ArticlePDF Available

Beta-Amyloid Peptides Induce Mitochondrial Dysfunction and Oxidative Stress in Astrocytes and Death of Neurons through Activation of NADPH Oxidase

Authors:

Abstract and Figures

Beta-amyloid (betaA) peptide is strongly implicated in the neurodegeneration underlying Alzheimer's disease, but the mechanisms of neurotoxicity remain controversial. This study establishes a central role for oxidative stress by the activation of NADPH oxidase in astrocytes as the cause of betaA-induced neuronal death. betaA causes a loss of mitochondrial potential in astrocytes but not in neurons. The mitochondrial response consists of Ca2+-dependent transient depolarizations superimposed on a slow collapse of potential. The slow response is both prevented by antioxidants and, remarkably, reversed by provision of glutamate and other mitochondrial substrates to complexes I and II. These findings suggest that the depolarization reflects oxidative damage to metabolic pathways upstream of mitochondrial respiration. Inhibition of NADPH oxidase by diphenylene iodonium or 4-hydroxy-3-methoxy-acetophenone blocks betaA-induced reactive oxygen species generation, prevents the mitochondrial depolarization, prevents betaA-induced glutathione depletion in both neurons and astrocytes, and protects neurons from cell death, placing the astrocyte NADPH oxidase as a primary target of betaA-induced neurodegeneration.
␤ -Amyloid causes depolarization of mitochondria in astrocytes but not in neurons. Changes in ⌬ ␺ m were measured using Rh123 in dequench mode; the loss of potential is seen as an increase in fluorescence. A , Changes in ⌬ ␺ m from a neuron (black line) and an astrocyte (gray symbols) in a mixed culture from rat hippocampus (15 DIV) after exposure to 50 ␮ M ␤ A 25-35. In the example shown, ␤ A 25-35 caused a slow progressive collapse of ⌬ ␺ m in the astrocyte, whereas no change at all was seen in a nearby neuron. Application of 300 ␮ M glutamate caused a rapid collapse of ⌬ ␺ m in the neuron but promoted recovery of ⌬ ␺ m in the astrocyte. On subsequent removal of external Ca 2 ϩ , neuronal ⌬ ␺ m recovered, showing that mitochondrial injury was still reversible. In this and all subsequent records, the protonophore FCCP was added at the end of the experiment to determine the extent of the Rh123 signal in response to complete mitochondrial depolarization. The Rh123 fluorescence signal in these traces is normalized between 0, representing the resting Rh123 fluorescence, and 100, representing the maximal increase in Rh123 fluorescence in response to complete mitochondrial depolarization by 1 ␮ M FCCP. In traces in which the Rh123 signal was lost in some cells during exposure to ␤ A, this normalization was not possible, and so the data are shown normalized only to a baseline set at 100%. Thus, the trace shown in A is actually very similar to some of the traces in B , in which FCCP responses were maintained. In B is shown an example of an experiment using astrocytes in culture in response to the full peptide ␤ A 1-42 (5 ␮ M ). The cells responded with a slow loss of mitochondrial potential on which were superimposed abrupt depolarizing transitions. Some of these were reversible, but the larger depolarizations were followed by loss of the dye, suggesting cell death. Some examples of these traces are extracted from another data set to illustrate these three types of response in C . The series of images shown in D illustrate examples from an extended time sequence showing the transient increases in signal and the gradual increase in basal signal in response to ␤ A 25-35. The image after FCCP saturates the display at this range because it is much brighter than the rest of the sequence and so is not shown. The time of each extracted image is indicated in minutes. In Eii and Eii are shown records from two astrocytes co-loaded with fura-2 and Rh123 to measure [Ca 2 ϩ ] c (black triangles) and ⌬ ␺ m (gray lines and filled circles) simultaneously during exposure to 50 ␮ M ␤ A 25-35. As we have shown previously, ␤ A caused fluctuations in [Ca 2 ϩ ] c in the astrocytes. Although the abrupt mitochondrial depolarizations were clearly associated with [Ca 2 ϩ ] c signals, they did not show a tight or fixed correlation in time or amplitude. Thus, large changes in ⌬ ␺ m followed large changes in [Ca 2 ϩ ] c , and some small transient depolar- izations of ⌬ ␺ m could be seen associated with [Ca 2 ϩ ] c transients (asterisks in Eii ).
… 
Content may be subject to copyright.
Neurobiology of Disease
-Amyloid Peptides Induce Mitochondrial Dysfunction and
Oxidative Stress in Astrocytes and Death of Neurons through
Activation of NADPH Oxidase
Andrey Y. Abramov,
1
Laura Canevari,
2
and Michael R. Duchen
1
1
Mitochondrial Biology Group, Department of Physiology, University College London, London WC1E 6BT, United Kingdom, and
2
Division of
Neurochemistry, Institute of Neurology, London WC1N 3BG, United Kingdom
-Amyloid (
A) peptide is strongly implicated in the neurodegeneration underlying Alzheimer’s disease, but the mechanisms of neu-
rotoxicity remain controversial. This study establishes a central role for oxidative stress by the activation of NADPH oxidase in astrocytes
as the cause of
A-induced neuronal death.
A causes a loss of mitochondrial potential in astrocytes but not in neurons. The mitochon-
drial response consists of Ca
2
-dependent transient depolarizations superimposed on a slow collapse of potential. The slow response is
both prevented by antioxidants and, remarkably, reversed by provision of glutamate and other mitochondrial substrates to complexes I
and II. These findings suggest that the depolarization reflects oxidative damage to metabolic pathways upstream of mitochondrial
respiration. Inhibition of NADPH oxidase by diphenylene iodonium or 4-hydroxy-3-methoxy-acetophenone blocks
A-induced reactive
oxygen species generation, prevents the mitochondrial depolarization, prevents
A-induced glutathione depletion in both neurons and
astrocytes, and protects neurons from cell death, placing the astrocyte NADPH oxidase as a primary target of
A-induced
neurodegeneration.
Key words: Alzheimer; astrocyte; astroglia; calcium; mitochondria; NADPH; neuron
Introduction
The deposition of
-amyloid (
A) plays a central role in the
pathogenesis of Alzheimer’s disease (AD). Accumulation of
A
in neuritic plaques is the defining feature for diagnosis of the
disease, and the amyloid load correlates well with the degree of
cognitive impairment (Naslund et al., 2000). The 1-40 and 1-42
amino acid forms of
A are neurotoxic, a property conserved in
the shorter peptide form 25-35, whereas the reverse peptide, 35-
25, is innocuous. We have found recently in neurons and astro-
cytes grown in coculture, in cultures of cortical astrocytes, or in
organotypic hippocampal slices that
A causes a delayed rise in
intracellular free calcium ([Ca
2
]
c
) in astrocytes but not in neu-
rons (Abramov et al., 2003). These data suggested that
A forms
a pore in astrocyte (but not neuronal) membranes that permits
Ca
2
influx from the extracellular space. The combined effect of
high [Ca
2
]
c
with oxidative stress may damage mitochondrial
function and has been implicated as a pathophysiological mech-
anism in many systems (Duchen, 2000). Therefore, we have now
investigated the connection between changes in [Ca
2
]
c
and
changes in mitochondrial function in
A neurotoxicity.
Mitochondrial degeneration is one of the earliest signs of Alz-
heimer pathology, appearing before neurofibrillary tangles are
evident (Hirai et al., 2001). It is also clearly associated with the
overexpression of the amyloid precursor protein (APP) (Askanas
et al., 1996) or expression of the APP
751
form in cultured cells
(Grant et al., 1999). Recently, it has also been shown that APP
may also be targeted to mitochondria, causing mitochondrial
dysfunction (Anandatheerthavarada et al., 2003), in addition to
established localization in the plasma membrane and endoplas-
mic reticulum. The mechanism of the mitochondrial involve-
ment in
A-induced neurotoxicity, however, remains unclear. As
well as being a target of oxidative damage, mitochondria may be
a source of endogenous production of reactive oxygen species
(ROS), and
A may increase mitochondrial ROS production
(Sheehan et al., 1997), causing further impairment of mitochon-
drial function (Arias et al., 2002). Enzyme systems containing
iron–sulfur centers, including several enzyme complexes of the
respiratory chain,
-ketoglutarate dehydrogenase, and aconitase,
are particularly vulnerable to damage by both
A and ROS (Blass
and Gibson, 1991; Casley et al., 2002b; Longo et al., 2000).
A also
has multiple direct effects on isolated mitochondria, causing al-
terations in enzyme activity, damage to the respiratory chain, and
opening of the mitochondrial permeability transition pore
(Canevari et al., 1999; Parks et al., 2001; Shevtzova et al., 2001;
Kim et al., 2002; Moreira et al., 2002). The latter may trigger cell
death, either promoting cytochrome crelease and apoptosis or
causing energetic failure and necrosis. Furthermore,
A-induced
apoptosis requires an intact respiratory chain, because
A did not
Received Sept. 2, 2003; revised Nov. 13, 2003; accepted Nov. 14, 2003.
This work was supported by the Wellcome Trust, The Medical Research Council, and the Miriam Marks Fund. We
especially thank Dr. Jake Jacobson for invaluable suggestions and discussion while the work was in progress, and we
thank Professors John Clark, Stephen Bolsover, and Rosario Rizzuto for helpful discussion and comments on this
manuscript.
Correspondence should be addressed to Michael R. Duchen, Department of Physiology, University College Lon-
don, Gower Street, London WC1E 6BT, UK. E-mail: m.duchen@ucl.ac.uk.
DOI:10.1523/JNEUROSCI.4042-03.2004
Copyright © 2004 Society for Neuroscience 0270-6474/04/240565-11$15.00/0
The Journal of Neuroscience, January 14, 2004 24(2):565–575 565
cause apoptosis in cells lacking mitochondrial DNA (Morais Car-
doso et al., 2002). Even in intact cortical neurons,
A causes
mitochondrial dysfunction, reducing ATP levels (Casley et al.,
2002a), mostly through inhibition of complex I, causing both
mitochondrial depolarization and a loss of mitochondrial mass
(Casley et al., 2002b). Therefore, we have now investigated the
nature and mechanism of
A-induced mitochondrial
dysfunction.
Materials and Methods
Cell culture. Mixed cultures of hippocampal neurons and glial cells were
prepared as described previously (Abramov et al., 2003) from Sprague
Dawley rat pups 24 d postpartum [University College London (UCL)
breeding colony]. Hippocampi were removed into ice-cold Geys salt
solution (Invitrogen, Paisley, UK) with 20
g/ml gentamycin. The tissue
was minced and trypsinized (0.1% for 15 min at 37°C), triturated, plated
on poly-D-lysine-coated coverslips, and cultured in Neurobasal medium
(Invitrogen) supplemented with B-27 (Invitrogen) and 2 mM
L-glutamine. Cultures were maintained at 37°C in a humidified atmo-
sphere of 5% CO
2
and 95% air, fed twice a week, and maintained for a
minimum of 10 d before experimental use to ensure the expression of
glutamate and other receptors. Neurons were easily distinguishable from
glia: they appeared phase bright, had smooth rounded somata and dis-
tinct processes, and lay just above the focal plane of the glial layer. Cells
were used at 10 20din vitro (DIV) unless stated otherwise.
Isolated cortical astrocytes were prepared as described previously
(Abramov et al., 2003). Cerebra taken from adult Sprague Dawley rats
(UCL breeding colony), were chopped and triturated until homoge-
neous, passed through a 297
m mesh, and trypsinized (50,000 U/ml
porcine pancreas; Sigma, Gillingham, UK) with 336 U/ml DNase 1 (bo-
vine pancreas; Sigma), and 1.033 U/ml collagenase (Sigma, Gillingham,
UK) at 37°C for 15 min. After addition of fetal bovine serum (10% of final
volume) and filtering through 140
Mmesh, the tissue was centrifuged
through 0.4 Msucrose (400 gm, 10 min), and the resulting pellet was
transferred to Minimal Essential Medium supplemented with 5% fetal
calf serum, 2 mMglutamine, and 1 mMmaleate in tissue culture flasks
precoated with 0.01% poly-D-lysine. The cells reached confluency at
1214 DIV and were harvested and reseeded onto 24-mm-diameter glass
coverslips (BDH, Poole, UK) precoated with 0.01% poly-D-lysine for
fluorescence measurements and used during 24 d. Purity of cell type
was confirmed by immunohistochemistry using antibodies directed to
glial acidic fibrillary protein (GFAP). In excess of 99% of the cells exam-
ined stained positively for GFAP.
In all experiments we have used both hippocampal neurons and astro-
cytes in cocultures and preparations of purified astrocytes from hip-
pocampus or cortex. The former was used routinely as a physiological
model in which we could ensure the selective effects of agents on neurons
or astrocytes. The astrocyte cultures were used because the predominant
effects seen were in astrocytes. We need to be sure that these effects were
not secondary to some action on neurons that was not being measured,
and also we could exclude contributions to the response from other cell
types, especially microglia, because we have ascertained that the micro-
glial contamination of the astrocyte cultures is extremely small. In this
paper, we have referred to data from the different preparations inter-
changeably because we have seen no difference in astrocytes responses
from any of the cultures systems that we have used.
Peptides and treatments.
A 25-35,
A 1-42, and
A 35-25 (Bachem,
St. Helens, UK) were dissolved at 15m
Min sterile ultrapure water
(Milli-Q standard; Millipore, Watford, UK) and kept frozen until use.
The peptides were added to cells during experimental recordings, except
for the neurotoxicity measurements, in which they were added 24 hr
before the assays of cell death (see below).
A 25-35 and 35-25 were used
at concentrations of up to 50
Mto ensure that it was present in molar
excess compared with inhibitors and so would exclude any direct inter-
action, and
A 1-42 was used at concentrations of 15
M. The mono-
clonal antibody CD36 (10
g/ml; clone FA6152; Immunotech, Mar-
seille, France) was incubated with cells 30 min before experiments at
37°C and was also added at the time of experiments. The 25-35 peptide is
easy to use and relatively inexpensive, but the 1-42 peptide probably
represents a more physiologically appropriate model, and so all experi-
ments have been repeated with this peptide. All experiments were also
repeated using the 35-25 reverse peptide as a control. We have never seen
any significant response to this treatment.
Imaging [Ca
2
]
c
and
m
and ROS generation. Cells were loaded for
30 min at room temperature with 5
Mfura-2 AM (Molecular Probes,
Eugene, OR) and 0.005% pluronic acid in a HEPES-buffered salt solution
(HBSS) composed of (in mM): 156 NaCl, 3 KCl, 2 MgSO
4
, 1.25 KH
2
PO
4
,
2 CaCl
2
, 10 glucose, and 10 HEPES, pH adjusted to 7.35 with NaOH. For simul-
taneous measurement of [Ca
2
]
i
and
m
, rhodamine (Rh) 123 (10
M;Mo
-
lecular Probes) was added into the cultures during the last 15 min of the fura-2
loading period. For measurements of rates of ROS generation, cells were incu-
bated with 20
M2,7-dichlorodihydrofluorescein diacetate (Molecular
Probes) in HBSS for 40 min at room temperature.
Fluorescence measurements were obtained on an epifluorescence in-
verted microscope equipped with a 20fluorite objective. [Ca
2
]
i
and
m
were monitored in single cells using excitation light provided by a
Xenon arc lamp with the beam passing sequentially through 10 nm band-
pass filters centered at 340, 380, and 490 nm housed in a computer-
controlled filter wheel (Cairn Research, Kent, UK). Emitted fluorescence
light was reflected through a 515 nm long-pass filter to a frame transfer
cooled CCD camera (Orca ER; Hamamatsu, Welwyn Garden City, UK).
All imaging data were collected and analyzed using Kinetic Imaging soft-
ware (Kinetic Imaging, Wirral, UK). The fluorescence data were acquired
at intervals of 10 15 sec. The fura2 data have not been calibrated in
terms of [Ca
2
]
i
because of the uncertainty arising from the use of dif-
ferent calibration techniques. Accumulation of Rh123 in polarized mi-
tochondria quenches the fluorescent signal; in response, emission is de-
quenched. An increase in Rh123 signal therefore signals mitochondrial
depolarization (Duchen and Biscoe, 1992). We have presented these data
normalized between 0, representing the resting Rh123 fluorescence, and
100, representing the maximal increase in Rh123 fluorescence in re-
sponse to complete mitochondrial depolarization by 1
Mcarbonyl cya-
nide 4-(trifluoromethoxy)phenylhydazone (FCCP), which was done
routinely at the end of every experiment (indicated by a label Rh123
fluorescence, scaled). In some experiments, cells showed a loss of Rh123
and so showed no, or only small, responses to FCCP. In such traces, this
normalization was not possible, and so the data are shown normalized
only to a baseline set at 100% (indicated as Rh123 fluorescence, percent-
age). All presented data were obtained from at least five coverslips and
two to three different cell preparations.
In all experiments in the present paper, we have used Rh123 in de-
quench: mode. We have also conducted equivalent experiments using
the dye tetramethyl rhodamine methyl ester (TMRM) in redistribution
mode,in which the dye was present continuously at 20 nMand allowed
to equilibrate. Mitochondrial depolarization is then seen as the move-
ment of dye from mitochondria into the cytosol. Responses to
A were
seen showing a slow and progressive decline in mitochondrially localized
dye on which were superimposed more dramatic and reversible decreases
in mitochondrial-specific signal, but high-resolution imaging of mito-
chondria over this long period of time was confounded by movement of
mitochondria and swelling of cells associated with the calcium transients.
We therefore have limited the data presented to the Rh123 data set,
although the TMRM experiments show qualitatively the same
phenomenology.
Cells loaded with 2,7-dichlorodihydrofluorescein (DCF) were illu-
minated at 490 nm at low light levels to avoid auto-oxidation. Changes in
the rate of rise of the signal were interpreted as changes in rates of ROS
generation and could be fitted with a regression coefficient of 0.98 by a
linear regression line (Origin; Microcal Software Inc., Northampton,
MA).
Toxicity experiments. For toxicity assays cells were loaded simulta-
neously with 20
Mpropidium iodide, which is excluded from viable cells
but exhibits a red fluorescence after a loss of membrane integrity, and 4.5
MHoechst 33342 (Molecular Probes), which gives a blue staining to
566 J. Neurosci., January 14, 2004 24(2):565–575 Abramov et al.
A, Mitochondria, and Oxidative Stress in Astrocytes
chromatin, to count the total number of cells.
With use of phase-contrast optics, a bright-field
image allowed identification of neurons, which
look quite different from the flatter astrocytes
and also lie in a different focal plane, above the
astrocytic layer. A total number of 600800
neurons or astrocytes were counted in 2025
fields of each coverslip. Each experiment was
repeated five or more times using separate
cultures.
Statistical analysis. Statistical analysis and
exponential curve fitting were performed us-
ing Origin 7 (Microcal Software Inc.) soft-
ware. Results are expressed as means SEM
(SEM).
Results
Application of the
A peptide fragment
25-35 (550
M) or the full-length peptide
1-42 (0.55
M) to hippocampal neurons
and astrocytes in coculture caused changes
in
m
of astrocytes (n510 cells) after a
delay of 310 min, whereas mitochon-
drial potential remained completely stable
in adjacent neurons (n298 cells) (Fig.
1A). Neurons were positively identified by
the application of 100
Mglutamate to-
ward the end of the experiment, because
glutamate routinely causes a collapse of
mitochondrial potential in neurons at this
stage in culture (Vergun et al., 1999), with-
out altering mitochondrial potential in as-
trocytes. Remarkably, although glutamate
caused the predicted collapse of potential
in neurons, in astrocytes it reversed the
mitochondrial depolarization caused by
A (Fig. 1A) (and see below), further ac-
centuating the difference in the responses
of these two cell types. Switching to a
Ca
2
-free saline at the end of the experi-
ment revealed that the mitochondrial de-
polarization in the neurons was reversible
at this stage (Khodorov et al., 1996; Ver-
gun et al., 1999), despite the presence of
A, suggesting that under these conditions
A does not significantly enhance the neu-
rotoxicity of glutamate.
Both the full-length peptide
A 1-42
and the peptide fragment 25-35 produced
changes in
m
in the vast majority
(95.2 1.3%) of cells in a culture of cor-
tical astrocytes and in 89.6 2.1% of as-
trocytes in coculture with neurons. The
nontoxic reverse peptide
A 35-25 had no
effect at all on
m
either in hippocampal
neurons (n123 cells) or in cortical or
hippocampal astrocytes (n191 cells).
The
A-induced changes in
m
of astro-
cytes could be divided into three catego-
ries, some examples of which are also pre-
sented in Figure 1, Band C: (1) a slow and
progressive mitochondrial depolarization;
(2) large, transient, and reversible abrupt
mitochondrial depolarizations, such as we
Figure1.
-Amyloidcausesdepolarizationofmitochondriainastrocytes but not in neurons. Changes in
m
weremeasured
using Rh123 in dequench mode; the loss of potential is seen as an increase in fluorescence. A, Changes in
m
from a neuron
(black line) and an astrocyte (gray symbols) in a mixed culture from rat hippocampus (15 DIV) after exposure to 50
M
A 25-35.
In the example shown,
A 25-35 caused a slow progressive collapse of
m
in the astrocyte, whereas no change at all was seen
in a nearby neuron. Application of 300
Mglutamate caused a rapid collapse of
m
in the neuron but promoted recovery of
m
in the astrocyte. On subsequent removal of external Ca
2
, neuronal
m
recovered, showing that mitochondrial injury
was still reversible. In this and all subsequent records, the protonophore FCCP was added at the end of the experiment to
determine the extent of the Rh123 signal in response to complete mitochondrial depolarization. The Rh123 fluorescence signal in
these traces is normalized between 0, representing the resting Rh123 fluorescence, and 100, representing the maximal increase
inRh123fluorescencein response to complete mitochondrial depolarizationby1
MFCCP.Intracesin which the Rh123 signal was
lost in some cells during exposure to
A, this normalization was not possible, and so the data are shown normalized only
to a baseline set at 100%. Thus, the trace shown in Ais actually very similar to some of the traces in B, in which FCCP
responses were maintained. In Bis shown an example of an experiment using astrocytes in culture in response to the full
peptide
A 1-42 (5
M). The cells responded with a slow loss of mitochondrial potential on which were superimposed
abrupt depolarizing transitions. Some of these were reversible, but the larger depolarizations were followed by loss of the
dye, suggesting cell death. Some examples of these traces are extracted from another data set to illustrate these three
types of response in C. The series of images shown in Dillustrate examples from an extended time sequence showing the
transient increases in signal and the gradual increase in basal signal in response to
A 25-35. The image after FCCP
saturates the display at this range because it is much brighter than the rest of the sequence and so is not shown. The time
of each extracted image is indicated in minutes. In Eii and Eii are shown records from two astrocytes co-loaded with fura-2
and Rh123 to measure [Ca
2
]
c
(black triangles) and
m
(gray lines and filled circles) simultaneously during exposure to
50
M
A 25-35. As we have shown previously,
A caused fluctuations in [Ca
2
]
c
in the astrocytes. Although the abrupt
mitochondrial depolarizations were clearly associated with [Ca
2
]
c
signals, they did not show a tight or fixed correlation
in time or amplitude. Thus, large changes in
m
followed large changes in [Ca
2
]
c
, and some small transient depolar-
izations of
m
could be seen associated with [Ca
2
]
c
transients (asterisks in Eii).
Abramov et al.
A, Mitochondria, and Oxidative Stress in Astrocytes J. Neurosci., January 14, 2004 24(2):565–575 567
have never witnessed before under any
conditions, often superimposed on a slow
progressive depolarization; and (3) abrupt
mitochondrial depolarizations that failed
to recover and were followed by the com-
plete release of Rh123 from the mitochon-
dria (and showed no further response to
FCCP application), probably signifying
cell death. The traces shown in Figure 1B
show a range of examples of cell responses
from one experiment, and in Figure 1Cwe
have extracted three examples from an-
other experiment to illustrate these three
patterns of response. For the majority of
traces that were used as illustrative mate-
rial we have chosen traces that show both
sustained and transient components, be-
cause these show how the components are
modulated separately. It should be clear
from the distribution histograms, how-
ever, that a substantial proportion of cells
routinely show only graded sustained re-
sponses. Examples of the images from
which these data were taken are shown in
Figure 1D, in which sample images from a
time series have been extracted. Note how
the Rh123 signal becomes bright in individ-
ual cells only to recover later. These cells usu-
ally responded again to FCCP, which was ap-
plied routinely at the end of the experiment.
Role of [Ca
2
]
c
in amyloid-induced
changes in mitochondrial potential
We have shown previously that
A in-
duces, after a delay, the appearance of os-
cillatory [Ca
2
]
c
signals in astrocytes
(Abramov et al., 2003) and not in neurons.
Because changes in [Ca
2
]
c
may initiate
changes in mitochondrial membrane po-
tential (for review, see Duchen, 1999), we
conducted experiments in which
m
and
[Ca
2
]
c
were measured simultaneously.
These experiments revealed an association
between the transient increases in [Ca
2
]
c
and the transient mitochondrial depolar-
izations. Close scrutiny of traces from in-
dividual cells (Fig. 1Ei,Eii) shows that the
transient mitochondrial responses are
clearly routinely associated with [Ca
2
]
c
transients, although not all
[Ca
2
]
c
transients cause mitochondrial depolarizations.
We have shown previously that
A-induced changes in
[Ca
2
]
c
are dependent on the presence of external calcium, re-
flecting
A-induced Ca
2
influx from the extracellular space
(Abramov et al., 2003). In keeping with this observation, changes
in [Ca
2
]
c
were abolished on the removal of external Ca
2
(Fig.
2Ai,Bi). In the absence of external Ca
2
,
A 25-35 or 1-42 still
induced a slow mitochondrial depolarization in hippocampal
and cortical astrocytes (n231) (Fig. 2Aii), but transient depo-
larizations were much reduced and irreversible abrupt depolar-
izations were abolished. This is further emphasized by the analy-
sis of each component of the responses illustrated in Figure 2C.
Washing the cells with a saline containing Ca
2
after a period in
Ca
2
-free conditions caused an abrupt [Ca
2
]
c
increase in the
astrocytes (Fig. 2Bi) and triggered a rapid loss of mitochondrial
potential in many cells (Fig. 2Bii). Thus, the slow progressive
mitochondrial depolarization is apparently relatively indepen-
dent of [Ca
2
]
c
, whereas the rapid transient changes in potential
appear to be more closely associated with
A-induced changes in
[Ca
2
]
c
(Fig. 2C). The average change in Rh123 signal caused by
30 min of exposure to
A in the absence of Ca
2
was significantly
smaller than seen in the standard Ca
2
-containing saline ( p
0.05; from 64.1 13.7% in standard buffer to 29.2 6.3% in
Ca
2
-free) (Fig. 2D). These differences in signal were primarily
attributable to the absence of large rapid transients or irreversible
depolarizations in the Ca
2
-free saline.
Zn
2
(1 mM) and the heavy metal chelator clioquinol (2
M) both block the
A-induced [Ca
2
]
c
fluctuation in astro-
cytes (Abramov et al., 2003). We therefore examined their
actions on the mitochondrial response and found that both of
Figure 2. Transient mitochondrial depolarizations are dependent on
A-induced Ca
2
influx. Simultaneous measurements
of [Ca
2
]
c
and
m
were made from astrocytes in mixed hippocampal culture (14 DIV) co-loaded with fura-2 (Ai,Bi) and Rh123
(Aii,Bii). Traces are shown from two cells in each case. A, As shown previously (Abramov et al., 2003), the [Ca
2
]
c
signals were
abolished in the absence of external Ca
2
(a saline with no added Ca
2
and with the addition of 500
MEGTA). The transient
mitochondrial depolarizations were also suppressed, but a significant slow depolarization was still evident as shown by the
analysis of different components of the response shown in C. Note that the appearance of irreversible abrupt depolarizations was
abolished and that of transient reversible depolarizations much reduced, whereas the slow increase in signal was unaffected. It
may appear to increase simply because cells are not undergoing abrupt loss of signal. B, When
A was added in the absence of
external Ca
2
,asinA,no change in [Ca
2
]
c
and only a modest mitochondrial depolarization were seen. Washing with a Ca
2
containing saline then caused the abrupt appearance of [Ca
2
]
c
fluctuations and the appearance of large mitochondrial depo-
larizations. The data shown in Dsummarize the mean changes in the normalized Rh123 signal measured at 30 min exposure to 50
M
A25-35inthepresence(Control) or absence of calcium and in the presence of 1mMZn
2
and2
Mclioquinol,whichinhibit
completely the [Ca
2
]
c
response to
A (Abramov et al., 2003). In all three cases, the mitochondrial depolarization was reduced
significantly. *p0.01; **p0.001.
568 J. Neurosci., January 14, 2004 24(2):565–575 Abramov et al.
A, Mitochondria, and Oxidative Stress in Astrocytes
the agents also completely blocked the effect of
A on mito-
chondria (Fig. 2D). The most probable explanation is that
Zn
2
and clioquinol may prevent the aggregation of
A and
its incorporation into membranes, thereby preventing its ac-
tion either on calcium signaling or on mitochondrial mem-
brane potential.
Mitochondrial substrates protect mitochondria from the
effects of
A
As shown in Figure 1A, the slow progressive mitochondrial de-
polarization seen in astrocytes could surprisingly be reversed by
the application of glutamate (from a mean change in Rh123 sig-
nal of 64.1 13.7 to 4.9 3.2%; n108 cells). This action
was unaffected by (S)-()-
-methyl-4-
carboxyphenylglycine (50
M;n178)
or ()-5-methyl- 10,11-dihydroxy -5H-
dibenzo(a,d)cyclohepten-5,10-imine (15
M;n97), inhibitors of the metabo-
tropic and ionotropic glutamate receptors,
respectively (data not shown). Glutamate
was most effective at concentrations of
15m
M, whereas it had smaller effects at
the lower (1300
M) concentrations ex-
pected to saturate receptors, suggesting
that glutamate is acting as a metabolic
substrate rather than as a neurotrans-
mitter. We therefore explored the ac-
tions of substrates for mitochondrial
complex I (1 mMglutamate; also pyruvate
and malate, 5 mMeach), complex II [10 mM
methyl-succinate (me-succinate)], a mem-
brane-permeant metabolizable form of suc-
cinate (Maechler and Wollheim, 1999),
and complex IV [2,3,5,6-tetramethyl-p-
phenylenediamine (TMPD) (200
M)with1
mMascorbate]. These all increased the mito-
chondrial potential at rest (seen as a de-
crease, an increased quench, of the Rh123
signal) when added before the
A (Fig.
3A,B) (TMPD trace not shown). Glutamate,
pyruvate/malate, and me-succinate all sup-
pressed the slow mitochondrial depolariza-
tion in response to
A, as seen in the traces of
Figure 3, Aand B. Some images extracted
from a sequential image series in the pres-
ence of me-succinate are in Figure 3Dto
show the remaining transient mitochondrial
depolarizations. Note that cells that light
upintermittently during the exposure to
A also become bright with FCCP, showing
that recovery of signal must represent recov-
ery of potential and reaccumulation of dye
into the mitochondria. Additional mito-
chondrial substrate also dramatically re-
duced the percentage of cells that failed to
recover from abrupt mitochondrial depolar-
ization shown in an analysis of the compo-
nents of the response for glutamate in Figure
3E. Glutamate (500
M-1 mM,), pyruvate/
malate (5 mMof each), and me-succinate
(p0.01; n99) all significantly reduced
the
A-induced mitochondrial depolariza-
tion measured as the mean increase of Rh123
signal after 30 min of exposure to
A from an increase of Rh123
signal by 64.1 13.7 to 7.12 2.34% for glutamate (n194 cells;
p0.01), to 6.1 2.8%, for pyruvate/malate (n78 cells; p
0.01), and to 18.6 4.17% for me-succinate, (n99; p0.01) (Fig.
3F) in cortical and hippocampal astrocytes. In contrast, and perhaps
surprisingly, TMPD/ascorbate (n87) had no significant effect on
the changes in
m
signal induced by
A in either cortical or hip-
pocampal astrocytes (Fig. 3F) (this will be discussed later). The com-
bined effect of A
causing calcium influx associated with fast tran-
sient mitochondrial depolarizations, and a metabolic effect that
causes a slow loss of potential, is emphasized by the experiment
illustrated in Figure 3C, in which A
was applied in the presence of 1
Figure 3. The slow mitochondrial depolarization induced by
A is reversed by mitochondrial substrates. Changes in Rh123
signal were measured in a culture of cortical astrocytes in response to
A 25-35 (50
M) in the presence of 1 mMglutamate ( A),
10 mMmethyl-succinate ( B), and 1 mMglutamate in a Ca
2
-free saline with the addition of 500
MEGTA ( C). In the presence of
either glutamate or me-succinate, fast reversible transient mitochondrial depolarizations were still seen, but the slow progressive
mitochondrial depolarization was abolished almost completely; compare with the day-matched control trace shown in Aas a
dashed line. In the presence of glutamate and the absence of external Ca
2
, the entire response was abolished. In Dare shown a
series of images extracted from a time sequence showing changes in
m
in response to
A in the presence of methyl succinate.
Note that the Rh123 signal in some cells becomes transiently very bright but then is restored and shows a further increase with
FCCP in the final image. The final image with FCCP shows the first image acquired immediately after application of FCCP, because
the signal continued to get even brighter with a saturation of the display at this range. The time of each image in the sequence is
indicated (minutes). An analysis of the effect of glutamate on different components of the response is shown in E, and the
measurements of the mean increase in Rh123 fluorescence at 30 min in response to
A in the presence of glutamate, methyl
succinate, and TMPD/ascorbate are shown in F.*p0.01; **p0.001.
Abramov et al.
A, Mitochondria, and Oxidative Stress in Astrocytes J. Neurosci., January 14, 2004 24(2):565–575 569
mMglutamate and in the absence of external
calcium. Under these conditions, the effect
of A
on the mitochondria was abolished
almost completely (Fig. 3C)(n79).
In all of these experiments, [Ca
2
]
c
and
mitochondrial potential were measured si-
multaneously. The [Ca
2
]
c
signal was not
altered significantly by the presence of glu-
tamate or other mitochondrial substrates.
Role of ROS in the mitochondrial
depolarization induced by
A
As discussed above, there is ample evi-
dence to implicate ROS in the toxicity of
A
. We wished to explore whether the
change in
m
was a response to or a cause
of increased ROS generation. To investi-
gate the role of ROS in the collapse of
m
of cortical or hippocampal astrocytes pro-
duced by
A, we used the antioxidant
compound TEMPO. TEMPO is a cell-
permeable nitroxide that acts to catalyze
O
2
dismutation (Samuni et al., 1988),
and it has been shown to remove oxidant
species and protect cells from oxidative damage in some model
systems (Grinberg and Samuni, 1994; Vergun et al., 2001). To
avoid the toxic action of H
2
O
2
generated by O
2
dismutation,
TEMPO (500
M) was used in combination with catalase (250
U/ml), a scavenger of H
2
O
2
. The cells were incubated with
TEMPO and catalase for 30 min and then exposed to
A 1-42 or
25-35 in the presence of these antioxidants. Although the catalase
acts extracellularly, this combination has been shown to be effec-
tive in scavenging superoxide generated by xanthine/xanthine
oxidase in neurons (Vergun et al., 2001). Figure 4Ashows the
changes in
m
in response to
A 25-35 in the presence of
TEMPO plus catalase. The major effect of the antioxidants was to
suppress the slow mitochondrial depolarization (Fig. 4 B), similar
to the action of glutamate or methyl-succinate. The average
change in Rh123 signal measured after 30 min of exposure to
A
in astrocytes pretreated with TEMPO and catalase was signifi-
cantly smaller ( p0.01) than in control cells (from an increase
by 64.1 13.7% in controls to an increase of just 14.5 5.2%;
n156 cells) (Fig. 4C). We have shown previously that antioxi-
dants do not alter the [Ca
2
]
c
response of astrocytes to
A
(Abramov et al., 2003), and so this effect seems specific to the
metabolic response to A
.
Role of the mitochondrial permeability transition pore in the
mitochondrial response to
A
The fast large transient mitochondrial depolarizations seen in
cortical and hippocampal astrocytes during the
A exposure
were remarkable and unprecedented. We wondered whether they
might be induced by transient opening of the mitochondrial per-
meability transition pore (MPTP) or whether
A might be inter-
nalized and form a pore directly in the mitochondrial inner
membrane, as it does in the plasma membrane (Lin et al., 2001;
Abramov et al., 2003). The hippocampal astrocytes were incu-
bated with cyclosporin A (CsA) (0.5
M), the classic inhibitor of
MPTP opening, for 30 min and then exposed to
A 25-35, still in
the presence of the inhibitor. The data shown in Figure 5, Aand
D, show that CsA did not significantly alter the slow mitochon-
drial depolarization in
A-treated cells and only suppressed
slightly the transient depolarizations, but it completely blocked
the large abrupt and irreversible depolarizations (Fig. 5D). CsA
has no significant effect on the calcium signal (data not shown),
which was always measured simultaneously with the mitochon-
drial signals. The average amplitude of the mitochondrial depo-
larization decreased from 64.1 13.7% in control to 38.43
7.3% in CsA-treated astrocytes (n147) (Fig. 5E).
The major mechanism that triggers MPTP opening is the syn-
ergistic action of raised intramitochondrial calcium with oxida-
tive stress (for review, see Duchen, 2000; Crompton, 2000).
When CsA was combined with the antioxidant combination of
TEMPO with catalase (n107), the action of
Aon
m
was
suppressed almost completely (Fig. 5 B,E), again without altering
the calcium response (data not shown). Similarly, removing both
triggers to MPTP opening by using TEMPO plus catalase in the
absence of extracellular Ca
2
(n67 cells) also completely
blocked the
A-triggered changes in
m
(Fig. 5C). Thus, the full
effect of
A on mitochondrial potential appears to be produced
by the combined action of ROS and Ca
2
.
Sources of
A-induced ROS: a role for the NADPH oxidase
Many mechanisms have been proposed to account for increase in
ROS formation induced by
-amyloid. We have recently found
that the plasmalemmal NADPH oxidase, an enzyme expressed
predominantly in neutrophils and macrophages (for review, see
Droge, 2002), is expressed also in astrocytes (Shimohama et al.,
2000; A. Abramov and J. Jacobson, unpublished observations).
The NADPH oxidase inhibitor DPI (0.5
M) completely pre-
vented the
Ainduced mitochondrial depolarization, although
it had no apparent effect on the
A-induced changes in [Ca
2
]
c
(Fig. 6). Because DPI is not fully selective for the NADPH oxidase
but also inhibits mitochondrial NADH dehydrogenases and
therefore may inhibit mitochondrial respiration, we examined
the effect of this concentration of the drug on
m
of untreated
astrocytes and found no significant effect (n79 cells). Further-
more, because the mitochondrial ATPase could maintain the po-
tential in the face of mitochondrial inhibition by DPI, we tested
the effect of oligomycin (2.5
g/ml) on the Rh123 signal in the
presence of DPI, and this had no detectable effect (n43 cells).
This demonstrates that DPI had no significant effect on the mi-
Figure 4. The mitochondrial response to
A is suppressed by antioxidants. A, Cortical astrocytes co-loaded with Rh123 and
fura-2 were preincubated for 30 min with 500
MTEMPO plus catalase (250 U/ml) before exposure to 50
M
A 25-35. Only the
Rh123traceisshown; the fura-2 record was indistinguishable fromcontrols.Theantioxidantsremained in the chamber during the
exposure to
A. All aspects of the response were much reduced, as indicated in the analysis in B, and the mean mitochondrial
depolarization measured after 30 min in the presence of
A was reduced significantly, as shown in C.*p0.01; **p0.001.
570 J. Neurosci., January 14, 2004 24(2):565–575 Abramov et al.
A, Mitochondria, and Oxidative Stress in Astrocytes
tochondrial dehydrogenases at this concentration and appears to
act only on the plasma membrane. Another specific inhibitor of
NADPH oxidase, 4-hydroxy-3-methoxy-acetophenone (apocy-
nin), used at 1 mM(Gao et al., 2002), which binds to a different
site on the enzyme from DPI, also completely inhibited
A-
induced mitochondrial depolarization (n63 cells), further
confirming the involvement of NADPH oxidase in
A-induced
mitochondrial depolarization.
Effect of
A on ROS generation
To explore further the increase in ROS induced by
A, we used
DCF to examine directly the rate of ROS generation. DCF is
oxidized to a fluorescent product by ROS, and so we measured
the rate of increase of signal as a measure of the rate of ROS
generation.
A 25-35 (50
M)or
A 1-42 (2.5
M) caused a rapid
and substantial increase in the rate of rise of the DCF signal with
a slope of 3.19 0.18 arbitrary fluorescence units per minute, in
both cortical and hippocampal astrocytes (n88 cells) (Fig.
7A,D). It is interesting that although there was a 510 min delay
before the onset of either [Ca
2
]
c
signals or mitochondrial depo-
larization, the DCF signal increased almost immediately after the
application of
A. The rate of rise of the DCF signal induced by
A was significantly reduced when cells were preincubated in
Ca
2
-free saline (from 3.19 0.18 arbi-
trary fluorescence units per minute to
1.18 0.11; p0.001; n101 cells) (Fig.
7B). The
A-stimulated ROS production
in astrocytes was inhibited almost com-
pletely by 0.5
MDPI (from 3.19 0.18
arbitrary fluorescence units per minute to
0.61 0.14; n114 cells; p0.001) (Fig.
7C)or1m
Mapocynin (from 3.19 0.18
arbitrary fluorescence units per minute to
0.97 0.11;. n68), strongly suggesting
that the response is mediated mainly by
the activation of the NADPH oxidase.
It has been reported that, in microglia,
A activates the NADPH oxidase through
an interaction of fibrillar peptide with the
scavenger receptor CD36 (Coraci et al.,
2002). To test this, we preincubated cells
with an antibody to CD36 (5
g/ml FA6
152) for 30 min. We saw no significant dif-
ference in the rate of rise of the DCF signal
in cortical astrocytes on three separate cul-
tures (54 cells) or in hippocampal astro-
cytes (two cultures, 40 cells).
We have shown previously that
A
causes a loss of glutathione (GSH) in as-
trocytes and also in neurons in coculture
(Abramov et al., 2003), but we had not
previously identified the source of
A-
induced oxidative stress. We therefore
asked whether the depletion of GSH in as-
trocytes and neurons was related to activa-
tion of the NADPH oxidase. Using mono-
chlorobimane (MCB) (Keelan et al., 2001)
to measure GSH levels in adjacent neurons
and astrocytes in coculture, we found as
before that
A caused GSH depletion both
in astrocytes and in neurons in hippocam-
pal cocultures. This response was abol-
ished almost completely by treatment with
DPI (n127 neurons and n198 astrocytes; p0.05 for both
types of cells) (Fig. 8A,B), which had no significant effect on the
resting GSH levels in the cells. Furthermore, GSH levels could be
better maintained in the face of
A exposure by provision of the
GSH precursor 1 mM
-glutamyl cysteine (
-GluCys), which sig-
nificantly increased MCB fluorescence in
A-treated astrocytes
from 497 21 to 847 63 (n295) and in neurons from 415
23 to 556 27 (n136; p0.05) (Fig. 8A,B).
Effect of
A on cell viability and rescue by GSH precursors
We then examined the effect of incubation with
A on cell via-
bility of neurons and astrocytes. Exposure of mixed hippocampal
cultures to
A 25-35 for 24 hr caused neuronal cell death of
38.4 9.3% above that seen in untreated cells (which was 9.8
3.2%; n13 experiments) (Fig. 8C).
A caused significantly less
cell death among the astrocytes, with an increase of 15.4 4.2%
above a background of 7.8 2.9% (n13 experiments) (Fig.
8D). Interestingly, two compounds that are normally highly neu-
rotoxic,
A 25-35 and glutamate (500
M), did not have a syner-
gistic effect on neuronal cell death. In fact, glutamate was signif-
icantly protective in astrocytes, reducing
A-induced cell death
to 4.9 2.2% above background ( p0.05) (Fig. 8D). The
antioxidant combination of TEMPO and catalase afforded signif-
Figure 5. Role of the mitochondrial permeability transition pore in the mitochondrial depolarization induced by
A. A,
Hippocampal astrocytes in coculture were loaded with Rh123 and preincubated with 0.5
Mcyclosporin A before exposure to 50
M
A 25-35. Large irreversible mitochondrial depolarizations were suppressed completely, and the incidence of smaller tran-
sient and reversible depolarizations was also reduced, whereas the slow mitochondrial depolarization was not affected, as
indicatedbythe analysis of the components oftheresponse shown in D.The tracesinBshowthat combined treatment of cellswith
CsA and the antioxidants TEMPO/catalase almost completely suppressed any response to
A, as did a combination of TEMPO and
catalase combined with the removal of extracellular Ca
2
, as illustrated in C.Esummarizes measurements of the normalized
Rh123 signal at 30 min of exposure to
A with each of these manipulations. These data strongly suggest that the mitochondrial
response to
A involves the synergistic action of Ca
2
and oxidative stress leading together to the activation of the MPTP. *p
0.01; **p0.001.
Abramov et al.
A, Mitochondria, and Oxidative Stress in Astrocytes J. Neurosci., January 14, 2004 24(2):565–575 571
icant protection from
A-induced cell death in both cell types [to
16.8 4.6% for neurons ( p0.05) and 4.6 2.1% for astro-
cytes ( p0.05) for both cell types; n5 experiments], but the
most effective protection was given by TEMPO and catalase com-
bined with CsA (to 4.7 2.9% for neurons and 0.3 0.3% for
astrocytes; p0.001 for neurons and astrocytes; n5). DPI also
dramatically protected both classes of cells (to 11.1 4.3% for
neurons and to 5.1 1.1% for astrocytes; p0.05 for both type
of the cells; n4) but was less effective than the antioxidant/CsA
combination, suggesting that DPI, despite inhibiting
A-induced
ROS generation very effectively, may have a marginal toxicity.
The importance of neuronal glutathione is underscored by
some experiments in which GSH in the cultures was enriched
with the provision of the GSH precursor 1 mM
-GluCys (and see
above). Cultures were preincubated with
-GluCys for 6 hr and
then exposed to
A still in the presence of the
-GluCys for 24 hr.
This treatment significantly reduced cell death in neurons (from
40.1 8.5 to 20.4 3.2%; p0.05; n4 experiments) and in
astrocytes (from 15.4 4.2 to 7.1 1.3%; p0.05; n4
experiments) (Fig. 8C,D).
Discussion
We have found that
A treatment causes dramatic changes in
mitochondrial potential in cultures of neurons and astrocytes
from rat hippocampus. There are several remarkable features of
these responses. First, the mitochondrial responses were re-
stricted to the astrocytes, whereas no significant changes in mito-
chondrial function were seen in neurons over the time frame
investigated. The mitochondrial response appears to be a com-
plex response to several different processes taking place within
the astrocytes that are triggered by A
and seems to reflect
changes in several different but related aspects of mitochondrial
function, with oxidative stress as a common underlying factor.
The slow mitochondrial depolarization, occurring over a time
scale of minutes, is attributable to the effect of oxidative stress on
enzymes that ensure mitochondrial substrate supply. The re-
sponse could be reversed or blocked by a supply of substrates to
complex I (glutamate) or complex II (me-succinate), by antioxi-
dants, by DPI, and by apocynin. These observations show that the
response is caused by an effect of oxidative stress as a consequence
of the activation of NADPH oxidase. Enzymes of the TCA cycle,
aconitase and
-ketoglutarate dehydrogenase, have been shown
to be impaired both by
A and by oxidative stress in various
model systems (Blass and Gibson, 1991; Chinopoulos et al., 1999;
Figure 6. The mitochondrial response to
A but not the change in [Ca
2
]
c
is mediated by
the NADPH oxidase. Astrocytes were co-loaded with fura-2 and Rh123 for simultaneous mea-
surement of [Ca
2
]
c
and
m
. Cells were preincubated with 0.5
MDPI for 30 min. [Ca
2
]
c
transients were seen as usual after application of 50 mM
A, whereas the mitochondrial re-
sponse was suppressed completely.
Figure 7.
A-increased ROS generation in astrocytes is dependent on Ca
2
and activation
of the NADPH oxidase. Hippocampal cultures were loaded with DCF. A, Addition of 5
M
A
1-42 caused a clear increase in the rate of appearance of the fluorescent product, suggesting
increased ROS generation. The line shown was fitted by a linear regression with a correlation
coefficient of 0.98. The rate of increase of ROS production was reduced significantly in the
absence of external Ca
2
(B) and was blocked almost completely by 0.5
MDPI ( C). The data
are summarized in D,which shows the mean rate of rise of DCF signal under these conditions
and also after exposure to 1 mMapocynin, another inhibitor of the NADPH oxidase. *p0.01.
572 J. Neurosci., January 14, 2004 24(2):565–575 Abramov et al.
A, Mitochondria, and Oxidative Stress in Astrocytes
Casley et al., 2002b). The observation that the mitochondrial
potential could be recovered almost completely by substrates of
complex I argues that that this effect was not caused by impaired
activity of the respiratory complexes and that it must lie upstream
of mitochondrial metabolism. Both glucose uptake and glycolysis
are impaired by exposure to
A (Mark et al., 1997; Pereira et al.,
1999), providing the most plausible mechanism for these obser-
vations. TMPD/ascorbate should bypass other components of
the respiratory chain but failed to reverse the effect. It has been
shown previously that the K
m
of cytochrome oxidase for cyto-
chrome cmay be raised by
A (Casley et al., 2002b). Because the
TMPD/ascorbate pair supplies electrons to cytochrome c, this
may account for the failure of the TMPD/ascorbate combination
to restore potential. It seems that
A-induced oxidative stress
causes impairments of metabolism at multiple sites through in-
termediary metabolism, but the failure of the glycolytic supply of
mitochondrial substrate appears to be the first to become func-
tionally significant. We have also shown previously that
A
causes impaired activity of several complexes of the mitochon-
drial respiratory chain (Canevari et al., 1999; Casley et al., 2002a).
This in turn may act as an auto-amplifying mechanism, because
impaired mitochondrial respiration may act as yet another source
of intramitochondrial free radical generation.
The other dramatic feature of the mitochondrial response was
the occurrence of profound abrupt and reversible mitochondrial
depolarizations. These effects were remarkable and unprece-
dented. There seem to be two probable explanations for such
abrupt and reversible changes in potential, both based on tran-
sient openings of a large conductance pore. One is that the
Ais
internalized and inserts into the mitochondrial membrane where
it acts as a pore much as it does at the plasma membrane. This is
difficult to test, but seems unlikely, because the protein would
have to cross the outer mitochondrial membrane and insert itself
into the inner membrane. The other, perhaps more probable
explanation is that the combination of transient pulses of high
[Ca
2
]
c
with an underlying oxidative stress suffices to cause tran-
sient openings of the MPTP. CsA suppressed the large irreversible
mitochondrial depolarizations almost completely, suggesting
that these reflect MPTP opening and cell death. The effect of CsA
on the smaller, reversible transient depolarizations was unim-
pressive, however, suggesting that these might still reflect another
process. The observation that the transients were suppressed so
completely by those manipulations that suppressed oxidative
stressDPI, apocynin, TEMPO/catalasewithout altering the
[Ca
2
]
c
signal argues strongly that the events cannot be a sim-
pleconsequence of Ca
2
influx and also argues strongly against
the insertion of
A as a pore in the mitochondrial membrane.
The events were also inhibited effectively in the absence of external
calcium and appeared to be temporally related to [Ca
2
]
c
transients,
again suggesting a possible role for the MPTP, although the calcium
dependence could also reflect the calcium dependence of the rate of
ROS generation. These experiments show clearly that the transient
depolarizations are strongly dependent both on calcium and on ox-
idative stress and that, although both are required, neither alone is
sufficient to cause them.
Thus, the underlying key to the response seems to be an oxi-
dative stress generated by
A, and the various effects on mito-
chondria are attributable to the consequences of oxidative stress
on different aspects of mitochondrial function in the astrocytes.
The key question then becomes what is the mechanism of oxida-
tive stress, of the selectivity of the
A action for astrocytes, and
finally of neuronal cell death. We have found recently that astro-
cytes selectively express the plasmalemmal NADPH oxidase (our
unpublished observations). It is not clear what normally regu-
lates the activity of this enzyme or what its role in the CNS might
be. Nevertheless, the dramatic protection afforded by DPI, at
concentrations that do not affect other potential targets (mostly
mitochondrial dehydrogenases), and by apocynin, an agent that
binds at a separate site on the target enzyme, strongly suggests
that this enzyme represents a major source of radical species in
response to
A and that the selective expression of the enzyme
might account in some measure for the selective effects of
Aon
astrocyte, and not neuronal, mitochondria. The importance of
the NADPH oxidase in
A toxicity has already been suggested
but with reference to the expression of the enzyme in microglia
(Bianca et al., 1999; Qin et al., 2002). In the present study, how-
ever, astrocytic ROS generation seems to play a critical role in
A
neurotoxicity, suggesting that the astrocyte may have been over-
looked as a source of ROS. The imaging of DCF appears to us to
provide unequivocal evidence identifying the astrocytes as the
source of ROS. Thus, we measure ROS generation from astro-
cytes in pure astrocyte cultures (99% GFAP positive). In mixed
cultures, an increase in DCF signal is seen over astrocytes and not
Figure 8. GSH depletion and cell death induced by
A are dependent on free radical gener-
ation by NADPH oxidase.
A 25-35 or 1-42 caused GSH depletion in neurons ( A) and astrocytes
(B) measured in hippocampal cocultures loaded with monochlorobimane. This was prevented
almost completely by 0.5
MDPI and by 1 mMapocynin and was reversed by provision of 1 mM
-glutamyl-cysteine.
A 25-35 or 1-42 also increased cell death of neurons ( C)to40% and
astrocytes ( D)to15% above background. In both cell types, cell death was reduced signifi-
cantly by all manipulations that also suppressed the mitochondrial depolarization in astrocytes:
in the absence of external Ca
2
, with the addition of antioxidants, and in the presence of 0.5
MDPI or 1 mMapocynin. The most effective protection was afforded by the combination of
antioxidants and CsA. The addition of 1 mMglutamate suppressed
A-induced cell death in
astrocytes and made no significant difference to cell death in the neurons. Both neurons and
astrocytes were also substantially protected from death induced by
A by previous incubation
with 1 mM
-GluCys (see Results). *p0.01; **p0.001.
Abramov et al.
A, Mitochondria, and Oxidative Stress in Astrocytes J. Neurosci., January 14, 2004 24(2):565–575 573
over nearby neurons, showing that the response cannot arise as
some diffusion (itself unlikely) of ROS from occasional contam-
inating microglia.
We do not yet understand how
A might activate the NADPH
oxidase. The dependence of the rate of rise of ROS generation on
calcium suggests that the
A-induced increase in [Ca
2
]
c
might
play a role in activating the enzyme. It also seemed possible that
the
A might interact with the enzyme through the CD36 scav-
enger receptors as reported for microglia (Coraci et al., 2002). We
did not see any effect of antibodies to CD36 on the
A-induced
increase in astrocyte ROS generation in these experiments, again
highlighting the importance of expression of the enzyme in as-
trocytes as well as in microglia. One isoform of NADPH oxidase,
nox5, expressed in testis, spleen, and lymph nodes (Banfi et al.,
2001), is activated by a rise in [Ca
2
]
c
. Our data do not clearly
resolve this issue. That the rate of ROS generation tends to in-
crease almost immediately on
A application whereas the
[Ca
2
]
c
changes occur with a delay of 510 min strongly argues
against a Ca
2
dependence of NADPH oxidase. However, the
Ca
2
dependence of the
A-induced ROS generation and of
GSH oxidation is consistent with Ca
2
-dependent modulation
of the NADPH oxidase. This issue will require additional work to
clarify the specific relationship between changes in [Ca
2
]
c
and
ROS generation. These data nevertheless suggest that these two
armsof the response to
A, the influx of Ca
2
and the activa-
tion of NADPH oxidase and generation of oxidative stress, and
their impact on intracellular metabolic pathways and mitochon-
drial function are coupled.
Why do neurons die more than the astrocytes even though the
apparent pathophysiology takes place in astrocytes? We have
shown previously that
A reduces astrocyte glutathione concen-
trations. Astrocytes supply amino acid precursors for neuronal
GSH synthesis, and neurons depleted of GSH die through an
inability to withstand endogenous pro-oxidants. In turn, astro-
cytes can sustain themselves adequately with glycolytic metabo-
lism, seem not to be significantly affected by loss of mitochon-
drial function, and are also far more resistant to oxidative stress
than are the neurons. The importance of GSH as a protective
mechanism is again emphasized by the protection afforded by the
GSH precursor
-glutamyl cysteine. It remains possible that neu-
ronal death is further mediated by other mechanisms, perhaps
through cytokines released from damaged astrocytes or perhaps
through increased vulnerability of neurons to other stimuli.
These data suggest, however, that a major part of
A neurotox-
icity reflects the neuronal dependence on astrocyte support and
results from a complex interplay between actions of the peptide
on astrocytes, altered astrocyte [Ca
2
]
c
signaling, free radical
generation, and resultant oxidative stress. This in turn impairs
the ability of astrocytes to maintain neuronal integrity and results
eventually in neuronal attrition.
References
Abramov AY, Canevari L, Duchen MR (2003) Changes in [Ca
2
]
c
and glu-
tathione in astrocytes as the primary mechanism of amyloid neurotoxic-
ity. J Neurosci 23:50885095.
Anandatheerthavarada HK, Biswas G, Robin MA, Avadhani NG (2003) Mi-
tochondrial targeting and a novel transmembrane arrest of Alzheimers
amyloid precursor protein impair mitochondrial function in neuronal
cells. J Cell Biol 161:114.
Arias C, Montiel T, Quiroz-Baez R, Massieu L (2002) beta-Amyloid neuro-
toxicity is exacerbated during glycolysis inhibition and mitochondrial
impairment in the rat hippocampus in vivo and in isolated nerve termi-
nals: implications for Alzheimers disease. Exp Neurol 176:163174.
Askanas V, McFerrin J, Baque S, Alvarez RB, Sarkozi E, Engel WK (1996)
Transfer of beta-amyloid precursor protein gene using adenovirus vector
causes mitochondrial abnormalities in cultured normal human muscle.
Proc Natl Acad Sci USA 93:13141319.
Banfi B, Molnar G, Maturana A, Steger K, Hegedus B, Demaurex N, Krause
KH (2001) A Ca
2
-activated NADPH oxidase in testis, spleen, and
lymph nodes. J Biol Chem 276:3759437601.
Bianca VD, Dusi S, Bianchini E, Dal Pra I, Rossi F (1999) beta-Amyloid
activates the O-2 forming NADPH oxidase in microglia, monocytes, and
neutrophils. A possible inflammatory mechanism of neuronal damage in
Alzheimers disease. J Biol Chem 274:1549315499.
Blass JP, Gibson SA (1991) The role of oxidative abnormalities in the patho-
physiology of Alzheimers disease. Rev Neurol (Paris) 147:513525.
Canevari L, Clark JB, Bates TE (1999)
-Amyloid fragment 2535 selec-
tively decreases complex IV activity in isolated mitochondria. FEBS Lett
457:131134.
Casley C, Land J, Sharpe M, Clark J, Duchen M, Canevari L (2002a)
-Amyloid fragment 2535 causes mitochondrial dysfunction in primary
cortical neurons. Neurobiol Dis 10:258267.
Casley CS, Canevari L, Land JM, Clark JB, Sharpe MA (2002b)
-Amyloid
inhibits integrated mitochondrial respiration and key enzyme activities.
J Neurochem 80:91100.
Chinopoulos C, Tretter L, Adam-Vizi V (1999) Depolarization of in situ
mitochondria due to hydrogen peroxide-induced oxidative stress in nerve
terminals: inhibition of alpha-ketoglutarate dehydrogenase. J Neuro-
chem 73:220228.
Coraci IS, Husemann J, Berman JW, Hulette C, Dufour JH, Campanella GK,
Luster AD, Silverstein SC, El-Khoury JB (2002) CD36, a class B scaven-
ger receptor, is expressed on microglia in Alzheimers disease brains and
can mediate production of reactive oxygen species in response to beta-
amyloid fibrils. Am J Pathol 160:101112.
Crompton M (2000) . Mitochondrial intermembrane junctional complexes
and their role in cell death. J Physiol (Lond) 529:1121.
Droge W (2002) Free radicals in the physiological control of cell function.
Physiol Rev 82:4795.
Duchen MR (1999) Contributions of mitochondria to animal physiology:
from homeostatic sensor to calcium signaling and cell death. J Physiol
(Lond) 516:117.
Duchen MR (2000) Mitochondria and calcium: from cell signaling to cell
death. J Physiol (Lond) 529:5768.
Duchen MR, Biscoe TJ (1992) Relative mitochondrial membrane potential
and [Ca
2
]
i
in type I cells isolated from the rabbit carotid body. J Physiol
(Lond) 450:3361.
Gao HM, Hong JS, Zhang W, Liu B (2002) Distinct role for microglia in
rotenone-induced degeneration of dopaminergic neurons. J Neurosci
22:782790.
Grant SM, Shankar SL, Chalmers-Redman RM, Tatton WG, Szyf M, Cuello
AC (1999) Mitochondrial abnormalities in neuroectodermal cells stably
expressing human amyloid precursor protein (hAPP751). NeuroReport
10:4146.
Grinberg LN, Samuni A (1994) Nitroxide stable radical prevents
primaquine-induced lysis of red blood cell. BBA 1201:284288.
Hirai K, Aliev G, Nunomura A, Fujioka H, Russell RM, Atwood CS, Johnson
AB, Kress Y, Vinters HV, Tabaton M, Shimohama S, Cash AD, Siedlak SL,
Harris PL, Jones PK, Petersen PB, Perry G, Smith MA (2001) Mitochon-
drial abnormalities in Alzheimers disease. J Neurosci 21:30173023.
Keelan J, Allen NJ, Antcliffe D, Pal S, Duchen MR (2001) Quantitative im-
aging of glutathione in hippocampal neurons and glia in culture using
monochlorobimane. J Neurosci Res 66:873884.
Khodorov B, Pinelis V, Vergun O, Storozhevykh T, Vinskaya N (1996) Mi-
tochondrial deenergization underlies neuronal calcium overload follow-
ing a prolonged glutamate challenge. FEBS Lett 397:230234.
Kim HS, Lee JH, Lee JP, Kim EM, Chang KA, Park CH, Jeong CJ, Wittendorp
MC, Seo JH, Choi SH, Suh YH (2002) Amyloid
peptide induces cyto-
chrome crelease from isolated mitochondria. NeuroReport
13:19891993.
Lin H, Bhatia R, Lal R (2001) Amyloid
protein forms ion channels: impli-
cations for Alzheimers disease pathophysiology. FASEB J 15:24332444.
Longo VD, Viola KL, Klein WL, Finch CE (2000) Reversible inactivation of
superoxide-sensitive aconitase in Abeta142-treated neuronal cell lines.
J Neurochem 75:19771985.
Maechler P, Wollheim CB (1999) Mitochondrial glutamate acts as a mes-
senger in glucose-induced insulin exocytosis. Nature 402:685689.
Mark RJ, Pang Z, Geddes JW, Uchida K, Mattson MP (1997) Amyloid beta-
574 J. Neurosci., January 14, 2004 24(2):565–575 Abramov et al.
A, Mitochondria, and Oxidative Stress in Astrocytes
peptide impairs glucose transport in hippocampal and cortical neurons:
involvement of membrane lipid peroxidation. J Neurosci 17:10461054.
Morais Cardoso S, Swerdlow RH, Oliveira CR (2002) Induction of cyto-
chrome c-mediated apoptosis by amyloid beta 25-35 requires functional
mitochondria. Brain Res 931:117125.
Moreira PI, Santos MS, Moreno A, Rego AC, Oliveira C (2002) Effect of
amyloid beta-peptide on permeability transition pore: a comparative
study. J Neurosci Res 69:257267.
Naslund J, Haroutunian V, Mohs R, Davis KL, Davies P, Greengard P, Bux-
baum JD (2000) Correlation between elevated levels of amyloid beta-
peptide in the brain and cognitive decline. JAMA 283:15711577.
Parks JK, Smith TS, Trimmer PA, Bennett JP, Parker WD (2001) Neuro-
toxic Abeta peptides increase oxidative stress in vivo through NMDA-
receptor and nitric-oxide-synthase mechanisms, and inhibit complex IV
activity and induce a mitochondrial permeability transition in vitro.
J Neurochem 76:10501056.
Pereira C, Santos MS, Oliveira C (1999) Involvement of oxidative stress on
the impairment of energy metabolism induced by A beta peptides on
PC12 cells: protection by antioxidants. Neurobiol Dis 6:209219.
Qin L, Liu Y, Cooper C, Liu B, Wilson B, Hong JS (2002) Microglia enhance
beta-amyloid peptide-induced toxicity in cortical and mesencephalic
neurons by producing reactive oxygen species. J Neurochem 83:973983.
Samuni A, Krishna CM, Riesz P, Finkelstein E, Russo A (1988) A novel
metal-free low molecular weight superoxide dismutase mimic. J Biol
Chem 263:1792117924.
Sheehan JP, Swerdlow RH, Miller SW, Davis RE, Parks JK, Parker WD, Tuttle
JB (1997) Calcium homeostasis and reactive oxygen species production
in cells transformed by mitochondria from individuals with sporadic Alz-
heimers disease. J Neurosci 17:46124622.
Shevtzova EF, Kireeva EG, Bachurin SO (2001) Effect of beta-amyloid pep-
tide fragment 2535 on nonselective permeability of mitochondria. Bull
Exp Biol Med 132:11731176.
Shimohama S, Tanino H, Kawakami N, Okamura N, Kodama H, Yamaguchi
T, Hayakawa T, Nunomura A, Chiba S, Perry G, Smith MA, Fujimoto S
(2000) Activation of NADPH oxidase in Alzheimers disease brains. Bio-
chem Biophys Res Commun 273:59.
Vergun O, Keelan J, Khodorov BI, Duchen MR (1999) Glutamate-induced
mitochondrial depolarization and perturbation of calcium homeostasis
in cultured rat hippocampal neurones. J Physiol (Lond) 519:451466.
Vergun O, Sobolevsky AI, Yelshansky MV, Keelan J, Khodorov BI, Duchen
MR (2001) Exploration of the role of reactive oxygen species in gluta-
mate neurotoxicity in rat hippocampal neurones in culture. J Physiol
(Lond) 531:147163.
Abramov et al.
A, Mitochondria, and Oxidative Stress in Astrocytes J. Neurosci., January 14, 2004 24(2):565–575 575
... Indeed, it has been reported that exposure to neurotoxic Aβ peptides can result in many of the observed components of AD, including excitotoxicity, neuroinflammation, oxidative stress, and heightened neurodegeneration [14][15][16]. Intracellular accumulation of Aβ can directly dysregulate several processes including, but not limited to, cellular respiration (via mitochondrial dysfunction and oxidative stress) [17][18][19], autophagy-lysosomal pathways [20,21], and kinase signaling [22][23][24]. Critically, the relationship between Aβ aggregation and ensuing disease pathology is bidirectional and cyclical, constituting a vicious cycle that promotes neuronal death and loss of neuronal network connectivity. ...
... Mitochondrial dysfunction is a major pathogenic event underlying Aβ toxicity [2,17,18,41,42]. To assess whether the neuroprotective effects of fosgonimeton could attenuate Aβ-mediated mitochondrial dysfunction and characterize a potential mechanism by which fosgonimeton mitigates neurotoxicity, primary cortical neurons were treated with fosgo-AM for 15 min and challenged with Aβ 1-42 for 4 h (a time point at which Aβ-induced neuronal death has yet to occur [19]) and immunoassayed with anti-MAP2 and either MitoSox (an indicator of ROS generated by the mitochondria) ( Fig. 2a and b), or anti-CytC (marker of cytochrome c) ( Fig. 2c and d). ...
... Aβ 1-42 -induced cytotoxicity has been associated with disruption of mitochondrial membrane potential and production of ROS, leading to the induction of apoptotic mechanisms [2,17,18,41,42]. Upregulation of ROS and apoptotic signaling trigger several neurodegenerative events including synapse loss, neurite degeneration, pathological protein accumulation, and neuroinflammation. ...
Article
Full-text available
Positive modulation of hepatocyte growth factor (HGF) signaling may represent a promising therapeutic strategy for Alzheimer's disease (AD) based on its multimodal neurotrophic, neuroprotective, and anti-inflammatory effects addressing the complex pathophysiology of neurodegeneration. Fosgonimeton is a small-molecule positive modulator of the HGF system that has demonstrated neurotrophic and pro-cognitive effects in preclinical models of dementia. Herein, we evaluate the neuroprotective potential of fosgonimeton, or its active metabolite, fosgo-AM, in amyloid-beta (Aβ)-driven preclinical models of AD, providing mechanistic insight into its mode of action. In primary rat cortical neurons challenged with Aβ (Aβ1-42), fosgo-AM treatment significantly improved neuronal survival, protected neurite networks, and reduced tau hyperphosphorylation. Interrogation of intracellular events indicated that cortical neurons treated with fosgo-AM exhibited a significant decrease in mitochondrial oxidative stress and cytochrome c release. Following Aβ injury, fosgo-AM significantly enhanced activation of pro-survival effectors ERK and AKT, and reduced activity of GSK3β, one of the main kinases involved in tau hyperphosphorylation. Fosgo-AM also mitigated Aβ-induced deficits in Unc-like kinase 1 (ULK1) and Beclin-1, suggesting a potential effect on autophagy. Treatment with fosgo-AM protected cortical neurons from glutamate excitotoxicity, and such effects were abolished in the presence of an AKT or MEK/ERK inhibitor. In vivo, fosgonimeton administration led to functional improvement in an intracerebroventricular Aβ25-35 rat model of AD, as it significantly rescued cognitive function in the passive avoidance test. Together, our data demonstrate the ability of fosgonimeton to counteract mechanisms of Aβ-induced toxicity. Fosgonimeton is currently in clinical trials for mild-to-moderate AD (NCT04488419; NCT04886063).
... Changes in the balance between production of ROS and efficiency of antioxidant system in brain cells can be a trigger for pathology and cell death that also can possibly be regulated by intracellular signaling [6]. One of the major producers of ROS in the brain cells, NADPH oxidase, is involved in processes of physiology and pathology [7][8][9]. Although activation of NADPH oxidase (NOX) in different cells is shown through receptor dependent and receptor independent pathways, as a result of involvement in phagocytic processes, it can be activated by abnormal peptides and proteins, including Ab [8,10]. ...
... One of the major producers of ROS in the brain cells, NADPH oxidase, is involved in processes of physiology and pathology [7][8][9]. Although activation of NADPH oxidase (NOX) in different cells is shown through receptor dependent and receptor independent pathways, as a result of involvement in phagocytic processes, it can be activated by abnormal peptides and proteins, including Ab [8,10]. RAGE was shown to be able to increase expression of the major NOX isoforms in the brain: NOX2, NOX 4 and NOX1 [11][12][13]. ...
... Dialogue about the possible regulation of NOX in brain cells by RAGE activation has been ongoing for a relatively long time considering the increase in ROS production in response to damage-associated and pathogen-associated molecules [22][23][24]. However, most of the damage-associated and pathogenassociated molecules able to activate RAGE are shown to be able to induce ROS in NOX through various mechanisms excluding RAGE, but, at the same time, inhibitors of this enzyme are protective in AGE-pathologies [8,25,26]. In the present study, we used short peptides from the RAGE V-domain that are unable to induce any cell death in primary cocultures of neurons and astrocytes (Fig. 7). ...
Article
Full-text available
The transmembrane receptor for advanced glycation end products (RAGE) is a signaling receptor for many damage‐ and pathogen‐associated molecules. Activation of RAGE is associated with inflammation and an increase in reactive oxygen species (ROS) production. Although several sources of ROS have been previously suggested, how RAGE induces ROS production is still unclear, considering the multiple targets of pathogen‐associated molecules. Here, using acute brain slices and primary co‐culture of cortical neurons and astrocytes, we investigated the effects of a range of synthetic peptides corresponding to the fragments of the RAGE V‐domain on redox signaling. We found that the synthetic fragment (60–76) of the RAGE V‐domain induces activation of ROS production in astrocytes and neurons from the primary co‐culture and acute brain slices. This effect occurred through activation of RAGE and could be blocked by a RAGE inhibitor. Activation of RAGE by the synthetic fragment stimulates ROS production in NADPH oxidase (NOX). This RAGE‐induced NOX activation produced only minor decreases in glutathione levels and increased the rate of lipid peroxidation, although it also reduced basal and β‐amyloid induced cell death in neurons and astrocytes. Thus, specific activation of RAGE induces redox signaling through NOX, which can be a part of a cell protective mechanism.
... In AD, Aβ plays a pivotal role in initiating oxidative stress, and astrocytes seem to be involved in this process. When exposed to Aβ in co-cultures with neurons, rat astrocytes exhibit a decline in mitochondrial membrane potential and an increase in ROS production [84]. Importantly, this surge in ROS is attributed to the activation of astrocytic nicotinamide adenine dinucleotide phosphate (NADPH) oxidase, as demonstrated by the fact that inhibiting this enzyme blocks ROS production and provides protection to neurons. ...
Article
Full-text available
Introduction: Historically, astrocytes were seen primarily as a supportive cell population within the brain; with neurodegenerative disease research focusing exclusively on malfunctioning neurons. However, astrocytes perform numerous tasks that are essential for maintenance of the central nervous system`s complex processes. Disruption of these functions can have negative consequences; hence, it is unsurprising to observe a growing amount of evidence for the essential role of astrocytes in the development and progression of neurodegenerative diseases. Targeting astrocytic functions may serve as a potential disease-modifying drug therapy in the future. Areas covered: The present review emphasizes the key astrocytic functions associated with neurodegenerative diseases and explores the possibility of pharmaceutical interventions to modify these processes. In addition, the authors provide an overview of current advancement in this field by including studies of possible drug candidates. Expert opinion: Glial research has experienced a significant renaissance in the last quarter-century. Understanding how disease pathologies modify or are caused by astrocyte functions is crucial when developing treatments for brain diseases. Future research will focus on building advanced models that can more precisely correlate to the state in the human brain, with the goal of routinely testing therapies in these models.
Article
Alteration of mitochondrial metabolism by various mutations or toxins leads to various neurological conditions. Age-related changes in energy metabolism could also play the role of a trigger for neurodegenerative disorders. Nonetheless, it is not clear if restoration of ATP production or supplementation of brain cells with substrates for energy production could be neuroprotective. Using primary neurons and astrocytes, and neurons with familial forms of neurodegenerative disorders we studied whether various substrates of energy metabolism could improve mitochondrial metabolism and stimulate ATP production, and whether increased ATP levels could protect cells against glutamate excitotoxicity and neurodegeneration. We found that supplementation of neurons with several substrates, or combination thereof, for the TCA cycle and cellular respiration, and oxidative phosphorylation resulted in an increase in mitochondrial NADH level and in mitochondrial membrane potential and led to an increased level of ATP in neurons and astrocytes. Subsequently, these cells were protected against energy deprivation during ischemia or glutamate excitotoxicity. Provision of substrates for energy metabolism to cells with familial forms of Parkinson's disease also prevented triggering of cell death. Thus, restoration of energy metabolism and increase of ATP production can play neuroprotective role in neurodegeneration. A combination of a succinate salt of choline and nicotinamide provided the best results.
Article
Significance: The NADPH oxidase (NOX) enzyme family, located in the central nervous system (CNS), is recognized as a source of reactive oxygen species (ROS) in the brain. Despite its importance in cellular processes, excessive ROS generation leads to cell death and is involved in the pathogenesis of neurodegenerative disorders. Recent advances: NOX enzymes contribute to the development of neurodegenerative diseases, such as Parkinson's disease (PD), Alzheimer's disease (AD), amyotrophic lateral sclerosis (ALS), and stroke, highlighting their potential as targets for future therapeutic development. This review will discuss NOX's contribution and therapeutic targeting potential in neurodegenerative diseases, focusing on PD, AD, ALS, and Stroke. Critical issues: Homeostatic and physiological levels of ROS are crucial for regulating several processes, such as development, memory, neuronal signaling, and vascular homeostasis. However, NOX-mediated excessive ROS generation is deeply involved in the damage of DNA, proteins, and lipids, leading to cell death in the pathogenesis of a wide range of diseases, namely neurodegenerative diseases. Future directions: It is essential to understand the role of NOX homologs in neurodegenerative disorders and the pathological mechanisms undergoing neurodegeneration mediated by increased levels of ROS. This further knowledge will allow the development of new specific NOX inhibitors and their application for neurodegenerative disease therapeutics.
Article
Full-text available
Neurodegenerative diseases are chronic conditions occurring when neurons die in specific brain regions that lead to loss of movement or cognitive functions. Despite the progress in understanding the mechanisms of this pathology, currently no cure exists to treat these types of diseases: for some of them the only help is alleviating the associated symptoms. Mitochondrial dysfunction has been shown to be involved in the pathogenesis of most the neurodegenerative disorders. The fast and transient permeability of mitochondria (the mitochondrial permeability transition, mPT) has been shown to be an initial step in the mechanism of apoptotic and necrotic cell death, which acts as a regulator of tissue regeneration for postmitotic neurons as it leads to the irreparable loss of cells and cell function. In this study, we review the role of the mitochondrial permeability transition in neuronal death in major neurodegenerative diseases, covering the inductors of mPTP opening in neurons, including the major ones—free radicals and calcium—and we discuss perspectives and difficulties in the development of a neuroprotective strategy based on the inhibition of mPTP in neurodegenerative disorders.
Article
Full-text available
Mitochondria are thought to have become incorporated within the eukaryotic cell approximately 2 billion years ago and play a role in a variety of cellular processes, such as energy production, calcium buffering and homeostasis, steroid synthesis, cell growth, and apoptosis, as well as inflammation and ROS production. Considering that mitochondria are involved in a multitude of cellular processes, mitochondrial dysfunction has been shown to play a role within several age-related diseases, including cancers, diabetes (type 2), and neurodegenerative diseases, although the underlying mechanisms are not entirely understood. The significant increase in lifespan and increased incidence of age-related diseases over recent decades has confirmed the necessity to understand the mechanisms by which mitochondrial dysfunction impacts the process of aging and age-related diseases. In this review, we will offer a brief overview of mitochondria, along with structure and function of this important organelle. We will then discuss the cause and consequence of mitochondrial dysfunction in the aging process, with a particular focus on its role in inflammation, cognitive decline, and neurodegenerative diseases, such as Huntington’s disease, Parkinson’s disease, and Alzheimer’s disease. We will offer insight into therapies and interventions currently used to preserve or restore mitochondrial functioning during aging and neurodegeneration.
Article
Background Cataract is the leading cause of blindness around the world. Previous investigations have assessed the relationship between cataract, cataract surgery and dementia risk, but their results remain controversial. Herein, we conducted a meta‐analysis to evaluate the associations between cataract, cataract surgery and the risk of dementia. Methods We systemically screened the literature from three electronic databases PubMed, EMBASE and CENTRAL until April 2023. The data were collected by two independent researchers. The hazard ratios (HRs) or odds ratios (ORs) from eligible studies with 95% confidence intervals (CIs) were adjusted into the risk ratios (RRs), which were pooled using the random‐effects model. Results A total of nine studies with 448,140 participants reported the associations between cataract or cataract surgery and the risk of dementia were included in this meta‐analysis. The outcomes of our pooled analysis indicated that cataract was associated with an increased risk of all‐cause dementia (RR = 1.24, 95% CI, 1.14–1.35, p < .00001), Alzheimer's disease (RR = 1.22, 95% CI, 1.10–1.35, p = .0002) and vascular dementia (RR = 1.29, 95% CI, 1.01–1.66, p = .04). Cataract surgery is associated with a reduction of the dementia risk (RR = 0.74, 95% CI, 0.67–0.81, p < .00001). Conclusions Current evidence from the existing studies supports that cataract is associated with an increased risk of dementia, and cataract surgery may be instrumental in reducing the risk of dementia in patients with cataract.
Article
Full-text available
The finding that oxidative damage, including that to nucleic acids, in Alzheimer's disease is primarily limited to the cytoplasm of susceptible neuronal populations suggests that mitochondrial abnormalities might be part of the spectrum of chronic oxidative stress of Alzheimer's disease. In this study, we used in situ hybridization to mitochondrial DNA (mtDNA), immunocytochemistry of cytochrome oxidase, and morphometry of electron micrographs of biopsy specimens to determine whether there are mitochondrial abnormalities in Alzheimer's disease and their relationship to oxidative damage marked by 8-hydroxyguanosine and nitrotyrosine. We found that the same neurons showing increased oxidative damage in Alzheimer's disease have a striking and significant increase in mtDNA and cytochrome oxidase. Surprisingly, much of the mtDNA and cytochrome oxidase is found in the neuronal cytoplasm and in the case of mtDNA, the vacuoles associated with lipofuscin. Morphometric analysis showed that mitochondria are significantly reduced in Alzheimer's disease. The relationship shown here between the site and extent of mitochondrial abnormalities and oxidative damage suggests an intimate and early association between these features in Alzheimer's disease.
Article
Full-text available
2-Ethyl-1-hydroxy-2,5,5-trimethyl-3-oxazolidine (OXANOH), the one-electron reduction product of the stable nitroxide radical, 2-ethyl-2,5,5-trimethyl-3-oxazolidinoxyl (OXANO), is reportedly oxidized by superoxide, and its oxidation has been proposed as a method for assaying superoxide. We find that superoxide can both reduce OXANO and oxidize OXANOH. The respective rate constants, k1 and k2, were determined using two superoxide-generating systems (xanthine oxidase/xanthine as well as ionizing radiation). OXANOH oxidation and OXANO reduction are both inhibitable by superoxide dismutase, pH-dependent (4.5-9.3), and result in a steady state distribution of [OXANO] and [OXANOH], independent of their initial concentrations, i.e. the OXANO/OXANOH couple exhibits a metal-independent superoxide dismutase-like function. Thus it provides a prototype for future development of improved low molecular weight superoxide dismutase mimics which will also function in cellular hydrophobic (aprotic) compartments such as membranes.
Article
Accumulating data suggest a central role for mitochondria and oxidative stress in neurodegenerative apoptosis. We previously demonstrated that amyloid-beta peptide 25-35 (Abeta 25-35) toxicity in cultured cells is mediated by its effects on functioning mitochondria. In this study, we further explored the hypothesis that AD 25-35 might induce apoptotic cell death by altering mitochondrial physiology. Mitochondria in Ntera2 (NT2 rho+) human teratocarcinoma cells exposed to either staurosporine (STS) or Abeta 25-35 were found to release cytochrome c. with subsequent activation of caspases 9 and 3. However, NT2 cells depleted of mitochondrial DNA (rho0 cells), which maintain a normal mitochondrial membrane potential (Deltapsi(m)) despite the absence of a functional electron transport chain (ETC), demonstrated cytochrome c release and caspase activation only with STS. We further observed increased reactive oxygen species (ROS) production and decreased reduced glutathione (GSH) levels in rho(+) and rho(0) cells treated with STS, but only in rho(+) cells treated with Abeta 25-35. We conclude that under in vitro conditions, Abeta can induce oxidative stress and apoptosis only when a functional mitochondrial ETC is present.
Article
Beta amyloid (Aβ) peptides accumulate in Alzheimer's disease and are neurotoxic possibly through the production of oxygen free radicals. Using brain microdialysis we characterized the ability of Aβ to increase oxygen radical production in vivo. The 1–40 Aβ fragment increased 2,3-dehydroxybenzoic acid efflux more than the 1–28 fragment, in a manner dependent on nitric oxide synthase and NMDA receptor channels. We then examined the effects of Aβ peptides on mitochondrial function in vitro. Induction of the mitochondrial permeability transition in isolated rat liver mitochondria by Aβ(25–35) and Aβ(35–25) exhibited dose dependency and required calcium and phosphate. Cyclosporin A prevented the transition as did ruthenium red, chlorpromazine, or N-ethylmaleimide. ADP and magnesium delayed the onset of mitochondrial permeability transition. Electron microscopy confirmed the presence of Aβ aggregates and swollen mitochondria and preservation of mitochondrial structure by inhibitors of mitochondrial permeability transition. Cytochrome c oxidase (COX) activity was selectively inhibited by Aβ(25–35) but not by Aβ(35–25). Neurotoxic Aβ peptide can increase oxidative stress in vivo through mechanisms involving NMDA receptors and nitric oxide sythase. Increased intracellular Aβ levels can further exacerbate the genetically driven complex IV defect in sporadic Alzheimer's disease and may precipitate mitochondrial permeability transition opening. In combination, our results provide potential mechanisms to support the feed-forward hypothesis of Aβ neurotoxicity.
Article
Defects in mitochondrial oxidative metabolism, in particular decreased activity of cytochrome c oxidase, have been demonstrated in Alzheimer’s disease, and after the expression of the amyloid precursor protein (APP) in cultured cells, suggesting that mitochondria might be involved in β-amyloid toxicity. Recent evidence suggests that the proteolysis of APP to generate β-amyloid is at least in part intracellular, preceding the deposition of extracellular fibrils. We have therefore investigated the effect of incubation of isolated rat brain mitochondria with the β-amyloid fragment 25–35 (100 μM) on the activities of the mitochondrial respiratory chain complexes I, II–III, IV (cytochrome c oxidase) and citrate synthase. The peptide caused a rapid, dose-dependent decrease in the activity of complex IV, while it had no effect on the activities on any of the other enzymes tested. The reverse sequence peptide (35–25) had no effect on any of the activities measured. We conclude that inhibition of mitochondrial complex IV might be a contributing factor to the pathogenesis of Alzheimer’s disease.
Article
1. In the accompanying paper (Duchen & Biscoe, 1992) we have described graded changes in autofluorescence derived from mitochondrial NAD(P)H in type I cells of the carotid body in response to changes of PO2 over a physiologically significant range. These observations suggest that mitochondrial function in these cells is unusually sensitive to oxygen and could play a role in oxygen sensing. We have now explored further the relationships between hypoxia, mitochondrial membrane potential (delta psi m) and [Ca2+]i. 2. The fluorescence of Rhodamine 123 (Rh 123) accumulated within mitochondria is quenched by delta psi m. Mitochondrial depolarization thus increases the fluorescence signal. Blockade of electron transport (CN-, anoxia, rotenone) and uncoupling agents (e.g. carbonyl cyanide p-trifluoromethoxy-phenylhydrazone; FCCP) increased fluorescence by up to 80-120%, while fluorescence was reduced by blockade of the F0 proton channel of the mitochondrial ATP synthase complex (oligomycin). 3. delta psi m depolarized rapidly with anoxia, and was usually completely dissipated within 1-2 min. The depolarization of delta psi m with anoxia (or CN-) and repolarization on reoxygenation both followed a time course well characterized as the sum of two exponential processes. Oligomycin (0.2-2 micrograms/ml) hyperpolarized delta psi m and abolished the slower components of both the depolarization with anoxia and of the subsequent repolarization. These data (i) illustrate the role of the F1-F0 ATP synthetase in slowing the rate of dissipation of delta psi m on cessation of electron transport, (ii) confirm blockade of the ATP synthetase by oligomycin at these concentrations, and (iii) indicate significant accumulation of intramitochondrial ADP during 1-2 min of anoxia. 4. Depolarization of delta psi m was graded with graded changes in PO2 below about 60 mmHg. The stimulus-response curves thus constructed strongly resemble those for [Ca2+]i and NAD(P)H with PO2. The change in delta psi m closely followed changes in PO2 with time. 5. The rate of rise of [Ca2+]i in response to anoxia is strongly temperature sensitive. The rate of depolarization of delta psi m with anoxia similarly increased at least two- to fivefold on warming from 22 to 36 degrees C. The change with FCCP was not significantly altered by temperature. 6. These data show that the mitochondrial membrane potential changes over a physiological range of PO2 values in type I cells. This contrasts with the behaviour in dissociated chromaffin cells and sensory neurons, in which no change was measurable until the PO2 fell close to zero.(ABSTRACT TRUNCATED AT 400 WORDS)
Article
Several lines of evidence suggest that abnormalities in oxidative metabolism and specifically in mitochondria may play an important role in Alzheimer's disease. Abnormalities of oxidative metabolism exist in this disorder. They have been demonstrated in brain studied in vivo, ex vivo (biopsies), at autopsy, and in non-neural tissues including cultured cells. The abnormalities include a profound deficit in the activity of the ketoglutarate dehydrogenase complex (KGDHC), which is likely to lead to impaired metabolism of glutamate and might contribute to selective neuronal cell death by excitotoxic mechanisms as well as by direct effects on energy metabolism through its role in the tricarboxylic acid cycle. Abnormalities in oxidative metabolism may be related to the pathophysiology of Alzheimer's disease by plausible mechanisms for which there is evidence at least in model systems. Hypoxia is known to induce neuropsychological impairments analogous to those which occur in dementing syndromes. Brain, and specifically neurons, are likely to be particularly vulnerable to impairments of oxidative metabolism because of their demonstrated tight dependence on continuous oxidation of glucose to maintain their structure and function. Neuroanatomic studies as well as recent data from CT, PET, and SPECT scanning agree with formulations that suggest the brain areas of greatest vulnerability in Alzheimer's disease include those particularly sensitive to oxidative impairments. Although the mechanisms of accumulation of the classical neuropathological hallmarks of Alzheimer's disease (paired helical filaments, amyloid plaques) are not known, experimental data suggest that metabolic stresses may contribute to the accumulation of these materials. These data include the accumulation of immunoreactivity of anti-paired helical filament antibodies in cells exposed to a mitochondrial poison, the uncoupler CCCP. Impairments of oxidative metabolism are known to impair the metabolism of neurotransmitters involved in Alzheimer's disease; the synthesis of acetylcholine, which is characteristically involved, is exquisitely sensitive to oxidative abnormalities. Experimental evidence suggests that abnormalities of cellular calcium homeostasis, which have been demonstrated in Alzheimer cells, may mediate key deleterious effects of the abnormalities in oxidative metabolism in this disorder. Experimental studies in animals indicate that age potentiates the effects of inherent oxidative abnormalities on the brain, as does cerebrovascular disease. These observations might help to explain the increasing clinical expression of the gene for Alzheimer's disease with age. They are also in accord with difficulties in separating the role of vascular from that of inherent degenerative factors in dementia of later onset. Treatment with L-carnitine, a manipulation designed to mitigate consequences of a mitochondrial abnormality, normalized several non-mitochondrial abnormalities in cultured Alzheimer cells.(ABSTRACT TRUNCATED AT 400 WORDS)