Next Article in Journal
Enhanced Apoptosis and Loss of Cell Viability in Melanoma Cells by Combined Inhibition of ERK and Mcl-1 Is Related to Loss of Mitochondrial Membrane Potential, Caspase Activation and Upregulation of Proapoptotic Bcl-2 Proteins
Next Article in Special Issue
Dopamine Transmission Imbalance in Neuroinflammation: Perspectives on Long-Term COVID-19
Previous Article in Journal
Towards Controlling the Local Bone Tissue Remodeling—Multifunctional Injectable Composites for Osteoporosis Treatment
Previous Article in Special Issue
Recent Insights into the Functional Role of AMPA Receptors in the Oligodendrocyte Lineage Cells In Vivo
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Status of Metabolomic Measurement for Insights in Alzheimer’s Disease Progression—What Is Missing?

1
Metabolomics and Analytics Centre, Leiden Academic Centre for Drug Research, Leiden University, 2333 CC Leiden, The Netherlands
2
Division of Systems Pharmacology and Pharmacy, Leiden Academic Centre for Drug Research, Leiden University, 2333 CC Leiden, The Netherlands
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(5), 4960; https://doi.org/10.3390/ijms24054960
Submission received: 31 January 2023 / Revised: 24 February 2023 / Accepted: 2 March 2023 / Published: 4 March 2023
(This article belongs to the Special Issue Latest Review Papers in Neurobiology 2023)

Abstract

:
Alzheimer’s disease (AD) is an aging-related neurodegenerative disease, leading to the progressive loss of memory and other cognitive functions. As there is still no cure for AD, the growth in the number of susceptible individuals represents a major emerging threat to public health. Currently, the pathogenesis and etiology of AD remain poorly understood, while no efficient treatments are available to slow down the degenerative effects of AD. Metabolomics allows the study of biochemical alterations in pathological processes which may be involved in AD progression and to discover new therapeutic targets. In this review, we summarized and analyzed the results from studies on metabolomics analysis performed in biological samples of AD subjects and AD animal models. Then this information was analyzed by using MetaboAnalyst to find the disturbed pathways among different sample types in human and animal models at different disease stages. We discuss the underlying biochemical mechanisms involved, and the extent to which they could impact the specific hallmarks of AD. Then we identify gaps and challenges and provide recommendations for future metabolomics approaches to better understand AD pathogenesis.

1. Introduction

Alzheimer’s disease (AD), the most common form of dementia in aging population, leads to the progressive loss of memory and other cognitive functions [1]. In 1907, Dr. Alois Alzheimer discovered the first patient with senile plaques and NFTs, which represent the major hallmarks of AD which has become a major public health problem due to the increase in the elder population worldwide [2]. AD can be divided into two types: familial AD (5%) and sporadic AD (95%). Familial AD (FAD) has an early onset (<65 years of age) and it is caused by mutations in the genes encoding amyloid precursor protein (APP) and presenilin 1 and 2 (PS1 and PS2) [3]. Age is a major risk factor for AD, but inactivity (lack of exercise), obesity, diabetes, high blood pressure, high cholesterol, and too high alcohol consumption also increase the incidence of AD. Furthermore, it seems that low educational level, social isolation, and cognitive inactivity also contributes to AD [4]. Today, diagnosis of AD is based on several neuropsychological tests, imaging, and biological analyses, which all indicate AD in a later stage. Currently, there is no efficient treatment available, although treatment by the recently approved drug lecanemab seems to delay AD progression [5], and provides some hope.
Currently, we know that brain extracellular amyloid deposits, called neuritic senile or amyloid-β plaques, and fibrillary protein deposits inside neurons, known as neurofibrillary bundles or tau tangles, appear mainly in the frontal and temporal lobes and contribute to AD progression [6]. However, there are still many questions about how AD initiates and how it progresses.
Information on factors in and mechanism of initiation and progression, therefore, is crucial for earlier diagnosis, as well as to find targets to treat AD in a stage-dependent manner. AD is very complex, and a systems biological approach is warranted. Metabolomics allows such an approach as it can be used to measure biochemical alterations underlying pathological processes thus offering great potential for the diagnosis and prognosis of neurodegenerative diseases. This is because a subject’s metabolome reflects alterations in genetic, transcript, and protein profiles as well as influences from the environment [7]. Moreover, since metabolic pathways are largely conserved between species, metabolomics could improve the translation of preclinical research conducted in animal models of AD into humans. Furthermore, the brain has a high lipid content, which indicates that lipidomics may be a highly valuable omics technique as well, to provide novel insights into AD pathogenesis.
In this review, we provide an overview of the current state of application of metabolomics (including lipidomics) research on AD in human and animal models, together with the metabolite information that has been obtained from plasma, brain, and cerebrospinal fluid (CSF) samples in human studies and animal studies. We then analyze this information using MetaboAnalyst to find the disturbed pathways among different sample types in human and animal models of different disease stages. Then we identify gaps and challenges and provide recommendations for future metabolomics approaches to better understand AD pathogenesis.

2. Metabolomics

Metabolites are substrates, intermediates, and products of metabolic body processes, which typically are small molecules with a molecular weight of less than ~1.5 kDa [8]. Since low molecular weight metabolites are intermediates or end products of cellular metabolism, metabolomics, or the study of metabolism can be considered one of the core disciplines of systems biology. It can help in improving our understanding of changes in biochemical pathways, revealing crucial information that is closely related to human disease or therapeutic status [9,10,11,12,13].
Metabolomics allows the systematic study of unique metabolomic fingerprints that result from the body functioning in different conditions, such as healthy and diseased. These fingerprints can be viewed as biomarkers of normal biological processes, pathological processes, or pharmacological responses to a therapeutic intervention [14,15,16].
Metabolism refers to the biochemical reactions that occur throughout the body within each cell and that provide the body with energy. This energy is needed for vital body processes and the synthesis of new body components [17]. Some of these are mediated by enzymes, with specialized functions in anabolism and catabolism. To that end, the body needs nutrients and energy that come from the diet. Metabolism is affected by many factors such as sex, race, exercise, diet, age, and diseases such as Parkinson’s or Alzheimer’s. The biggest impacts on metabolites are genetics and environment.
Lipidomics, a subfield of metabolomics, is the study of the lipidome, i.e., all the lipids within cells, organs, or biological systems. Lipids are vital in the biological processes of living organisms. They not only serve as structural components of cell membranes, but also play an important role in the source of chemical energy and cell signaling molecules [18]. Apart from adipose tissue, which is the most lipid-rich organ, the brain is the body’s second lipid-rich organ; 10% to 12% of the fresh weight and more than 50% of the dry weight is composed of lipids [19]. With significant structural diversity, major lipid species in the brain can be categorized as sphingolipids, glycerolipids, glycerophospholipids, fatty acids, cholesterol, and cholesterol ester. Among them, phospholipids account for 50% of total lipid content [20]. Plasma abnormal lipid profiles have been known to be associated with AD for several decades [21,22,23]. Lipidomics profiling of plasma and tissues has the potential to discover biomarkers of aging or AD, which can contribute to understanding the pathological mechanism of AD.

2.1. Metabolomics Platforms for Identification of AD Biomarkers

There are a number of metabolomics platforms available, each with its specific advantages and disadvantages, most notably differences in sensitivity, reproducibility, and equipment costs [24]. Due to the diversity of the metabolome and the complexity of biological systems, it is impossible to give a fully comprehensive metabolite profile of a biological sample by using a single analytical platform [25]. Therefore, the choice of analytical platforms will be determined by the nature of the biological specimen to be analyzed, the goal of the analysis, the nature of the compounds under investigation (i.e., polar or apolar, volatile or non-volatile), and the resources of the laboratory [24,26,27]. Numerous metabolomics platforms are commonly used in both targeted and untargeted studies, and include gas chromatography-mass spectrometry (GC-MS), liquid chromatography-mass spectrometry (LC-MS), capillary electrophoresis-mass spectrometry (CE-MS), direct-infusion mass spectrometry (DI-MS), and nuclear magnetic resonance (NMR).
GC-MS is generally considered a versatile platform given its excellent separation power, sensitivity, and reproducibility [28]. This platform often requires a chemical derivatization procedure to create volatile compounds, which means the compounds profiled are limited to those that are volatile or can be made volatile using this complex and time-consuming procedure [29]. In addition, derivatization can improve volatility, thermal stability, sensitivity, chromatographic selectivity, and peak shapes [30]. GC-MS uses electron ionization (EI) or chemical ionization (CI) to analyze volatile metabolites. EI-MS is a hard ionization technique that does not suffer from ion suppression, which means it can generate quantitative data and extensive and predictable fragmentation for the structural characterization of metabolites [31]. As EI mass spectra are consistent across instruments and laboratories, sample identification in GC-MS is based on the use of EI mass spectral libraries by matching mass spectral fragment ion patterns, which can be seen as a compound-specific “fingerprint” [32,33].
LC-MS, which is also known as high-performance LC-MS (HPLC-MS) or ultra-HPLC (UHPLC or UPLC), is predominantly used in metabolomics and can provide analysis of thermally non-volatile, unstable, high- or low-molecular weight compounds with wide polarity range. LC-MS does not need a derivatization step, which makes sample preparation simpler and more amenable to high throughput analysis [25]. The columns used in liquid chromatography separate metabolites based on the physical properties of the molecules. Two classes of stationary phases commonly used in metabolomics analysis are hydrophilic interaction liquid chromatography (HILIC) and reversed-phase (RP). HILIC is good at analyzing highly hydrophilic and ionic compounds and therefore suitable for profiling polar metabolites, whereas RP with C8 or C18 columns is widely used in providing good separation of non-polar or weakly polar compounds [34].
CE-MS has been recognized recently as an attractive complementary technique for metabolomic studies and is particularly suitable for the separation of polar and ionic compounds based on a charge-to-mass ratio. The separation of CE is fast and highly efficient and does not need extensive sample pretreatment [35]. In addition, CE-MS only needs very low or even no organic solvents. A drawback of CE is the poor concentration sensitivity due to the limited sample volume. The currently available CE-MS techniques only allow sample loading of up to 1µL, and usually only utilize 10–100 nL [36]. In addition, migration times of metabolites can fluctuate with changing environmental temperatures, which can lead to reduced reproducibility [37].
DI-MS is a high throughput method with a short analysis time where the sample is directly introduced into the ESI source without chromatographic separation by using a syringe pump or nanospray chip [38]. However, its quantitative performance is inferior to LC-MS because of the strong matrix effect. A stable isotope labeling strategy has been applied to overcome the matrix effect [39].
NMR spectroscopy is an analytical technique based on the exploitation of the magnetic properties of atomic nuclei such as 1H, 13C, and 31P, allowing the identification of different atomic nuclei based on their resonant frequencies, which are dependent on their location in the molecule [14,24,40]. The NMR technique can uniquely and simultaneously quantify a wide range of organic compounds as well as provide unbiased information about metabolic profiles [29]. The applications of NMR spectroscopy are not only limited to liquid samples [41,42,43,44] but can also be used on solid [45,46] and tissue samples [47,48,49,50]. This platform is straightforward, largely automated, and non-destructive, so samples can be reused for further studies [15,29]. The major limitation of NMR for comprehensive metabolite profiling is its relatively low sensitivity, which makes it inappropriate for analyzing low-abundance metabolites [29].
Metabolomics techniques can be divided into untargeted and targeted. Untargeted metabolomics is a global, unbiased analysis of all small-molecule metabolites within a biological system, under a given set of conditions [51]. It measures hundreds of metabolites to identify metabolic changes, in a relative or non-quantitative way, and may serve to identify changed pathways for hypothesis building and further targeted studies [52]. Compared to targeted metabolomics, it is impossible to quantify all metabolites in untargeted metabolomics due to the large number of variables as well as the identity of metabolites is often unknown [53,54,55]. An important advantage of the untargeted approach is that it may also identify new metabolism areas [56,57]. The principal challenges of untargeted metabolomics lie in several aspects: (i) the protocols and time required to process the generated a large amount of raw data, (ii) the bias towards detection of molecules in high-abundance, (iii) the reliance on the intrinsic analytical coverage of the platform used, (iv) identifying and characterizing unknown small molecules [58]. In contrast, targeted metabolomics is the (semi-)quantitative measurement of a predefined set of metabolites [7]. It is commonly driven by a hypothesis or a specific biochemical question [59]. Targeted metabolomics can be effectively used for a pharmacokinetic study of drug metabolism as well as for measuring the influence of therapeutics or genetic modifications on specific enzymes [60].
There are many public databases available for metabolomics studies, such as the Human Metabolome Database (HMDB), METLIN, PubChem, and the Kyoto Encyclopedia of Genes and Genomes (KEGG) [61]. MetaboAnalyst (https://www.metaboanalyst.ca/) (accessed on 1 October 2022) is a powerful tool designed for processing and analyzing LC-MS-based global metabolomics data including spectral processing, functional interpretation, statistical analysis with complex metadata, and multi-omics integration [62].

2.2. Metabolomics Studies Related to AD

To investigate the current status of knowledge on metabolomics and insights into AD, for this review, an advanced literature search was performed using the following words: “Alzheimer’s disease [Title] AND (metabolomics OR lipidomics)”, until October 2022. It gave 498 results. We included experimental articles which compared the results of metabolomics analysis performed in biological samples taken from controls and from pathological conditions, both in AD animal models and in AD subjects. Case reports, reviews, editorials, conference summaries, and communications articles were excluded.

2.2.1. Metabolomics in AD Human Studies

We found 44 articles that reported the outcomes of metabolomics analyses on CSF, plasma, saliva, and brain tissue samples from human subjects (Table 1) [63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106]. Among them, 5 articles used CSF sample alone, 25 articles used plasma samples alone, 8 articles used brain samples alone, 4 articles used plasma along with CSF samples, 1 article used postmortem CSF samples, and 1 article used saliva samples alone. These human studies have used metabolomics to establish disease-related plasma, brain, and CSF metabolite differences between cognitively normal (CN) individuals, mild cognitive impairment (MCI), and AD patients as predictors of AD progression.
In humans, CSF is the only fluid that can be sampled from the CNS, and therefore CSF is often used to obtain some information about what is happening in the brain. The composition of CSF reflects, to some extent, the composition of the brain’s extracellular space environment, which is in close connection with the metabolic processes occurring in this fluid and in the brain cells [107]. Several studies have shown that high CSF-total tau (tTau) and CSF-hyperphosphorylated tau (pTau) levels were found in early AD [108,109]. It is hypothesized that the metabolome in CSF may be altered in MCI or AD [64]. For that reason, the number of metabolomics-based studies investigating CSF composition is rapidly increasing. However, the collection of CSF through a lumbar puncture procedure is invasive and can be painful. It requires a patient’s cooperation which may be challenging especially for elderly people [15].
Blood samples are collected more easily compared to CSF, and would reduce the need for expensive, invasive, and time-consuming tests [110]. However, the difficulty in developing blood-based biomarkers for AD is underscored by the often-unknown ability of the molecule to pass through blood–brain barrier (BBB) and the difficulty in directly linking peripheral markers with brain processes [15]. Nevertheless, it is generally stated that the BBB is disrupted, which will increase permeability, with aging and in AD. Moreover, BBB disruption worsens as cognitive impairment increases, which means the relationships between metabolite concentrations in blood and the brain are strengthened [111]. However, there needs to be a somewhat better specification on what is meant by BBB permeability and conclusions on altered BBB transport [112].
Below we describe the results that have been reported in the metabolomics studies and summarize them in Table 1.
Maffioli et al. [103] explored the metabolome of healthy (n = 20) and AD-affected (n = 23) individuals by performing an untargeted metabolomics analysis on hippocampal samples. They detected 126 metabolites in total; 13 and 11 were up- and down-regulated, respectively, when comparing HC with AD samples. Enrichment analysis revealed that the most significantly upregulated pathways in AD samples were Arg/Pro metabolism and the pentose phosphate pathway. In contrast, the most significantly downregulated ones in AD samples were Ala/Asp/Glu metabolism, pyruvate metabolism, glycolysis/gluconeogenesis, pyrimidine metabolism, and aminoacyl-tRNA biosynthesis. In addition, gender-specific hallmarks of AD were explored. They found women with AD display a decrease in the D-serine/total serine ratio compared with men with AD.
To investigate the association between fatty acid metabolism and AD, Snowden et al. [82] conducted an untargeted metabolomics study. The samples were brain tissue from 43 individuals (14 AD, 14 healthy controls (HC), and 15 asymptomatic AD), ranging from 57 to 95 years old. They found that lower tissue levels of linoleic acid, linolenic acid, eicosapentaenoic acid, oleic acid, and arachidonic acid were related to worse cognitive performance, whereas higher brain DHA levels were associated with poorer cognitive performance.
Three potential biomarkers (ornithine, uracil, lysine) were identified by CE-TOF-MS on plasma samples from 40 HC, 26 MCI, and 40 AD patients [104]. For ornithine, there were significant differences between the HC and AD groups and between the MCI and AD groups. For uracil and lysine, there were significant differences between the HC and AD groups. They also measured mRNA expression levels of the metabolic enzymes in ornithine pathways, spermine synthase (SMS), nitric oxide synthase2 (NOS2), and ornithine transcarbamylase (OTC) mRNA levels were significantly different among the three groups.
Gonzalez-Dominguez et al. [68] utilized CE-TOF-MS to discover the early diagnostic biomarkers of Alzheimer’s disease. The serum samples were obtained from different stages of the disease (42 AD, 14 MCI, 37 HC). They found that with the progression of the disease, the levels of choline, creatinine, asymmetric dimethyl-arginine, homocysteine-cysteine disulfide, phenylalanyl-phenylalanine, and different medium-chain acylcarnitines were observed to increase significantly, while asparagine, methionine, histidine, carnitine, acetyl-spermidine, and C5-carnitine levels were reduced. Those metabolites are related to oxidative stress and defects in energy metabolism.
Shao et al. [96] used RP-UPLC-MS based untargeted metabolomics to measure the concentration of plasma metabolites among AD (n = 44) and cognitively normal control (n = 94) groups. Then, another cohort (43 HC, 31 neurological disease controls, 30 AD) was used to validate the result. They identified five metabolites that were able to distinguish AD patients from the HC group, which are allocholic acid, cholic acid, chenodeoxycholic acid, indolelactic acid, and tryptophan. This finding suggested that altered bile acid profiles in AD and MCI might indicate an early risk for AD development.
Wang et al. [71] applied UPLC-MS and GC-MS to analyze plasma samples from HC, MCI, and AD patients. They found a biomarker panel consisting of six metabolites (glutamine, glutamic acid, arachidonic acid, N,N-dimethylglycine, thymine, and cytidine) that can discriminate AD patients from control. Another panel of five metabolites (arachidonic acid, N,N-dimethylglycine, 2-aminoadipic acid, thymine, and 5,8-tetradecadienoic acid) was able to differentiate MCI patients from control subjects.
Peña-Bautista et al. [105] performed an untargeted lipidomics analysis on plasma samples from 20 healthy participants, 31 MCI-AD, and 11 preclinical AD. Statistically significant differences in the levels of Cer, lysophosphatidylethanolamine (LPE), lysophosphatidylcholine (LPC), and monoglyceride (MG) were observed between the preclinical AD and healthy groups. Statistically significant differences were also observed in the levels of diglycerol (DG), MG, and phosphatidylethanolamines (PE) between healthy groups and MCI-AD. In addition, LPE (18:1) showed significant differences between healthy participants and early AD (MCI and preclinical).
To determine whether the lipidome in AD has racial and ethnic disparities, Khan et al. [102] conducted a targeted lipidomics analysis of plasma samples from 54 HC and 59 AD from African American/Black (n = 56) and non-Hispanic White (n = 57) backgrounds. Five lipids (PS (18:0/18:0), PS (18:0/20:0), PC (16:0/22:6), PC (18:0/22:6), and PS (18:1/22:6)) were altered between the AD and HC sample groups. As for racial analysis, PS (20:0/20:1) was found reduced in AD in samples from non-Hispanic White but not altered in African American/Black samples.
Chouraki et al. [79] conducted a longitudinal assessment for 2067 participants with an average period of 15.6 ± 5.2 years to identify the novel biomarkers association with AD. Among 2067 participants, 93 developed dementia, including 68 with AD. Plasma samples were collected every 4 to 8 years. Four candidate plasma biomarkers were found for dementia through the metabolomics technique. Anthranilic acid, glutamic acid, taurine, and hypoxanthine levels were found to be associated with the risk of dementia.
Van der Lee [87] studied 299 metabolites in two discovery cohorts (n = 55,658) to find the associations with cognition. A total of 15 metabolites were discovered and replicated associated with cognition including subfractions of high-density lipoprotein (HDL), docosahexaenoic acid, ornithine, glutamine, and glycoprotein acetyls. Moreover, fish (oil) intake was found to be strongly associated with DHA blood concentrations (p = 9.9 × 10−53). Physical activity was found to be associated with increased (p < 0.05) levels of metabolites that were associated with higher cognitive function (medium and large HDL subfractions) and decreased levels of metabolites that were associated with lower cognitive function (glycoprotein acetyls, ornithine, and glutamine). Smokers were found to have decreased concentrations of all HDL subfractions associated with higher cognitive function and increased concentrations of metabolites associated with decreased cognitive function.
The Alzheimer’s Disease Neuroimaging Initiative (ADNI) began in 2004, and unites researchers who are investigating longitudinal data in AD. Here, clinical, neuroimaging, cognitive, biofluid biomarkers, and genetic data were collected to define the progression of AD [113]. Horgusluoglu et al. [101] systematically interrogated metabolomic, genetic, transcriptomic, proteomic, and clinical data from ADNI and found that short-chain acylcarnitines/amino acids and medium/long-chain acylcarnitines are most associated with AD clinical severity. Arnold et al. [114] used metabolomics data from 1571 participants of ADNI to investigate AD group-specific metabolic alterations. Fifteen metabolites were found to be associated with the female sex and APOE ε4 genotype: For CSF Aβ1–42, threonine showed a sex-specific effect with a greater effect size in males, while valine showed a larger effect in females. For CSF p-tau, acylcarnitines C5-DC (C6-OH), C8, C10, C2, and histidine showed stronger associations in females, whereas the related ether-containing PCs, PC ae C36:1, PC ae C36:2, asparagine, glycine, and one hydroxy-SM (SM (OH) C16:1) yielded stronger associations in males. MahmoudianDehkordi et al. [90] measured 15 primary and secondary bile acids in serum levels of 1464 subjects (37 CN older adults, 284 early mild cognitive impaired patients, 505 late mild cognitive impaired patients, and 305 AD). Primary bile acid cholic acid was found to be significantly lower serum levels in AD patients compared to CN subjects, whereas higher levels of secondary bile acids deoxycholic acid (DCA) and its conjugated forms (glycodeoxycholic acid (GDCA), glycolithocholic acid (GLCA), and taurolithocholic acid (TLCA)) were significantly associated with worse cognitive function.
Taken together, metabolomics allows the detection of metabolic alterations by monitoring multiple metabolites simultaneously. Multiple human studies have used metabolomics to distinguish age and sex-specific changes in plasma, brain, or CSF samples between CN subjects, MCI, and AD patients as predictors of AD progression which were then tested in animal models of AD to identify possible underlying causal mechanisms.

2.2.2. Metabolomics in Animal Models of AD

More insights have been obtained on changes in metabolic pathways in AD, however, the lack of substantial time course data in humans is hindering understanding the sequence of disease stage-dependent changes in metabolic pathways. In other words, if we do not follow the change of metabolites and lipids over time, we cannot understand what changes along with disease progression and what may be important for developing adequate (stage-dependent) treatment of AD. This knowledge gap in biomarkers indicates that the onset and early stage of AD cannot be bridged by human studies since extensively obtaining human samples is by far more costly and time-consuming than obtaining samples from animals, it is also impossible to obtain human brain samples in longitudinal studies from the human aging population. Therefore, there is a need for alternative approaches to obtain the relation between the changes in biomarkers and AD stages, which can be achieved through animal model studies [115].
The most commonly used experimental animal models are transgenic mice that express human genes associated with familial AD (FAD) that result in the formation of amyloid plaques (by expression of human APP alone or in combination with human PSEN1), whereas human familial AD accounts for only 5% of cases [115,116,117,118,119]. Though not ideal, animal models provide an opportunity to study the early pathological disease mechanisms that can help to unravel processes associated with the development of AD. Additionally, animal models allow the investigation of (brain) tissues and fluids, and longitudinal studies can be performed to track disease progression, which cannot be accomplished in humans. Currently, along with the popular animal models of FAD including APP (Tg2576), APP/PS1, or 3xTg AD mice, the development of humanized mouse models expressing genetic risk factors, such as APOE ε4 allele, allows researchers to study mechanisms of late-onset sporadic AD [120,121,122].
In the present review, we specifically summarized AD mouse model research as only mouse research included age information in young mice. As summarized in Table 2, we found 42 articles [123,124,125,126,127,128,129,130,131,132,133,134,135,136,137,138,139,140,141,142,143,144,145,146,147,148,149,150,151,152,153,154,155,156,157,158,159,160,161,162,163,164] that carried out metabolomics analyses on samples of the brain, plasma, feces, spleen, and pancreas in mouse models of AD. Among them, nine articles studied plasma samples. A total of 21 articles only studied brain samples, and 9 articles conducted metabolomics experiments on plasma and brain samples. One study profiled spleen and brain samples, one study used serum and pancreas samples, one study used pancreas and serum samples, and one article used serum, brain, and feces samples on AD research. The results are described below.
Speers et al. [157] conducted metabolomics analysis of cortical tissue in 8-month-old male and female 5xFAD mice and their aged-match wild-type littermates. Sex differences were observed: 12 metabolites (betaine, lysine, pyridoxamine, urate, NAD, erythritol, spermine, N-acetylmannosamine, glycerol 2-phosphate, lauroyl-L-carnitine, sodium taurocholate, nicotinamide hypoxanthine dinucleotide) were significantly altered by the transgene in the female 5xFAD mice, whereas only five (homocysteine, betaine, N-Acetyl-mannosamine, S-Adenosyl-homocysteine, Adenosine 3′,5′-cyclic monophosphate) significantly altered metabolites in male 5xFAD.
Zhao et al. [159] applied targeted metabolomics on the hippocampi of 2- and 6-month-old triple transgenic AD male mice and age-sex-matched wild-type mice (WT). A total of 70 differential metabolites were identified, among them 24 metabolites were found changed in 2-month-old AD mice compared to WT, 60 metabolites were found changed in 6-month-old AD mice compared to WT. Fourteen metabolites were found in common, which are 7-methylguanosine, adenosine, adenosine 3′,5′-cyclic monophosphate (cAMP), cis-4-Hydroxy-d-proline, deoxycytidine, cytidine, deoxyadenosine monophosphate (dAMP), ethanolamine, glycerophosphocholine (GPC), L-2-aminoadipic acid, L-methionine, N-acetyl-D-glucosamine (GlcNAc), N-acetyl-L-tyrosine, and riboflavin (VB2). These results highlight the involvement of abnormal purine, pyrimidine, arginine, and proline metabolism, along with glycerophospholipid metabolism in the early pathology of AD.
Dejakaisaya et al. [156] identified alterations in cerebral metabolites and metabolic pathways in cortex, hippocampus, and serum samples from the Tg2576 AD mice model. Eleven metabolites showed significant differences in the cortex, including hydroxyphenyllactate (linked to oxidative stress) and phosphatidylserine (linked to lipid metabolism). For the network analysis, the authors used weighted correlation network analysis (WGCNA) to investigate the metabolite-group corrections. They identified five pathways, including alanine, aspartate, and glutamate metabolism, and mitochondria electron transport chain, that were significantly correlated with AD genotype.
Kim et al. [150] used untargeted metabolomics to investigate alterations in metabolite profiles of hippocampal tissues in 6-, 8- and 12-month-old wild-type and 5xfamiliar AD (5xFAD) mice. They found nicotinamide and adenosine monophosphate levels significantly decreased while lysophosphatidylcholine (LysoPC) (16:0), LysoPC (18:0), and lysophosphatidylethanolamine (LysoPE) (16:0) levels significantly increased in the hippocampi from 5xFAD mice at 8 months or 12 months of age when compared to age-matched wild type mice. In addition, the authors assumed that the primary neurons from 5xFAD reflect the hippocampal pathophysiological characteristics of 5xFAD. They treated the primary neurons with nicotinamide and found that treatment with nicotinamide rescued synaptic deficits in hippocampal primary neurons derived from 5xFAD mice. This finding indicated that decreased hippocampal nicotinamide levels could be linked with AD pathogenesis.
In 2020, Hunsberger et al. [149] collected prefrontal cortex, hippocampus, and spleen samples in 6-, 12- and 24-month-old APP/PS1 mice and age-matched wild-type mice. They conducted untargeted metabolomics analysis to investigate metabolomic alterations in naturally aged and APP/PS1 (AD) mice. Pathway analysis of changed metabolites revealed that across age, histidine metabolism was affected in all tissue samples, whereas amino acid metabolism and energy metabolism were altered in the prefrontal cortex, and AD significantly altered protein synthesis and oxidative stress in the hippocampus. Moreover, they found age-related metabolic changes occur earlier in the spleen compared to the CNS.
Zheng et al. [141] explored metabolic changes in six different brain regions between transgenic APP/PS1 mice and wild-type mice at 1, 5, and 10 months of age by using an NMR-based metabolomics approach to explore the metabolic mechanism that underlies the progression of amyloid pathology. They found the concentrations of glycerolphosphorylcholine, phosphocholine, and myo-inositol increased significantly in the hypothalamus of APP/PS1 mice when compared to WT mice, which indicated that the hypothalamus may be the main hypermetabolic region in the brain.
In conclusion, considering that biochemical pathways are largely conserved between humans and rodents [165], animal research is considered a valuable addition to human studies as human samples are costly and—especially in the case of brain samples—not available for longitudinal study. Different animal models of AD closely mimic the changes in metabolic networks associated with disease progression in humans.

3. AD Metabolic Pathways Analysis

MetaboAnalyst 5.0 (https://www.metaboanalyst.ca/) (accessed on 1 October 2022) is a website that provides software to analyze metabolomics data. It processes raw MS spectra, normalizes comprehensive data, and provides statistical analysis, functional analysis, meta-analysis, and metabolic pathway analysis. Metabolic pathway analysis identifies which metabolic pathways have compounds (from the user’s input list) that are over-represented and returns pathway impact. Pathway impact is calculated as the sum of the importance measures of matched metabolites normalized by the sum of the importance measures of all metabolites in each pathway [166]. Using this software, we analyzed the metabolic pathways to identify altered metabolites in the brain, plasma, and CSF in the literature studies described above.
The metabolic pathways were represented as circles according to their scores from enrichment (y-axis) and topology analyses (pathway impact, x-axis). Darker circle colors indicated more significant changes in metabolites in the corresponding pathway. The size of the circle corresponds to the pathway impact score and was correlated with the centrality of the involved metabolites.

3.1. Metabolic Pathway Analysis among Human Studies

After inputting all the altered metabolites from 43 human studies among different AD stages (MCI and AD) in brain, plasma, and CSF samples, 13 main metabolic pathways had a p-value less than 0.05 and impact value greater than 0.5 (Figure 1), including phenylalanine, tyrosine and tryptophan biosynthesis, taurine and hypotaurine metabolism, alanine, aspartate and glutamate metabolism, cysteine and methionine metabolism, arginine and proline metabolism, phenylalanine metabolism, tryptophan metabolism, arginine biosynthesis, beta-alanine metabolism, histidine metabolism, tyrosine metabolism, glycine, serine and threonine metabolism, and D-glutamine and D-glutamate metabolism. Details of pathway information are presented in Supplementary Table S1.

3.2. Common Regulated Metabolic Pathways among Human Plasma and CSF Sample

Furthermore, we investigated the altered metabolic pathways in common among different sample types in MCI and AD patients. After inputting all the altered metabolites from human MCI plasma samples, 12 metabolic pathways had a p-value less than 0.05 and impact values greater than 0. Using all the altered metabolites from the human AD plasma sample, 16 metabolic pathways had a p-value less than 0.05 and an impact value greater than 0. As shown in Figure 2A, a total of eight pathways were shared in the comparison between MCI VS. CN and AD VS. CN in plasma samples, which indicated that the metabolic mechanisms of AD and MCI share similar pathological alterations.
Similarly, all the altered metabolites from CSF samples were analyzed among MCI and AD groups. As shown in Figure 2B, a total of five pathways were shared in the comparison between MCI VS. CN and AD VS. CN. It is noted that all the MCI pathways overlapped with AD pathways in CSF samples.
We next used the combined (MCI + AD) altered metabolites found in plasma and CSF samples to understand important pathways in common between these two sample matrices. As shown in Figure 2C, a total of 13 pathways were shared between the plasma and CSF samples. Details of pathway information are presented in Supplementary Table S2.

3.3. Significantly Altered Metabolic Pathways among AD Mouse Models

For all the altered metabolites from the 35 mouse studies among different ages (2 months–24 months) in brain and plasma samples, when comparing to the control group, 13 main metabolic pathways were found with a p-value less than 0.05 and an impact value greater than 0.5 (Figure 3), including phenylalanine, tyrosine and tryptophan biosynthesis, linoleic acid metabolism, synthesis and degradation of ketone bodies, alanine, aspartate and glutamate metabolism, glycine, serine and threonine metabolism, arachidonic acid metabolism, phenylalanine metabolism, beta-alanine metabolism, arginine biosynthesis, glycerophospholipid metabolism, histidine metabolism, arginine and proline metabolism, and glyoxylate and dicarboxylate metabolism. Details of pathway information are presented in Supplementary Table S3.

3.4. Common Regulated Metabolic Pathways among Mouse Plasma and Brain Sample

Literature studies were combined to find disturbed pathways at different ages of AD mouse models, using all the altered metabolites at different ages in plasma and brain samples to perform pathway analysis. We included the pathways that meet a p-value less than 0.05 and an impact value greater than 0. The results showed that 31 pathways were significantly altered in mouse brain and plasma samples, the pathway impact values are shown as a heatmap (Figure 4A).
Furthermore, we investigated the altered metabolic pathways in common including all ages in plasma and brain samples. After including all the altered metabolites from mouse plasma samples, 18 metabolic pathways were found with a p-value less than 0.05 and an impact value greater than 0. Using all the altered metabolites from mouse brain samples, 18 metabolic pathways had a p-value less than 0.05 and an impact value greater than 0. As shown in Figure 4B, a total of 10 pathways were overlapping between the brain and plasma, indicating that the metabolic mechanisms seen in mouse plasma and brain share similar pathological alterations. Details of pathway information are presented in Supplementary Table S4.

4. Main Metabolic Pathways and Main Lipid Species with Respect to AD

The previous Section 3.1 and Section 3.3 highlighted that there was a total of 13 significant (p-value < 0.05 and impact value > 0.5) metabolic pathways altered in all matrices measured across all studies, including human and mouse research. Of these, eight altered metabolic pathways were found to be in common between AD mouse models and human AD subjects over all matrices: (i) alanine, aspartate, and glutamate metabolism, (ii) arginine and proline metabolism, (iii) arginine biosynthesis, (iv) β-alanine metabolism, (v) glycine, serine, and threonine metabolism, (vi) phenylalanine metabolism, vii) histidine metabolism, and (viii) phenylalanine, tyrosine, and tryptophan biosynthesis. Section 3.2 highlighted metabolic pathways in human plasma and CSF samples. Section 3.4 highlighted metabolic pathways in mouse brain and plasma samples. The matrix commonly investigated in both human and mouse research is plasma. There were 17 significant (p-value < 0.05 and impact value > 0.5) metabolic pathways in human plasma samples, whereas there were 18 significant (p-value < 0.05 and impact value > 0.5) metabolic pathways in mouse plasma samples. Of all of these, there were 12 metabolic pathways that were altered in both AD mouse models and human AD subjects in plasma: (i) alanine, aspartate, and glutamate metabolism, (ii) arginine and proline metabolism, (iii) arginine biosynthesis, (iv) butanoate metabolism, (v) citrate cycle (TCA cycle), (vi) glutathione metabolism, (vii) glycerophospholipid metabolism, (viii) glycine, serine, and threonine metabolism, (ix) glyoxylate and dicarboxylate metabolism, (x) linoleic acid metabolism, (xi) phenylalanine metabolism, and (xii) phenylalanine, tyrosine, and tryptophan biosynthesis. The following section is a detailed description of these important metabolic pathways that emerged as significantly altered after consolidating the analysis results.

4.1. Arginine Metabolism

L-arginine is a semi-essential amino acid that can be metabolized to form a number of bioactive molecules [167] (Figure 5). It is synthesized from proline or glutamate, with the ultimate synthetic step catalyzed by argininosuccinate lyase [168]. L-arginine can be metabolized by arginases, nitric oxide synthases (NOS), and possibly also by arginine decarboxylase (ADC), resulting ultimately in the production of agmatine, ornithine, nitric oxide (NO), or urea [168]. The expression of several of these enzymes can be regulated at transcriptional and translational levels by changes in the concentration of L-arginine itself [169].
L-ornithine is the arginase-mediated metabolite of L-arginine, with urea as the by-product. L-ornithine can be further metabolized to form putrescine, spermidine, and spermine polyamine, which are essential for normal cell growth and functioning, or via a separate pathway to form glutamine and cell-signaling molecule, GABA [167]. Previous research has reported decreased glutamate and GABA levels in AD brains and increased glutamine synthase (GS) levels in the lumbar cerebrospinal fluid of AD patients [170,171]. In peripheral organs and also CNS, arginine can also be metabolized by ADC to produce agmatine, a neurotransmitter that plays an important role in the learning and memory process [172].
NO is a gaseous signaling molecule produced by NOS. NO, derived from neuronal NOS (nNOS), plays an important role in synaptic plasticity and learning, and memory [173,174,175]. Moreover, L-arginine and NO affect the cardiovascular system as endogenous antiatherogenic molecules that protect the endothelium, modulate vasodilatation, and interact with the vascular wall and circulating blood cells [176,177,178,179,180].

4.2. Alanine, Aspartate, and Glutamate Metabolism

Glutamate is the principal excitatory neurotransmitter of the brain [181]. Most neurons and glia are likely to be influenced by glutamate since they have receptors for glutamate. Glutamate is considered the main neurotransmitter of neocortical and hippocampal pyramidal neurons and is involved in higher mental functions such as cognition and memory [182]. Disturbance of excitatory glutamatergic neurotransmission is believed to be associated with many neurological disorders, including Alzheimer’s disease (AD) [182], ischemic brain damage [183], and motor neuron disease [184].
Glutamate receptors can be divided into two classes: ionotropic (N-methyl-D-aspartate, NMDA) and α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA)/kainite subtypes and metabotropic [185]. The role of glutamate and glutamate receptors in learning and memory is widely recognized. For instance, NMDA antagonists impair learning and memory while NMDA agonists and facilitators improve memory [182]; likewise, AMPAKines (positive modulators of receptor function) facilitate learning and memory [186]. Circumstantial evidence of the involvement of glutamatergic pathways derives from the well-known role of structures such as the hippocampus in learning and memory [187]. More specifically, lesions of certain glutamatergic pathways impair learning and memory [188]. Moreover, glutamate and glutamate receptors are involved in mechanisms of synaptic plasticity, which are considered to underlie learning and memory [189,190,191].

4.3. Purine Metabolism

Purines and pyrimidines are components of many key molecules in living organisms. The primary purines adenine and guanosine and the pyrimidines cytosine, thymidine, and uracyl are the core of DNA, RNA, nucleosides, and nucleotides involved in energy transfer (ATP, GTP) [192,193]. Several studies indirectly suggested that purine metabolism has altered in AD. Energy metabolism, which depends on mitochondrial function and ATP production, is markedly altered in AD [194,195]. In addition, oxidative damage to DNA and RNA, as revealed by the increase in 8-hydroxyguanosine, is found in the brain samples of AD [196,197,198,199]. Direct alterations of purine metabolism in AD have been detected by metabolomics in postmortem ventricular CSF [200] and in the spinal cord CSF of living individuals [201,202,203]. Only a limited number of metabolomics studies have been carried out in AD brains [203].

4.4. Taurine and Hypotaurine Metabolism

Taurine is the second most abundant endogenous amino acid in the central nervous system (CNS) and has multiple roles in our body: thermoregulation [204], stabilization in regulating protein folding [205], anti-inflammatory effects [206], antioxidation [207], osmoregulation [208], and calcium homeostasis [209]. Recently, taurine has shown therapeutic effects as a cognitive enhancer in animal models of non-AD neurological disorders [210,211,212,213]. Taurine protected mice from the memory disruption induced by alcohol, pentobarbital, sodium nitrite, and cycloheximide but had no obvious effect on other behaviors including motor coordination, exploratory activity, and locomotor activity [210]. Intravenously injected taurine significantly improves post-injury functional impairments of traumatic brain injury in rats [211]. The intracerebroventricular (ICV) administration of taurine protects mice from learning impairment induced by hypoxia. Neither beta-alanine nor saccharose was able to mimic the effects of taurine [212]. In streptozotocin-induced sporadic dementia rat models, cognitive impairment and deterioration of neurobehavioral activities are ameliorated by taurine [213].
Taurine also has multiple disease-modifying roles to cease or prevent AD neuropathology. During the development of AD, amyloid-β (Aβ) progressively misfolded into toxic aggregates, which are strongly associated with neuronal loss, synaptic damage, and brain atrophy. An electron microscopy study indicates that taurine slightly decreases β-amyloid peptide aggregation in the brain at a millimolar concentration [214]. Taurine also has anti-inflammatory and antioxidant properties; it can provide protection for neuronal cells and mitochondria from the neurotoxicity of Aβ. By activating GABA and glycine receptors, taurine inhibits excitotoxicity caused by Aβ-induced glutamatergic transmission activation [215].

4.5. Cholinergic System

As acetylcholine (ACh) plays a vital role in cognitive processes, the cholinergic system is considered an important factor in AD [216]. The brain regions most affected by a loss of elements of the acetylcholine system include the hippocampus, cortex, and entorhinal [217]. Cholinesterase inhibitors are one of the few drug therapies available in the clinic for the treatment of AD, and it was inspired by the fact that cholinesterase inhibitors increase the availability of acetylcholine at brain synapses [218]. The validation of the cholinergic system was seen as an important therapeutic target in the disease.

4.6. Fatty Acids

Fatty acids are the basic building blocks of more complex lipids and can be classified by the number of double bonds as saturated fatty acids (SFAs) and unsaturated fatty acids. SFAs do not include any double bonds, whereas unsaturated fatty acids contain at least one (monounsaturated fatty acids, MUFAs) or two or more (polyunsaturated fatty acids, PUFAs) double bonds [219,220]. Altered unsaturated fatty acids have been associated with AD in multiple studies. The brain is especially enriched with two PUFAs: docosahexaenoic acid (DHA) and arachidonic acid (AA). DHA, as one of omega-3 PUFAs, is the predominant structural fatty acid in the mammalian brain and plays an essential role in brain functioning, especially in cognitive function; DHA levels were lower in AD brains [221,222] or plasma [69], and increased intake of DHA from fish or marine oils may lower AD risk [223,224,225]. AA of the ω-6 fatty acid family appears to play critical mediator roles in amyloid (Aβ)-induced pathogenesis, leading to learning, memory, and behavioral impairments in AD [226]. The levels of free AA have been found to increase in AD patient brain samples [82], whereas the levels of AA in phospholipids are reduced in the hippocampus of AD subjects [227].

4.7. Glycerolipids

Glycerolipids can be categorized into triacylglycerols (TAG, also known as triglycerides, TG), monoacylglycerol (MAG), and diacylglycerol (DAG) based on the number of acyl groups in the structure. TAG, the most predominant glycerolipids, are esters composed of a glycerol backbone and three fatty acids. TAG levels are found not to be changed in the serum of AD patients when compared to control subjects. [228]. However, MAG and DAG are elevated in both the prefrontal cortex and plasma of AD and MCI subjects in comparison to controls [229,230]. Moreover, MAG and DAG are elevated in the grey matter of MCI and AD patients, suggesting that these biochemical changes may play a role in the development of MCI and in the transition from MCI to AD [231].

4.8. Glycerophospholipids

Glycerophospholipids (GPs), also referred to as phospholipids (PLs), are typically amphipathic and make up the characteristic lipid bilayer structure of biological membranes. Moreover, GPs are the major type of lipids that make up cell membranes and account for 50–60% of the total membrane mass along with cholesterol and glycolipids [232]. GPs include phosphatidylethanolamine (PE), phosphatidic acid (PA), phosphatidylserine (PS), phosphatidylglycerol (PG), phosphatiylcholine (PC), phosphatidylinositol (PI), sphingomyelin (SM), and cardiolipin (CL) [233]. Studies on GP composition indicate that levels of PC, PE, and PI are significantly decreased in neural membranes from different regions of AD patients compared to age-matched control brains [234,235,236,237,238,239].
Phosphatidylethanolamine (PE) is converted to lysophosphatidylethanolamine (lyso-PE) by phospholipase A2 (PLA2), an important inflammatory mediator that is dysregulated in AD. PLA2 level has been found to be elevated in the human cerebral cortex [240] or decreased in the human parietal and frontal cortex [241]. Moreover, PLA2 influences the processing and secretion of amyloid precursor protein, which gives rise to the β-amyloid peptide, the major component of the amyloid plaque in AD [241]. Moreover, PLA2 has been found to play an important role in memory retrieval [242].
Phosphatidylserine (PS) is the major acidic phospholipid class that accounts for 13–15% of the phospholipids in the human cerebral cortex [243]. PS is known as a “brain nutrient”, as it can not only nourish the brain, but also enhance brain functions such as improving cognition, memory, and reaction force [244]. In six double-blind trials, PS has been found effective for AD. At daily doses of 200–300 mg for up to six months, PS consistently improved clinical global impression and activities of daily living [245]. In milder cases, PS improved orientation, concentration, learning, and memory for names, locations, and recent events. In the largest trial, involving 425 elderly patients (aged between 65 and 93 years) with moderate to severe cognitive decline, PS significantly improved memory, learning motivation, and socialization, suggesting that it has a vital impact on the quality of life of such elderly patients.
Phosphatidylcholine (PC) is an essential component of cell membranes and makes up approximately 95% of the total choline compound pool in most tissues [246,247]. Its function is defined primarily by chain length since chain length differences can affect cell membrane fluidity [248]. Three PCs (PC 16:0/20:5, PC 16:0/22:6, and PC 18:0/22:6) have been found significantly diminished in AD patients [249].
Lysophosphatidic acids (LPAs) are phospholipids derivatives that can act as signaling molecules [250]. Ahmad et al. [95] investigated the association between LPAs and CSF biomarkers of AD, Aβ-42, p-tau, and total tau levels overall and with MCI to AD progression. Five LPAs (LPA C16:0, LPA C16:1, LPA C22:4, LPA C22:6, and isomer-LPA C 22:5) correlated significantly and positively with CSF biomarkers of AD, Aβ-42, p-tau, and total tau. Additionally, LPA C16:0 and LPA C16:1 showed associations with MCI to AD dementia progression.

4.9. Sphingolipids

Sphingolipids, a class of membrane biomolecules, include sphingosine 1-phosphates (S1P), Cers, SMs, and glycosphingolipids, which are vital for maintaining cell integrity and signal transduction processes [251]. Cers, the basic structural units of the sphingolipid class, have been seen as key contributors to the pathology of AD as they are able to affect both Aβ generation and tau phosphorylation [252]. Filippov et al. found elevated levels of ceramides Cer16, Cer18, Cer20, and Cer24 in the brains of AD patients. Two saturated ceramides, Cer (d18:1/18:0) and Cer (d18:1/20:0) were significantly increased in the senile plaques [253]. High ceramide levels were also found in AD serum [254] and CSF samples [255]. The greatest genetic risk factor for late-onset AD is the ε4 allele of apolipoprotein E (ApoE). ApoE regulates the secretion of the potent neuroprotective signaling lipid S1P [256]. S1P is derived by phosphorylation of sphingosine, catalyzed by sphingosine kinases 1 and 2 (SphK1 and 2). SphK1 positively regulates glutamate secretion and synaptic strength in hippocampal neurons. Reduced levels of S1P have been found in AD brains compared to controls [256,257]. All these studies mentioned above suggested that sphingolipid metabolism plays a critical role in AD pathology.

4.10. Cholesterol and Cholesteryl Esters

Despite the brain occupying only 2% of total body weight, it contains 25% of the body’s cholesterol. Due to the BBB, cholesterol metabolism in the CNS is largely separated from that in the periphery and cholesterol is de novo synthesized in the CNS [258]. Studies have found that brain cholesterol was significantly increased in AD patients than in controls [259,260]. Moreover, cholesterol showed abnormal accumulation in the senile plaques of the human brain, a hallmark neuropathological feature of AD [261].

5. Conclusions and Future Directions

Considering the dramatic aging of populations worldwide, it is of great importance to explore AD pathogenesis. Metabolic changes associated with AD progression occur prior to the development of clinical symptoms; metabolomics by itself or in conjunction with the additional currently available biomarkers for AD diagnosis could serve as an additional tool to increase the accuracy of diagnosis, to predict the disease progression, and to monitor the efficacy of therapeutic intervention.
Metabolomic studies have demonstrated the dramatic impact of AD pathogenesis and progression on metabolites and related metabolic pathways, including energy-related metabolism, fatty acid metabolism, abnormal lipid metabolism, altered amino acids metabolism (e.g., arginine, glutamate), and some others. In the present review, we summarized the metabolomics studies that were performed in biological samples of AD subjects and AD mouse models. The results of rats were too sparse and not suitable for further analysis. The mouse research shows that 12- and 24-months, middle and old age in AD mouse models, can be equivalated to the MCI and AD late stage in humans, respectively. As obtaining human body samples is costly, limited in possible samples sites by ethics (i.e., brain), and time-consuming for the long life span of humans, the above indicates that animal research may be considered a valuable addition, as it can be designed in a longitudinal fashion and with samples from multiple sites of the body to obtain time-course information and interrelationships, to gain insights that support research on AD in humans.
The disturbed pathways by AD were analyzed based on metabolite data collected from the literature. Eight disturbed metabolic pathways were found in common between AD mouse research and AD human research. These pathways are alanine, aspartate, and glutamate metabolism, arginine and proline metabolism, arginine biosynthesis, β-Alanine metabolism, glycine, serine and threonine metabolism, phenylalanine metabolism, histidine metabolism, phenylalanine, tyrosine, and tryptophan biosynthesis.
Our analysis of the literature studies has several limitations. The coverage of metabolites varied among different studies due to the detection sensitivity differences, as analytical platforms (e.g., NMR, GC-MS, LC-MS) and analytical methods are diverse in different laboratories. In addition, it is hard to reproduce various metabolomics results because of different sample sources (brain tissue or plasma) from either deceased or living patients and diverse distribution about sex, age, and suffering from other diseases. Moreover, the methods used for obtaining samples, such as CSF and brain, varied among different studies. For example, delays between removing and freezing animal or human brain tissue can affect metabolomics analysis.
Altogether, in this review, we summarized and analyzed existing metabolomics data and the relation between plasma, CSF, and brain for animals and plasma and CSF for human data. We identified missing longitudinal information, which would be difficult to be obtained from humans (high costs, long direction) while also in the human brain cannot be sampled. Longitudinal and multi-body site information, however, is important to understand the processes in AD. This is where animal research may support AD research in humans to provide new insights on disease biomarkers patterns and biological pathways, that will support AD stage diagnosis in humans but also the discovery of AD future therapeutic targets.

Supplementary Materials

The supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms24054960/s1.

Author Contributions

Conceptualization, E.C.M.d.L.; manuscript structure, E.C.M.d.L. and C.Y.; literature investigation and data extraction, C.Y.; writing—original draft preparation, C.Y.; writing—review and editing, E.C.M.d.L., A.C.H., A.K. and T.H.; supervision, E.C.M.d.L., A.C.H., A.K. and T.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Chunyuan Yin acknowledges the China Scholarship Council for the funding No. 202006550003.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AA: arachidonic acid; Ach: acetylcholine; Ac-PUT: N1-acetyl-putrescine; Ac-SPM: N1-acetyl-spermine; AD: Alzheimer’s disease; ADC: arginine decarboxylase; ADNI: Alzheimer’s Disease Neuroimaging Initiative; AMPA: α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid; ApoE: apolipoprotein E; APP: amyloid precursor protein; BBB: blood–brain barrier; CE-MS: capillary electrophoresis-mass spectrometry; Cers: ceramides; CI: chemical ionization; CL: cardiolipin; CN: cognitively normal; CNS: central nervous system; DAG: diacylglycerol; dAMP: deoxyadenosine monophosphate; DCA: deoxycholic acid; DHA: docosahexaenoic acid; DI-MS: direct infusion mass spectrometry; EI: electron ionization; FAD: familial AD; GABA: γ-aminobutyric acid; GC-MS: gas chromatography-mass spectrometry; GDCA: glycodeoxycholic acid; GLCA: glycolithocholic acid; GlcNAc: N-Acetyl-D-glucosamine; GPC: glycerophosphocholine; GPs: glycerophospholipids; GS: glutamine synthase; HILIC: hydrophilic interaction liquid chromatography; HMDB: Human Metabolome Database; HPLC-MS: high-performance LC-MS; ICV: intracerebroventricular; KEGG: Kyoto Encyclopedia of Genes and Genomes; LC-MS: liquid chromatography-mass spec-trometry; LPAs: Lysophosphatidic acids; LysoPC or LPC: lysophosphatidylcholine; LysoPE or LPE: lysophosphatidylethanolamine; MAG: monoacylglycerol; MCI: mild cognitive impairment; MG: monoglyceride; MUFAs: monounsaturated fatty acids; NAD: nicotinamide adenine dinucleotide; NMDA: N-methyl-D-aspartate; NMR: nuclear magnetic resonance; nNOS: neuronal NOS; NO: nitric oxide; NOS2: nitric oxide synthase2; OTC: ornithine transcarbam-ylase; PA: phosphatidic acid; PC: phosphatiylcholine; PCae: acyl-alkyl phosphatidylcholines; PE: phosphatidylethanolamines; PG: phosphatidylglycerol; PI: phosphatidylinositol; PLA2: phospholipase A2; PLs: phospholipids; PS: phosphatidylserine; PS1 and PS2: presenilin 1 and 2; pTau: hyperphosphorylated tau; RP: reversed-phase; S1P: sphingosine 1-phosphates; SAH: S-adenosyl-homocysteine; SDMA: symmetric dimethylarginine; SFAs: saturated fatty acids; SM: sphingomyelin; SMS: spermine synthase; SphK1 and 2: sphingosine kinases 1 and 2; TAG: triacylglycerols; TCA cycle: citrate cycle; TLCA: taurolithocholic acid; TQ-MS: triple quadrupole-mass spectrometry; tTau: total tau; UHPLC or UPLC: ultra-HPLC; UPLC-MS: ultraperformance liquid chromatography-mass spectrometry; VB2: riboflavin; WGCNA: weighted correlation network analysis.

References

  1. Altiné-Samey, R.; Antier, D.; Mavel, S.; Dufour-Rainfray, D.; Balageas, A.C.; Beaufils, E.; Emond, P.; Foucault-Fruchard, L.; Chalon, S. The contributions of metabolomics in the discovery of new therapeutic targets in Alzheimer’s disease. Fundam. Clin. Pharmacol. 2021, 35, 582–594. [Google Scholar] [CrossRef]
  2. Lane, C.A.; Hardy, J.; Schott, J.M. Alzheimer’s disease. Eur. J. Neurol. 2018, 25, 59–70. [Google Scholar] [CrossRef]
  3. Querfurth, H.W.; LaFerla, F.M. Alzheimer’s disease. N. Engl. J. Med. 2010, 362, 329–344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Livingston, G.; Huntley, J.; Sommerlad, A.; Ames, D.; Ballard, C.; Banerjee, S.; Brayne, C.; Burns, A.; Cohen-Mansfield, J.; Cooper, C.; et al. Dementia prevention, intervention, and care: 2020 report of the Lancet Commission. Lancet 2020, 396, 413–446. [Google Scholar] [CrossRef] [PubMed]
  5. van Dyck, C.H.; Swanson, C.J.; Aisen, P.; Bateman, R.J.; Chen, C.; Gee, M.; Kanekiyo, M.; Li, D.; Reyderman, L.; Cohen, S.; et al. Lecanemab in Early Alzheimer’s Disease. N. Engl. J. Med. 2023, 388, 9–21. [Google Scholar] [CrossRef] [PubMed]
  6. Serrano-Pozo, A.; Frosch, M.P.; Masliah, E.; Hyman, B.T. Neuropathological alterations in Alzheimer disease. Cold Spring Harb. Perspect. Med. 2011, 1, a006189. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Wilkins, J.M.; Trushina, E. Application of Metabolomics in Alzheimer’s Disease. Front. Neurol. 2017, 8, 719. [Google Scholar] [CrossRef] [Green Version]
  8. Rinschen, M.M.; Ivanisevic, J.; Giera, M.; Siuzdak, G. Identification of bioactive metabolites using activity metabolomics. Nat. Rev. Mol. Cell Biol. 2019, 20, 353–367. [Google Scholar] [CrossRef] [PubMed]
  9. Gowda, G.A.; Zhang, S.; Gu, H.; Asiago, V.; Shanaiah, N.; Raftery, D. Metabolomics-based methods for early disease diagnostics. Expert Rev. Mol. Diagn. 2008, 8, 617–633. [Google Scholar] [CrossRef] [Green Version]
  10. Dunn, W.B.; Broadhurst, D.I.; Atherton, H.J.; Goodacre, R.; Griffin, J.L. Systems level studies of mammalian metabolomes: The roles of mass spectrometry and nuclear magnetic resonance spectroscopy. Chem. Soc. Rev. 2011, 40, 387–426. [Google Scholar] [CrossRef]
  11. Goodacre, R.; Vaidyanathan, S.; Dunn, W.B.; Harrigan, G.G.; Kell, D.B. Metabolomics by numbers: Acquiring and understanding global metabolite data. Trends Biotechnol. 2004, 22, 245–252. [Google Scholar] [CrossRef]
  12. Hollywood, K.; Brison, D.R.; Goodacre, R. Metabolomics: Current technologies and future trends. Proteomics 2006, 6, 4716–4723. [Google Scholar] [CrossRef] [PubMed]
  13. Mamas, M.; Dunn, W.B.; Neyses, L.; Goodacre, R. The role of metabolites and metabolomics in clinically applicable biomarkers of disease. Arch. Toxicol. 2011, 85, 5–17. [Google Scholar] [CrossRef] [PubMed]
  14. Spratlin, J.L.; Serkova, N.J.; Eckhardt, S.G. Clinical applications of metabolomics in oncology: A review. Clin. Cancer Res. 2009, 15, 431–440. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Enche Ady, C.N.A.; Lim, S.M.; Teh, L.K.; Salleh, M.Z.; Chin, A.V.; Tan, M.P.; Poi, P.J.H.; Kamaruzzaman, S.B.; Abdul Majeed, A.B.; Ramasamy, K. Metabolomic-guided discovery of Alzheimer’s disease biomarkers from body fluid. J. Neurosci. Res. 2017, 95, 2005–2024. [Google Scholar] [CrossRef] [PubMed]
  16. Griffin, J.L.; Shockcor, J.P. Metabolic profiles of cancer cells. Nat. Rev. Cancer 2004, 4, 551–561. [Google Scholar] [CrossRef]
  17. Arrese, E.L.; Soulages, J.L. Insect fat body: Energy, metabolism, and regulation. Annu. Rev. Entomol. 2010, 55, 207–225. [Google Scholar] [CrossRef] [Green Version]
  18. Ooi, K.M.; Vacy, K.; Boon, W.C. Fatty acids and beyond: Age and Alzheimer’s disease related changes in lipids reveal the neuro-nutraceutical potential of lipids in cognition. Neurochem. Int. 2021, 149, 105143. [Google Scholar] [CrossRef] [PubMed]
  19. Yin, F. Lipid metabolism and Alzheimer’s disease: Clinical evidence, mechanistic link and therapeutic promise. FEBS J. 2022; Early View. [Google Scholar] [CrossRef]
  20. Sastry, P.S. Lipids of nervous tissue: Composition and metabolism. Prog. Lipid Res. 1985, 24, 69–176. [Google Scholar] [CrossRef]
  21. Kuo, Y.-M.; Emmerling, M.R.; Bisgaier, C.L.; Essenburg, A.D.; Lampert, H.C.; Drumm, D.; Roher, A.E. Elevated Low-Density Lipoprotein in Alzheimer’s Disease Correlates with Brain Aβ 1–42 Levels. Biochem. Biophys. Res. Commun. 1998, 252, 711–715. [Google Scholar] [CrossRef]
  22. Merched, A.; Xia, Y.; Visvikis, S.; Serot, J.M.; Siest, G. Decreased high-density lipoprotein cholesterol and serum apolipoprotein AI concentrations are highly correlated with the severity of Alzheimer’s disease☆. Neurobiol. Aging 2000, 21, 27–30. [Google Scholar] [CrossRef]
  23. Roher, A.E.; Kuo, Y.M.; Kokjohn, K.M.; Emmerling, M.R.; Gracon, S. Amyloid and lipids in the pathology of Alzheimer disease. Amyloid 1999, 6, 136–145. [Google Scholar] [CrossRef] [PubMed]
  24. Davis, V.W.; Bathe, O.F.; Schiller, D.E.; Slupsky, C.M.; Sawyer, M.B. Metabolomics and surgical oncology: Potential role for small molecule biomarkers. J. Surg. Oncol. 2011, 103, 451–459. [Google Scholar] [CrossRef] [PubMed]
  25. Bujak, R.; Struck-Lewicka, W.; Markuszewski, M.J.; Kaliszan, R. Metabolomics for laboratory diagnostics. J. Pharm. Biomed. Anal. 2015, 113, 108–120. [Google Scholar] [CrossRef]
  26. Weiss, R.H.; Kim, K. Metabolomics in the study of kidney diseases. Nat. Rev. Nephrol. 2011, 8, 22–33. [Google Scholar] [CrossRef]
  27. Manach, C.; Hubert, J.; Llorach, R.; Scalbert, A. The complex links between dietary phytochemicals and human health deciphered by metabolomics. Mol. Nutr. Food Res. 2009, 53, 1303–1315. [Google Scholar] [CrossRef]
  28. Beale, D.J.; Pinu, F.R.; Kouremenos, K.A.; Poojary, M.M.; Narayana, V.K.; Boughton, B.A.; Kanojia, K.; Dayalan, S.; Jones, O.A.H.; Dias, D.A. Review of recent developments in GC-MS approaches to metabolomics-based research. Metabolomics 2018, 14, 152. [Google Scholar] [CrossRef]
  29. Zhang, A.; Sun, H.; Wang, P.; Han, Y.; Wang, X. Modern analytical techniques in metabolomics analysis. Analyst 2012, 137, 293–300. [Google Scholar] [CrossRef]
  30. Farajzadeh, M.A.; Nouri, N.; Khorram, P. Derivatization and microextraction methods for determination of organic compounds by gas chromatography. TrAC Trends Anal. Chem. 2014, 55, 14–23. [Google Scholar] [CrossRef]
  31. Gathungu, R.M.; Bird, S.S.; Sheldon, D.P.; Kautz, R.; Vouros, P.; Matson, W.R.; Kristal, B.S. Identification of metabolites from liquid chromatography-coulometric array detection profiling: Gas chromatography-mass spectrometry and refractionation provide essential information orthogonal to LC-MS/microNMR. Anal. Biochem. 2014, 454, 23–32. [Google Scholar] [CrossRef] [Green Version]
  32. Alon, T.; Amirav, A. Isotope abundance analysis methods and software for improved sample identification with supersonic gas chromatography/mass spectrometry. Rapid Commun. Mass Spectrom. RCM 2006, 20, 2579–2588. [Google Scholar] [CrossRef]
  33. Halket, J.M.; Waterman, D.; Przyborowska, A.M.; Patel, R.K.; Fraser, P.D.; Bramley, P.M. Chemical derivatization and mass spectral libraries in metabolic profiling by GC/MS and LC/MS/MS. J. Exp. Bot. 2005, 56, 219–243. [Google Scholar] [CrossRef] [Green Version]
  34. Tang, D.Q.; Zou, L.; Yin, X.X.; Ong, C.N. HILIC-MS for metabolomics: An attractive and complementary approach to RPLC-MS. Mass Spectrom. Rev. 2016, 35, 574–600. [Google Scholar] [CrossRef]
  35. Ramautar, R.; Somsen, G.W.; de Jong, G.J. CE-MS in metabolomics. Electrophoresis 2009, 30, 276–291. [Google Scholar] [CrossRef] [PubMed]
  36. Mischak, H.; Schanstra, J.P. CE-MS in biomarker discovery, validation, and clinical application. Proteom. Clin. Appl. 2011, 5, 9–23. [Google Scholar] [CrossRef] [PubMed]
  37. Khakimov, B.; Bak, S.; Engelsen, S.B. High-throughput cereal metabolomics: Current analytical technologies, challenges and perspectives. J. Cereal Sci. 2014, 59, 393–418. [Google Scholar] [CrossRef]
  38. Putri, S.P.; Yamamoto, S.; Tsugawa, H.; Fukusaki, E. Current metabolomics: Technological advances. J. Biosci. Bioeng. 2013, 116, 9–16. [Google Scholar] [CrossRef]
  39. Giavalisco, P.; Hummel, J.; Lisec, J.; Inostroza, A.C.; Catchpole, G.; Willmitzer, L. High-Resolution Direct Infusion-Based Mass Spectrometry in Combination with Whole 13C Metabolome Isotope Labeling Allows Unambiguous Assignment of Chemical Sum Formulas. Anal. Chem. 2008, 80, 9417–9425. [Google Scholar] [CrossRef]
  40. 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]
  41. Du, Z.; Shen, A.; Huang, Y.; Su, L.; Lai, W.; Wang, P.; Xie, Z.; Xie, Z.; Zeng, Q.; Ren, H.; et al. 1H-NMR-based metabolic analysis of human serum reveals novel markers of myocardial energy expenditure in heart failure patients. PLoS ONE 2014, 9, e88102. [Google Scholar] [CrossRef] [PubMed]
  42. Yilmaz, A.; Geddes, T.; Han, B.; Bahado-Singh, R.O.; Wilson, G.D.; Imam, K.; Maddens, M.; Graham, S.F. Diagnostic Biomarkers of Alzheimer’s Disease as Identified in Saliva using 1H NMR-Based Metabolomics. J. Alzheimers Dis. 2017, 58, 355–359. [Google Scholar] [CrossRef] [PubMed]
  43. Zordoky, B.N.; Sung, M.M.; Ezekowitz, J.; Mandal, R.; Han, B.; Bjorndahl, T.C.; Bouatra, S.; Anderson, T.; Oudit, G.Y.; Wishart, D.S.; et al. Metabolomic fingerprint of heart failure with preserved ejection fraction. PLoS ONE 2015, 10, e0124844. [Google Scholar] [CrossRef]
  44. Wang, J.; Li, Z.; Chen, J.; Zhao, H.; Luo, L.; Chen, C.; Xu, X.; Zhang, W.; Gao, K.; Li, B.; et al. Metabolomic identification of diagnostic plasma biomarkers in humans with chronic heart failure. Mol. Biosyst. 2013, 9, 2618–2626. [Google Scholar] [CrossRef] [PubMed]
  45. Eddy, M.T.; Belenky, M.; Sivertsen, A.C.; Griffin, R.G.; Herzfeld, J. Selectively dispersed isotope labeling for protein structure determination by magic angle spinning NMR. J. Biomol. NMR 2013, 57, 129–139. [Google Scholar] [CrossRef] [Green Version]
  46. Khan, M.T.; Busch, M.; Molina, V.G.; Emwas, A.H.; Aubry, C.; Croue, J.P. How different is the composition of the fouling layer of wastewater reuse and seawater desalination RO membranes? Water Res. 2014, 59, 271–282. [Google Scholar] [CrossRef]
  47. Atherton, H.J.; Bailey, N.J.; Zhang, W.; Taylor, J.; Major, H.; Shockcor, J.; Clarke, K.; Griffin, J.L. A combined 1H-NMR spectroscopy- and mass spectrometry-based metabolomic study of the PPAR-alpha null mutant mouse defines profound systemic changes in metabolism linked to the metabolic syndrome. Physiol. Genom. 2006, 27, 178–186. [Google Scholar] [CrossRef]
  48. Risa, O.; Melø, T.M.; Sonnewald, U. Quantification of amounts and (13)C content of metabolites in brain tissue using high-resolution magic angle spinning (13)C NMR spectroscopy. NMR Biomed. 2009, 22, 266–271. [Google Scholar] [CrossRef] [PubMed]
  49. Komoroski, R.A.; Pearce, J.M.; Mrak, R.E. 31P NMR spectroscopy of phospholipid metabolites in postmortem schizophrenic brain. Magn. Reson Med. 2008, 59, 469–474. [Google Scholar] [CrossRef]
  50. Lutz, N.W.; Béraud, E.; Cozzone, P.J. Metabolomic analysis of rat brain by high resolution nuclear magnetic resonance spectroscopy of tissue extracts. J. Vis. Exp. 2014, 91, 51829. [Google Scholar]
  51. Naz, S.; Vallejo, M.; Garcia, A.; Barbas, C. Method validation strategies involved in non-targeted metabolomics. J. Chromatogr. A 2014, 1353, 99–105. [Google Scholar] [CrossRef]
  52. Schrimpe-Rutledge, A.C.; Codreanu, S.G.; Sherrod, S.D.; McLean, J.A. Untargeted Metabolomics Strategies-Challenges and Emerging Directions. J. Am. Soc. Mass Spectrom. 2016, 27, 1897–1905. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Sana, T.R.; Roark, J.C.; Li, X.; Waddell, K.; Fischer, S.M. Molecular formula and METLIN Personal Metabolite Database matching applied to the identification of compounds generated by LC/TOF-MS. J. Biomol. Tech. 2008, 19, 258–266. [Google Scholar]
  54. Draper, J.; Enot, D.P.; Parker, D.; Beckmann, M.; Snowdon, S.; Lin, W.; Zubair, H. Metabolite signal identification in accurate mass metabolomics data with MZedDB, an interactive m/z annotation tool utilising predicted ionisation behaviour ‘rules’. BMC Bioinform. 2009, 10, 227. [Google Scholar] [CrossRef] [Green Version]
  55. Kind, T.; Fiehn, O. Metabolomic database annotations via query of elemental compositions: Mass accuracy is insufficient even at less than 1 ppm. BMC Bioinform. 2006, 7, 234. [Google Scholar] [CrossRef] [Green Version]
  56. Chen, C.; Krausz, K.W.; Idle, J.R.; Gonzalez, F.J. Identification of novel toxicity-associated metabolites by metabolomics and mass isotopomer analysis of acetaminophen metabolism in wild-type and Cyp2e1-null mice. J. Biol. Chem. 2008, 283, 4543–4559. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Viant, M.R.; Rosenblum, E.S.; Tieerdema, R.S. NMR-based metabolomics: A powerful approach for characterizing the effects of environmental stressors on organism health. Environ. Sci. Technol. 2003, 37, 4982–4989. [Google Scholar] [CrossRef]
  58. Roberts, L.D.; Souza, A.L.; Gerszten, R.E.; Clish, C.B. Targeted metabolomics. Curr. Protoc. Mol. Biol. 2012, 98, 30.2.1–30.2.24. [Google Scholar] [CrossRef]
  59. 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] [Green Version]
  60. Nicholson, J.K.; Connelly, J.; Lindon, J.C.; Holmes, E. Metabonomics: A platform for studying drug toxicity and gene function. Nat. Rev. Drug Discov. 2002, 1, 153–161. [Google Scholar] [CrossRef] [PubMed]
  61. Go, E.P. Database Resources in Metabolomics: An Overview. J. Neuroimmune Pharmacol. 2010, 5, 18–30. [Google Scholar] [CrossRef]
  62. Pang, Z.; Zhou, G.; Ewald, J.; Chang, L.; Hacariz, O.; Basu, N.; Xia, J. Using MetaboAnalyst 5.0 for LC–HRMS spectra processing, multi-omics integration and covariate adjustment of global metabolomics data. Nat. Protoc. 2022, 17, 1735–1761. [Google Scholar] [CrossRef]
  63. Czech, C.; Berndt, P.; Busch, K.; Schmitz, O.; Wiemer, J.; Most, V.; Hampel, H.; Kastler, J.; Senn, H. Metabolite profiling of Alzheimer’s disease cerebrospinal fluid. PLoS ONE 2012, 7, e31501. [Google Scholar] [CrossRef] [PubMed]
  64. Ibanez, C.; Simo, C.; Martin-Alvarez, P.J.; Kivipelto, M.; Winblad, B.; Cedazo-Minguez, A.; Cifuentes, A. Toward a predictive model of Alzheimer’s disease progression using capillary electrophoresis-mass spectrometry metabolomics. Anal. Chem. 2012, 84, 8532–8540. [Google Scholar] [CrossRef] [PubMed]
  65. Sato, Y.; Suzuki, I.; Nakamura, T.; Bernier, F.; Aoshima, K.; Oda, Y. Identification of a new plasma biomarker of Alzheimer’s disease using metabolomics technology. J. Lipid Res. 2012, 53, 567–576. [Google Scholar] [CrossRef] [Green Version]
  66. Ibáñez, C.; Simó, C.; Barupal, D.K.; Fiehn, O.; Kivipelto, M.; Cedazo-Mínguez, A.; Cifuentes, A. A new metabolomic workflow for early detection of Alzheimer’s disease. J. Chromatogr. A 2013, 1302, 65–71. [Google Scholar] [CrossRef] [PubMed]
  67. Iuliano, L.; Pacelli, A.; Ciacciarelli, M.; Zerbinati, C.; Fagioli, S.; Piras, F.; Orfei, M.D.; Bossù, P.; Pazzelli, F.; Serviddio, G.; et al. Plasma fatty acid lipidomics in amnestic mild cognitive impairment and Alzheimer’s disease. J. Alzheimers Dis. 2013, 36, 545–553. [Google Scholar] [CrossRef]
  68. González-Domínguez, R.; García, A.; García-Barrera, T.; Barbas, C.; Gómez-Ariza, J.L. Metabolomic profiling of serum in the progression of Alzheimer’s disease by capillary electrophoresis-mass spectrometry. Electrophoresis 2014, 35, 3321–3330. [Google Scholar] [CrossRef]
  69. Gonzalez-Dominguez, R.; Garcia-Barrera, T.; Gomez-Ariza, J.L. Using direct infusion mass spectrometry for serum metabolomics in Alzheimer’s disease. Anal. Bioanal. Chem. 2014, 406, 7137–7148. [Google Scholar] [CrossRef]
  70. González-Domínguez, R.; García-Barrera, T.; Gómez-Ariza, J.L. Metabolomic study of lipids in serum for biomarker discovery in Alzheimer’s disease using direct infusion mass spectrometry. J. Pharm. Biomed. Anal. 2014, 98, 321–326. [Google Scholar] [CrossRef]
  71. Wang, G.; Zhou, Y.; Huang, F.J.; Tang, H.D.; Xu, X.H.; Liu, J.J.; Wang, Y.; Deng, Y.L.; Ren, R.J.; Xu, W.; et al. Plasma metabolite profiles of Alzheimer’s disease and mild cognitive impairment. J. Proteome Res. 2014, 13, 2649–2658. [Google Scholar] [CrossRef] [PubMed]
  72. Gonzalez-Dominguez, R.; Garcia-Barrera, T.; Gomez-Ariza, J.L. Application of a novel metabolomic approach based on atmospheric pressure photoionization mass spectrometry using flow injection analysis for the study of Alzheimer’s disease. Talanta 2015, 131, 480–489. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. González-Domínguez, R.; García-Barrera, T.; Gómez-Ariza, J.L. Metabolite profiling for the identification of altered metabolic pathways in Alzheimer’s disease. J. Pharm. Biomed. Anal. 2015, 107, 75–81. [Google Scholar] [CrossRef]
  74. Graham, S.F.; Chevallier, O.P.; Elliott, C.T.; Holscher, C.; Johnston, J.; McGuinness, B.; Kehoe, P.G.; Passmore, A.P.; Green, B.D. Untargeted metabolomic analysis of human plasma indicates differentially affected polyamine and L-arginine metabolism in mild cognitive impairment subjects converting to Alzheimer’s disease. PLoS ONE 2015, 10, e0119452. [Google Scholar] [CrossRef]
  75. González-Domínguez, R.; Rupérez, F.J.; García-Barrera, T.; Barbas, C.; Gómez-Ariza, J.L. Metabolomic-Driven Elucidation of Serum Disturbances Associated with Alzheimer’s Disease and Mild Cognitive Impairment. Curr. Alzheimer Res. 2016, 13, 641–653. [Google Scholar] [CrossRef]
  76. Paglia, G.; Stocchero, M.; Cacciatore, S.; Lai, S.; Angel, P.; Alam, M.T.; Keller, M.; Ralser, M.; Astarita, G. Unbiased Metabolomic Investigation of Alzheimer’s Disease Brain Points to Dysregulation of Mitochondrial Aspartate Metabolism. J. Proteome Res. 2016, 15, 608–618. [Google Scholar] [CrossRef] [Green Version]
  77. Vaňková, M.; Hill, M.; Velíková, M.; Včelák, J.; Vacínová, G.; Dvořáková, K.; Lukášová, P.; Vejražková, D.; Rusina, R.; Holmerová, I.; et al. Preliminary evidence of altered steroidogenesis in women with Alzheimer’s disease: Have the patients “OLDER” adrenal zona reticularis? J. Steroid. Biochem. Mol. Biol. 2016, 158, 157–177. [Google Scholar] [CrossRef]
  78. Xu, J.; Begley, P.; Church, S.J.; Patassini, S.; Hollywood, K.A.; Jüllig, M.; Curtis, M.A.; Waldvogel, H.J.; Faull, R.L.; Unwin, R.D.; et al. Graded perturbations of metabolism in multiple regions of human brain in Alzheimer’s disease: Snapshot of a pervasive metabolic disorder. Biochim. Biophys. Acta 2016, 1862, 1084–1092. [Google Scholar] [CrossRef]
  79. Chouraki, V.; Preis, S.R.; Yang, Q.; Beiser, A.; Li, S.; Larson, M.G.; Weinstein, G.; Wang, T.J.; Gerszten, R.E.; Vasan, R.S.; et al. Association of amine biomarkers with incident dementia and Alzheimer’s disease in the Framingham Study. Alzheimers Dement. 2017, 13, 1327–1336. [Google Scholar] [CrossRef] [PubMed]
  80. de Leeuw, F.A.; Peeters, C.F.W.; Kester, M.I.; Harms, A.C.; Struys, E.A.; Hankemeier, T.; van Vlijmen, H.W.T.; van der Lee, S.J.; van Duijn, C.M.; Scheltens, P.; et al. Blood-based metabolic signatures in Alzheimer’s disease. Alzheimers Dement. 2017, 8, 196–207. [Google Scholar] [CrossRef] [PubMed]
  81. Oberacher, H.; Arnhard, K.; Linhart, C.; Diwo, A.; Marksteiner, J.; Humpel, C. Targeted Metabolomic Analysis of Soluble Lysates from Platelets of Patients with Mild Cognitive Impairment and Alzheimer’s Disease Compared to Healthy Controls: Is PC aeC40:4 a Promising Diagnostic Tool? J. Alzheimers Dis. 2017, 57, 493–504. [Google Scholar] [CrossRef] [PubMed]
  82. Snowden, S.G.; Ebshiana, A.A.; Hye, A.; An, Y.; Pletnikova, O.; O’Brien, R.; Troncoso, J.; Legido-Quigley, C.; Thambisetty, M. Association between fatty acid metabolism in the brain and Alzheimer disease neuropathology and cognitive performance: A nontargeted metabolomic study. PLoS Med. 2017, 14, e1002266. [Google Scholar] [CrossRef] [Green Version]
  83. Hao, L.; Wang, J.; Page, D.; Asthana, S.; Zetterberg, H.; Carlsson, C.; Okonkwo, O.C.; Li, L. Comparative Evaluation of MS-based Metabolomics Software and Its Application to Preclinical Alzheimer’s Disease. Sci. Rep. 2018, 8, 9291. [Google Scholar] [CrossRef] [Green Version]
  84. Kim, S.H.; Yang, J.S.; Lee, J.C.; Lee, J.Y.; Lee, J.Y.; Kim, E.; Moon, M.H. Lipidomic alterations in lipoproteins of patients with mild cognitive impairment and Alzheimer’s disease by asymmetrical flow field-flow fractionation and nanoflow ultrahigh performance liquid chromatography-tandem mass spectrometry. J. Chromatogr. A 2018, 1568, 91–100. [Google Scholar] [CrossRef]
  85. Muguruma, Y.; Tsutsui, H.; Noda, T.; Akatsu, H.; Inoue, K. Widely targeted metabolomics of Alzheimer’s disease postmortem cerebrospinal fluid based on 9-fluorenylmethyl chloroformate derivatized ultra-high performance liquid chromatography tandem mass spectrometry. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2018, 1091, 53–66. [Google Scholar] [CrossRef]
  86. Nasaruddin, M.L.; Pan, X.; McGuinness, B.; Passmore, P.; Kehoe, P.G.; Hölscher, C.; Graham, S.F.; Green, B.D. Evidence That Parietal Lobe Fatty Acids May Be More Profoundly Affected in Moderate Alzheimer’s Disease (AD) Pathology Than in Severe AD Pathology. Metabolites 2018, 8, 69. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. van der Lee, S.J.; Teunissen, C.E.; Pool, R.; Shipley, M.J.; Teumer, A.; Chouraki, V.; Melo van Lent, D.; Tynkkynen, J.; Fischer, K.; Hernesniemi, J.; et al. Circulating metabolites and general cognitive ability and dementia: Evidence from 11 cohort studies. Alzheimers Dement. 2018, 14, 707–722. [Google Scholar]
  88. Kim, Y.H.; Shim, H.S.; Kim, K.H.; Lee, J.; Chung, B.C.; Kowall, N.W.; Ryu, H.; Lee, J. Metabolomic Analysis Identifies Alterations of Amino Acid Metabolome Signatures in the Postmortem Brain of Alzheimer’s Disease. Exp. Neurobiol. 2019, 28, 376–389. [Google Scholar] [CrossRef]
  89. Lin, C.N.; Huang, C.C.; Huang, K.L.; Lin, K.J.; Yen, T.C.; Kuo, H.C. A metabolomic approach to identifying biomarkers in blood of Alzheimer’s disease. Ann. Clin. Transl. Neurol. 2019, 6, 537–545. [Google Scholar] [CrossRef] [Green Version]
  90. MahmoudianDehkordi, S.; Arnold, M.; Nho, K.; Ahmad, S.; Jia, W.; Xie, G.; Louie, G.; Kueider-Paisley, A.; Moseley, M.A.; Thompson, J.W.; et al. Altered bile acid profile associates with cognitive impairment in Alzheimer’s disease-An emerging role for gut microbiome. Alzheimers Dement. 2019, 15, 76–92. [Google Scholar] [CrossRef] [PubMed]
  91. Marksteiner, J.; Oberacher, H.; Humpel, C. Acyl-Alkyl-Phosphatidlycholines are Decreased in Saliva of Patients with Alzheimer’s Disease as Identified by Targeted Metabolomics. J. Alzheimers Dis. 2019, 68, 583–589. [Google Scholar] [CrossRef] [PubMed]
  92. Peña-Bautista, C.; Roca, M.; Hervás, D.; Cuevas, A.; López-Cuevas, R.; Vento, M.; Baquero, M.; García-Blanco, A.; Cháfer-Pericás, C. Plasma metabolomics in early Alzheimer’s disease patients diagnosed with amyloid biomarker. J. Proteom. 2019, 200, 144–152. [Google Scholar] [CrossRef] [PubMed]
  93. Snowden, S.G.; Ebshiana, A.A.; Hye, A.; Pletnikova, O.; O’Brien, R.; Yang, A.; Troncoso, J.; Legido-Quigley, C.; Thambisetty, M. Neurotransmitter Imbalance in the Brain and Alzheimer’s Disease Pathology. J. Alzheimers Dis. 2019, 72, 35–43. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. van der Velpen, V.; Teav, T.; Gallart-Ayala, H.; Mehl, F.; Konz, I.; Clark, C.; Oikonomidi, A.; Peyratout, G.; Henry, H.; Delorenzi, M.; et al. Systemic and central nervous system metabolic alterations in Alzheimer’s disease. Alzheimers Res. Ther. 2019, 11, 93. [Google Scholar] [CrossRef] [Green Version]
  95. Ahmad, S.; Orellana, A.; Kohler, I.; Frölich, L.; de Rojas, I.; Gil, S.; Boada, M.; Hernández, I.; Hausner, L.; Bakker, M.H.M.; et al. Association of lysophosphatidic acids with cerebrospinal fluid biomarkers and progression to Alzheimer’s disease. Alzheimers Res. Ther. 2020, 12, 124. [Google Scholar] [CrossRef]
  96. Shao, Y.; Ouyang, Y.; Li, T.; Liu, X.; Xu, X.; Li, S.; Xu, G.; Le, W. Alteration of Metabolic Profile and Potential Biomarkers in the Plasma of Alzheimer’s Disease. Aging Dis. 2020, 11, 1459–1470. [Google Scholar] [CrossRef]
  97. Byeon, S.K.; Madugundu, A.K.; Jain, A.P.; Bhat, F.A.; Jung, J.H.; Renuse, S.; Darrow, J.; Bakker, A.; Albert, M.; Moghekar, A.; et al. Cerebrospinal fluid lipidomics for biomarkers of Alzheimer’s disease. Mol. Omics 2021, 17, 454–463. [Google Scholar] [CrossRef]
  98. Liu, P.; Yang, Q.; Yu, N.; Cao, Y.; Wang, X.; Wang, Z.; Qiu, W.Y.; Ma, C. Phenylalanine Metabolism Is Dysregulated in Human Hippocampus with Alzheimer’s Disease Related Pathological Changes. J. Alzheimers Dis. 2021, 83, 609–622. [Google Scholar] [CrossRef]
  99. Liu, Y.; Thalamuthu, A.; Mather, K.A.; Crawford, J.; Ulanova, M.; Wong, M.W.K.; Pickford, R.; Sachdev, P.S.; Braidy, N. Plasma lipidome is dysregulated in Alzheimer’s disease and is associated with disease risk genes. Transl. Psychiatry 2021, 11, 344. [Google Scholar] [CrossRef]
  100. Nielsen, J.E.; Maltesen, R.G.; Havelund, J.F.; Færgeman, N.J.; Gotfredsen, C.H.; Vestergård, K.; Kristensen, S.R.; Pedersen, S. Characterising Alzheimer’s disease through integrative NMR- and LC-MS-based metabolomics. Metab. Open 2021, 12, 100125. [Google Scholar] [CrossRef]
  101. 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. Alzheimers Dement. 2022, 18, 1260–1278. [Google Scholar] [CrossRef]
  102. Khan, M.J.; Chung, N.A.; Hansen, S.; Dumitrescu, L.; Hohman, T.J.; Kamboh, M.I.; Lopez, O.L.; Robinson, R.A.S. Targeted Lipidomics To Measure Phospholipids and Sphingomyelins in Plasma: A Pilot Study To Understand the Impact of Race/Ethnicity in Alzheimer’s Disease. Anal. Chem. 2022, 94, 4165–4174. [Google Scholar] [CrossRef]
  103. Maffioli, E.; Murtas, G.; Rabattoni, V.; Badone, B.; Tripodi, F.; Iannuzzi, F.; Licastro, D.; Nonnis, S.; Rinaldi, A.M.; Motta, Z.; et al. Insulin and serine metabolism as sex-specific hallmarks of Alzheimer’s disease in the human hippocampus. Cell Rep. 2022, 40, 111271. [Google Scholar] [CrossRef]
  104. Ozaki, T.; Yoshino, Y.; Tachibana, A.; Shimizu, H.; Mori, T.; Nakayama, T.; Mawatari, K.; Numata, S.; Iga, J.I.; Takahashi, A.; et al. Metabolomic alterations in the blood plasma of older adults with mild cognitive impairment and Alzheimer’s disease (from the Nakayama Study). Sci. Rep. 2022, 12, 15205. [Google Scholar] [CrossRef] [PubMed]
  105. Peña-Bautista, C.; Álvarez-Sánchez, L.; Roca, M.; García-Vallés, L.; Baquero, M.; Cháfer-Pericás, C. Plasma Lipidomics Approach in Early and Specific Alzheimer’s Disease Diagnosis. J. Clin. Med. 2022, 11, 5030. [Google Scholar] [CrossRef] [PubMed]
  106. Weng, J.; Muti, I.H.; Zhong, A.B.; Kivisäkk, P.; Hyman, B.T.; Arnold, S.E.; Cheng, L.L. A Nuclear Magnetic Resonance Spectroscopy Method in Characterization of Blood Metabolomics for Alzheimer’s Disease. Metabolites 2022, 12, 181. [Google Scholar] [CrossRef]
  107. Roche, S.; Gabelle, A.; Lehmann, S. Clinical proteomics of the cerebrospinal fluid: Towards the discovery of new biomarkers. Proteom. Clin. Appl. 2008, 2, 428–436. [Google Scholar] [CrossRef] [PubMed]
  108. Schraen-Maschke, S.; Sergeant, N.; Dhaenens, C.M.; Bombois, S.; Deramecourt, V.; Caillet-Boudin, M.L.; Pasquier, F.; Maurage, C.A.; Sablonnière, B.; Vanmechelen, E.; et al. Tau as a biomarker of neurodegenerative diseases. Biomark. Med. 2008, 2, 363–384. [Google Scholar] [CrossRef] [Green Version]
  109. Andreasen, N.; Blennow, K. CSF biomarkers for mild cognitive impairment and early Alzheimer’s disease. Clin. Neurol. Neurosurg. 2005, 107, 165–173. [Google Scholar] [CrossRef]
  110. Guzman-Martinez, L.; Maccioni, R.B.; Farías, G.A.; Fuentes, P.; Navarrete, L.P. Biomarkers for Alzheimer’s Disease. Curr. Alzheimer Res. 2019, 16, 518–528. [Google Scholar] [CrossRef] [Green Version]
  111. Henriksen, K.; O’Bryant, S.E.; Hampel, H.; Trojanowski, J.Q.; Montine, T.J.; Jeromin, A.; Blennow, K.; Lönneborg, A.; Wyss-Coray, T.; Soares, H.; et al. The future of blood-based biomarkers for Alzheimer’s disease. Alzheimers Dement. 2014, 10, 115–131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. de Lange, E.C.M.; Hammarlund Udenaes, M. Understanding the Blood-Brain Barrier and Beyond: Challenges and Opportunities for Novel CNS Therapeutics. Clin. Pharmacol. Ther. 2022, 111, 758–773. [Google Scholar] [CrossRef]
  113. Veitch, D.P.; Weiner, M.W.; Aisen, P.S.; Beckett, L.A.; DeCarli, C.; Green, R.C.; Harvey, D.; Jack, C.R., Jr.; Jagust, W.; Landau, S.M.; et al. Using the Alzheimer’s Disease Neuroimaging Initiative to improve early detection, diagnosis, and treatment of Alzheimer’s disease. Alzheimers Dement. 2022, 18, 824–857. [Google Scholar] [CrossRef]
  114. Arnold, M.; Nho, K.; Kueider-Paisley, A.; Massaro, T.; Huynh, K.; Brauner, B.; MahmoudianDehkordi, S.; Louie, G.; Moseley, M.A.; Thompson, J.W.; et al. Sex and APOE ε4 genotype modify the Alzheimer’s disease serum metabolome. Nat. Commun. 2020, 11, 1148. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Qin, T.; Prins, S.; Groeneveld, G.J.; Van Westen, G.; de Vries, H.E.; Wong, Y.C.; Bischoff, L.J.M.; de Lange, E.C.M. Utility of Animal Models to Understand Human Alzheimer’s Disease, Using the Mastermind Research Approach to Avoid Unnecessary Further Sacrifices of Animals. Int. J. Mol. Sci. 2020, 21, 3158. [Google Scholar] [CrossRef]
  116. Drummond, E.; Wisniewski, T. Alzheimer’s disease: Experimental models and reality. Acta Neuropathol. 2017, 133, 155–175. [Google Scholar] [CrossRef] [Green Version]
  117. Boutajangout, A.; Wisniewski, T. Tau-based therapeutic approaches for Alzheimer’s disease—A mini-review. Gerontology 2014, 60, 381–385. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Dujardin, S.; Colin, M.; Buée, L. Invited review: Animal models of tauopathies and their implications for research/translation into the clinic. Neuropathol. Appl. Neurobiol. 2015, 41, 59–80. [Google Scholar] [CrossRef]
  119. Puzzo, D.; Gulisano, W.; Palmeri, A.; Arancio, O. Rodent models for Alzheimer’s disease drug discovery. Expert Opin. Drug Discov. 2015, 10, 703–711. [Google Scholar] [CrossRef] [Green Version]
  120. Kitazawa, M.; Medeiros, R.; Laferla, F.M. Transgenic mouse models of Alzheimer disease: Developing a better model as a tool for therapeutic interventions. Curr. Pharm. Des. 2012, 18, 1131–1147. [Google Scholar] [CrossRef] [Green Version]
  121. Hamanaka, H.; Katoh-Fukui, Y.; Suzuki, K.; Kobayashi, M.; Suzuki, R.; Motegi, Y.; Nakahara, Y.; Takeshita, A.; Kawai, M.; Ishiguro, K.; et al. Altered cholesterol metabolism in human apolipoprotein E4 knock-in mice. Hum. Mol. Genet. 2000, 9, 353–361. [Google Scholar] [CrossRef] [PubMed]
  122. Tai, L.M.; Youmans, K.L.; Jungbauer, L.; Yu, C.; Ladu, M.J. Introducing Human APOE into Aβ Transgenic Mouse Models. Int. J. Alzheimers Dis. 2011, 2011, 810981. [Google Scholar] [PubMed] [Green Version]
  123. Salek, R.M.; Xia, J.; Innes, A.; Sweatman, B.C.; Adalbert, R.; Randle, S.; McGowan, E.; Emson, P.C.; Griffin, J.L. A metabolomic study of the CRND8 transgenic mouse model of Alzheimer’s disease. Neurochem. Int. 2010, 56, 937–947. [Google Scholar] [CrossRef]
  124. Hu, Z.P.; Browne, E.R.; Liu, T.; Angel, T.E.; Ho, P.C.; Chan, E.C. Metabonomic profiling of TASTPM transgenic Alzheimer’s disease mouse model. J. Proteome Res. 2012, 11, 5903–5913. [Google Scholar] [CrossRef]
  125. Trushina, E.; Nemutlu, E.; Zhang, S.; Christensen, T.; Camp, J.; Mesa, J.; Siddiqui, A.; Tamura, Y.; Sesaki, H.; Wengenack, T.M.; et al. Defects in mitochondrial dynamics and metabolomic signatures of evolving energetic stress in mouse models of familial Alzheimer’s disease. PLoS ONE 2012, 7, e32737. [Google Scholar] [CrossRef] [Green Version]
  126. Tajima, Y.; Ishikawa, M.; Maekawa, K.; Murayama, M.; Senoo, Y.; Nishimaki-Mogami, T.; Nakanishi, H.; Ikeda, K.; Arita, M.; Taguchi, R.; et al. Lipidomic analysis of brain tissues and plasma in a mouse model expressing mutated human amyloid precursor protein/tau for Alzheimer’s disease. Lipids Health Dis. 2013, 12, 68. [Google Scholar] [CrossRef] [Green Version]
  127. González-Domínguez, R.; García-Barrera, T.; Vitorica, J.; Gómez-Ariza, J.L. Region-specific metabolic alterations in the brain of the APP/PS1 transgenic mice of Alzheimer’s disease. Biochim. Biophys. Acta 2014, 1842 Pt A, 2395–2402. [Google Scholar] [CrossRef] [Green Version]
  128. Kim, E.; Jung, Y.S.; Kim, H.; Kim, J.S.; Park, M.; Jeong, J.; Lee, S.K.; Yoon, H.G.; Hwang, G.S.; Namkoong, K. Metabolomic signatures in peripheral blood associated with Alzheimer’s disease amyloid-beta-induced neuroinflammation. J. Alzheimers Dis. 2014, 42, 421–433. [Google Scholar] [CrossRef]
  129. Lalande, J.; Halley, H.; Balayssac, S.; Gilard, V.; Déjean, S.; Martino, R.; Francés, B.; Lassalle, J.M.; Malet-Martino, M. 1H NMR metabolomic signatures in five brain regions of the AβPPswe Tg2576 mouse model of Alzheimer’s disease at four ages. J. Alzheimers Dis. 2014, 39, 121–143. [Google Scholar] [CrossRef]
  130. Gonzalez-Dominguez, R.; Garcia-Barrera, T.; Vitorica, J.; Gomez-Ariza, J.L. Application of metabolomics based on direct mass spectrometry analysis for the elucidation of altered metabolic pathways in serum from the APP/PS1 transgenic model of Alzheimer’s disease. J. Pharm. Biomed. Anal. 2015, 107, 378–385. [Google Scholar] [CrossRef]
  131. González-Domínguez, R.; García-Barrera, T.; Vitorica, J.; Gómez-Ariza, J.L. Deciphering metabolic abnormalities associated with Alzheimer’s disease in the APP/PS1 mouse model using integrated metabolomic approaches. Biochimie 2015, 110, 119–128. [Google Scholar] [CrossRef] [Green Version]
  132. González-Domínguez, R.; García-Barrera, T.; Vitorica, J.; Gómez-Ariza, J.L. Metabolomic screening of regional brain alterations in the APP/PS1 transgenic model of Alzheimer’s disease by direct infusion mass spectrometry. J. Pharm. Biomed. Anal. 2015, 102, 425–435. [Google Scholar] [CrossRef] [PubMed]
  133. Li, N.; Zhou, L.; Li, W.; Liu, Y.; Wang, J.; He, P. Protective effects of ginsenosides Rg1 and Rb1 on an Alzheimer’s disease mouse model: A metabolomics study. J. Chromatogr. B 2015, 985, 54–61. [Google Scholar] [CrossRef]
  134. Li, N.; Liu, Y.; Li, W.; Zhou, L.; Li, Q.; Wang, X.; He, P. A UPLC/MS-based metabolomics investigation of the protective effect of ginsenosides Rg1 and Rg2 in mice with Alzheimer’s disease. J. Ginseng Res. 2016, 40, 9–17. [Google Scholar] [CrossRef] [Green Version]
  135. Pan, X.; Nasaruddin, M.B.; Elliott, C.T.; McGuinness, B.; Passmore, A.P.; Kehoe, P.G.; Hölscher, C.; McClean, P.L.; Graham, S.F.; Green, B.D. Alzheimer’s disease-like pathology has transient effects on the brain and blood metabolome. Neurobiol. Aging 2016, 38, 151–163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Nuriel, T.; Angulo, S.L.; Khan, U.; Ashok, A.; Chen, Q.; Figueroa, H.Y.; Emrani, S.; Liu, L.; Herman, M.; Barrett, G.; et al. Neuronal hyperactivity due to loss of inhibitory tone in APOE4 mice lacking Alzheimer’s disease-like pathology. Nat. Commun. 2017, 8, 1464. [Google Scholar] [CrossRef] [Green Version]
  137. Pan, X.; Elliott, C.T.; McGuinness, B.; Passmore, P.; Kehoe, P.G.; Hölscher, C.; McClean, P.L.; Graham, S.F.; Green, B.D. Metabolomic Profiling of Bile Acids in Clinical and Experimental Samples of Alzheimer’s Disease. Metabolites 2017, 7, 28. [Google Scholar] [CrossRef] [Green Version]
  138. Bergin, D.H.; Jing, Y.; Mockett, B.G.; Zhang, H.; Abraham, W.C.; Liu, P. Altered plasma arginine metabolome precedes behavioural and brain arginine metabolomic profile changes in the APPswe/PS1ΔE9 mouse model of Alzheimer’s disease. Transl. Psychiatry 2018, 8, 108. [Google Scholar] [CrossRef] [Green Version]
  139. Gao, H.L.; Zhang, A.H.; Yu, J.B.; Sun, H.; Kong, L.; Wang, X.Q.; Yan, G.L.; Liu, L.; Wang, X.J. High-throughput lipidomics characterize key lipid molecules as potential therapeutic targets of Kaixinsan protects against Alzheimer’s disease in APP/PS1 transgenic mice. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2018, 1092, 286–295. [Google Scholar] [CrossRef]
  140. Sun, L.M.; Zhu, B.J.; Cao, H.T.; Zhang, X.Y.; Zhang, Q.C.; Xin, G.Z.; Pan, L.M.; Liu, L.F.; Zhu, H.X. Explore the effects of Huang-Lian-Jie-Du-Tang on Alzheimer’s disease by UPLC-QTOF/MS-based plasma metabolomics study. J. Pharm. Biomed. Anal. 2018, 151, 75–83. [Google Scholar] [CrossRef]
  141. Zheng, H.; Zhou, Q.; Du, Y.; Li, C.; Xu, P.; Lin, L.; Xiao, J.; Gao, H. The hypothalamus as the primary brain region of metabolic abnormalities in APP/PS1 transgenic mouse model of Alzheimer’s disease. Biochim. Biophys. Acta BBA-Mol. Basis Dis. 2018, 1864, 263–273. [Google Scholar] [CrossRef]
  142. Zhou, Q.; Zheng, H.; Chen, J.; Li, C.; Du, Y.; Xia, H.; Gao, H. Metabolic fate of glucose in the brain of APP/PS1 transgenic mice at 10 months of age: A (13)C NMR metabolomic study. Metab. Brain Dis. 2018, 33, 1661–1668. [Google Scholar] [CrossRef] [PubMed]
  143. Chang, K.L.; Wong, L.R.; Pee, H.N.; Yang, S.; Ho, P.C. Reverting Metabolic Dysfunction in Cortex and Cerebellum of APP/PS1 Mice, a Model for Alzheimer’s Disease by Pioglitazone, a Peroxisome Proliferator-Activated Receptor Gamma (PPARγ) Agonist. Mol. Neurobiol. 2019, 56, 7267–7283. [Google Scholar] [CrossRef] [PubMed]
  144. Li, G.; Zhang, N.; Geng, F.; Liu, G.; Liu, B.; Lei, X.; Li, G.; Chen, X. High-throughput metabolomics and ingenuity pathway approach reveals the pharmacological effect and targets of Ginsenoside Rg1 in Alzheimer’s disease mice. Sci. Rep. 2019, 9, 7040. [Google Scholar] [CrossRef] [Green Version]
  145. Liu, X.; Wang, W.; Chen, H.L.; Zhang, H.Y.; Zhang, N.X. Interplay between Alzheimer’s disease and global glucose metabolism revealed by the metabolic profile alterations of pancreatic tissue and serum in APP/PS1 transgenic mice. Acta Pharmacol. Sin. 2019, 40, 1259–1268. [Google Scholar] [CrossRef]
  146. Liu, Y.; Du, T.; Zhang, W.; Lu, W.; Peng, Z.; Huang, S.; Sun, X.; Zhu, X.; Chen, C.; Qian, L.; et al. Modified Huang-Lian-Jie-Du Decoction Ameliorates Aβ Synaptotoxicity in a Murine Model of Alzheimer’s Disease. Oxid. Med. Cell Longev. 2019, 2019, 8340192. [Google Scholar] [CrossRef] [Green Version]
  147. Pan, X.; Green, B.D. Temporal Effects of Neuron-specific beta-secretase 1 (BACE1) Knock-in on the Mouse Brain Metabolome: Implications for Alzheimer’s Disease. Neuroscience 2019, 397, 138–146. [Google Scholar] [CrossRef] [Green Version]
  148. Rong, W.; Ding, K.; Guo, S.; Xie, F.; Li, Q.; Bi, K. Metabolomics analysis of Xanthoceras sorbifolia husks protection of rats against Alzheimer’s disease using liquid chromatography mass spectrometry. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2019, 1126–1127, 121739. [Google Scholar] [CrossRef]
  149. Hunsberger, H.C.; Greenwood, B.P.; Tolstikov, V.; Narain, N.R.; Kiebish, M.A.; Denny, C.A. Divergence in the metabolome between natural aging and Alzheimer’s disease. Sci. Rep. 2020, 10, 12171. [Google Scholar] [CrossRef]
  150. Kim, H.; Kim, B.; Kim, H.S.; Cho, J.Y. Nicotinamide attenuates the decrease in dendritic spine density in hippocampal primary neurons from 5xFAD mice, an Alzheimer’s disease animal model. Mol. Brain 2020, 13, 17. [Google Scholar] [CrossRef]
  151. Sun, W.; Liu, C.; Zhou, X.; Li, X.; Chu, X.; Wang, X.; Han, F. Serum lipidomics study reveals protective effects of Rhodiola crenulata extract on Alzheimer’s disease rats. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2020, 1158, 122346. [Google Scholar] [CrossRef] [PubMed]
  152. Tondo, M.; Wasek, B.; Escola-Gil, J.C.; de Gonzalo-Calvo, D.; Harmon, C.; Arning, E.; Bottiglieri, T. Altered Brain Metabolome Is Associated with Memory Impairment in the rTg4510 Mouse Model of Tauopathy. Metabolites 2020, 10, 69. [Google Scholar] [CrossRef] [Green Version]
  153. Yi, M.; Zhang, C.; Zhang, Z.; Yi, P.; Xu, P.; Huang, J.; Peng, W. Integrated Metabolomic and Lipidomic Analysis Reveals the Neuroprotective Mechanisms of Bushen Tiansui Formula in an Aβ1-42-Induced Rat Model of Alzheimer’s Disease. Oxid. Med. Cell. Longev. 2020, 2020, 5243453. [Google Scholar] [CrossRef]
  154. Zhang, X.; Liu, W.; Cao, Y.; Tan, W. Hippocampus Proteomics and Brain Lipidomics Reveal Network Dysfunction and Lipid Molecular Abnormalities in APP/PS1 Mouse Model of Alzheimer’s Disease. J. Proteome Res. 2020, 19, 3427–3437. [Google Scholar] [CrossRef] [PubMed]
  155. Zhang, X.; Liu, W.; Zan, J.; Wu, C.; Tan, W. Untargeted lipidomics reveals progression of early Alzheimer’s disease in APP/PS1 transgenic mice. Sci. Rep. 2020, 10, 14509. [Google Scholar] [CrossRef]
  156. Dejakaisaya, H.; Harutyunyan, A.; Kwan, P.; Jones, N.C. Altered metabolic pathways in a transgenic mouse model suggest mechanistic role of amyloid precursor protein overexpression in Alzheimer’s disease. Metabolomics 2021, 17, 42. [Google Scholar] [CrossRef]
  157. Speers, A.B.; García-Jaramillo, M.; Feryn, A.; Matthews, D.G.; Lichtenberg, T.; Caruso, M.; Wright, K.M.; Quinn, J.F.; Stevens, J.F.; Maier, C.S.; et al. Centella asiatica Alters Metabolic Pathways Associated With Alzheimer’s Disease in the 5xFAD Mouse Model of ß-Amyloid Accumulation. Front. Pharmacol. 2021, 12, 788312. [Google Scholar] [CrossRef] [PubMed]
  158. Sun, W.; Liu, C.; Wang, Y.; Zhou, X.; Sui, W.; Zhang, Y.; Zhang, Q.; Han, J.; Li, X.; Han, F. Rhodiola crenulata protects against Alzheimer’s disease in rats: A brain lipidomics study by Fourier-transform ion cyclotron resonance mass spectrometry coupled with high-performance reversed-phase liquid chromatography and hydrophilic interaction liquid chromatography. Rapid Commun. Mass Spectrom. RCM 2021, 35, e8969. [Google Scholar]
  159. Zhao, Y.; Chen, H.; Iqbal, J.; Liu, X.; Zhang, H.; Xiao, S.; Jin, N.; Yao, F.; Shen, L. Targeted metabolomics study of early pathological features in hippocampus of triple transgenic Alzheimer’s disease male mice. J. Neurosci. Res. 2021, 99, 927–946. [Google Scholar] [CrossRef]
  160. Cheng, X.; Tan, Y.; Li, H.; Huang, J.; Zhao, D.; Zhang, Z.; Yi, M.; Zhu, L.; Hui, S.; Yang, J.; et al. Fecal 16S rRNA sequencing and multi-compartment metabolomics revealed gut microbiota and metabolites interactions in APP/PS1 mice. Comput. Biol. Med. 2022, 151 Pt A, 106312. [Google Scholar] [CrossRef]
  161. Dai, Z.; Hu, T.; Su, S.; Liu, J.; Ma, Y.; Zhuo, Y.; Fang, S.; Wang, Q.; Mo, Z.; Pan, H.; et al. Comparative Metabolomics Analysis Reveals Key Metabolic Mechanisms and Protein Biomarkers in Alzheimer’s Disease. Front. Pharmacol. 2022, 13, 904857. [Google Scholar] [CrossRef] [PubMed]
  162. Dehghan, A.; Pinto, R.C.; Karaman, I.; Huang, J.; Durainayagam, B.R.; Ghanbari, M.; Nazeer, A.; Zhong, Q.; Liggi, S.; Whiley, L.; et al. Metabolome-wide association study on ABCA7 indicates a role of ceramide metabolism in Alzheimer’s disease. Proc. Natl. Acad. Sci. USA 2022, 119, e2206083119. [Google Scholar] [CrossRef] [PubMed]
  163. Dunham, S.J.B.; McNair, K.A.; Adams, E.D.; Avelar-Barragan, J.; Forner, S.; Mapstone, M.; Whiteson, K.L. Longitudinal Analysis of the Microbiome and Metabolome in the 5xfAD Mouse Model of Alzheimer’s Disease. mBio 2022, 13, e0179422. [Google Scholar] [CrossRef] [PubMed]
  164. Sun, P.; Zhu, H.; Li, X.; Shi, W.; Guo, Y.; Du, X.; Zhang, L.; Su, L.; Qin, C. Comparative Metagenomics and Metabolomes Reveals Abnormal Metabolism Activity Is Associated with Gut Microbiota in Alzheimer’s Disease Mice. Int. J. Mol. Sci. 2022, 23, 11560. [Google Scholar] [CrossRef] [PubMed]
  165. Trushina, E.; Mielke, M.M. Recent advances in the application of metabolomics to Alzheimer’s Disease. Biochim. Biophys. Acta BBA-Mol. Basis Dis. 2014, 1842, 1232–1239. [Google Scholar] [CrossRef] [Green Version]
  166. Xia, J.; Wishart, D.S. MetPA: A web-based metabolomics tool for pathway analysis and visualization. Bioinformatics 2010, 26, 2342–2344. [Google Scholar] [CrossRef] [Green Version]
  167. Wu, G.; Morris, S.M., Jr. Arginine metabolism: Nitric oxide and beyond. Biochem. J. 1998, 336 Pt 1, 1–17. [Google Scholar] [CrossRef]
  168. Morris, S.M., Jr. Enzymes of Arginine Metabolism. J. Nutr. 2004, 134, 2743S–2747S. [Google Scholar] [CrossRef] [Green Version]
  169. Aktan, F. iNOS-mediated nitric oxide production and its regulation. Life Sci. 2004, 75, 639–653. [Google Scholar] [CrossRef]
  170. Ellison, D.W.; Beal, M.F.; Mazurek, M.F.; Bird, E.D.; Martin, J.B. A postmortem study of amino acid neurotransmitters in Alzheimer’s disease. Ann. Neurol. 1986, 20, 616–621. [Google Scholar] [CrossRef]
  171. Tumani, H.; Shen, G.; Peter, J.B.; Brück, W. Glutamine synthetase in cerebrospinal fluid, serum, and brain: A diagnostic marker for Alzheimer disease? Arch. Neurol. 1999, 56, 1241–1246. [Google Scholar] [CrossRef] [Green Version]
  172. Satriano, J. Arginine pathways and the inflammatory response: Interregulation of nitric oxide and polyamines: Review article. Amino Acids 2004, 26, 321–329. [Google Scholar] [CrossRef]
  173. Zhou, L.; Zhu, D.Y. Neuronal nitric oxide synthase: Structure, subcellular localization, regulation, and clinical implications. Nitric Oxide 2009, 20, 223–230. [Google Scholar] [CrossRef] [PubMed]
  174. Susswein, A.J.; Katzoff, A.; Miller, N.; Hurwitz, I. Nitric oxide and memory. Neuroscientist 2004, 10, 153–162. [Google Scholar] [CrossRef] [PubMed]
  175. Feil, R.; Kleppisch, T. NO/cGMP-dependent modulation of synaptic transmission. In Pharmacology of Neurotransmitter Release; Handbook of Experimental Pharmacology; Springer: Berlin/Heidelberg, Germany, 2008; pp. 529–560. [Google Scholar]
  176. Böger, R.H.; Bode-Böger, S.M.; Frölich, J.C. The L-arginine-nitric oxide pathway: Role in atherosclerosis and therapeutic implications. Atherosclerosis 1996, 127, 1–11. [Google Scholar] [CrossRef] [PubMed]
  177. Cooke, J.P.; Dzau, V.J. Nitric oxide synthase: Role in the genesis of vascular disease. Annu. Rev. Med. 1997, 48, 489–509. [Google Scholar] [CrossRef] [Green Version]
  178. Cooke, J.P. The pivotal role of nitric oxide for vascular health. Can. J. Cardiol. 2004, 20 (Suppl. B), 7b–15b. [Google Scholar]
  179. Li, X.A.; Everson, W.; Smart, E.J. Nitric oxide, caveolae, and vascular pathology. Cardiovasc. Toxicol. 2006, 6, 1–13. [Google Scholar] [CrossRef] [PubMed]
  180. Napoli, C.; de Nigris, F.; Williams-Ignarro, S.; Pignalosa, O.; Sica, V.; Ignarro, L.J. Nitric oxide and atherosclerosis: An update. Nitric Oxide 2006, 15, 265–279. [Google Scholar] [CrossRef] [PubMed]
  181. Fonnum, F. Glutamate: A neurotransmitter in mammalian brain. J. Neurochem. 1984, 42, 1–11. [Google Scholar] [CrossRef]
  182. Francis, P.T.; Sims, N.R.; Procter, A.W.; Bowen, D.M. Cortical pyramidal neurone loss may cause glutamatergic hypoactivity and cognitive impairment in Alzheimer’s disease: Investigative and therapeutic perspectives. J. Neurochem. 1993, 60, 1589–1604. [Google Scholar] [CrossRef]
  183. Bruno, V.; Battaglia, G.; Copani, A.; D’Onofrio, M.; Di Iorio, P.; De Blasi, A.; Melchiorri, D.; Flor, P.J.; Nicoletti, F. Metabotropic glutamate receptor subtypes as targets for neuroprotective drugs. J. Cereb. Blood Flow Metab. 2001, 21, 1013–1033. [Google Scholar] [CrossRef] [Green Version]
  184. Gadea, A.; López-Colomé, A.M. Glial transporters for glutamate, glycine and GABA I. Glutamate transporters. J. Neurosci. Res. 2001, 63, 453–460. [Google Scholar] [CrossRef] [PubMed]
  185. Ozawa, S.; Kamiya, H.; Tsuzuki, K. Glutamate receptors in the mammalian central nervous system. Prog. Neurobiol. 1998, 54, 581–618. [Google Scholar] [CrossRef] [PubMed]
  186. Lynch, G. Memory and the brain: Unexpected chemistries and a new pharmacology. Neurobiol. Learn. Mem. 1998, 70, 82–100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Riedel, G.; Micheau, J. Function of the hippocampus in memory formation: Desperately seeking resolution. Prog. Neuropsychopharmacol. Biol. Psychiatry 2001, 25, 835–853. [Google Scholar] [CrossRef]
  188. Myhrer, T. Effects of selective perirhinal and postrhinal lesions on acquisition and retention of a visual discrimination task in rats. Neurobiol. Learn. Mem. 2000, 73, 68–78. [Google Scholar] [CrossRef] [Green Version]
  189. Baudry, M.; Lynch, G. Remembrance of arguments past: How well is the glutamate receptor hypothesis of LTP holding up after 20 years? Neurobiol. Learn. Mem. 2001, 76, 284–297. [Google Scholar] [CrossRef] [Green Version]
  190. Jay, T.M.; Zilkha, E.; Obrenovitch, T.P. Long-term potentiation in the dentate gyrus is not linked to increased extracellular glutamate concentration. J. Neurophysiol. 1999, 81, 1741–1748. [Google Scholar] [CrossRef]
  191. Scannevin, R.H.; Huganir, R.L. Postsynaptic organization and regulation of excitatory synapses. Nat. Rev. Neurosci. 2000, 1, 133–141. [Google Scholar] [CrossRef]
  192. Ansoleaga, B.; Jové, M.; Schlüter, A.; Garcia-Esparcia, P.; Moreno, J.; Pujol, A.; Pamplona, R.; Portero-Otín, M.; Ferrer, I. Deregulation of purine metabolism in Alzheimer’s disease. Neurobiol. Aging 2015, 36, 68–80. [Google Scholar] [CrossRef] [PubMed]
  193. Ipata, P.L.; Camici, M.; Micheli, V.; Tozz, M.G. Metabolic network of nucleosides in the brain. Curr. Top. Med. Chem. 2011, 11, 909–922. [Google Scholar] [CrossRef] [PubMed]
  194. Ferreira, I.L.; Resende, R.; Ferreiro, E.; Rego, A.C.; Pereira, C.F. Multiple defects in energy metabolism in Alzheimer’s disease. Curr. Drug Targets 2010, 11, 1193–1206. [Google Scholar] [CrossRef] [PubMed]
  195. Ferrer, I. Altered mitochondria, energy metabolism, voltage-dependent anion channel, and lipid rafts converge to exhaust neurons in Alzheimer’s disease. J. Bioenerg. Biomembr. 2009, 41, 425–431. [Google Scholar] [CrossRef]
  196. Lovell, M.A.; Markesbery, W.R. Oxidatively modified RNA in mild cognitive impairment. Neurobiol. Dis. 2008, 29, 169–175. [Google Scholar] [CrossRef] [Green Version]
  197. Lovell, M.A.; Soman, S.; Bradley, M.A. Oxidatively modified nucleic acids in preclinical Alzheimer’s disease (PCAD) brain. Mech. Ageing Dev. 2011, 132, 443–448. [Google Scholar] [CrossRef] [Green Version]
  198. Markesbery, W.R.; Lovell, M.A. DNA oxidation in Alzheimer’s disease. Antioxid. Redox Signal. 2006, 8, 2039–2045. [Google Scholar] [CrossRef] [PubMed]
  199. Nunomura, A.; Tamaoki, T.; Motohashi, N.; Nakamura, M.; McKeel, D.W., Jr.; Tabaton, M.; Lee, H.G.; Smith, M.A.; Perry, G.; Zhu, X. The earliest stage of cognitive impairment in transition from normal aging to Alzheimer disease is marked by prominent RNA oxidation in vulnerable neurons. J. Neuropathol. Exp. Neurol. 2012, 71, 233–241. [Google Scholar] [CrossRef]
  200. Kaddurah-Daouk, R.; Rozen, S.; Matson, W.; Han, X.; Hulette, C.M.; Burke, J.R.; Doraiswamy, P.M.; Welsh-Bohmer, K.A. Metabolomic changes in autopsy-confirmed Alzheimer’s disease. Alzheimers Dement. 2011, 7, 309–317. [Google Scholar] [CrossRef] [Green Version]
  201. Isobe, C.; Abe, T.; Terayama, Y. Levels of reduced and oxidized coenzyme Q-10 and 8-hydroxy-2’-deoxyguanosine in the CSF of patients with Alzheimer’s disease demonstrate that mitochondrial oxidative damage and/or oxidative DNA damage contributes to the neurodegenerative process. J. Neurol. 2010, 257, 399–404. [Google Scholar] [CrossRef]
  202. Kaddurah-Daouk, R.; Zhu, H.; Sharma, S.; Bogdanov, M.; Rozen, S.G.; Matson, W.; Oki, N.O.; Motsinger-Reif, A.A.; Churchill, E.; Lei, Z.; et al. Alterations in metabolic pathways and networks in Alzheimer’s disease. Transl. Psychiatry 2013, 3, e244. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Jové, M.; Portero-Otín, M.; Naudí, A.; Ferrer, I.; Pamplona, R. Metabolomics of human brain aging and age-related neurodegenerative diseases. J. Neuropathol. Exp. Neurol. 2014, 73, 640–657. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Frosini, M.; Sesti, C.; Saponara, S.; Ricci, L.; Valoti, M.; Palmi, M.; Machetti, F.; Sgaragli, G. A specific taurine recognition site in the rabbit brain is responsible for taurine effects on thermoregulation. Br. J. Pharmacol. 2003, 139, 487–494. [Google Scholar] [CrossRef] [Green Version]
  205. Bhat, M.A.; Ahmad, K.; Khan, M.S.A.; Bhat, M.A.; Almatroudi, A.; Rahman, S.; Jan, A.T. Expedition into Taurine Biology: Structural Insights and Therapeutic Perspective of Taurine in Neurodegenerative Diseases. Biomolecules 2020, 10, 863. [Google Scholar] [CrossRef]
  206. Qaradakhi, T.; Gadanec, L.K.; McSweeney, K.R.; Abraham, J.R.; Apostolopoulos, V.; Zulli, A. The Anti-Inflammatory Effect of Taurine on Cardiovascular Disease. Nutrients 2020, 12, 2847. [Google Scholar] [CrossRef]
  207. Schaffer, S.W.; Azuma, J.; Mozaffari, M. Role of antioxidant activity of taurine in diabetes. Can. J. Physiol. Pharmacol. 2009, 87, 91–99. [Google Scholar] [CrossRef]
  208. Schaffer, S.; Takahashi, K.; Azuma, J. Role of osmoregulation in the actions of taurine. Amino Acids 2000, 19, 527–546. [Google Scholar] [CrossRef] [PubMed]
  209. Foos, T.M.; Wu, J.Y. The role of taurine in the central nervous system and the modulation of intracellular calcium homeostasis. Neurochem. Res. 2002, 27, 21–26. [Google Scholar] [CrossRef]
  210. Vohra, B.P.; Hui, X. Improvement of impaired memory in mice by taurine. Neural Plast. 2000, 7, 245–259. [Google Scholar] [CrossRef]
  211. Su, Y.; Fan, W.; Ma, Z.; Wen, X.; Wang, W.; Wu, Q.; Huang, H. Taurine improves functional and histological outcomes and reduces inflammation in traumatic brain injury. Neuroscience 2014, 266, 56–65. [Google Scholar] [CrossRef]
  212. Malcangio, M.; Bartolini, A.; Ghelardini, C.; Bennardini, F.; Malmberg-Aiello, P.; Franconi, F.; Giotti, A. Effect of ICV taurine on the impairment of learning, convulsions and death caused by hypoxia. Psychopharmacology 1989, 98, 316–320. [Google Scholar] [CrossRef]
  213. Javed, H.; Khan, A.; Vaibhav, K.; Moshahid Khan, M.; Ahmad, A.; Ejaz Ahmad, M.; Ahmad, A.; Tabassum, R.; Islam, F.; Safhi, M.M.; et al. Taurine ameliorates neurobehavioral, neurochemical and immunohistochemical changes in sporadic dementia of Alzheimer’s type (SDAT) caused by intracerebroventricular streptozotocin in rats. Neurol. Sci. 2013, 34, 2181–2192. [Google Scholar] [CrossRef] [PubMed]
  214. Santa-María, I.; Hernández, F.; Moreno, F.J.; Avila, J. Taurine, an inducer for tau polymerization and a weak inhibitor for amyloid-beta-peptide aggregation. Neurosci. Lett. 2007, 429, 91–94. [Google Scholar] [CrossRef] [PubMed]
  215. Pan, C.; Prentice, H.; Price, A.L.; Wu, J.Y. Beneficial effect of taurine on hypoxia- and glutamate-induced endoplasmic reticulum stress pathways in primary neuronal culture. Amino Acids 2012, 43, 845–855. [Google Scholar] [CrossRef]
  216. Ferreira-Vieira, T.H.; Guimaraes, I.M.; Silva, F.R.; Ribeiro, F.M. Alzheimer’s disease: Targeting the Cholinergic System. Curr. Neuropharmacol. 2016, 14, 101–115. [Google Scholar] [CrossRef] [Green Version]
  217. Kása, P.; Rakonczay, Z.; Gulya, K. The cholinergic system in Alzheimer’s disease. Prog. Neurobiol. 1997, 52, 511–535. [Google Scholar] [CrossRef]
  218. Hampel, H.; Mesulam, M.M.; Cuello, A.C.; Farlow, M.R.; Giacobini, E.; Grossberg, G.T.; Khachaturian, A.S.; Vergallo, A.; Cavedo, E.; Snyder, P.J.; et al. The cholinergic system in the pathophysiology and treatment of Alzheimer’s disease. Brain 2018, 141, 1917–1933. [Google Scholar] [CrossRef] [PubMed]
  219. de Carvalho, C.; Caramujo, M.J. The Various Roles of Fatty Acids. Molecules 2018, 23, 2583. [Google Scholar] [CrossRef] [Green Version]
  220. Kao, Y.C.; Ho, P.C.; Tu, Y.K.; Jou, I.M.; Tsai, K.J. Lipids and Alzheimer’s Disease. Int. J. Mol. Sci. 2020, 21, 1505. [Google Scholar] [CrossRef]
  221. Fonteh, A.N.; Cipolla, M.; Chiang, A.J.; Edminster, S.P.; Arakaki, X.; Harrington, M.G. Polyunsaturated Fatty Acid Composition of Cerebrospinal Fluid Fractions Shows Their Contribution to Cognitive Resilience of a Pre-symptomatic Alzheimer’s Disease Cohort. Front. Physiol. 2020, 11, 83. [Google Scholar] [CrossRef]
  222. Cunnane, S.C.; Schneider, J.A.; Tangney, C.; Tremblay-Mercier, J.; Fortier, M.; Bennett, D.A.; Morris, M.C. Plasma and brain fatty acid profiles in mild cognitive impairment and Alzheimer’s disease. J. Alzheimers Dis. 2012, 29, 691–697. [Google Scholar] [CrossRef]
  223. Kalmijn, S.; Launer, L.J.; Ott, A.; Witteman, J.C.; Hofman, A.; Breteler, M.M. Dietary fat intake and the risk of incident dementia in the Rotterdam Study. Ann. Neurol. 1997, 42, 776–782. [Google Scholar] [CrossRef] [Green Version]
  224. Morris, M.C.; Evans, D.A.; Bienias, J.L.; Tangney, C.C.; Bennett, D.A.; Wilson, R.S.; Aggarwal, N.; Schneider, J. Consumption of fish and n-3 fatty acids and risk of incident Alzheimer disease. Arch. Neurol. 2003, 60, 940–946. [Google Scholar] [CrossRef]
  225. Huang, T.L.; Zandi, P.P.; Tucker, K.L.; Fitzpatrick, A.L.; Kuller, L.H.; Fried, L.P.; Burke, G.L.; Carlson, M.C. Benefits of fatty fish on dementia risk are stronger for those without APOE epsilon4. Neurology 2005, 65, 1409–1414. [Google Scholar] [CrossRef]
  226. Sanchez-Mejia, R.O.; Mucke, L. Phospholipase A2 and arachidonic acid in Alzheimer’s disease. Biochim. Biophys. Acta 2010, 1801, 784–790. [Google Scholar] [CrossRef] [Green Version]
  227. Prasad, M.R.; Lovell, M.A.; Yatin, M.; Dhillon, H.; Markesbery, W.R. Regional membrane phospholipid alterations in Alzheimer’s disease. Neurochem. Res. 1998, 23, 81–88. [Google Scholar] [CrossRef]
  228. Proitsi, P.; Kim, M.; Whiley, L.; Simmons, A.; Sattlecker, M.; Velayudhan, L.; Lupton, M.K.; Soininen, H.; Kloszewska, I.; Mecocci, P.; et al. Association of blood lipids with Alzheimer’s disease: A comprehensive lipidomics analysis. Alzheimers Dement. 2017, 13, 140–151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  229. Chan, R.B.; Oliveira, T.G.; Cortes, E.P.; Honig, L.S.; Duff, K.E.; Small, S.A.; Wenk, M.R.; Shui, G.; Di Paolo, G. Comparative lipidomic analysis of mouse and human brain with Alzheimer disease. J. Biol. Chem. 2012, 287, 2678–2688. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  230. Wood, P.L.; Medicherla, S.; Sheikh, N.; Terry, B.; Phillipps, A.; Kaye, J.A.; Quinn, J.F.; Woltjer, R.L. Targeted Lipidomics of Fontal Cortex and Plasma Diacylglycerols (DAG) in Mild Cognitive Impairment and Alzheimer’s Disease: Validation of DAG Accumulation Early in the Pathophysiology of Alzheimer’s Disease. J. Alzheimers Dis. 2015, 48, 537–546. [Google Scholar] [CrossRef] [Green Version]
  231. Wood, P.L.; Barnette, B.L.; Kaye, J.A.; Quinn, J.F.; Woltjer, R.L. Non-targeted lipidomics of CSF and frontal cortex grey and white matter in control, mild cognitive impairment, and Alzheimer’s disease subjects. Acta Neuropsychiatr. 2015, 27, 270–278. [Google Scholar] [CrossRef] [PubMed]
  232. Farooqui, A.A.; Horrocks, L.A.; Farooqui, T. Glycerophospholipids in brain: Their metabolism, incorporation into membranes, functions, and involvement in neurological disorders. Chem. Phys. Lipids 2000, 106, 1–29. [Google Scholar] [CrossRef] [PubMed]
  233. Bargui, R.; Solgadi, A.; Prost, B.; Chester, M.; Ferreiro, A.; Piquereau, J.; Moulin, M. Phospholipids: Identification and Implication in Muscle Pathophysiology. Int. J. Mol. Sci. 2021, 22, 8176. [Google Scholar] [CrossRef]
  234. Pettegrew, J.W.; Panchalingam, K.; Hamilton, R.L.; McClure, R.J. Brain membrane phospholipid alterations in Alzheimer’s disease. Neurochem. Res. 2001, 26, 771–782. [Google Scholar] [CrossRef] [PubMed]
  235. Igarashi, M.; Ma, K.; Gao, F.; Kim, H.W.; Rapoport, S.I.; Rao, J.S. Disturbed choline plasmalogen and phospholipid fatty acid concentrations in Alzheimer’s disease prefrontal cortex. J. Alzheimers Dis. 2011, 24, 507–517. [Google Scholar] [CrossRef] [PubMed]
  236. Han, X.; Holtzman, D.M.; McKeel, D.W., Jr. Plasmalogen deficiency in early Alzheimer’s disease subjects and in animal models: Molecular characterization using electrospray ionization mass spectrometry. J. Neurochem. 2001, 77, 1168–1180. [Google Scholar] [CrossRef]
  237. Guan, Z.; Wang, Y.; Cairns, N.J.; Lantos, P.L.; Dallner, G.; Sindelar, P.J. Decrease and structural modifications of phosphatidylethanolamine plasmalogen in the brain with Alzheimer disease. J. Neuropathol. Exp. Neurol. 1999, 58, 740–747. [Google Scholar] [CrossRef] [Green Version]
  238. Wells, K.; Farooqui, A.A.; Liss, L.; Horrocks, L.A. Neural membrane phospholipids in Alzheimer disease. Neurochem. Res. 1995, 20, 1329–1333. [Google Scholar] [CrossRef]
  239. Söderberg, M.; Edlund, C.; Kristensson, K.; Dallner, G. Fatty acid composition of brain phospholipids in aging and in Alzheimer’s disease. Lipids 1991, 26, 421–425. [Google Scholar] [CrossRef]
  240. Stephenson, D.T.; Lemere, C.A.; Selkoe, D.J.; Clemens, J.A. Cytosolic phospholipase A2 (cPLA2) immunoreactivity is elevated in Alzheimer’s disease brain. Neurobiol. Dis. 1996, 3, 51–63. [Google Scholar] [CrossRef] [Green Version]
  241. Gattaz, W.F.; Maras, A.; Cairns, N.J.; Levy, R.; Förstl, H. Decreased phospholipase A2 activity in Alzheimer brains. Biol. Psychatry 1995, 37, 13–17. [Google Scholar] [CrossRef]
  242. Schaeffer, E.L.; Gattaz, W.F. Requirement of hippocampal phospholipase A2 activity for long-term memory retrieval in rats. J. Neural Transm. 2007, 114, 379–385. [Google Scholar] [CrossRef] [PubMed]
  243. Kim, H.Y.; Huang, B.X.; Spector, A.A. Phosphatidylserine in the brain: Metabolism and function. Prog. Lipid Res. 2014, 56, 1–18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. Zhang, Y.Y.; Yang, L.Q.; Guo, L.M. Effect of phosphatidylserine on memory in patients and rats with Alzheimer’s disease. Genet. Mol. Res. 2015, 14, 9325–9333. [Google Scholar] [CrossRef]
  245. Kidd, P.M. Omega-3 DHA and EPA for cognition, behavior, and mood: Clinical findings and structural-functional synergies with cell membrane phospholipids. Altern. Med. Rev. 2007, 12, 207–227. [Google Scholar]
  246. Frisardi, V.; Panza, F.; Seripa, D.; Farooqui, T.; Farooqui, A.A. Glycerophospholipids and glycerophospholipid-derived lipid mediators: A complex meshwork in Alzheimer’s disease pathology. Prog. Lipid Res. 2011, 50, 313–330. [Google Scholar] [CrossRef] [PubMed]
  247. Schaeffer, E.L.; da Silva, E.R.; Novaes Bde, A.; Skaf, H.D.; Gattaz, W.F. Differential roles of phospholipases A2 in neuronal death and neurogenesis: Implications for Alzheimer disease. Prog. Neuropsychopharmacol. Biol. Psychiatry 2010, 34, 1381–1389. [Google Scholar] [CrossRef]
  248. Perttu, E.K.; Kohli, A.G.; Szoka, F.C., Jr. Inverse-phosphocholine lipids: A remix of a common phospholipid. J. Am. Chem. Soc. 2012, 134, 4485–4488. [Google Scholar] [CrossRef] [Green Version]
  249. Whiley, L.; Sen, A.; Heaton, J.; Proitsi, P.; García-Gómez, D.; Leung, R.; Smith, N.; Thambisetty, M.; Kloszewska, I.; Mecocci, P.; et al. Evidence of altered phosphatidylcholine metabolism in Alzheimer’s disease. Neurobiol. Aging 2014, 35, 271–278. [Google Scholar] [CrossRef]
  250. Kang, S.; Han, J.; Song, S.Y.; Kim, W.S.; Shin, S.; Kim, J.H.; Ahn, H.; Jeong, J.H.; Hwang, S.J.; Sung, J.H. Lysophosphatidic acid increases the proliferation and migration of adipose-derived stem cells via the generation of reactive oxygen species. Mol. Med. Rep. 2015, 12, 5203–5210. [Google Scholar] [CrossRef] [Green Version]
  251. Jones, E.E.; Dworski, S.; Canals, D.; Casas, J.; Fabrias, G.; Schoenling, D.; Levade, T.; Denlinger, C.; Hannun, Y.A.; Medin, J.A.; et al. On-tissue localization of ceramides and other sphingolipids by MALDI mass spectrometry imaging. Anal. Chem. 2014, 86, 8303–8311. [Google Scholar] [CrossRef] [Green Version]
  252. Jazvinšćak Jembrek, M.; Hof, P.R.; Šimić, G. Ceramides in Alzheimer’s Disease: Key Mediators of Neuronal Apoptosis Induced by Oxidative Stress and Aβ Accumulation. Oxid. Med. Cell. Longev. 2015, 2015, 346783. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  253. Panchal, M.; Gaudin, M.; Lazar, A.N.; Salvati, E.; Rivals, I.; Ayciriex, S.; Dauphinot, L.; Dargère, D.; Auzeil, N.; Masserini, M.; et al. Ceramides and sphingomyelinases in senile plaques. Neurobiol. Dis. 2014, 65, 193–201. [Google Scholar] [CrossRef]
  254. Mielke, M.M.; Bandaru, V.V.; Haughey, N.J.; Xia, J.; Fried, L.P.; Yasar, S.; Albert, M.; Varma, V.; Harris, G.; Schneider, E.B.; et al. Serum ceramides increase the risk of Alzheimer disease: The Women’s Health and Aging Study II. Neurology 2012, 79, 633–641. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  255. Satoi, H.; Tomimoto, H.; Ohtani, R.; Kitano, T.; Kondo, T.; Watanabe, M.; Oka, N.; Akiguchi, I.; Furuya, S.; Hirabayashi, Y.; et al. Astroglial expression of ceramide in Alzheimer’s disease brains: A role during neuronal apoptosis. Neuroscience 2005, 130, 657–666. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Couttas, T.A.; Kain, N.; Daniels, B.; Lim, X.Y.; Shepherd, C.; Kril, J.; Pickford, R.; Li, H.; Garner, B.; Don, A.S. Loss of the neuroprotective factor Sphingosine 1-phosphate early in Alzheimer’s disease pathogenesis. Acta Neuropathol. Commun. 2014, 2, 9. [Google Scholar] [CrossRef] [Green Version]
  257. He, X.; Huang, Y.; Li, B.; Gong, C.X.; Schuchman, E.H. Deregulation of sphingolipid metabolism in Alzheimer’s disease. Neurobiol. Aging 2010, 31, 398–408. [Google Scholar] [CrossRef] [Green Version]
  258. Feringa, F.M.; van der Kant, R. Cholesterol and Alzheimer’s Disease; From Risk Genes to Pathological Effects. Front. Aging Neurosci. 2021, 13, 690372. [Google Scholar] [CrossRef]
  259. Cutler, R.G.; Kelly, J.; Storie, K.; Pedersen, W.A.; Tammara, A.; Hatanpaa, K.; Troncoso, J.C.; Mattson, M.P. Involvement of oxidative stress-induced abnormalities in ceramide and cholesterol metabolism in brain aging and Alzheimer’s disease. Proc. Natl. Acad. Sci. USA 2004, 101, 2070–2075. [Google Scholar] [CrossRef] [Green Version]
  260. Heverin, M.; Bogdanovic, N.; Lütjohann, D.; Bayer, T.; Pikuleva, I.; Bretillon, L.; Diczfalusy, U.; Winblad, B.; Björkhem, I. Changes in the levels of cerebral and extracerebral sterols in the brain of patients with Alzheimer’s disease. J. Lipid Res. 2004, 45, 186–193. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  261. Mori, T.; Paris, D.; Town, T.; Rojiani, A.M.; Sparks, D.L.; Delledonne, A.; Crawford, F.; Abdullah, L.I.; Humphrey, J.A.; Dickson, D.W.; et al. Cholesterol accumulates in senile plaques of Alzheimer disease patients and in transgenic APP(SW) mice. J. Neuropathol. Exp. Neurol. 2001, 60, 778–785. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Metabolic pathway analysis in biological samples (CSF, brain, and plasma) of 43 human AD studies. The size of the circle corresponds to the pathway impact score and was correlated with the centrality of the involved metabolites. Darker circle colors indicated more significant changes in metabolites in the corresponding pathway.
Figure 1. Metabolic pathway analysis in biological samples (CSF, brain, and plasma) of 43 human AD studies. The size of the circle corresponds to the pathway impact score and was correlated with the centrality of the involved metabolites. Darker circle colors indicated more significant changes in metabolites in the corresponding pathway.
Ijms 24 04960 g001
Figure 2. Disturbed metabolic pathways analysis in humans, comparing different disease stages with healthy controls: (A) Intersection analysis of metabolic pathways among MCI and AD groups in plasma samples. Yellow circles are all the disturbed pathways in MCI plasma samples, blue circles are all the disturbed pathways in AD plasma samples, and the middle intersection is pathways in common among MCI and AD in plasma samples. (B) Intersection analysis of metabolic pathways among MCI and AD groups in CSF samples. Yellow circles are all the disturbed pathways in MCI CSF samples, blue circles are all the disturbed pathways in AD CSF samples. All the MCI pathways overlapped with AD pathways in CSF samples. (C) Intersection analysis of metabolic pathways among plasma and CSF samples. Yellow circles are all the disturbed pathways in plasma samples, blue circles are all the disturbed pathways in CSF samples, and the middle intersection is pathways in common among plasma and CSF samples.
Figure 2. Disturbed metabolic pathways analysis in humans, comparing different disease stages with healthy controls: (A) Intersection analysis of metabolic pathways among MCI and AD groups in plasma samples. Yellow circles are all the disturbed pathways in MCI plasma samples, blue circles are all the disturbed pathways in AD plasma samples, and the middle intersection is pathways in common among MCI and AD in plasma samples. (B) Intersection analysis of metabolic pathways among MCI and AD groups in CSF samples. Yellow circles are all the disturbed pathways in MCI CSF samples, blue circles are all the disturbed pathways in AD CSF samples. All the MCI pathways overlapped with AD pathways in CSF samples. (C) Intersection analysis of metabolic pathways among plasma and CSF samples. Yellow circles are all the disturbed pathways in plasma samples, blue circles are all the disturbed pathways in CSF samples, and the middle intersection is pathways in common among plasma and CSF samples.
Ijms 24 04960 g002
Figure 3. Metabolic pathway analysis in biological samples (brain and plasma) of 42 mouse studies. The size of the circle corresponds to the pathway impact score and was correlated with the centrality of the involved metabolites. Darker circle colors indicated more significant changes in metabolites in the corresponding pathway.
Figure 3. Metabolic pathway analysis in biological samples (brain and plasma) of 42 mouse studies. The size of the circle corresponds to the pathway impact score and was correlated with the centrality of the involved metabolites. Darker circle colors indicated more significant changes in metabolites in the corresponding pathway.
Ijms 24 04960 g003
Figure 4. Altered metabolic pathways in mouse plasma and brain samples, comparing AD and control animals. (A) Heatmap of the changes of significantly altered metabolic pathways at different ages of AD mouse in brain and plasma samples. The impact value is calculated as the sum of the importance measures of matched metabolites normalized by the sum of the importance measures of all metabolites in each pathway. (B) Altered pathways in AD mouse plasma and brain samples combining all age groups. Yellow circles are all the disturbed pathways in mouse plasma samples, blue circles are all the disturbed pathways in mouse brain samples, and the middle intersection is pathways in common among plasma and brain samples in mouse research.
Figure 4. Altered metabolic pathways in mouse plasma and brain samples, comparing AD and control animals. (A) Heatmap of the changes of significantly altered metabolic pathways at different ages of AD mouse in brain and plasma samples. The impact value is calculated as the sum of the importance measures of matched metabolites normalized by the sum of the importance measures of all metabolites in each pathway. (B) Altered pathways in AD mouse plasma and brain samples combining all age groups. Yellow circles are all the disturbed pathways in mouse plasma samples, blue circles are all the disturbed pathways in mouse brain samples, and the middle intersection is pathways in common among plasma and brain samples in mouse research.
Ijms 24 04960 g004
Figure 5. Arginine metabolic pathways. L-arginine can be metabolized by phosphatidic acid (PA), nitric oxide synthase (NOS), arginase, and arginine decarboxylase (ADC) to form several bioactive molecules. (ADC, Arginine decarboxylase; ADMA, NG-dimethyl-L-arginine; DDAH, dimethylarginine dimethylaminohydrolase; GABA, γ-aminobutyric acid; ODC, ornithine decarboxylase;).
Figure 5. Arginine metabolic pathways. L-arginine can be metabolized by phosphatidic acid (PA), nitric oxide synthase (NOS), arginase, and arginine decarboxylase (ADC) to form several bioactive molecules. (ADC, Arginine decarboxylase; ADMA, NG-dimethyl-L-arginine; DDAH, dimethylarginine dimethylaminohydrolase; GABA, γ-aminobutyric acid; ODC, ornithine decarboxylase;).
Ijms 24 04960 g005
Table 1. Summary of the included articles related to metabolomics studies performed on human subjects.
Table 1. Summary of the included articles related to metabolomics studies performed on human subjects.
No.ReferenceStudy
Population
Sample
Type
Analytical PlatformAltered Metabolites
1Czech, et al., 2012 [63]51 HC,
79 AD
CSFLC-MS, GC-MSCitrulline, Cortisol, Creatinine, Cysteine, Dopamine, Erythrol, Galactitol, Histamine, Methionine, Noradrenaline, Normetanephrine, Phenylalanine, Pseudouridine, Pyruvate (incl. Phosphoenolpyruvate), Quinic acid (incl. Chlorogenic acid), Ribonic acid, Scyllo-inositol, Serine, Sorbitol (incl. Mannitol, Galactitol), Tyrosine, Uridine, Ornithine (incl. Arginine, Citrulline)
2Ibáñez, et al., 2012 [64]19 HC,
22 MCI,
9 MCI-AD,
23 AD
CSFCE-TOF-MSCholine, Valine, Arginine, Tripeptide, Carnitine, Dimethy-L-arginine, Creatine
3Sato, et al., 2012 [65]10 HC,
10 AD
Plasma, CSFLC/APCI-MSCholesterol, Desmosterol
4Ibáñez, et al., 2013 [66]21 HC,
21 MCI-S,
12 MCI-AD,
21 AD
CSFUHPLC-TOF-MSUracil, Xanthine, Uridine, Tyrosyl-serine, Methylsalsolinol, Nonanoylglycine, Dopamine-quinone, Caproic acid, Vanylglycol, Histidine, Pipecolic acid, Hydroxyphosphinyl-pyruvate, Creatinine, Taurine, Sphingosine-1-phosphate, Tryptophan, Methylthioadenosine
5Luliano, et al., 2013 [67]30 HC,
14 MCI,
30 AD
Plasma GC-MSArachidic acid, Cerotic acid, cis-Vaccenic acid, Erucic acid, Linoleic acid, Mead acid
6González-Domínguez, et al., 2014 [68]37 HC,
14 MCI,
42 AD
Plasma CE-MSCholine, Creatinine, Asparagine, Proline betaine, Methionine, Histidine, Carnitine, N-acetyl-spermidine, Asymmetric dimethyl-Arginine, Tripeptide
7González-Domínguez, et al., 2014 [69]18 HC,
22 AD
PlasmaDIMS/MSCaprylic acid, Capric acid, Lauric acid, Myristic acid, Palmitoleic acid, Palmitic acid, Linoleic acid, Docosahexaenoic acid, Leukotriene B4, Prostaglandin, Choline, Valine, Creatine, Glutamine, Glutamate, Dopamine, Histidine, Carnitine, Arginine, N-acetyl glutamine, Glucose, Glycerophosphocholine, Lyso-phospholipids, Phospholipids
8González-Domínguez, et al., 2014 [70]18 HC,
22 AD
PlasmaESI-Q-TOFMSArginine, Guanidine, Histidine, Imidazole, Kynurenine, Oleamide, P18:0/C22:6-PlsEt, P18:1/C20:4-PlsEt, Prostaglandins, Putrescine, Taurine
9Wang, et al., 2014 [71]57 HC,
58 MCI,
57 AD
Plasma UPLC-QTOF-MSPanel for MCI: Thymine, Arachidonic acid, 2-Aminoadipic acid, N,N-dimethylglycine, 5,8-Tetradecadienoic acid
Panel for AD: Arachidonic acid, N,N-dimethylglycine, Thymine, Glutamine, Glutamic acid, Cytidine
10González-Domínguez, et al., 2015 [72]30 HC,
30 AD
PlasmaFIA-APPI-QTOFMSPalmitoleamide, Palmitamide, Linolenamide, Linoleamide, Oleamide, Stearamide, Palmitoleic acid, Palmitic acid, Oleic acid, Urea, Alanine, Taurine, Picolinic acid, Creatine, Malic acid, Dopamine, Serotonin, Ceramides (Cers), Diacylglycerols (DAG)
11González-Domínguez, et al., 2015 [73]21 HC,
23 AD
PlasmaGC-MSAdenosine, Asparagine, Aspartic acid, Cholesterol, Cystine, Glucose, Glutamine, Histidine, Isocitric acid, Lactic acid, Oleic acid, Ornithine, Palmitic acid, Phenylalanine, Pipecolic acid, Pyroglutamic acid, Stearic acid, Tryptophan, Tyrosine, Urea, Uric acid, Valine, α-Ketoglutarate
12Graham, et al., 2015 [74]37 HC,
16 MCI,
19 MCI-AD
Plasma LC-QTOF-MS4-Aminobutanal, Creatine, γ-aminobutyric acid (GABA), L-ornithine, N1-diacetlyspermine, N-acetylputrescine, Spermine, L-arginine, Methylthioadenosine, N1-acetyl-spermidine, Putrescine, Spermidine
13González-Domínguez, et al., 2016 [75]45 HC,
17 MCI,
75 AD
PlasmaUPLC-QTOF-MSLyso-phospholipids, Phospholipids, S1P, Cers, Sphingomyelins, Glycosphingolipids, Monoglycerides, Acyl-carnitines, Histidine, Phenyl-acetyl-glutamine, Oleamide
14Paglia, et al., 2016 [76]19 HC,
21 AD
Frontal cortexUPLC-MSAcetylaspartic acid, Acetylglutamic acid, ADP, ADP-ribose, Alanine, AMP, Arginine, Asparagine, Aspartic acid, Choline, Cystine, Glutamic acid, Glutamine, GMP, Guanosine, Hydroxyproline, Hypoxanthine, IMP, Inosine, Methionine, Pantothenic acid, Pentose 5-phosphate, Proline, Pyruvate, SAH, SAMe, Serine, Succinic acid, Threonine, Tryptophan, Uric acid, Valine, Xanthine, Xanthosine
15Vaňková, et al., 2016 [77] 22 HC,
16 AD
PlasmaGC-MS17-Hydroxypregnenolone, 17-Hydroxyprogesterone, 20α-Dihydroprogesterone, Allopregnanolone, Allopregnanolone sulfate, Androstenedione, Conjugated 5α-androstane-3b,17b-diol, Isopregnanolone, Pregnanolone, Pregnenolone sulfate, steroid
16Xu, et al., 2016 [78]9 HC,
9 AD
Brain GC-MS55 altered metabolites belonging to glucose utilization/clearance and brain energetics metabolism, and urea and amino-acid metabolism
17Chouraki, et al., 2017 [79] 1974HC,
93AD
PlasmaLC-MSAnthranilic acid, Glutamic acid, Taurine, Hypoxanthine
18de Leeuw, et al., 2017 [80] 121 HC,
127 AD
PlasmaMultiple Mass spectrometry platformsTyrosine, Glycylglycine, Glutamine, Lysophosphatic acid C18:2
19Oberacher, et al., 2017 [81]18 HC,
15 MCI,
21 AD
Plasma FIA-MS/MSPCaa C32:0, PCaa C34:1, PCaa C36:5, PCaa C36:6, PCaa C38:0, PCaa C38:3, PCae C32:1, PCae C32:2, PCae C34:1, PCae C40:4, lysoPC aC16:0, lysoPC aC18:1, lysoPC aC18:2, SM (OH) C14:1
20Snowden, et al., 2017 [82] 14 HC,
14 AD
Brain LC-MS and GC-MSLinoleic acid, Linolenic acid, Eicosapentaenoic acid, Oleic acid, Arachidonic acid
21Hao, et al., 2018 [83]14 HC,
16 AD
CSF LC-MS/MS142 metabolites belong to carboxylic acids, amino acids, fatty acyls, fatty acids and conjugates, pyrimidines, nucleosides, and analogs. E.g.: 3-Oxododecanoic acid, Dodecanedioic acid, Methylerythritol phosphate, Glutamine, Dihydrothioctic acid, Deoxyinosine, Succinyl-glutamate
22Kim, et al., 2018 [84]13 HC,
23 MCI,
14 AD
Plasma UHPLC-ESI-MS/MSPE (34:2), PE (36:2), PE (38:4), PE (38:5), PA (18:1/22:6), PI (18:0/20:4), PI (18:1/16:1), PI (18:2/16:1), Cer (d18:1/22:0), Cer (d18:1/24:1), HexCer (d18:1/24:0), DG (18:0/22:6), DG (18:1/18:1), TG (50:1)
23Muguruma, et al.,
2018 [85]
10 HC,
10 AD
Postmortem CSF (pCSF)UHPLC-MS/MSPolyamine and tryptophan-kynurenine (Trp-Kyn) metabolisms, such as methionine sulfoxide, 3-methoxy-anthranilate, cadaverine, guanine, and histamine
24Nasaruddin, et al.,
2018 [86]
27 HC,
16 AD
Brain GC-MSArachidic acid, Arachidonic acid, Behenic acid, Cis-10-heptanoic acid, Cis-11,14,17-eicosatrienoic acid, Cis-13,16-docosadienoic acid, Erucic acid, Lignoceric acid, Linolenic acid, Nervonic acid, Oleic acid, Palmitic acid, Stearic acid
25van der Lee, et al.,
2018 [87]
23,882 HC,
1356 AD
Plasma NMR spectroscopyHigh-density lipoprotein subfractions, Docosahexaenoic acid, Ornithine, Glutamine, Glycoprotein acetyls
26Kim, et al., 2019 [88]9 HC,
9 AD
Postmortem, brain UPLC-MSHypotaurine, Myo-inositol, Oxo-proline, Glutamate, N-acetyl-aspartate, Cortisol, N-acetylaspartate (NAA), N-acetylaspartylglutamate (NAAG), Acetylcholine, Alanine
27Lin, et al., 2019 [89]15 HC,
10 MCI,
15 AD
Plasma LC-MS/MSPropionylcarnitine, Valerylcarnitine, Glutarylcarnitine/Hydroxyhexanoylcarnitine, Arginine, Phenylalanine, Creatinine, Symmetric dimethylarginine (SDMA)
28MahmoudianDehkordi, et al., 2019 [90]37 HC,
284 early MCI, 505 late MCI, 305 AD
PlasmaUPLC-MS/MSCholic acid, Deoxycholic acid (DCA), Glycodeoxycholic acid (GDCA), Glycolithocholic acid (GLCA), Taurolithocholic acid (TLCA)
29Marksteiner, et al.,
2019 [91]
25 HC,
26 MCI,
27 AD
SalivaFIA-MS/MSPCae C34:1-2, PCae C36:1-2-3, PCaeC38:1-3, PCae C40:2-3, PCae C36: (1-2-3)
30Peña-Bautista, et al., 2019 [92]29 HC,
29 AD
Plasma UPLC-Q-TOF-MSCholine, L-carnitine, 4-Deoxyphysalolactone, Rescinnamine, Chlorohydrin, Brassinin, Nicotinamide ribotide, Cyasterone
31Snowden, et al., 2019, [93] 14 HC,
14 AD
Brain LC-MSAminobutanal, Arginine, Aspartate, Dihydroxy-phenylalanine, Dopamine, Gamma-aminobutanoate, Glutamate, Glycine, Guanidinobutanoate, Guanosine, Ornithine, Serotonin, Tryptophan, Tyrosine
32van der Velpen, et al., 2019 [94]34 HC,
40 AD
Plasma,
CSF
LC-MS/MSAcetylcarnitine, Acylcarnitines C14, Acylcarnitines C16, Acylcarnitines C18, cis-Aconitate, Citrate, Creatinine, Decanoylcarnitine, Hexanoylcarnitine, Kynurenic acid, Lauroylcarnitine, L-carnitine, Octanoylcarnitine, Quinolinic acid, Tryptophan, α-Ketoglutarate
33Ahmad, et al., 2020
[95]
Two independent cohorts: 142 MCI and 40 MCICSF and plasmaUHPLC-MS/MSLPA (C16:0), LPA (C16:1), LPA (C22:4), LPA (C22:6), and isomer-LPA (C22:5)
34Shao, et al., 2020 [96]94 HC,
44 AD
PlasmaUPLC-MSPolyunsaturated fatty acids (PUFAs), Docosahexaenoic acid (DHA, C22:6), medium- and long-chain Acyl-Carnitines, Cholic acid (CA), Chenodeoxycholic acid (CDCA), Allocholic acid, Tryptophan, Serotonin, Indolelactic acid
35Byeon, et al., 2021 [97]18 HC,
15 MCI,
17 AD
CSFLC-MSLPC, PC, LdMePE, LPE, dMePE, PE class
36Liu, et al., 2021 [98]19 HC,
25 AD
Brain tissue LC-MS/MS3-Methylguanine, Acetylcholine, Cytosine, Glycylproline, Guanosine, Histidinyl-aspartate, Imidazoleacetic acid, Indole-3-propionic acid, Inosine, Ketoleucine, Linoleamide, L-methionine, L-norleucine, L-valine, N6-methyladenosine, N-acetylglutamic acid, N-acetyl-L-aspartic acid, N-acetyl-L-phenylalanine, Palmitoleic acid, Phenylalanine, Phenylpyruvic acid, Piperidine, Sarcosine, Serylglycine, S-formylglutathione, Sphingosine, Theaflavin
37Liu, et al., 2021 [99]42 HC,
40 AD
Plasma LC-MS/MSCer, ChE, DG, LPC, PC, PE, PI, SM, and TG class
38Nielsen, et al., 2020 [100]10 HC,
10 MCI,
10 AD
PlasmaLC-MS and NMR, spectroscopyValine, Histidine, Allopurinol riboside, Inosine, 4-Pyridoxic acid, Guanosine
39Horgusluoglu, et al., 2022 [101] 362 HC,
270 early MCI,
494 late MCI,
298 AD
PlasmaHPLC-MS/MSShort-chain Acylcarnitines/amino acids and medium/long-chain Acylcarnitines
40Khan, et al., 2022 [102]54 HC,
59 AD
PlasmaLC-MS/MSPS (18:0/18:0), PS (18:0/20:0), PC (16:0/22:6), PC (18:0/22:6), PS (18:1/22:6)
41Maffioli, et al., 2022 [103]20 HC,
23 AD
Hippocampal tissueLC-MS and GC-MSGlycerol 3 phosphate, Erythrose 4 phosphate, Glucose, Deoxyuridine, Lactic acid, Saccharopine, Uridine 5′ diphosphate, Oxidized glutathione, Urea, Uracil, Succinic acid, Arginine, Beta D Glucose 6 phosphate, Glucosamine 6 phosphate, AICAR, Citrulline, Alanine, Pyruvic acid, Glutathione, Guanidoacetic acid, Mevalonate P, Lysine
42Ozaki, et al., 2022 [104]40 HC,
26 MCI,
40 AD
PlasmaCE-TOF-MSOrnithine, Uracil, Lysine
43Peña-Bautista, et al., 2022 [105] 20 HC,
11 AD,
31 MCI-AD
PlasmaUPLC-TOF/MSCer, LPE, LPC, MG, and SM were observed as being altered significantly between the preclinical AD and healthy groups. DG, MG, and PE were observed as being altered significantly between the MCI-AD and healthy groups.
44Weng, et al., 2022 [106] 19 HC,
16 AD
PlasmaNMR spectroscopy3-Phosphoglycerate, Fructose-6-phosphate, Glucose-6-phosphate, Betaine, Methyl-histidine, Glycerylphosphorylcholine, 2-Oxoglutarate, Citrate, Malate, Ergothioneine, Glutathione disulfide, Taurine, Carnosine, Ornithine, Glycine, Alanine, Serine, Glutamine, Tryptophan, Valine
Table 2. Summary of the included articles related to metabolomics studies performed on animal models.
Table 2. Summary of the included articles related to metabolomics studies performed on animal models.
ReferenceStudy PopulationSample
Type
Analytical PlatformAltered Metabolites
1Salek, et al., 2010 [123] Two age groups (2–3 months and 12–13 months) of transgenic CRND8 APP 695 and non-transgenic littermates (controls)Brain samples with seven different brain regionsNMR-based metabolomicsLactate, Aspartate, Glycine, Alanine, Leucine, Iso-leucine, Valine, N-acetyl-L-aspartate, Glutamate, Glutamine, Taurine, Gamma-amino butyric acid, Choline, Phosphocholine, Creatine, Phosphocreatine, Succinate
2Hu, et al., 2012 [124]50-week-old TASTPM transgenic AD mice and 5-month old wild type C57BL/6J micePlasma and brain tissuesGC-MSAA, Androstenedione, Cortisol, D-fructose, D-galactose, D-glucose, Gluconic acid, Linoleic acid, L-serine, L-threonine, L-valine, Palmitic acid
3Trushina, et al., 2012 [125] APP/PS1 mice at 16 months of age and age-matched nTg controlsHippocampusGC-MSThreonic acid, Ethanolamine, Alanine, Mannitol, Glycerol 3-P, Pyroglutamic acid, N-acetyl-aspartate, Lactic Acid
4Tajima, et al., 2013 [126]APP/tau mice at 4, 10, and 15 months of age and age-matched wild-type miceBrain tissue, PlasmaRPLC-ESI-TOFMS12-HETE, 19,20-diHDoPE, 17,18-diHETE, 19,20-EpDPE, 17,18-EpETE, Prostaglandin D2, 15-HETE, Phosphatidylcholines, PEs with Polyunsaturated fatty acids, Docosahexaenoyl (22:6) Cholesterol Ester (ChE), Ethanolamine Plasmalogens (pPEs), Sphingomyelins (SMs)
5González-Domínguez, et al., 2014 [127]6-month-old APP/PS1 mice and age-matched controlsBrain tissue with different regionsGC-MS and UPLC-MS40 significant differences in metabolites related to abnormal purine metabolism, bioenergetic failures, dyshomeostasis of amino acids, and disturbances in membrane lipids
6Kim, et al., 2014 [128] Model: mice receive an intracerebroventricular infusion of Aβ, Control: age-matched untreated micePlasmaNMR-MSNiacinamide, AMP, Hypoxanthine, Citrate, Lactate, Pyruvate, Creatine, Choline, Acetate, Phenylalanine, Glycine, Valine, Tyrosine, Alanine, Glucose, Glutamine
7Lalande, et al., 2014 [129]Single-transgenic Tg2576 mice at 1, 3, 6, and 11 months of age, and age-matched controlsBrain tissues with five different brain regionsNMR-MSγ-Aminobutyric acid, Glutamate, N-acetylaspartate (NAA), Myo-Inositol, Creatine, Phosphocholine, Glu, Creatine, Taurine
8González-Domínguez, et al., 2015 [130]APP/PS1 mice at 6 months old, and age-matched nTg controlsPlasmaFI-APPI-MS, DI-ESI-MSCholine, Serine, Valine, Threonine, Pyroglutamate, Creatine, Phosphoethanolamine, Histidine, Carnitine, Glucose, Tyrosine, Tryptophan, GPE, Inosine, Urea, Myristic Acid, Palmitoleic acid, HEPE, Lysophosphocholines, Phosphocholines, Diacylglycerols, Cholesteryl esters, Triacylglycerols
9González-Domínguez, et al., 2015 [131]APP/PS1 mice at 6 months of age and age-matched nTg controlsPlasmaGC-MS, UHPLC-MSPhosphoethanolamine, Adenosine monophosphate, Citrulline, Citric acid, Monostearin, Bile acids, Eicosanoids, Fatty acid amides, Sphingoid bases, Lyso-phospholipids, Phospholipids, Sphingomyelins, Lactic acid, B-hydroxybutyric acid, Urea, Phosphoric acid, Glycine, Succinic acid, Threonine, Malic acid, Pyroglutamic Acid, Creatinine, Proline, Glycerol-3-phosphate, Citric acid, Glucose, Tyrosine, 6Myoinositol, Uric Acid, Glucose-6-phosphate, Tryptophan, Stearic acid, Myoinositol-1-phosphate, Serotonin, 1,3-Bisphosphoglycerate, Cholesterol
10González-Domínguez, et al., 2015 [132] 6-month-old APP/PS1 mice and age-matched controlsBrain tissue with different regionsLC-QTOF-MS52 significant differences in metabolites including phospholipids, fatty acids, purine and pyrimidine metabolites, acylcarnitines, sterols, and amino acids which related to the homeostasis of lipids, energy management, and metabolism of amino acids and nucleotides
11Li, et al., 2015 [133] Model: mice receive hippocampal region infusion of Aβ1-42, Control: age-matched untreated micePlasma, Brain tissuesUPLC-MS/MSPhenylalanine, Tryptophan, Dihydrosphingosine, LPC (18:2), LPC (20:4), LPC (22:6), LPC (16:0), LPC (16:0), LPC (18:1), LPC (18:0)
12Li, et al., 2016 [134] APP/PS1 mice at 1 month old and age-matched nTg controlsBrain tissuesUPLC-MSHypoxanthine, Hexadecasphinganine, Dihydrosphingosine, Phytosphingosine, LPC (13:0), LPC (15:1), LPC (15:0), LPC (16:0), LPC (18:3), LPC (18:1), LPC (18:0)
13Pan, et al., 2016 [135] APP/PS1 mice at 1, 8, 10, 12, and 18 months of age, and age-matched nTg controlsPlasma and brain tissuesUPLC-TQ-MSPhenylalanine, Trypotophan, Tyrosine, a-Aminoadipic acid, Asparagine, Histidine, Phenylalanine, Valine, Isoleucine, Methionine, Tyrosine, Methionine sulfoxide, Serotonin, Taurine
14Nuriel, et al., 2017 [136]APOE mice at 14.5 months of age and age-matched controlsBrain tissuesUPLC-MSMyristic acid, DHA, Stearic acid, 12-Hydroxydodecanoic acid, Arachidic acid, Palmitic acid, Trisaccharide, Tetrasaccharide, Disaccharide, Phylloquinone, Tocopherol, Dehydroascorbic acid, Inosine 5′-monophosphate (IMP), D-fructose 6-phosphate, Succinoadenosine, Carnitine, Citric acid/Isocitric acid, Malic acid, ATP, Lanosterol, Cholesteryl 19acetate, Leucine, Proline, Glycine, Quinaldic acid, Kynurenine, Kynurenic acid, S-Adenosylhomocysteine, Carnosine, 4-Oxoproline, Tyramine, Thymidine, Uracil, GMP, Methylglutarylcarnitine, Trimethylamine N-oxide, 2-Hydroxypyridine, N-acetylneuraminic acid, Hydroxybutyric acid
15Pan, et al., 2017 [137] APP/PS1 mice at 6 and 12 months of age and age-matched controlsPlasma and brain tissuesLC-MS/MSCholic acid, Hyodeoxycholic acid, Lithocholic acid, Taurocholic acid, Tauromuricholic acid (α and β), Tauroursodeoxycholic acid, β-Muricholic acid, Ω-Muricholic acid
16Bergin, et al., 2018 [138]APP/PS1 mice at 7 and 13 months of age and age-matched nTg controlsBrain tissues, PlasmaHPLC-MSArginine, Citrulline, Ornithine, Agmatine, Putrescine, Spermine, Spermidine
17Gao, et al., 2018 [139]APP/PS1 mice at 12 months of age and age-matched nTg controlsPlasmaUPLC-MSAcetoacetyl-CoA, 21-Deoxycortisol, 9 (10)-EpOME, 3α-Hydroxy-5β-androstan-17-one, PGF2α, Dehydroepiandrosterone, 3α,21-Dihydroxy-5β-pregnane-11,20-dione, Sphinganine, Sphingosine, DHA, Linoleic acid, 4,4-Dimethyl-5α-cholest-7-en-3β-ol, Arachidonic acid, PC (20:4 (5Z,8Z,11Z,14Z)/0:0), 9-Cis-retinoic acid, LTA4
18Sun, et al., 2018 [140] 21-week-old APP/PS1 mice and age-matched controlsPlasmaUPLC-QTOF/MSArachidonic acid, Cer (d18:0/12:0), Glycerophosphocholine, Leukotriene B4, Linoleic acid, L-tryptophan, Phosphorylcholine, Phytosphingosine, Prostaglandin F2a, Prostaglandin G2, Sphinganine
19Zheng, et al., 2018 [141] APP/PS1 mice at 5 and 10 months of age and age-matched controlsBrain tissuesNMR-based metabolomicsAdenosine monophosphate, ADP, Alanine, Aspartate, Choline, Creatine/phosphocreatine, Glutamate, Glutamine, Glycerolphorylcholine, Glycine, Inosine, Inosine monophosphate, Lactate, Myo-inositol, N-acetylaspartate, Nicotinamide adenine dinucleotide, Phosphocholine, Succinate, γ-Aminobutyric acid
20Zhou, et al., 2018 [142] 10-month-old APP/PS1 double transgenic mice and age-matched controlsBrain tissues13C NMR based metabolomicGlutamate, Glutamine, γ-Aminobutyric acid, Aspartate, Succinate, Lactate, Alanine
21Chang, et al., 2019 [143]5-month-old APP/PS1 mice and age-matched controlsBrain tissue with different regionsGC-TOF-MS2-Hydroxyglutaric acid, 4-Guanidinobutyric acid, Aspartic acid, Beta-alanine, Citric acid, Creatinine, Ethanolamine, Fumaric acid, GABA, Glucose, Glucose-6-phosphate, Glutamic acid, Glycerol, Glycine, Inosine, Lactic acid, Leucine, Malic acid, N-acetyl-L-aspartic acid, Oleic acid, Proline, Pyroglutamic acid, Ribose-5-phosphate, Serine, Succinic acid, Threonine, Threonolactone, Tyrosine, Uracil, Urea, Valine
22Li, et al., 2019 [144] 3xTg-AD mice at 19 months of age and age-matched nTg controlsPlasmaUPLC-Q/TOF-MSLinoleic acid, 9 (10)-EpOME, DHA, Arachidonic acid, 11b-PGF2a, Sphingosine, Pyruvate, L-Leucine, Ursocholic acid, Corticosterone, L-Tryptophan, Dodecanoic acid, L-lysine
23Liu, et al., 2019 [145] APP/PS1 mice at 6 and 9 months of age and age-matched controlsPancreas and serumNMR-based metabolomics2-Aminobutyrate, Acetate, Adenosine-monophosphate, AMP, Adenosine-triphosphate, ATP, Alanine, Allantoin, Aspartate, Betaine, Choline, Citrate, Fumarate, Glucose, Glucose 1-phosphate, Glutamine, Glycine, Lactate, Leucine, Malate, Nicotinurate, Pyruvate, Reduced nicotinamide adenine dinucleotide phosphate, NADPH, Serine, Succinate, Taurine, Valine
24Liu, et al., 2019 [146] 3-month-old C57BL/6 mouse with local administration of Aβ peptide into brain, and 3-month-old controlsHippocampusUPLC-QTOF/MSAdenine, Adenosine, Citrulline, Inosine, L-Arginine, L-Isoleucine, L-Threonine, Normetanephrine
25Pan, et al., 2019 [147] 8-month-old PLB4 mice and age-matched controlsBrain tissuesLC-MSLeucine, Creatinine, Putrescine and species of Acylcarnitines, LysoPC, PCs, Sphingomyelin
26Rong, et al., 2019 [148]Model: SD rats (200 ± 20 g) induced by D-Gal and Aβ25–35 injection, Control: age-matched untreated ratsPlasma, Brain tissueUPLC-Q/TOF-MSLysoPC (17:0), LysoPC (20:2), LysoPC (20:4), LysoPC (16:0), LysoPC (18:0), LysoPE (20:4), PE (18:1), PE (22:4)
27Hunsberger, et al., 2020 [149] APP/PS1 mice at 6, 12, and 24 months of age and age-matched controlsPrefrontal cortex, Hippocampus, SpleenLC-MS and GC-MS1-Methyl nicotinamide, 2-Methylbutyroylcarnitine, 7-Methylguanine, Arginine, Argininosuccinic acid, Azelaic acid, Dimethylglycine, Formiminoglutamic acid, Glutathione, Glycerate, Glycolic acid, Guanidinobutanoic acid, Hydroxyproline, Hydroxypyruvic acid, Isovalerylcarnitine, L-glutamic acid, L-methionine, L-tyrosine, Methylhistamine, Methylhistidine, N6-acetyl-L-lysine, N-acetyl L-aspartic acid, N-acetylneuraminic acid, Orotate, Phenylpyruvic acid, Phosphocreatine, Propionylcarnitine, Pyroglutamic acid, Pyruvate, S-adenosylmethionine, SAICAR, Sphingosine-1-phosphate, Uracil, Uridine
28Kim, et al., 2020 [150]5xFAD mice at 8 and 12 months of age and age-matched controlsHippocampusUPLC-MSNicotinamide, Adenosine monophosphate, LysoPC (16:0), LysoPC (18:0), and LysoPE (16:0)
29Sun, et al., 2020 [151]Model: SD rats (weight 260 ± 20 g) receive an infusion of Aβ1-42. Control: age-matched untreated rats.PlasmaHPLC-MSArachidic acid, FA (18:0), FA (18:1), FA (18:2), FA (18:3), CE (20:5), CE (22:6), Cer (d18:0/20:0), Cer (d18:1/24:0), MG (16:0/0:0/0:0), DG (18:0/16:0/0:0), TG (16:0/18:0/18:1), LPC (18:0), LPC (20:4), LPC (20:5), LPC (20:1), LPC (O-18:0), LPC (P-16:0), LPC (P-18:0), PC (18:2/18:0), PC (18:2/18:2), PC (18:2/20:4), PC (35:1), PC (35:4), PC (36:1), PC (O-32:0), SM (d18:0/16:1), SM (d40:2)
30Tondo, et al., 2020 [152] 4-month-old 16 rTg4510 mice and age-matched controlsBrain LC-MSGlutamine, Serotonin, Sphingomyelin C18:0
31Yi, et al., 2020 [153]SD rats (180 ± 40 g) were induced by Aβ1-42 injection, Control: age-matched untreated ratsPrefrontal cortexUPLC-TripleTOF/MSCer (d18:1/16:0), Cer (d18:1/24:1 (15Z)), GlcCer (d14:1/20:0), Galbeta-cer (d18:1/20:0), Cer (d18:1/18:0), PI (18:0/0:0), LysoPE (16:1 (9Z)/0:0), Glucosylceramide (d18:1/18:0), PE (18:1 (9Z)/16:0), LysoPC (16:1 (9Z)/0:0), PI (20:4 (5Z,8Z,11Z,14Z)/0:0), PE (16:1 (9Z)/P-16:0), 4-Nitrophenol, LysoPE (0:0/22:6 (4Z,7Z,10Z,13Z,16Z,19Z)), PE (15:0/14:1 (9Z)), Cer (d18:1/20:0), PC (18:1 (11Z)/18:2 (9Z,12Z)), PS (20:1 (11Z)/18:0), PC (18:1 (11Z)/20:4 (5Z,8Z,11Z,14Z)), PE (18:3 (9Z,12Z,15Z)/22:1 (13Z)), Cer (d18:0/22:0), PE (15:0/22:0)
32Zhang, et al., 2020 [154]APP/tau mice at 7 months of age and age-matched wild-type miceBrain tissueUHPLC-QTOF/MSHydroxy-E4-neuroprostane, Prostaglandin D3, PS (18:0/22:5), PC (14:0/16:0), PC (16:0/19:0), PE (22:5/22:6), PE (15:0/22:4), PE (18:1/20:2), Leukotriene D4, Cer (d18:1/18:1)
33Zhang, et al., 2020 [155]APP/tau mice at 2, 3, and 7 months of age and age-matched wild-type miceBrain tissue, PlasmaUHPLC-QTOF/MSLysophospholipids, PCs, Pes, Cer, Fatty acids, Diacylglycerols (DGs), Triacylglycerols (TGs)
34Dejakaisaya, et al., 2021 [156] Six-month-old mixed-sex Tg2576 mice and c57 × SJL (WT) littermatesCortexLC-MS2′-Deoxyguanosine 5′-monophosphate, Lysophosphatidylethanolamines (18:0), Diglyceride (P-32:1), 3–4-Hydroxyphenyllactate, Npi-methyl-l-histidine, O-butanoylcarnitine, O-propanoylcarnitine, Phosphatidylethanolamine (44:3), Phosphatidylserine (36:2), Phosphatidylserine (40:7), Phosphatidylserine (44:12)
35Speers, et al., 2021 [157] 8-month-old 5xFAD mice and age-matched controlsCortical tissueHPLC-MSBetaine, Lysine, Pyridoxamine, Urate, NAD, Erythritol, Spermine, N-Acetylmannosamine, Glycerol 2-phosphate, Lauroyl-L-carnitine, Nicotinamide hypoxanthine Dinucleotide, Sodium taurocholate, Homocysteine, Betaine, N-acetyl-mannosamine, S-adenosyl-homocysteine, Adenosine 3′5′-cyclic monophosphate
36Sun, et al., 2021 [158]Model: SD rats with Aβ1-42 protein injection. Control: age-matched untreated rats.Brain tissueHPLC/FT-ICR MSDG (20:4/16:0), DG (22:6/16:0), TG (10:0/18:0/16:0), TG (16:0/16:0/18:1), TG (43:1), TG (16:0/14:0/16:1), Glucosylceramide (d18:1/24:1), PA (22:6/22:5), PG (18:0/22:4), PG (40:5), PG (43:6), PI (18:0/20:4), PI (36:4), PI (18:1/20:4), PS (18:0/20:4), LPE (22:6), PC (17:1/18:1), PC (20:0), PC (34:5), Adrenic acid
37Zhao, et al., 2021 [159] 3×Tg-AD mice at 2 and 6 months of age and age-matched nTg controlsHippocampusUPLC-MRM-MS84 significant differences in metabolites related to abnormal purine, pyrimidine, arginine, and proline metabolism, glycerophospholipid metabolism
38Cheng, et al., 2022 [160] 9-month-old male APP/PS1 mice and age-matched controlsPlasmaUHPLC-MS/MS3-Indole carboxylic acid glucuronide, Cyclic dAMP, (5Z,9Z)-2-Methoxy-hexacosadienoic acid, Caffeoyl aspartic acid, Saccharin, (Z)-13-Oxo-9-octadecenoic acid, 5C-Aglycone, PS (O-16:0/20:2 (11Z,14Z)), Bisacurone Epoxide, PS (22:0/15:0), 6-Hydroxymelatonin glucuronide, Methyl 2-undecynoate, Cis-3-hexenyl trans-4-hexenoate
39Dai, et al., 2022 [161] 4-month old APPswe/PS1∆E9 (PAP) transgenic mice and age-matched controlsBrain and plasmaLC-MS23-Nordeoxycholic acid, 23-Norcholic acid, 7-Ketodeoxycholic acid, Deoxycholic acid, β-muricholic acid, Cholic acid, Cholecalciferol, Pyridoxal, Flavin adenine dinucleotide, Riboflavin, Arachidic acid, Eicosapentaenoic acid, Neuronic acid, Erucic acid, Acetoacetate, Fumaric acid, Estradiol, Progesterone, Testosterone
40Dehghan, et al., 2022 [162] 50-week-old Abca7 knockout mice and age-matched controlsBrain UPLC-MSLacCer, Sphingomyelins, Cers, Hexosylceramides
41Dunham, et al., 2022 [163] 4, 8, 12, and 18 months 5xFAD and wild-type littermatesPlasma UPLC-MRMsGlycine, Carnosine, Serine, SM C24:1, Serotonin, Serotonin, Spermidine, Spermine, Aspartic acid phosphatidylcholines (PCs), α-Aminoadipic acid, Serotonin, Glutamic acid
42Sun, et al., 2022 [164] 4-month-old APP/PS1 mice and age-matched controlsSerum, Brain Tissues, FecesLC-MS88 significant differences in metabolites related to neurotransmitters metabolism, lipid metabolism, aromatic amino acids metabolism, energy metabolism, vitamin digestion and absorption, and bile metabolism
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

Yin, C.; Harms, A.C.; Hankemeier, T.; Kindt, A.; de Lange, E.C.M. Status of Metabolomic Measurement for Insights in Alzheimer’s Disease Progression—What Is Missing? Int. J. Mol. Sci. 2023, 24, 4960. https://doi.org/10.3390/ijms24054960

AMA Style

Yin C, Harms AC, Hankemeier T, Kindt A, de Lange ECM. Status of Metabolomic Measurement for Insights in Alzheimer’s Disease Progression—What Is Missing? International Journal of Molecular Sciences. 2023; 24(5):4960. https://doi.org/10.3390/ijms24054960

Chicago/Turabian Style

Yin, Chunyuan, Amy C. Harms, Thomas Hankemeier, Alida Kindt, and Elizabeth C. M. de Lange. 2023. "Status of Metabolomic Measurement for Insights in Alzheimer’s Disease Progression—What Is Missing?" International Journal of Molecular Sciences 24, no. 5: 4960. https://doi.org/10.3390/ijms24054960

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