ArticlePDF Available

Separating NADH and NADPH fluorescence in live cells and tissues using FLIM

Authors:

Abstract and Figures

NAD is a key determinant of cellular energy metabolism. In contrast, its phosphorylated form, NADP, plays a central role in biosynthetic pathways and antioxidant defence. The reduced forms of both pyridine nucleotides are fluorescent in living cells but they cannot be distinguished, as they are spectrally identical. Here, using genetic and pharmacological approaches to perturb NAD(P)H metabolism, we find that fluorescence lifetime imaging (FLIM) differentiates quantitatively between the two cofactors. Systematic manipulations to change the balance between oxidative and glycolytic metabolism suggest that these states do not directly impact NAD(P)H fluorescence decay rates. The lifetime changes observed in cancers thus likely reflect shifts in the NADPH/NADH balance. Using a mathematical model, we use these experimental data to quantify the relative levels of NADH and NADPH in different cell types of a complex tissue, the mammalian cochlea. This reveals NADPH-enriched populations of cells, raising questions about their distinct metabolic roles.
Content may be subject to copyright.
ARTICLE
Received 11 Feb 2014 | Accepted 23 Apr 2014 | Published 29 May 2014
Separating NADH and NADPH fluorescence
in live cells and tissues using FLIM
Thomas S. Blacker
1,2,3
, Zoe F. Mann
2,4
, Jonathan E. Gale
2,4
, Mathias Ziegler
5
, Angus J. Bain
3
,
Gyorgy Szabadkai
2,6,
* & Michael R. Duchen
2,
*
NAD is a key determinant of cellular energy metabolism. In contrast, its phosphorylated form,
NADP, plays a central role in biosynthetic pathways and antioxidant defence. The reduced
forms of both pyridine nucleotides are fluorescent in living cells but they cannot be
distinguished, as they are spectrally identical. Here, using genetic and pharmacological
approaches to perturb NAD(P)H metabolism, we find that fluorescence lifetime imaging
(FLIM) differentiates quantitatively between the two cofactors. Systematic manipulations to
change the balance between oxidative and glycolytic metabolism suggest that these states do
not directly impact NAD(P)H fluorescence decay rates. The lifetime changes observed in
cancers thus likely reflect shifts in the NADPH/NADH balance. Using a mathematical model,
we use these experimental data to quantify the relative levels of NADH and NADPH in
different cell types of a complex tissue, the mammalian cochlea. This reveals NADPH-enri-
ched populations of cells, raising questions about their distinct metabolic roles.
DOI: 10.1038/ncomms4936
OPEN
1
Centre for Mathematics and Physics in the Life Sciences and Experimental Biology, University College London, London WC1E 6BT, UK.
2
Research
Department of Cell & Developmental Biology, University College London, London WC1E 6BT, UK.
3
Department of Physics and Astronomy, University College
London, London WC1E 6BT, UK.
4
UCL Ear Institute, University College London, London WC1X 8EE, UK.
5
Department of Molecular Biology, University of
Bergen, N-5008 Bergen, Norway.
6
Department of Biomedical Sciences, University of Padua and CNR Neuroscience Institute, Padua 35121, Italy. * These
authors contributed equally to this work. Correspondence and requests for materials should be addressed to M.R.D. (email: m.duchen@ucl.ac.uk).
NATURE COMMUNICATIONS | 5:3936 | DOI: 10.1038/ncomms4936 | www.nature.com/naturecommunications 1
& 2014 Macmillan Publishers Limited. All rights reserved.
T
he cellular redox state is the central regulator of energy
production and intermediary metabolism, playing a
crucial role in health and disease
1
. The nicotinamide
adenine dinucleotide (NAD
þ
/NADH) and nicotinamide adenine
dinucleotide phosphate (NADP
þ
/NADPH) redox couples are
the major determinants of redox state in the cell. However, these
engage in distinct metabolic pathways. NAD drives ATP
production in the cytosol by glycolysis and in the mitochondria
by oxidative phosphorylation, while the phosphorylated analogue
NADP governs lipid, amino acid and nucleotide biosynthetic
pathways and the defence against reactive oxygen species by
glutathione (GSH)
2
. Free radical generation is therefore
determined by the redox state of NAD, while NADP redox
state is key to antioxidant defence
3
. The relative abundance of the
two pyridine nucleotides and their redox balance thus mediates
cell fate in a wide range of diseases, including cancer, diabetes and
neurodegeneration. Quantifying their behaviour is therefore
essential in understanding the role of metabolism in these
diseases. However, separating the contributions of the two pools
in intact tissues has proven technically challenging
4
.
In the 1960s, Britton Chance et al.
5
showed that live tissues
illuminated with ultraviolet light emit blue fluorescence, arising
primarily from mitochondrial NADH. The nicotinamide moiety
of NADH absorbs light of wavelength 340
±
30 nm and emits
fluorescence at 460
±
50 nm. As NADP is phosphorylated at a
remote site of the molecule, the fluorescence properties of the
nicotinamide ring of NADPH are identical to those of NADH
6,7
.
Thus, changes in autofluorescence intensity may reflect changes
in either [NADH] or [NADPH], often denoted as NAD(P)H to
indicate the uncertain origin of the signal
8,9
.
Fluorescence lifetime imaging (FLIM) allows the study
of NAD(P)H photochemistry inside living tissue
10–12
.
This technique measures the rates of fluorescence decay of
NAD(P)H, an excited-state process occurring over nanosecond
time scales, which is highly sensitive to the immediate
environment of the fluorophore. We have therefore investigated
whether the fluorescence lifetime of cellular NADPH differs from
that of cellular NADH, reflecting the different set of enzymes to
which the two cofactors bind. Indeed, a substantial literature
documents variations in the fluorescence lifetime of NAD(P)H
in a range of physiological and pathological conditions
13–18
.
However, despite the potential clinical applications of
autofluorescence lifetime measurements, such as delineating the
boundaries of accessible cancers
14
, the biochemical basis for these
variations remains unknown. Moreover, the issue of whether
FLIM may permit separation of NADH and NADPH
fluorescence to resolve cellular specializations and dynamics
in intact tissues is yet to be addressed. In the present study,
we have explored the fluorescence decay properties of NADH and
NADPH in living cells and tissues. Combined with computational
and mathematical modelling, we have found that NAD(P)H
fluorescence lifetime characteristics discriminate between NADH
and NADPH. This provides a unique approach to identify cells
within complex tissues that are enriched in NADPH, thus raising
questions about their metabolic roles and specialization. We have
also analysed the impact of altered metabolic state on NAD(P)H
fluorescence decay characteristics, helping to place changes in
lifetimes observed in transformed neoplastic cells on a firm
biochemical footing.
Results
NAD(P)H fluorescence decay reflects bound NADPH/NADH
ratio. We have previously shown that the fluorescence lifetime
of NADPH is identical to that of NADH in solution
19
,
demonstrating that fluorescence from the free, unbound
pyridine nucleotides cannot be discriminated on the basis of
their fluorescence lifetime. However, it has long been known that
this value increases from 0.3–0.8 to 1–6.5 ns on binding to an
enzyme, depending on the target to which the cofactor binds
10,20
,
as well as by the simultaneous presence of substrate molecules
21
.
Canonical NAD(P)H FLIM studies in live cells and tissues resolve
two fluorescence lifetimes at each pixel; one of the order of 0.4 ns
and the other larger and more variable at B2 ns or more
11,12,22
.
These represent the freely diffusing (t
free
) and enzyme-bound
NAD(P)H (t
bound
) pools, respectively, confirmed by time-
resolved anisotropy imaging studies
23
. As the distribution of
bound NAD(P)H species will inevitably be more heterogeneous
than implied by the single long lifetime, we suggest that t
bound
is a
weighted mean of the fluorescence lifetimes of enzyme-bound
species, a conclusion supported by computational modelling (see
Supplementary Fig. 1a–m and Supplementary Note 1). It is
primarily this parameter that has previously been observed to
vary with changes in metabolism
11,24–26
.
Next, to understand the effect of varying [NADPH]/[NADH]
ratios on t
bound
, we acquired FLIM images of NAD(P)H in live
HEK293 cells in which NADPH levels were genetically and
pharmacologically manipulated. In wild-type cells, the fluores-
cence decays obtained at each pixel required a two-component
model to yield good fits (Fig. 1a,b), as expected, with w
2
r
values
close to unity. As such, addition of further components was not
appropriate, to avoid inaccurate weightings due to overfitting
27
.
Identical fluorescence decay parameters were observed in the
cytosol and mitochondria, with mean (
±
s.d.) values of
t
bound
¼ 2.7
±
0.2 ns, a
bound
¼ 0.19
±
0.01 (fraction of bound
NAD(P)H) and t
free
¼ 0.36
±
0.04 ns (Supplementary Table 1).
Interestingly, t
bound
was significantly smaller in the nucleus
at 2.3
±
0.2 ns (P ¼ 1E 8, two-tailed Student’s t-test, n ¼ 17
images).
To explain reported variations in t
bound
, we hypothesized that,
since NADH and NADPH are associated with different binding
site structures inside the cell
28
, the measured t
bound
value could
reflect the proportion of the two cofactors present. We therefore
measured NAD(P)H fluorescence lifetimes in HEK293 cell lines
in which NAD
þ
kinase (NADK) was either overexpressed
(NADK þ ) or knocked down (NADK )
29
. NADK is the key
determinant of NADPH concentration inside the cell, and its
overexpression results in a 10- to 15-fold higher [NADPH] in
NADK þ relative to NADK cells, leaving [NADH] relatively
unaffected
29
. This was reflected by an NAD(P)H fluorescence
intensity B10-fold brighter in NADK þ compared with
NADK cells (Supplementary Fig. 2b). FLIM revealed that
t
bound
was significantly larger in the NADK þ cells
(t
bound
¼ 3.6
±
0.2 ns in the mitochondria, 3.8
±
0.2 ns in the
cytosol) compared with the NADK cells (t
bound
¼ 2.7
±
0.1 ns
in both subcellular regions, P ¼ 1E 11 and P ¼ 5E 12,
two-tailed Student’s t-test, n ¼ 9, Fig. 1c and Supplementary
Table 1).
These data suggest that increased concentrations of bound
NADPH result in increased values of t
bound
. To verify this
interpretation, we explored the impact of epigallocatechin gallate
(EGCG) on the NAD(P)H FLIM parameters. EGCG is a potent
competitive inhibitor of NADPH binding
30
with no effect
on NADH binding indicated by the BRENDA database
31
.
Preferential competition for NADPH-binding sites by EGCG
would decrease the bound NADPH population, leaving the
bound NADH population unaffected. The value of a
bound
should
therefore be more sensitive to EGCG treatment in NADK þ cells
than in NADK cells, as bound NADPH will form a greater
fraction of the total population of enzyme-bound NAD(P)H
species. As predicted, a
bound
decreased significantly from
0.18
±
0.02 to 0.13
±
0.02 in both the mitochondria and cytosol
ARTICLE NATURE COMMUNICATIONS | DOI: 10.1038/ncomms4936
2 NATURE COMMUNICATIONS | 5:3936 | DOI: 10.1038/ncomms4936 | www.nature.com/naturecommunications
& 2014 Macmillan Publishers Limited. All rights reserved.
(P ¼ 4E 10, two-tailed Student’s t-test, n ¼ 9) on EGCG
treatment in NADK þ cells but remained constant in NADK
cells, confirming the NADPH specificity of this compound (see
Fig. 1d). Exposure to EGCG decreased t
bound
to 3.1
±
0.2 ns
in both the mitochondria and cytosol of NADK þ cells
(P ¼ 2E 12, two-tailed Student’s t-test, n ¼ 9) and to
2.5
±
0.2 ns in their nuclei (P ¼ 3E 7, two-tailed Student’s
t-test, n ¼ 9), and did not significantly affect NADK cells
(Fig. 1c). In addition, treatment with this compound did not
affect t
free
(Supplementary Fig. 2c,d). Thus, these data are
consistent with a model where t
bound
reports the amount of
enzyme-bound NADPH relative to bound NADH.
These data suggest that t
bound
can be used to quantify enzyme-
bound NADPH/NADH ratios. We therefore developed a
numerical model to quantify this phenomenon and generate
predictions (see Supplementary Fig. 1k–o and Supplementary
Note 2). By combining the enzyme-bound NAD(P)H fluores-
cence lifetimes measured in the NADK þ and NADK cells, the
biochemically quantified [NADH] and [NADPH] values in each
cell line
29
and a mathematical model in which NADH and
NADPH were assumed to possess discrete and distinct
fluorescence lifetimes when bound inside the cell, we found
that t
bound
would describe the [NADPH]/[NADH] ratio by
NADPH½
NADH½
¼
t
bound
nsðÞ1:5
4:4 t
bound
nsðÞ
ð1Þ
Application of this model to the NADK þ and NADK data
(see Supplementary Note 3) showed that the concentration of
bound NADH remained constant on EGCG application, while
the concentration of bound NADPH decreased B3-fold with
small differences in each subcellular compartment, supporting
our hypothesis.
Lifetime changes reflect mechanism of metabolic perturbation.
In previous work
11
, shortening of t
bound
observed in tumours has
been attributed to a shift from oxidative to glycolytic metabolism
which occurs in many cancers, the so-called Warburg effect. To
investigate this hypothesis in the light of the results described
here, FLIM images of wild-type HEK293 cells were acquired
following a range of manipulations that alter the balance of ATP
production by aerobic or anaerobic pathways. Dependence on
glycolysis was achieved by inhibition of mitochondrial oxidative
phosphorylation using rotenone (10 mM) or uncoupling using
Mitochondria
NADK+
NADK+
*
*
*
*
*
*
*
*
*
*
*
*
NADK+ NADK+
w/EGCG
NADK–
w/EGCG
NADK–
NADK–
w/EGCG
NADK+ NADK+
w/EGCG
NADK–
NADK–
NADK–
ControlControl
w/EGCGw/EGCG
4.0
2.0
0.3
bound
(ns)
4.0
3.5
3.0
2.5
2.0
0.0
0.0
0.0
0.3
0.2
0.1
Cytosol Nucleus
bound
IRF
Biexponential fit
IRF
Monoexponential fit
Time (ns)
500
0
10
–10
500
0
0
10
–10
0
0
123
4567
Weighted
residual
χ
2
= 1.1 χ
2
= 5.0
Time (ns)
0
123
456
7
Photon counts
Photon counts
Figure 1 | s
bound
reflects the enzyme-bound NADPH/NADH ratio in intact cells. (a,b) A biexponential decay model adequately described the NAD(P)H
fluorescence decay measured in wtHEK293 cells (IRF, instrument response function). The mean w
2
r
was 1.24
±
0.08 compared with 4
±
1 with a
monoexponential fit (representative data from n ¼ 17 experiments). (c,d) Representative colour-coded images and mean t
bound
and a
bound
values in
NADK þ and NADK HEK293 cells prior and following treatment with EGCG (100 mM), a competitive inhibitor of NADPH binding. Scale bar, 20 mm.
Error bars indicate
±
s.d., *Po0.05 (two-tailed Student’s t-test, n ¼ 9).
NATURE COMMUNICATIONS | DOI: 10.1038/ncomms4936 ARTICLE
NATURE COMMUNICATIONS | 5:3936 | DOI: 10.1038/ncomms4936 | www.nature.com/naturecommunications 3
& 2014 Macmillan Publishers Limited. All rights reserved.
FCCP (1 mM). The cells were driven to a more oxidative
phenotype by inhibition of glycolysis by glucose deprivation in
the presence of deoxyglucose (10 mM), while pyruvate (1 mM)
or lactate (10 mM) were provided as mitochondrial-specific
substrates.
Ultraviolet confocal microscopy was used to establish the time
taken for the redox state of the NAD(P) pool to reach a new
steady state following each treatment (Fig. 2a–d). Inhibition of
glycolysis decreased the cellular NAD(P)H fluorescence intensity
by 26
±
6 and 22
±
8% with pyruvate and lactate supplied,
respectively. Rotenone increased steady-state NAD(P)H fluores-
cence intensity by 20
±
5% while FCCP caused a decrease by
38
±
2% (each n ¼ 3). FLIM images were acquired at the steady-
state fluorescence intensity levels following each treatment
(Fig. 2e–g and Supplementary Table 2). Rotenone caused t
bound
to decrease significantly from 2.7
±
0.2 to 2.52
±
0.05 ns in both
the mitochondria and cytosol (P ¼ 6E 4, two-tailed Student’s
t-test, n ¼ 9), suggesting that this treatment caused the concen-
tration of NADH present in the cell to increase relative to the
concentration of NADPH, following inhibition of NADH
oxidation by complex I. In contrast, t
bound
did not change in
response to FCCP (P40.05, two-tailed Student’s t-test, n ¼ 13).
This lack of change in t
bound
suggested that uncoupling caused
the oxidation of both NAD and NADP pools in equal measure.
Increased oxidation of the NAD pool was to be expected on
uncoupling due to the increased complex I activity, and the equal
oxidation of the NADP pool was likely caused by the action of the
mitochondrial transhydrogenase. In respiring mitochondria, this
inner mitochondrial membrane protein transfers hydride from
NADH to NADP
þ
powered by translocation of protons from the
intermembrane space to the mitochondrial matrix
32
. However, on
uncoupling, NADPH is oxidized and the hydride is passed to
Time (min)
Time (min)
NAD(P)H intensity (a.u.)
150
50
0
100
NAD(P)H intensity (a.u.)
150
50
0
100
NAD(P)H intensity (a.u.)
150
50
0
100
NAD(P)H intensity (a.u.)
150
50
0
100
Time (min)
Time (min)
Glucose Rotenone
FCCP Pyruvate Lactate
(Glucose deprivation)
4.0
2.0
bound
(ns)
bound
(ns)
bound
bound
0.3
0.0
Mitochondria Cytosol Nucleus
*
0.2
0.3
0.1
0.0
*
*
*
*
Glucose Rotenone FCCP Pyruvate Lactate
(Glucose deprivation)
Glucose Rotenone FCCP Pyruvate Lactate
(Glucose deprivation)
*
*
*
*
*
*
*
*
*
Rotenone
Pyruvate
0246810
02468
10
0246810
0246810
Lactate
FCCP
0.0
2.0
2.5
3.0
4.0
3.5
Figure 2 | NAD(P)H fluorescence decay responses to metabolic perturbation are mechanism dependent. (ad) Time series of NAD(P)H
fluorescence intensity following treatments chosen to perturb oxidative (a: respiratory chain inhibition by rotenone, b: uncoupling with FCCP) and glycolytic
metabolism (glucose replaced by deoxyglucose with c: pyruvate or d: lactate supplied as substrate). (e) Colour-coded images and (f,g) quantification
of changes in t
bound
and a
bound
on application of treatment. Scale bar, 20 mm. Error bars indicate
±
s.d., *Po0.05 (two-tailed Student’s t-test, n ¼ 9).
ARTICLE NATURE COMMUNICATIONS | DOI: 10.1038/ncomms4936
4 NATURE COMMUNICATIONS | 5:3936 | DOI: 10.1038/ncomms4936 | www.nature.com/naturecommunications
& 2014 Macmillan Publishers Limited. All rights reserved.
NAD
þ
to produce NADH
33
. Relative to controls, glucose
deprivation in the presence of either pyruvate or lactate supply
failed to significantly alter t
bound
(P40.05, two-tailed Student’s
t-test, both n ¼ 9). However, the mean value of t
bound
in both the
mitochondria and cytosol of 2.70
±
0.01 ns in the presence of
pyruvate was significantly larger than its value of 2.57
±
0.04 ns in
the presence of lactate (P ¼ 7E 6, two-tailed Student’s t-test,
n ¼ 9). As lactate dehydrogenase promotes NADH production
during the conversion of lactate to pyruvate, the significantly
smaller t
bound
value in the presence of lactate relative to pyruvate
supports the interpretation developed in this work that this
parameter reflects NADPH/NADH ratio, assuming NADPH
production was identical in the presence of the two substrates.
While changes in t
bound
, and thus the NADPH/NADH ratio,
reflected the specific treatment causing a defect in OXPHOS
or glycolysis, we observed that each treatment causing a net
oxidation of the combined NAD(P) pools caused a
bound
to
increase (Fig. 2g). This is in support of previous suggestions that
this parameter, reflecting the enzyme-bound population fraction
of NAD(P)H, reports acute changes in the redox state of the
cell
25
. However, confirming whether this parameter reflects
phenotypic differences in the redox state of different cell types
requires further investigation. Interestingly, each oxidizing
treatment also caused an increase in t
free
(Supplementary
Fig. 2e,f). However, as these lifetimes lie close to the time
resolution of the FLIM system, such small differences may be an
artefact of the fitting process, such as an interdependency between
t
free
and a
bound
(Supplementary Fig. 1j).
The lack of significant change in t
bound
in response to FCCP
treatment or replacement of glucose with pyruvate or lactate,
along with the very small, if significant, change in this parameter
following rotenone treatment relative to the pathophysiological
variations reported in the literature, suggests that changes in
NAD(P)H fluorescence decay cannot be simply attributed to
alterations in the balance between oxidative and glycolytic
metabolism. Accordingly, no correlation between t
bound
and the
balance of ATP production by glycolytic or oxidative means was
observed by measuring the rates of oxygen consumption and
lactate release in the wild-type, NADK þ and NADK HEK293
cell lines (see Supplementary Fig. 3 and Supplementary Tables 3
and 4). Altogether, these data strongly suggest that any variation
in the NAD(P)H fluorescence decay parameter t
bound
will be
specific to the mechanism of metabolic perturbation, such as
mitochondrial dysfunction caused either by respiratory chain
inhibition or uncoupling, or the utilization of different substrates
following inhibition of glycolysis. Such specificity has the
potential to aid the study of the variety of possible metabolic
rearrangements that may occur in cancer and other pathologies.
FLIM reveals metabolic variations in complex tissues. The
abundance of NADPH and NADH in live cells can be measured
Untreated
OHCs
Intensity Intensity
OHCs
OHCs
OHCs
OHCs
OHCs
OPCs
OPCs
OPCs
OPCs
OHCs
OPCs
OPC
OPCs
w/EGCG
w/EGCG
*
*
*
*
*
bound
(ns)
4.0
3.0
2.0
2.0 4.0 4.02.0
MCB fluorescence (a.u.)
0.0
0.5
1.0
1.5
2.0
*
bound
0.0
0.1
0.2
0.3
0.0
bound
(ns)
bound
(ns)
Figure 3 | Supporting cells in the mammalian cochlea exhibit increased enzyme-bound NADPH. (a) Mean t
bound
and (b) a
bound
values in outer hair
cells (OHC’s) and adjacent outer pillar ‘supporting’ cells (OPC’s) under control conditions and following application of EGCG (200 mM). Error bars
indicate
±
s.d., *Po0.05 (two-tailed Student’s t-test, n ¼ 11). (c,d) Corresponding representative FLIM images colour coded for the mean parameter
value in each cell. Scale bar, 25 mm. ( e) Schematic diagram showing organ of Corti in the cochlear explants, indicating the positions of OHC’s and OPC’s.
(f) Mean fluorescence intensity in OPC’s and OHC’s. Error bars indicate
±
s.d., *Po0.05 (Wilcoxon signed-rank test, n ¼ 17). (g) Representative image
following MCB staining for GSH concentration.
NATURE COMMUNICATIONS | DOI: 10.1038/ncomms4936 ARTICLE
NATURE COMMUNICATIONS | 5:3936 | DOI: 10.1038/ncomms4936 | www.nature.com/naturecommunications 5
& 2014 Macmillan Publishers Limited. All rights reserved.
in tissues by biochemical means such as high performance liquid
chromatography
29,34
. However, a microscopy-based imaging
approach permits measurement of NADH and NADPH in
discrete compartments within a cell or in different cell types
within a complex tissue. Indeed, we have observed consistently
smaller values of t
bound
in the nucleus compared with the rest of
the cell. Combining enzyme localization data
31
with the analysis
and model provided here, NAD(P)H FLIM indicates that nuclear
glucose metabolism favours NADH-producing glycolysis rather
than NADPH-producing pentose phosphate pathway (PPP),
resulting in a more oxidized NADP pool in this region (see
Supplementary Note 4).
To explore the application of NAD(P)H FLIM in multicellular
specimens, the redox metabolism of the mammalian cochlea was
investigated. The cochlea is a structure containing a highly
ordered system of functionally distinct cell types, which are easily
identifiable from their architecture (Fig. 3e). FLIM imaging of
cochlear explant cultures revealed that t
bound
differed markedly
between different cell types. The glia-like outer pillar ‘supporting’
cells (OPCs) exhibited an extended fluorescence lifetime of
3.5
±
0.1 ns, while the mean value of t
bound
measured in the
neighbouring sensory receptors, the outer hair cells (OHCs),
was significantly lower at 2.9
±
0.1 ns (P ¼ 2E 15, two-tailed
Student’s t-test, n ¼ 19, Fig. 3a–d). EGCG treatment decreased
t
bound
in supporting cells by 0.6
±
0.2 ns (P ¼ 1E 6, two-tailed
Student’s t-test, n ¼ 11) to a mean of 2.9
±
0.2 ns, but by only
0.2
±
0.2 ns (P ¼ 8E 4, two-tailed Student’s t-test, n ¼ 11) to a
mean of 2.7
±
0.1 ns in hair cells (Supplementary Table 5).
These data confirm that the differences in fluorescence lifetime
between cell types were a measure of differences in the bound
NADPH/NADH ratio.
Application of the mathematical model developed in this work
(Supplementary equations 10 and 11) implied that the total
concentration of reduced pyridine nucleotides were equal in the
OHCs and OPCs. In the OHCs, the absolute concentrations of
NADH and NADPH were equal at 1
±
0.2 a.u. However, in the
OPCs, the concentration of NADPH was significantly larger at
1.4
±
0.3 a.u. (P ¼ 4E 4, Wilcoxon signed-rank test, n ¼ 19),
with a correspondingly smaller concentration of NADH (0.7
±
0.2
a.u., P ¼ 2E 4, Wilcoxon signed-rank test, n ¼ 19). To assess the
functional significance of this finding, we measured the distribu-
tion of GSH in the tissue using monochlorobimane (MCB)
staining, since a major demand for NADPH in glial cells arises
from GSH turnover
35
. The glia-like OPCs showed significantly
higher [GSH] compared with OHCs (P ¼ 0.02, Wilcoxon signed-
rank test, n ¼ 17, Fig. 3f,g), consistent with a functional
requirement for NADPH enrichment. Thus, our interpretation
of t
bound
is consistent with alternative indicators of cellular redox
state.
Discussion
This work shows that FLIM can be used effectively to differentiate
between NADH and NADPH at the level of the single cell or
organelle. The results suggest that enzyme-bound NADPH
possesses a significantly larger fluorescence lifetime than
enzyme-bound NADH within the cellular environment, so that
the proportion of enzyme-bound NADPH and NADH present in
live tissue determines the lifetime of their combined fluorescence
decay. By making the simplifying assumption that bound NADH
and bound NADPH possess finite and distinct fluorescence
lifetimes inside the cell, the relative contribution of each cofactor
to the combined fluorescence signal could be calculated. With
excitation at 700 nm and a 435–485 nm detection window, the
intracellular fluorescence lifetimes of NADH and NADPH
were predicted to be 1.5
±
0.2 and 4.4
±
0.2 ns respectively. It is
reasonable to hypothesize that these conclusions can be extended
to all cell types as the fluorescence lifetime of NAD(P)H when
bound to an enzyme is determined by its local environment in the
binding site
21
, and the NADH and NADPH-binding sites are two
of the most highly conserved in all biology
36,37
. Indeed, we have
observed values of t
bound
that are similar in magnitude across a
range of cell types, including isolated ventricular cardiomyocytes
and neurons in culture or in brain slices (data not shown).
The analysis and model presented herein are consistent with
previously published NAD(P)H FLIM studies. In 2008, Niesner
et al.
38
performed a novel study on the decay of NAD(P)H
fluorescence in granulocytes in the presence of Aspergillus
fumigatus fungus. The parameter t
bound
was B2 ns within
the bulk cytosol of the granulocytes. However, localized
subplasmalemmal regions of the cytosol in contact with the
fungus displayed increased values of 3.7 ns. This was attributed to
a unique fluorescence lifetime of NADPH when bound to the
NADPH oxidase activated in response to pathogenic exposure.
However, the computational simulations performed here showed
that t
bound
is a weighted average of the fluorescence lifetimes of
the enzyme-bound NAD(P)H species present. For NADPH
oxidase alone to cause an increase in t
bound
as large as that
observed in the granulocytes would require this enzyme to be
present at a greater concentration than the NADH-binding
enzymes of the cytosol. The Model Organism Protein Expression
Database (MOPED)
39
within the GeneCards human gene
compendium
40
shows that expression of glyceraldehyde
3-phosphate dehydrogenase outweighs that of the NADPH-
binding subunit of NADPH oxidase (neutrophil cytosolic factor
2)
41
by around 100 to one in neutrophils. It is therefore more
likely that the NAD(P)H fluorescence lifetime observed in the
regions of the cytosol where NADPH oxidase was activated was
due to increased local NADPH production, as implied from the
results reported here. Indeed, the large quantities of superoxide
produced by NADPH oxidase requires plentiful supply of
NADPH. Activation of this enzyme is thus associated with
increased flux of glucose through the PPP
42
.
Changes in the fluorescence decay of NAD(P)H have recently
been observed in applications ranging from wound healing
17
to
stem cell differentiation
18
and necrotic deterioration of skin
15
to
staurosporine-induced apoptosis
16
. The large number of studies
reporting differences between the fluorescence lifetime of
NAD(P)H in healthy control cells and cells at different stages
of carcinogenesis
11,43,44
have prompted the design of clinical
instruments for the detection, diagnosis and staging of
accessible tumours using time-resolved autofluorescence
measurements
14,45–50
. While the Warburg effect is the most
well known of the metabolic shifts occurring during cancer
development, tumorigenesis is also associated with variations in
glucose flux through the PPP and various biosynthetic pathways
utilizing NADPH
51
. As none of the severe pharmacological
perturbations to cytosolic or mitochondrial ATP production
applied in this work could reproduce responses in t
bound
of the
magnitude caused by an increased NADPH/NADH ratio, our
results strongly suggest that NAD(P)H fluorescence lifetime
differences observed between healthy and pathological states
reflect shifts in the NADPH/NADH balance.
In this work, termination of oxidative phosphorylation by
uncoupling and of glycolysis by glucose deprivation could not be
resolved on the basis of the fluorescence decay parameters
measured under these conditions. In addition, two of the cell lines
studied here with similar fluorescence decay parameters were
shown to differ in their reliance on aerobic and anaerobic
metabolism (wild-type HEK293 and NADK ). This implies that
changes in the NAD(P)H fluorescence decay do not simply report
shifts between an oxidative or glycolytic phenotype, but reflect the
ARTICLE NATURE COMMUNICATIONS | DOI: 10.1038/ncomms4936
6 NATURE COMMUNICATIONS | 5:3936 | DOI: 10.1038/ncomms4936 | www.nature.com/naturecommunications
& 2014 Macmillan Publishers Limited. All rights reserved.
differential response of the NAD and NADP pools following a
metabolic alteration. For example, both uncoupling and electron
transport chain inhibition terminate aerobic ATP production yet
have opposing effects on the overall NAD(P) redox state and
thus, perhaps on a
bound
. Changes in t
bound
will be induced by
metabolic transitions that cause divergent effects on NAD- and
NADP-associated pathways. Thus, neither parameter alone will
permit the detection of oxidative or glycolytic states using
NAD(P)H fluorescence decay measurements. However, a method
for assessing the pathways used for ATP production by
measuring autofluorescence may be possible by combining a
number of observables, including a
bound
and t
bound
, alongside
fluorescence intensity information and similar measurements
from flavoprotein fluorescence as part of a spectrally resolved
lifetime imaging tool
24
.
In the present study, we have investigated the canonical form
of FLIM in which fluorescence lifetimes t
free
and t
bound
, along
with the weighting parameter a
bound
, are extracted by least-
squares fitting of a biexponential decay at each pixel
11,24–26
.
Drawbacks of this method of analysis have been identified, such
as the computational burden of fitting 10
4
–10
5
independent decay
curves in a single image, difficulties in resolving decay
components with closely spaced lifetimes and correlations
between the fit parameters
52
. Other groups have therefore
focussed on developing novel analytical techniques. For
example, Yaseen et al.
53
recently applied global analysis to the
application of NAD(P)H FLIM of the rat cortex. In this approach,
the lifetimes of four decay components were shared across
the image and a novel computational algorithm recovered the
optimum amplitudes of each species to describe the fluorescence
decay at each pixel. Interestingly, two enzyme-bound components
were identified in the tissue with lifetimes of 1.7 and 3.2 ns,
perhaps corresponding to enzyme-bound NADH and NADPH.
Another approach to which our conclusions may be strongly
applicable is the phasor method developed by Digman et al.
52
Here, the real and imaginary parts of the Fourier transform of the
fluorescence decay at each pixel of a FLIM image define the
coordinates of a point in a two-dimensional phase space.
The relative abundance of two or more fluorescent species with
different lifetimes can then be inferred from the location of the
pixels in the phasor plot. Application of this ‘fit-free’ approach to
separating NADH and NADPH fluorescence will be the subject of
further work.
Altogether, we have shown that the fluorescence lifetime
characteristics of NAD(P)H in live cells and tissues can be used to
discriminate between NADH and NADPH fluorescence, provid-
ing, for the first time, a biochemical framework for interpretation
of NAD(P)H FLIM studies. Such a technique permits the
separation of NADH and NADPH redox signalling without
disrupting the sample on the addition of external probes
54,55
,
allowing complex tissue preparations to be investigated. The
approach revealed previously unknown cellular metabolic
specializations in the mammalian cochlea, highlighting a
subpopulation of cells characterized by high NADPH levels,
opening up new avenues of research to understand the functional
significance of redox pathways with respect to the physiological
roles of these cells.
Methods
Cell culture. HEK293 cells were obtained from the American Type Culture
Collection and grown in advanced Dulbecco’s modified Eagle Medium (Gibco)
supplemented with 10% fetal bovine serum, 2 mM GlutaMAX, 100 U ml
1
penicillin and 100 mgml
1
streptomycin (Gibco). Production of the NADK þ and
NADK cell lines has been reported previously
29
. These cultures were
additionally supplemented with 0.1 mg ml
1
G-418 selective antibiotic (Gibco).
All cells were grown as monolayers in sterile 75 cm
2
tissue culture flasks (Thermo
Scientific Nunc) in a 37 °C, 5% CO
2
incubator.
Cochlea explant cultures
. Cochlear coils were isolated from male and female
post-natal day 2–3 Sprague Dawley rats as previously described
56
. Briefly, auditory
bullae were removed and transferred into Medium 199 with Hank’s balanced salts
(Life Technologies). The cartilaginous wall of the bulla was opened and the whole
cochlea extracted. The stria vascularis and Reissner’s membrane were removed and
the cochlea cut into three coils. The cochlear coils were placed onto Cell-Tak cell
and tissue adhesive (BD Biosciences)-coated dishes (MatTek). For coating, cell
adhesive was diluted to 70 mgml
1
in 0.1 mM NaHCO
3
. The cochlear explants
were incubated overnight in DMEM/F12 (Gibco), supplemented with 1% fetal
bovine serum (Life Technologies) in a 37 °C, 5% CO
2
incubator. The isolation was
performed in accordance with the United Kingdom Animals (Scientific
Procedures) Act of 1986 and in compliance with the Biological Services
Management Group and the Biological Services Ethical Committee, University
College London.
Live-cell microscopy
. On the microscope stage, coverslips containing 300,000 cells
were maintained at 37 °C in a metal ring and bathed in DMEM solution at pH 7.4
containing 10 mM HEPES and 2 mM GlutaMAX. Glucose (25 mM) was present
under contro l conditions and during EGCG treatment (100 mM in cell lines,
200 mM for cochlea). Pharmacological perturbations to metabolism were applied by
the dropwise addition of working concentrations of each compound, diluted from
stock solutions in DMEM recording medium.
NAD(P)H FLIM was performed on an upright LSM 510 microscope (Carl
Zeiss) with a 1.0 NA 40 water-dipping objective using a 650-nm short-pass
dichroic and 460
±
25 nm emission filter. Two-photon excitation was provided by a
Chameleon (Coherent) Ti:sapphire laser tuned to 700 nm, with on-sample powers
kept below 10 mW. Spectral controls (see Supplementary Fig. 4 and Supplementary
Note 5) confirmed the NAD(P)H specificity of this excitation wavelength and
emission filtering. Photodamage controls (see Supplementary Tables 6 and 7 and
Supplementary Note 6) demonstrated that FLIM parameters were not varying over
the course of imaging. Emission events were registered by an external detector
(HPM-100, Becker & Hickl) attached to a commercial time-correlated single
photon counting electronics module (SPC-830, Becker & Hickl) contained on a
PCI board in a desktop computer. Scanning was performed continuously for 4 min
with a pixel dwell time of 1.6 ms. To identify mitochondrial and cytosolic regions,
the mitochondrially targeted fluorescent dye tetramethylrhodamine methyl ester
(TMRM) was added to the recording medium, at a final concentra tion of 25 nM,
20 min before imaging. TMRM fluorescence was collected for a 10-s burst using a
610
±
30 nm emission filter. Excitation was provided at the same wavelength as
NAD(P)H to avoid possible chromatic aberration. The 585
±
15 nm emission
spectrum of TMRM ensured its fluorescence did not contaminate the NAD(P)H
FLIM images.
NAD(P)H fluorescence intensity time series and MCB imaging were performed
on an inverted LSM 510 laser scanning confocal microscope (Carl Zeiss) with
351 nm illumination from an argon ion laser (Coherent Enterprise UV). MCB
(30 min loading at 50 mM) and NAD(P)H fluorescence were observed using a
351-nm long-pass dichroic and 460
±
25 nm band-pass emission filter with a 40,
1.3 NA quartz oil immersion objective. Images (12-bit 512 512) were obtained
with a pixel dwell time of 1.6 ms. Time series measurements were obtained at 1 min
intervals. To reduce noise, the image recorded at each time point was a mean of
four consecutive scans. Fluorescence intensity levels were extracted using ImageJ
(NIH).
Metabolic controls
. For the measurement of oxygen consumption, cells were
trypsinized and resuspended at B1 million cells per ml in DMEM buffered with
10 mM HEPES and supplemented with 25 mM glucose and 2 mM GlutaMAX.
Respiration rates were measured in triplicate at 37 °C with the high-resolution
Oxygraph (Oroboros Oxygraph-2k). State 4 respiration values were obtained in the
presence of 2.5 mM oligomycin, maximal oxidative capacities were determined in
the presence of 2 mM FCCP and non-mitochondrial background oxygen con-
sumption was determined in the presence of 2.5 mM antimycin A. Lactate release
rates were measured in triplicate by removing a sample of serum-free DMEM from
70% confluent cell cultures at 1 h intervals for 5 h total and determining the
concentration of lactate present using a commercially available plate-reader assay
(Sigma Aldrich) in absorption mode. For normalization of both oxygen
consumption and lactate release rates, final cell counts were performed using a
haemocytometer.
FLIM data analysis
. Following 5 5 binning of photon counts at each pixel,
fluorescence decay curves of the form
ItðÞ¼Z þ I
0
1 a
bound
ðÞexp
t
t
free

þ a
bound
exp
t
t
bound

ð2Þ
were fit to the FLIM images using iterative reconvolution in SPCImage (Becker &
Hickl), where Z allows for time-uncorrelated background noise. The instrument
response function of the FLIM system was obtained by measuring the fluorescence
decay profile of second harmonic generation by a potassium dihydrogen phosphate
(KDP) crystal at 920 nm, grown by leaving a molar solution of KDP in water on a
coverslip to evaporate. Matrices of the fit parameters t
free
, a
bound
and t
bound
, along
NATURE COMMUNICATIONS | DOI: 10.1038/ncomms4936 ARTICLE
NATURE COMMUNICATIONS | 5:3936 | DOI: 10.1038/ncomms4936 | www.nature.com/naturecommunications 7
& 2014 Macmillan Publishers Limited. All rights reserved.
with the total photons counted at each pixel, were exported from SPCImage.
In MATLAB (The Mathworks), a 16-bit grayscale image was produced for each
parameter matrix in which the intensity of each pixel was proportional to the
parameter value at that location. ImageJ was then used to measure the grayscale
intensity in the parameter images and the values obtained converted back to
parameter values using the scaling factors applied in their production. Masks
identifying the location of mitochondrial and cytosolic pixels were created by
importing images of the TMRM distribution in the cell and a nuclear mask was
defined by hand, allowing the parameters describing the fluorescence decay of
NAD(P)H in each of these regions to be extracted separately (Supplementary
Fig. 2a).
Statistical analysis
. NAD(P)H fluorescence lifetime parameters obtained using
FLIM are reported as a mean over at least three regions of at least three separate
cultures. Uncertainties in these values were taken as the s.d. of the measurements.
Differences between the lifetime parameters measured under different conditions
were tested for statistical significance (Po0.05) using a two-tailed Student’s t-test.
Statistically significant differences between data sets normalized to be expressed in
relative arbitrary units (fluorescence intensity, NADH/NADPH concentration)
were assessed using a Wilcoxon signed-rank test. For the estimation of uncer-
tainties in the predictions of the numerical model, the standard formula for the
calculation of the error s
2
z
in an arbitrary function Z ¼ Z(A, B, C,...) was applied,
s
2
z
¼
@Z
@A

2
s
2
A
þ
@Z
@B

2
s
2
B
þ
@Z
@C

2
s
2
C
::: ð3Þ
References
1. Collins, Y. et al. Mitochondrial redox signalling at a glance. J. Cell Sci. 125,
801–806 (2012).
2. Ying, W. NAD þ /NADH and NADP þ /NADPH in cellular functions and cell
death: regulation and biological consequences. Antioxid. Redox Signal. 10,
179–206 (2008).
3. Mailloux, R., Lemire, J. & Appanna, V. Metabolic networks to combat oxidative
stress in Pseudomonas fluorescens. Antonie Van Leeuwenhoek 99, 433–442
(2011).
4. Rocheleau, J. V., Head, W. S. & Piston, D. W. Quantitative NAD(P)H/
flavoprotein autofluorescence imaging reveals metabolic mechanisms of
pancreatic islet pyruvate response. J. Biol. Chem. 279, 31780–31787 (2004).
5. Chance, B., Schoener, B., Oshino, R., Itshak, F. & Nakase, Y. Oxidation-
reduction ratio studies of mitochondria in freeze-trapped samples. NADH and
flavoprotein fluorescence signals. J. Biol. Chem. 254, 4764–4771 (1979).
6. Patterson, G. H., Knobel, S. M., Arkhammar, P., Thastrup, O. & Piston, D. W.
Separation of the glucose-stimulated cytoplasmic and mitochondrial NAD(P)H
responses in pancreatic islet beta cells. Proc. Natl Acad. Sci. USA 97, 5203–5207
(2000).
7. De Ruyck, J. et al. Towards the understanding of the absorption spectra of
NAD(P)H/NAD(P) þ as a common indicator of dehydrogenase enzymatic
activity. Chem. Phys. Lett. 450, 119–122 (2007).
8. Duchen, M. R., Surin, A. & Jacobson, J. Imaging mitochondrial function in
intact cells. Methods Enzymol. 361, 353–389 (2003).
9. Spat, A. & Pitter, J. G. The effect of cytoplasmic Ca2 þ signal on the redox state
of mitochondrial pyridine nucleotides. Mol. Cell. Endocrinol. 215, 115–118
(2004).
10. Vishwasrao, H. D., Heikal, A. A., Kasischke, K. A. & Webb, W. W.
Conformational dependence of intracellular NADH on metabolic state revealed
by associated fluorescence anisotropy. J. Biol. Chem. 280, 25119–25126 (2005).
11. Skala, M. C. et al. In vivo multiphoton fluorescence lifetime imaging of protein-
bound and free nicotinamide adenine dinucleotide in normal and precancerous
epithelia. J. Biomed. Opt. 12, 24014 (2007).
12. Yu, Q. & Heikal, A. A. Two-photon autofluorescence dynamics imaging reveals
sensitivity of intracellular NADH concentration and conformation to cell
physiology at the single-cell level. J. Photochem. Photobiol. B 95, 46–57 (2009).
13. Evans, N. D., Gnudi, L., Rolinski, O. J., Birch, D. J. S. & Pickup, J. C. Glucose-
dependent changes in NAD(P)H-related fluorescence lifetime of adipocytes and
fibroblasts in vitro: potential for non-invasive glucose sensing in diabetes
mellitus. J. Photochem. Photobiol. B 80, 122–129 (2005).
14. Butte, P. V. et al. Intraoperative delineation of primary brain tumors using
time-resolved fluorescence spectroscopy. J. Biomed. Opt. 15, 27008 (2010).
15. Sanchez, W. Y., Prow, T. W., Sanchez, W. H., Grice, J. E. & Roberts, M. S.
Analysis of the me tabolic deterioration of ex vivo skin from ischemic necrosis
through the imaging of intracellular NAD(P)H by multiphoton tomography
and fluorescence lifetime imaging microscopy. J. Biomed. Opt. 15, 46008–46011
(2010).
16. Yu, J.-S., Guo, H.-W., Wang, C.-H., Wei, Y.-H. & Wang, H.-W. Increase of
reduced nicotinamide adenine dinucleotide fluorescence lifetime precedes
mitochondrial dysfunction in staurosporine-induced apoptosis of HeLa cells.
J. Biomed. Opt. 16, 36008 (2011).
17. Gehlsen, U. et al. Two-photon fluorescence lifetime imaging monitors
metabolic changes during wound healing of corneal epithelial cells in vitro.
Graefes Arch. Clin. Exp. Ophthalmol. 250, 1293–1302 (2012).
18. Stringari, C. et al. Metabolic trajectory of cellular differentiation in small
intestine by Phasor Fluorescence Lifetime Microscopy of NADH. Sci. Rep. 2,
568 (2012).
19. Blacker, T. S., Marsh, R. J., Duchen, M. R. & Bain, A. J. Activated barrier
crossing dynamics in the non-radiative decay of NADH and NADPH. Chem.
Phys. 422, 184–194 (2013).
20. Scott, T. G., Spencer, R. D., Leonard, N. J. & Weber, G. Synthetic spectroscopic
models related to coenzymes and base pairs. V. Emission properties of NADH.
Studies of fluorescence lifetimes and quantum efficiencies of NADH,
AcPyADH, [reduced acetylpyridineadenine dinucleotide] and simplified
synthetic models. J. Am. Chem. Soc. 92, 687–695 (1970).
21. Gafni, A. & Brand, L. Fluorescence decay studies of reduced nicotinamide
adenine dinucleotide in solution and bound to liver alcohol dehydrogenase.
Biochemistry 15, 3165–3171 (1976).
22. Wang, H. W. et al. Differentiation of apoptosis from necrosis by dynamic
changes of reduced nicotinamide adenine dinucleotide fluorescence lifetime in
live cells. J. Biomed. Opt. 13, 054011 (2008).
23. Zheng, W., Li, D. & Qu, J. Y. Monitoring changes of cellular metabolism and
microviscosity in vitro based on time-resolved endogenous fluorescence and its
anisotropy decay dynamics. J. Biomed. Opt. 15, 37011–37013 (2010).
24. Skala, M. C. et al. In vivo multiphoton microscopy of NADH and FAD redox
states, fluorescence lifetimes, and cellular morphology in precancerous
epithelia. Proc. Natl Acad. Sci. USA 104, 19494–19499 (2007).
25. Bird, D. K. et al. Metabolic mapping of MCF10A human breast cells via
multiphoton fluorescence lifetime imaging of the coenzyme NADH. Cancer
Res. 65, 8766–8773 (2005).
26. Pradhan, A. et al. Steady state and time-resolved fluorescence properties
of metastatic and non-metastatic malignant cells from different species.
J. Photochem. Photobiol. B 31, 101–112 (1995).
27. Grinvald, A. & Steinberg, I. Z. On the analysis of fluorescence decay kinetics by
the method of least-squares. Anal. Biochem. 59, 583–598 (1974).
28. Scrutton, N. S., Berry, A. & Perham, R. N. Redesign of the coenzyme specificity
of a dehydrogenase by protein engineering. Nature 343, 38–43 (1990).
29. Pollak, N., Niere, M. & Ziegler, M. NAD kinase levels control the NADPH
concentration in human cells. J. Biol. Chem. 282, 33562–33571 (2007).
30. Shin, E. S. et al. Catechin gallates are NADP þ -competitive inhibitors of
glucose-6-phosphate dehydrogenase and other enzymes that employ NADP þ
as a coenzyme. Bioorg. Med. Chem. 16, 3580–3586 (2008).
31. Schomburg, I. et al. BRENDA: a resource for enzyme data and metabolic
information. Trends Biochem. Sci. 27, 54–56 (2002).
32. Huxley, L., Quirk, P. G., Cotton, N. P. J., White, S. A. & Jackson, J. B.
The specificity of proton-translocating transhydrogenase for nicotinamide
nucleotides. Biochim. Biophys. Acta 1807, 85–94 (2011).
33. Enander, K. & Rydstrom, J. Energy-linked nicotinamide nucleotide
transhydrogenase. Kinetics and regulation of purified and reconstituted
transhydrogenase from beef heart mitochondria. J. Biol. Chem. 257,
14760–14766 (1982).
34. Klaidman, L. K., Leung, A. C. & Adams, J. D. High-performance liquid
chromatography analysis of oxidized and reduced pyridine dinucleotides in
specific brain regions. Anal. Biochem. 228, 312–317 (1995).
35. Keelan, J., Allen, N. J., Antcliffe, D., Pal, S. & Duchen, M. R. Quantitative
imaging of glutathione in hippocampal neurons and glia in culture using
monochlorobimane. J. Neurosci. Res. 66, 873–884 (2001).
36. Rossmann, M. G., Moras, D. & Olsen, K. W. Chemical and biological evolution
of a nucleotide-binding protein. Nature 250, 194–199 (1974).
37. Carugo, O. & Argos, P. NADP-dependent enzymes. I: conserved
stereochemistry of cofactor binding. Proteins 28, 10–28 (1997).
38. Niesner, R. et al. Selective detection of NADPH oxidase in polymorphonuclear
cells by means of NAD(P)H-based fluorescence lifetime imaging. J. Biophys.
2008, 602639 (2008).
39. Kolker, E. et al. MOPED: Model Organism Protein Expression Database.
Nucleic Acids Res. 40, D1093–D1099 (2012).
40. Rebhan, M., Chalifa-Caspi, V., Prilusky, J. & Lancet, D. GeneCards: integrating
information about genes, proteins and diseases. Trends Genet. 13, 163 (1997).
41. Shatwell, K. P. & Segal, A. W. NADPH oxidase. Int. J. Biochem. Cell Biol. 28,
1191–1195 (1996).
42. Iyer, G. Y. N., Islam, M. F. & Quastel, J. H. Biochemical aspects of phagocytosis.
Nature 192, 535–541 (1961).
43. Tadrous, P. J. et al. Fluorescence lifetime imaging of unstained tissues: early
results in human breast cancer. J. Pathol. 199, 309–317 (2003).
44. Sud, D., Zhong, W., Beer, D. G. & Mycek, M. A. Time-resolved optical imaging
provides a molecular snapshot of altered metabolic function in living human
cancer cell models. Opt. Express 14, 4412–4426 (2006).
45. Colasanti, A. et al. MS-2 fibrosarcoma characterization by laser induced
autofluorescence. Lasers Surg. Med. 26, 441–448 (2000).
ARTICLE NATURE COMMUNICATIONS | DOI: 10.1038/ncomms4936
8 NATURE COMMUNICATIONS | 5:3936 | DOI: 10.1038/ncomms4936 | www.nature.com/naturecommunications
& 2014 Macmillan Publishers Limited. All rights reserved.
46. Butte, P. V. et al. Diagnosis of meningioma by time-resolved fluorescence
spectroscopy. J. Biomed. Opt. 10, 64026–64029 (2005).
47. De Beule, P. A. et al. A hyperspectral fluorescence lifetime probe for skin cancer
diagnosis. Rev. Sci. Instrum. 78, 123101 (2007).
48. Sun, Y. et al. Fluorescence lifetime imaging microscopy: in vivo application to
diagnosis of oral carcinoma. Opt. Lett. 34, 2081–2083 (2009).
49. Sun, Y. et al. Fluorescence lifetime imaging microscopy for brain tumor image-
guided surgery. J. Biomed. Opt. 15, 56022 (2010).
50. McGinty, J. et al. Wide-field fluorescence lifetime imaging of cancer. Biomed.
Opt. Express 1, 627–640 (2010).
51. Vander Heiden, M. G., Cantley, L. C. & Thompson, C. B. Understanding the
Warburg effect: the metabolic requirements of cell proliferation. Science 324,
1029–1033 (2009).
52. Digman, M. A., Caiolfa, V. R., Zamai, M. & Gratton, E. The phasor approach to
fluorescence lifetime imaging analysis. Biophys. J. 94, L14–L16 (2008).
53. Yaseen, M. A. et al. In vivo imaging of cerebral energy metabolism with
two-photon fluorescence lifetime microscopy of NADH. Biomed. Opt. Express
4, 307–321 (2013).
54. Zhao, Y. et al. Genetically encoded fluorescent sensors for intracellular NADH
detection. Cell Metab. 14, 555–566 (2011).
55. Hung, Y. P., Albeck, J. G., Tantama, M. & Yellen, G. Imaging cytosolic
NADH-NAD þ redox state with a genetically encoded fluorescent biosensor.
Cell Metab. 14, 545–554 (2011).
56. Mann, Z. F., Duchen, M. R. & Gale, J. E. Mitochondria modulate the
spatio-temporal properties of intra- and intercellular Ca2 þ signals in cochlear
supporting cells. Cell Calcium 46, 136–146 (2009).
Acknowledgements
We thank Dr Will Kotiadis for assistance with respirometry measurements, Dr Zhi Yao
for assistance with the lactate release assay and Dr Laura Osellame for proofreading
of the manuscript. Action on Hearing Loss supported the work of Z.F.M. and J.E.G.
G.S. was funded by Parkinson’s UK, British Heart Foundation, Wellcome Trust-UCL
Therapeutic Innovation Fund, Telethon (Italy, GEP1206) and the Italian Association for
Cancer Research (AIRC). M.R.D. acknowledges strategic fu nding from the Wellcome
Trust/MRC Joint Call in Neurodegeneration Award (WT089698). T.S.B. is grateful
for the award of an EPSRC studentship and the inaugural Professor Anne Warner
postdoctoral fellowship through the CoMPLEX Doctoral Training Centre at UCL.
Authors contributions
M.Z. provided the NADK þ and NADK– cell lines. J.E.G. prepared cochlea cultures and
oversaw their imaging. Z.F.M., G.S. and M.R.D. performed preliminary experiments.
T.S.B. performed the experiments, data analysis and modelling. A.J.B., G.S. and M.R.D.
supervised the work. T.S.B., G.S. and M.R.D. wrote the manuscript.
Additional information
Supplementary Information accompanies this paper at http://www.nature.com/
naturecommunications
Competing financial interests: The authors declare no competing financial interests.
Reprints and permission information is available online at http://npg.nature.com/
reprintsandpermissions/
How to cite this article: Blacker, T. S. et al. Separating NADH and NADPH fluorescence
in live cells and tissues using FLIM. Nat. Commun. 5:3936 doi: 10.1038/ncomms4936
(2014).
This work is licensed under a Creative Commons Attribution 3.0
Unported License. The images or other third party material in this
article are included in the article’s Creative Commons license, unless indicated otherwise
in the credit line; if the material is not included under the Creative Commons license,
users will need to obtain permission from the license holder to reproduce the material.
To view a copy of this license, visit http://creativecommons.org/licenses/by/3.0/
NATURE COMMUNICATIONS | DOI: 10.1038/ncomms4936 ARTICLE
NATURE COMMUNICATIONS | 5:3936 | DOI: 10.1038/ncomms4936 | www.nature.com/naturecommunications 9
& 2014 Macmillan Publishers Limited. All rights reserved.
... These include the pH, which is especially variable in the mitochondrial matrix 10 , and protein composition, which impacts the lifetime of the protein-bound NADH 14 . Additionally, NADPH autofluorescence cannot be spectrally distinguished from NADH fluorescence 15 . As a result, these factors must be taken into account when utilizing NADH imaging to evaluate cellular metabolism. ...
... Comparing the relative amounts of NADPH versus NADH revealed that NR treatment did not change the ratio of NADPH to NADH, but FK866 treatment resulted in higher NADPH/ NADH ratio (Fig. S5c). To verify our biochemical quantifications, we also performed analysis developed by Blacker et al. to determine the NADPH/ NADH ratio in treated and untreated cells using lifetime components from FLIM 15 (Fig. S5d, e). This analysis showed close correlations between the optically quantified NADPH/NADH ratio in both mitochondria and the nucleus and whole cell biochemical quantifications of NADPH/NADH, suggesting no significant differential effects on the subcellular level ( Fig. S5f, g). ...
... One underlying reason for these different characteristics could be that changes in the NAD(H) pool size affect NADP(H) pool size to a lesser extent, altering the ratio of NADPH to NADH. It has been reported that protein-bound NADPH has a longer lifetime compared to NADH 15 . Thus, the change in the NADPH to NADH ratio could explain the influence of τ2 on pool-size induced lifetime alterations. ...
Article
Full-text available
NADH autofluorescence imaging is a promising approach for visualizing energy metabolism at the single-cell level. However, it is sensitive to the redox ratio and the total NAD(H) amount, which can change independently from each other, for example with aging. Here, we evaluate the potential of fluorescence lifetime imaging microscopy (FLIM) of NADH to differentiate between these modalities. We perform targeted modifications of the NAD(H) pool size and ratio in cells and mice and assess the impact on NADH FLIM. We show that NADH FLIM is sensitive to NAD(H) pool size, mimicking the effect of redox alterations. However, individual components of the fluorescence lifetime are differently impacted by redox versus pool size changes, allowing us to distinguish both modalities using only FLIM. Our results emphasize NADH FLIM’s potential for evaluating cellular metabolism and relative NAD(H) levels with high spatial resolution, providing a crucial tool for our understanding of aging and metabolism.
... In the cases studied here, the NAD(P)H and FAD reference lifetimes can be inferred from the observed linear arrangement of the phasorswhich were in agreement with reported values. 36 Because of shot noise and other sources of variance (e.g., differing bound state populations of NAD(P)H), 37 orthogonal projection of each phasor onto the line connecting both references is undertaken to obtain the phasor ratio r expressing the relative distance of the phasor to the two reference locations, as follows: E Q -T A R G E T ; t e m p : i n t r a l i n k -; e 0 0 2 ; 1 1 7 ; 3 5 8 8 > > < > > : ...
Article
Full-text available
Significance: To enable non-destructive longitudinal assessment of drug agents in intact tumor tissue without the use of disruptive probes, we have designed a label-free method to quantify the health of individual tumor cells in excised tumor tissue using multiphoton fluorescence lifetime imaging microscopy (MP-FLIM). Aim: Using murine tumor fragments which preserve the native tumor microenvironment, we seek to demonstrate signals generated by the intrinsically fluorescent metabolic co-factors nicotinamide adenine dinucleotide phosphate [NAD(P)H] and flavin adenine dinucleotide (FAD) correlate with irreversible cascades leading to cell death. Approach: We use MP-FLIM of NAD(P)H and FAD on tissues and confirm viability using standard apoptosis and live/dead (Caspase 3/7 and propidium iodide, respectively) assays. Results: Through a statistical approach, reproducible shifts in FLIM data, determined through phasor analysis, are shown to correlate with loss of cell viability. With this, we demonstrate that cell death achieved through either apoptosis/necrosis or necroptosis can be discriminated. In addition, specific responses to common chemotherapeutic treatment inducing cell death were detected. Conclusions: These data demonstrate that MP-FLIM can detect and quantify cell viability without the use of potentially toxic dyes, thus enabling longitudinal multi-day studies assessing the effects of therapeutic agents on tumor fragments.
... The contrast for imaging is based on the lifetime of individual fluorophores rather than their emission spectra, which is affected by the environment of the fluorophore. This allows for a more sensitive approach to detect autofluorescence as it enables discrimination between fluorescent molecules integral to embryo metabolism that share similar spectral characteristics such as bound-and free-NADH, enabling differentiation of NADH from NADPH [96]. Therefore, this imaging approach has the potential for label-free imaging of dynamic molecular processes in the embryo. ...
Article
Full-text available
Embryo quality is an important determinant of successful implantation and a resultant live birth. Current clinical approaches for evaluating embryo quality rely on subjective morphology assessments or an invasive biopsy for genetic testing. However, both approaches can be inherently inaccurate and crucially, fail to improve the live birth rate following the transfer of in vitro produced embryos. Optical imaging offers a potential non-invasive and accurate avenue for assessing embryo viability. Recent advances in various label-free optical imaging approaches have garnered increased interest in the field of reproductive biology due to their ability to rapidly capture images at high-resolution, delivering both morphological and molecular information. This burgeoning field holds immense potential for further development, with profound implications for clinical translation. Here, our review aims to: 1) describe the principles of various imaging systems, distinguishing between approaches that capture morphological and molecular information, 2) highlight the recent application of these technologies in the field of reproductive biology, and 3) assess their respective merits and limitations concerning the capacity to evaluate embryo quality. Additionally, the review summarizes challenges in the translation of optical imaging systems into routine clinical practice, providing recommendations for their future development. Finally, we identify suitable imaging approaches for interrogating the mechanisms underpinning successful embryo development.
... Using a single fluorophore on the biosensor simplifies multiplexed imaging. [10][11][12][13] FLIM was previously explored for intracellular biosensor imaging, including studies of NAD(P)H, 14 calcium, 15,16 sodium, 17 and protein conformation. For the latter, FLIM was used largely to image FRET, 8,9,[18][19][20][21] but FLIM biosensors based on single dyes have the potential for brighter emission than FRET because the dyes are directly excited. ...
Article
Full-text available
In this work, we examine the use of environment-sensitive fluorescent dyes in fluorescence lifetime imaging microscopy (FLIM) biosensors. We screened merocyanine dyes to find an optimal combination of environment-induced lifetime changes, photostability, and brightness at wavelengths suitable for live-cell imaging. FLIM was used to monitor a biosensor reporting conformational changes of endogenous Cdc42 in living cells. The ability to quantify activity using phasor analysis of a single fluorophore (e.g., rather than ratio imaging) eliminated potential artifacts. We leveraged these properties to determine specific concentrations of activated Cdc42 across the cell.
Article
Full-text available
Autofluorophores are endogenous fluorescent compounds that naturally occur in the intra and extracellular spaces of all tissues and organs. Most have vital biological functions – like the metabolic cofactors NAD(P)H and FAD⁺, as well as the structural protein collagen. Others are considered to be waste products – like lipofuscin and advanced glycation end products – which accumulate with age and are associated with cellular dysfunction. Due to their natural fluorescence, these materials have great utility for enabling non‐invasive, label‐free assays with direct ties to biological function. Numerous technologies, with different advantages and drawbacks, are applied to their assessment, including fluorescence lifetime imaging microscopy, hyperspectral microscopy, and flow cytometry. Here, the applications of label‐free autofluorophore assessment are reviewed for clinical and health‐research applications, with specific attention to biomaterials, disease detection, surgical guidance, treatment monitoring, and tissue assessment – fields that greatly benefit from non‐invasive methodologies capable of continuous, in vivo characterization.
Article
Microbial production of L‐malic acid from renewable carbon sources has attracted extensive attention. The reduced cofactor NADPH plays a key role in biotransformation because it participates in both biosynthetic reactions and cellular stress responses. In this study, NADPH or its precursors nicotinamide and nicotinic acid were added to the fermentation medium of Aspergillus niger RG0095, which significantly increased the yield of malic acid by 11%. To further improve the titer and productivity of L‐malic acid, we increased the cytoplasmic NADPH levels of A. niger by upregulating the NAD kinases Utr1p and Yef1p. Biochemical analyses demonstrated that overexpression of Utr1p and Yef1p reduced oxidative stress, while also providing more NADPH to catalyze the conversion of glucose into malic acid. Notably, the strain overexpressing Utr1p reached a malate titer of 110.72 ± 1.91 g L ⁻¹ after 108 h, corresponding to a productivity of 1.03 ± 0.02 g L ⁻¹ h ⁻¹ . Thus, the titer and productivity of malate were increased by 24.5% and 44.7%, respectively. The strategies developed in this study may also be useful for the metabolic engineering of fungi to produce other industrially relevant bulk chemicals.
Article
Currently, much attention in oncology is devoted to the issues of tumor heterogeneity, which creates serious problems in the diagnosis and therapy of malignant neoplasms. Intertumoral and intratumoral differences relate to various characteristics and aspects of the vital activity of tumor cells, including cellular metabolism. This review provides general information about the tumor metabolic heterogeneity with a focus on energy metabolism, its causes, mechanisms and research methods. Among the methods, fluorescence lifetime imaging is described in more detail as a new promising method for observing metabolic heterogeneity at the cellular level. The review demonstrates the importance of studying the features of tumor metabolism and identifying intra- and intertumoral metabolic differences.
Article
Full-text available
The recording of oxidation-reduction-related fluorescence signals of oxidized flavoprotein (Fp) and reduced pyridine nucleotide (PN) from isolated mitochondria at temperatures below -80 degrees C can be accompanished with a high degree of accuracy and a wide dynamic range. The specific low temperature enhancement of the fluorescence signals due to increased quantum yield and to multiple scattering affords increased accuracy and less interference due to screening pigments such as hemoglobin and myoglobin. Since the metabolic processes are arrested and the recording speed can be greatly diminished, the technique can operate with a much smaller concentration of mitochondria than is needed at room temperature, and the method is suitable for localized oxidation-reduction measurements. The Fp and PN signals originate from the mitochondrial matrix space in which they represent the major fluorochromes. Since Fp and PN are near oxidation-reduction equilibrium, the ratio of the two fluorescence intensities, suitably normalized, approximates the oxidation-reduction ratio of oxidized flavoprotein/reduced pyridine nucleotide. Thus, this technique affords a foundation for the resolution of oxidation-reduction states in two and three dimensions.
Article
Full-text available
Minimally invasive, specific measurement of cellular energy metabolism is crucial for understanding cerebral pathophysiology. Here, we present high-resolution, in vivo observations of autofluorescence lifetime as a biomarker of cerebral energy metabolism in exposed rat cortices. We describe a customized two-photon imaging system with time correlated single photon counting detection and specialized software for modeling multiple-component fits of fluorescence decay and monitoring their transient behaviors. In vivo cerebral NADH fluorescence suggests the presence of four distinct components, which respond differently to brief periods of anoxia and likely indicate different enzymatic formulations. Individual components show potential as indicators of specific molecular pathways involved in oxidative metabolism.
Article
The ubiquitous redox cofactors nicotinamide adenine dinucleotides [NAD and NADP] are very similar molecules, despite their participation in substantially different biochemical processes. NADP differs from NAD in only the presence of an additional phosphate group esterified to the 2′-hydroxyl group of the ribose at the adenine end and yet NADP is confined with few exceptions to the reactions of reductive biosynthesis, whereas NAD is used almost exclusively in oxidative degradations. The discrimination between NAD and NADP is therefore an impressive example of the power of molecular recognition by proteins. The many known tertiary structures of NADP complexes affords the possibility for an analysis of their discrimination. A systematic analysis of several crystal structures of NAD(P)-protein complexes show that: 1) the NADP coenzymes are more flexible in conformation than those of NAD; 2) although the protein-cofactor interactions are largely conserved in the NAD complexes, they are quite variable in those of NADP; and 3) in both cases the pocket around the nicotinamide moiety is substrate dependent. The conserved and variable interactions between protein and cofactors in the respective binding pockets are reported in detail. Discrimination between NAD and NADP is essentially a consequence of the overall pocket and not of a few residues. A clear fingerprint in NAD complexes is a carboxylate side chain that chelates the diol group at the ribose near the adenine, whereas in NADP complexes an arginine side chain faces the adenine plane and interacts with the phosphomonoester. The latter type of interaction might be a general feature of recognition of nucleotides by proteins. Other features such as strand-like hydrogen bonding between the NADP diphosphate moeties and the protein are also significant. The NADP binding pocket properties should prove useful in protein engineering and design. © 1997 Wiley-Liss Inc.
Article
The ubiquitous redox cofactors nicotinamide adenine dinucleotides [NAD and NADP] are very similar molecules, despite their participation in substantially different biochemical processes. NADP differs from NAD in only the presence of an additional phosphate group esterified to the 2′-hydroxyl group of the ribose at the adenine end and yet NADP is confined with few exceptions to the reactions of reductive biosynthesis, whereas NAD is used almost exclusively in oxidative degradations. The discrimination between NAD and NADP is therefore an impressive example of the power of molecular recognition by proteins. The many known tertiary structures of NADP complexes affords the possibility for an analysis of their discrimination. A systematic analysis of several crystal structures of NAD(P)-protein complexes show that: 1) the NADP coenzymes are more flexible in conformation than those of NAD; 2) although the protein-cofactor interactions are largely conserved in the NAD complexes, they are quite variable in those of NADP; and 3) in both cases the pocket around the nicotinamide moiety is substrate dependent. The conserved and variable interactions between protein and cofactors in the respective binding pockets are reported in detail. Discrimination between NAD and NADP is essentially a consequence of the overall pocket and not of a few residues. A clear fingerprint in NAD complexes is a carboxylate side chain that chelates the diol group at the ribose near the adenine, whereas in NADP complexes an arginine side chain faces the adenine plane and interacts with the phosphomonoester. The latter type of interaction might be a general feature of recognition of nucleotides by proteins. Other features such as strand-like hydrogen bonding between the NADP diphosphate moeties and the protein are also significant. The NADP binding pocket properties should prove useful in protein engineering and design. © 1997 Wiley-Liss Inc.
Article
The aim of this study was to test the hypothesis that glucose can be monitored non-invasively by measuring NAD(P)H-related fluorescence lifetime of cells in an in vitro cell culture model. Autofluorescence decay functions were measured in 3T3-L1 adipocytes by time-correlated single-photon counting (excitation 370nm, emission 420-480nm). Free NADH had a two-exponential decay but cell autofluorescence fitted best to a three-exponential decay. Addition of 30mM glucose caused a 29% increase in autofluorescence intensity, a significantly shortened mean lifetime (from 7.23 to 6.73ns), and an increase in the relative amplitude and fractional intensity of the short-lifetime component at the expense of the two longer-lifetime components. Similar effects were seen with rotenone, an agent that maximizes mitochondrial NADH. 3T3-L1 fibroblasts stained with the fluorescent mitochondrial marker, rhodamine 123 showed a 16% quenching of fluorescence intensity when exposed to 30mM glucose, and an increase in the relative amplitude and fractional intensity of the short lifetime at the expense of the longer lifetime component. We conclude that, though the effect size is relatively small, glucose can be measured non-invasively in cells by monitoring changes in the lifetimes of cell autofluorescence or of a dye marker of mitochondrial metabolism. Further investigation and development of fluorescence intensity and lifetime sensing is therefore indicated for possible non-invasive metabolic monitoring in human diabetes.
Article
In live tissue, alterations in metabolism induce changes in the fluorescence decay of the biological coenzyme NAD(P)H, the mechanism of which is not well understood. In this work, the fluorescence and anisotropy decay dynamics of NADH and NADPH were investigated as a function of viscosity in a range of water–glycerol solutions. The viscosity dependence of the non-radiative decay is well described by Kramers and Kramers–Hubbard models of activated barrier crossing over a wide viscosity range. Our combined lifetime and anisotropy analysis indicates common mechanisms of non-radiative relaxation in the two emitting states (conformations) of both molecules. The low frequencies associated with barrier crossing suggest that non-radiative decay is mediated by small scale motion (e.g. puckering) of the nicotinamide ring. Variations in the fluorescence lifetimes of NADH and NADPH when bound to different enzymes may therefore be attributed to differing levels of conformational restriction upon binding.
Article
The spectrophotometric evaluation of NAD(P) dehydrogenase enzymatic activity is very popular as the reduced form (NAD(P)H) absorbs at 340 nm, while the aromatic oxidised form does not. In this joint theoretical and experimental investigation, we identify the chromophoric unit of both the NAD(P)+ and NAD(P)H forms. Rather than a modification of size or shape of the frontier orbitals, the sharp variation in the absorption upon reduction is mainly related to the stabilisation of the occupied orbitals due to the positive charge. In addition, TD-DFT nicely reproduces the experimental wavelengths.