ArticlePDF Available

Copper-binding-site-null SOD1 causes ALS in transgenic mice: Aggregates of non-native SOD1 delineate a common feature

Authors:

Abstract and Figures

Cu/Zn superoxide dismutase (SOD1), a crucial cellular antioxidant, can in certain settings mediate toxic chemistry through its Cu cofactor. Whether this latter property explains why mutations in SOD1 cause FALS has been debated. Here, we demonstrate motor neuron disease in transgenic mice expressing a SOD1 variant that mutates the four histidine residues that coordinately bind Cu. In-depth analyses of this new mouse model, previously characterized models and FALS human tissues revealed that the accumulation of detergent-insoluble forms of SOD1 is a common feature of the disease. These insoluble species include full-length SOD1 proteins, peptide fragments, stable oligomers and ubiquitinated entities. Moreover, chaperones Hsp25 and alphaB-crystallin specifically co-fractionated with insoluble SOD1. In cultured cells, all 11 of the FALS variants tested produced insoluble forms of mutant SOD1. Importantly, expression of recombinant peptide fragments of wild-type SOD1 in cultured cells also produced insoluble species, suggesting that SOD1 possesses elements with an intrinsic propensity to aggregate. Thus, modifications to the protein, such as FALS mutations, fragmentation and possibly covalent modification, may simply act to augment a natural, but potentially toxic, propensity to aggregate.
Content may be subject to copyright.
Copper-binding-site-null SOD1 causes ALS in
transgenic mice: aggregates of non-native
SOD1 delineate a common feature
Jiou Wang
1
, Hilda Slunt
1
, Victoria Gonzales
1
, David Fromholt
1
, Michael Coonfield
1
,
Neal G. Copeland
3
, Nancy A. Jenkins
3
and David R. Borchelt
1,2,
*
1
Department of Pathology and
2
Department of Neuroscience, The Johns Hopkins University School of Medicine,
Baltimore, MD 21205, USA and
3
Mouse Cancer Genetics Program, NCI-Frederick Cancer Research
and Development Center, Frederick, MD 21702, USA
Received June 3, 2003; Revised August 19, 2003; Accepted August 29, 2003
Cu/Zn superoxide dismutase (SOD1), a crucial cellular antioxidant, can in certain settings mediate toxic
chemistry through its Cu cofactor. Whether this latter property explains why mutations in SOD1 cause FALS
has been debated. Here, we demonstrate motor neuron disease in transgenic mice expressing a SOD1 variant
that mutates the four histidine residues that coordinately bind Cu. In-depth analyses of this new mouse
model, previously characterized models and FALS human tissues revealed that the accumulation of
detergent-insoluble forms of SOD1 is a common feature of the disease. These insoluble species include full-
length SOD1 proteins, peptide fragments, stable oligomers and ubiquitinated entities. Moreover, chaperones
Hsp25 and aB-crystallin specifically co-fractionated with insoluble SOD1. In cultured cells, all 11 of the FALS
variants tested produced insoluble forms of mutant SOD1. Importantly, expression of recombinant peptide
fragments of wild-type SOD1 in cultured cells also produced insoluble species, suggesting that SOD1
possesses elements with an intrinsic propensity to aggregate. Thus, modifications to the protein, such as
FALS mutations, fragmentation and possibly covalent modification, may simply act to augment a natural, but
potentially toxic, propensity to aggregate.
INTRODUCTION
Mutations in Cu/Zn SOD1 have been linked to familial
amyotrophic lateral sclerosis (FALS) (1), a degenerative disease
characterized by loss of lower and upper motor neurons in the
motor cortex, brain stem and spinal cord (2). Over 70 different
missense substitutions at more than 50 residues of the 153 amino
acid protein have been described in individuals and kindreds
affected by SOD1-linked FALS (www.alsod.org). A small
number of mutations lead to early translation termination in
the last of the five exons. Although intensely studied, the nature
of the common abnormality by which all these mutations in
SOD1 cause FALS has not been definitively identified. Most
investigators accept the notion that the SOD1 mutations cause
the disease through a gain-of-property mechanism (3–7), but just
what this toxic property is has not be resolved.
One hypothesis has focused on the metal binding properties
of Cu/Zn SOD1. The ‘Cu hypothesis’ holds that mutations
diminish the normal shielding of the Cu cofactor in the
enzyme, enhancing entry of less favored substrates including
H
2
O
2
and ONOO. Reactions between the Cu cofactor and
H
2
O
2
can produce potentially toxic radical species (8–11)
whereas reaction with ONOO can catalyze the covalent
nitration of tyrosine residues (12–14). However, whether these
reactions are central to the toxicity of mutant SOD1 has been
debated (15,16).
Another hypothesis suggests that aggregation of SOD1 may
selectively injure motor neurons. Abnormal inclusions have
been observed in human ALS spinal cords from both familial
and sporadic cases (17–19), in spinal cords from transgenic mice
expressing FALS variants (7,19), and in cytoplasm of cultured
COS cells, or motor neurons, expressing FALS-SOD1 variants
(20,21). Detergent-insoluble forms of mutant SOD1 have been
detected in cultured cells and mice expressing FALS-SOD1
variants (22–24). The principal criticisms of this hypothesis are
that relatively few mutants have been examined in model
*To whom correspondence should be addressed at: Department of Pathology, The Johns Hopkins University School of Medicine, 720 Rutland Ave.,
Room 558, Baltimore, MD 21205, USA. Tel: þ1 4105025174; Fax: þ1 4109559777; Email: drbor@jhmi.edu
Human Molecular Genetics, 2003, Vol. 12, No. 21 2753–2764
DOI: 10.1093/hmg/ddg312
Human Molecular Genetics, Vol. 12, No. 21 #Oxford University Press 2003; all rights reserved
by guest on December 13, 2016http://hmg.oxfordjournals.org/Downloaded from
systems, and in some of the transgenic mouse models, SOD1
positive inclusions in affected spinal cords are relatively rare
(19,24).
To test the Cu hypothesis, we have focused on disease
associated mutations that occur in the Cu-binding site of
SOD1. There are four histidines that coordinately bind Cu
[histidines 46, 48, 63, and 120 (25)], and disease causing
mutations have been described at histidines 46 and 48 (H46R
and H48Q). Experimental mutants of SOD1 that combine these
two mutations can induce motor neuron disease in transgenic
mice (26). To eliminate the Cu binding site, we combined the
two disease-causing mutations at histidine 46 and 48 with two
experimental mutations at histidines 63 and 120 (H63G and
H120G), yielding a protein (SOD1-Quad) that was surprisingly
stable but completely inactive. When we expressed this variant
in mice, via a mutated version of the human genomic fragment
of the sod1 gene (27), we observed motor neuron disease
similar in clinical and pathological appearance to other mouse
models of SOD1-linked FALS. We interpret this outcome as
evidence that motor-neuron specific toxicity does not require a
normally configured active site to produce a protein toxic to
motor neurons. In analyzing diseased tissues from mice
expressing SOD1-Quad, we found evidence of aggregated
forms of mutant protein by both size exclusion filter assay, and
differential detergent extraction and centrifugation. We utilized
the latter approach to compare aggregating SOD1 species in
this new model to four other previously characterized models,
demonstrating that SOD1-Quad acquires characteristics similar
to natural FALS-SOD1 mutants. We also developed cell culture
assays to examine 11 different FALS variants, finding that
misfolding to produce detergent insolubility is a common
feature. We suggest that misfolded and aggregating species of
SOD1 are strong candidates for the toxic species in the disease.
RESULTS
Motor neuron disease and SOD1 aggregates in
H46R/H48Q/H63G/H120G mice
Previous characterizations of SOD1 reactivity with H
2
O
2
and
ONOO have dealt with enzymes where Cu is loaded into its
normal active site (8–11,13,14,28). There are four histidine
residues that coordinately bind Cu, residues 46, 48, 63 and 120
(25). Mutations at histidines 46 and 48 have each been reported
in FALS patients (www.also.org), and mice expressing an
SOD1 with a double mutation at histidines 46 and 48 (H46R/
H48Q) develop motor neuron disease (26). Enzymes harboring
H46R and H48Q mutations alone, or in combination, lack
detectable superoxide scavenging activity (26). Hence, partial
destruction of the Cu-binding site in SOD1 renders the enzyme
inactive, while retaining the capacity to cause motor neuron
disease.
To completely destroy the Cu-binding site, we combined the
H46R/H48Q double FALS mutation with additional mutations
at histidines 63 and 120 (H46R/H48Q/H63G/H120G; the Quad
mutant). Two lines (87 and 125) were identified that developed
typical motor neuron disease, which progressed to hindlimb
paralysis by 8–12 months of age (Fig. 1A and B). The levels of
SOD1-Quad expression that were required to induce disease at
these ages were similar to that of other FALS-SOD1 variants
(Fig. 1D). In a native gel assay of SOD1 activity in spinal cords
of non-transgenic mice (NTg), mice expressing high levels of
wild-type human SOD1 (WT), and mice expressing SOD1-
Quad, no bands corresponding to human SOD1 were noted in
tissue extracts from the SOD1-Quad mice (Fig. 1C). Moreover,
the band of activity corresponding to endogenous mouse SOD1
homodimers was neither diminished in its intensity nor altered
in its migration in the SOD1-Quad mice, suggesting the SOD1-
Quad protein may not dimmerize with the mouse enzyme.
Pathological examination of symptomatic SOD1-Quad mice
demonstrated marked motor neuron loss (by silver impregna-
tion) and dramatically increased glial fibrillary acidic protein
(GFAP) immunostaining (Supplementary Material Fig. S1).
Spinal cords and brain stems from affected mice also showed
conspicuous inclusion body pathology, which was revealed by
immunostaining with ubiquitin antibodies or thioflavin-S
staining (Fig. 2 and Supplementary Material Fig. S2). These
pathologies were absent in cerebral cortex, cerebellum, or other
parts of the forebrain. As controls, age-matched wild-type
SOD1 transgenics and non-transgenic littermates remained
healthy and free of abnormal inclusions. Although immunos-
taining with SOD1 antibodies failed to detect obvious
inclusions (data not shown), by size exclusion filter assay
aggregated species of SOD1-Quad were abundant in homo-
genates of spinal cord and brain stem tissues from paralyzed
SOD1-Quad mice (Fig. 2E). These structures were detected in
cerebellum and cerebral cortex, but at much lower levels.
Figure 1. Transgenic mice expressing H46R/H48Q/H63G/H120G (Quad)
mutant SOD1 develop motor neuron disease. (A) Hindlimb paralysis in
SOD1-Quad mice (line 125). (B) SOD1-Quad mice (lines 125 and 87) deve-
loped complete hind limb paralysis at the age of 9–12 months, and 8–11
months, respectively. (C) Mice expressing SOD1-Quad do not produce excess
superoxide scavenging activity. Only activity associated with endogenous
SOD1 equal to that in nontransgenic (NTg) control mice is detected; 100 mg
of protein from spinal cord homogenates of 2-month-old mice were tested for
SOD1 activity in the native gel assays. (D) Disease progression is correlated
with protein expression levels in different lines of SOD1-Quad mice; the higher
the expression level (line 87>125>121), the earlier the onset of paralysis. A
similar dose dependency for disease onset in mice expressing other SOD1
mutants including, the H46R/H48Q mutant (e.g. line 139>58), has been noted
previously (26). Protein levels were determined by western blotting of spinal
cord homogenates from 2-month-old mice.
2754 Human Molecular Genetics, 2003, Vol. 12, No. 21
by guest on December 13, 2016http://hmg.oxfordjournals.org/Downloaded from
Overexpressing WT SOD1 in mice did not induce formation of
these structures [Fig. 2E and Wang et al. (24)].
SOD1 monomers, fragments, and SDS-resistant
oligomers are components of detergent-insoluble
complexes in FALS mice and human patients
The appearance of motor neuron disease and aggregated
species of mutant SOD1 in the SOD1-Quad mice suggested to
us that aggregation of the mutant protein may be the key to
disease. To characterize further the SOD1 species accumulating
in the symptomatic mice, we used differential detergent
extraction and high speed centrifugation to enrich for forms
of the mutant protein in non-native structure (see Materials and
Methods). Proteins insoluble in non-ionic detergent (P2) and
SDS (P3) were then solubilized by boiling in the SDS sample
buffer before being analyzed by western blotting with two
polyclonal antibodies: hSOD1 antibody, a peptide antiserum
binding to amino acids 24–36 (not conserved between mouse
and human SOD1); and m/hSOD1 antibody, a peptide
antiserum recognizing amino acids 124–136 (completely
conserved between mouse and human protein; Fig. 3A).
Spinal cord extracts from five different strains of FALS mice
were examined; G37R (line 29), G85R (line 164), G93A (line
1—high expression), H46R/H48Q (line 139), and SOD1-Quad
(line 125). As compared with either non-transgenic mice, or
mice expressing high levels of wild-type human SOD1 (line
76), spinal cords from the paralyzed mutant mice contained
large amounts of detergent-insoluble SOD1 species (Fig. 3A).
However, the majority of both wild-type and mutant SOD1,
from all animals, was soluble in non-ionic detergent [Fig. 3B;
the soluble (S1) fraction contained 600 mg total protein of
which only 1 mg was analyzed by SDS-PAGE; the P2 fraction
contained 50 mg of total protein of which 3 mg was analyzed
by immunoblot—see legend to Fig. 3].
Although the predominant SOD1 species solubilized after
boiling in SDS consisted of full-length SOD1 monomers (open
arrow), there were distinct immunoreactive peptide fragments
(solid arrow) and high molecular weight (solid arrowhead)
species in these insoluble fractions (Fig. 3A). Notably, the G85R
variant always migrates, in SDS–PAGE, slightly faster than
wild-type SOD1 or other FALS variants (4). This characteristic
was also evident in 28 and 38 kDa species detected in the G85R
mice. Interestingly, detergent-insoluble, SOD1-immunoreactive
species in the spinal cords of the paralyzed mice displayed
patterns that were not entirely identical when examined by
immunoblot with the two different antibodies. The hSOD1
antibody recognized, with variable abundance, a set of 7 kDa
fragments that were particularly abundant in the H46R/H48Q
mice. Additionally, there were several species that appeared to be
slightly smaller than the full-length protein; these fragments
were not recognized by the m/hSOD1 antibody, suggesting that
these fragments lack C-terminal residues. By contrast, the m/
hSOD1 antibody specifically recognized several fragments in the
range of 9–15 kDa; these fragments were particularly abundant
in spinal cords from the G85R and H46R/H48Q mice. Both
antibodies recognized a pair of prominent SOD1 species
(marked by asterisks), which have relative molecular masses
of 28 and 38 kDa. Note that the hSOD1 antibody detected, at
lower abundance, additional species just below these 28 and
38 kDa species and since these were not recognized by the
m/hSOD1 antibody we infer that these molecules are truncated
at the C-terminus.
Next, we examined spinal cord tissues from an FALS patient
known to harbor an SOD1-A4V mutation, together with two non-
SOD1 FALS patients, 12 sporadic ALS patients, and three
controls. Only tissue homogenates from the patient harboring
the A4V mutation showed enhanced levels of detergent-insoluble
SOD1 monomers, fragments, and high molecular weight species
(Fig. 3C). By contrast, little or no insolubleSOD1 was detected in
spinal cord tissues from patients with sporadic and non-SOD1
familial ALS (Supplementary Material Fig. S3).
To determine whether covalent ubiquitination of any of these
SOD1 species might be contributing to the complexity of the
SOD1-immunoreactive profiles seen in immunoblots probed
Figure 2. Fibrillar inclusions and SOD1-containing aggregates in symptomatic
SOD1-Quad mice. (Aand B) Immunostains of spinal cords with antibodies to
ubiquitin revealed inclusions (open arrow) and degenerating cell bodies/neuro-
pil (solid arrow) in SOD1-Quad transgenic mice (A) as compared with NTg lit-
termate controls (B). (Cand D) Sagittal sections through the juncture of brain
stem and spinal cord stained with thioflavin-S revealed long fibrils in sympto-
matic SOD1-Quad mice (C) as compared with NTg controls (D). Scale bars:
100 mm. (E) Size exclusion filter trapping and immunodetection (described in
Materials and Methods and 24) demonstrated high molecular weight forms
of SOD1-Quad in spinal cord and brain stem of symptomatic mice. The same
amount of protein from transgenic mice overexpressing wild-type SOD1 or
NTg mice was filtered and immunoblotted as controls.
Human Molecular Genetics, 2003, Vol. 12, No. 21 2755
by guest on December 13, 2016http://hmg.oxfordjournals.org/Downloaded from
with SOD1 antibodies, we probed immunoblots of the SDS
insoluble pellet (P3) with monoclonal antibodies to ubiquitin.
In all five of the SOD1 FALS models we examined, we found
that the insoluble pellets contain ubiquitinated entities includ-
ing two species with apparent molecular mass of mono and di-
ubiquitinated mutant SOD1 (Fig. 4A). To confirm whether
ubiquitinated SOD1 was indeed present in these fractions, we
dissociated the P2 pellet fraction by boiling in 2% SDS, and
then diluted the solubilized protein to 0.2% SDS and immuno-
precipitated SOD1 protein with the m/hSOD1 antibody.
Immunoprecipitated proteins were then subjected to immuno-
blot with ubiquitin antibody (Fig. 4B), noting the detection of
two prominent proteins with relative molecular masses of 28
and 38 kDa (Fig. 4B), which are roughly the predicted sizes for
mono- and di-ubiquitinated SOD1 proteins.
FALS mutations promote non-native folding to
form detergent-insoluble complexes
As noted above, and from our previous work (24), high-
molecular weight, detergent-insoluble, mutant SOD1 is most
abundant in tissues that develop visible pathology: brain stem
and spinal cord. In this setting, it is impossible to determine
whether the formation of these altered species of SOD1 is an
inherent property of the protein or consequence of the disease
process. To address this issue, we studied the detergent
solubility of different SOD1 mutants expressed at high levels
in cultured cells. Wild-type SOD1 and 11 FALS mutant cDNAs
[A4V, G37R, G41D, H46R, H48Q, H46R/H48Q, G85R, G93C,
I113T, Quad, and FS126 (frame shift 126–stop 131)] were
inserted into the pEF-BOS expression vector (4) before transient
transfection of human HEK-293 cells. Northern analysis of the
transfected cultures showed similar SOD1 mRNA levels (data
not shown) for each variant. As expected, the accumulated
steady-state levels of the different mutants varied according to
the intrinsic half-life of the individual proteins (Fig. 5A) (4,29).
Twenty-four hours after transfection, cells were harvested and
differentially extracted as described in Materials and Methods.
The non-ionic detergent-insoluble pellets (P2) and supernatant
(S1) were separated, and SOD1 levels were quantified by
western blotting with the hSOD1 antibody and
125
I-labeled
secondary antibodies, followed by Phosphorimager detection
and quantification (Fig. 5A and B). All of the cells transfected
with FALS variants contained forms of mutant SOD1 that were
insoluble in non-ionic detergents (Fig. 5A) and SDS (data not
shown). Notably, however, wild-type human SOD1 was barely
detectable in the insoluble fractions, even though the wild-type
protein accumulated to the highest steady-state level in the
soluble fraction. Comparing the relative amounts of SOD1 in
the pellet and supernatant fractions, we found that each mutant
has a statistically higher aggregation potential than wild-type
SOD1 (Fig. 5B). Interestingly, the A4V, G41D, G85R and
FS126 variants, all short-lived (4,29), produced more detergent
insoluble species than any of the other mutants. The FS126
variant, the only truncated protein, appeared to be the most
aggregation prone. These results strongly suggest that FALS
variants are prone to adopting non-native conformations that
lead to detergent-insoluble structures, and that this property is
intrinsic to the mutants and does not require extrinsic events
such as may occur in degenerating and diseased tissues.
Intrigued by the appearance of detergent insoluble fragments
of SOD1 in the mouse and human tissues (Fig. 3), we used
the cell culture transfection system to ask whether these
molecules may also be inherently prone to adopt non-native
Figure 3. Accumulation of detergent-insoluble SOD1 monomers, fragments, and high molecular weight species in spinal cords of affected mice and human
patients. (A) Mutant SOD1 from the spinal cords of affected mice fractionated in the detergent-insoluble pellet fraction (P3). Spinal cords were differentially
extracted and sedimented in non-ionic detergent and SDS as described in Materials and Methods. P3 is the final pellet after sequential extraction in non-ionic
detergent and SDS. Fractions were separated by tricine-peptide SDS–PAGE. Western blotting was done with the hSOD1 antibody, which binds to residues 24–
36 specific to human SOD1, or the m/hSOD1 antibody, which recognizes residues 124–136 common to both the human and mouse protein. Bound primary anti-
bodies were detected by protein A-horseradish-peroxidase and chemiluminescence. In mice expressing wild-type human SOD1 and in normal NTg controls, all
SOD1 was fully soluble in detergent. By contrast, insoluble species of SOD1 were abundant in symptomatic mutant mice. SOD1 monomers (open arrow), frag-
ments (solid arrow), and high molecular weight species (solid arrow head) are indicated. Apparent dimers (asterisk) and trimers (double asterisk) are noted. (B) The
majority of SOD1 in both mutant and control mice is fully soluble in detergent (S1). For every 600 mg of total protein in the supernatant S1 fraction, roughly 200 mg
of protein fractionated in P2 and 50 mg of protein fractionated in P3. One microgram of S1 protein was loaded in (B), compared with 3 mg of P3 protein loaded in
(A). With the same blotting condition and antibody titration, the exposure time for (A) is three times that for (B), which was <5s. The immunoblot in (B) was
probed with the m/hSOD1 antiserum. (C) The spinal cord of a patient with the A4V mutation contains SOD1 monomers, fragments and oligomers that are inso-
luble in non-ionic detergent (P2). A representative age-matched non-ALS control case was examined in parallel. Immunoblots were probed with hSOD1 orm/
hSOD1 antiserum as described in (A).
2756 Human Molecular Genetics, 2003, Vol. 12, No. 21
by guest on December 13, 2016http://hmg.oxfordjournals.org/Downloaded from
conformations. We examined the aggregation potential of
several experimental variants of wild-type SOD1 truncated at
the N- and C-terminus to approximate the sizes of fragments
observed in mouse and human tissues. Although the steady-
state levels of each recombinant fragment were much lower
than wild-type protein, each fragment produced detergent
insoluble species (Fig. 6). These results support the notion that
structural elements in SOD1 have an unusually high potential
to aggregate, and that in the wild-type protein these elements
must be effectively ordered in the native structure to avoid
aggregation.
Up-regulation of Hsp25 and aB-crystallin in
symptomatic SOD1 mutant mice
Because molecular chaperones are crucial participants in
protein folding, degradation and aggregation, we used a panel
of antibodies against heat shock proteins to determine whether
any of the known chaperones were up-regulated in the spinal
cords of FALS mice. Using non-ionic detergent extraction
protocols described above, we also asked whether any of these
chaperones were also associated with the detergent-insoluble
SOD1 complexes (Fig. 7). Although we found no appreciable
differences between control and symptomatic mice in analyzing
the levels and solubility of Hsp40, Hsp60, Hsp70 and Hsp90,
we observed a dramatic up-regulation in the levels of Hsp25. In
affected mice from five different SOD1 mouse models, the
level of Hsp25 was elevated in both the soluble and insoluble
fractions. In the soluble fraction, Hsp 25 appeared as a doublet
band, whereas in the pellet fraction only a single species of
HSP25 was detected (Fig. 7A). Another small chaperone
protein, aB-crystallin, was found to be significantly up-
regulated in the mutant mice, with the majority of the induced
component being localized to the detergent-insoluble fraction.
Because these two proteins are known to form highly organized
structures (30), we do not know whether they are integral
components of SOD1 aggregates. However, both chaperones
co-fractionated with SOD1 aggregates in the SDS insoluble
fraction P3 (data not shown).
Immunocytochemical examination of paralyzed mice loca-
lized Hsp25 to both neuronal and glial cells and aB-crystallin
to oligodendrocytes. In neither case, were neuropil inclusions
(recognized by ubiquitin antibodies) specifically marked (data
not shown). Hence, we are uncertain as to whether these
chaperones are directly binding protein contained in large
visible inclusions, or protein organized in smaller structures not
visible by light microscopy.
DISCUSSION
Our demonstration of pathologic and behavioral symptoms of
motor neuron disease in strains of mice expressing SOD1-Quad
proves that disease can be induced by mutant forms of SOD1
lacking the histidine residues that coordinately bind Cu in the
active site. The pathologic features of disease in the SOD1-
Quad mice were identical to that of other FALS-SOD1 mice;
these features include ubiquitin-immunoreactive inclusions,
thioflavin-S positive inclusions, robust astrogliosis, and loss of
motor neurons. As we have previously demonstrated in other
FALS-SOD1 mouse models (24), diseased tissues contained
high molecular weight forms of SOD1-Quad that could be
trapped by size exclusion filter trap. Similarly, behavioral
phenotypes were identical; initial symptom of hindlimb
weakness progressing to hindlimb paralysis before involvement
of forelimbs. Thus by all criteria we have used to characterize
FALS-SOD1 mice, the SOD1-Quad mice appear to have
authentic motor neuron disease.
Have we definitively disproved the Cu hypothesis?
The Cu hypothesis is an umbrella term that covers hypotheses
that involve the ability of SOD1 to covalently nitrate tyrosine
residues in protein (12) or to react with H
2
O
2
to produce radical
species (8). These toxic reactions have been described for
Figure 4. High molecular weight ubiquitinated entities, including ubiquitinated SOD1, are enriched in the detergent-insoluble fractions extracted from spinal cords
of mutant SOD1 mice. (A) Spinal cord homogenates from symptomatic mutant SOD1 mice and normal control mice were detergent extracted and western blotted
with anti-ubiquitin antibodies. High molecular weight, ubiquitin-immunoreactive entities fractionated in the SDS-insoluble fractions (P3) from the spinal cords of
symptomatic mutant mice as compared with mice expressing wild-type SOD1 and NTg controls (left). The majority of monomeric ubiquitin was completely
solubilized by non-ionic detergent (S1; right). Prominent 28 and 38 kDa bands, with a slight down-shift in size from the G85R mice, were found in samples from
all mutant mice. (B) The non-ionic detergent pellets (P2) of spinal cord homogenates were adjusted to 2% SDS and boiled for 10min. Then the diluted samples
were immunoprecipitated with the m/hSOD1 antibody and western blotted with the anti-ubiquitin monoclonal antibody. The two distinct protein species at 28 and
38 kDa seen in (A) clearly contain SOD1. The relative size of the two proteins is consistent with mono- and di-ubiquitinated SOD1.
Human Molecular Genetics, 2003, Vol. 12, No. 21 2757
by guest on December 13, 2016http://hmg.oxfordjournals.org/Downloaded from
wild-type SOD1, with evidence that FALS mutations augment
these activities. The nitration chemistries attributed to both
wild-type SOD1 (28) and FALS variants (13,14) by other
investigators were described in enzymes with intact active sites.
There is no evidence that inactive enzymes can catalyze these
reactions.
Similarly, the peroxidative activity of SOD1 was originally
described for wild-type SOD1 with Cu bound in the active site
(9,31). As was the case for the nitration reaction, mutations distal
to the active site were found to augment the peroxidative reaction
(8,10). Moreover, enzyme harboring the H48Q mutation within
the active site was found to retain the peroxidative chemistry.
Although altering one of the histidine residues responsible for
binding Cu is not sufficient to eliminate this chemistry, there is
no evidence that enzyme lacking all four of the histidine residues
necessary for the coordinate binding of Cu could perform the
peroxidative chemistry.
When one sums all of the studies that have tested the Cu-
hypothesis by experimentally lowering the level of Cu binding
in the active site [the present study, the work by Subramaniam
et al. (15) to delete CCS and diminish the loading of Cu into
mutant SOD1, and our previous study of mice expressing a
double histidine mutant (H46R/H48Q) (26)], then the weight of
evidence does not favor the hypothesis as it pertains to Cu
correctly loaded in the Cu-binding pocket. It remains to be
determined whether Cu bound to other parts of the protein (the
Zn site or elsewhere) could produce SOD1 enzymes that are
capable of generating the same type, or some other type, of
toxic chemical reaction (16).
Are aggregated forms of mutant SOD1 the
toxic species?
We have examined, in great detail, five different mouse models
of SOD1-linked FALS and found that disease is always
associated with the accumulation of high molecular weight
forms of mutant SOD1 that are retained by size exclusion
filtration [Fig. 2 and (24)]. We have also found that the tissues
that are most devastated in mice (brain stem and spinal cord)
contain non-native, detergent insoluble, species of mutant
protein. We have never observed full-length wild-type protein
to possess these properties. Further, in a cell culture assay, we
have found that of 11 different FALS mutants tested, all were
found to be inherently prone to adopt non-native, detergent-
insoluble, structures. In many other neurodegenerative disease
Figure 5. Adopting structures insoluble in detergent is an inherent property of
FALS SOD1 variants. (A) Images of representative immunoblots of detergent-
soluble and insoluble fractions from HEK 293 cells transiently transfected with
wild-type and mutant SOD1 constructs. The levels of SOD1 in the pellets (P2)
and supernatants (S1) were measured by immunoblots using the hSOD1 anti-
serum and
125
I-labeled secondary antibodies followed by Phosphorimager
detection. Although highly expressed, wild-type SOD1 was not detected in
the insoluble fractions. In contrast, all tested mutants accumulated, to varying
degrees, in insoluble fractions (P2). (B) The relative ratio of detergent-insoluble
to soluble SOD1 in pellets and supernatants was used to calculate the aggrega-
tion potential of tested variants. The data represent meansSEM of nine dis-
tinct transfection experiments. Significance levels were determined by paired
Student’s t-test. All mutants were statistically different from wild-type SOD1;
*significance of difference between wild-type and FALS variants P<0.02;
**significance of difference between A4V, G41D, and G85R variants and all
other variants except FS126 P<0.04; ***significance of difference between
the FS126 variant and all the other variants P<0.02.
Figure 6. Peptide fragments of SOD1 form detergent-insoluble structures. In
addition to an authentic FALS truncation, FS126, we generated the following
variants: 1, Frag(1–75), C-terminally truncated ending at the K75 residue in
the large loop domain; 2, Frag(38–153) and Frag(49–153), which are two
N-terminally truncated peptides that start at residue L38 and E49 located at
the ends of the third and fourth beta strands, respectively. The fragments,
together with controls, were transiently expressed by the pEF-BOS vector in
HEK 293 cells. After non-ionic detergent extraction, P2 and S1 fractions
were analyzed by western blotting with the hSOD1 antibody for N-terminal
fragments and controls (lanes 1–4) and the m/hSOD1 antibody for C-terminal
fragments and controls (lanes 5–8). The relative position of SOD1 wild-type
monomers (open arrow), peptide fragments (solid arrow), and oligomers
(solid arrowhead) are indicated. In addition to FS126, all expressed wild-type
peptide fragments were preferentially enriched in the non-ionic detergent pellet
fractions (*next to lane 4; and **next to lanes 7 and 8).
2758 Human Molecular Genetics, 2003, Vol. 12, No. 21
by guest on December 13, 2016http://hmg.oxfordjournals.org/Downloaded from
settings, loss of detergent solubility by a given protein or peptide
is a hall mark of aggregation. In prion disease a distinguishing
feature of the disease-specific isoforms of prion protein is loss of
detergent solubility (32). Similarly, aggregated forms of the
Alzheimer’ b-amyloid peptide become insoluble in detergents
(33,34). In Parkinson’s disease, a-synuclein acquires detergent
insolubility (34,35) and in Alzheimer’s disease and Fronto-
temporal dementia with Parkinson’s the tau protein adopts
detergent-insoluble conformations (34,36,37). A similar change
in solubility occurs in protein harboring expansions of
polyglutamine tracts such as occurs in Huntington’s disease
(38). In each of these cases, the appearance of detergent
insoluble entities is associated with the pathologic accumulation
of aggregated forms of the particular protein.
Although we have not been able to visualize large SOD1-
immunoreactive aggregates in the FALS mice with our SOD1
antibodies, we have noted that affected tissues contain large
amounts of thioflavin-S positive inclusions. Thioflavin-S is
thought to be relatively specific for b-sheet structures and will
recognize b-amyloid peptide aggregates and aggregated forms
of tau protein (39,40). The abundance of detergent insoluble
forms of SOD1 in the brain stem and spinal cord of mice that
show thioflavin-S positive inclusions leads us to believe that
aggregated forms of mutant SOD1 are one of the targets for this
dye in tissue sections from the FALS mice.
Features of the SOD1 FALS mouse models also implicate
aggregation in the toxic process. Studies of b-peptide and poly-
glutamine protein aggregation have demonstrated that aggrega-
tion is a slow progress where nucleation and seeding is the
slowest component of the process, which is highly influenced
by the concentration of the monomeric species (41–44). In
studies of several lines of transgenic mice expressing varied
levels of SOD1-G37R (6) and SOD1-H46R/H48Q (26), a strong
correlation between the age of onset and the expression level
of mutant proteins was observed. Moreover, in the G37R (6),
G93A (3), H46R/H48Q (26) and SOD1-Quad mice, hyper-
expression of the mutant protein was required to induce the
disease. Although sometimes criticized as a flaw of these models,
the requirement for high-level expression fits well with the notion
that aggregation is crucial; at lower levels of expression the
thermodynamics of the aggregation process would preclude any
appreciable nucleation before mice reach their natural life
expectancy. It is noteworthy that mice expressing the G85R
variant develop disease, accompanied by aggregated species of
mutant SOD1, when the steady state levels of protein are much
lower—equal to or less than endogenous mouse SOD1 (7,24).
Our demonstration that G85R variant has one of the highest
aggregation potentials appears to explain how this mutant
appears to possess a greater toxicity per unit of protein than
other FALS variants tested in mice.
In our view, there is much evidence that SOD1-FALS shares
features common to other neurodegenerative diseases where the
aggregation of specific proteins is a pathologic hallmark.
However, just what the toxic species is in these protein
aggregation diseases has been hotly debated (reviewed in 45).
Are the large aggregates the toxic species, or are there small
oligomeric structures that are the real culprits? We note that, in
every mouse model we have looked at, aggregated forms of
mutant SOD1 increase in abundance as symptoms worsen (24).
However, it remains possible that the large aggregates are
biomarkers for another species of toxic protein whose detection
is more elusive.
In our study of heat shock response in FALS mice, we noted
that Hsp25 and aB-crystallin were dramatically induced in the
spinal cords of affected mice but that antibodies to these
proteins did not decorate the large ubiquitin-positive inclusions
that are prominent in affected spinal cords of the FALS mice
(data not shown). It is possible that these chaperones
are associated with much smaller SDS-insoluble structures;
these small heat shock proteins are known to bind to proteins
Figure 7. Elevated levels of Hsp25 and aB-crystallin in the spinal cords of symptomatic mutant SOD1 mice. (A) Spinal cord homogenates from symptomatic
mutant SOD1 mice and normal control mice were extracted in non-ionic detergent before immunoblot analysis with antibodies specific to Hsp25. Substantially
elevated levels of Hsp25 were observed in spinal cords from all mutant SOD1 mice, but not in cords from mice expressing wild-type SOD1 or from NTg controls.
Hsp25 was up-regulated in both supernatant and detergent-insoluble pellet fractions. (B)aB-crystallin was also found to be significantly up-regulated in the symp-
tomatic mutant mice; however, most of the elevated aB-crystallin was located in the detergent-insoluble fractions.
Human Molecular Genetics, 2003, Vol. 12, No. 21 2759
by guest on December 13, 2016http://hmg.oxfordjournals.org/Downloaded from
in non-native conformations (46). The cellular pattern of
staining by antibodies to Hsp25 and aB-crystallin suggested
that non-neuronal cell types, astroglia in the case of Hsp25 and
oligodendrocytes in the case of aB-crystallin, were the major
sources of these proteins. Hence, from these data we could
argue that something other than a large aggregate has induced
a response in non-neuronal cells.
One of the attractive aspects of the Cu-hypothesis was the
notion that the gained-property (or activity) by SOD1 when
carrying FALS mutations need not involve the acquisition of
some entirely new function. Because wild-type SOD1 could, at
low levels, catalyze toxic peroxidative and nitration activities
(9,10,28), the disease associated mutations need only augment
these natural secondary activities for the protein to acquire
toxicity (8,12). Here, we demonstrate, through study of defined
peptide fragments of wild-type SOD1, that there are elements
in SOD1 that are prone to induce aggregation. Hence, once
again we could argue that the FALS mutations simply augment
a natural, potentially toxic, feature of the protein.
Ubiquitin inclusion pathology in FALS mice
The accumulation of ubiquitinated protein inclusions are
hallmark pathologies of many neurodegenerative diseases
(reviewed in 47). The appearance of ubiquitinated inclusions
has been construed as evidence that there may be diminished
proteasome function. A direct test of this possibility in a cell
culture paradigm demonstrated that cells containing large
aggregates of polyglutamine protein have lower proteasome
function (48). We and others have noted that ubiquitin
immunoreactive inclusions are frequently observed in spinal
cords of paralyzed FALS mice (6,7,19,26). Our results indicate
that most of the accumulated ubiquitin-immunoreactive
material in the affected tissues of our mice fractionates in the
detergent insoluble fraction (Fig. 4A). Some of this ubiquitin is
conjugated to SOD1 protein (Fig. 4B), and thus it is possible
that the ubiquitin-immunoreactive inclusion pathology seen in
the mouse models reflects the abundance of ubiquitinated
SOD1 in aggregates, which are likely to be poorly degraded by
the cell. Whether the appearance of the large ubiquitinated
inclusions reflects a general dysfunction of the ubiquitin/
proteasome system is unclear.
Why do FALS variants of SOD1 tend to aggregate?
One simple mechanism could involve interactions between
specific domains exposed only when non-native structures are
adopted. An example of this type of interaction has been
termed domain swapping (49,50). If FALS mutations were to
heighten spontaneous unfolding of the protein, then alternative
conformations could form oligomeric species. We have noted
that the distribution of disease-causing mutations is not entirely
random. Mutations are most preponderant at residues con-
served across many phyla and occur at a much higher
frequency in b-strand domains (Wang and Borchelt, unpub-
lished data), suggesting that these mutations may destabilize
native structures. Consistent with this notion, Rodriquez et al.
(51) reported that 14 FALS mutants have decreased thermal
stability, a property augmented when Cu and Zn were removed
(52). Thermal stability was also eroded by eliminating the
single disulfide bond in the protein (53). In an unfolded state,
interactions between hydrophobic b-strands could occur
between two individual molecules rather than within individual
molecules in a domain-swapping type of interaction (49,50). In
such a scenario, highly stable oligomeric molecules could form
(Fig. 8).
It is possible that ubiquitination or fragmentation of SOD1
increases its potential to form aggregating structures.
Ubiquitinated forms of mutant SOD1 were detected in the
detergent-insoluble fraction of affected spinal cords from each
of the FALS mouse models tested. However, in our cell culture
assay, ubiquitinated forms of SOD1 were less prevalent in the
detergent insoluble fraction. Thus, how ubiquitination may
affect solubility is not clear. By contrast, it appears that
fragmentation can certainly change the solubility of the protein.
The FS126, truncation, variant was the most aggregation-prone
of all variants tested. Experimentally truncated variants of wild-
type SOD1 likewise demonstrated very high aggregation
potential. Therefore it is possible that fragmentation of mutant
SOD1 may actually potentiate aggregation.
Does aggregation potential correlate with a clinical
aspect of SOD1-linked FALS?
In studies of SOD1-linked FALS, it has been noted that disease
onset is highly variable, even within a particular kindred, but
disease severity (as indicated by duration) appears to be more
predictable by the location of the mutation (54). However, with
the exception of the A4V and D90A mutations, the number of
affected individuals with a given mutation is usually quite
small. It has been noted that patients harboring the A4V
mutation tend to have very short durations (<2 years) (55–57),
suggesting a more severe mutation. Likewise, patients harbor-
ing the FS126 mutation also show short duration (3 years)
(58). Both of these mutants show very high aggregation
potential in our cell culture assay. However, disease of short
duration has also been noted in patients with the H48Q (59)
and I113T (54–57) mutations, variants that appear to have a
lower aggregation potential in our cell culture assay. Notably,
the measured half-lives of the H48Q and I113T variants are
substantially longer than that of the A4V and FS126 variants
(4,29). We note that, in our cell culture assay, the strongest
correlate to aggregation potential was the measured half-life of
the individual variant. The FS126 truncation variant, which has
the shortest recorded half-life (<5 h) (29), displayed the highest
aggregation potential. Other short-lived variants, including the
A4V, G41D and G85R variants, along with the FS126 variant,
clearly segregated into a distinct group of proteins that were
much more prone to form detergent-insoluble structures
(Fig. 5). In previous studies, we have established that these
four variants exhibit the shortest half-lives when expressed in
cultured cells (29,60). If the accumulation of mis-folded,
aggregating, SOD1 is a key pathologic process, then disease
severity may be due to a complex interaction between the
half-life of the protein and its potential to aggregate. What is
most striking to us is that proteins with relatively short half-
lives have the greatest aggregation potential. Hence, it becomes
easier to explain how these unstable proteins could possess
toxicity that is equivalent to the longer-lived variants.
2760 Human Molecular Genetics, 2003, Vol. 12, No. 21
by guest on December 13, 2016http://hmg.oxfordjournals.org/Downloaded from
CONCLUSIONS
We demonstrate that an experimental variant of SOD1, lacking
all crucial residues for the coordinate binding of Cu, induces an
ALS-like disease in mice that is indistinguishable from the
disease phenotypes of mice expressing natural FALS variants
of SOD1. Disease in these mice, as well as all other models of
FALS mice expressing human SOD1, is accompanied by the
accumulation of high-molecular-weight SOD1 aggregates, and
non-native, detergent-insoluble species of mutant protein. In all
of these respects, the SOD1 models of FALS bear striking
resemblance to other neurodegenerative diseases where the
accumulation of mis-folded and aggregating protein culprits
features in the pathogenesis of disease.
MATERIALS AND METHODS
Transgenic mice, human tissues and cell
culture models
The H46R/H48Q/H63G/H120G mutation was engineered into
a 12 kb fragment of human genomic DNA, encompassing the
entire SOD1 gene (27), by PCR-based oligonucleotide primer-
directed mutagenesis. Each coding exon and at least 50 bp of
intronic sequences on either side of each exon were
subsequently verified by sequence analysis before injection
into mouse embryos (C3H/HeJ C57BL/6J F2). Transgenic
founder mice were identified by PCR amplification of DNA
extracted from tail biopsies. Lines were maintained by crossing
transgenic males to non-transgenic (C57BL/6J C3/HeJ F1)
females (Jackson Laboratories, Bar Harbor, ME, USA). All the
other lines of SOD1 transgenic animals used in this study have
been previously characterized; the G93A variant [B6SJL-
TgN(SOD1-G93A)1Gur; Jackson Laboratory, Bar Harbor,
ME, USA], the G85R variant (line 164) (7), the G37 variant
(line 29) (6), the H46R/H48Q variant (lines 139 and 58) (26),
and the wild-type protein (line 76) (6). The animal use protocol
was approved by the Animal Care and Use Committee of The
Johns Hopkins Medical Institutions.
Human post-mortem spinal cord tissues from a patient
harboring the A4V variant, two familial ALS patients with
unknown causes, 12 patients with sporadic ALS, and three
control cases were obtained from Johns Hopkins Alzheimer’s
Disease Research Center (n¼12) and J. Rothstein at Johns
Hopkins University, School of Medicine (n¼6). HEK 293
cells were transfected with expression plasmids encoding
engineered human SOD1 cDNAs. All point mutations and
fragments were generated based on PCR strategies and expressed
by the pEF-BOS vector (61); many of the constructs used here
have been described previously (4). Disease mutant FS126
comprises a 2bp deletion at codon 126 resulting in a shift in
reading frame and termination five amino acids downstream
from residue 125 (58). Confluent (90%) HEK 293 cells, cultured
by standard procedures in 60 mm wells, were transfected by
using Lipofectamine 2000 (Invitrogen, Carlsbad, CA, USA) and
harvested 24h later for analyses as described below.
Differential detergent extraction and centrifugation
Tissues and cultured cells were homogenized in 1TEN
(10 mM Tris–HCl pH 8.0, 1 mM EDTA pH 8.0, and 100 mM
NaCl); (for tissues, volume : weight ¼10 : 1; for 60 mm culture
dishes, 100 ml lysis buffer per well). Tissue homogenates or cell
lysates were mixed at 1 : 1 with buffer A [1TEN; 1% Nonidet
P40; proteinase inhibitor cocktail 1 : 100 dilution (P 8340,
Sigma, St Louis, MO, USA)]. The mix was sonicated [50%
output for 30 s with a probe sonicator (70 W; TEKMAR,
Cincinnati, OH, USA)] and then centrifuged at >100 000gfor
Figure 8. Hypothetical mechanism of mutant SOD1 aggregation by a domain-swapping mechanism. The asterisk denotes the positions of FALS mutations. Each
arrow represents a b-strand domain of SOD1. The dashed arrows are b-strands contained in portions of the protein deleted by certain FALS-associated mutations.
Human Molecular Genetics, 2003, Vol. 12, No. 21 2761
by guest on December 13, 2016http://hmg.oxfordjournals.org/Downloaded from
5 min to separate supernatant S1 and pellet P1 in a Beckman
Airfuge. The P1 pellet was washed with buffer B (1TEN;
0.5% Nonidet P40) by sonication (50% for 30 s), and
centrifuged at >100 000gfor 5 min to obtain pellet P2, which
was resuspended in buffer C (1TEN; 0.5% Nonidet P40;
0.5% deoxycholic acid; 0.25% SDS). The P2 fraction was
further extracted by sonication (50% for 30 s) and centrifuga-
tion (>100 000gfor 5 min) to sediment P3, which was
resuspended by buffer D (1TEN; 0.5% Nonidet P40; 0.5%
deoxycholic acid; 2% SDS).
Western blotting and immunoprecipitation
Proteins were fractionated by SDS–PAGE on either 8–16%
Tris-glycine or 16.5% tricine–peptide (for SOD1 fragments)
Criterion gels (Bio-Rad, Hercules, CA, USA). Samples were
boiled for 5 min in Laemmli sample buffer containing 2.5%
b-mercaptoethanol. SOD1 was detected by immunoblotting
with rabbit polyclonal antibody hSOD1 (7), or rabbit polyclonal
antibody m/hSOD1 (4). A
125
I-labeled donkey-anti-rabbit
antibody (Amersham, Piscataway, NJ, USA) was used in
quantitative assessment of SOD1; between 1 and 10 mgof
protein were loaded per lane. Protein amounts were measured
by a BCA kit (Pierce, Rockford, IL, USA). The signals were
read by a Phosphorimager (Bio-Rad). A monoclonal ubiquitin
antibody (5–25) was used in western blotting for ubiquitinated
proteins (Signet, Dedham, MA, USA). Antibodies against
Hsp25 (SPA-801), aB-crystallin (SPA-222), Hsp40 (SPA-450),
Hsp60 (SPA-806), Hsp70 (SPA-812), and Hsp90 (SPA-830)
were from Stressgen (Victoria, BC, Canada).
For immunoprecipitation, the m/hSOD1 polyclonal antibody
was cross-linked to immobilized protein A by using the Seize
X Protein A immunoprecipitation kit (Pierce). Forty micro-
grams of P2 fraction protein were adjusted to 2% SDS, boiled
for 10 min, diluted to 0.2% SDS, immunoprecipitated with the
above protein A/SOD1 antibody complex, separated by SDS–
PAGE, and blotted with the monoclonal ubiquitin antibody
(5–25; Signet).
SOD1 activity gel
SOD1 activities were determined as previously described (4,62).
Spinal cords were homogenized in 1phosphate buffered saline
(PBS; volume : weight ¼10 : 1), and then centrifuged at 10 000g
for 5 min. One hundred micrograms of supernatant proteins were
electrophoresed on a 7.5% polyacrylamide native gel, where an
in-gel reaction was carried out in the final step.
Filter trap assay
The assay was carried out as previously described (24). Tissue
homogenates were centrifuged at 800gfor 5 min in a micro-
centrifuge. The supernatant was frozen at 70C until used.
Protein concentrations were determined by BCA before filtering
through acetate membranes (0.2 mm pore size; Schleicher and
Schuell, Keene, NH, USA). The samples containing 100 mgof
protein were thawed and diluted with 20 vols of 1PBS (pH
7.4) containing 1% SDS. The solution was sonicated (70 W,
50% output, 30 s) before membrane filtering using a 96-well
dot-blot apparatus (Bio-Rad). Each well was washed twice with
1PBS and the membranes were immunostained with the m/
hSOD1 antiserum, in a fashion similar to a standard immuno-
blot, before enhanced chemiluminescence detection.
Histopathology and immunocytochemistry
Mice anesthetized with ethyl ether were sacrificed by trans-
cardial perfusion with 1PBS (pH 7.4), followed by 4%
paraformaldehyde in 1PBS. Brains and spinal cords were
removed, post-fixed in the same fixative, embedded in paraffin,
and sectioned for histologic and immunologic staining. Sagittal
sections of the brain and brain stem and cross sections of spinal
cord (10 mm) were stained with hematoxylin and eosin, silver
impregnation and thioflavin-S. Deparaffinized sections were
processed for immunocytochemistry, using antibodies against
ubiquitin (DAKO, Carpenteria, CA, USA). The immune
reaction was visualized with diaminobenzidine, and sections
were counterstained with hematoxylin.
SUPPLEMENTARY MATERIAL
Supplementary Material is available at HMG Online.
ACKNOWLEDGEMENTS
We are very grateful to Ms Debbie Swing for her help in
transgene injections. We thank Drs Wenxue Li, Michael K. Lee,
Guilian Xu and Alena V. Savonenko for their helpful discussion
and assistance. This study was supported by the ALS
Association, the Muscular Dystrophy Association, and by the
Robert Packard Center for ALS Research at Johns Hopkins
University.
REFERENCES
1. Rosen, D.R., Siddique, T., Patterson, D., Figlewicz, D.A., Sapp, P.,
Hentati, A., Donaldson, D., Goto, J., O’Regan, J.P., Deng, H.-X., et al.
(1993) Mutations in Cu/Zn superoxide dismutase gene are associated with
familial amyotrophic lateral sclerosis. Nature,362, 59–62.
2. Rowland, L.P. (1991) Amyotrophic Lateral Sclerosis and Other Motor
Neuron Diseases. Raven Press, New York.
3. Gurney, M.E., Pu, H., Chiu, A.Y., Dal Canto, M.C., Polchow, C.Y.,
Alexander, D.D., Caliendo, J., Hentati, A., Kwon, Y.W., Deng, H.-X., et al.
(1994) Motor neuron degeneration in mice that express a human Cu,Zn
superoxide dismutase mutation. Science,264, 1772–1775.
4. Borchelt, D.R., Lee, M.K., Slunt, H.H., Guarnieri, M., Xu, Z.-S.,
Wong, P.C., Brown, R.H., Jr, Price, D.L., Sisodia, S.S. and Cleveland, D.W.
(1994) Superoxide dismutase 1 with mutations linked to familial
amyotrophic lateral sclerosis possesses significant activity. Proc. Natl Acad.
Sci. USA,91, 8292–8296.
5. Ripps, M.E., Huntley, G.W., Hof, P.R., Morrison, J.H. and Gordon, J.W.
(1995) Transgenic mice expressing an altered murine superoxide dismutase
gene provide an animal model of amyotrophic lateral sclerosis. Proc. Natl
Acad. Sci. USA,92, 689–693.
6. Wong, P.C., Pardo, C.A., Borchelt, D.R., Lee, M.K., Copeland, N.G.,
Jenkins, N.A., Sisodia, S.S., Cleveland, D.W. and Price, D.L. (1995) An
adverse property of a familial ALS-linked SOD1 mutation causes motor
neuron disease characterized by vacuolar degeneration of mitochondria.
Neuron,14, 1105–1116.
7. Bruijn, L.I., Becher, M.W., Lee, M.K., Anderson, K.L., Jenkins, N.A.,
Copeland, N.G., Sisodia, S.S., Rothstein, J.D., Borchelt, D.R., Price, D.L.
and Cleveland, D.W. (1997) ALS-linked SOD1 mutant G85R mediates
damage to astrocytes and promotes rapidly progressive disease with
SOD1-containing inclusions. Neuron,18, 327–338.
2762 Human Molecular Genetics, 2003, Vol. 12, No. 21
by guest on December 13, 2016http://hmg.oxfordjournals.org/Downloaded from
8. Wiedau-Pazos, M., Goto, J.J., Rabizadeh, S., Gralla, E.B., Roe, J.A.,
Lee, M.K., Valentine, J.S. and Bredesen, D.E. (1996) Altered reactivity of
superoxide dismutase in familial amyotrophic lateral sclerosis. Science,
271, 515–518.
9. Yim, M.B., Chock, P.B. and Stadtman, E.R. (1993) Enzyme function of
copper, zinc superoxide dismutase as a free radical generator. J. Biol.
Chem.,268, 4099–4105.
10. Yim, M.B., Kang, J.H., Yim, H.S., Kwak, H.S., Chock, P.B. and
Stadtman, E.R. (1996) A gain-of-function of an amyotrophic lateral
sclerosis-associated Cu,Zn-superoxide dismutase mutant: An enhancement
of free radical formation due to a decrease in K
m
for hydrogen peroxide.
Proc. Natl Acad. Sci. USA,93, 5709–5714.
11. Liochev, S.I., Chen, L.L., Hallewell, R.A. and Fridovich, I. (1997)
Superoxide-dependent peroxidase activity of H48Q: a superoxide
dismutase variant associated with familial amyotrophic lateral sclerosis.
Arch. Biochem. Biophys.,346, 263–268.
12. Beckman, J.S., Carson, M., Smith, C.D. and Koppenol, W.H. (1993)
ALS, SOD and peroxynitrite. Nature,364, 584.
13. Crow, J.P., Sampson, J.B., Zhuang, Y., Thompson, J.A. and
Beckman, J.S. (1997) Decreased zinc affinity of amyotrophic lateral
sclerosis-associated superoxide dismutase mutants leads to enhanced
catalysis of tyrosine nitration by peroxynitrite. J. Neurochem.,69,
1936–1944.
14. Este
´vez, A.G., Crow, J.P., Sampson, J.B., Reiter, C., Zhuang, Y.,
Richardson, G.J., Tarpey, M.M., Barbeito, L. and Beckman, J.S. (1999)
Induction of nitric oxide-dependent apoptosis in motor neurons by zinc-
deficient superoxide dismutase. Science,286, 2498–2500.
15. Subramaniam, J.R., Lyons, W.E., Liu, J., Bartnikas, T.B., Rothstein, J.,
Price, D.L., Cleveland, D.W., Gitlin, J.D. and Wong, P.C. (2002) Mutant
SOD1 causes motor neuron disease independent of copper chaperone-
mediated copper loading. Nat. Neurosci.,5, 301–307.
16. Bush, A.I. (2002) Is ALS caused by an altered oxidative activity of
mutant superoxide dismutase? Nat. Neurosci.,5, 919–920.
17. Shibata, N., Hirano, M. and Kobayashi, K. (1993) Immunohistochemical
demonstration of Cu/Zn superoxide dismutase in the spinal cord of patients
with familial amyotrophic lateral sclerosis. Acta Histochem. Cytochem.,26,
619–622.
18. Shibata, N., Hirano, A., Kobayashi, M., Sasaki, S., Kato, T., Matsumoto, S.,
Shiozawa, Z., Komori, T., Ikemoto, A., Umahara, T. and Asayama, K.
(1994) Cu/Zn superoxide dismutase-like immunoreactivity in Lewy
body-like inclusions of sporadic amyotrophic lateral sclerosis. Neurosci.
Lett.,179, 149–152.
19. Watanabe, M., Dykes-Hoberg, M., Cizewski, C., V, Price, D.L., Wong, P.C.
and Rothstein, J.D. (2001) Histological evidence of protein aggregation in
mutant SOD1 transgenic mice and in amyotrophic lateral sclerosis neural
tissues. Neurobiol. Dis.,8, 933–941.
20. Koide, T., Igarashi, S., Kikugawa, K., Nakano, R., Inuzuka, T., Yamada, M.,
Takahashi, H. and Tsuji, S. (1998) Formation of granular cytoplasmic
aggregates in COS7 cells expressing mutant Cu-Zn superoxide dismutase
associated with familial amyotrophic lateral sclerosis. Neurosci. Lett.,257,
29–32.
21. Durham, H.D., Roy, J., Dong, L. and Figlewicz, D.A. (1997) Aggregation
of mutation Cu/Zn superoxide dismutase proteins in a culture model of
ALS. J. Neuropathol. Exp. Neurol.,56, 523–530.
22. Johnston, J.A., Dalton, M.J., Gurney, M.E. and Kopito, R.R. (2000)
Formation of high molecular weight complexes of mutant Cu, Zn-
superoxide dismutase in a mouse model for familial amyotrophic lateral
sclerosis. Proc. Natl Acad. Sci. USA,97, 12571–12576.
23. Shinder, G.A., Lacourse, M.C., Minotti, S. and Durham, H.D. (2001)
Mutant Cu/Zn-superoxide dismutase proteins have altered solubility and
interact with heat shock/stress proteins in models of amyotrophic lateral
sclerosis. J. Biol. Chem.,276, 12791–12796.
24. Wang, J., Xu, G. and Borchelt, D.R. (2002) High molecular weight
complexes of mutant superoxide dismutase 1: age- dependent and
tissue-specific accumulation. Neurobiol. Dis.,9, 139–148.
25. Parge, H.E., Hallewell, R.A. and Tainer, J.A. (1992) Atomic structures of
wild-type and thermostable mutant recombinant human Cu,Zn superoxide
dismutase. Proc. Natl Acad. Sci. USA,89, 6109–6113.
26. Wang, J., Xu, G., Gonzales, V., Coonfield, M., Fromholt, D.,
Copeland, N.G., Jenkins, N.A. and Borchelt, D.R. (2002) Fibrillar
inclusions and motor neuron degeneration in transgenic mice expressing
superoxide dismutase 1 with a disrupted copper-binding site.
Neurobiol. Dis.,10, 128–138.
27. Groner, Y., Lieman-Hurwitz, J., Dafni, N., Sherman, L., Levanon, D.,
Bernstein, Y., Danciger, E. and Elroy-Stein, O. (1985) Molecular structure
and expression of the gene locus on chromosome 21 encoding the Cu/Zn
superoxide dismutase and its relevance to Down syndrome. Ann. N.Y. Acad.
Sci.,450, 133–156.
28. Ischiropoulos, H., Zhu, L., Chen, J., Tsai, M., Martin, J.C., Smith, C.D. and
Beckman, J.S. (1992) Peroxynitrite-mediated tyrosine nitration catalyzed by
superoxide dismutase. Arch. Biochem. Biophys.,298, 431–437.
29. Ratovitski, T., Corson, L.B., Strain, J., Wong, P., Cleveland, D.W.,
Culotta, V.C. and Borchelt, D.R. (1999) Variation in the biochemical/
biophysical properties of mutant superoxide dismutase 1 enzymes and the
rate of disease progression in familial amyotrophic lateral sclerosis
kindreds. Hum. Mol. Genet.,8, 1451–1460.
30. Narberhaus, F. (2002) a-Crystallin-type heat shock proteins: socializing
minichaperones in the context of a multichaperone network. Microbiol.
Mol. Biol. Rev.,66, 64–93.
31. Yim, M.B., Chock, P.B. and Stadtman, E.R. (1990) Copper, zinc superoxide
dismutase catalyzes hydroxyl radical production from hydrogen peroxide.
Proc. Natl Acad. Sci. USA,87, 5006–5010.
32. McKinley, M.P., Meyer, R.K., Kenaga, L., Rahbar, F., Cotter, R., Serban, A.
and Prusiner, S.B. (1991) Scrapie prion rod formation in vitro requires both
detergent extraction and limited proteolysis. J. Virol.,65, 1340–1351.
33. Pike, C.J., Walencewicz, A.J., Glabe, C.G. and Cotman, C.W. (1991)
In vitro aging of b-amyloid protein causes peptide aggregation and
neurotoxicity. Brain Res.,563, 311–314.
34. Xu, G., Gonzales, V. and Borchelt, D.R. (2002) Rapid detection of protein
aggregates in the brains of Alzheimer patients and transgenic mouse models
of amyloidosis. Alzheimer Dis. Assoc. Disord.,16, 191–195.
35. Giasson, B.I., Uryu, K., Trojanowski, J.Q. and Lee, V.M. (1999) Mutant and
wild-type human alpha-synucleins assemble into elongated filaments with
distinct morphologies in vitro.J. Biol. Chem.,274, 7619–7622.
36. Ishihara, T., Hong, M., Zhang, B., Nakagawa, Y., Lee, M.K.,
Trojanowski, J.Q. and Lee, V.M. (1999) Age-dependent emergence and
progression of a tauopathy in transgenic mice overexpressing the shortest
human tau isoform. Neuron,24, 751–762.
37. Ishihara, T., Zhang, B., Higuchi, M., Yoshiyama, Y., Trojanowski, J.Q. and
Lee, V.M. (2001) Age-dependent induction of congophilic neurofibrillary
tau inclusions in tau transgenic mice. Am. J. Pathol.,158, 555–562.
38. Scherzinger, E., Sittler, A., Schweiger, K., Heiser, V., Lurz, R.,
Hasenbank, R., Lehrach, H. and Wanker, E.E. (1999) Self-assembly of
polyglutamine-containing huntingtin fragments into amyloid-like fibrils:
implications for Huntington’s disease pathology. Proc. Natl Acad. Sci. USA,
96, 4604–4609.
39. Guntern, R., Bouras, C., Hof, P.R. and Vallet, P.G. (1992) An improved
thioflavine S method for staining neurofibrillary tangles and senile plaques
in Alzheimer’s disease. Experientia,48, 8–10.
40. Yamamoto, T. and Hirano, A. (1986) A comparative study of modified
Bielschowsky, Bodian and thioflavin S stains on Alzheimer’s neurofibrillary
tangles. Neuropathol. Appl. Neurobiol.,12, 3–9.
41. Jarrett, J.T. and Lansbury, P.T., Jr (1993) Seeding ‘one-dimensional
crystallization’ of amyloid: a pathogenic mechanism in Alzheimer’s disease
and scrapie? Cell,73, 1055–1058.
42. Burdick, D., Soreghan, B., Kwon, M., Kosmoski, J., Knauer, M.,
Henschen, A., Yates, J., Cotman, C. and Glabe, C. (1992) Assembly and
aggregation properties of synthetic Alzheimer’s A4/bamyloid peptide
analogs. J. Biol. Chem.,267, 546–554.
43. Poirier, M.A., Li, H., Macosko, J., Cai, S., Amzel, M. and Ross, C.A.
(2002) Huntingtin spheroids and protofibrils as precursors in polyglutamine
fibrilization. J. Biol. Chem.,277, 41032–41037.
44. Thakur, A.K., and Wetzel, R. (2002) Mutational analysis of the structural
organization of polyglutamine aggregates. Proc. Natl Acad. Sci. USA,99,
17014–17019.
45. Caughey, B. and Lansbury, P.T., Jr (2003) Protofibrils, pores, fibrils, and
neurodegeneration: separating the responsible protein aggregates from the
Innocent Bystanders. A. Rev. Neurosci.,26, 276–298.
46. Stromer, T., Ehrnsperger, M., Gaestel, M. and Buchner, J. (2003) Analysis
of the interaction of small heat shock proteins with unfolding proteins.
J. Biol. Chem.,278, 18015–18021.
47. Berke, S.J., and Paulson, H.L. (2003) Protein aggregation and the ubiquitin
proteasome pathway: gaining the UPPer hand on neurodegeneration. Curr.
Opin. Genet. Dev.,13, 253–261.
48. Bence, N.F., Sampat, R.M. and Kopito, R.R. (2001) Impairment of the
ubiquitin–proteasome system by protein aggregation. Science,292, 1552–1555.
Human Molecular Genetics, 2003, Vol. 12, No. 21 2763
by guest on December 13, 2016http://hmg.oxfordjournals.org/Downloaded from
49. Janowski, R., Kozak, M., Jankowska, E., Grzonka, Z., Grubb, A.,
Abrahamson, M. and Jaskolski, M. (2001) Human cystatin C, an
amyloidogenic protein, dimerizes through three- dimensional domain
swapping. Nat. Struct. Biol.,8, 316–320.
50. Jaskolski, M. (2001) 3D domain swapping, protein oligomerization, and
amyloid formation. Acta Biochim. Pol.,48, 807–827.
51. Rodriguez, J.A., Valentine, J.S., Eggers, D.K., Roe, J.A., Tiwari, A.,
Brown, R.H., Jr and Hayward, L.J. (2002) Familial amyotrophic lateral
sclerosis-associated mutations decrease the thermal stability of distinctly
metallated species of human copper/zinc superoxide dismutase. J. Biol.
Chem.,277, 15932–15937.
52. Lindberg, M.J., Tibell, L. and Oliveberg, M. (2002) Common denominator
of Cu/Zn superoxide dismutase mutants associated with amyotrophic lateral
sclerosis: Decreased stability of the apo state. Proc. Natl Acad. Sci. USA,
99, 16607–16612.
53. Tiwari, A. and Hayward, L.J. (2003) Familial amyotrophic lateral sclerosis
mutants of copper/zinc superoxide dismutase are susceptible to disulfide
reduction. J. Biol. Chem.,278, 5984–5992.
54. Cleveland, D.W., Laing, N., Hurse, P.V. and Brown, R.H., Jr (1995) Toxic
mutants in Charcot’s sclerosis. Nature,378, 342–343.
55. Cudkowicz, M.E., McKenna-Yasek, D., Sapp, P.E., Chin, W., Geller, B.,
Hayden, D.L., Schoenfeld, D.A., Hosler, B.A., Horvitz, H.R. and
Brown, R.H. (1997) Epidemiology of mutations in superoxide dismutase
in amyotrophic lateral sclerosis. Ann. Neurol.,41, 210–221.
56. Radunovic, A. and Leigh, P.N. (1996) Cu/Zn superoxide dismutase gene
mutations in amyotrophic lateral sclerosis: correlation between genotype
and clinical features. J. Neurol. Neurosurg. Psychiat.,61, 565–572.
57. Juneja, T., Pericak-Vance, M.A., Laing, N.G., Dave, S. and Siddique, T.
(1997) Prognosis in familial amyotrophic lateral sclerosis: progression and
survival in patients with glu100gly and ala4val mutations in Cu,Zn
superoxide dismutase. Neurology,48, 55–57.
58. Pramatarova, A., Goto, J., Nanba, E., Nakashima, K., Takahashi, K.,
Takagi, A., Kanazawa, I., Figlewicz, D.A. and Rouleau, G.A. (1994) A two
basepair deletion in the SOD 1 gene causes familial amyotrophic lateral
sclerosis. Hum. Mol. Genet.,3, 2061–2062.
59. Enayat, Z.E., Orrell, R.W., Claus, A., Ludolph, A., Bachus, R.,
Brockmu¨ller, J., Ray-Chaudhuri, K., Radunovic, A., Shaw, C., Wilkinson, J.
et al. (1995) Two novel mutations in the gene for copper zinc superoxide
dismutase in UK families with amyotrophic lateral sclerosis. Hum. Mol.
Genet.,4, 1239–1240.
60. Borchelt, D.R., Guarnieri, M., Wong, P.C., Lee, M.K., Slunt, H.S., Xu, Z.-S.,
Sisodia, S.S., Price, D.L. and Cleveland, D.W. (1995) Superoxide dismutase 1
subunits with mutations linked to familial amyotrophic lateral sclerosis do
not affect wild-type subunit function. J. Biol. Chem.,270, 3234–3238.
61. Mizushima, S. and Nagata, S. (1990) pEF-BOS, a powerful mammalian
expression vector. Nucl. Acids Res.,18, 5322.
62. Beauchamp, C. and Fridovich, I. (1971) Superoxide dismutase: improved
assays and an assay applicable to acrylamide gels. Anal. Biochem.,44, 276–287.
2764 Human Molecular Genetics, 2003, Vol. 12, No. 21
by guest on December 13, 2016http://hmg.oxfordjournals.org/Downloaded from
... Differential extractions were used to assess SOD1 aggregation as previously described [14,15]. Cells were scraped Content courtesy of Springer Nature, terms of use apply. ...
... The accumulation of detergent-insoluble SOD1 aggregates is a common feature of ALS that may act as a transmissible agent between neurons [53]. As such, both the detergent-soluble and -insoluble SOD1 may be potentially toxic and are considered one of the mechanisms by which mutant/ misfolded SOD1 causes ALS [15]. Using aggregation assay, we demonstrated that mutant SOD1 was prone to form the detergent-insoluble aggregates 48 h after transfection with G93A-hSOD1 plasmid. ...
Article
Full-text available
Background Amyotrophic lateral sclerosis (ALS) is a devastating neurodegenerative disease. There is no cure currently. The discovery that mutations in the gene SOD1 are a cause of ALS marks a breakthrough in the search for effective treatments for ALS. SOD1 is an antioxidant that is highly expressed in motor neurons. Human SOD1 is prone to aberrant modifications. Familial ALS-linked SOD1 variants are particularly susceptible to aberrant modifications. Once modified, SOD1 undergoes conformational changes and becomes misfolded. This study aims to determine the effect of selective removal of misfolded SOD1 on the pathogenesis of ALS. Methods Based on the chaperone-mediated protein degradation pathway, we designed a fusion peptide named CT4 and tested its efficiency in knocking down intracellularly misfolded SOD1 and its efficacy in modifying the pathogenesis of ALS. Results Expression of the plasmid carrying the CT4 sequence in human HEK cells resulted in robust removal of misfolded SOD1 induced by serum deprivation. Co-transfection of the CT4 and the G93A-hSOD1 plasmids at various ratios demonstrated a dose-dependent knockdown efficiency on G93A-hSOD1, which could be further increased when misfolding of SOD1 was enhanced by serum deprivation. Application of the full-length CT4 peptide to primary cultures of neurons expressing the G93A variant of human SOD1 revealed a time course of the degradation of misfolded SOD1; misfolded SOD1 started to decrease by 2 h after the application of CT4 and disappeared by 7 h. Intravenous administration of the CT4 peptide at 10 mg/kg to the G93A-hSOD1 reduced human SOD1 in spinal cord tissue by 68% in 24 h and 54% in 48 h in presymptomatic ALS mice. Intraperitoneal administration of the CT4 peptide starting from 60 days of age significantly delayed the onset of ALS and prolonged the lifespan of the G93A-hSOD1 mice. Conclusions The CT4 peptide directs the degradation of misfolded SOD1 in high efficiency and specificity. Selective removal of misfolded SOD1 significantly delays the onset of ALS, demonstrating that misfolded SOD1 is the toxic form of SOD1 that causes motor neuron death. The study proves that selective removal of misfolded SOD1 is a promising treatment for ALS.
... SOD1 aggregation and thiol oxidation assay. Differential extractions were used to assess SOD1 aggregation as previously described [12,13]. Cells were scraped from the culture dish in PBS and centrifuged to pellet the cells before the pellets were resuspended in 100 µl 1 × TEN (10 mM Tris, 1 mM EDTA, and 100 mM NaCl). ...
... The accumulation of detergent-insoluble SOD1 aggregates is a common feature of ALS that may act as a transmissible agent between neurons [45]. As such, both of the detergent-soluble and -insoluble SOD1 may be potentially toxic and are considered as one of the mechanisms by which mutant/misfolded SOD1 causes ALS [13]. Using aggregation assay, we demonstrated that mutant SOD1 was prone to form the detergent-insoluble aggregates 48 h after transfection with SOD1-G93A plasmid. ...
Preprint
Full-text available
Background Amyotrophic lateral sclerosis (ALS) is a devastating neurodegenerative disease. There is no cure currently. The discovery that mutations in the gene SOD1 are a cause of ALS marks a breakthrough for the search of effective treatments for ALS. SOD1 is an antioxidant that is highly expressed in motor neurons. Human SOD1 is prone to aberrant modifications. Familial ALS-linked SOD1 variants are particularly susceptible to aberrant modifications. Once modified, SOD1 undergoes conformational changes and becomes misfolded. This study aims to determine the effect of selective removal of misfolded SOD1 on the pathogenesis of ALS. Methods Based on chaperone-mediated protein degradation pathway, we designed a fusion peptide named CT4, and tested its efficiency in knocking down intracellularly misfolded SOD1 and its efficacy in modifying pathogenesis of ALS. Results Expression of plasmid carrying the CT4 sequence in human HEK cells resulted in robust removal of misfolded SOD1 induced by serum deprivation. Co-transfection of the CT4 and the human SOD1 G93A plasmids at various ratios in rat PC12 cells demonstrated a dose-dependent knockdown efficiency on G93A, which could be further increased when misfolding of SOD1 was enhanced by serum deprivation. Application of the full length CT4 peptide to primary cultures of neurons expressing the G93A variant of human SOD1 revealed a time-course of the degradation of misfolded SOD1; misfolded SOD1 started to decrease by 2 h after the application of CT4 and disappeared by 7 h. Intravenous administration of the CT4 peptide at 10 mg/kg to the G93A mice at the age of 4 months old induced reduction of human SOD1 in spinal cord tissue by 68% in 24 h and 54% in 48 h. Intraperitoneal administration of the CT4 peptide starting from 60 days of age significantly delayed the onset of ALS and prolonged the lifespan of the G93A mice. Conclusions The CT4 peptide directs degradation of misfolded SOD1 in high efficiency and specificity. Selective removal of misfolded SOD1 significantly delays the onset of ALS, demonstrating that misfolded SOD1 is the toxic form of SOD1 that causes motor neuron death. The study provides a proof of concept that selective removal of misfolded SOD1 is a promising treatment for ALS.
... This result is in favor of a contribution of this gene to a reduction in neuronal survival in vivo. HSPD1 expression was not significantly decreased, which is consistent with previous observations in a mouse model of ALS and PD [156,157]. Nicotinamide phosphoribosyltransferase (NAMPT) is a catalyzer of nicotinamide adenine dinucleotide (NAD). Decreased levels of NAD, a cofactor in numerous redox reactions, have been observed in the aging brain and are closely associated with classical ALS and PD progression [75]. ...
Article
Full-text available
Amyotrophic Lateral Sclerosis/Parkinsonism-Dementia Complex (ALS/PDC), a rare and complex neurological disorder, is predominantly observed in the Western Pacific islands, including regions of Japan, Guam, and Papua. This enigmatic condition continues to capture medical attention due to affected patients displaying symptoms that parallel those seen in either classical amyotrophic lateral sclerosis (ALS) or Parkinson’s disease (PD). Distinctly, postmortem examinations of the brains of affected individuals have shown the presence of α-synuclein aggregates and TDP-43, which are hallmarks of PD and classical ALS, respectively. These observations are further complicated by the detection of phosphorylated tau, accentuating the multifaceted proteinopathic nature of ALS/PDC. The etiological foundations of this disease remain undetermined, and genetic investigations have yet to provide conclusive answers. However, emerging evidence has implicated the contribution of astrocytes, pivotal cells for maintaining brain health, to neurodegenerative onset, and likely to play a significant role in the pathogenesis of ALS/PDC. Leveraging advanced induced pluripotent stem cell technology, our team cultivated multiple astrocyte lines to further investigate the Japanese variant of ALS/PDC (Kii ALS/PDC). CHCHD2 emerged as a significantly dysregulated gene when disease astrocytes were compared to healthy controls. Our analyses also revealed imbalances in the activation of specific pathways: those associated with astrocytic cilium dysfunction, known to be involved in neurodegeneration, and those related to major neurological disorders, including classical ALS and PD. Further in-depth examinations revealed abnormalities in the mitochondrial morphology and metabolic processes of the affected astrocytes. A particularly striking observation was the reduced expression of CHCHD2 in the spinal cord, motor cortex, and oculomotor nuclei of patients with Kii ALS/PDC. In summary, our findings suggest a potential reduction in the support Kii ALS/PDC astrocytes provide to neurons, emphasizing the need to explore the role of CHCHD2 in maintaining mitochondrial health and its implications for the disease.
... Differential extractions were used to assess SOD1 aggregation as previously described [21,22]. Cells were scraped from the culture dish in PBS and centrifuged to pellet the cells before the pellets were re-suspended in 100 μl 1 × TEN (10 mM Tris, 1 mM EDTA, and 100 mM NaCl). ...
Article
Full-text available
Background Neural stem cells (NSCs), especially human NSCs, undergo cellular senescence characterized by an irreversible proliferation arrest and loss of stemness after prolonged culture. While compelling correlative data have been generated to support the oxidative stress theory as one of the primary determinants of cellular senescence of NSCs, a direct cause-and-effect relationship between the accumulation of oxidation-mediated damage and cellular senescence of NSCs has yet to be firmly established. Human SOD1 (hSOD1) is susceptible to oxidation. Once oxidized, it undergoes aberrant misfolding and gains toxic properties associated with age-related neurodegenerative disorders. The present study aims to examine the role of oxidized hSOD1 in the senescence of NSCs. Methods NSCs prepared from transgenic mice expressing the wild-type hSOD1 gene were maintained in culture through repeated passages. Extracellular vesicles (EVs) were isolated from culture media at each passage. To selectively knock down oxidized SOD1 in NSCs and EVs, we used a peptide-directed chaperone-mediated protein degradation system named CT4 that we developed recently. Results In NSCs expressing the hSOD1 from passage 5, we detected a significant increase of oxidized hSOD1 and an increased expression of biomarkers of cellular senescence, including upregulation of P53 and SA-β-Gal and cytoplasmic translocation of HMGB1. The removal of oxidized SOD1 remarkably increased the proliferation and stemness of the NSCs. Meanwhile, EVs derived from senescent NSCs carrying the wild-type hSOD1 contained high levels of oxidized hSOD1, which could accelerate the senescence of young NSCs and induce the death of cultured neurons. The removal of oxidized hSOD1 from the EVs abolished their senescence-inducing activity. Blocking oxidized SOD1 on EVs with the SOD1 binding domain of the CT4 peptide mitigated its toxicity to neurons. Conclusion Oxidized hSOD1 is a causal factor in the cellular senescence of NSCs. The removal of oxidized hSOD1 is a strategy to rejuvenate NSCs and to improve the quality of EVs derived from senescent cells.
... Since protein misfolding and aggregation are considered as central events in ALS pathogenesis, therapeutic strategies targeting these processes hold promise (Elliott et al., 2020). The understanding of SOD1 aggregation kinetics has introduced new possibilities for ALS therapy, by the discovery of small molecules that bind and stabilize either (i) the native state of SOD1 or (ii) the intermediate aggregate species (misfolded monomers, low molecular SOD1 aggregates) (Banerjee et al., 2016;Wang et al., 2003), (Rakhit et al., 2004). ...
Preprint
Cu/Zn Superoxide Dismutase 1 (SOD1) is a 32-kDa cytosolic dimeric metalloenzyme that neutralizes superoxide anions into harmless oxygen and hydrogen peroxide. Mutations in SOD1 are associated with ALS, a disease causing motor neuron atrophy and subsequent mortality. These mutations exert their harmful effects through a gain of function mechanism, rather than loss of function. Despite extensive research, the specific mechanism causing selective motor neuron death still remains unclear. A defining feature of ALS pathogenesis is protein misfolding and aggregation, evidenced by ubiquitinated protein inclusions containing SOD1 in motor neurons. This work aims to identify compounds countering SOD1(A4V) misfolding and aggregation, potentially aiding ALS treatment. The approach employed is drug repurposing and in vitro screening of a 1280 pharmacologically active compounds library, LOPAC®. Using Differential Scanning Fluorimetry Technique (DSF), compounds were tested for their impact on SOD1(A4V) thermal stability. Screening revealed one compound raising protein-ligand Tm by 7oC, eight inducing a higher second Tm, suggesting stabilzation effect, and five reducing Tm up to 18oC, suggesting possible interactions or non-specific binding.
... Detergent insoluble SOD1 aggregates were measured in spinal cord homogenates as previously described [5,28]. Briefly, spinal cords from control or Atp1a2 ASO-treated SOD1*G93A mice were weighed and mixed with 10 volumes of 1x TEN buffer (10mM Tris, pH-7.5; ...
Article
Full-text available
Astrocyte-specific ion pump α2-Na ⁺ /K ⁺ -ATPase plays a critical role in the pathogenesis of amyotrophic lateral sclerosis (ALS). Here, we test the effect of Atp1a2 mRNA-specific antisense oligonucleotides (ASOs) to induce α2-Na ⁺ /K ⁺ -ATPase knockdown in the widely used ALS animal model, SOD1*G93A mice. Two ASOs led to efficient Atp1a2 knockdown and significantly reduced SOD1 aggregation in vivo . Although Atp1a2 ASO-treated mice displayed no off-target or systemic toxicity, the ASO-treated mice exhibited an accelerated disease onset and shorter lifespan than control mice. Transcriptomics studies reveal downregulation of genes involved in oxidative response, metabolic pathways, trans-synaptic signaling, and upregulation of genes involved in glutamate receptor signaling and complement activation, suggesting a potential role for these molecular pathways in de-coupling SOD1 aggregation from survival in Atp1a2 ASO-treated mice. Together, these results reveal a role for α2-Na ⁺ /K ⁺ -ATPase in SOD1 aggregation and highlight the critical effect of temporal modulation of genetically validated therapeutic targets in neurodegenerative diseases.
... Without copper, SOD1 is less stable under physiological conditions (Kabuta et al., 2006). Moreover, copper-deficient SOD1 monomers provide a precursor to the formation of SOD1 aggregates (Furukawa, 2013) and copper insufficiency causes the protein to misfold in a fashion similar to other amyloidogenic proteins associated with neurodegenerative disease (Furukawa et al., 2008;Wang et al., 2003). ...
Article
Full-text available
Since the first description of Parkinson's disease (PD) over two centuries ago, the recognition of rare types of atypical parkinsonism has introduced a spectrum of related PD‐like diseases. Among these is progressive supranuclear palsy (PSP), a neurodegenerative condition that clinically differentiates through the presence of additional symptoms uncommon in PD. As with PD, the initial symptoms of PSP generally present in the sixth decade of life when the underpinning neurodegeneration is already significantly advanced. The causal trigger of neuronal cell loss in PSP is unknown and treatment options are consequently limited. However, converging lines of evidence from the distinct neurodegenerative conditions of PD and amyotrophic lateral sclerosis (ALS) are beginning to provide insights into potential commonalities in PSP pathology and opportunity for novel therapeutic intervention. These include accumulation of the high abundance cuproenzyme superoxide dismutase 1 (SOD1) in an aberrant copper‐deficient state, associated evidence for altered availability of the essential micronutrient copper, and evidence for neuroprotection using compounds that can deliver available copper to the central nervous system. Herein, we discuss the existing evidence for SOD1 pathology and copper imbalance in PSP and speculate that treatments able to provide neuroprotection through manipulation of copper availability could be applicable to the treatment of PSP. image
Article
Protein misfolding and amyloid formation are hallmarks of numerous diseases, including amyotrophic lateral sclerosis (ALS), in which hSOD1 aggregation is involved in pathogenesis. We used two point mutations in the electrostatic loop, G138E and T137R, to analyze charge distribution under destabilizing circumstances to gain more about how ALS-linked mutations affect SOD1 protein stability or net repulsive charge. We show that protein charge is important in the ALS disease process using bioinformatics and experiments. The MD simulation findings demonstrate that the mutant protein differs significantly from WT SOD1, which is consistent with the experimental evidence. The specific activity of the wild type was 1.61 and 1.48 times higher than that of the G138E and T137R mutants, respectively. Under amyloid induction conditions, the intensity of intrinsic and ANS fluorescence in both mutants reduced. Increasing the content of β-sheet structures in mutants can be attributed to aggregation propensity, which was confirmed using CD polarimetry and FTIR spectroscopy. Our findings show that two ALS-related mutations promote the formation of amyloid-like aggregates at near physiological pH under destabilizing conditions, which were detected using spectroscopic probes such as Congo red and ThT fluorescence, and also further confirmation of amyloid-like species by TEM. Overall, our results provide evidence supporting the notion that negative charge changes combined with other destabilizing factors play an important role in increasing protein aggregation by reducing repulsive negative charges.
Chapter
In recent years, medical developments have resulted in an increase in human life expectancy. Some developed countries now have a larger population of individuals aged over 64 than those under 14. One consequence of the ageing population is a higher incidence of certain neurodegenerative disorders. In order to prevent these, we need to learn more about them. This book provides up-to-date information on the use of transgenic mouse models in the study of neurodegenerative disorders such as Alzheimer's and Huntington's disease. By reproducing some of the pathological aspects of the diseases, these studies could reveal the mechanism for their onset or development. Some of the transgenic mice can also be used as targets for testing new compounds with the potential to prevent or combat these disorders. The editors have extensive knowledge and experience in this field and the book is aimed at undergraduates, postgraduates and academics. The chapters cover disorders including: Alzheimer's disease, Parkinson's disease, Huntington's and other CAG diseases, amyotrophic lateral sclerosis (ALS), recessive ataxias, disease caused by prions, and ischemia.
Article
Full-text available
Alpha-crystallins were originally recognized as proteins contributing to the transparency of the mammalian eye lens. Subsequently, they have been found in many, but not all, members of the Archaea, Bacteria, and Eucarya. Most members of the diverse alpha-crystallin family have four common structural and functional features: (i) a small monomeric molecular mass between 12 and 43 kDa; (ii) the formation of large oligomeric complexes; (iii) the presence of a moderately conserved central region, the so-called alpha-crystallin domain; and (iv) molecular chaperone activity. Since alpha-crystallins are induced by a temperature upshift in many organisms, they are often referred to as small heat shock proteins (sHsps) or, more accurately, alpha-Hsps. Alpha-crystallins are integrated into a highly flexible and synergistic multichaperone network evolved to secure protein quality control in the cell. Their chaperone activity is limited to the binding of unfolding intermediates in order to protect them from irreversible aggregation. Productive release and refolding of captured proteins into the native state requires close cooperation with other cellular chaperones. In addition, alpha-Hsps seem to play an important role in membrane stabilization. The review compiles information on the abundance, sequence conservation, regulation, structure, and function of alpha-Hsps with an emphasis on the microbial members of this chaperone family.
Article
Nitro blue tetrazolium has been used to intercept O2⁻ generated enzymically or photochemically. The reduction of NBT by O2⁻ has been utilized as the basis of assays for superoxide dismutase, which exposes its presence by inhibiting the reduction of NBT. Superoxide dismutase could thus be assayed either in crude extracts or in purified protein fractions. The assays described are sensitive to ng/ml levels of super-oxide dismutase and were applicable in free solution or on polyacrylamide gels. The staining procedure for localizing superoxide dismutase on polyacrylamide electrophoretograms has been applied to extracts obtained from a variety of sources. E. coli has been found to contain two superoxide dismutases whereas bovine heart, brain, lung, and erthrocytes contain only one.
Article
Mutations in copper, zinc superoxide dismutase (SOD) have been implicated in the selective death of motor neurons in 2 percent of amyotrophic lateral sclerosis (ALS) patients. The loss of zinc from either wild-type or ALS-mutant SODs was sufficient to induce apoptosis in cultured motor neurons. Toxicity required that copper be bound to SOD and depended on endogenous production of nitric oxide. When replete with zinc, neither ALS-mutant nor wild-type copper, zinc SODs were toxic, and both protected motor neurons from trophic factor withdrawal. Thus, zinc-deficient SOD may participate in both sporadic and familial ALS by an oxidative mechanism involving nitric oxide.
Article
AMYOTROPHIC lateral sclerosis (ALS) is a degenerative disorder of motor neurons in the cortex, brainstem and spinal cord1,2. Its cause is unknown and it is uniformly fatal, typically within five years3. About 10% of cases are inherited as an autosomal dominant trait, with high penetrance after the sixth decade4,5. In most instances, sporadic and autosomal dominant familial ALS (FALS) are clinically similar4,6,7. We have previously shown that in some but not all FALS pedigrees the disease is linked to a genetic defect on chromosome 21q (refs 8, 9). Here we report tight genetic linkage between FALS and a gene that encodes a cytosolic, Cu/Zn-binding superoxide dismutase (SOD1), a homodimeric metalloenzyme that catalyzes the dismutation of the toxic superoxide anion O2.- to O2 and H2O2 (ref. 10). Given this linkage and the potential role of free radical toxicity in other neurodenegerative disorders11, we investigated SOD1 as a candidate gene in FALS. We identified 11 different SOD1 missense mutations in 13 different FALS families.
Article
Some human diseases selectively affect motor neurons of the spinal cord, brain stem and cerebral cortex. The best known of these is amyotrophic lateral sclerosis (ALS), a disease of unknown cause that poses a special challenge to neuroscientists because it involves principles of general biological importance, such as ‘selective vulnerability’ and degeneration of particular sets of neurons in the ageing brain. In the past few years there have been advances in understanding toxic, viral, genetic and immunological syndromes that involve motor neurons. Better comprehension of these disorders may give us clues about ALS itself.
Article
In the past five years, a great diversity of motor neuron diseases has emerged from new evidence of different etiologies. The theory of persistent viral infection has been revived by a case attributed to injections of human growth hormone, by cases of myelopathy attributed to HIV infections, and by the role of HTLV-1 in tropical spastic paraparesis. Dietary factors — calcium deprivation, ingestion of cycad, and lathyrism — seem to be important in different parts of the world. An immunological disorder is suggested in some cases by inordinate frequency of paraproteinemia. Finally, genetic forms are recognized in hexosaminidase deficiency, and the chromosomal locus of X-linked recessive motor neuron disease has been identified. The tasks now are to understand the pathogenesis of these (and other) different forms of motor neuron diseases, and to devise rational therapy.
Article
Mutations in copper, zinc superoxide dismutase (SOD) have been implicated in the selective death of motor neurons in 2 percent of amyotrophic lateral sclerosis (ALS) patients. The loss of zinc from either wild-type or ALS-mutant SODs was sufficient to induce apoptosis in cultured motor neurons. Toxicity required that copper be bound to SOD and depended on endogenous production of nitric oxide. When replete with zinc, neither ALS-mutant nor wild-type copper, zinc SODs were toxic, and both protected motor neurons from trophic factor withdrawal. Thus, zinc-deficient SOD may participate in both sporadic and familial ALS by an oxidative mechanism involving nitric oxide.
Article
Submitted to the Department of Biological Sciences. Copyright by the author. Thesis (Ph. D.)--Stanford University, 2003.
Article
Large differences are usually observed when standard staining methods for a number of pathological lesions in neurodegenerative disorders are compared. With the modified thioflavine S method presented here (easy and cheap to perform), the morphological appearance of the stained neurofibrillary tangles (NFT) and senile plaques (SP) is greatly improved. Furthermore, the intense contrast between stained lesions and background obtained with this technique permits an accurate automatic quantification of NFT and SP using a computer-assisted image analysis system.