SlideShare a Scribd company logo
1 of 8
Download to read offline
Evaluation of 1
H NMR Metabolic Profiling Using Biofluid Mixture
Design
Toby J. Athersuch,*,†,‡
Shahid Malik,§
Aalim Weljie,∥,⊥
Jack Newton,∥
and Hector C. Keun†
†
Computational and Systems Medicine, Department of Surgery and Cancer, Faculty of Medicine, Sir Alexander Fleming Building,
Imperial College London, South Kensington, SW7 2AZ, U.K.
‡
MRC-PHE Centre for Environment and Health, Department of Epidemiology and Biostatistics, School of Public Health, Faculty of
Medicine, Imperial College London, Norfolk Place, London, W2 1PG, U.K.
§
Chenomx Inc., Suite 800, 10050 112 Street, Edmonton, Alberta, T5K 2J1, Canada
∥
Department of Biological Sciences, Bio-NMR Center, University of Calgary, Calgary, Alberta, T2N 1N4, Canada
⊥
Institute for Translational Medicine and Therapeutics, Perelman School of Medicine, University of Pennsylvania, 10-113
Translational Research Center, 3400 Civic Center Boulevard, Building 421, Philadelphia, Pennsylvania 19104, United States
*S Supporting Information
ABSTRACT: A strategy for evaluating the performance of quantitative
spectral analysis tools in conditions that better approximate background
variation in a metabonomics experiment is presented. Three different urine
samples were mixed in known proportions according to a {3, 3} simplex
lattice experimental design and analyzed in triplicate by 1D 1
H NMR spec-
troscopy. Fifty-four urinary metabolites were subsequently quantified from
the sample spectra using two methods common in metabolic profiling
studies: (1) targeted spectral fitting and (2) targeted spectral integration.
Multivariate analysis using partial least-squares (PLS) regression showed
the latent structure of the spectral set recapitulated the experimental mix-
ture design. The goodness-of-prediction statistic (Q2
) of each metabolite
variable in a PLS model was calculated as a metric for the reliability of
measurement, across the sample compositional space. Several metabolites
were observed to have low Q2
values, largely as a consequence of their spectral resonances having low s/n or strong overlap with
other sample components. This strategy has the potential to allow evaluation of spectral features obtained from metabolic
profiling platforms in the context of the compositional background found in real biological sample sets, which may be subject to
considerable variation. We suggest that it be incorporated into metabolic profiling studies to improve the estimation of matrix
effects that confound accurate metabolite measurement. This novel method provides a rational basis for exploiting information
from several samples in an efficient manner and avoids the use of multiple spike-in authentic standards, which may be difficult
to obtain.
■ INTRODUCTION
Metabolic Profiling. Metabolic profiling (metabonomics/
metabolomics) has become a key platform in systems biology;
the application of spectroscopy or spectrometry to biological
samples provides a multicomponent metabolic phenotype that
reflects a large number of interacting upstream processes including
gene expression, cellular status, and organism function.1−3
Metabolite profiles are currently being used in a wide variety of
contexts including “bench-to-bedside” translational medicine,4,5
real-time profiling for enhanced biomarker-based decision tools
for clinicians/surgeons,6
and high-throughput metabolic
phenotyping in large-scale molecular epidemiological studies
aimed at understanding chronic disease risk and etiology.7,8
Nuclear magnetic resonance (NMR) spectroscopy is a core
analytical platform used to characterize biological matrices in
metabolic profiling studies as it provides quantitative spectra
that capture concentration information on multiple metabolites
simultaneously, in a highly robust and reproducible manner.9
A large number of studies have used NMR as the primary
analytical tool for exploring a wide range of research
questions, including those in drug toxicity testing,10
efficacy
assessment,11
and population studies.12,13
Metabolic profiles
obtained by NMR spectroscopy typically contain hundreds
or thousands of real spectral features, along with confound-
ing signals (including noise), that span a wide dynamic range
(>8 orders of magnitude). Accurate quantification of metab-
olites using biofluid spectra obtained by NMR spectroscopy
can be relatively difficult, as a consequence of signal overlap,
resulting from insufficient spectral resolution, that may reduce
accuracy of integral measurements made as an estimator of
Received: February 10, 2013
Accepted: June 3, 2013
Published: June 3, 2013
Article
pubs.acs.org/ac
© 2013 American Chemical Society 6674 dx.doi.org/10.1021/ac400449f | Anal. Chem. 2013, 85, 6674−6681
concentration. In addition to peak-alignment based methods
of spectral deconvolution,14
spectral fitting approaches, that
use individual template spectra matching for each metabolite, are
commonly used in an attempt to reduce the influence of signal
overlap on quantification.15
By virtue of the spectral matching
process (that typically uses multiple peak fits in the spectrum to
provide a best estimate of concentration), spectral fitting has the
advantage oversimple spectral integration in that it is inherently
less affected by background variation arising from the sample
matrix, and from spectral artifacts such as the residual water peak
in aqueous sample spectra. However, spectral fitting is more
time-consuming and may be prone to user error or subjectivity.
A common approach used to assess the quantification
accuracy in biofluid spectra is the use of traditional “spike-in”
experiments, whereby authentic standards are added to the
sample in known concentration. However, these experiments
are conducted typically in the context of an invariant back-
ground, which is often not representative of the “real world”
scenario where baseline signals from different samples vary due
to numerous matrix effects.
Additional measures of the spectral quality and reliability of
individual measurements made in metabolic profiling studies,
that characterize performance in real sample sets, are therefore
of potential utility to the metabolic profiling community.
Mixture Design. Mixture design experiments are routinely
used for the selection of optimal criteria for production pro-
cesses, formulation, and more generally in the characterization
of relationships between response and system composition.
There are numerous designs that can be used, depending on
the constraints placed on the mixture components; a simplex-
lattice design reflects one of the simplest designs, and is
described as follows: “A {q, m} simplex-lattice design for q com-
ponents consists of points defined by the following coordinate
settings: the proportions assumed by each component take the
m+1 equally spaced values from 0 to 1
= =x m m i q0, 1/ , 2/ , ..., 1 for 1, 2, ...,i
...and all possible combinations (mixtures) of the proportions
from this equation are used.”16
The proportions must sum to
unity. For example, a {3,3} simplex lattice design represents
three components (q = 3), each of which have four (m + 1 = 4),
equally spaced, different possible levels (0, 1/3, 2/3, 1), and
therefore will have ten possible mixture combinations.
We propose that mixing different biofluid samples in known
proportions according to a mixture design (such as a simplex
lattice) will produce a sample set that enables metabolite behavior
across the sample compositional space to be characterized by
regression of the design against the metabolite response. In an
ideal situation, the observed response of an individual metabolite
will exactly follow the mixture design, and a perfect fit will be
achieved. In reality, matrix effects and confounding signal overlap
may reduce the accuracy of metabolite responses and reduce the
correspondence with the mixture design. Thus, this approach
allows the reproducibility of individual metabolites to be assessed,
and those that are adversely affected by matrix effects or signal
overlap to be identified.
Here we have applied this strategy of mixing intact biofluids,
according to a predetermined experimental mixture design, to
compare the performance of two commonly used metabolite
quantification methods in the context of “real world” 1
H NMR
metabonomic analysis. The potential benefits of incorporating
a designed mixture component in metabonomic analyses, as a
method of assessing the accuracy of metabolite quantification,
are discussed. We suggest that this strategy may have general
benefits and applicability in metabolic profiling studies.
■ MATERIALS AND METHODS
Chemicals. D2O was obtained from Goss Scientific
(Nantwich, U.K.). All other reagents were of analytical grade
and obtained from SigmaAldrich (Poole, U.K.).
Experimental Design. A schematic of the experimental
design is shown in (Figure 1), with details of discussed in turn
below.
Sample Collection and Preparation. Urine samples were
obtained from an existing large-scale toxicological study
resource.10,17,18
Sprague−Dawley rats (n = 7) were individually
housed in standard metabolism cages (21 ± 3 ◦C, relative
humidity 55 ± 15%) and acclimatized for six days prior to the
start of the study (t = 0 h). A standard diet (Purina chow 5002)
and fresh water (acidified to pH 2.5 using HCl to prevent
microbial growth) was available to each animal ad libitum.
Urine samples used in the current study were collected during
Figure 1. Schematic showing the overall approach described. Three
different urines were mixed in known proportions according to a
mixture design (1, 2). Concentrations of metabolites were determined
by 1
H NMR spectroscopy (3). Spectral fitting and spectral integration
were both used for quantification (4). The mixture design data (Y block)
were used in a PLS regression against the metabolite concentration data
(X block) to generate model metrics (5).
Table 1. Sample Composition for Designed Biofluid
Mixtures Used in This Study Following a {3,3} Simplex-
Lattice Mixture Design
volume (μL)
rat urine human urine
sample number 0−8 h 8−24 h spot sample sodium phospate buffer
1 300 0 0 300
2 0 300 0 300
3 0 0 300 300
4 200 100 0 300
5 200 0 100 300
6 100 200 0 300
7 0 200 100 300
8 100 0 200 300
9 0 100 200 300
10 100 100 100 300
Analytical Chemistry Article
dx.doi.org/10.1021/ac400449f | Anal. Chem. 2013, 85, 6674−66816675
Table 2. 1
H NMR Spectral Regions Used for Integration, Estimated Metabolite Concentrations Derived from Spectral Fitting
Using Chenomx Software, and PLS Model Metricsa
spectral region (ppm) estimated concentration (μM) Q2
ID metabolite high low width min max mean STDEV A B C D
1 1-methylnicotinamide 4.47 4.46 0.01 11 182 115 55 0.99 0.20 0.91 −0.10
2 1,3-dimethyluratec,f
3.29 3.28 0.01 24 312 149 93 0.98 0.99 0.90 0.73
3 1,6-anhydro-β-D-glucosef
5.45 5.44 0.01 24 70 44 15 0.89 0.01 0.25 −0.11
4 2-hydroxyisobutyratea,c,f
1.35 1.34 0.01 0 36 16 15 0.42 0.98 0.39 0.61
5 2-oxoglutarate 3.02 2.98 0.04 104 4759 2737 1491 0.99 0.99 0.89 0.84
6 2-oxoisocaproateb,c,f
0.94 0.92 0.02 0 76 41 23 0.94 0.99 0.82 0.83
7 3-hydroxyisovalerate 1.26 1.25 0.01 20 32 26 4 0.80 0.98 0.85 0.63
8 3-indoxylsulfate 7.51 7.48 0.03 75 288 175 64 0.98 0.96 −0.01 0.65
9 acetate 1.91 1.91 0.01 10 107 73 32 0.98 0.99 0.89 0.72
10 alanine 1.48 1.46 0.02 65 106 92 14 0.94 0.98 0.80 0.56
11 allantoinc
5.42 5.34 0.08 19 3929 2530 1162 0.73 0.96 −0.06 0.70
12 arginine 1.95 1.87 0.08 129 184 153 17 0.55 0.98 0.66 0.25
13 betainec
3.90 3.88 0.02 29 462 233 133 0.92 0.99 0.69 0.84
14 choline 3.18 3.18 0.01 15 50 30 13 0.88 0.95 0.87 0.80
15 cis-aconitate 3.11 3.09 0.02 161 459 321 89 0.51 0.99 0.66 0.23
16 citrate 2.56 2.50 0.06 1775 14420 8951 4025 0.94 0.99 0.88 0.85
17 creatinine 3.05 3.02 0.02 1873 5154 3388 1028 0.98 0.98 0.85 0.79
18 dimethylamine 2.72 2.70 0.02 125 452 285 100 0.98 0.98 −0.05 0.59
19 ethanolc,d,f
1.19 1.16 0.03 9 36 20 9 0.97 0.98 −0.07 0.45
20 ethanolamine 3.15 3.12 0.03 68 204 144 41 0.84 0.98 0.83 0.66
21 formate 8.45 8.44 0.01 30 321 184 88 0.83 0.98 0.76 0.87
22 fucosef
1.25 1.23 0.02 53 197 135 42 0.97 0.97 0.68 0.64
23 fumarateb,d
6.51 6.51 0.01 0 42 26 13 0.96 0.97 0.71 0.80
24 glucosed,e
4.65 4.63 0.03 198 641 385 140 0.90 −0.04 0.01 −0.01
25 glycine 3.56 3.55 0.01 149 458 270 95 0.95 0.96 0.81 0.74
26 glycolated
3.94 3.93 0.01 104 198 152 29 0.97 0.98 0.82 0.61
27 guanidinoacetated
3.79 3.78 0.01 96 332 190 79 0.87 0.99 −0.07 0.36
28 hippurate 7.85 7.81 0.04 256 3696 2158 1060 0.81 0.99 0.88 0.86
29 isoleucine 1.01 0.99 0.02 9 15 12 2 0.99 0.97 0.73 0.56
30 lactatec
1.33 1.31 0.03 65 106 81 15 0.62 0.98 0.75 0.45
31 leucined
0.97 0.94 0.03 11 16 14 1 0.87 0.98 0.89 −0.11
32 malonated,f
3.12 3.11 0.01 8 43 26 11 0.25 0.98 0.03 0.43
33 methanol 3.35 3.34 0.01 11 87 50 27 0.70 0.92 0.49 0.78
34 methylmalonate 1.24 1.22 0.02 58 156 95 27 0.56 0.94 0.72 0.77
35 methylsuccinatef
1.08 1.06 0.02 14 52 35 12 0.81 0.97 0.03 0.46
36 N,N-dimethylformamidea,f
2.86 2.85 0.01 14 94 48 25 0.99 0.98 0.74 0.44
37 N,N-dimethylglycinef
2.92 2.91 0.01 0 59 11 23 0.75 0.94 0.01 0.76
38 N-acetylglycinee
2.03 2.03 0.01 27 266 131 78 −0.15 0.99 0.93 0.90
39 O-phosphocholinee,f
3.19 3.18 0.01 7 42 25 11 0.96 0.98 0.82 0.65
40 oxaloacetatea,c,d
3.67 3.66 0.01 0 308 76 125 0.95 0.98 −0.21 0.29
41 phenylacetylglycinef
7.43 7.39 0.04 103 235 160 41 0.52 0.98 0.34 0.06
42 pyruvated
2.37 2.36 0.01 11 58 32 15 0.87 0.99 0.25 0.57
43 succinate 2.40 2.39 0.02 17 1100 537 335 0.77 0.99 0.90 0.89
44 taurine 3.43 3.39 0.04 466 1843 1277 448 0.99 0.99 0.85 0.24
45 threoninec
1.33 1.31 0.02 9 62 43 19 0.98 0.98 0.71 0.48
46 trans-aconitate 6.59 6.56 0.03 15 1834 845 566 0.96 0.99 0.91 0.90
47 trigonelline 4.43 4.42 0.01 54 294 168 75 0.88 0.72 0.76 −0.09
48 trimethylamine N-oxidec
3.26 3.25 0.01 124 481 280 116 1.00 0.99 0.30 0.64
49 tryptophana,d
7.76 7.73 0.03 0 41 24 17 0.98 0.92 0.89 0.62
50 tyrosine 6.90 6.87 0.02 34 48 39 5 0.96 0.94 0.68 −0.01
51 uracilb,e
7.54 7.52 0.03 0 123 66 36 0.14 0.98 0.67 0.82
52 ureae
5.95 5.61 0.34 90366 207739 152730 44253 0.70 0.91 0.45 0.65
53 valine 0.99 0.97 0.03 17 24 20 3 0.91 0.97 0.69 0.54
54 xylose 4.59 4.56 0.03 102 561 287 151 0.66 0.00 0.85 0.11
a
The Q2
statistic is given for each of four models: (A) spectral fitting data set normalized to TSP, (B) spectral integration data set normalized to
TSP, (C) spectral fitting data set normalized using PQN, and (D) spectral integration data set normalized using PQN. a
Metabolite present in only
one of the component samples. b
Metabolite absent in one of the component samples. c
Overlapped signal. d
Low s/n. e
Spectral artifact present in
region. f
Tentative assignment.
Analytical Chemistry Article
dx.doi.org/10.1021/ac400449f | Anal. Chem. 2013, 85, 6674−66816676
two periods (0−8 h, 8−24 h) from control animals. Urine
voided by the animals was collected in the metabolism cage
into a container cooled by dry ice. Samples were subsequently
stored at −40 °C. Each sample underwent two freeze−thaw
cycles before use in this work as a consequence of realiquoting.
Further study information has previously been published.17
Additionally, a spot urine sample (5 mL) was obtained from a
healthy human volunteer, according to established protocols,
including filtration to remove cellular material (0.2 μm Minisart
16534K, Sartorius, Germany), and immediate storage at −40 °C
until required for analysis. Urine samples were prepared follow-
ing established protocols for NMR metabolome analysis.19
Urine samples were defrosted, vortex mixed (30 s, RT), and
centrifuged (16000 g, RT, 10 min) to remove particulate matter.
To provide sufficient total sample, for each collection period,
450 μL of each rat urine sample was pooled (total volume 3150 μL
per collection period). The three urines (two pooled rat urine,
one human spot urine) were mixed according to a {3,3} simplex-
lattice mixture design (Table 1), with each mixed sample having
a volume of 300 μL. These mixed samples were then buffered by
the addition of 300 μL sodium phosphate buffer (pH 7.4, 0.2 M,
80:20 H2O:D2O (v/v)) containing sodium 3-(trimethylsilyl)-
[2,2,3,3-2
H4]propionate (TSP, 1 mM). Samples were vortex
mixed (30 s, RT), and a 550 μL aliquot transferred to a 96-well
autosampler plate. The mixed samples were prepared in triplicate
from the pooled rat samples and the human spot urine. The
preparation order was randomized.
NMR Spectral Acquisition and Processing. 1
H NMR
spectra were acquired on a Bruker AVANCE DRX600 NMR
spectrometer (Bruker Biospin, Rheinstetten, Germany) operat-
ing at 14.1 T (600.29 MHz 1
H NMR frequency) using a PH FI
TXI 600SB 5 mm probe maintained at 300 K. Samples were
introduced to the probe using a BEST flow-injection system
(Bruker) in a randomized order. Gradient shimming was used
immediately prior to spectral acquisition to ensure high field
homogeneity. Spectral acquisition was made using standard a
standard 1D pulse sequence (RD-90°-t1-90°-tm-90°-AQ).20
The
t1 delay and the mixing time (tm) were set to 3 μs and 100 ms
respectively. All spectra were collected as the sum of 128 free
induction decays (FIDs) were collected into 32K complex
data points. The spectral width of 12019.23 Hz (20 ppm)
giving the FID a native resolution of 0.366 Hz/pt, and an
acquisition time (AQ) of 1.36 s. A 2 s relaxation delay (RD)
was used between pulses. A presaturation pulse was applied
to the water resonance (δH = 4.7 ppm) during RD and tm.
Processing of the raw NMR data for analysis using a targeted
integration approach was carried out using XWINNMR
software (Bruker Biospin, Rheinstetten, Germany), with each
FID being multiplied by an exponential weighting func-
tion equivalent to a line broadening of 1 Hz prior to Fourier
transformation. Resulting frequency-domain spectra were
referenced to TSP (δH = 0.00 ppm) and interpolated from
32K to ∼42K data points using a cubic spline function
to regularize the abscissa and improve calibration accuracy
Figure 2. 1
H NMR spectra of urine samples used in this study: (A) Pooled rat urine 0−8 h collection (n = 7), (B) pooled rat urine 8−24 h collection
(n = 7), and (C) human urine spot collection. Spectra were acquired at an observation frequency of 600 MHz using a standard 1D pulse sequence
with water presaturation.
Analytical Chemistry Article
dx.doi.org/10.1021/ac400449f | Anal. Chem. 2013, 85, 6674−66816677
(final resolution 0.29 Hz/pt) prior to analysis using in-house
scripts running in the Matlab (The Mathworks, Natick) computing
environment.
Metabolite Quantification. Fifty-four metabolites were
quantified using both the spectral fitting approach and a
targeted spectral integration approach. Spectral fitting was
performed using Chenomx NMR Suite 4.6 (Chenomx Inc.,
Edmonton, Canada); reference spectra from the Chenomx
600 MHz library were combined so as to best approximate each
acquired urine spectrum, and the relative concentrations of
each metabolite present determined by reference to the internal
TSP standard15
(Table 2). For the targeted spectral integration
approach, spectral regions were defined for each of the
metabolites of interest (Table 2), with a width sufficient to
encapsulate the majority of the peak across the entire set of
spectra (determined manually by spectral overlay). The integral
area of these regions was calculated using an in-house routine
in Matlab. Probabilistic quotient normalization21
(PQN) was
applied to remove variation originating from intersample dif-
ferences in urinary dilution. Chemometric analysis of meta-
bolite concentration data was completed using Simca P+12
(Umetrics, Umea, Sweden). Principal component analysis
(PCA, using metabolite concentration data) was conducted.
Partial least-squares regression (PLS, using metabolite concen-
tration data and the experimental design matrix) allowed
goodness-of-fit (R2
) and goodness-of-prediction (Q2
) estimates
to be made for each metabolite.22
■ RESULTS AND DISCUSSION
Quantification of Metabolites Using NMR. As detailed
in the Materials and Methods section, three different urines
were mixed according to a {3,3} simplex lattice design, and
analyzed in triplicate by 1
H NMR spectroscopy. Representative
spectra are shown in Figure 2. A total of 54 metabolites were
successfully quantified in these samples using both a targeted
spectral fitting approach (involving the fitting of individual
reference metabolite spectra to the spectra acquired for each
sample mixture), and a targeted spectral integration approach
(involving the integration of a representative spectral region).
Metabolite data are given in Table 2.
Comparison of the metabolite concentrations of metabolites
across the three samples containing only one of the three urine
components (i.e., the corners of the simplex lattice design) showed
considerable variation between the rat and human samples
(Supporting Information Figure S1). Some metabolites were
present or absent in only one of the samples (2-hydroxyisobutyrate,
2-oxoisocaproate, fumarate, N,N-dimethylglycine, oxaloacetate,
tryptophan). Other metabolites spanned up to 2 orders of
magnitude in their absolute concentration in these samples;
those with the greatest variation included 1-methylnicotina-
mide, 1,3-dimethylurate, betaine, and allantoin.
PCA/PLS Modeling and Effect of Normalization.
Principal component analysis (PCA) and partial least-squares
(PLS) regression are widely used multivariate analysis tools
based on latent variable methods.23,24
For each quantification
approach, the metabolite concentrations (X-matrix) were
modeled by PCA in an unsupervised manner, and also
modeled against the experimental mixture design (Y-
matrix) using PLS.
PCA of the data set before and after PQN was conducted
and showed differences in the variation captured by the two
methods of quantification (Supporting Information, Figure S2).
Prior to normalization, the largest variation in each data set was
attributable to the sample dilution. Upon normalization, the
next biggest variation in the spectral fitting data set was revealed
to be the mixture design, whereas in the spectral targeting data
set, it was related to nuisance variation, driven by outlying
samples, and attributable to inferior water suppression.
PLS analysis was used to assess the fit of the relative
metabolite variation to the mixture design. The component
scores for the PLS models (Figure 3) clearly showed the experi-
mental design as anticipated. Switching the X and Y matrices
of the PLS model also allowed a calculation of the goodness-
of-fit (R2
) and goodness-of-prediction (Q2
) values for each
metabolite against the {3,3} simplex-lattice design (Figure 3
and Table 2). It can be seen that a large proportion have high
Q2
values indicating that the latent structure of these data
follow that of the mixture proportions. It was recognized that in
these models, overall urinary dilution would have the effect of
artificially enhancing some of these values as a consequence of
introducing structure into the concentration profiles. PQN
concentrations were subsequently modeled and indicated that
once the global dilution factor was removed from the con-
centration data, several metabolites displayed greatly reduced
Q2
values. Removing the dominating dilution difference means
that the Q2
statistic reported for each metabolite gives a more
realistic representation of the fit to the design of the metabolite
response, in the presence of a variable background.
There are several reasons that explain low Q2
value including
(a) changes in chemical shift as a consequence of pH variation,
Figure 3. Partial least-squares regression analysis scores plots
indicating latent structure of the TSP normalized data (A) spectral
integration data set and (B) spectral fitting data set. It can be seen that
the samples recapitulate the {3,3} simplex-lattice mixture design in the
score space. Samples (triplicates) are colored according to proportions
of the three component urines (Table 1).
Analytical Chemistry Article
dx.doi.org/10.1021/ac400449f | Anal. Chem. 2013, 85, 6674−66816678
(b) the absence of a metabolite in one or more of the
component samples (in this case the rat and human urines),
(c) low s/n in the measurement, and (d) overlap of spectral
features. Where identified during analysis, these influences are
indicated (Table 2).
Comparison of the PQN data revealed that several
metabolites exhibited high Q2
values in PLS models from
both quantification approaches, as shown in Figure 4. Of these,
the highest were N,N-dimethylglycine, succinate, trans-
aconitate, and 2-oxoglutarate. The Q2
value for several
Figure 4. continued
Analytical Chemistry Article
dx.doi.org/10.1021/ac400449f | Anal. Chem. 2013, 85, 6674−66816679
metabolites was substantially different when comparing the two
methods. Those performing well only in the targeted spectral
fitting approach included 1-methylnicotinamide, trigonelline,
and tyrosine. Conversely, those performing well only in the
targeted integration approach included N,N-dimethylformamide,
allantoin, 3-indoxysulfate, dimethylamine, malonate, guanidinoa-
cetate, and ethanol. Oxaloacetate, phenylacetylglycine, 1,6-andro-
β-D-glucose, and glucose exhibited low, or subzero Q2
values in
both models.
In summary, metabolites artificially well modeled as a con-
sequence of the urinary dilution factor may be shown to be poorly
modeled following normalization (e.g., an overlapped peak on a
variable background). Conversely, metabolites apparently
poorly modeled may have their concentration structure across
the experimental design revealed (e.g., a small peak on a variable
background). Metabolite resonances in spectral regions with
little background variation, and that are well resolved should be
well modeled and exhibit a high Q2
value. It should be noted
Figure 4. Goodness-of-fit (R2
, green bars) and goodness-of-prediction (Q2
, blue bars) metrics generated across the {3,3} simplex-lattice mixture
design data for 54 metabolites for (A) spectral fitting data set normalized to TSP, (B) spectral integration data set normalized to TSP, (C) spectral
fitting data set normalized using PQN, and (D) spectral integration data set normalized using PQN.
Analytical Chemistry Article
dx.doi.org/10.1021/ac400449f | Anal. Chem. 2013, 85, 6674−66816680
that metabolites that are invariant across all the component
samples might artificially appear to be poorly quantified as a
consequence of their low difference in signal relative to the
background noise. In this study, we deliberately included samples
of the same type (urine), but of varying similarity (rat (day) vs
rat (night) vs human) to produce a spectral set with contrasting
metabolite concentrations and background matrix effects.
We chose a simple mixture design as an exemplar of the
approach, but other designs are possible. For example, this might
have particular value when substantial changes in the back-
ground/matrix are expected to change between the samples
in each class (e.g., toxicological interventions that result in
proteinuria, clinical samples containing high concentrations of
treatment excipients). Samples pooled according to class and
titrated to give linear combinations with known proportions
would characterize the sample compositional space between these
classes. In practice, the approach described could be adapted
for use in larger sample sets; post hoc selection of samples that
are identified by their profiles as being at contrasting extremes
could be used in this way to characterize the individual metabolite
behavior (linearity and matrix effect in relation to a varying
background) across the sample compositional space.
■ CONCLUSION
The approach we report combines the use of sample mixing to
encode sample spectra according to a known experimental
design, with multivariate analysis that allows the theoretical and
observed responses to be compared. We propose that a Q2
statistic is a suitable index with which to make this comparison.
This statistic provides an unbiased estimate of how reliable the
quantification of a particular spectral feature is across the
sample compositional space, and thus which can be safely
interpreted from the urinary data. We found PQN suitable to
remove nuisance variation attributable to gross sample dilution,
and this procedure helped reveal the variation of interest, that
related to the experimental design. Broad agreement between
targeted spectral fitting and targeted spectral integration approaches
was observed, but differences in the response of metabolites with
peaks in overlapped or baseline-dominated spectral regions. This
approach, which efficiently exploits the information contained in
several samples simultaneously, has general applicability, can be
used as an additional metric for profile quality assessment when
conducting biomarker discovery research using spectroscopic
platforms. We suggest that the method offers good complemen-
tarity to measures of analytical reproducibility obtained by replicate
analysis of individual samples.
■ ASSOCIATED CONTENT
*S Supporting Information
Additional material as described in the text. This material is
available free of charge via the Internet at http://pubs.acs.org.
■ AUTHOR INFORMATION
Corresponding Author
*E-mail: toby.athersuch@imperial.ac.uk.
Notes
The authors declare no competing financial interest.
■ ACKNOWLEDGMENTS
The authors wish to acknowledge the Consortium of
Metabonomic Toxicology (COMET) - comprising Bristol-
Myers-Squibb, Hoffman-La Roche Pharmaceuticals, Pfizer Inc.,
Eli Lilly & Co. and NovoNordisk - for the provision of rat urine
samples.
■ REFERENCES
(1) Nicholson, J. K.; Lindon, J. C.; Holmes, E. Xenobiotica 1999, 29,
1181−1189.
(2) Nicholson, J. K.; Connelly, J.; Lindon, J. C.; Holmes, E. Nat. Rev.
Drug Discovery 2002, 1, 153−161.
(3) Nicholson, J. K.; Lindon, J. C. Nature 2008, 455, 1054−1056.
(4) Keun, H. C. Biomarker discovery for drug development and
translational medicine using metabonomics. Oncogenes Meet Metabo-
lism: From Deregulated Genes to a Broader Understanding of Tumour
Physiology; Kroemer, G., Mumberg, D., Keun, K., Riefke, B., Steger-
Hartman, T., Petersen, K., Eds.; Springer: Berlin, 2008; Vol. 4, pp 79−
98.
(5) Keun, H. C.; Athersuch, T. J. Pharmacogenomics 2007, 8, 731−
741.
(6) Nicholson, J. K.; Holmes, E.; Kinross, J. M.; Darzi, A. W.; Takats,
Z.; Lindon, J. C Nature 2012, 491, 384−392.
(7) Holmes, E.; Loo, R. L.; Stamler, J.; Bictash, M.; Yap, I. K.; Chan,
Q.; Ebbels, T.; De Iorio, M.; Brown, I. J.; Veselkov, K. A.; Daviglus, M.
L.; Kesteloot, H.; Ueshima, H.; Zhao, L.; Nicholson, J. K.; Elliott, P.
Nature 2008, 453, 396−400.
(8) Athersuch, T. J. Bioanalysis 2012, 4, 2207−2212.
(9) Keun, H. C.; Ebbels, T. M.; Antti, H.; Bollard, M. E.; Beckonert,
O.; Schlotterbeck, G.; Senn, H.; Niederhauser, U.; Holmes, E.; Lindon,
J. C.; Nicholson, J. K. Chem. Res. Toxicol. 2002, 15, 1380−1386.
(10) Lindon, J. C.; Keun, H. C.; Ebbels, T. M. D.; Pearce, J. M. T.;
Holmes, E.; Nicholson, J. K. Pharmacogenomics 2005, 6, 691−699.
(11) Forgue, P.; Halouska, S.; Werth, M.; Xu, K.; Harris, S.; Powers,
R. J. Proteome Res. 2006, 5, 1916−1923.
(12) Holmes, E.; Nicholson, J. K. Ernst Schering Found. Symp. Proc.
2007, 227−249.
(13) Ellis, J. K.; Athersuch, T. J.; Thomas, L. D.; Teichert, F.; Perez-
Trujillo, M.; Svendsen, C.; Spurgeon, D. J.; Singh, R.; Jarup, L.; Bundy,
J. G.; Keun, H. C. BMC Med. 2012, 10, 61.
(14) Veselkov, K. A.; Lindon, J. C.; Ebbels, T. M.; Crockford, D.;
Volynkin, V. V.; Holmes, E.; Davies, D. B.; Nicholson, J. K. Anal.
Chem. 2009, 81, 56−66.
(15) Weljie, A. M.; Newton, J.; Mercier, P.; Carlson, E.; Slupsky, C.
M. Anal. Chem. 2006, 78, 4430−4442.
(16) accessed January 2013.
(17) Athersuch, T. J.; Keun, H.; Tang, H.; Nicholson, J. K. J. Pharm.
Biomed Anal. 2006, 40, 410−416.
(18) Ebbels, T. M. D.; Keun, H. C.; Beckonert, O. P.; Bollard, M. E.;
Lindon, J. C.; Holmes, E.; Nicholson, J. K. J. Proteome Res. 2007, 6,
4407−4422.
(19) Beckonert, O.; Keun, H. C.; Ebbels, T. M.; Bundy, J.; Holmes,
E.; Lindon, J. C.; Nicholson, J. K. Nat. Protoc. 2007, 2, 2692−2703.
(20) Neuhaus, D.; Ismail, I. M.; Chung, C. W. J. Magn. Reson. Series A
1996, 118, 256−263.
(21) Dieterle, F.; Ross, A.; Schlotterbeck, G.; Senn, H. Anal. Chem.
2006, 78, 4281−4290.
(22) Keun, H. C.; Ebbels, T. M. D.; Antti, H.; Bollard, M. E.;
Beckonert, O.; Holmes, E.; Lindon, J. C.; Nicholson, J. K Anal. Chim.
Acta 2003, 490, 265−276.
(23) Wold, S.; Esbensen, K.; Geladi, P. Chemometr. Intell. Lab. 1987,
2, 37−52.
(24) Wold, S.; Ruhe, A.; Wold, H.; Dunn, W. J. Siam J. Sci. Comput.
1984, 5, 735−743.
Analytical Chemistry Article
dx.doi.org/10.1021/ac400449f | Anal. Chem. 2013, 85, 6674−66816681

More Related Content

What's hot

QSAR Studies of the Inhibitory Activity of a Series of Substituted Indole and...
QSAR Studies of the Inhibitory Activity of a Series of Substituted Indole and...QSAR Studies of the Inhibitory Activity of a Series of Substituted Indole and...
QSAR Studies of the Inhibitory Activity of a Series of Substituted Indole and...inventionjournals
 
Insilico methods for design of novel inhibitors of Human leukocyte elastase
Insilico methods for design of novel inhibitors of Human leukocyte elastaseInsilico methods for design of novel inhibitors of Human leukocyte elastase
Insilico methods for design of novel inhibitors of Human leukocyte elastaseJayashankar Lakshmanan
 
CoMFA CoMFA Comparative Molecular Field Analysis)
CoMFA CoMFA Comparative Molecular Field Analysis)CoMFA CoMFA Comparative Molecular Field Analysis)
CoMFA CoMFA Comparative Molecular Field Analysis)Pinky Vincent
 
2015_Nature Chem
2015_Nature Chem2015_Nature Chem
2015_Nature ChemXimin He
 
The use of functionalized afm tips as molecular sensors in the detection of p...
The use of functionalized afm tips as molecular sensors in the detection of p...The use of functionalized afm tips as molecular sensors in the detection of p...
The use of functionalized afm tips as molecular sensors in the detection of p...Grupo de Pesquisa em Nanoneurobiofisica
 
Antibacterial Activity of Schiff Bases Derived from OrthoDiaminocyclohexane, ...
Antibacterial Activity of Schiff Bases Derived from OrthoDiaminocyclohexane, ...Antibacterial Activity of Schiff Bases Derived from OrthoDiaminocyclohexane, ...
Antibacterial Activity of Schiff Bases Derived from OrthoDiaminocyclohexane, ...inventionjournals
 
quantitative structure activity relationship studies of anti proliferative ac...
quantitative structure activity relationship studies of anti proliferative ac...quantitative structure activity relationship studies of anti proliferative ac...
quantitative structure activity relationship studies of anti proliferative ac...IJEAB
 
NEAFS_DSA_FTIR_Image_NBOMes and LSD_Poster
NEAFS_DSA_FTIR_Image_NBOMes and LSD_PosterNEAFS_DSA_FTIR_Image_NBOMes and LSD_Poster
NEAFS_DSA_FTIR_Image_NBOMes and LSD_PosterAmanda Moore
 
3 d qsar approaches structure
3 d qsar approaches structure3 d qsar approaches structure
3 d qsar approaches structureROHIT PAL
 
Ligand based drug design
Ligand based drug designLigand based drug design
Ligand based drug designSatyendra Yadav
 
techniques used in Metabolite profiling of bryophytes ppt
techniques used in Metabolite  profiling of bryophytes ppttechniques used in Metabolite  profiling of bryophytes ppt
techniques used in Metabolite profiling of bryophytes pptUnnatiChopra1
 
Non biopsy diagnosis of acute rejection of Renal allograft
  Non biopsy diagnosis of acute rejection of Renal allograft  Non biopsy diagnosis of acute rejection of Renal allograft
Non biopsy diagnosis of acute rejection of Renal allograftBakshish Singh
 
Spectroscopic Study on the Interaction between Mobic and Lysozyme
Spectroscopic Study on the Interaction between Mobic and LysozymeSpectroscopic Study on the Interaction between Mobic and Lysozyme
Spectroscopic Study on the Interaction between Mobic and LysozymeAI Publications
 
Application of Multivariate Adaptive Regression Splines (MARSplines) for Pred...
Application of Multivariate Adaptive Regression Splines (MARSplines) for Pred...Application of Multivariate Adaptive Regression Splines (MARSplines) for Pred...
Application of Multivariate Adaptive Regression Splines (MARSplines) for Pred...Maciej Przybyłek
 
Evaluation of the mass attenuation coefficient and Effective Atomic Number of...
Evaluation of the mass attenuation coefficient and Effective Atomic Number of...Evaluation of the mass attenuation coefficient and Effective Atomic Number of...
Evaluation of the mass attenuation coefficient and Effective Atomic Number of...IOSRJAP
 

What's hot (20)

QSAR Studies of the Inhibitory Activity of a Series of Substituted Indole and...
QSAR Studies of the Inhibitory Activity of a Series of Substituted Indole and...QSAR Studies of the Inhibitory Activity of a Series of Substituted Indole and...
QSAR Studies of the Inhibitory Activity of a Series of Substituted Indole and...
 
Insilico methods for design of novel inhibitors of Human leukocyte elastase
Insilico methods for design of novel inhibitors of Human leukocyte elastaseInsilico methods for design of novel inhibitors of Human leukocyte elastase
Insilico methods for design of novel inhibitors of Human leukocyte elastase
 
CoMFA CoMFA Comparative Molecular Field Analysis)
CoMFA CoMFA Comparative Molecular Field Analysis)CoMFA CoMFA Comparative Molecular Field Analysis)
CoMFA CoMFA Comparative Molecular Field Analysis)
 
2015_Nature Chem
2015_Nature Chem2015_Nature Chem
2015_Nature Chem
 
Metabolomics
MetabolomicsMetabolomics
Metabolomics
 
The use of functionalized afm tips as molecular sensors in the detection of p...
The use of functionalized afm tips as molecular sensors in the detection of p...The use of functionalized afm tips as molecular sensors in the detection of p...
The use of functionalized afm tips as molecular sensors in the detection of p...
 
Antibacterial Activity of Schiff Bases Derived from OrthoDiaminocyclohexane, ...
Antibacterial Activity of Schiff Bases Derived from OrthoDiaminocyclohexane, ...Antibacterial Activity of Schiff Bases Derived from OrthoDiaminocyclohexane, ...
Antibacterial Activity of Schiff Bases Derived from OrthoDiaminocyclohexane, ...
 
quantitative structure activity relationship studies of anti proliferative ac...
quantitative structure activity relationship studies of anti proliferative ac...quantitative structure activity relationship studies of anti proliferative ac...
quantitative structure activity relationship studies of anti proliferative ac...
 
NEAFS_DSA_FTIR_Image_NBOMes and LSD_Poster
NEAFS_DSA_FTIR_Image_NBOMes and LSD_PosterNEAFS_DSA_FTIR_Image_NBOMes and LSD_Poster
NEAFS_DSA_FTIR_Image_NBOMes and LSD_Poster
 
3 d qsar approaches structure
3 d qsar approaches structure3 d qsar approaches structure
3 d qsar approaches structure
 
Matrix Effect
Matrix EffectMatrix Effect
Matrix Effect
 
Ligand based drug design
Ligand based drug designLigand based drug design
Ligand based drug design
 
Metabolomics
MetabolomicsMetabolomics
Metabolomics
 
techniques used in Metabolite profiling of bryophytes ppt
techniques used in Metabolite  profiling of bryophytes ppttechniques used in Metabolite  profiling of bryophytes ppt
techniques used in Metabolite profiling of bryophytes ppt
 
Non biopsy diagnosis of acute rejection of Renal allograft
  Non biopsy diagnosis of acute rejection of Renal allograft  Non biopsy diagnosis of acute rejection of Renal allograft
Non biopsy diagnosis of acute rejection of Renal allograft
 
Spectroscopic Study on the Interaction between Mobic and Lysozyme
Spectroscopic Study on the Interaction between Mobic and LysozymeSpectroscopic Study on the Interaction between Mobic and Lysozyme
Spectroscopic Study on the Interaction between Mobic and Lysozyme
 
Application of Multivariate Adaptive Regression Splines (MARSplines) for Pred...
Application of Multivariate Adaptive Regression Splines (MARSplines) for Pred...Application of Multivariate Adaptive Regression Splines (MARSplines) for Pred...
Application of Multivariate Adaptive Regression Splines (MARSplines) for Pred...
 
Metabolomics
MetabolomicsMetabolomics
Metabolomics
 
Evaluation of the mass attenuation coefficient and Effective Atomic Number of...
Evaluation of the mass attenuation coefficient and Effective Atomic Number of...Evaluation of the mass attenuation coefficient and Effective Atomic Number of...
Evaluation of the mass attenuation coefficient and Effective Atomic Number of...
 
Food metabolomics Arapitsas 2017
Food metabolomics Arapitsas 2017Food metabolomics Arapitsas 2017
Food metabolomics Arapitsas 2017
 

Similar to Chenomx

Applications of mass spectrometry.seminar.pptx
Applications of mass spectrometry.seminar.pptxApplications of mass spectrometry.seminar.pptx
Applications of mass spectrometry.seminar.pptxAsif Shaikh
 
Metabolic engineering approaches in medicinal plants
Metabolic engineering approaches in medicinal plantsMetabolic engineering approaches in medicinal plants
Metabolic engineering approaches in medicinal plantsN Poorin
 
JBEI Research Highlights - August 2017
JBEI Research Highlights - August 2017 JBEI Research Highlights - August 2017
JBEI Research Highlights - August 2017 Irina Silva
 
JBEI Research Highlights - November 2018
JBEI Research Highlights - November 2018 JBEI Research Highlights - November 2018
JBEI Research Highlights - November 2018 Irina Silva
 
EXPERIMENTAL IMPLEMENTATION OF EMBARRASINGLY PARALLEL PROCESS IN ANALYSIS OF ...
EXPERIMENTAL IMPLEMENTATION OF EMBARRASINGLY PARALLEL PROCESS IN ANALYSIS OF ...EXPERIMENTAL IMPLEMENTATION OF EMBARRASINGLY PARALLEL PROCESS IN ANALYSIS OF ...
EXPERIMENTAL IMPLEMENTATION OF EMBARRASINGLY PARALLEL PROCESS IN ANALYSIS OF ...ijesajournal
 
Preparation and evaluation of electrospun nanofibers containing pectin and ti...
Preparation and evaluation of electrospun nanofibers containing pectin and ti...Preparation and evaluation of electrospun nanofibers containing pectin and ti...
Preparation and evaluation of electrospun nanofibers containing pectin and ti...Nanomedicine Journal (NMJ)
 
Journal Combinatorial Chemistry 2006 v8 p820
Journal Combinatorial Chemistry 2006 v8 p820Journal Combinatorial Chemistry 2006 v8 p820
Journal Combinatorial Chemistry 2006 v8 p820Peter Tidswell
 
Publication - Alternative Surfactants for Improved Efficiency of In Situ Tryp...
Publication - Alternative Surfactants for Improved Efficiency of In Situ Tryp...Publication - Alternative Surfactants for Improved Efficiency of In Situ Tryp...
Publication - Alternative Surfactants for Improved Efficiency of In Situ Tryp...Nathan Marshall
 
Nucl_Acids_Res-2015-Yang-NAR_gkv060
Nucl_Acids_Res-2015-Yang-NAR_gkv060Nucl_Acids_Res-2015-Yang-NAR_gkv060
Nucl_Acids_Res-2015-Yang-NAR_gkv060Bing Yang
 
Ligand interaction by kk sahu
Ligand interaction by kk sahuLigand interaction by kk sahu
Ligand interaction by kk sahuKAUSHAL SAHU
 
Use of fluorescence lifetime technology to provide efficient protection from ...
Use of fluorescence lifetime technology to provide efficient protection from ...Use of fluorescence lifetime technology to provide efficient protection from ...
Use of fluorescence lifetime technology to provide efficient protection from ...Dmitry Gakamsky
 
2013-Blomquist-Targeted RNA-sequencing with competitive multiplex-PCR amplico...
2013-Blomquist-Targeted RNA-sequencing with competitive multiplex-PCR amplico...2013-Blomquist-Targeted RNA-sequencing with competitive multiplex-PCR amplico...
2013-Blomquist-Targeted RNA-sequencing with competitive multiplex-PCR amplico...Ji-Youn Yeo
 
Nanobiosensors based on chemically modified afm probes a useful tool for met...
Nanobiosensors based on chemically modified afm probes  a useful tool for met...Nanobiosensors based on chemically modified afm probes  a useful tool for met...
Nanobiosensors based on chemically modified afm probes a useful tool for met...Grupo de Pesquisa em Nanoneurobiofisica
 
Weber_FinalProposal
Weber_FinalProposalWeber_FinalProposal
Weber_FinalProposalAnna Weber
 
Radiolytic Modification of Basic Amino Acid Residues in Peptides : Probes for...
Radiolytic Modification of Basic Amino Acid Residues in Peptides : Probes for...Radiolytic Modification of Basic Amino Acid Residues in Peptides : Probes for...
Radiolytic Modification of Basic Amino Acid Residues in Peptides : Probes for...Keiji Takamoto
 

Similar to Chenomx (20)

Applications of mass spectrometry.seminar.pptx
Applications of mass spectrometry.seminar.pptxApplications of mass spectrometry.seminar.pptx
Applications of mass spectrometry.seminar.pptx
 
Metabolic engineering approaches in medicinal plants
Metabolic engineering approaches in medicinal plantsMetabolic engineering approaches in medicinal plants
Metabolic engineering approaches in medicinal plants
 
JBEI Research Highlights - August 2017
JBEI Research Highlights - August 2017 JBEI Research Highlights - August 2017
JBEI Research Highlights - August 2017
 
JBEI Research Highlights - November 2018
JBEI Research Highlights - November 2018 JBEI Research Highlights - November 2018
JBEI Research Highlights - November 2018
 
c4ra02698e
c4ra02698ec4ra02698e
c4ra02698e
 
EXPERIMENTAL IMPLEMENTATION OF EMBARRASINGLY PARALLEL PROCESS IN ANALYSIS OF ...
EXPERIMENTAL IMPLEMENTATION OF EMBARRASINGLY PARALLEL PROCESS IN ANALYSIS OF ...EXPERIMENTAL IMPLEMENTATION OF EMBARRASINGLY PARALLEL PROCESS IN ANALYSIS OF ...
EXPERIMENTAL IMPLEMENTATION OF EMBARRASINGLY PARALLEL PROCESS IN ANALYSIS OF ...
 
Preparation and evaluation of electrospun nanofibers containing pectin and ti...
Preparation and evaluation of electrospun nanofibers containing pectin and ti...Preparation and evaluation of electrospun nanofibers containing pectin and ti...
Preparation and evaluation of electrospun nanofibers containing pectin and ti...
 
Journal Combinatorial Chemistry 2006 v8 p820
Journal Combinatorial Chemistry 2006 v8 p820Journal Combinatorial Chemistry 2006 v8 p820
Journal Combinatorial Chemistry 2006 v8 p820
 
Publication - Alternative Surfactants for Improved Efficiency of In Situ Tryp...
Publication - Alternative Surfactants for Improved Efficiency of In Situ Tryp...Publication - Alternative Surfactants for Improved Efficiency of In Situ Tryp...
Publication - Alternative Surfactants for Improved Efficiency of In Situ Tryp...
 
Metabolomics
MetabolomicsMetabolomics
Metabolomics
 
Nucl_Acids_Res-2015-Yang-NAR_gkv060
Nucl_Acids_Res-2015-Yang-NAR_gkv060Nucl_Acids_Res-2015-Yang-NAR_gkv060
Nucl_Acids_Res-2015-Yang-NAR_gkv060
 
Ligand interaction by kk sahu
Ligand interaction by kk sahuLigand interaction by kk sahu
Ligand interaction by kk sahu
 
HTS by mukesh
HTS by mukeshHTS by mukesh
HTS by mukesh
 
Use of fluorescence lifetime technology to provide efficient protection from ...
Use of fluorescence lifetime technology to provide efficient protection from ...Use of fluorescence lifetime technology to provide efficient protection from ...
Use of fluorescence lifetime technology to provide efficient protection from ...
 
2013-Blomquist-Targeted RNA-sequencing with competitive multiplex-PCR amplico...
2013-Blomquist-Targeted RNA-sequencing with competitive multiplex-PCR amplico...2013-Blomquist-Targeted RNA-sequencing with competitive multiplex-PCR amplico...
2013-Blomquist-Targeted RNA-sequencing with competitive multiplex-PCR amplico...
 
journal published article
journal published articlejournal published article
journal published article
 
Metabolomics
Metabolomics Metabolomics
Metabolomics
 
Nanobiosensors based on chemically modified afm probes a useful tool for met...
Nanobiosensors based on chemically modified afm probes  a useful tool for met...Nanobiosensors based on chemically modified afm probes  a useful tool for met...
Nanobiosensors based on chemically modified afm probes a useful tool for met...
 
Weber_FinalProposal
Weber_FinalProposalWeber_FinalProposal
Weber_FinalProposal
 
Radiolytic Modification of Basic Amino Acid Residues in Peptides : Probes for...
Radiolytic Modification of Basic Amino Acid Residues in Peptides : Probes for...Radiolytic Modification of Basic Amino Acid Residues in Peptides : Probes for...
Radiolytic Modification of Basic Amino Acid Residues in Peptides : Probes for...
 

Chenomx

  • 1. Evaluation of 1 H NMR Metabolic Profiling Using Biofluid Mixture Design Toby J. Athersuch,*,†,‡ Shahid Malik,§ Aalim Weljie,∥,⊥ Jack Newton,∥ and Hector C. Keun† † Computational and Systems Medicine, Department of Surgery and Cancer, Faculty of Medicine, Sir Alexander Fleming Building, Imperial College London, South Kensington, SW7 2AZ, U.K. ‡ MRC-PHE Centre for Environment and Health, Department of Epidemiology and Biostatistics, School of Public Health, Faculty of Medicine, Imperial College London, Norfolk Place, London, W2 1PG, U.K. § Chenomx Inc., Suite 800, 10050 112 Street, Edmonton, Alberta, T5K 2J1, Canada ∥ Department of Biological Sciences, Bio-NMR Center, University of Calgary, Calgary, Alberta, T2N 1N4, Canada ⊥ Institute for Translational Medicine and Therapeutics, Perelman School of Medicine, University of Pennsylvania, 10-113 Translational Research Center, 3400 Civic Center Boulevard, Building 421, Philadelphia, Pennsylvania 19104, United States *S Supporting Information ABSTRACT: A strategy for evaluating the performance of quantitative spectral analysis tools in conditions that better approximate background variation in a metabonomics experiment is presented. Three different urine samples were mixed in known proportions according to a {3, 3} simplex lattice experimental design and analyzed in triplicate by 1D 1 H NMR spec- troscopy. Fifty-four urinary metabolites were subsequently quantified from the sample spectra using two methods common in metabolic profiling studies: (1) targeted spectral fitting and (2) targeted spectral integration. Multivariate analysis using partial least-squares (PLS) regression showed the latent structure of the spectral set recapitulated the experimental mix- ture design. The goodness-of-prediction statistic (Q2 ) of each metabolite variable in a PLS model was calculated as a metric for the reliability of measurement, across the sample compositional space. Several metabolites were observed to have low Q2 values, largely as a consequence of their spectral resonances having low s/n or strong overlap with other sample components. This strategy has the potential to allow evaluation of spectral features obtained from metabolic profiling platforms in the context of the compositional background found in real biological sample sets, which may be subject to considerable variation. We suggest that it be incorporated into metabolic profiling studies to improve the estimation of matrix effects that confound accurate metabolite measurement. This novel method provides a rational basis for exploiting information from several samples in an efficient manner and avoids the use of multiple spike-in authentic standards, which may be difficult to obtain. ■ INTRODUCTION Metabolic Profiling. Metabolic profiling (metabonomics/ metabolomics) has become a key platform in systems biology; the application of spectroscopy or spectrometry to biological samples provides a multicomponent metabolic phenotype that reflects a large number of interacting upstream processes including gene expression, cellular status, and organism function.1−3 Metabolite profiles are currently being used in a wide variety of contexts including “bench-to-bedside” translational medicine,4,5 real-time profiling for enhanced biomarker-based decision tools for clinicians/surgeons,6 and high-throughput metabolic phenotyping in large-scale molecular epidemiological studies aimed at understanding chronic disease risk and etiology.7,8 Nuclear magnetic resonance (NMR) spectroscopy is a core analytical platform used to characterize biological matrices in metabolic profiling studies as it provides quantitative spectra that capture concentration information on multiple metabolites simultaneously, in a highly robust and reproducible manner.9 A large number of studies have used NMR as the primary analytical tool for exploring a wide range of research questions, including those in drug toxicity testing,10 efficacy assessment,11 and population studies.12,13 Metabolic profiles obtained by NMR spectroscopy typically contain hundreds or thousands of real spectral features, along with confound- ing signals (including noise), that span a wide dynamic range (>8 orders of magnitude). Accurate quantification of metab- olites using biofluid spectra obtained by NMR spectroscopy can be relatively difficult, as a consequence of signal overlap, resulting from insufficient spectral resolution, that may reduce accuracy of integral measurements made as an estimator of Received: February 10, 2013 Accepted: June 3, 2013 Published: June 3, 2013 Article pubs.acs.org/ac © 2013 American Chemical Society 6674 dx.doi.org/10.1021/ac400449f | Anal. Chem. 2013, 85, 6674−6681
  • 2. concentration. In addition to peak-alignment based methods of spectral deconvolution,14 spectral fitting approaches, that use individual template spectra matching for each metabolite, are commonly used in an attempt to reduce the influence of signal overlap on quantification.15 By virtue of the spectral matching process (that typically uses multiple peak fits in the spectrum to provide a best estimate of concentration), spectral fitting has the advantage oversimple spectral integration in that it is inherently less affected by background variation arising from the sample matrix, and from spectral artifacts such as the residual water peak in aqueous sample spectra. However, spectral fitting is more time-consuming and may be prone to user error or subjectivity. A common approach used to assess the quantification accuracy in biofluid spectra is the use of traditional “spike-in” experiments, whereby authentic standards are added to the sample in known concentration. However, these experiments are conducted typically in the context of an invariant back- ground, which is often not representative of the “real world” scenario where baseline signals from different samples vary due to numerous matrix effects. Additional measures of the spectral quality and reliability of individual measurements made in metabolic profiling studies, that characterize performance in real sample sets, are therefore of potential utility to the metabolic profiling community. Mixture Design. Mixture design experiments are routinely used for the selection of optimal criteria for production pro- cesses, formulation, and more generally in the characterization of relationships between response and system composition. There are numerous designs that can be used, depending on the constraints placed on the mixture components; a simplex- lattice design reflects one of the simplest designs, and is described as follows: “A {q, m} simplex-lattice design for q com- ponents consists of points defined by the following coordinate settings: the proportions assumed by each component take the m+1 equally spaced values from 0 to 1 = =x m m i q0, 1/ , 2/ , ..., 1 for 1, 2, ...,i ...and all possible combinations (mixtures) of the proportions from this equation are used.”16 The proportions must sum to unity. For example, a {3,3} simplex lattice design represents three components (q = 3), each of which have four (m + 1 = 4), equally spaced, different possible levels (0, 1/3, 2/3, 1), and therefore will have ten possible mixture combinations. We propose that mixing different biofluid samples in known proportions according to a mixture design (such as a simplex lattice) will produce a sample set that enables metabolite behavior across the sample compositional space to be characterized by regression of the design against the metabolite response. In an ideal situation, the observed response of an individual metabolite will exactly follow the mixture design, and a perfect fit will be achieved. In reality, matrix effects and confounding signal overlap may reduce the accuracy of metabolite responses and reduce the correspondence with the mixture design. Thus, this approach allows the reproducibility of individual metabolites to be assessed, and those that are adversely affected by matrix effects or signal overlap to be identified. Here we have applied this strategy of mixing intact biofluids, according to a predetermined experimental mixture design, to compare the performance of two commonly used metabolite quantification methods in the context of “real world” 1 H NMR metabonomic analysis. The potential benefits of incorporating a designed mixture component in metabonomic analyses, as a method of assessing the accuracy of metabolite quantification, are discussed. We suggest that this strategy may have general benefits and applicability in metabolic profiling studies. ■ MATERIALS AND METHODS Chemicals. D2O was obtained from Goss Scientific (Nantwich, U.K.). All other reagents were of analytical grade and obtained from SigmaAldrich (Poole, U.K.). Experimental Design. A schematic of the experimental design is shown in (Figure 1), with details of discussed in turn below. Sample Collection and Preparation. Urine samples were obtained from an existing large-scale toxicological study resource.10,17,18 Sprague−Dawley rats (n = 7) were individually housed in standard metabolism cages (21 ± 3 ◦C, relative humidity 55 ± 15%) and acclimatized for six days prior to the start of the study (t = 0 h). A standard diet (Purina chow 5002) and fresh water (acidified to pH 2.5 using HCl to prevent microbial growth) was available to each animal ad libitum. Urine samples used in the current study were collected during Figure 1. Schematic showing the overall approach described. Three different urines were mixed in known proportions according to a mixture design (1, 2). Concentrations of metabolites were determined by 1 H NMR spectroscopy (3). Spectral fitting and spectral integration were both used for quantification (4). The mixture design data (Y block) were used in a PLS regression against the metabolite concentration data (X block) to generate model metrics (5). Table 1. Sample Composition for Designed Biofluid Mixtures Used in This Study Following a {3,3} Simplex- Lattice Mixture Design volume (μL) rat urine human urine sample number 0−8 h 8−24 h spot sample sodium phospate buffer 1 300 0 0 300 2 0 300 0 300 3 0 0 300 300 4 200 100 0 300 5 200 0 100 300 6 100 200 0 300 7 0 200 100 300 8 100 0 200 300 9 0 100 200 300 10 100 100 100 300 Analytical Chemistry Article dx.doi.org/10.1021/ac400449f | Anal. Chem. 2013, 85, 6674−66816675
  • 3. Table 2. 1 H NMR Spectral Regions Used for Integration, Estimated Metabolite Concentrations Derived from Spectral Fitting Using Chenomx Software, and PLS Model Metricsa spectral region (ppm) estimated concentration (μM) Q2 ID metabolite high low width min max mean STDEV A B C D 1 1-methylnicotinamide 4.47 4.46 0.01 11 182 115 55 0.99 0.20 0.91 −0.10 2 1,3-dimethyluratec,f 3.29 3.28 0.01 24 312 149 93 0.98 0.99 0.90 0.73 3 1,6-anhydro-β-D-glucosef 5.45 5.44 0.01 24 70 44 15 0.89 0.01 0.25 −0.11 4 2-hydroxyisobutyratea,c,f 1.35 1.34 0.01 0 36 16 15 0.42 0.98 0.39 0.61 5 2-oxoglutarate 3.02 2.98 0.04 104 4759 2737 1491 0.99 0.99 0.89 0.84 6 2-oxoisocaproateb,c,f 0.94 0.92 0.02 0 76 41 23 0.94 0.99 0.82 0.83 7 3-hydroxyisovalerate 1.26 1.25 0.01 20 32 26 4 0.80 0.98 0.85 0.63 8 3-indoxylsulfate 7.51 7.48 0.03 75 288 175 64 0.98 0.96 −0.01 0.65 9 acetate 1.91 1.91 0.01 10 107 73 32 0.98 0.99 0.89 0.72 10 alanine 1.48 1.46 0.02 65 106 92 14 0.94 0.98 0.80 0.56 11 allantoinc 5.42 5.34 0.08 19 3929 2530 1162 0.73 0.96 −0.06 0.70 12 arginine 1.95 1.87 0.08 129 184 153 17 0.55 0.98 0.66 0.25 13 betainec 3.90 3.88 0.02 29 462 233 133 0.92 0.99 0.69 0.84 14 choline 3.18 3.18 0.01 15 50 30 13 0.88 0.95 0.87 0.80 15 cis-aconitate 3.11 3.09 0.02 161 459 321 89 0.51 0.99 0.66 0.23 16 citrate 2.56 2.50 0.06 1775 14420 8951 4025 0.94 0.99 0.88 0.85 17 creatinine 3.05 3.02 0.02 1873 5154 3388 1028 0.98 0.98 0.85 0.79 18 dimethylamine 2.72 2.70 0.02 125 452 285 100 0.98 0.98 −0.05 0.59 19 ethanolc,d,f 1.19 1.16 0.03 9 36 20 9 0.97 0.98 −0.07 0.45 20 ethanolamine 3.15 3.12 0.03 68 204 144 41 0.84 0.98 0.83 0.66 21 formate 8.45 8.44 0.01 30 321 184 88 0.83 0.98 0.76 0.87 22 fucosef 1.25 1.23 0.02 53 197 135 42 0.97 0.97 0.68 0.64 23 fumarateb,d 6.51 6.51 0.01 0 42 26 13 0.96 0.97 0.71 0.80 24 glucosed,e 4.65 4.63 0.03 198 641 385 140 0.90 −0.04 0.01 −0.01 25 glycine 3.56 3.55 0.01 149 458 270 95 0.95 0.96 0.81 0.74 26 glycolated 3.94 3.93 0.01 104 198 152 29 0.97 0.98 0.82 0.61 27 guanidinoacetated 3.79 3.78 0.01 96 332 190 79 0.87 0.99 −0.07 0.36 28 hippurate 7.85 7.81 0.04 256 3696 2158 1060 0.81 0.99 0.88 0.86 29 isoleucine 1.01 0.99 0.02 9 15 12 2 0.99 0.97 0.73 0.56 30 lactatec 1.33 1.31 0.03 65 106 81 15 0.62 0.98 0.75 0.45 31 leucined 0.97 0.94 0.03 11 16 14 1 0.87 0.98 0.89 −0.11 32 malonated,f 3.12 3.11 0.01 8 43 26 11 0.25 0.98 0.03 0.43 33 methanol 3.35 3.34 0.01 11 87 50 27 0.70 0.92 0.49 0.78 34 methylmalonate 1.24 1.22 0.02 58 156 95 27 0.56 0.94 0.72 0.77 35 methylsuccinatef 1.08 1.06 0.02 14 52 35 12 0.81 0.97 0.03 0.46 36 N,N-dimethylformamidea,f 2.86 2.85 0.01 14 94 48 25 0.99 0.98 0.74 0.44 37 N,N-dimethylglycinef 2.92 2.91 0.01 0 59 11 23 0.75 0.94 0.01 0.76 38 N-acetylglycinee 2.03 2.03 0.01 27 266 131 78 −0.15 0.99 0.93 0.90 39 O-phosphocholinee,f 3.19 3.18 0.01 7 42 25 11 0.96 0.98 0.82 0.65 40 oxaloacetatea,c,d 3.67 3.66 0.01 0 308 76 125 0.95 0.98 −0.21 0.29 41 phenylacetylglycinef 7.43 7.39 0.04 103 235 160 41 0.52 0.98 0.34 0.06 42 pyruvated 2.37 2.36 0.01 11 58 32 15 0.87 0.99 0.25 0.57 43 succinate 2.40 2.39 0.02 17 1100 537 335 0.77 0.99 0.90 0.89 44 taurine 3.43 3.39 0.04 466 1843 1277 448 0.99 0.99 0.85 0.24 45 threoninec 1.33 1.31 0.02 9 62 43 19 0.98 0.98 0.71 0.48 46 trans-aconitate 6.59 6.56 0.03 15 1834 845 566 0.96 0.99 0.91 0.90 47 trigonelline 4.43 4.42 0.01 54 294 168 75 0.88 0.72 0.76 −0.09 48 trimethylamine N-oxidec 3.26 3.25 0.01 124 481 280 116 1.00 0.99 0.30 0.64 49 tryptophana,d 7.76 7.73 0.03 0 41 24 17 0.98 0.92 0.89 0.62 50 tyrosine 6.90 6.87 0.02 34 48 39 5 0.96 0.94 0.68 −0.01 51 uracilb,e 7.54 7.52 0.03 0 123 66 36 0.14 0.98 0.67 0.82 52 ureae 5.95 5.61 0.34 90366 207739 152730 44253 0.70 0.91 0.45 0.65 53 valine 0.99 0.97 0.03 17 24 20 3 0.91 0.97 0.69 0.54 54 xylose 4.59 4.56 0.03 102 561 287 151 0.66 0.00 0.85 0.11 a The Q2 statistic is given for each of four models: (A) spectral fitting data set normalized to TSP, (B) spectral integration data set normalized to TSP, (C) spectral fitting data set normalized using PQN, and (D) spectral integration data set normalized using PQN. a Metabolite present in only one of the component samples. b Metabolite absent in one of the component samples. c Overlapped signal. d Low s/n. e Spectral artifact present in region. f Tentative assignment. Analytical Chemistry Article dx.doi.org/10.1021/ac400449f | Anal. Chem. 2013, 85, 6674−66816676
  • 4. two periods (0−8 h, 8−24 h) from control animals. Urine voided by the animals was collected in the metabolism cage into a container cooled by dry ice. Samples were subsequently stored at −40 °C. Each sample underwent two freeze−thaw cycles before use in this work as a consequence of realiquoting. Further study information has previously been published.17 Additionally, a spot urine sample (5 mL) was obtained from a healthy human volunteer, according to established protocols, including filtration to remove cellular material (0.2 μm Minisart 16534K, Sartorius, Germany), and immediate storage at −40 °C until required for analysis. Urine samples were prepared follow- ing established protocols for NMR metabolome analysis.19 Urine samples were defrosted, vortex mixed (30 s, RT), and centrifuged (16000 g, RT, 10 min) to remove particulate matter. To provide sufficient total sample, for each collection period, 450 μL of each rat urine sample was pooled (total volume 3150 μL per collection period). The three urines (two pooled rat urine, one human spot urine) were mixed according to a {3,3} simplex- lattice mixture design (Table 1), with each mixed sample having a volume of 300 μL. These mixed samples were then buffered by the addition of 300 μL sodium phosphate buffer (pH 7.4, 0.2 M, 80:20 H2O:D2O (v/v)) containing sodium 3-(trimethylsilyl)- [2,2,3,3-2 H4]propionate (TSP, 1 mM). Samples were vortex mixed (30 s, RT), and a 550 μL aliquot transferred to a 96-well autosampler plate. The mixed samples were prepared in triplicate from the pooled rat samples and the human spot urine. The preparation order was randomized. NMR Spectral Acquisition and Processing. 1 H NMR spectra were acquired on a Bruker AVANCE DRX600 NMR spectrometer (Bruker Biospin, Rheinstetten, Germany) operat- ing at 14.1 T (600.29 MHz 1 H NMR frequency) using a PH FI TXI 600SB 5 mm probe maintained at 300 K. Samples were introduced to the probe using a BEST flow-injection system (Bruker) in a randomized order. Gradient shimming was used immediately prior to spectral acquisition to ensure high field homogeneity. Spectral acquisition was made using standard a standard 1D pulse sequence (RD-90°-t1-90°-tm-90°-AQ).20 The t1 delay and the mixing time (tm) were set to 3 μs and 100 ms respectively. All spectra were collected as the sum of 128 free induction decays (FIDs) were collected into 32K complex data points. The spectral width of 12019.23 Hz (20 ppm) giving the FID a native resolution of 0.366 Hz/pt, and an acquisition time (AQ) of 1.36 s. A 2 s relaxation delay (RD) was used between pulses. A presaturation pulse was applied to the water resonance (δH = 4.7 ppm) during RD and tm. Processing of the raw NMR data for analysis using a targeted integration approach was carried out using XWINNMR software (Bruker Biospin, Rheinstetten, Germany), with each FID being multiplied by an exponential weighting func- tion equivalent to a line broadening of 1 Hz prior to Fourier transformation. Resulting frequency-domain spectra were referenced to TSP (δH = 0.00 ppm) and interpolated from 32K to ∼42K data points using a cubic spline function to regularize the abscissa and improve calibration accuracy Figure 2. 1 H NMR spectra of urine samples used in this study: (A) Pooled rat urine 0−8 h collection (n = 7), (B) pooled rat urine 8−24 h collection (n = 7), and (C) human urine spot collection. Spectra were acquired at an observation frequency of 600 MHz using a standard 1D pulse sequence with water presaturation. Analytical Chemistry Article dx.doi.org/10.1021/ac400449f | Anal. Chem. 2013, 85, 6674−66816677
  • 5. (final resolution 0.29 Hz/pt) prior to analysis using in-house scripts running in the Matlab (The Mathworks, Natick) computing environment. Metabolite Quantification. Fifty-four metabolites were quantified using both the spectral fitting approach and a targeted spectral integration approach. Spectral fitting was performed using Chenomx NMR Suite 4.6 (Chenomx Inc., Edmonton, Canada); reference spectra from the Chenomx 600 MHz library were combined so as to best approximate each acquired urine spectrum, and the relative concentrations of each metabolite present determined by reference to the internal TSP standard15 (Table 2). For the targeted spectral integration approach, spectral regions were defined for each of the metabolites of interest (Table 2), with a width sufficient to encapsulate the majority of the peak across the entire set of spectra (determined manually by spectral overlay). The integral area of these regions was calculated using an in-house routine in Matlab. Probabilistic quotient normalization21 (PQN) was applied to remove variation originating from intersample dif- ferences in urinary dilution. Chemometric analysis of meta- bolite concentration data was completed using Simca P+12 (Umetrics, Umea, Sweden). Principal component analysis (PCA, using metabolite concentration data) was conducted. Partial least-squares regression (PLS, using metabolite concen- tration data and the experimental design matrix) allowed goodness-of-fit (R2 ) and goodness-of-prediction (Q2 ) estimates to be made for each metabolite.22 ■ RESULTS AND DISCUSSION Quantification of Metabolites Using NMR. As detailed in the Materials and Methods section, three different urines were mixed according to a {3,3} simplex lattice design, and analyzed in triplicate by 1 H NMR spectroscopy. Representative spectra are shown in Figure 2. A total of 54 metabolites were successfully quantified in these samples using both a targeted spectral fitting approach (involving the fitting of individual reference metabolite spectra to the spectra acquired for each sample mixture), and a targeted spectral integration approach (involving the integration of a representative spectral region). Metabolite data are given in Table 2. Comparison of the metabolite concentrations of metabolites across the three samples containing only one of the three urine components (i.e., the corners of the simplex lattice design) showed considerable variation between the rat and human samples (Supporting Information Figure S1). Some metabolites were present or absent in only one of the samples (2-hydroxyisobutyrate, 2-oxoisocaproate, fumarate, N,N-dimethylglycine, oxaloacetate, tryptophan). Other metabolites spanned up to 2 orders of magnitude in their absolute concentration in these samples; those with the greatest variation included 1-methylnicotina- mide, 1,3-dimethylurate, betaine, and allantoin. PCA/PLS Modeling and Effect of Normalization. Principal component analysis (PCA) and partial least-squares (PLS) regression are widely used multivariate analysis tools based on latent variable methods.23,24 For each quantification approach, the metabolite concentrations (X-matrix) were modeled by PCA in an unsupervised manner, and also modeled against the experimental mixture design (Y- matrix) using PLS. PCA of the data set before and after PQN was conducted and showed differences in the variation captured by the two methods of quantification (Supporting Information, Figure S2). Prior to normalization, the largest variation in each data set was attributable to the sample dilution. Upon normalization, the next biggest variation in the spectral fitting data set was revealed to be the mixture design, whereas in the spectral targeting data set, it was related to nuisance variation, driven by outlying samples, and attributable to inferior water suppression. PLS analysis was used to assess the fit of the relative metabolite variation to the mixture design. The component scores for the PLS models (Figure 3) clearly showed the experi- mental design as anticipated. Switching the X and Y matrices of the PLS model also allowed a calculation of the goodness- of-fit (R2 ) and goodness-of-prediction (Q2 ) values for each metabolite against the {3,3} simplex-lattice design (Figure 3 and Table 2). It can be seen that a large proportion have high Q2 values indicating that the latent structure of these data follow that of the mixture proportions. It was recognized that in these models, overall urinary dilution would have the effect of artificially enhancing some of these values as a consequence of introducing structure into the concentration profiles. PQN concentrations were subsequently modeled and indicated that once the global dilution factor was removed from the con- centration data, several metabolites displayed greatly reduced Q2 values. Removing the dominating dilution difference means that the Q2 statistic reported for each metabolite gives a more realistic representation of the fit to the design of the metabolite response, in the presence of a variable background. There are several reasons that explain low Q2 value including (a) changes in chemical shift as a consequence of pH variation, Figure 3. Partial least-squares regression analysis scores plots indicating latent structure of the TSP normalized data (A) spectral integration data set and (B) spectral fitting data set. It can be seen that the samples recapitulate the {3,3} simplex-lattice mixture design in the score space. Samples (triplicates) are colored according to proportions of the three component urines (Table 1). Analytical Chemistry Article dx.doi.org/10.1021/ac400449f | Anal. Chem. 2013, 85, 6674−66816678
  • 6. (b) the absence of a metabolite in one or more of the component samples (in this case the rat and human urines), (c) low s/n in the measurement, and (d) overlap of spectral features. Where identified during analysis, these influences are indicated (Table 2). Comparison of the PQN data revealed that several metabolites exhibited high Q2 values in PLS models from both quantification approaches, as shown in Figure 4. Of these, the highest were N,N-dimethylglycine, succinate, trans- aconitate, and 2-oxoglutarate. The Q2 value for several Figure 4. continued Analytical Chemistry Article dx.doi.org/10.1021/ac400449f | Anal. Chem. 2013, 85, 6674−66816679
  • 7. metabolites was substantially different when comparing the two methods. Those performing well only in the targeted spectral fitting approach included 1-methylnicotinamide, trigonelline, and tyrosine. Conversely, those performing well only in the targeted integration approach included N,N-dimethylformamide, allantoin, 3-indoxysulfate, dimethylamine, malonate, guanidinoa- cetate, and ethanol. Oxaloacetate, phenylacetylglycine, 1,6-andro- β-D-glucose, and glucose exhibited low, or subzero Q2 values in both models. In summary, metabolites artificially well modeled as a con- sequence of the urinary dilution factor may be shown to be poorly modeled following normalization (e.g., an overlapped peak on a variable background). Conversely, metabolites apparently poorly modeled may have their concentration structure across the experimental design revealed (e.g., a small peak on a variable background). Metabolite resonances in spectral regions with little background variation, and that are well resolved should be well modeled and exhibit a high Q2 value. It should be noted Figure 4. Goodness-of-fit (R2 , green bars) and goodness-of-prediction (Q2 , blue bars) metrics generated across the {3,3} simplex-lattice mixture design data for 54 metabolites for (A) spectral fitting data set normalized to TSP, (B) spectral integration data set normalized to TSP, (C) spectral fitting data set normalized using PQN, and (D) spectral integration data set normalized using PQN. Analytical Chemistry Article dx.doi.org/10.1021/ac400449f | Anal. Chem. 2013, 85, 6674−66816680
  • 8. that metabolites that are invariant across all the component samples might artificially appear to be poorly quantified as a consequence of their low difference in signal relative to the background noise. In this study, we deliberately included samples of the same type (urine), but of varying similarity (rat (day) vs rat (night) vs human) to produce a spectral set with contrasting metabolite concentrations and background matrix effects. We chose a simple mixture design as an exemplar of the approach, but other designs are possible. For example, this might have particular value when substantial changes in the back- ground/matrix are expected to change between the samples in each class (e.g., toxicological interventions that result in proteinuria, clinical samples containing high concentrations of treatment excipients). Samples pooled according to class and titrated to give linear combinations with known proportions would characterize the sample compositional space between these classes. In practice, the approach described could be adapted for use in larger sample sets; post hoc selection of samples that are identified by their profiles as being at contrasting extremes could be used in this way to characterize the individual metabolite behavior (linearity and matrix effect in relation to a varying background) across the sample compositional space. ■ CONCLUSION The approach we report combines the use of sample mixing to encode sample spectra according to a known experimental design, with multivariate analysis that allows the theoretical and observed responses to be compared. We propose that a Q2 statistic is a suitable index with which to make this comparison. This statistic provides an unbiased estimate of how reliable the quantification of a particular spectral feature is across the sample compositional space, and thus which can be safely interpreted from the urinary data. We found PQN suitable to remove nuisance variation attributable to gross sample dilution, and this procedure helped reveal the variation of interest, that related to the experimental design. Broad agreement between targeted spectral fitting and targeted spectral integration approaches was observed, but differences in the response of metabolites with peaks in overlapped or baseline-dominated spectral regions. This approach, which efficiently exploits the information contained in several samples simultaneously, has general applicability, can be used as an additional metric for profile quality assessment when conducting biomarker discovery research using spectroscopic platforms. We suggest that the method offers good complemen- tarity to measures of analytical reproducibility obtained by replicate analysis of individual samples. ■ ASSOCIATED CONTENT *S Supporting Information Additional material as described in the text. This material is available free of charge via the Internet at http://pubs.acs.org. ■ AUTHOR INFORMATION Corresponding Author *E-mail: toby.athersuch@imperial.ac.uk. Notes The authors declare no competing financial interest. ■ ACKNOWLEDGMENTS The authors wish to acknowledge the Consortium of Metabonomic Toxicology (COMET) - comprising Bristol- Myers-Squibb, Hoffman-La Roche Pharmaceuticals, Pfizer Inc., Eli Lilly & Co. and NovoNordisk - for the provision of rat urine samples. ■ REFERENCES (1) Nicholson, J. K.; Lindon, J. C.; Holmes, E. Xenobiotica 1999, 29, 1181−1189. (2) Nicholson, J. K.; Connelly, J.; Lindon, J. C.; Holmes, E. Nat. Rev. Drug Discovery 2002, 1, 153−161. (3) Nicholson, J. K.; Lindon, J. C. Nature 2008, 455, 1054−1056. (4) Keun, H. C. Biomarker discovery for drug development and translational medicine using metabonomics. Oncogenes Meet Metabo- lism: From Deregulated Genes to a Broader Understanding of Tumour Physiology; Kroemer, G., Mumberg, D., Keun, K., Riefke, B., Steger- Hartman, T., Petersen, K., Eds.; Springer: Berlin, 2008; Vol. 4, pp 79− 98. (5) Keun, H. C.; Athersuch, T. J. Pharmacogenomics 2007, 8, 731− 741. (6) Nicholson, J. K.; Holmes, E.; Kinross, J. M.; Darzi, A. W.; Takats, Z.; Lindon, J. C Nature 2012, 491, 384−392. (7) Holmes, E.; Loo, R. L.; Stamler, J.; Bictash, M.; Yap, I. K.; Chan, Q.; Ebbels, T.; De Iorio, M.; Brown, I. J.; Veselkov, K. A.; Daviglus, M. L.; Kesteloot, H.; Ueshima, H.; Zhao, L.; Nicholson, J. K.; Elliott, P. Nature 2008, 453, 396−400. (8) Athersuch, T. J. Bioanalysis 2012, 4, 2207−2212. (9) Keun, H. C.; Ebbels, T. M.; Antti, H.; Bollard, M. E.; Beckonert, O.; Schlotterbeck, G.; Senn, H.; Niederhauser, U.; Holmes, E.; Lindon, J. C.; Nicholson, J. K. Chem. Res. Toxicol. 2002, 15, 1380−1386. (10) Lindon, J. C.; Keun, H. C.; Ebbels, T. M. D.; Pearce, J. M. T.; Holmes, E.; Nicholson, J. K. Pharmacogenomics 2005, 6, 691−699. (11) Forgue, P.; Halouska, S.; Werth, M.; Xu, K.; Harris, S.; Powers, R. J. Proteome Res. 2006, 5, 1916−1923. (12) Holmes, E.; Nicholson, J. K. Ernst Schering Found. Symp. Proc. 2007, 227−249. (13) Ellis, J. K.; Athersuch, T. J.; Thomas, L. D.; Teichert, F.; Perez- Trujillo, M.; Svendsen, C.; Spurgeon, D. J.; Singh, R.; Jarup, L.; Bundy, J. G.; Keun, H. C. BMC Med. 2012, 10, 61. (14) Veselkov, K. A.; Lindon, J. C.; Ebbels, T. M.; Crockford, D.; Volynkin, V. V.; Holmes, E.; Davies, D. B.; Nicholson, J. K. Anal. Chem. 2009, 81, 56−66. (15) Weljie, A. M.; Newton, J.; Mercier, P.; Carlson, E.; Slupsky, C. M. Anal. Chem. 2006, 78, 4430−4442. (16) accessed January 2013. (17) Athersuch, T. J.; Keun, H.; Tang, H.; Nicholson, J. K. J. Pharm. Biomed Anal. 2006, 40, 410−416. (18) Ebbels, T. M. D.; Keun, H. C.; Beckonert, O. P.; Bollard, M. E.; Lindon, J. C.; Holmes, E.; Nicholson, J. K. J. Proteome Res. 2007, 6, 4407−4422. (19) Beckonert, O.; Keun, H. C.; Ebbels, T. M.; Bundy, J.; Holmes, E.; Lindon, J. C.; Nicholson, J. K. Nat. Protoc. 2007, 2, 2692−2703. (20) Neuhaus, D.; Ismail, I. M.; Chung, C. W. J. Magn. Reson. Series A 1996, 118, 256−263. (21) Dieterle, F.; Ross, A.; Schlotterbeck, G.; Senn, H. Anal. Chem. 2006, 78, 4281−4290. (22) Keun, H. C.; Ebbels, T. M. D.; Antti, H.; Bollard, M. E.; Beckonert, O.; Holmes, E.; Lindon, J. C.; Nicholson, J. K Anal. Chim. Acta 2003, 490, 265−276. (23) Wold, S.; Esbensen, K.; Geladi, P. Chemometr. Intell. Lab. 1987, 2, 37−52. (24) Wold, S.; Ruhe, A.; Wold, H.; Dunn, W. J. Siam J. Sci. Comput. 1984, 5, 735−743. Analytical Chemistry Article dx.doi.org/10.1021/ac400449f | Anal. Chem. 2013, 85, 6674−66816681