ArticlePDF Available

Systemic Alterations in the Metabolome of Diabetic NOD Mice Delineate Increased Oxidative Stress Accompanied by Reduced Inflammation and Hypertriglyceridemia

Authors:

Abstract and Figures

Non-obese diabetic (NOD) mice are a commonly-used model of type 1 diabetes (T1D). However, not all animals will develop overt diabetes despite undergoing similar autoimmune insult. In this study, a comprehensive metabolomic approach, consisting of gas chromatography time-of-flight (GC-TOF) mass spectrometry (MS), ultra high performance liquid chromatography accurate mass quadruple time-of-flight (UHPLC-qTOF) MS and targeted UHPLC-tandem mass spectrometry -based methodologies, was used to capture metabolic alterations in the metabolome and lipidome of plasma from NOD mice progressing or not progressing to T1D. Using this multi-platform approach, we identified >1000 circulating lipids and metabolites in male and female progressor and non-progressor animals (n=71). Statistical and multivariate analyses were used to identify age- and sex-independent metabolic markers, which best differentiated metabolic profiles of progressors and non-progressors. Key T1D-associated perturbations were related with 1) increases in oxidation products glucono delta lactone and galactonic acid, and reductions in cysteine methionine and threonic acid, suggesting increased oxidative stress; 2) reductions in circulating polyunsaturated fatty acids and lipid signaling mediators, most notably arachidonic acid (AA) and AA-derived eicosanoids, implying impaired states of systemic inflammation; 3) elevations in circulating triacylglyercides reflective of hypertriglyceremia; and 4) reductions in major structural lipids, most notably lysophosphatidylcholines and phosphatidylcholines. Taken together, our results highlight the systemic perturbations that accompany a loss of glycemic control and development of overt T1D. Copyright © 2015, American Journal of Physiology - Endocrinology and Metabolism.
Content may be subject to copyright.
Systemic alterations in the metabolome of diabetic NOD mice delineate
increased oxidative stress accompanied by reduced inflammation
and hypertriglyceremia
Johannes Fahrmann,
1
* Dmitry Grapov,
1
* Jun Yang,
2
Bruce Hammock,
2
Oliver Fiehn,
1
Graeme I. Bell,
3
and Manami Hara
3
1
National Institutes of Health West Coast Metabolomics Center, University of California Davis, Davis, California;
2
Department of Entomology and Cancer Center, University of California Davis, Davis, California; and
3
Department of
Medicine, The University of Chicago, Chicago, Illinois
Submitted 16 January 2015; accepted in final form 1 April 2015
Fahrmann J, Grapov D, Yang J, Hammock B, Fiehn O, Bell GI,
Hara M. Systemic alterations in the metabolome of diabetic NOD mice
delineate increased oxidative stress accompanied by reduced inflammation
and hypertriglyceremia. Am J Physiol Endocrinol Metab 308: E978 –E989,
2015. First published April 8, 2015; doi:10.1152/ajpendo.00019.2015.—
Nonobese diabetic (NOD) mice are a commonly used model of type
1 diabetes (T1D). However, not all animals will develop overt diabe-
tes despite undergoing similar autoimmune insult. In this study, a
comprehensive metabolomic approach, consisting of gas chromatog-
raphy time-of-flight (GC-TOF) mass spectrometry (MS), ultra-high-
performance liquid chromatography-accurate mass quadruple time-of-
flight (UHPLC-qTOF) MS and targeted UHPLC-tandem mass spec-
trometry-based methodologies, was used to capture metabolic
alterations in the metabolome and lipidome of plasma from NOD
mice progressing or not progressing to T1D. Using this multi-platform
approach, we identified 1,000 circulating lipids and metabolites in
male and female progressor and nonprogressor animals (n71).
Statistical and multivariate analyses were used to identify age- and
sex-independent metabolic markers, which best differentiated meta-
bolic profiles of progressors and nonprogressors. Key T1D-associated
perturbations were related with 1) increases in oxidation products
glucono--lactone and galactonic acid and reductions in cysteine,
methionine and threonic acid, suggesting increased oxidative stress;
2) reductions in circulating polyunsaturated fatty acids and lipid
signaling mediators, most notably arachidonic acid (AA) and AA-
derived eicosanoids, implying impaired states of systemic inflamma-
tion; 3) elevations in circulating triacylglyercides reflective of hyper-
triglyceridemia; and 4) reductions in major structural lipids, most
notably lysophosphatidylcholines and phosphatidylcholines. Taken
together, our results highlight the systemic perturbations that accom-
pany a loss of glycemic control and development of overt T1D.
diabetic mice; metabolomics; inflammation; oxidative stress
TYPE 1 DIABETES (T1D) is an autoimmune disease characterized
by the selective destruction of pancreatic -cells. Currently, it
is considered that the autoimmune insult is the primary driver
of -cell destruction that leads to development of T1D. How-
ever, it is becoming increasingly evident that early metabolic
perturbations are inherently involved in the development and
progression of T1D (39, 40, 53). These metabolic alterations
may potentiate or attenuate -cell loss and dysfunction. Oresic
et al. (40) reported that alterations in branched-chain amino
acids (BCAA) and lipid metabolism preceded the appearance
of autoantibodies in children who later progressed to T1D.
Pflueger et al. (44) found that, independent of age-related
differences, autoantibody-positive children had higher levels of
odd-chain triglycerides and polyunsaturated fatty acid (PUFA)-
containing phospholipids than autoantibody-negative children.
Furthermore, it was found that children who developed auto-
antibodies by age 2 yr had a significantly lower concentration
of circulating methionine than those who developed autoanti-
bodies late in childhood or remained autoantibody negative
(44). Collectively, these studies demonstrate an intrinsic rela-
tionship between metabolic perturbations, autoimmune insult,
and -cell destruction.
Metabolomics is the study of small molecules and biochem-
ical intermediates (metabolites), which are highly relevant to
other regulatory mechanisms (e.g., genomics, transcriptome,
and proteome) and sensitive to environmental stimuli, forming
detailed representations of organismal phenotypes. Over the
past decade, the application of metabolomics has been used to
gain new insights into the pathology of numerous diseases
including type 2 diabetes, develop methods predictive of dis-
ease onset, and reveal new biomarkers associated with diag-
nosis and prognosis (20, 51). Therefore, the application of
metabolomics to study T1D pathophysiology represents a
promising avenue of research to identify candidate biomarkers
related to disease development and progression.
Using a nonobese diabetic (NOD) mouse model, we recently
demonstrated marked heterogeneity in pancreatic -cell loss
regardless of age or sex. Importantly, we found that chronic
hyperglycemia (250 mg/dl) and overt T1D manifested only
in mice that lost 70% of their total -cell mass (23). Using an
untargeted metabolomics approach, we performed an initial
evaluation of primary metabolites in plasma of young (25 wk)
and old (25 wk) nonprogressors (nondiabetic) vs. progressors
(diabetic) to reveal distinct metabolomic pathways involved in
disease progression (23). In the present study, we largely extended
these investigations to capture alterations in primary metabolism,
complex lipids, and lipid signaling mediators in the plasma of
nonprogressors and progressors with the overall aim(s) of identi-
fying circulating factors linked to disease progression and, poten-
tially, -cell destruction and dysfunction.
METHODS
Metabolomic, Lipidomic, and Oxylipin Analysis
NOD mice (n71) were assessed as diabetic or nondiabetic based
on their fasting (4 h) blood glucose levels at death, which defined 31
* J. Fahrmann and D. Grapov contributed equally to this work.
Address for reprint requests and other correspondence: M. Hara, Dept. of
Medicine, Univ. of Chicago, 5841 South Maryland Ave., MC1027, Chicago,
IL 60637 (e-mail: mhara@midway.uchicago.edu).
Am J Physiol Endocrinol Metab 308: E978–E989, 2015.
First published April 8, 2015; doi:10.1152/ajpendo.00019.2015.
http://www.ajpendo.orgE978
Downloaded from journals.physiology.org/journal/ajpendo (193.160.077.067) on January 9, 2021.
hyperglycemic (glucose 250 mg/dl) and 40 normoglycemic ani-
mals. All procedures involving mice were approved by the University
of Chicago Institutional Animal Care and Use Committee. Character-
istics of our defined cohort are described in Table 1.
The MiniX database (47) was used as a laboratory information
management system (LIMS) and for sample randomization prior to
all analytical procedures. Detailed information on sample prepa-
ration and data acquisition for all analytical platforms are reported
in the APPENDIX.
For analysis of primary metabolism, plasma aliquots (30 l), stored
at 80°C, were thawed, extracted, and derivatized, and metabolite
levels were quantified by gas chromatography time-of-flight (GC-
TOF) mass spectrometry as previously described (19). All samples
were analyzed in one batch, throughout which data quality and
instrument performance were monitored using quality control and
reference plasma samples [National Institute of Standards and Tech-
nology (NIST)]. Quality controls (n8), comprising a mixture of
standards and analyzed every 10 samples, were monitored for changes
in the ratio of analyte peak heights and used to ensure equivalent
instrumental conditions (P0.05, t-test comparing observed to
expected ratios of analyte response factors) over the duration of the
sample acquisition (19). Acquired spectra were further processed
using the BinBase database (18, 47). Briefly, output results (30) were
filtered based on multiple parameters to exclude noisy or inconsistent
peaks. Detailed criteria for peak reporting, including mass spectral
matching, spectral purity, signal to noise, and retention time, are
discussed in detail elsewhere (17). Known artifact peaks such as
polysiloxanes or phthalates were excluded from data export in Bin-
Base. Missing values were replaced by investigating the extracted ion
traces of the raw data subtracted by the local background noise. All
entries in BinBase were matched against the Fiehn mass spectral
library of 1,200 authentic metabolite spectra, using retention index
and mass spectrum information or the NIST11 commercial library.
Metabolites were reported if present in at least 50% of the diabetic or
nondiabetic samples. Data reported as quantitative ion peak heights
were normalized by the sum intensity of all annotated metabolites and
used for further statistical analysis.
For analysis of the complex lipids, plasma aliquots (20 l), stored
at 80°C, were thawed and extracted using a modified liquid-liquid
phase extraction approach purposed by Matyash et al. (37) and
analyzed on an Agilent 1290A Infinity Ultra High Performance Liquid
Chromatography system with an Agilent Accurate Mass-6530-QTOF
(UHPLC-QTOF). Data quality and instrument performance were
monitored throughout the data acquisition using quality control (in-
ternal STDS) and reference pooled plasma samples. The data were
processed using MZmine 2.10 software. Complex lipids were identi-
fied by searching against a precursor accurate mass and retention time
library in conjunction with matching tandem mass spectra against the
LipidBlast virtual MS/MS database(29). Metabolites were reported if
positively detected at very high mass spectral confidence in at least
50% of the diabetic or nondiabetic samples. Data, reported as peak
heights for the quantification ion (m/z) at the specific retention time
for each annotated and unknown metabolite, was normalized to the
class-specific internal standard (annotated) or to the internal standard
that had the closest retention time (unknowns).
For analysis of oxylipins, samples were extracted in accordance
with previously described protocols (60) and analyzed by Agilent
1200SL-AB Sciex 4000 QTrap ultrahigh performance liquid chroma-
tography tandem mass spectrometry (UPLC-MS/MS). Optimized
conditions and the MRM transitions, as well as extraction efficiencies,
were reported previously (60). Quality control samples were analyzed
at minimum calibration throughout the analysis. Analyst software v.
1.4.2 was used to quantify peaks according to their corresponding
standard curves. Absolute concentrations of oxylipins are presented as
nanomoles per liter.
Data Analysis
Statistical analyses. Statistical analyses were carried out on age and
sex covariate adjusted metabolite values. A linear model was used to
describe differences in metabolite abundances due to animal age and
sex, the residuals from which were tested for differences between T1D
and control animals by use of a nonparametric Mann-Whitney U-test.
The significance levels (Pvalues) were adjusted for multiple hypoth-
esis testing according to Benjamini and Hochberg (4) at a false
discovery rate (FDR) of 5% (abbreviated pFDR 0.05). Statistical
tests were calculated separately for primary metabolite, lipidomic,
oxylipin, and nonannotated measurements.
Multivariate modeling. Multivariate modeling was conducted using
orthogonal signal correction partial least squares discriminant analysis
(O-PLS-DA) (52) to identify robust predictors of metabolic differ-
ences between T1D and control animals. O-PLS-DA models were
fitted to age and sex adjusted, natural logarithm transformed, and
autoscaled data. The 71 animals were split between two-thirds training
and one-third test data sets while the proportion of T1D and control
animals in the full data was preserved. Training data were used for
model optimization and feature selection. Final model performance
was determined based on predictions for the held-out test set and
Monte Carlo cross-validation results from the training data.
Model optimization and feature selection were carried out indepen-
dently for primary metabolite, lipidomic, and oxylipin training data.
Model latent variable (LV) number and orthogonal LV (OLV) number
were selected using leave-one-out cross-validation. Feature selection
was used to identify the top 10% of all metabolic predictors for T1D
from each biochemical domain. The full variable set was filtered to
retain analytes that dispalyed 1) significant correlation with model
scores (Spearman’s pFDR 0.05) (58) and 2) model loadings on LV1
in the top 90th quantile in magnitude (41) and Mann-Whitney U-test
Pvalues 0.05.
The top 10% selected features (n44; 12 primary metabolites, 29
complex lipids, and 3 oxylipins) from each biochemical domain were
combined and evaluated using Monte Carlo training and testing
cross-validation and permutation testing. Internal training and testing
were performed by further splitting the training set into two-thirds
pseudo-training and one-third pseudo-test sets, while the proportion of
T1D and control animals was preserved. This split was randomly
repeated 1,000 times and used to estimate the distributions for the
O-PLS-DA model performance statistics: the cross-validated fit to the
training data (Q2), root mean squared error of prediction (RMSEP),
area under the receiver operator characteristic curve (AUC), sensitiv-
ity (true positive rate), and specificity (true negative rate). The
probability of achieving the models’ predictive performance was
estimated through comparison of performance statistics to those of
permuted models (random class labels) (45), which were calculated by
replicating the internal training and testing procedures described
above. Additionally, model performance was compared between the
selected (n44; primary metabolites, 12; lipids, 29; oxylipins, 3) and
excluded (bottom 90%, n351; primary metabolites, 168; lipids,
132; oxylipins, 51) feature sets.
Final model classification performance was validated through pre-
diction of class labels for the originally held-out test set and are
Table 1. Nonobese diabetic mice characteristics
Nondiabetic Diabetic
Female 18 6
Male 23 24
Age, wk 36 (26,40) 38 (26,40)
Weigh, g 27.3 4.8 19.5 4.8*
Glucose, mg/dl 94 34 513 101*
Values are reported as means SD (median, minimum,maximum). *Un-
paired two-sample t-test, P0.05.
E979SYSTEMIC ALTERATIONS IN THE METABOLOME OF DIABETIC MICE
AJP-Endocrinol Metab doi:10.1152/ajpendo.00019.2015 www.ajpendo.org
Downloaded from journals.physiology.org/journal/ajpendo (193.160.077.067) on January 9, 2021.
reported as sensitivity, specificity, and the area under the receiver
operator characteristic curve.
Network analysis. Network analysis was used to investigate statis-
tical and multivariate modeling results within a biochemical context.
A biochemical and chemical similarity network (2) was calculated for
all measured metabolites with KEGG (31) and PubChem CID (5)
identifiers using MetaMapR (22). Enzymatic interactions were deter-
mined based on product-precursor relationships defined in the KEGG
RPAIR database. Molecules not directly participating in biochemical
transformations but sharing many structural properties, based on
PubChem Substructure Fingerprints (7), were connected at a threshold
of Tanimoto similarity 0.7.
Partial correlation. A partial correlation network was calculated to
analyze empirical dependencies among the top metabolic discrimi-
nants between T1D and control animals. Partial correlations were
calculated between covariate adjusted (age and sex) and natural
logarithm transformed O-PLS-DA selected features (top 10% feature
set). To limit the scope of the partial correlation analysis to the
strongest empirical relationships, q-order partial correlations (8) were
first calculated to identify direct from indirect metabolite associations.
q-Order partial correlations (q1, 11, 22, 32) were calculated using
1,000 replications each, and were used to estimate the average
nonrejection rate () for all pairwise relationships. Analysis of the
relationship between vertex number, edge degree (connections), and
was used to select ␤⫽0.4 for edge acceptance. Coefficients of partial
correlation and Pvalues were calculated for all qvalue-identified
connections. Significant metabolite relationships were determined
based on FDR-adjusted (4) partial correlations (pFDR 0.05).
Network mapping. Network mapping was used in Cytoscape (49) to
encode statistical and multivariate modeling results through network
edge and node attributes.
Hierarchical cluster analysis. Hierarchical cluster analysis was
used to summarize relationships between classes of measured mole-
cules, glucose, and BMI. Relationships were evaluated based on
Spearman correlations between total (sum) of metabolite levels for
various biochemical subdomains, glucose, and BMI. Correlations
were visualized using a heatmap organized on the basis of hierarchal
clustering calculated on Euclidean distances and Wards agglomera-
tion.
RESULTS
Analysis of Plasma Metabolites in NOD Mice
Unbiased GC-TOF- and UHPLC-qTOF-based metabolo-
mics coupled with targeted UPLC-MS/MS analysis of oxylip-
ins was used to compare the metabolome of 41 nondiabetic and
30 diabetic animals (Table 1). A total of 1,041 metabolomic
peaks were detected, of which 395 were structurally annotated
metabolites. For GC-TOF- and UHPLC-qTOF-based metabo-
lomics, measured peak heights are reported in normalized units
using the diagnostic “unique ion” for each compound. Lipid
signaling mediators (oxylipins) are reported as concentrations
(nmol/l).
Comparison of NOD Mice’s Physical and Biochemical
Characteristics
Statistical analyses were used to compare physical charac-
teristics and metabolomic measurements between diabetic and
nondiabetic NOD mice. Diabetic mice displayed significant
elevations in fasting glucose but reduced body weight com-
pared with nondiabetic mice (Table 1).
A nonparametric Mann-Whitney U-test with FDR correction
was used to identify 493 (47%) significantly altered age- and
sex-adjusted metabolic features (P
adj
0.05) between diabetic
and nondiabetic mice (see Supplemental Table S1 in SUPPLE-
MENTAL MATERIALS, linked to the online version of this article).
Statistical comparisons of major classes of molecules and
biochemical subdomains identified general T1D-dependent in-
creases in the majority of carbohydrates, aromatics, amino
acids or amides, nucleotides, and triacylglycerides (TGs)
whereas prostacyclins, C18-diols, C18- and C20-ketones, C20-
hydroxy acids, triols, sterols, acylcarnitines (ACs), phosphati-
dylethanolamines (PEs), phosphatidylcholines (PCs), phos-
phatidylinositols (PIs), sphingomyelins (SMs), and lysophos-
phatidylcholines (LPCs) showed a general decrease (Table 2).
Table 2. Significantly altered biochemical domains in T1D vs. control animals
Biochemical Class Control Diabetic FC PValue*
Increase
Amino acid or amide (47)† 1440000 570000‡ 1730000 5e 05 1.2 0.02
Aromatic (6) 3070 1800 4200 2300 1.37 0.01
Carbohydrate (55) 2690000 9e 05 4560000 640000 1.69 0.00001
Nucleotide (14) 29100 11000 37700 17000 1.3 0.03
TG (50) 21500000 8900000 32600000 2.3e 07 1.52 0.03
Decrease
AC (4) 134000 88000 106000 1e 05 0.79 0.02
C18-diol (4) 124 80 85 89 0.69 0.00
C18-ketone (2) 9920 11000 5480 3600 0.55 0.001
C20-hydroxy (10) 9480 5600 4230 5200 0.45 0.00001
C20-ketone (3) 3050 2400 1260 1600 0.41 0.00001
LPC (12) 7970000 2600000 3270000 1800000 0.41 0.00001
PC (50) 53500000 1.1e 07 39300000 1.6e 07 0.73 0.00001
PE (10) 210000 78000 141000 79000 0.67 0.002
PI (6) 1050000 290000 478000 320000 0.46 0.00001
Prostacyclin (4) 125 82 49.8 42 0.4 0.00001
SM (20) 2380000 660000 1410000 600000 0.59 0.00001
Sterol (9) 1030000 240000 656000 280000 0.64 0.00001
Triol (2) 23 23 10.7 9.6 0.47 0.01
FC, fold change in means in diabetic vs. control animals. *Based on Mann-Whitney U-test. †Number of measured molecules. ‡Oxylipins are reported in nM;
all others in peak area heights.
E980 SYSTEMIC ALTERATIONS IN THE METABOLOME OF DIABETIC MICE
AJP-Endocrinol Metab doi:10.1152/ajpendo.00019.2015 www.ajpendo.org
Downloaded from journals.physiology.org/journal/ajpendo (193.160.077.067) on January 9, 2021.
Multivariate classification modeling using O-PLS-DA cou-
pled with feature selection was used to identify the top meta-
bolic determinants of the T1D plasma phenotype. A subset of
44 metabolites (10%; Supplemental Table S1) was selected as
part of the validated model (see Appendix Table A1) for
discrimination between diabetic and nondiabetic NOD mice.
For primary metabolites and oxylipins, biochemical and
chemical similarity network analysis was conducted to calcu-
late and visualize relationships between precursor and product
metabolite reactant pairs and molecules sharing a high degree
of structural similarity (Fig. 1). A heatmap based on hierarchi-
cal cluster analysis was used to summarize the relationships
between classes of measured molecules, fasting blood glucose,
and body weight (Fig. 2).
A Gaussian graphic model network was calculated to iden-
tify conditionally independent relationships (partial correla-
tion, P
adj
0.05) between all significant (P
adj
0.05) T1D-
associated metabolic perturbations (Fig. 3).
Diabetic NOD Mice Display Altered Carbohydrate Profiles
Total plasma carbohydrate levels were increased by 169% in
diabetic mice (Table 2) and were generally positively corre-
lated with measured fasting blood glucose and negatively
correlated with body weight (Fig. 2). Moreover, circulating
levels of glucose, gluconic acid lactone, glucuronic acid, idonic
acid, ribose, cellobiose, and 2-deoxytetronic acid, all of which
were elevated in diabetic mice compared with nondiabetic mice,
were selected as being within the top 10% discriminants of the
diabetic phenotype (Supplemental Table S1). Elevations in ribose,
2-deoxytetronic acid and cellobiose were shown to be both bio-
chemically (Fig. 1) and empirically (Fig. 3) related to general
alterations in most carbohydrates. Similarly, we observed T1D-
associated increases in organic acids and metabolites related to
energetics including a 240% increase in 2-hydroxy-2-methylbu-
tanoic acid, which was found to be the single most discriminatory
metabolite of the T1D phenotype (Supplemental Table S1). 2-Hy-
droxy-2-methylbutanoic acid was positively related to the ob-
served increase in indole-3-lactate and ribose and negatively
correlated to PC (38:2), whereas indole-3-lactate was also posi-
tively correlated to the T1D-associated increase in isocitric acid
(Fig. 3). Whereas most carbohydrates were found to be elevated in
diabetic mice, we found significant T1D-dependent decreases
in circulating 1,5-anhydroglucitol, glycerol-3-galactoside,
and threonic acid. Threonic acid was similarly identified as
a top descriptor of T1D and paralleled an observed decrease
in methionine (Fig. 3).
Diabetic NOD Mice Display Alterations in Amino Acid Profiles
Overall, 10 of the 47 measured amino acid or amine metab-
olites (21%) were significantly altered in diabetic compared
with nondiabetic animals (Fig. 1). Of these changes, six were
significantly decreased in diabetic mice, including creatinine,
Fig. 1. Biochemical network displaying metabolic differences between diabetic and nondiabetic NOD mice. Metabolites are connected based on biochemical
relationships (orange, KEGG RPAIRS) or structural similarity (blue, Tanimoto coefficient 0.7). Metabolite size and color represent the importance (O-PLS-DA
model loadings, LV1) and relative change (gray, P
adj
0.05; green, decrease; red, increase) in diabetic vs. nondiabetic NOD mice. Shapes display metabolites’
molecular classes or biochemical subdomains, and top descriptors of type 1 diabetes (T1D)-associated metabolic perturbations (Supplemental Table S1) are
highlighted with thick black borders.
E981SYSTEMIC ALTERATIONS IN THE METABOLOME OF DIABETIC MICE
AJP-Endocrinol Metab doi:10.1152/ajpendo.00019.2015 www.ajpendo.org
Downloaded from journals.physiology.org/journal/ajpendo (193.160.077.067) on January 9, 2021.
5-methoxytryptamine, methionine, cysteine, tyrosine, and tryp-
tophan (Fig. 1). Inversely, four metabolites showed T1D-
associated elevations, including a 180% increase in pan-
thothenic acid and 140% increases in valine, isoleucine, and
5-hydroxyindole-3-acetic acid (Supplemental Table S1). Of the
noted shifts, only methionine was found to be included as a top
predictor of the T1D phenotype. Overall, amino acids were
negatively correlated with body weight but not fasting blood
glucose (Fig. 2).
Diabetic NOD Mice Display Perturbed Lipid Profiles
Cumulatively, alterations in lipid metabolism best delineated
diabetic from nondiabetic mice, encompassing 28 of the top 44
most discriminatory metabolites (64%) between diabetic and
nondiabetic animals (Supplemental Table S1). Jointly, diabetic
animals displayed significant elevations in virtually all mea-
sured triacylglyericides with TG (49:1), TG (50:1), TG (12:0/
18:2/20:5), TG (51:1), TG (54:1), TG (54:2), TG (54:4), TG
(58:3), and TG (58:10) all being ranked in the top 10% of
discriminatory metabolites (Supplemental Table S1) and being
empirically related to each other (Fig. 3). Only TG (50:3) and
TG (56:4) were found to be significantly lower in diabetic than
in nondiabetic mice (Supplemental Table S1). The increase in
TG (54:1) was negatively correlated to LPC (16:1), whereas
TG (58:3) and TG (16:0/18:2/20:5) were both negatively re-
lated to CE (20:4) (Fig. 3). Notably, all measured LPCs were
significantly reduced in diabetic compared with nondiabetic
animals (Supplemental Table S1 and Table 2). The T1D-
associated reduction in LPC was positively related to similar
reductions in several other lipid species, including SM (d40:2)
A, SM (d18:2/23:0), and PI (40:6) (Fig. 3). Intriguingly, LPC
(16:1) was positively associated with methionine, which was
reduced in diabetic mice and positively correlated to the
T1D-associated reduction in LPC (20:4) (Fig. 3). LPC (20:4)
coordinately correlated with LPC (16:0), which served as a
central hub and was positively related to numerous PC and SM
lipid species, including PC (36:4), PC (p-36:3) or PC (o-36:4),
PC (38:5), SM (d18:1), and SM (d18:2/23:0) (Fig. 3). In accord
with the observed T1D-associated decrease in all measured
LPCs, nearly all measured PCs and SMs were similarly re-
duced in diabetic compared with nondiabetic mice (Table 2).
Only SM (d16:1/20:1), PC (35:2), PC (35:2) B, PC (p-36:2) or
PC (o-36:3), and PC (36:5) were found to be elevated in
diabetic mice (Supplemental Table S1). The T1D-associated
reduction in PC (38:5) was also positively associated with
similar reductions in PC (38:2), PC (37:2), and PI (38:4).
Fig. 2. Correlation between body weight,
fasting blood glucose, and major classes of
measured metabolites and lipids. Heatmap is
shown displaying hierarchical clustered Spear-
man correlations between animal characteris-
tics (weight and fasting blood glucose) and
major classes of measured metabolites and
lipids.
E982 SYSTEMIC ALTERATIONS IN THE METABOLOME OF DIABETIC MICE
AJP-Endocrinol Metab doi:10.1152/ajpendo.00019.2015 www.ajpendo.org
Downloaded from journals.physiology.org/journal/ajpendo (193.160.077.067) on January 9, 2021.
Likewise, the T1D-dependent decrease in PE (38:6) B was
positively related to PE (38:6), which positively correlated
with PI (38:4). Comparable to PCs and SMs, many of the
measured PEs and PIs were decreased in diabetic animals
(Table 2). On a global scale, PCs, SMs, PIs, LPCs, and PEs
were all positively correlated with body weight and negatively
correlated with fasting blood glucose, whereas TGs were
negatively correlated with body weight and positively corre-
lated with fasting blood glucose (Fig. 2).
Plasma free fatty acids were generally lower in diabetic
mice, led by a 60% reduction in palmitoleic acid and 40%
reductions in palmitic acid and arachidonic acid (Supplemental
Table S1). The reductions in circulating free fatty acids were
negatively correlated with fasting blood glucose (Fig. 2).
Diabetic NOD Mice Display Perturbations in Lipid
Signaling Mediators
Overall, 24 of the 54 measured oxylipins (43%) were found
to be significantly altered between diabetic and nondiabetic
animals. Furthermore, all 24 oxylipins were found to be sig-
nificantly decreased in the plasma of diabetic compared with
nondiabetic mice (Fig. 1). Particularly interesting was that
many oxylipins derived from arachidonate metabolism were
significantly reduced in diabetic mice, including LTB
4
, PGE
2
,
PGD
2
, TXB
2
, 8-HETE, 12-HETE, 15-HETE, 12-OxoETE, and
15-OxoETE (Fig. 4). Moreover, global concentrations of cir-
culating 20C-ketones, 20C-hydroxy acids, and prostacyclins
were all positively correlated with body weight and negatively
correlated with fasting blood glucose (Fig. 2 and Supplemental
Table S1). Only 12-oxo-ETE, 9-oxo-ODE, and PGF
2
were
identified as being within the top 10% of discriminant metab-
olites (Supplemental Table S1). The T1D-associated decrease
in 12-oxo-ETE paralleled the observed reductions in PC (38:2)
and threonic acid (Fig. 3).
DISCUSSION
In the previous study, using NOD mice, we characterized
progressors (diabetic) and nonprogressors (nondiabetic) to
Fig. 3. Partial correlation network displaying associations between top 10% T1D-dependent metabolic perturbations. All significantly top 10% discriminant
metabolites (n44) of the T1D-phenotype are connected based on partial correlations (P
adj
0.5, Supplemental Table S1). Edge width displays the absolute
magnitude and color the direction (orange, positive; blue, negative) of the partial coefficient of correlation. Metabolite size and color represent the importance
(O-PLS-DA model loadings, LV 1) and relative change (green, decrease; red, increase) in diabetic vs. nondiabetic mice. Shapes display metabolites’ molecular
classes or biochemical subdomains (see Fig. 4 legend).
E983SYSTEMIC ALTERATIONS IN THE METABOLOME OF DIABETIC MICE
AJP-Endocrinol Metab doi:10.1152/ajpendo.00019.2015 www.ajpendo.org
Downloaded from journals.physiology.org/journal/ajpendo (193.160.077.067) on January 9, 2021.
T1D and performed an initial screening of primary metabolites
in the plasma of these mice with the aims of identifying
circulating factors linked with -cell viability and/or T1D
progression (23). In the current study, we aimed to expand
upon our initial observations by using a multiplatform ap-
proach to recapitulate our initial metabolomic findings in
addition to capturing perturbations in the lipidome and lipid
signaling mediators. Using this unique multiplatform ap-
proach, we detected more than 1,000 lipids and metabolites
(396 annotated) in the plasma of progressors and nonprogres-
sors.
Our analysis of the primary metabolites (carbohydrates,
amino acids, organic acids, nucleotides, free fatty acids) re-
vealed numerous age- and sex-adjusted T1D-dependent altera-
tions that extended to multiple biochemical domains. Importantly,
29 of the 60 annotated metabolites distinguishing progressors
from nonprogressors were also found to be significantly different
in our initial study (23) (see Appendix Table A2). Furthermore, all
29 metabolites maintained the same direction in change providing
validation for both our initial and current study (see Appendix
Table A2).
In comparison with nonprogressors, diabetic mice indicated
a 69% increase in global circulating carbohydrates. Most
notably, diabetic mice displayed significant elevations in cir-
culating glucose and significant reductions in 1,5-anhydroglu-
citol (1,5-AG), a marker of glycemic control (28). The eleva-
tion in glucose and reduction in 1,5-AG is consistent with the
diabetic model and provides assurance in our analytical anal-
ysis.
In concurrence with the increases in circulating carbohy-
drates, carbohydrate oxidation products glucono--lactone
(GDL), an oxidation product of glucose leading to the produc-
tion of hydrogen peroxide (33), and galactonic acid, an oxida-
tion product of galactose, were elevated in diabetic compared
with nondiabetic mice (Fig. 1). The elevation in circulating
GDL and galactonic acid may reflect elevated states of oxida-
tive stress, a well-known hallmark of T1D (36). Consistent
with this notion, we observed significant T1D-dependent re-
ductions in circulating levels of methionine and cysteine (Fig.
1). Methionine and cysteine are both components of the trans-
sulfuration pathway, which leads to the synthesis of glutathi-
one (not measured), an important intracellular antioxidant.
Previous studies have demonstrated that insulin-deprived T1D
subjects have reduced rates of homocysteine-methionine rem-
ethylation and increased rates of transsulfuration compared
with control subjects (1). Accordingly, patients with poor
glycemic control have been shown to have depleted glutathione
pools and reduced erythrocyte free cysteine concentrations
(11). The reduction in circulating methionine and cysteine
may, therefore, reflect an increased flux into the transsulfura-
tion pathway and glutathione synthesis. In the current investi-
gation, the reduction in methionine paralleled the observed
reduction in threonic acid, a breakdown product of the antiox-
idant ascorbic acid (vitamin C; Fig. 4) (54). The T1D-associ-
ated reduction in circulating threonic acid suggests impaired
ascorbic acid metabolism. This is a noteworthy observation
given that ascorbic acid uptake into the tissue of diabetic
animals is often depressed during states of hyperglycemia (10),
rendering cells more susceptible to oxidative stress. Addition-
ally, diabetic mice indicated a significant (raw Pvalue 0.017)
33% reduction in -tocopherol, a lipophilic antioxidant (see
Supplemental Table S2). Collectively, the above-mentioned
metabolic aberrations point toward a heightened state of oxi-
dative stress in mice that progress to T1D. In this sense, it is
Fig. 4. Significantly perturbed eicosanoids (P
adj
0.05) in the KEGG arachidonate metabolism pathway. Pathway enrichment was determined based on FDR
adjusted hypergeometric t-test, P0.05 for KEGG pathways for Mus musculus. Figure displays relative fold changes (blue, decrease; yellow, increase) in means
between diabetic and nondiabetic mice. *Values reported as means SD unless otherwise noted; †values reported median (minimum, maximum); ‡unpaired
two-sample t-test, P0.05.
E984 SYSTEMIC ALTERATIONS IN THE METABOLOME OF DIABETIC MICE
AJP-Endocrinol Metab doi:10.1152/ajpendo.00019.2015 www.ajpendo.org
Downloaded from journals.physiology.org/journal/ajpendo (193.160.077.067) on January 9, 2021.
interesting to note that cysteine, threonic acid, and methionine
were negatively correlated with fasting blood glucose and
positively correlated with 1,5-AG (see Appendix Table A3).
We therefore postulate that reductions in circulating cysteine,
threonic acid, and methionine serve as candidate biomarkers
for monitoring oxidative stress and predicting T1D develop-
ment and progression.
Although it was evident that there were many T1D-associ-
ated alterations in primary metabolism, one of the more pro-
nounced differences reflective of the progression to T1D was
observed in lipid species and lipid signaling mediators. Con-
sistent with our initial finding, diabetic NOD mice displayed
significant reductions in selected free fatty acids, most notably
arachidonic acid (AA), palmitoleic acid (PA) and oleic acid
(OA) compared with nonprogressors (Fig. 1). Free fatty acids
exhibit complex interactions with -cells, acting as both secre-
tagogues (14) and agents of apoptotic signaling (9, 26, 42, 55).
Moreover, the development of T1D is known to have a strong
inflammatory component (3, 34, 43). Cyclooxygenase
(COX)-2 and 12-lipoxygenase (LOX) derived products of AA,
such as prostaglandin E2 (PGE2) and 12-hydroxyeicosatetra-
enoic acid (12-HETE), have been shown to play critical roles
in cytokine-induced human -cell destruction (9, 26, 42, 55).
Thus, one may anticipate that the reduction in circulating AA
represents a shift in equilibrium toward eicosanoid production.
On the contrary, we found significant T1D-associated reduc-
tions in several AA-derived eicosanoids, including thrombox-
ane A2 (TXA2), leukotriene B4 (LTB4), PGD
2
, PGE2, 11-,12-
and 15-HETE, and 12-oxo-ETE. This suggests a reduced
bioavailability of AA rather than an increased flux into these
lipid signaling pathways (Fig. 4). The reduction in circulating
oxylipins, particularly those derived from AA, also suggests
that mice that have progressed to T1D are under a reduced state
of systemic inflammation. Kriegel et al. (32) previously re-
ported marked infiltration of CD11cCD11bdendritic cells
(DC) into the pancreatic islets of NOD mice. Interestingly,
these DCs exhibited a hypoactive phenotype, failed to induce
proliferation of diabetogenic CD4T-cells in vitro, and were
potent suppressors of diabetes development (32). Additionally,
Bouma et al. (6) demonstrated that NOD mice exhibit impaired
recruitment of leukocytes into sites of inflammation in the
peritoneum and subcutaneously elicited air pouches. It should
be noted that in both studies the cohort of NOD mice were
under 8 wk of age and, as such, cannot be exclusively com-
pared with our diabetic cohort. Despite this limitation, both
studies demonstrate marked modulation of overall immune and
inflammatory status in NOD mice and T1D development.
Moreover, both studies hint toward the fact that reduced
recruitment of inflammatory mediators accompanies T1D de-
velopment. However, it is important to note that we evaluated
systemic alterations of inflammatory mediators as opposed to
pancreas-specific changes; thus, changes in circulation reflect
the contributions all organ types. Despite these findings, AA, in
addition to PA and OA, has been also been shown to attenuate
the deleterious effects of high glucose on human pancreatic
-cell turnover and function (12, 13, 35). Similarly, reductions
in 9,10-DiHODE and 11- and 12-HETE have been observed in
men with hyperlipidemia, a frequent T1D-associated compli-
cation (15, 48). Future studies will be required to confirm our
initial observations and fully elucidate how the observed re-
ductions in circulating lipid mediators and free fatty acids
relate to T1D development and progression.
The coexistence of lipid dysfunction and hyperlipidemia is a
known to be intertwined with T1D pathophysiology. Hypertri-
glyceridemia is strongly linked to poor glycemic control (21,
46). Our analysis of the lipidome revealed significant T1D-
associated increases in virtually all circulating triglycerides
(Supplemental Table S1). Moreover, global abundances of
triglycerides were positively correlated with fasting blood
glucose (Fig. 3), indicative of hypertriglyceridemia. Inverse to
the observed increases in circulating triglycerides, diabetic
NOD mice displayed striking reductions in several lyso-
phophospholipids, phospholipids, and sphingolipids (Supple-
mental Table S1). It is important to note that these major lipid
classes were highly correlated with body weight, which was
found to be significantly lower in our diabetic cohort, and
negatively correlated with fasting blood glucose (Fig. 3). T1D
is characterized by increases in amino acid catabolism (25) and
fatty acid oxidation (24, 27), a consequence of the inability to
utilize glucose as a source of energy. Weight loss is frequently
observed in T1D, which is attributed to several T1D-associated
complications such as muscle atrophy and dehydration (57,
59). Consequently, the observed reduction in major structural
lipids may be a consequence of increased fatty acid -oxida-
tion and reduced bioavailability of fatty acids for phospholipid/
structural lipid synthesis, a notion that is partially supported by
the observed decrease in circulating free fatty acids and weight
loss. Despite the strong association between body weight and
major structural lipids, alterations in lipid metabolism are
involved in the etiology of T1D. Transiently elevated LPC
serum levels have been positively associated with seroconver-
sion to islet autoantibody positivity in children who subse-
quently progress to T1D (40). Furthermore, LPCs have been
shown to improve blood glucose levels in mouse models of
T1D and T2D (61) and act as insulin secretagogues (50).
Sysi-Aho et al. (53) found that female NOD mice that did not
progress to T1D after 8 wk of age despite being insulin
autoantibody (IAA) positive had elevated levels of selected
serum LPC species compared with progressors. Additionally, it
was demonstrated that IAA positivity and LPC concentrations
were able to reasonably discriminate between progressors and
nonprogressors (53). Collectively, the findings by Sysi-Aho et
al. and others suggest a protective effect by LPCs. It is
important to note that the studies by Sysi-Aho and colleagues
focused primarily on metabolic alterations that occur prior and
during seroconversion in young (25 wk) NOD mice. The
cohort of NOD mice used in the current investigation extended
beyond 25 wk and included NOD mice that escaped chronic
hyperglycemia, a group that has been largely excluded in
previous studies. The inclusion of mice that avoided develop-
ing overt T1D was particularly important, since it allowed us to
directly compare between progressors and nonprogressors both
undergoing similar autoimmune insults. In our model system,
LPCs were negatively correlated with fasting blood glucose
(Fig. 3) corroborating the notion that LPCs are associated with
glycemic control and -cell dysfunction/destruction. Interest-
ingly, selected LPCs were positively related to the T1D-
associated reduction in methionine (Fig. 4). Pflueger et al. (44)
previously found that concentrations of methionine were lower
in children who developed autoantibodies by age 2 yr com-
pared with those who developed autoantibodies later in child-
E985SYSTEMIC ALTERATIONS IN THE METABOLOME OF DIABETIC MICE
AJP-Endocrinol Metab doi:10.1152/ajpendo.00019.2015 www.ajpendo.org
Downloaded from journals.physiology.org/journal/ajpendo (193.160.077.067) on January 9, 2021.
hood or remained autoantibody negative. Methionine is a vital
component in one-carbon metabolism and serves as the pre-
cursor for s-adenosylmethionine synthesis. s-Adenosyl methi-
onine in turn acts as a methyl donor for the synthesis of PCs
from PEs (38). One-carbon metabolism has been shown to be
perturbed in animal models of T1D (38). The observed reduc-
tion in methionine in diabetic mice may therefore not only be
an indicator of increased flux into the transsulfuration pathway
but may also partially account for the T1D-dependent reduc-
tions in PCs and related LPCs. This is a notable observation as
Oresic et al. (40) demonstrated that children who developed
T1D had reduced serum levels of PCs at birth and were
consistently low in progressors compared with nonprogressors.
Similarly, analysis of umbilical cord serum lipidome in infants
who later developed T1D revealed distinct reductions in major
choline-containing phospholipids, including sphingomyelins
and PCs (39). Our results suggest that the reduction in circu-
lating LPC and PC species can be used as potential biomarkers
for T1D onset and progression and -cell loss.
In conclusion, a comprehensive metabolomics approach was
used to identify 493 (192 annotated) significant age- and
sex-independent metabolic alterations associated with T1D
development and progression. Our findings suggest that pro-
gression to T1D is characterized by increased oxidative stress,
perturbed states of inflammation, and altered lipid metabolism.
More importantly, these findings serve as a basis for the
identification of metabolic states (i.e., altered amino acid pro-
files in conjunction with altered lipid profiles) that can be used
as diagnostic and prognostic indicators of T1D pathophysiol-
ogy and consequently -cell death. In particular, we propose
that reductions in circulating threonic acid, methionine, and
cysteine may serve as stable markers of oxidative stress and
T1D progression, whereas reductions in phosposphatidylcho-
lines and related lysophosphatidylcholines may serve as can-
didate biomarkers of glycemic control, T1D onset, and -cell
destruction.
APPENDIX
Analysis of the Metabolome, Lipidome, and Lipid Signaling
Mediators
For analysis of primary metabolites, 30-l plasma aliquots, which
were extracted with 1 ml of degassed acetonitrile-isopropanol-water
(3:3:2) at 20°C and centrifuged, the supernatant was removed, and
solvents were evaporated to dryness under reduced pressure. To
remove membrane lipids and triglycerides, dried samples were recon-
stituted with acetonitrile-water (1:1), decanted, and taken to dryness
under reduced pressure. Internal standards, C8 –C30 fatty acid methyl
esters (FAMEs), were added to samples and derivatized with me-
thoxyamine hydrochloride in pyridine and subsequently by MSTFA
(Sigma-Aldrich) for trimethylsilylation of acidic protons and analyzed
by GC-TOF-MS. An Agilent 7890A gas chromatograph (Santa Clara,
CA) was used with a 30 m long, 0.25 mm ID Rtx5Sil-MS column with
0.25-m 5% diphenyl film; an additional 10-m integrated guard
column was used (Restek, Bellefonte, PA) (16, 30, 56). A Gerstel
MPS2 automatic liner exchange system (ALEX) was used to elimi-
nate sample cross-contamination during the GC-TOF analysis. A
0.5-l of sample was injected at 50°C (ramped to 250°C) in splitless
mode with a 25-s splitless time. The chromatographic gradient con-
sisted of a constant flow of 1 ml/min, ramping the oven temperature
from 50°C to 330°C over 22 min. Mass spectrometry was done using
a Leco Pegasus IV TOF mass spectrometer, 280°C transfer line
temperature, electron ionization at 70 V, and an ion source temper-
ature of 250°C. Mass spectra were acquired at 1,525 V detector
voltage at m/z 85–500 with 17 spectra/s. Acquired spectra were further
processed using the BinBase database (18, 47). Briefly, output results
(30) were filtered based on multiple parameters to exclude noisy or
inconsistent peaks. All entries in BinBase were matched against the
Fiehn mass spectral library of 1,200 authentic metabolite spectra,
using retention index and mass spectrum information or the NIST11
commercial library.
For analysis of the lipidome, plasma aliquots (20 l), stored at
80°C, were thawed and extracted using a modified liquid-liquid
phase extraction approach purposed by Matyash et al. (37). Briefly,
225 l of chilled methanol containing an internal standard mixture
[PE(17:0/17:0); PG(17:0/17:0); PC(17:0/0:0); C17 sphingosine; C17
ceramide; SM (d18:0/17:0); palmitic acid-d3; PC (12:0/13:0); choles-
terol-d7; TG (17:0/17:1/17:0)-d5; DG (12:0/12:0/0:0); DG (18:1/2:0/
0:0); MG (17:0/0:0/0:0); PE (17:1/0:0); LPC (17:0); LPE (17:1)], and
750 l of chilled MTBE containing the internal standard 22:1 cho-
lesteryl ester was added to 10-l aliquots of sample. Samples were
shaken for 6 min at 4°C using an Orbital Mixing Chilling/Heating
Plate (Torrey Pines Scientific Instruments) followed by the addition of
188 l of room temperature distilled water. Samples were vortexed
and centrifuged, and the upper layer was transferred to a new 1.5-ml
Eppendorf tube. An aliquot was dried to completeness using a Lab-
conco Centrivap. Upon complete dryness, samples were resuspended
in methanol-toluene (90:10) with 50 ng/ml CUDA ([[12-(cyclohexy-
lamino)carbonyl]amino]-dodecanoic acid, Cayman Chemical). Sam-
ples were vortexed, sonicated for 5 min, and centrifuged, whereafter
100 l of the sample was transferred to an amber glass vial (National
Scientific C4000-2W) with a microinsert (Supelco 27400-U). Lipid
extracts were subsequently analyzed on an Agilent 1290A Infinity
Ultra High Performance Liquid Chromatography system with an
Agilent Accurate Mass-6530-QTOF in both positive and negative
modes. The column (65°C) was a Waters Acquity UPLC CSH C18
(100 mm length 2.1 mm ID; 1.7 M particles) coupled with a
Waters Acquity VanGuard CSH C18 1.7 M precolumn. The solvent
system included (A) 60:40 vol/vol acetonitrile-water (LCMS grade)
containing 10 mM ammonium formate and 0.1% formic acid and (B)
90:10 vol/vol isopropanol-acetonitrile containing 10 mM ammonium
formate and 0.1% formic acid. The gradient started from 0 min 15%
(B), 0 –2 min 30% (B), 2–2.5 min 48% (B), 2.5–11 min 82% (B),
11–11.5 min 99% (B), 11.5–12 min 99% (B), 12–12.1 min 15% (B),
and 12.1–15 min 15% (B). The flow rate was 0.6 ml/min and with an
injection volume of 1.67 l for ESI () and 5 l for ESI () mode
acquisition. ESI capillary voltage was 3.5 kV and 3.5 kV with
collision energies of 25 and 40 eV for MS/MS collection in positive
and negative acquisition modes, respectively. Data were collected at a
mass range of m/z 60 –1700 Da with a spectral acquisition speed of 2
spectra/s. Method blanks and pooled sterile human plasma samples
were included to serve as additional quality controls. CUDA was used
to monitor instrument performance. Data were processed using
MZmine 2.10. All peak intensities are representative of peak heights.
Annotations were completed by matching experimental accurate mass
MS/MS spectra to MS/MS libraries, including Metlin-MSMS,
NIST12, and LipidBlast (29). Spectral matching was automated using
the MSPepSearch tool and manually curated using The NIST Mass
Spectral Search Program v. 2.0g. Metabolite libraries were created, in
positive and negative ionization modes, containing all confirmed
identified compounds. MZmine’s Custom Database Search tool was
used to assign annotations based on accurate mass and retention time
matching.
For analysis of lipid signaling mediators (oxylipins), samples were
extracted in accordance with previously described protocols (60).
Briefly, plasma samples underwent solid-phase extraction (SPE) on
60-mg Waters Oasis-HLB cartidges (Milford, MA). The elutions from
the SPE cartridges were evaporated to dryness using a Speedvac
(Jouan, St-Herblain, France) and reconstituted in 200 nM CUDA in
methanol and analyzed by UPLC-MS/MS. The LC system used for
E986 SYSTEMIC ALTERATIONS IN THE METABOLOME OF DIABETIC MICE
AJP-Endocrinol Metab doi:10.1152/ajpendo.00019.2015 www.ajpendo.org
Downloaded from journals.physiology.org/journal/ajpendo (193.160.077.067) on January 9, 2021.
analysis was an Agilent 1200 SL (Palo Alto, CA) equipped with a
2.1 150 mm Eclipse Plus C18 column with a 1.8 M particle size
(Agilent). The autosample was kept at 4°C. Mobile phase A consisted
of water with 0.1% glacial acetic acid. Mobile phase B consisted of
LCMS-grade acetonitrile-methanol (84:16 vol/vol) with 0.1% glacial
acetic acid. Gradient elution was performed at a flow rate of 250
l/min. Chromatography was optimized to separate all analytes in
21.5 min according to their polarity, with the most polar analytes,
prostaglandins, and leukotrienes eluting first, followed by the hy-
droxyl and epoxy fatty acids. The column was connected to a 4000
QTrap tandem mass spectrometer (Applied Biosystems Instrument,
Foster City, CA) equipped with an electrospray source (Turbo V). The
instrument was operated in negative multiple reaction monitoring
(MRM) mode. The optimized conditions and the MRM transitions, as
well as extraction efficiencies, were reported previously (60). Quality
control samples were analyzed at minimum calibration throughout the
analysis. Analyst 1.4.2 software was used to quantify peaks according
to their standard curve.
Table A1. O-PLS-DA model performance and validation statistics.
Table A2. List of compounds significantly different in both the
initial and follow-up studies.
Supplemental Table S2. List of all identified peaks which were
significantly different in T1D compared to control animals.
Supplemental Table S3. List of all identified peaks which were not
significantly different in T1D compared to control animals.
Table A3. Spearman rank correlations between fasting blood glu-
cose, threonic acid, methionine, cysteine and 1,5-anhydroglucitol.
GRANTS
This research is funded by the National Institutes of Health Grant U24
DK-097154, and is partially supported by NIEHS R01 ES-002710 and P42
ES-004699.
DISCLOSURES
No conflicts of interest, financial or otherwise, are declared by the
author(s).
AUTHOR CONTRIBUTIONS
Author contributions: J.F. and J.Y. performed experiments; J.F., D.G., and
J.Y. analyzed data; J.F., D.G., J.Y., and M.H. interpreted results of experi-
ments; J.F. and D.G. prepared figures; J.F., D.G., and M.H. drafted manuscript;
J.F., D.G., J.Y., B.H., O.F., G.I.B., and M.H. approved final version of
manuscript; J.Y., B.H., O.F., and G.I.B. edited and revised manuscript; O.F.,
G.I.B., and M.H. conception and design of research.
Table A1. O-PLS-DA model performance and validation statistics
Model Q
2
RMSEP AUC Sensitivity Specificity
Training data 0.691 0.05* 0.341 0.06* 0.844 0.09* 0.902 0.08** 0.786 0.17*
Test data 0.652 0.277 0.944 1 0.889
Q
2
, cross-validated fit to the training data; RMSEP, root mean squared error of prediction; AUC, area under the receiver operator characteristic curve.
*Performance worse than only 5% of all permuted models. **Performance worse than only 10% of all permuted models.
Table A2. List of compounds significantly different in both the initial and follow-up studies
Initial Study (23) Current Study
Metabolite† FC‡ Direction§ Padj** FC‡ Direction§ Padj**
1,5-Anhydroglucitol 0.4 Down 0.00000 0.13 Down 0.00000
2-Deoxytetronic acid 1.4 Up 0.04390 1.12 Up 0.00384
2-Hydroxy-2-methylbutanoic acid 2.2 Up 0.00061 2.44 Up 0.00000
3,6-Anhydrogalactose 4.3 Up 0.00000 1.82 Up 0.00002
4-Hydroxybutyric acid 2.5 Up 0.00000 1.23 Up 0.02300
5-Hydroxyindole-3-acetic acid 2 Up 0.00002 1.41 Up 0.03900
Adenosine-5-phosphate 0.5 Down 0.00113 0.52 Down 0.00021
Arabinose 1.6 Up 0.01850 1.59 Up 0.00755
Arachidonic acid 0.7 Down 0.00136 0.57 Down 0.00065
Creatinine 0.6 Down 0.01130 0.36 Down 0.00374
Cysteine 0.7 Down 0.03990 0.67 Down 0.03400
Erythritol 1.6 Up 0.01140 1.85 Up 0.00391
Galactinol 2.6 Up 0.00964 1.15 Up 0.04980
Galactonic acid 2.2 Up 0.00000 1.61 Up 0.00531
Glucose 3 Up 0.00000 2.08 Up 0.00000
Glycerol--phosphate 0.6 Down 0.00003 0.52 Down 0.00091
Indole-3-lactate 2.2 Up 0.00114 2.08 Up 0.00379
Isocitric acid 1.4 Up 0.02390 1.56 Up 0.00564
Isoleucine 1.9 Up 0.01870 1.43 Up 0.01250
Isothreonic acid 1.5 Up 0.00584 1.69 Up 0.00415
Lactobionic acid 3.2 Up 0.00044 2.86 Up 0.00000
Maltose 3.1 Up 0.00000 2.50 Up 0.00000
Oleic acid 0.7 Down 0.03470 0.68 Down 0.04830
Pipecolic acid 2 Up 0.00901 1.85 Up 0.03890
Saccharic acid 3.3 Up 0.00000 1.64 Up 0.01020
Sorbitol 2.4 Up 0.00048 2.27 Up 0.00003
Threonic acid 0.7 Down 0.00209 0.36 Down 0.00000
Tyrosine 0.9 Down 0.03990 0.65 Down 0.03700
Valine 2 Up 0.00785 1.43 Up 0.04440
*GC-TOF. †Metabolite name reported as a BinBase identifier. ‡Fold change of means in diabetic vs. nondiabetic animals. §Direction in change relative to
nondiabetic animals. **False discovery rate adjusted Pvalue.
E987SYSTEMIC ALTERATIONS IN THE METABOLOME OF DIABETIC MICE
AJP-Endocrinol Metab doi:10.1152/ajpendo.00019.2015 www.ajpendo.org
Downloaded from journals.physiology.org/journal/ajpendo (193.160.077.067) on January 9, 2021.
REFERENCES
1. Abu-Lebdeh HS, Barazzoni R, Meek SE, Bigelow ML, Persson XM,
Nair KS. Effects of insulin deprivation and treatment on homocysteine
metabolism in people with type 1 diabetes. J Clin Endocrinol Metab 91:
3344 –3348, 2006.
2. Barupal DK, Haldiya PK, Wohlgemuth G, Kind T, Kothari SL,
Pinkerton KE, Fiehn O. MetaMapp: mapping and visualizing metabo-
lomic data by integrating information from biochemical pathways and
chemical and mass spectral similarity. BMC Bioinform 13: 99, 2012.
3. Baumann B, Salem HH, Boehm BO. Anti-inflammatory therapy in type
1 diabetes. Curr Diabet Reports 12: 499 –509, 2012.
4. Benjamini Y, Hochberg Y. Controlling the false discovery rate: a
practical and powerful approach to multiple testing. J Royal Stat Soc B
(Methodological) 289 –300, 1995.
5. Bolton EE, Wang Y, Thiessen PA, Bryant SH. PubChem: integrated
platform of small molecules and biological activities. Ann Rep Comput
Chem 4: 217–241, 2008.
6. Bouma G, Nikolic T, Coppens JM, van Helden-Meeuwsen CG, Leenen
PJ, Drexhage HA, Sozzani S, Versnel MA. NOD mice have a severely
impaired ability to recruit leukocytes into sites of inflammation. Eur J
Immunol 35: 225–235, 2005.
7. Cao Y, Charisi A, Cheng LC, Jiang T, Girke T. ChemmineR: a
compound mining framework for R. Bioinformatics 24: 1733–1734, 2008.
8. Castelo R, Roverato A. Reverse engineering molecular regulatory net-
works from microarray data with qp-graphs. J Comput Biol 16: 213–227,
2009.
9. Chen M, Yang ZD, Smith KM, Carter JD, Nadler JL. Activation of
12-lipoxygenase in proinflammatory cytokine-mediated cell toxicity.
Diabetologia 48: 486 –495, 2005.
10. Cunningham JJ. The glucose/insulin system and vitamin C: implications
in insulin-dependent diabetes mellitus. J Am Coll Nutr 17: 105–108, 1998.
11. Darmaun D, Smith SD, Sweeten S, Sager BK, Welch S, Mauras N.
Evidence for accelerated rates of glutathione utilization and glutathione
depletion in adolescents with poorly controlled type 1 diabetes. Diabetes
54: 190 –196, 2005.
12. Das UN. Arachidonic acid and lipoxin A4 as possible endogenous anti-
diabetic molecules. Prostaglandins Leukot Essential Fatty Acids 88:
201–210, 2013.
13. Dixon G, Nolan J, McClenaghan NH, Flatt PR, Newsholme P. Ara-
chidonic acid, palmitic acid and glucose are important for the modulation
of clonal pancreatic -cell insulin secretion, growth and functional integ-
rity. Clin Sci (London, 1979) 106: 191–199, 2004.
14. Dobbins RL, Chester MW, Stevenson BE, Daniels MB, Stein DT,
McGarry JD. A fatty acid- dependent step is critically important for both
glucose- and non-glucose-stimulated insulin secretion. J Clin Invest 101:
2370 –2376, 1998.
15. Dunn FL. Management of hyperlipidemia in diabetes mellitus. Endocri-
nol METAB CLIN NORTH AM 21: 395–414, 1992.
16. Fiehn O. Extending the breadth of metabolite profiling by gas chroma-
tography coupled to mass spectrometry. Trends Analyt Chem 27: 261–269,
2008.
17. Fiehn O, Wohlgemuth G, Scholz M. Setup and annotation of metabo-
lomic experiments by integrating biological and mass spectrometric meta-
data. In: Data Integration in the Life Sciences, edited by Ludäscher B,
Raschid L. Springer Berlin Heidelberg: 2005, p. 224 –239.
18. Fiehn O, Wohlgemuth G, Scholz M. Setup and annotation of metabo-
lomic experiments by integrating biological and mass spectrometric meta-
data. Data Integr Life Sci Proc 3615: 224 –239, 2005.
19. Fiehn O, Wohlgemuth G, Scholz M, Kind T, Lee do Y, Lu Y, Moon S,
Nikolau B. Quality control for plant metabolomics: reporting MSI-
compliant studies. Plant J 53: 691–704, 2008.
20. Friedrich N. Metabolomics in diabetes research. J Endocrinol 215:
29 –42, 2012.
21. Goldberg IJ. Clinical review 124: Diabetic dyslipidemia: causes and
consequences. J Clin Endocrinol Metab 86: 965–971, 2001.
22. Grapov D. MetaMapR: Metabolomic Mapping and Analysis Tools. ver-
sion 131. https://github.com/dgrapov/MetaMapR: 2014.
23. Grapov DF, Hwang J, Poudel J, Jo A, Periwal J, Fiehn V, Hara O, M.
Diabetes associated metabolomic perturbations in NOD mice. Metabolo-
mics 11: 425–437, 2015.
24. Guder WG, Schmolke M, Wirthensohn G. Carbohydrate and lipid
metabolism of the renal tubule in diabetes mellitus. Eur J Clin Chem Clin
Biochem 30: 669 –674, 1992.
25. Hebert SL, Nair KS. Protein and energy metabolism in type 1 diabetes.
Clin Nutr (Edinburgh) 29: 13–17, 2010.
26. Heitmeier MR, Kelly CB, Ensor NJ, Gibson KA, Mullis KG, Corbett
JA, Maziasz TJ. Role of cyclooxygenase-2 in cytokine-induced -cell
dysfunction and damage by isolated rat and human islets. J Biol Chem
279: 53145–53151, 2004.
27. Herrero P, Peterson LR, McGill JB, Matthew S, Lesniak D, Dence C,
Gropler RJ. Increased myocardial fatty acid metabolism in patients with
type 1 diabetes mellitus. J Am Coll Cardiol 47: 598 –604, 2006.
28. Kim WJ, Park CY. 1,5-Anhydroglucitol in diabetes mellitus. Endocrine
43: 33–40, 2013.
29. Kind T, Liu KH, Lee DY, DeFelice B, Meissen JK, Fiehn O. LipidBlast
in silico tandem mass spectrometry database for lipid identification. Nat
Methods 10: 755–758, 2013.
30. Kind T, Tolstikov V, Fiehn O, Weiss RH. A comprehensive urinary
metabolomic approach for identifying kidney cancerr. Anal Biochem 363:
185–195, 2007.
31. Kotera M, Hirakawa M, Tokimatsu T, Goto S, Kanehisa M. The
KEGG databases and tools facilitating omics analysis: latest developments
involving human diseases and pharmaceuticals. In: Next Generation Mi-
croarray Bioinformatics. Berlin: Springer, 2012, p. 19 –39.
32. Kriegel MA, Rathinam C, Flavell RA. Pancreatic islet expression of
chemokine CCL2 suppresses autoimmune diabetes via tolerogenic
CD11cCD11bdendritic cells. Proc Natl Acad Sci USA 109: 3457–
3462, 2012.
33. Lindsay RM, Smith W, Lee WK, Dominiczak MH, Baird JD. The
effect of delta-gluconolactone, an oxidised analogue of glucose, on the
nonenzymatic glycation of human and rat haemoglobin. Clin Chim Acta
263: 239 –247, 1997.
34. Luo P, Wang MH. Eicosanoids, -cell function, and diabetes. Prosta-
glandins Other Lipid Mediators 95: 1–10, 2011.
35. Maedler K, Oberholzer J, Bucher P, Spinas GA, Donath MY. Mono-
unsaturated fatty acids prevent the deleterious effects of palmitate and high
glucose on human pancreatic -cell turnover and function. Diabetes 52:
726 –733, 2003.
36. Matough FA, Budin SB, Hamid ZA, Alwahaibi N, Mohamed J. The
role of oxidative stress and antioxidants in diabetic complications. Sultan
Qaboos Univ Med J 12: 5–18, 2012.
37. Matyash V, Liebisch G, Kurzchalia TV, Shevchenko A, Schwudke D.
Lipid extraction by methyl-tert-butyl ether for high-throughput lipidomics.
J Lipid Res 49: 1137–1146, 2008.
38. Nieman KM, Schalinske KL. Insulin administration abrogates perturba-
tion of methyl group and homocysteine metabolism in streptozotocin-
treated type 1 diabetic rats. Am J Physiol Endocrinol Metab 301: E560
E565, 2011.
39. Oresic M, Gopalacharyulu P, Mykkanen J, Lietzen N, Makinen M,
Nygren H, Simell S, Simell V, Hyoty H, Veijola R, Ilonen J, Sysi-Aho
M, Knip M, Hyotylainen T, Simell O. Cord serum lipidome in prediction
of islet autoimmunity and type 1 diabetes. Diabetes 62: 3268 –3274, 2013.
40. Oresic M, Simell S, Sysi-Aho M, Nanto-Salonen K, Seppanen-Laakso
T, Parikka V, Katajamaa M, Hekkala A, Mattila I, Keskinen P,
Yetukuri L, Reinikainen A, Lahde J, Suortti T, Hakalax J, Simell T,
Hyoty H, Veijola R, Ilonen J, Lahesmaa R, Knip M, Simell O.
Dysregulation of lipid and amino acid metabolism precedes islet autoim-
munity in children who later progress to type 1 diabetes. J Exp Med 205:
2975–2984, 2008.
41. Palermo G, Piraino P, Zucht HD. Performance of PLS regression
coefficients in selecting variables for each response of a multivariate PLS
for omics-type data. Adv Applic Bioinform Chem 2: 57, 2009.
Table A3. Spearman rank correlations between fasting
blood glucose, threonic acid, methionine, cysteine and 1,5-
anhydroglucitol
Correlation Matrix
(Spearman)
1
Blood
Glucose
Threonic
Acid Methionine Cysteine
Blood glucose
Threonic acid 0.696*
Methionine 0.496* 0.732*
Cysteine 0.441* 0.636* 0.643*
1,5-Anhydroglucitol 0.839* 0.753* 0.553* 0.537*
1
Values represent Spearman rank correlation coefficients (R). *Pvalue 0.001.
E988 SYSTEMIC ALTERATIONS IN THE METABOLOME OF DIABETIC MICE
AJP-Endocrinol Metab doi:10.1152/ajpendo.00019.2015 www.ajpendo.org
Downloaded from journals.physiology.org/journal/ajpendo (193.160.077.067) on January 9, 2021.
42. Parazzoli S, Harmon JS, Vallerie SN, Zhang T, Zhou H, Robertson
RP. Cyclooxygenase-2, not microsomal prostaglandin E synthase-1, is the
mechanism for interleukin-1-induced prostaglandin E2 production and
inhibition of insulin secretion in pancreatic islets. J Biol Chem 287:
32246 –32253, 2012.
43. Pedicino D, Liuzzo G, Trotta F, Giglio AF, Giubilato S, Martini F,
Zaccardi F, Scavone G, Previtero M, Massaro G, Cialdella P, Cardillo
MT, Pitocco D, Ghirlanda G, Crea F. Adaptive immunity, inflammation,
and cardiovascular complications in type 1 and type 2 diabetes mellitus. J
Diabetes Res 2013: 184258, 2013.
44. Pflueger M, Seppanen-Laakso T, Suortti T, Hyotylainen T, Achen-
bach P, Bonifacio E, Oresic M, Ziegler AG. Age- and islet autoimmu-
nity-associated differences in amino acid and lipid metabolites in children
at risk for type 1 diabetes. Diabetes 60: 2740 –2747, 2011.
45. Phipson B, Smyth GK. Permutation P-values should never be zero:
calculating exact P-values when permutations are randomly drawn. Stat
Applic Genet Mol Biol 9: Article 39, 2010.
46. Reusch JE. Diabetes, microvascular complications, and cardiovascular
complications: what is it about glucose? J Clin Invest 112: 986 –988, 2003.
47. Scholz M, Fiehn O. SetupX–a public study design database for metabo-
lomic projects. Pac Symp Biocomput 169 –180, 2007.
48. Schuchardt JP, Schmidt S, Kressel G, Dong H, Willenberg I, Ham-
mock BD, Hahn A, Schebb NH. Comparison of free serum oxylipin
concentrations in hyper- vs. normolipidemic men. Prostaglandins Leukot
Essent Fatty Acids 89: 19 –29, 2013.
49. Shannon P, Markiel A, Ozier O, Baliga NS, Wang JT, Ramage D,
Amin N, Schwikowski B, Ideker T. Cytoscape: a software environment
for integrated models of biomolecular interaction networks. Genome Res
13: 2498 –2504, 2003.
50. Soga T, Ohishi T, Matsui T, Saito T, Matsumoto M, Takasaki J,
Matsumoto S, Kamohara M, Hiyama H, Yoshida S, Momose K, Ueda
Y, Matsushime H, Kobori M, Furuichi K. Lysophosphatidylcholine
enhances glucose-dependent insulin secretion via an orphan G-protein-
coupled receptor. Biochem Biophys Res Commun 326: 744 –751, 2005.
51. Spratlin JL, Serkova NJ, Eckhardt SG. Clinical applications of metabo-
lomics in oncology: a review. Clin Cancer Res 15: 431–440, 2009.
52. Svensson OKT, MacGregor JF. An investigation of orthogonal signal
correction algorithms and their characteristics. J Chemom 16: 176 –188,
2002.
53. Sysi-Aho M, Ermolov A, Gopalacharyulu PV, Tripathi A, Seppanen-
Laakso T, Maukonen J, Mattila I, Ruohonen ST, Vahatalo L, Yetu-
kuri L, Harkonen T, Lindfors E, Nikkila J, Ilonen J, Simell O, Saarela
M, Knip M, Kaski S, Savontaus E, Oresic M. Metabolic regulation in
progression to autoimmune diabetes. PLoS Comput Biol 7: e1002257,
2011.
54. Thomas M, Hughes RE. A relationship between ascorbic acid and
threonic acid in guinea-pigs. Food Chem Toxicol 21: 449 –452, 1983.
55. Tran PO, Gleason CE, Robertson RP. Inhibition of interleukin-1-
induced COX-2 and EP3 gene expression by sodium salicylate enhances
pancreatic islet -cell function. Diabetes 51: 1772–1778, 2002.
56. Weckwerth W, Wenzel K, Fiehn O. Process for the integrated extraction,
identification and quantification of metabolites, proteins and RNA to
reveal their co-regulation in biochemical networks. Proteomics 4: 78–83,
2004.
57. Westerberg DP. Diabetic ketoacidosis: evaluation and treatment. Am Fam
Physician 87: 337–346, 2013.
58. Wiklund S, Johansson E, Sjöström L, Mellerowicz EJ, Edlund U,
Shockcor JP, Gottfries J, Moritz T, Trygg J. Visualization of GC/TOF-
MS-based metabolomics data for identification of biochemically interest-
ing compounds using OPLS class models. Anal Chem 80: 115–122, 2008.
59. Workeneh B, Bajaj M. The regulation of muscle protein turnover in
diabetes. Int J Biochem Cell Biol 45: 2239 –2244, 2013.
60. Yang J, Schmelzer K, Georgi K, Hammock BD. Quantitative pro-
filing method for oxylipin metabolome by liquid chromatography
electrospray ionization tandem mass spectrometry. Anal Chem 81:
8085–8093, 2009.
61. Yea K, Kim J, Yoon JH, Kwon T, Kim JH, Lee BD, Lee HJ, Lee SJ,
Kim JI, Lee TG, Baek MC, Park HS, Park KS, Ohba M, Suh PG, Ryu
SH. Lysophosphatidylcholine activates adipocyte glucose uptake and
lowers blood glucose levels in murine models of diabetes. J Biol Chem
284: 33833–33840, 2009.
E989SYSTEMIC ALTERATIONS IN THE METABOLOME OF DIABETIC MICE
AJP-Endocrinol Metab doi:10.1152/ajpendo.00019.2015 www.ajpendo.org
Downloaded from journals.physiology.org/journal/ajpendo (193.160.077.067) on January 9, 2021.
... In Aedes, other blood components, such as glucose, have also been shown to modulate mosquito vector capacity through AKT/TOR [23]. However, artificial single-point alterations of blood composition rarely reflect the blood's physiological modulations where several components are found altered because of metabolic syndromes and altered physiological states [24,25]. In that sense, the impact of metabolic syndromes and nutritional imbalance, such as those derived from Western-type diets on mosquito biology, remains mostly unknown. ...
Article
Full-text available
Background The high prevalence of metabolic syndrome in low- and middle-income countries is linked to an increase in Western diet consumption, characterized by a high intake of processed foods, which impacts the levels of blood sugar and lipids, hormones, and cytokines. Hematophagous insect vectors, such as the yellow fever mosquito Aedes aegypti, rely on blood meals for reproduction and development and are therefore exposed to the components of blood plasma. However, the impact of the alteration of blood composition due to malnutrition and metabolic conditions on mosquito biology remains understudied. Methods In this study, we investigated the impact of whole-blood alterations resulting from a Western-type diet on the biology of Ae. aegypti. We kept C57Bl6/J mice on a high-fat, high-sucrose (HFHS) diet for 20 weeks and followed biological parameters, including plasma insulin and lipid levels, insulin tolerance, and weight gain, to validate the development of metabolic syndrome. We further allowed Ae. aegypti mosquitoes to feed on mice and tracked how altered host blood composition modulated parameters of vector capacity. Results Our findings identified that HFHS-fed mice resulted in reduced mosquito longevity and increased fecundity upon mosquito feeding, which correlated with alteration in the gene expression profile of nutrient sensing and physiological and metabolic markers as studied up to several days after blood ingestion. Conclusions Our study provides new insights into the overall effect of alterations of blood components on mosquito biology and its implications for the transmission of infectious diseases in conditions where the frequency of Western diet-induced metabolic syndromes is becoming more frequent. These findings highlight the importance of addressing metabolic health to further understand the spread of mosquito-borne illnesses in endemic areas. Graphical Abstract
... Fatty acids derived from adipose tissue lipolysis are oxidized for providing energy primarily through mitochondrial β-oxidation. An increase in fatty acid oxidation induces excessive oxidative stress, and the observed increases in 2-hydroxybutyrate and uric acid reflected an imbalance between prooxidants and antioxidants (Ames et al. 1981;Gall et al. 2010;Fahrmann et al. 2015). Our previous study demonstrated a sustained state of oxidative stress in dairy goats during the peripartal period . ...
Article
Full-text available
Dairy goats experience metabolic stress during the peripartal period, and their ability to navigate this stage of lactation is related to the occurrence and development of metabolic diseases. Unlike dairy cows, there is a lack of comprehensive analysis of changes in the plasma profiles of peripartal dairy goats, particularly using high-throughput techniques. A subset of 9 clinically-healthy dairy goats were used from a cohort of 96 primiparous Guanzhong dairy goats (BCS, 2.75 ± 0.15). Blood samples were collected at seven time points around parturition (d 21, 14, 7 before parturition, the day of kidding, and d 7, 14, 21 postpartum), were analyzed using untargeted metabolomics and targeted lipidomics. The orthogonal partial least squares discriminant analysis model revealed a total of 31 differential metabolites including p-cresol sulfate, pyruvic acid, cholic acid, and oxoglutaric acid. The pathway enrichment analysis identified phenylalanine metabolism, aminoacyl-tRNA biosynthesis, and citrate cycle as the top three significantly-altered pathways. The Limma package identified a total of 123 differentially expressed lipids. Phosphatidylserine (PS), free fatty acids (FFA), and acylcarnitines (ACs) were significantly increased on the day of kidding, while diacylglycerols (DAG) and triacylglycerols (TAG) decreased. Ceramides (Cer) and lyso-phosphatidylinositols (LPI) were significantly increased during postpartum period, while PS, FFA, and ACs decreased postpartum and gradually returned to antepartum levels. Individual species of FFA and phosphatidylcholines (PC) were segregated based on the differences in the saturation and length of the carbon chain. Overall, this work generated the largest repository of the plasma lipidome and metabolome in dairy goats across the peripartal period, which contributed to our understanding of the multifaceted adaptations of transition dairy goats.
... Furthermore, there is evidence that low methionine levels are associated with increased oxidative stress. Significant T1D-dependent increases in circulating oxidation products and decreases in methionine and cysteine levels were detected, indicating increased oxidative stress [50]. Methionine is a component of the transsulfuration process, which leads to glutathione production, an important intracellular antioxidant. ...
Article
Full-text available
Glycemic variability (GV) in some patients with type 1 diabetes (T1D) remains heterogeneous despite comparable clinical indicators, and whether other factors are involved is yet unknown. Metabolites in the serum indicate a broad effect of GV on cellular metabolism and therefore are more likely to indicate metabolic dysregulation associated with T1D. To compare the metabolomic profiles between high GV (GV-H, coefficient of variation (CV) of glucose ≥ 36%) and low GV (GV-L, CV < 36%) groups and to identify potential GV biomarkers, metabolomics profiling was carried out on serum samples from 17 patients with high GV, 16 matched (for age, sex, body mass index (BMI), diabetes duration, insulin dose, glycated hemoglobin (HbA1c), fasting, and 2 h postprandial C-peptide) patients with low GV (exploratory set), and another 21 (GV-H/GV-L: 11/10) matched patients (validation set). Subsequently, 25 metabolites were significantly enriched in seven Kyoto Encyclopedia of Genes and Genomes (KEGG) pathways between the GV-H and GV-L groups in the exploratory set. Only the differences in spermidine, L-methionine, and trehalose remained significant after validation. The area under the curve of these three metabolites combined in distinguishing GV-H from GV-L was 0.952 and 0.918 in the exploratory and validation sets, respectively. L-methionine was significantly inversely related to HbA1c and glucose CV, while spermidine was significantly positively associated with glucose CV. Differences in trehalose were not as reliable as those in spermidine and L-methionine because of the relatively low amounts of trehalose and the inconsistent fold change sizes in the exploratory and validation sets. Our findings suggest that metabolomic disturbances may impact the GV of T1D. Additional in vitro and in vivo mechanistic studies are required to elucidate the relationship between spermidine and L-methionine levels and GV in T1D patients with different geographical and nutritional backgrounds.
... We illustrate our tool using a metabolomics dataset for mice with 100 metabolites from [28]. The data contains the metabolic profiles of 41 non-diabetic and 30 diabetic mice and has been recently analyzed in [29]. ...
Article
Full-text available
Background Differential correlation networks are increasingly used to delineate changes in interactions among biomolecules. They characterize differences between omics networks under two different conditions, and can be used to delineate mechanisms of disease initiation and progression. Results We present a new R package, CorDiffViz, that facilitates the estimation and visualization of differential correlation networks using multiple correlation measures and inference methods. The software is implemented in R, HTML and Javascript, and is available at https://github.com/sqyu/CorDiffViz. Visualization has been tested for the Chrome and Firefox web browsers. A demo is available at https://diffcornet.github.io/CorDiffViz/demo.html. Conclusions Our software offers considerable flexibility by allowing the user to interact with the visualization and choose from different estimation methods and visualizations. It also allows the user to easily toggle between correlation networks for samples under one condition and differential correlations between samples under two conditions. Moreover, the software facilitates integrative analysis of cross-correlation networks between two omics data sets.
... For this data, 112 KEGG signaling and metabolic pathways were analyzed. The final dataset is much smaller and measured metabolic profiles among 41 non-diabetic and 30 diabetic mice for 100 selected metabolites [25]. ...
Article
Full-text available
Existing software tools for topology-based pathway enrichment analysis are either computationally inefficient, have undesirable statistical power, or require expert knowledge to leverage the methods’ capabilities. To address these limitations, we have overhauled NetGSA, an existing topology-based method, to provide a computationally-efficient user-friendly tool that offers interactive visualization. Pathway enrichment analysis for thousands of genes can be performed in minutes on a personal computer without sacrificing statistical power. The new software also removes the need for expert knowledge by directly curating gene-gene interaction information from multiple external databases. Lastly, by utilizing the capabilities of Cytoscape, the new software also offers interactive and intuitive network visualization.
Article
Full-text available
Blood metabolomics profiling using mass spectrometry has emerged as a powerful approach for investigating non-cancer diseases and understanding their underlying metabolic alterations. Blood, as a readily accessible physiological fluid, contains a diverse repertoire of metabolites derived from various physiological systems. Mass spectrometry offers a universal and precise analytical platform for the comprehensive analysis of blood metabolites, encompassing proteins, lipids, peptides, glycans, and immunoglobulins. In this comprehensive review, we present an overview of the research landscape in mass spectrometry-based blood metabolomics profiling. While the field of metabolomics research is primarily focused on cancer, this review specifically highlights studies related to non-cancer diseases, aiming to bring attention to valuable research that often remains overshadowed. Employing natural language processing methods, we processed 507 articles to provide insights into the application of metabolomic studies for specific diseases and physiological systems. The review encompasses a wide range of non-cancer diseases, with emphasis on cardiovascular disease, reproductive disease, diabetes, inflammation, and immunodeficiency states. By analyzing blood samples, researchers gain valuable insights into the metabolic perturbations associated with these diseases, potentially leading to the identification of novel biomarkers and the development of personalized therapeutic approaches. Furthermore, we provide a comprehensive overview of various mass spectrometry approaches utilized in blood metabolomics research, including GC-MS, LC-MS, and others discussing their advantages and limitations. To enhance the scope, we propose including recent review articles supporting the applicability of GC×GC-MS for metabolomics-based studies. This addition will contribute to a more exhaustive understanding of the available analytical techniques. The Integration of mass spectrometry-based blood profiling into clinical practice holds promise for improving disease diagnosis, treatment monitoring, and patient outcomes. By unraveling the complex metabolic alterations associated with non-cancer diseases, researchers and healthcare professionals can pave the way for precision medicine and personalized therapeutic interventions. Continuous advancements in mass spectrometry technology and data analysis methods will further enhance the potential of blood metabolomics profiling in non-cancer diseases, facilitating its translation from the laboratory to routine clinical application.
Article
Full-text available
Computational analysis and interpretation of metabolomic profiling data remains a major challenge in translational research. Exploring metabolic biomarkers and dysregulated metabolic pathways associated with a patient phenotype could offer new opportunities for targeted therapeutic intervention. Metabolite clustering based on structural similarity has the potential to uncover common underpinnings of biological processes. To address this need, we have developed the MetChem package. MetChem is a quick and simple tool that allows to classify metabolites in structurally related modules, thus revealing their functional information. Availability MetChem is freely available from the R archive CRAN (http://cran.r-project.org). The software is distributed under the GNU General Public License (version 3 or later). Supplementary information Supplementary data are available at Bioinformatics online.
Article
Full-text available
Disturbance of circulating metabolites and disorders of the gut microbiota are involved in the progression of diabetic kidney disease (DKD). However, there is limited research on the relationship between serum metabolites and gut microbiota, and their involvement in DKD. In this study, using an experimental DKD rat model induced by combining streptozotocin injection and unilateral nephrectomy, we employed untargeted metabolomics and 16S rRNA gene sequencing to explore the relationship between the metabolic profile and the structure and function of gut microbiota. Striking alterations took place in 140 serum metabolites, as well as in the composition and function of rat gut microbiota. These changes were mainly associated with carbohydrate, lipid, and amino acid metabolism. In these pathways, isomaltose, D-mannose, galactonic acid, citramalic acid, and prostaglandin B2 were significantly upregulated. 3-(2-Hydroxyethyl)indole, 3-methylindole, and indoleacrylic acid were downregulated and were the critical metabolites in the DKD model. Furthermore, the levels of these three indoles were restored after treatment with the traditional Chinese herbal medicine Tangshen Formula. At the genera level, g_Eubacterium_nodatum_group, g_Lactobacillus, and g_Faecalibaculum were most involved in metabolic disorders in the progression of DKD. Notably, the circulating lipid metabolites had a strong relationship with DKD-related parameters and were especially negatively related to the mesangial matrix area. Serum lipid indices (TG and TC) and UACR were directly associated with certain microbial genera. In conclusion, the present research verified the anomalous circulating metabolites and gut microbiota in DKD progression. We also identified the potential metabolic and microbial targets for the treatment of DKD.
Article
Full-text available
Non-obese diabetic (NOD) mice are a widely-used model of type 1 diabetes (T1D). However, not all animals develop overt diabetes. This study examined the circulating metabolomic profiles of NOD mice progressing or not progressing to T1D. Total beta-cell mass was quantified in the intact pancreas using transgenic NOD mice expressing green fluorescent protein under the control of mouse insulin I promoter. While both progressor and non-progressor animals displayed lymphocyte infiltration and endoplasmic reticulum stress in the pancreas tissue, overt T1D did not develop until animals lost ~70 % of the total beta-cell mass. Gas chromatography time of flight mass spectrometry was used to measure >470 circulating metabolites in male and female progressor and non-progressor animals (n = 76) across a wide range of ages (neonates to >40-week). Statistical and multivariate analyses were used to identify age and sex independent metabolic markers which best differentiated progressor and non-progressor animals’ metabolic profiles. Key T1D-associated perturbations were related with: (1) increased plasma glucose and reduced 1,5-anhydroglucitol markers of glycemic control; (2) increased allantoin, gluconic acid and nitric acid-derived saccharic acid markers of oxidative stress; (3) reduced lysine, an insulin secretagogue; (4) increased branched-chain amino acids, isoleucine and valine; (5) reduced unsaturated fatty acids including arachidonic acid; and (6) perturbations in urea cycle intermediates suggesting increased arginine-dependent NO synthesis. Together these findings highlight the strength of the unique approach of comparing progressor and non-progressor NOD mice to identify metabolic perturbations involved in T1D progression.
Article
Full-text available
Current tandem mass spectral libraries for lipid annotations in metabolomics are limited in size and diversity. We provide a freely available computer-generated tandem mass spectral library of 212,516 spectra covering 119,200 compounds from 26 lipid compound classes, including phospholipids, glycerolipids, bacterial lipoglycans and plant glycolipids. We show platform independence by using tandem mass spectra from 40 different mass spectrometer types including low-resolution and high-resolution instruments.
Article
Full-text available
Diabetes mellitus (DM) is a pandemics that affects more than 170 million people worldwide, associated with increased mortality and morbidity due to coronary artery disease (CAD). In type 1 (T1) DM, the main pathogenic mechanism seems to be the destruction of pancreatic β -cells mediated by autoreactive T-cells resulting in chronic insulitis, while in type 2 (T2) DM primary insulin resistance, rather than defective insulin production due to β -cell destruction, seems to be the triggering alteration. In our study, we investigated the role of systemic inflammation and T-cell subsets in T1- and T2DM and the possible mechanisms underlying the increased cardiovascular risk associated with these diseases.
Article
Full-text available
Oxylipins, the oxidation products of unsaturated fatty acids (FA), are potent endogenous mediators being involved in the regulation of various biological processes such as inflammation, pain and blood coagulation. Compared to oxylipins derived from arachidonic acid (AA) by cyclooxygenase action, i.e. prostanoides, only limited information is available about the endogenous levels of hydroxy-, epoxy- and dihydroxy-FA of linoleic acid (LA), AA, α-linolenic acid (ALA), eicosapentaenoic acid (EPA) and docosahexaenoic acid (DHA) in humans. Particularly, it is unknown how metabolic disorders affect endogenous oxylipin levels in humans. Therefore, in the present study we compared the serum concentrations of 44 oxylipins in 20 normolipidemic with 20 hyperlipidemic (total cholesterol >200mg/dl; LDL-C>130mg/dl; TG>150mg/dl) men (age 29-51y). The serum concentration varied strongly among subjects. For most hydroxy-, epoxy- and dihydroxy-FA the concentrations were comparable to those in plasma reported in earlier studies. Despite the significant change in blood lipid levels the hyperlipidemic group showed only minor differences in oxylipin levels. The hyperlipidemic subjects had a slightly higher serum concentration of 8,9-DiHETrE, 5-HEPE, 10,11-DiHDPE, and a lower concentration of 12,13-DiHOME, 12-HETE, 9,10-DiHODE, and 12,13-DiHODE compared to normolipidemic subjects. Overall the hydroxy-, epoxy- and dihydroxy-FA levels were not changed suggesting that mild combined hyperlipidemia has no apparent effect on the concentration of circulating oxylipins. By contrast, serum levels of several hydroxy-, epoxy-, and dihydroxy-FA are dependent on the individual status of the parent FA. Particularly, a strong correlation between the EPA content in the erythrocyte membrane and the serum concentration of EPA derived oxylipins was observed. Given that the synthesis of EPA from other n-3 FA in humans is low; this suggests that oxylipin levels can be directly influenced by the diet.
Article
Full-text available
Earlier studies show that children who later progress to type 1 diabetes (T1D) have decreased pre-autoimmune concentrations of multiple phospholipids as compared to non-progressors. It is still unclear whether these changes associate with development of β-cell autoimmunity or specifically with clinical T1D. Here we studied umbilical cord serum lipidome in newborn infants who later developed T1D (N=33), infants who developed three or four (N=31), two (N=31), or one (N=48) islet autoantibody during the follow-up, and controls (N=143) matched for gender, HLA-DQB1 genotype, city and period of birth. The analyses of serum molecular lipids were performed using the established lipidomics platform based on Ultra Performance Liquid Chromatography(TM) coupled to mass spectrometry (UPLC-MS). We found that T1D progressors are characterized by a distinct cord blood lipidomic profile which includes reduced major choline-containing phospholipids including sphingomyelins and phosphatidylcholines. A molecular signature was developed comprising seven lipids which predicted high risk for progression to T1D, with an odds ratio of 5.94 (95% CI, 1.07 - 17.50). Reduction in choline-containing phospholipids in cord blood is therefore specifically associated with progression to T1D but not with development of β-cell autoimmunity in general.
Article
The common approach to the multiplicity problem calls for controlling the familywise error rate (FWER). This approach, though, has faults, and we point out a few. A different approach to problems of multiple significance testing is presented. It calls for controlling the expected proportion of falsely rejected hypotheses — the false discovery rate. This error rate is equivalent to the FWER when all hypotheses are true but is smaller otherwise. Therefore, in problems where the control of the false discovery rate rather than that of the FWER is desired, there is potential for a gain in power. A simple sequential Bonferronitype procedure is proved to control the false discovery rate for independent test statistics, and a simulation study shows that the gain in power is substantial. The use of the new procedure and the appropriateness of the criterion are illustrated with examples.
Article
Diabetes cannot be considered simply a disease of glucose dysregulation; it is a chronic inflammatory disease that affects nearly every biological process, including protein metabolism. Diabetes is associated with disturbances in muscle protein metabolism that results in decreased muscle mass and in some cases, loss in the activities of daily living, decreased productivity and diminished quality of life. Alteration in protein metabolism and its effect on muscle mass and function is one of the most challenging and least understood issues in the management of diabetes. Central among insulin action in muscle is suppression of protein degradation pathways and up-regulation of anabolic pathways. In type 1 diabetes, muscle wasting essentially results from insulin deficiency and this induces of genes involved in the ubiquitin proteasome pathway. On the other hand, the chief defect that leads to muscle atrophy in type 2 diabetes is decreased insulin responsiveness primarily in muscle. Decreased insulin responsiveness has been attributed to defects in the insulin signaling pathways secondary to inflammation (e.g., NF-κB activation and elevated levels of TNF-α, IL-1 and IL-6), metabolic acidosis, increased circulating free fatty acids and glucotoxicity. Furthermore, emerging pathways, such as myostatin/activin A system are beginning to be uncovered. We conclude with a discussion of possible interventions to slow, mitigate or reverse muscle wasting associated with diabetes.
Article
PubChem is an open repository for experimental data identifying the biological activities of small molecules. PubChem contents include more than: 1000 bioassays, 28 million bioassay test outcomes, 40 million substance contributed descriptions, and 19 million unique compound structures contributed from over 70 depositing organizations. PubChem provides a significant, publicly accessible platform for mining the biological information of small molecules.
Article
Diabetic ketoacidosis is characterized by a serum glucose level greater than 250 mg per dL, a pH less than 7.3, a serum bicarbonate level less than 18 mEq per L, an elevated serum ketone level, and dehydration. Insulin deficiency is the main precipitating factor. Diabetic ketoacidosis can occur in persons of all ages, with 14 percent of cases occurring in persons older than 70 years, 23 percent in persons 51 to 70 years of age, 27 percent in persons 30 to 50 years of age, and 36 percent in persons younger than 30 years. The case fatality rate is 1 to 5 percent. About one-third of all cases are in persons without a history of diabetes mellitus. Common symptoms include polyuria with polydipsia (98 percent), weight loss (81 percent), fatigue (62 percent), dyspnea (57 percent), vomiting (46 percent), preceding febrile illness (40 percent), abdominal pain (32 percent), and polyphagia (23 percent). Measurement of A1C, blood urea nitrogen, creatinine, serum glucose, electrolytes, pH, and serum ketones; complete blood count; urinalysis; electrocardiography; and calculation of anion gap and osmolar gap can differentiate diabetic ketoacidosis from hyperosmolar hyperglycemic state, gastroenteritis, starvation ketosis, and other metabolic syndromes, and can assist in diagnosing comorbid conditions. Appropriate treatment includes administering intravenous fluids and insulin, and monitoring glucose and electrolyte levels. Cerebral edema is a rare but severe complication that occurs predominantly in children. Physicians should recognize the signs of diabetic ketoacidosis for prompt diagnosis, and identify early symptoms to prevent it. Patient education should include information on how to adjust insulin during times of illness and how to monitor glucose and ketone levels, as well as information on the importance of medication compliance.