, 4446–4455. [Google Scholar] [CrossRef]
  76. Jiang, P.; Huang, Y.; Gutierrez Reyes, C.D.; Zhong, J.; Mechref, Y. Isomeric Separation of α2,3/α2,6-Linked 2-Aminobenzamide (2AB)-Labeled Sialoglycopeptides by C18-LC-MS/MS. Anal. Chem. 2023, 95, 18388–18397. [Google Scholar] [CrossRef]
  77. Zhong, J.; Huang, Y.; Mechref, Y. Derivatization of Sialylated Glycopeptides (DOSG) Enabling Site-Specific Isomeric Profiling Using LC-MS/MS. Anal. Chem. 2021, 93, 5763–5772. [Google Scholar] [CrossRef]
  78. Jiang, K.; Aloor, A.; Qu, J.; Xiao, C.; Wu, Z.; Ma, C.; Zhang, L.; Wang, P.G. Rapid and sensitive MALDI MS analysis of oligosaccharides by using 2-hydrazinopyrimidine as a derivative reagent and co-matrix. Anal. Bioanal. Chem. 2017, 409, 421–429. [Google Scholar] [CrossRef]
  79. Wu, Y.; Sha, Q.; Du, J.; Wang, C.; Zhang, L.; Liu, B.F.; Lin, Y.; Liu, X. Determination of N-glycans by high performance liquid chromatography using 6-aminoquinolyl-N-hydroxysuccinimidyl carbamate as the glycosylamine labeling reagent. J. Chromatogr. A 2018, 1535, 114–122. [Google Scholar] [CrossRef]
  80. Ruhaak, L.R.; Zauner, G.; Huhn, C.; Bruggink, C.; Deelder, A.M.; Wuhrer, M. Glycan labeling strategies and their use in identification and quantification. Anal. Bioanal. Chem. 2010, 397, 3457–3481. [Google Scholar] [CrossRef]
  81. Mechref, Y.; Kang, P.; Novotny, M.V. Solid-phase permethylation for glycomic analysis. Methods Mol. Biol. 2009, 534, 53–64. [Google Scholar] [CrossRef] [PubMed]
  82. Saraswat, M.; Mangalaparthi, K.K.; Garapati, K.; Pandey, A. TMT-Based Multiplexed Quantitation of N-Glycopeptides Reveals Glycoproteome Remodeling Induced by Oncogenic Mutations. ACS Omega 2022, 7, 11023–11032. [Google Scholar] [CrossRef] [PubMed]
  83. Dube, D.H.; Bertozzi, C.R. Glycans in cancer and inflammation—Potential for therapeutics and diagnostics. Nat. Rev. Drug Discov. 2005, 4, 477–488. [Google Scholar] [CrossRef] [PubMed]
  84. Taniguchi, N.; Kizuka, Y. Glycans and cancer: Role of N-glycans in cancer biomarker, progression and metastasis, and therapeutics. Adv. Cancer Res. 2015, 126, 11–51. [Google Scholar] [CrossRef] [PubMed]
  85. Gutierrez Reyes, C.D.; Huang, Y.; Atashi, M.; Zhang, J.; Zhu, J.; Liu, S.; Parikh, N.D.; Singal, A.G.; Dai, J.; Lubman, D.M.; et al. PRM-MS Quantitative Analysis of Isomeric N-Glycopeptides Derived from Human Serum Haptoglobin of Patients with Cirrhosis and Hepatocellular Carcinoma. Metabolites 2021, 11, 563. [Google Scholar] [CrossRef] [PubMed]
  86. Veillon, L.; Huang, Y.; Peng, W.; Dong, X.; Cho, B.G.; Mechref, Y. Characterization of isomeric glycan structures by LC-MS/MS. Electrophoresis 2017, 38, 2100–2114. [Google Scholar] [CrossRef]
  87. Kizuka, Y.; Kitazume, S.; Taniguchi, N. N-glycan and Alzheimer’s disease. Biochim. Biophys. Acta Gen. Subj. 2017, 1861, 2447–2454. [Google Scholar] [CrossRef]
  88. Tena, J.; Tang, X.; Zhou, Q.; Harvey, D.; Barajas-Mendoza, M.; Jin, L.W.; Maezawa, I.; Zivkovic, A.M.; Lebrilla, C.B. Glycosylation alterations in serum of Alzheimer’s disease patients show widespread changes in N-glycosylation of proteins related to immune function, inflammation, and lipoprotein metabolism. Alzheimers Dement. 2022, 14, e12309. [Google Scholar] [CrossRef] [PubMed]
  89. Pinho, S.S.; Alves, I.; Gaifem, J.; Rabinovich, G.A. Immune regulatory networks coordinated by glycans and glycan-binding proteins in autoimmunity and infection. Cell Mol. Immunol. 2023, 20, 1101–1113. [Google Scholar] [CrossRef]
  90. Chien, M.W.; Fu, S.H.; Hsu, C.Y.; Liu, Y.W.; Sytwu, H.K. The Modulatory Roles of N-glycans in T-Cell-Mediated Autoimmune Diseases. Int. J. Mol. Sci. 2018, 19, 780. [Google Scholar] [CrossRef]
  91. Unione, L.; Gimeno, A.; Valverde, P.; Calloni, I.; Coelho, H.; Mirabella, S.; Poveda, A.; Arda, A.; Jimenez-Barbero, J. Glycans in Infectious Diseases. A Molecular Recognition Perspective. Curr. Med. Chem. 2017, 24, 4057–4080. [Google Scholar] [CrossRef] [PubMed]
  92. Reyes, C.D.G.; Onigbinde, S.; Sanni, A.; Bennett, A.I.; Jiang, P.; Daramola, O.; Ahmadi, P.; Fowowe, M.; Atashi, M.; Sandilya, V.; et al. N-Glycome Profile of the Spike Protein S1: Systemic and Comparative Analysis from Eleven Variants of SARS-CoV-2. Biomolecules 2023, 13, 1421. [Google Scholar] [CrossRef]
  93. Freitas, D.; Campos, D.; Gomes, J.; Pinto, F.; Macedo, J.A.; Matos, R.; Mereiter, S.; Pinto, M.T.; Polónia, A.; Gartner, F.; et al. O-glycans truncation modulates gastric cancer cell signaling and transcription leading to a more aggressive phenotype. EBioMedicine 2019, 40, 349–362. [Google Scholar] [CrossRef] [PubMed]
  94. Amon, R.; Reuven, E.M.; Leviatan Ben-Arye, S.; Padler-Karavani, V. Glycans in immune recognition and response. Carbohydr. Res. 2014, 389, 115–122. [Google Scholar] [CrossRef] [PubMed]
  95. Liwosz, A.; Lei, T.; Kukuruzinska, M.A. N-glycosylation affects the molecular organization and stability of E-cadherin junctions. J. Biol. Chem. 2006, 281, 23138–23149. [Google Scholar] [CrossRef] [PubMed]
  96. Rodrigues, J.G.; Balmaña, M.; Macedo, J.A.; Poças, J.; Fernandes, Â.; de-Freitas-Junior, J.C.M.; Pinho, S.S.; Gomes, J.; Magalhães, A.; Gomes, C.; et al. Glycosylation in cancer: Selected roles in tumour progression, immune modulation and metastasis. Cell Immunol. 2018, 333, 46–57. [Google Scholar] [CrossRef] [PubMed]
  97. Zhang, Q.; Ma, C.; Chin, L.S.; Li, L. Integrative glycoproteomics reveals protein N-glycosylation aberrations and glycoproteomic network alterations in Alzheimer’s disease. Sci. Adv. 2020, 6, eabc5802. [Google Scholar] [CrossRef] [PubMed]
  98. Bennun, S.V.; Hizal, D.B.; Heffner, K.; Can, O.; Zhang, H.; Betenbaugh, M.J. Systems Glycobiology: Integrating Glycogenomics, Glycoproteomics, Glycomics, and Other ‘Omics Data Sets to Characterize Cellular Glycosylation Processes. J. Mol. Biol. 2016, 428, 3337–3352. [Google Scholar] [CrossRef] [PubMed]
  99. Peixoto, A.; Relvas-Santos, M.; Azevedo, R.; Santos, L.L.; Ferreira, J.A. Protein Glycosylation and Tumor Microenvironment Alterations Driving Cancer Hallmarks. Front. Oncol. 2019, 9, 380. [Google Scholar] [CrossRef]
  100. Guo, S.; Zhang, C.; Le, A. The limitless applications of single-cell metabolomics. Curr. Opin. Biotechnol. 2021, 71, 115–122. [Google Scholar] [CrossRef]
  101. Blanco, A.; Blanco, G. Chapter 5—Lipids. In Medical Biochemistry, 2nd ed.; Blanco, A., Blanco, G., Eds.; Academic Press: Cambridge, MA, USA, 2022; pp. 105–129. [Google Scholar]
  102. Li, Z.; Cheng, S.; Lin, Q.; Cao, W.; Yang, J.; Zhang, M.; Shen, A.; Zhang, W.; Xia, Y.; Ma, X.; et al. Single-cell lipidomics with high structural specificity by mass spectrometry. Nat. Commun. 2021, 12, 2869. [Google Scholar] [CrossRef]
  103. Yang, K.; Han, X. Lipidomics: Techniques, Applications, and Outcomes Related to Biomedical Sciences. Trends Biochem. Sci. 2016, 41, 954–969. [Google Scholar] [CrossRef] [PubMed]
  104. Hoang, G.; Udupa, S.; Le, A. Chapter Six—Application of metabolomics technologies toward cancer prognosis and therapy. In International Review of Cell and Molecular Biology; Montrose, D.C., Galluzzi, L., Eds.; Academic Press: Cambridge, MA, USA, 2019; Volume 347, pp. 191–223. [Google Scholar]
  105. Wang, Z.; Zhu, H.; Xiong, W. Advances in mass spectrometry-based multi-scale metabolomic methodologies and their applications in biological and clinical investigations. Sci. Bull. 2023, 68, 2268–2284. [Google Scholar] [CrossRef] [PubMed]
  106. Zhang, Z.; Bao, C.; Jiang, L.; Wang, S.; Wang, K.; Lu, C.; Fang, H. When cancer drug resistance meets metabolomics (bulk, single-cell and/or spatial): Progress, potential, and perspective. Front. Oncol. 2022, 12, 1054233. [Google Scholar] [CrossRef] [PubMed]
  107. Hanahan, D. Hallmarks of Cancer: New Dimensions. Cancer Discov. 2022, 12, 31–46. [Google Scholar] [CrossRef] [PubMed]
  108. Bhat, G.R.; Sethi, I.; Sadida, H.Q.; Rah, B.; Mir, R.; Algehainy, N.; Albalawi, I.A.; Masoodi, T.; Subbaraj, G.K.; Jamal, F.; et al. Cancer cell plasticity: From cellular, molecular, and genetic mechanisms to tumor heterogeneity and drug resistance. Cancer Metastasis Rev. 2024, 43, 197–228. [Google Scholar] [CrossRef] [PubMed]
  109. Zhang, W.; Xu, F.; Yao, J.; Mao, C.; Zhu, M.; Qian, M.; Hu, J.; Zhong, H.; Zhou, J.; Shi, X.; et al. Single-cell metabolic fingerprints discover a cluster of circulating tumor cells with distinct metastatic potential. Nat. Commun. 2023, 14, 2485. [Google Scholar] [CrossRef]
  110. Zhu, G.; Zhang, W.; Zhao, Y.; Chen, T.; Yuan, H.; Liu, Y.; Guo, G.; Liu, Z.; Wang, X. Single-Cell Metabolomics-Based Strategy for Studying the Mechanisms of Drug Action. Anal. Chem. 2023, 95, 4712–4720. [Google Scholar] [CrossRef]
  111. Zeng, Q.; Mousa, M.; Nadukkandy, A.S.; Franssens, L.; Alnaqbi, H.; Alshamsi, F.Y.; Safar, H.A.; Carmeliet, P. Understanding tumour endothelial cell heterogeneity and function from single-cell omics. Nat. Rev. Cancer 2023, 23, 544–564. [Google Scholar] [CrossRef]
  112. Chen, X.; Peng, Z.; Yang, Z. Metabolomics studies of cell–cell interactions using single cell mass spectrometry combined with fluorescence microscopy. Chem. Sci. 2022, 13, 6687–6695. [Google Scholar] [CrossRef]
  113. León-Letelier, R.A.; Dou, R.; Vykoukal, J.; Yip-Schneider, M.T.; Maitra, A.; Irajizad, E.; Wu, R.; Dennison, J.B.; Do, K.-A.; Zhang, J.; et al. Contributions of the Microbiome-Derived Metabolome for Risk Assessment and Prognostication of Pancreatic Cancer. Clin. Chem. 2024, 70, 102–115. [Google Scholar] [CrossRef] [PubMed]
  114. Zahra, W.; Rai, S.N.; Birla, H.; Singh, S.S.; Dilnashin, H.; Rathore, A.S.; Singh, S.P. The Global Economic Impact of Neurodegenerative Diseases: Opportunities and Challenges. In Bioeconomy for Sustainable Development; Keswani, C., Ed.; Springer: Singapore, 2020; pp. 333–345. [Google Scholar]
  115. Zheng, Y.; Liu, Z.; Xing, J.; Zheng, Z.; Pi, Z.; Song, F.; Liu, S. In situ analysis of single cell and biological samples with rGO-Cu functional probe ESI-MS spectrometry. Talanta 2020, 211, 120751. [Google Scholar] [CrossRef] [PubMed]
  116. Bao, L.; Wu, B.; Yang, C.; Xu, X.; Shi, Z.; Jiang, D. Targeted Metabolomics Analysis of Serum Amino Acids in T2DM Patients. Diabetes Metab. Syndr. Obes. 2024, 17, 203–212. [Google Scholar] [CrossRef] [PubMed]
  117. Yan, H.; Li, G.; Zhang, X.; Zhang, C.; Li, M.; Qiu, Y.; Sun, W.; Dong, Y.; Li, S.; Li, J. Targeted metabolomics-based understanding of the sleep disturbances in drug-naïve patients with schizophrenia. BMC Psychiatry 2024, 24, 355. [Google Scholar] [CrossRef] [PubMed]
  118. Lee, E.-J.; Kim, D.J.; Kang, D.-W.; Yang, W.; Jeong, H.-Y.; Kim, J.-M.; Ko, S.-B.; Lee, S.-H.; Yoon, B.-W.; Cho, J.-Y.; et al. Targeted Metabolomic Biomarkers for Stroke Subtyping. Transl. Stroke Res. 2024, 15, 422–432. [Google Scholar] [CrossRef] [PubMed]
  119. Wang, S.; He, T.; Wang, H. Non-targeted metabolomics study for discovery of hepatocellular carcinoma serum diagnostic biomarker. J. Pharm. Biomed. Anal. 2024, 239, 115869. [Google Scholar] [CrossRef]
  120. Sun, S.; Chen, M.; Zhang, T.; Wang, Y.; Shen, W.; Zhang, T.; Liu, J.; Lan, H.; Zhao, J.; Lin, F.; et al. Identification of Key Factors in Cartilage Tissue During the Progression of Osteoarthritis Using a Non-targeted Metabolomics Strategy. Phenomics 2024. [Google Scholar] [CrossRef]
  121. Chen, W.; Guo, W.; Li, Y.; Chen, M. Integrative analysis of metabolomics and transcriptomics to uncover biomarkers in sepsis. Sci. Rep. 2024, 14, 9676. [Google Scholar] [CrossRef] [PubMed]
  122. Wang, S.; Liu, G.; Xie, C.; Zhou, Y.; Yang, R.; Wu, J.; Xu, J.; Tu, K. Metabolomics Analysis of Different Quinoa Cultivars Based on UPLC-ZenoTOF-MS/MS and Investigation into Their Antioxidant Characteristics. Plants 2024, 13, 240. [Google Scholar] [CrossRef]
  123. Windarsih, A.; Arifah, M.F.; Utami, I.D.; Suratno; Rohman, A. Detection of goat milk adulteration in horse milk using LC-HRMS-based non-targeted metabolomics and chemometrics. Chem. Pap. 2024, 78, 809–821. [Google Scholar] [CrossRef]
  124. Tian, Y.; Wang, W.; Jiang, W.; Zhang, G.; He, J.; Dong, S.; Zhou, Y.; Yang, W.; Tang, Q.; Yu, Y.; et al. Non-targeted metabolomics provides insights into the distinct amino acid and lipid metabolism in liver tissues of rainbow trout (Oncorhynchus mykiss) cultured in seawater at different temperatures. Aquaculture 2024, 579, 740188. [Google Scholar] [CrossRef]
  125. Yang, S.; Pathak, S.; Tang, H.; Zhang, D.; Chen, Y.; Ntezimana, B.; Ni, D.; Yu, Z. Non-Targeted Metabolomics Reveals the Effects of Different Rolling Methods on Black Tea Quality. Foods 2024, 13, 325. [Google Scholar] [CrossRef] [PubMed]
  126. Feng, X.; Ma, R.; Wang, Y.; Tong, L.; Wen, W.; Mu, T.; Tian, J.; Yu, B.; Gu, Y.; Zhang, J. Non-targeted metabolomics identifies biomarkers in milk with high and low milk fat percentage. Food Res. Int. 2024, 179, 113989. [Google Scholar] [CrossRef] [PubMed]
  127. Mo, L.; Fang, L.; Yao, W.; Nie, J.; Dai, J.; Liang, Y.; Qin, L. LC-QTOF/MS-based non-targeted metabolomics to explore the toxic effects of di(2-ethylhexyl) phthalate (DEHP) on Brassica chinensis L. Sci. Total Environ. 2024, 918, 170817. [Google Scholar] [CrossRef] [PubMed]
  128. Song, Z.; Wu, H.; Fang, X.; Feng, X.; Zhou, L. The cardiovascular toxicity of polystyrene microplastics in rats: Based on untargeted metabolomics analysis. Front. Pharmacol. 2024, 15, 1336369. [Google Scholar] [CrossRef]
  129. Xiao, X.; Zhou, Y.; Li, X.; Jin, J.; Durham, J.; Ye, Z.; Wang, Y.; Hennig, B.; Deng, P. 13C-Stable isotope resolved metabolomics uncovers dynamic biochemical landscape of gut microbiome-host organ communications in mice. Microbiome 2024, 12, 90. [Google Scholar] [CrossRef] [PubMed]
  130. Wen, S.; Tu, X.; Zang, Q.; Zhu, Y.; Li, L.; Zhang, R.; Abliz, Z. Liquid chromatography–mass spectrometry-based metabolomics and fluxomics reveals the metabolic alterations in glioma U87MG multicellular tumor spheroids versus two-dimensional cell cultures. Rapid Commun. Mass. Spectrom. 2024, 38, e9670. [Google Scholar] [CrossRef]
  131. Peper, C.J.; Kilgore, M.D.; Jiang, Y.; Xiu, Y.; Xia, W.; Wang, Y.; Shi, M.; Zhou, D.; Dumont, A.S.; Wang, X.; et al. Tracing the path of disruption: 13C isotope applications in traumatic brain injury-induced metabolic dysfunction. CNS Neurosci. Ther. 2024, 30, e14693. [Google Scholar] [CrossRef]
  132. Osipenko, S.; Bashilov, A.; Vishnevskaya, A.; Rumiantseva, L.; Levashova, A.; Kovalenko, A.; Tupertsev, B.; Kireev, A.; Nikolaev, E.; Kostyukevich, Y. Investigating the Metabolism of Plants Germinated in Heavy Water, D2O, and H218O-Enriched Media Using High-Resolution Mass Spectrometry. Int. J. Mol. Sci. 2023, 24, 15396. [Google Scholar] [CrossRef]
  133. Brorsen, L.F.; McKenzie, J.S.; Tullin, M.F.; Bendtsen, K.M.S.; Pinto, F.E.; Jensen, H.E.; Haedersdal, M.; Takats, Z.; Janfelt, C.; Lerche, C.M. Cutaneous squamous cell carcinoma characterized by MALDI mass spectrometry imaging in combination with machine learning. Sci. Rep. 2024, 14, 11091. [Google Scholar] [CrossRef]
  134. Bag, S.; Oetjen, J.; Shaikh, S.; Chaudhary, A.; Arun, P.; Mukherjee, G. Impact of spatial metabolomics on immune-microenvironment in oral cancer prognosis: A clinical report. Mol. Cell. Biochem. 2024, 479, 41–49. [Google Scholar] [CrossRef] [PubMed]
  135. Vermeulen, I.; Rodriguez-Alvarez, N.; François, L.; Viot, D.; Poosti, F.; Aronica, E.; Dedeurwaerdere, S.; Barton, P.; Cillero-Pastor, B.; Heeren, R.M.A. Spatial omics reveals molecular changes in focal cortical dysplasia type II. Neurobiol. Dis. 2024, 195, 106491. [Google Scholar] [CrossRef] [PubMed]
  136. Wang, J.; Sun, N.; Kunzke, T.; Shen, J.; Zens, P.; Prade, V.M.; Feuchtinger, A.; Berezowska, S.; Walch, A. Spatial metabolomics identifies distinct tumor-specific and stroma-specific subtypes in patients with lung squamous cell carcinoma. Npj Precis. Oncol. 2023, 7, 114. [Google Scholar] [CrossRef] [PubMed]
  137. Conroy, L.R.; Clarke, H.A.; Allison, D.B.; Valenca, S.S.; Sun, Q.; Hawkinson, T.R.; Young, L.E.A.; Ferreira, J.E.; Hammonds, A.V.; Dunne, J.B.; et al. Spatial metabolomics reveals glycogen as an actionable target for pulmonary fibrosis. Nat. Commun. 2023, 14, 2759. [Google Scholar] [CrossRef] [PubMed]
  138. Züllig, T.; Trötzmüller, M.; Köfeler, H.C. Lipidomics from sample preparation to data analysis: A primer. Anal. Bioanal. Chem. 2020, 412, 2191–2209. [Google Scholar] [CrossRef]
  139. Kenwood, B.M.; Merrill, A.H. Lipidomics. In Encyclopedia of Cell Biology; Bradshaw, R.A., Stahl, P.D., Eds.; Academic Press: Waltham, MA, USA, 2016; pp. 147–159. [Google Scholar]
  140. Zhao, Y.-Y.; Cheng, X.-l.; Lin, R.-C. Chapter One—Lipidomics Applications for Discovering Biomarkers of Diseases in Clinical Chemistry. In International Review of Cell and Molecular Biology; Jeon, K.W., Ed.; Academic Press: Cambridge, MA, USA, 2014; Volume 313, pp. 1–26. [Google Scholar]
  141. Costa, A.C.; Riça, L.B.; van de Bilt, M.; Zandonadi, F.S.; Gattaz, W.F.; Talib, L.L.; Sussulini, A. Application of Lipidomics in Psychiatry: Plasma-Based Potential Biomarkers in Schizophrenia and Bipolar Disorder. Metabolites 2023, 13, 600. [Google Scholar] [CrossRef]
  142. Chiurchiù, V.; Tiberi, M.; Matteocci, A.; Fazio, F.; Siffeti, H.; Saracini, S.; Mercuri, N.B.; Sancesario, G. Lipidomics of Bioactive Lipids in Alzheimer’s and Parkinson’s Diseases: Where Are We? Int. J. Mol. Sci. 2022, 23, 6235. [Google Scholar] [CrossRef]
  143. Graham, Z.A.; Siedlik, J.A.; Toro, C.A.; Harlow, L.; Cardozo, C.P. Boldine Alters Serum Lipidomic Signatures after Acute Spinal Cord Transection in Male Mice. Int. J. Environ. Res. Public Health 2023, 20, 6591. [Google Scholar] [CrossRef]
  144. Senko, D.; Gorovaya, A.; Stekolshchikova, E.; Anikanov, N.; Fedianin, A.; Baltin, M.; Efimova, O.; Petrova, D.; Baltina, T.; Lebedev, M.A.; et al. Time-Dependent Effect of Sciatic Nerve Injury on Rat Plasma Lipidome. Int. J. Mol. Sci. 2022, 23, 15544. [Google Scholar] [CrossRef]
  145. Meikle, P.J.; Wong, G.; Barlow, C.K.; Kingwell, B.A. Lipidomics: Potential role in risk prediction and therapeutic monitoring for diabetes and cardiovascular disease. Pharmacol. Ther. 2014, 143, 12–23. [Google Scholar] [CrossRef]
  146. Rasmiena, A.A.; Ng, T.W.; Meikle, P.J. Metabolomics and ischaemic heart disease. Clin. Sci. 2012, 124, 289–306. [Google Scholar] [CrossRef] [PubMed]
  147. Perrotti, F.; Rosa, C.; Cicalini, I.; Sacchetta, P.; Del Boccio, P.; Genovesi, D.; Pieragostino, D. Advances in Lipidomics for Cancer Biomarkers Discovery. Int. J. Mol. Sci. 2016, 17, 1992. [Google Scholar] [CrossRef] [PubMed]
  148. Horgusluoglu, E.; Neff, R.; Song, W.-M.; Wang, M.; Wang, Q.; Arnold, M.; Krumsiek, J.; Galindo-Prieto, B.; Ming, C.; Nho, K.; et al. Integrative metabolomics-genomics approach reveals key metabolic pathways and regulators of Alzheimer’s disease. Alzheimer’s Dement. 2022, 18, 1260–1278. [Google Scholar] [CrossRef] [PubMed]
  149. Bernardes, J.P.; Mishra, N.; Tran, F.; Bahmer, T.; Best, L.; Blase, J.I.; Bordoni, D.; Franzenburg, J.; Geisen, U.; Josephs-Spaulding, J.; et al. Longitudinal Multi-omics Analyses Identify Responses of Megakaryocytes, Erythroid Cells, and Plasmablasts as Hallmarks of Severe COVID-19. Immunity 2020, 53, 1296–1314.e1299. [Google Scholar] [CrossRef] [PubMed]
  150. Xing, X.; Hu, E.; Ouyang, J.; Zhong, X.; Wang, F.; Liu, K.; Cai, L.; Zhou, Y.; Wang, Y.; Chen, G.; et al. Integrated omics landscape of hepatocellular carcinoma suggests proteomic subtypes for precision therapy. Cell Rep. Med. 2023, 4, 101315. [Google Scholar] [CrossRef] [PubMed]
  151. Liu, N.; Wu, Y.; Cheng, W.; Wu, Y.; Wang, L.; Zhuang, L. Identification of novel prognostic biomarkers by integrating multi-omics data in gastric cancer. BMC Cancer 2021, 21, 460. [Google Scholar] [CrossRef] [PubMed]
  152. Clark, C.; Dayon, L.; Masoodi, M.; Bowman, G.L.; Popp, J. An integrative multi-omics approach reveals new central nervous system pathway alterations in Alzheimer’s disease. Alzheimer’s Res. Ther. 2021, 13, 71. [Google Scholar] [CrossRef] [PubMed]
  153. Strefeler, A.; Jan, M.; Quadroni, M.; Teav, T.; Rosenberg, N.; Chatton, J.-Y.; Guex, N.; Gallart-Ayala, H.; Ivanisevic, J. Molecular insights into sex-specific metabolic alterations in Alzheimer’s mouse brain using multi-omics approach. Alzheimer’s Res. Ther. 2023, 15, 8. [Google Scholar] [CrossRef] [PubMed]
  154. Lewis, J.E.; Kemp, M.L. Integration of machine learning and genome-scale metabolic modeling identifies multi-omics biomarkers for radiation resistance. Nat. Commun. 2021, 12, 2700. [Google Scholar] [CrossRef]
  155. Nativio, R.; Lan, Y.; Donahue, G.; Sidoli, S.; Berson, A.; Srinivasan, A.R.; Shcherbakova, O.; Amlie-Wolf, A.; Nie, J.; Cui, X.; et al. An integrated multi-omics approach identifies epigenetic alterations associated with Alzheimer’s disease. Nat. Genet. 2020, 52, 1024–1035. [Google Scholar] [CrossRef]
  156. Heintz-Buschart, A.; May, P.; Laczny, C.C.; Lebrun, L.A.; Bellora, C.; Krishna, A.; Wampach, L.; Schneider, J.G.; Hogan, A.; de Beaufort, C.; et al. Integrated multi-omics of the human gut microbiome in a case study of familial type 1 diabetes. Nat. Microbiol. 2016, 2, 16180. [Google Scholar] [CrossRef] [PubMed]
  157. Lupberger, J.; Croonenborghs, T.; Roca Suarez, A.A.; Van Renne, N.; Jühling, F.; Oudot, M.A.; Virzì, A.; Bandiera, S.; Jamey, C.; Meszaros, G.; et al. Combined Analysis of Metabolomes, Proteomes, and Transcriptomes of Hepatitis C Virus–Infected Cells and Liver to Identify Pathways Associated With Disease Development. Gastroenterology 2019, 157, 537–551.e539. [Google Scholar] [CrossRef]
  158. Hudson, K.M.; Shiver, E.; Yu, J.; Mehta, S.; Jima, D.D.; Kane, M.A.; Patisaul, H.B.; Cowley, M. Transcriptomic, proteomic, and metabolomic analyses identify candidate pathways linking maternal cadmium exposure to altered neurodevelopment and behavior. Sci. Rep. 2021, 11, 16302. [Google Scholar] [CrossRef] [PubMed]
  159. Zheng, P.; Zhang, N.; Ren, D.; Yu, C.; Zhao, B.; Zhang, Y. Integrated spatial transcriptome and metabolism study reveals metabolic heterogeneity in human injured brain. Cell Rep. Med. 2023, 4, 101057. [Google Scholar] [CrossRef]
  160. Vicari, M.; Mirzazadeh, R.; Nilsson, A.; Shariatgorji, R.; Bjärterot, P.; Larsson, L.; Lee, H.; Nilsson, M.; Foyer, J.; Ekvall, M.; et al. Spatial multimodal analysis of transcriptomes and metabolomes in tissues. Nat. Biotechnol. 2023. [Google Scholar] [CrossRef] [PubMed]
  161. Lopez Soriano, V.; Dueñas Rey, A.; Mukherjee, R.; Inglehearn, C.F.; Coppieters, F.; Bauwens, M.; Willaert, A.; De Baere, E.; Genomics England Research, C. Multi-omics analysis in human retina uncovers ultraconserved cis-regulatory elements at rare eye disease loci. Nat. Commun. 2024, 15, 1600. [Google Scholar] [CrossRef]
  162. Pattaroni, C.; Begka, C.; Cardwell, B.; Jaffar, J.; Macowan, M.; Harris, N.L.; Westall, G.P.; Marsland, B.J. Multi-omics integration reveals a nonlinear signature that precedes progression of lung fibrosis. Clin. Transl. Immunol. 2024, 13, e1485. [Google Scholar] [CrossRef]
  163. Garcia-Segura, M.E.; Durainayagam, B.R.; Liggi, S.; Graça, G.; Jimenez, B.; Dehghan, A.; Tzoulaki, I.; Karaman, I.; Elliott, P.; Griffin, J.L. Pathway-based integration of multi-omics data reveals lipidomics alterations validated in an Alzheimer’s disease mouse model and risk loci carriers. J. Neurochem. 2023, 164, 57–76. [Google Scholar] [CrossRef]
  164. Qiu, C.; Yu, F.; Su, K.; Zhao, Q.; Zhang, L.; Xu, C.; Hu, W.; Wang, Z.; Zhao, L.; Tian, Q.; et al. Multi-omics Data Integration for Identifying Osteoporosis Biomarkers and Their Biological Interaction and Causal Mechanisms. iScience 2020, 23, 100847. [Google Scholar] [CrossRef]
  165. Byeon, S.K.; Madugundu, A.K.; Garapati, K.; Ramarajan, M.G.; Saraswat, M.; Kumar, M.P.; Hughes, T.; Shah, R.; Patnaik, M.M.; Chia, N.; et al. Development of a multiomics model for identification of predictive biomarkers for COVID-19 severity: A retrospective cohort study. Lancet Digit. Health 2022, 4, e632–e645. [Google Scholar] [CrossRef]
  166. Huang, C.; You, Z.; He, Y.; Li, J.; Liu, Y.; Peng, C.; Liu, Z.; Liu, X.; Sun, J. Combined transcriptomics and proteomics forecast analysis for potential biomarker in the acute phase of temporal lobe epilepsy. Front. Neurosci. 2023, 17, 1145805. [Google Scholar] [CrossRef] [PubMed]
  167. Suryadevara, R.; Gregory, A.; Lu, R.; Xu, Z.; Masoomi, A.; Lutz, S.M.; Berman, S.; Yun, J.H.; Saferali, A.; Ryu, M.H.; et al. Blood-based Transcriptomic and Proteomic Biomarkers of Emphysema. Am. J. Respir. Crit. Care Med. 2024, 209, 273–287. [Google Scholar] [CrossRef] [PubMed]
  168. Sivakumar, P.; Ammar, R.; Thompson, J.R.; Luo, Y.; Streltsov, D.; Porteous, M.; McCoubrey, C.; Cantu, E.; Beers, M.F.; Jarai, G.; et al. Integrated plasma proteomics and lung transcriptomics reveal novel biomarkers in idiopathic pulmonary fibrosis. Respir. Res. 2021, 22, 273. [Google Scholar] [CrossRef] [PubMed]
  169. Zhou, Y.; Fang, J.; Bekris, L.M.; Kim, Y.H.; Pieper, A.A.; Leverenz, J.B.; Cummings, J.; Cheng, F. AlzGPS: A genome-wide positioning systems platform to catalyze multi-omics for Alzheimer’s drug discovery. Alzheimer’s Res. Ther. 2021, 13, 24. [Google Scholar] [CrossRef] [PubMed]
  170. Wang, Z.; Li, Y.; Zhao, W.; Jiang, S.; Huang, Y.; Hou, J.; Zhang, X.; Zhai, Z.; Yang, C.; Wang, J.; et al. Integrative multi-omics and drug–response characterization of patient-derived prostate cancer primary cells. Signal Transduct. Target. Ther. 2023, 8, 175. [Google Scholar] [CrossRef] [PubMed]
  171. Leo, I.R.; Aswad, L.; Stahl, M.; Kunold, E.; Post, F.; Erkers, T.; Struyf, N.; Mermelekas, G.; Joshi, R.N.; Gracia-Villacampa, E.; et al. Integrative multi-omics and drug response profiling of childhood acute lymphoblastic leukemia cell lines. Nat. Commun. 2022, 13, 1691. [Google Scholar] [CrossRef] [PubMed]
  172. Kim, M.S.; Song, M.; Kim, B.; Shim, I.; Kim, D.S.; Natarajan, P.; Do, R.; Won, H.-H. Prioritization of therapeutic targets for dyslipidemia using integrative multi-omics and multi-trait analysis. Cell Rep. Med. 2023, 4, 101112. [Google Scholar] [CrossRef] [PubMed]
  173. Leinonen, R.; Sugawara, H.; Shumway, M.; International Nucleotide Sequence Database Collaboration. The Sequence Read Archive. Nucleic Acids Res. 2010, 39, D19–D21. [Google Scholar] [CrossRef]
  174. Barrett, T.; Wilhite, S.E.; Ledoux, P.; Evangelista, C.; Kim, I.F.; Tomashevsky, M.; Marshall, K.A.; Phillippy, K.H.; Sherman, P.M.; Holko, M.; et al. NCBI GEO: Archive for functional genomics data sets—Update. Nucleic Acids Res. 2012, 41, D991–D995. [Google Scholar] [CrossRef]
  175. Clark, K.; Karsch-Mizrachi, I.; Lipman, D.J.; Ostell, J.; Sayers, E.W. GenBank. Nucleic Acids Res. 2015, 44, D67–D72. [Google Scholar] [CrossRef]
  176. Wang, G.; Wu, S.; Xiong, Z.; Qu, H.; Fang, X.; Bao, Y. CROST: A comprehensive repository of spatial transcriptomics. Nucleic Acids Res. 2023, 52, D882–D890. [Google Scholar] [CrossRef] [PubMed]
  177. Vizcaíno, J.A.; Côté, R.; Reisinger, F.; Foster, J.M.; Mueller, M.; Rameseder, J.; Hermjakob, H.; Martens, L. A guide to the Proteomics Identifications Database proteomics data repository. Proteomics 2009, 9, 4276–4283. [Google Scholar] [CrossRef] [PubMed]
  178. Deutsch, E.W.; Bandeira, N.; Sharma, V.; Perez-Riverol, Y.; Carver, J.J.; Kundu, D.J.; García-Seisdedos, D.; Jarnuczak, A.F.; Hewapathirana, S.; Pullman, B.S.; et al. The ProteomeXchange consortium in 2020: Enabling ‘big data’ approaches in proteomics. Nucleic Acids Res 2020, 48, D1145–D1152. [Google Scholar] [CrossRef] [PubMed]
  179. Watanabe, Y.; Aoki-Kinoshita, K.F.; Ishihama, Y.; Okuda, S. GlycoPOST realizes FAIR principles for glycomics mass spectrometry data. Nucleic Acids Res. 2020, 49, D1523–D1528. [Google Scholar] [CrossRef] [PubMed]
  180. 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]
  181. Wishart, D.S.; Tzur, D.; Knox, C.; Eisner, R.; Guo, A.C.; Young, N.; Cheng, D.; Jewell, K.; Arndt, D.; Sawhney, S.; et al. HMDB: The Human Metabolome Database. Nucleic Acids Res. 2007, 35, D521–D526. [Google Scholar] [CrossRef] [PubMed]
  182. Conesa, A.; Beck, S. Making multi-omics data accessible to researchers. Sci. Data 2019, 6, 251. [Google Scholar] [CrossRef] [PubMed]
  183. Guo, X.; Chen, F.; Gao, F.; Li, L.; Liu, K.; You, L.; Hua, C.; Yang, F.; Liu, W.; Peng, C.; et al. CNSA: A data repository for archiving omics data. Database 2020, 2020, baaa055. [Google Scholar] [CrossRef] [PubMed]
  184. Athieniti, E.; Spyrou, G.M. A guide to multi-omics data collection and integration for translational medicine. Comput. Struct. Biotechnol. J. 2023, 21, 134–149. [Google Scholar] [CrossRef]
  185. Yamada, R.; Okada, D.; Wang, J.; Basak, T.; Koyama, S. Interpretation of omics data analyses. J. Hum. Genet. 2021, 66, 93–102. [Google Scholar] [CrossRef]
  186. Dunkler, D.; Sánchez-Cabo, F.; Heinze, G. Statistical analysis principles for Omics data. Methods Mol. Biol. 2011, 719, 113–131. [Google Scholar] [CrossRef] [PubMed]
  187. Rouam, S. False Discovery Rate (FDR). In Encyclopedia of Systems Biology; Dubitzky, W., Wolkenhauer, O., Cho, K.-H., Yokota, H., Eds.; Springer: New York, NY, USA, 2013; pp. 731–732. [Google Scholar]
  188. Krämer, A.; Green, J.; Pollard, J., Jr.; Tugendreich, S. Causal analysis approaches in Ingenuity Pathway Analysis. Bioinformatics 2014, 30, 523–530. [Google Scholar] [CrossRef] [PubMed]
  189. Ahsan, S.; Drăghici, S. Identifying Significantly Impacted Pathways and Putative Mechanisms with iPathwayGuide. Curr. Protoc. Bioinform. 2017, 57, 7.15.11–17.15.30. [Google Scholar] [CrossRef]
  190. Wu, L.; Liu, F.; Cai, H. IOAT: An interactive tool for statistical analysis of omics data and clinical data. BMC Bioinform. 2021, 22, 326. [Google Scholar] [CrossRef]
  191. Buy, M.; Digan, W.; Chen, X.; Husson, J.; Ménager, M.; Rieux-Laucat, F.; Garcelon, N. A Multi-Omics Common Data Model for Primary Immunodeficiencies. Stud. Health Technol. Inform. 2022, 290, 56–60. [Google Scholar] [CrossRef]
  192. Harbig, T.A.; Fratte, J.; Krone, M.; Nieselt, K. OmicsTIDE: Interactive exploration of trends in multi-omics data. Bioinform. Adv. 2023, 3, vbac093. [Google Scholar] [CrossRef]
  193. Fang, Z.; Ma, T.; Tang, G.; Zhu, L.; Yan, Q.; Wang, T.; Celedón, J.C.; Chen, W.; Tseng, G.C. Bayesian integrative model for multi-omics data with missingness. Bioinformatics 2018, 34, 3801–3808. [Google Scholar] [CrossRef]
  194. Ma, C.; Wu, M.; Ma, S. Analysis of cancer omics data: A selective review of statistical techniques. Brief. Bioinform. 2022, 23, bbab585. [Google Scholar] [CrossRef] [PubMed]
  195. Das, S.; Mukhopadhyay, I. TiMEG: An integrative statistical method for partially missing multi-omics data. Sci. Rep. 2021, 11, 24077. [Google Scholar] [CrossRef] [PubMed]
  196. Cifuentes, A. Food analysis and Foodomics. J. Chromatogr. A 2009, 1216, 7109. [Google Scholar] [CrossRef]
  197. Picó, C.; Serra, F.; Rodríguez, A.M.; Keijer, J.; Palou, A. Biomarkers of Nutrition and Health: New Tools for New Approaches. Nutrients 2019, 11, 1092. [Google Scholar] [CrossRef] [PubMed]
  198. Hill, E.B.; Tang, M.; Long, J.M.; Kemp, J.F.; Westcott, J.L.; Hendricks, A.E.; Reisdorph, N.A.; Campbell, W.W.; Krebs, N.F.; Schaefer, C.; et al. mini-MED: Study protocol for a randomized, multi-intervention, semi-controlled feeding trial of a Mediterranean-amplified vs. habitual Western dietary pattern for the evaluation of food-specific compounds and cardiometabolic health. Trials 2024, 25, 101. [Google Scholar] [CrossRef] [PubMed]
  199. Trimigno, A.; Khakimov, B.; Savorani, F.; Poulsen, S.K.; Astrup, A.; Dragsted, L.O.; Engelsen, S.B. Human urine 1H NMR metabolomics reveals alterations of protein and carbohydrate metabolism when comparing habitual Average Danish diet vs. healthy New Nordic diet. Nutrition 2020, 79–80, 110867. [Google Scholar] [CrossRef] [PubMed]
  200. Dilmore, A.H.; Martino, C.; Neth, B.J.; West, K.A.; Zemlin, J.; Rahman, G.; Panitchpakdi, M.; Meehan, M.J.; Weldon, K.C.; Blach, C.; et al. Effects of a ketogenic and low-fat diet on the human metabolome, microbiome, and foodome in adults at risk for Alzheimer’s disease. Alzheimer’s Dement. 2023, 19, 4805–4816. [Google Scholar] [CrossRef]
Figure 1. Multi omics. Inside the eukaryotic nucleus, spatial conformations of genomic material and interactions participate in regulating gene expression. Different RNAs are accumulated, and some splice variants are produced in response to internal and external stimuli. Transcripts can be rapidly degraded or preserved and can sustain high or low translation rates. Post translational modifications provide different protein functions that can modify ligand affinity, protein interactions, catalytic activity, etc. The cellular regulatory processes favor synthesis, degradation, modification, and interaction of a multitude of molecules. The right section of the figure shows high-throughput sequencing methods and technologies to measure gene expression in situ, and the common LC-MS/MS label-free quantitation “discovery stage” followed by the targeted “validation stage”.
Figure 1. Multi omics. Inside the eukaryotic nucleus, spatial conformations of genomic material and interactions participate in regulating gene expression. Different RNAs are accumulated, and some splice variants are produced in response to internal and external stimuli. Transcripts can be rapidly degraded or preserved and can sustain high or low translation rates. Post translational modifications provide different protein functions that can modify ligand affinity, protein interactions, catalytic activity, etc. The cellular regulatory processes favor synthesis, degradation, modification, and interaction of a multitude of molecules. The right section of the figure shows high-throughput sequencing methods and technologies to measure gene expression in situ, and the common LC-MS/MS label-free quantitation “discovery stage” followed by the targeted “validation stage”.
Cimb 46 00345 g001
Table 1. Complemental integrative omics applications.
Table 1. Complemental integrative omics applications.
Omics
Spotlight
Omics
Integration
ApplicationsReferences
Diagnosis and
prognosis
G+T+P+MAlzheimer’s disease diagnosis[148]
T+Me+scTEvaluation of Covid-19 prognosis[149]
G+T+P+pPHepatocellular carcinoma classification[150]
G+TIdentification of prognostic biomarkers for gastric cancer[151]
P+M+LSearch for pathway alterations in Alzheimer’s disease[152,153]
G+T+MIdentification of metabolite biomarkers for predicting radiation resistance[154]
T+P+EEpigenetic alterations associated with Alzheimer’s disease[155]
G+T+PGastrointestinal microbiome and diabetes mellitus (type 1)[156]
G+P+MLiver disease pathogenesis[157]
T+P+MEffect of heavy-metal exposure in neurodevelopment[158]
sT+sMSpatial resolution approaches to study brain injuries[159]
sT+sMMultimodal spatial approach used in Parkinson’s disease[160]
scT+EIdentification of genomic variants associated with eye diseases[161]
BiomarkerT+M+LProgression of lung fibrosis[162]
G+T+PMetabolic mapping of Alzheimer’s disease[163]
G+T+M+MeIdentification of osteoporosis biomarkers[164]
M+LPredictive biomarkers for COVID-19 severity[165]
G+T+P+pPExploration for Hepatocellular carcinoma biomarkers[150]
P+GlyIdentification of mild cognitive impairment biomarkers[53]
T+PTemporal lobe epilepsy biomarkers[166]
T+PEmphysema biomarkers[167]
T+PIdiopathic pulmonary fibrosis[168]
Drug targets and
therapeutics
G+T+PAlzheimer’s drug discovery[169]
G+T+PProstate cancer diagnosis and therapies[170]
P+GlyGlycosylation profile of the S1 protein of eleven SARS-CoV-2 variants[92]
G+TDrug response profiling of childhood acute lymphoblastic leukemia cell lines[171]
G+TPrioritization of therapeutic targets for dyslipidemia[172]
Abbreviations: G, genomics; T, transcriptomics; P, proteomics; M, metabolomics; L, lipidomics; Gly, glycomics; sc, single cell; p, phospho; Me, methylomics; E, epigenomics; and s, spatial.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gutierrez Reyes, C.D.; Alejo-Jacuinde, G.; Perez Sanchez, B.; Chavez Reyes, J.; Onigbinde, S.; Mogut, D.; Hernández-Jasso, I.; Calderón-Vallejo, D.; Quintanar, J.L.; Mechref, Y. Multi Omics Applications in Biological Systems. Curr. Issues Mol. Biol. 2024, 46, 5777-5793. https://doi.org/10.3390/cimb46060345

AMA Style

Gutierrez Reyes CD, Alejo-Jacuinde G, Perez Sanchez B, Chavez Reyes J, Onigbinde S, Mogut D, Hernández-Jasso I, Calderón-Vallejo D, Quintanar JL, Mechref Y. Multi Omics Applications in Biological Systems. Current Issues in Molecular Biology. 2024; 46(6):5777-5793. https://doi.org/10.3390/cimb46060345

Chicago/Turabian Style

Gutierrez Reyes, Cristian D., Gerardo Alejo-Jacuinde, Benjamin Perez Sanchez, Jesus Chavez Reyes, Sherifdeen Onigbinde, Damir Mogut, Irma Hernández-Jasso, Denisse Calderón-Vallejo, J. Luis Quintanar, and Yehia Mechref. 2024. "Multi Omics Applications in Biological Systems" Current Issues in Molecular Biology 46, no. 6: 5777-5793. https://doi.org/10.3390/cimb46060345

Article Metrics

Back to TopTop