ArticlePDF Available

First Evidence for Helical Transitions in Supercoiled DNA by Amyloid β Peptide (1-42) and Aluminum: A New Insight in Understanding Alzheimer's Disease

Authors:
  • Houston Methodist Hospital Research Institute
  • TA Sciences, Inc

Abstract

Previously, we evidenced a B --> Z helical change in Alzheimer's brain genomic DNA, leading to a hypothesis that Alzheimer's disease (AD) etiological factors such as aluminum (Al), amyloid beta (Abeta) peptide, and Tau might play a role in modulating DNA topology. In the present study, we investigated the interaction of Al and Abeta with DNA. Our results show that Abeta(1-42) could induce a B --> Psi (Psi) conformational change in pUC 18 supercoiled DNA (scDNA), Abeta(1-16) caused an altered B-form, whereas Al induced a complex B-C-A mixed conformation. Ethidium bromide binding and agarose gel electrophoresis studies revealed that Al uncoiled the DNAto a fully relaxed form, whereas Abeta(1-42) and Abeta(1-16) effected a partial uncoiling and also showed differential sensitivity toward chloroquine-induced topoisomer separation. Our findings show for the first time that Abeta and Al modulate both helicity and superhelicity in scDNA. A new hypothetical model explaining the potential toxicity of Abeta and Al in terms of their DNA binding properties leading to DNA conformational alteration is proposed.
First Evidence for Helical Transitions in Supercoiled DNA
by Amyloid βPeptide (1–42) and Aluminum
ANew Insight in Understanding Alzheimer’s Disease
Muralidhar L. Hegde,1Suram Anitha,1Kallur S. Latha,2
Mohammed S. Mustak,1Reuven Stein,3Rivka Ravid,4
and K. S. Jagannatha Rao*,1
1Department of Biochemistry and Nutrition, Central Food Technological Research Institute,
Mysore-570013, India; 2Centre for Human Genetics, Institute of Biotechnology, G-05,
Discoverer, ITPL, Whitefield Road, Bangalore, India; 3Department of Neurobiochemistry,
George S. Wise Faculty of Life Sciences, Tel-Aviv University, Ramat Aviv 69978, Israel;
and 4The Netherlands Brain Bank, The Netherlands
Received August 10, 2003; Accepted August 11, 2003
Abstract
Previously, we evidenced a B Z helical change in Alzheimer’s brain genomic DNA, leading to a hypoth-
esis that Alzheimer’s disease (AD) etiological factors such as aluminum (Al), amyloid β(Aβ) peptide, and Tau
might play a role in modulating DNA topology. In the present study, we investigated the interaction of Al and
Aβwith DNA. Our results show that Aβ(1–42) could induce a B →ψ(Psi) conformational change in pUC 18
supercoiled DNA (scDNA), Aβ(1–16) caused an altered B-form, whereas Al induced a complex B-C-A mixed
conformation. Ethidium bromide binding and agarose gel electrophoresis studies revealed that Al uncoiled the
DNA to a fully relaxed form, whereas Aβ(1–42) and Aβ(1–16) effected a partial uncoiling and also showed dif-
ferential sensitivity toward chloroquine-induced topoisomer separation. Our findings show for the first time
that Aβand Al modulate both helicity and superhelicity in scDNA. A new hypothetical model explaining the
potential toxicity of Aβand Al in terms of their DNAbinding properties leading to DNA conformational alter-
ation is proposed.
Index Entries: Alzheimer’s disease; pUC18 supercoiled DNA; helical transitions; topoisomer separation;
B-DNA; Z-DNA; ψ-DNA.
Journal of Molecular Neuroscience
Copyright © 2004 Humana Press Inc.
All rights of any nature whatsoever reserved.
ISSN0895-8696/04/22:19–31/$25.00
Journal of Molecular Neuroscience 19 Volume 22, 2004
Introduction
For the first time, we recently evidenced a left-
handed Z-DNA conformation in severely affected
Alzheimer’s disease (AD) brain hippocampal cells,
whereas normal hippocampal brain DNAexhibited
B-DNA conformation and moderately affected AD
brain DNA exhibited a B-Z intermediate conforma-
This paper is dedicated to the memory of the late Prof. M. A. Viswamitra, Indian Institute of Science, Bangalore, India.
M. L. Hegde and S. Anitha have contributed equally to this paper.
*Author to whom all correspondence and reprint requests should be addressed. E-mail: kjr4n@yahoo.com
20 Hegde et al.
Journal of Molecular Neuroscience Volume 22, 2004
tion (Anitha et al., 2002). Based on these findings,
we hypothesized that AD-associated molecules like
Al, amyloid β(Aβ) peptide, and Tau might play a
pivotal role in modulating DNA topology in AD
brain (Anitha et al., 2002). The puzzling question,
however, was how the nuclear localization of Aβ
might play a role in bringing about changes in DNA
topology. There were two previous reports on Aβ
immunoreactivity in the nuclear envelopes of P19
cells (Grant et al., 2000) and intranuclear accumula-
tion of Aβin AD brain (Gouras et al., 2000). In the
present study, we provided new evidence for Aβ
immunoreactivity in the vicinity of DNA in hip-
pocampal cells from AD brain. Nuclear localization
of Al has been well-established in AD brain (Crap-
per et al., 1980; Perl and Brody, 1980). Based on this
evidence for nuclear localization of Aβand Al in AD
brain, we hypothesized that Aβand Al might play
a role in modulating DNA topology and possibly
contribute to the B Z helical transition associated
with AD (Anitha et al., 2002). To investigate the above
hypothesis experimentally, we have studied the
interaction of Al, Aβ(1–42), and Aβ(1–16) with pUC
18 supercoiled DNA (scDNA). It is quite evocative
to study plasmid scDNA as a model system, in view
of the observation that a vast array of small scDNA
packets have been found to be present in animal and
human cells and are known to be involved in gene
expression (Bauer et al., 1980). These superhelical
packets are proposed to be analogous to plasmid
DNA supercoiling. Hence, the results can be corre-
lated or extended to human brain genomic DNA to
provide an insight into explaining the possible role
of Al and Aβin the progression of AD pathology
with reference to DNAtopology. Aβ, a hallmark fea-
ture of the senile plaque, is a proteolytic product of
the transmembrane β-amyloid precursor protein
(APP) (Balakrishnan et al., 1998). Al, Aβ(1–42) pep-
tide and its fragments, Aβ(1–16), and Aβ(1–28) are
reported to play a critical role in inducing the pathol-
ogy seen in AD (Kang et al., 1987; Savory et al., 1999).
The apparent role of Aβ, especially Aβ(1–42), is now
considered as a unifying pathological feature of
diverse forms of AD (Selkoe, 1996). The neurotoxi-
city of insoluble Aβaggregates has been widely
reported (Clements et al., 1996; Huang et al., 2000).
The present study attempts to investigate the poten-
tial role of Aβand Al in terms of DNA helical alter-
ation. Though the role of Al in AD is a highly
debatable topic, we investigated the effect of Al to
understand the biology of Al in terms of its effect on
DNA. The understanding of the complex biology of
Aβand Al with respect to DNA topological transi-
tion might provide an avenue to explore new insight
into AD pathology.
Materials and Methods
Materials
Three normal human brains (age 60–79 yr) (with
no history of long-term illness, dementia, or neuro-
logical disease), and six AD-affected (70–88 yr) brain
hippocampus samples were obtained from The
Netherlands Brain Bank (NBB). In normal and AD
brains, four each were male. Normal brains were
obtained from persons who died of either cardiac
arrest or accident. The average postmortem time on
average is 6 h. pUC 18 scDNA(cesium chloride-puri-
fied, 90% supercoiled structure) was purchased from
Bangalore Genei, India. Aβ(1–42), Aβ(1–16), agarose,
chloroquine, ethidium bromide (EtBr), and HEPES
were purchased from Sigma (USA). Anti-Aβ(1–42)
antibody was obtained from Chemicon International
(USA).
Al-Maltolate Synthesis
Al-maltolate (Al-maltol) was prepared in our lab-
oratory from aluminum chloride hexahydrate and
maltolate (3-hydroxy-2-methyl-4H-pyran-4-one),
following the method of Finneagan et al. (1986). In
the present study, Al-maltol was used to study DNA-
Al interactions. Al-maltol is highly soluble and
hydrolytically stable from pH 2.0 to 12.0. It over-
comes the hydrolytic speciation chemistry problem
of inorganic Al. This complex of Al can deliver a pre-
dicted amount of free aqueous Al at a physiological
pH (Martin, 1991). The thermodynamically pre-
dicted species that dominate at pH 7.0 for Al-maltol
are Al (mal)2+and Al (mal)22+.
Circular Dichroism (CD) Studies
scDNA was titrated against different concentra-
tions of Al-maltol (5 ×10–7 to 1 ×10–3 M), Aβ(1–42)
(0.8 ×10–7 to 0.8 ×10–4M), and Aβ(1–16) (0.8 ×10–7 to
0.8 ×10–4 M). CD spectra (200–330 nm) were recorded
for each concentration in 10–4 MHEPES buffer (pH
7.0). The CD signal of scDNA was also monitored in
the presence of chloroquine to examine the effect of
uncoiling the DNA on CD signals. The spectra at
each concentration represent an average of four
recordings. The CD spectra were recorded on a
JASCO J 715 Spectropolarimeter at 25°C, with 2-mm
cell length. DNAstructures were characterized using
Aβand Al Affect Supercoiled DNA 21
Journal of Molecular Neuroscience Volume 22, 2004
the references of Gray et al. (1978), Hanlon et al.
(1975), and Shin and Eichhorn (1984).
Agarose Gel Studies
DNA samples (scDNA, scDNA+Al-maltol,
scDNA+Aβ(1–42), and scDNA+Aβ(1–16) com-
plexes) were loaded on 1% agarose gels and elec-
trophoresed at 4 V/cm at room temperature. DNA
concentration loaded in all lanes was 1 µg. Topoiso-
mer separation of scDNA and complexes was car-
ried out in the presence of chloroquine. The samples
were stained with EtBr and photographed.
EtBr Binding Fluorescence Studies
EtBr binding patterns to scDNA, scDNA+Al
maltol, scDNA+Aβ(1–42), and scDNA+Aβ(1–16)
complexes were analyzed by mixing 1:1 (w/w)
DNA/EtBr before measuring fluorescence emission.
DNA/EtBr solutions were excited at 535 nm, and
emission was monitored at 600 nm using the Hitachi
F-2000 Fluorescence Spectrophotometer. The
amount of EtBr bound and the average number of
base pairs per bound EtBr molecule were calculated
using Scatchard’s equation.
Melting Profile
The melting profile curves (1°C/min, 25–95°C)
for scDNA, scDNA+Al maltol, scDNA+Aβ(1–42),
and scDNA+Aβ(1–16) complexes, in 10–4 MHEPES
buffer (pH 7.0) were recorded using the Gilford
Response II UV spectrophotometer with a thermo-
stat. The effect of chloroquine on melting profiles of
the above complexes has also been tested by treat-
ing the complexes with chloroquine (1–10 mg/mL
of sample). Melting temperature (Tm) values were
determined graphically from hyperchromicity
versus temperature plots. The precision in Tmvalues
estimated in triplicate was ± 0.5°C.
Clinical and Neuropathological Diagnostic
Criteria Applied in the Rapid Autopsy
System of the NBB
Human brain material is obtained via the rapid
autopsy system of the NBB, which supplies post-
mortem specimens from clinically well-documented
and neuropathologically confirmed cases. Autop-
sies are performed on donors from whom written
informed consent has been obtained either from the
donor or direct next of kin. The demented patients
are clinically assessed, and the diagnosis of “prob-
able Alzheimer’s disease” is based on exclusion of
other possible causes of dementia by history, phys-
ical examination, and laboratory tests.The clinical
diagnosis is performed according to NINCDS-
ADRDA criteria (McKhann et al., 1984), and the
severity of the dementia is estimated according to
the Global Deterioration Scale (Reisberg et al., 1982).
The control subjects have no known history or symp-
toms of neurological or psychiatric disorders.
Once the brain is removed, it is macroscopically
examined and immediately dissected following a
standard protocol. The specimens are frozen rapidly
in liquid nitrogen and stored at –80°C. All cases are
neuropathologically confirmed using conventional-
histopathological stains on formalin-fixed speci-
mens. The diagnosis is based on the presence and
distribution of the classical hallmarks for each of the
disease groups investigated. The NBB uses a scor-
ing system in which the density of senile plaques,
neurofibrillary tangles (NFTs), disrupted interneu-
ronal network (dINN), neuropil threads, congo-
phylic plaques, and vessels are estimated in Bodian
and Congo stains in four neocortical areas: frontal;
temporal; parietal; and occipital. For the staging of
the various pathological hallmarks, a combination
of a grading system and Braak staging is applied to
all specimens (Braak and Braak, 1991; Braak et al.,
1996; Mirra, 1997; Ravid et al., 1998).
In addition, ApoE allele frequency is determined
for each case as a possible risk factor for AD (Nielsen
et al., 2003). All cases in the disease groups and con-
trols are well matched for various factors, both ante-
mortem and postmortem. Antemortem factors
include age, sex, agonal state, seasonal alterations,
circadian variation, clock time of death, and med-
ication. Postmortem factors include postmortem
delay, fixation, and storage time and laterality (Ravid
et al., 1992, 1995; Ravid and Winblad, 1993). Sam-
ples are also controlled for quality by monitoring the
agonal state of the deceased prior to death (Ravid et
al., 1992, 1995).
Quantification of Apoptosis
in Hippocampal Neurons
Fifteen-micron brain sections were cut from the
hippocampal region of three normal and six AD-
affected human brains. Brain sections were first
stained for apoptosis using the fluorescence TUNEL
(terminal deoxynucleotidyl transferase [TdT]-
mediated dUTP nick-end labeling) technique
(Promega apoptosis detection kit, Madison, WI)
(Rao et al., 1998), costained for Aβ(1–42) using
22 Hegde et al.
Journal of Molecular Neuroscience Volume 22, 2004
monoclonal antibody, and developed by
3,3’-diaminobenzidine (DAB) reagent (Sigma,
Aldrich). Studies were also conducted for colocal-
ization of Aβin nonapoptotic cells. The apoptotic
detection system utilizes fluorescein, which mea-
sures the fragmented DNA of the apoptotic cells by
the catalytic incorporation of fluorescein-12-dUTP
at the 3’-OH DNA ends with the enzyme TdT. The
latter forms a polymeric tail by applying the TUNEL
assay. The fluorescein-12-dUTP-labeled DNAis then
visualized directly by fluorescence microscopy. Sec-
tions were photographed at 520 nm under ×400 mag-
nification using a fluorescence microscope. The slides
were scanned using Adobe Photoshop 3.0 and a
Nikon scanner (Nikon LS-3510 AF, 365-mm film scan-
ner, no. 6070192). The captured images were used
for quantification. The cells, which were stained with
flourescein (green), were counted as positive for
apoptosis based on the following criteria as reported
earlier by Rao et al. (1998): (1) DNA fragmentation
with no indication of chromatin margination or other
morphologic changes in the nucleus; (2) markedly
condensed nuclei with a nuclear diameter of 2.5 µm
or more; (3) swollen nuclei containing cytoplasmic
fragments of DNA; and (4) intracellular or extracel-
lular chromatin fragments. Aminimum of 18 fields
were photographed at an initial magnification of
×400 from hippocampal (CA1) regions of each brain
section. Six sections from each brain were evaluated.
Atotal area of 1 mm2for each section was evaluated.
The number of apoptotic cells was quantified by two
independent observers, and the interobserver vari-
ability approximated 2%. The number of apoptotic
cells was represented as cells/mm2.
Colocalization of Apoptosis
and AβImmunoreactivity
The hippocampus sections were first stained for
apoptosis using the TUNEL procedure, followed by
costaining for Aβimmunoreactivity, using anti-Aβ
monoclonal antibody, and reactivity was developed
using the ABC kit. In brief, the sections, prestained
for apoptosis, were incubated overnight at 4°C
with anti-Aβ(1–42) antibody. The bound immuno-
globulins were detected using the avidin-biotin-
peroxidase complex method (Vectastain ABC kit,
Vector Laboratories, Burlingame, CA) and visual-
ized by incubation with 0.05% DAB. The colocal-
ization of apoptosis and Aβin single hippocampal
neuron in CA1 region was imaged by a laser-fitted
confocal microscope.
Results
CD Studies
In an attempt to examine the potential role of Aβ
and Al in modulating DNAtopology with relevance
to the Z-DNA conformation found in AD brain, we
employed the CD method. scDNA was treated with
Al-maltol, and changes in DNA helicity were mon-
itored by analysis of CD spectra. As shown in Fig.1a,
CD spectra of scDNA showed B-DNAconformation
with a characteristic positive peak at 275 nm and
negative peak at 245 nm.Addition of increasing
concentrations of Al-maltol (from 5 ×10–7 to
1 ×10–3 M) to the DNAcaused a decrease in the pos-
itive band at 275 nm with no concomitant change in
the magnitude of the negative peak at 245 nm (Fig.
1b–d). In addition, there was a severalfold increase
in the magnitude of the negative peak at 210 nm.
This characteristic decrease in the positive peak at
275 nm, without change in the magnitude of the neg-
ative peak at 245 nm, reveals a C-DNA form; how-
ever, according to Maestre and Wang (1971), the
lowering of superhelical density of the scDNA also
involved a similar spectral change. To exclude this
possibility, we studied CD spectra of scDNA in the
presence of chloroquine (10 µg/mL), which uncoils
the DNA, like topoisomerase I. No change was
observed in the positive or negative peak on com-
plete uncoiling of the scDNA with chloroquine, and
the spectra were similar to that of control scDNACD
spectra (data not shown). It is clear from the above
experiments that the relaxation of supercoils is not
associated with decrease in positive peak intensity
at 275 nm; hence, we conclude that the changes
observed in the above spectra are not the result of a
simple uncoiling phenomenon but a conformational
change caused by Al-maltol. This conformation of
DNAbelongs to C family; however, the deep, narrow
negative peak at 210 nm is associated with A-DNA
(Gray et al., 1978). In view of this, we favor the inter-
pretation that the spectrum reflects the properties of
a B-C-A mixed conformation. A possible simultane-
ous helical transition of B C and B Acould have
been formed leading to a complex conformation.
Hanlon et al. (1975) have reported a similar pattern
of transition in calf thymus DNA.
Next we examined the effect of increasing con-
centrations of Aβ(1–42) on the ellipticity values of
scDNA. As shown in Fig. 2, substantial changes
of the CD spectra were observed upon interaction of
scDNA with Aβ(1–42) (0.8 ×10–7 to 0.8 ×10–4 M). At
Aβand Al Affect Supercoiled DNA 23
Journal of Molecular Neuroscience Volume 22, 2004
Aβ(1–42)/DNA molar ratios lower than 0.1, a DNA
secondary structural transition from the native B-
DNAto the C motif was observed. The spectral changes
involved decrease in a positive band at 275 nm with
no concomitant change in the magnitude of the neg-
ative peak at 245 nm (Fig. 2A). The spectral modifi-
cations of the positive peak were continuous with
the increasing Aβ(1–42) concentrations. The limit C-
DNA motif is characterized by a small positive CD
band at 275 nm and a long negative signal at 245 nm.
However, on the addition of a higher concentration
of Aβ(1–42) [Aβ(1–42)/DNA ratio > 0.1], the nega-
tive CD band extended in the nonabsorbing region
in the form of CD tails, with a large CD magnitude
compared to the intrinsic CD of scDNA, which is
intriguing. This spectral change presumably reflects
the asymmetric compaction of scDNA by Aβ(1–42)
to form ψ(+) DNA. This pattern of CD signal (Fig. 2B)
is a typical characteristic of the (+)ψform of DNA
(Shin and Eichhorn et al., 1984). All of the spectra
showed an isodichroic point at the cutting point of
215 nm. Examination of DNA ellipticities, as shown
in Fig. 2C,D, indicated the transformation of native
B-DNA to C-DNAto ψ-DNA in scDNA induced by
Aβ(1–42). Thus, Aβ(1–42) caused B C →ψ
conformation in scDNA.
Next we examined the effect of the different con-
centrations (0.8 ×10–7 to 0.8 ×10–4 M) of the shorter
Aβ(1–16) peptides on scDNA conformation. Higher
concentration (0.8 ×10–4 M) of Aβ(1–16) modi-
fied the scDNA to an unusual altered B-form
(Fig. 3). Addition of Aβ(1–16) caused a decrease in
the intensity of the positive peak centered around
275 nm, with a concomitant reduction of the nega-
tive peak around 245 nm. However, there was a sig-
nificant increase in the negative band intensity at
205 nm, indicating a modification in the usual B-sec-
ondary structure. The cutting point has been shifted
from 220 to 224 nm as the concentration of Aβ(1–16)
is increased. All of the spectra have two isodichroic
points at 228 and 252 nm.
Agarose Gel Electrophoresis Studies
These studies were conducted to monitor the
uncoiling process of scDNA. scDNA (1 µg) was
incubated with Al-maltol (1 ×10–3 M), Aβ(1–42)
(0.8 ×10–4 M), and Aβ(1–16) (0.8 ×10–4 M) and sub-
jected to agarose gel electrophoresis (Fig. 4).
Al-maltol uncoiled the scDNA nearer to the fully
relaxed form (lane B). Aβ(1–42) and Aβ(1–16) par-
tially uncoiled the DNA(lanes C and D, respectively).
To study the sensitivity of the complexes to chloro-
quine-induced topoisomer separation and to obtain
information on the stability of the complexes, gel elec-
trophoresis was carried out in the presence of chloro-
quine (Fig. 5). A fair degree of sensitivity was
Fig. 1. CD spectra of Al-maltol interaction with scDNA. Effect of Al-maltol binding on conformation of pUC 18 scDNA
in 10–4 MHEPES buffer (pH 7.0). scDNA was titrated against increasing concentrations of Al-maltol, and the changes in
molar ellipticities were monitored at 250°C in a 2-mm path length cell. The scDNA alone showed B-DNA conforma-
tion (a) On addition of increasing concentrations of Al-maltol (1 ×10–6, b; 1 ×10–4, c; 5 ×10–3, d MAl), the native
B-DNA conformation was transformed into a complex B-C-A mixed conformation.
24 Hegde et al.
Journal of Molecular Neuroscience Volume 22, 2004
observed toward chloroquine for DNA+Aβcom-
plexes (lanes C and D, respectively) compared to
DNA alone (lane A). The topoisomer separation was
observed for scDNA at 1 µg/mL of chloroquine,
whereas Aβ(1–42) and Aβ(1–16) DNA complexes
were relaxed to topoisomers at much lower amounts
of chloroquine (0.8 µg/mL). Chloroquine had no
effect on the DNA+Al complex, as Al-maltol had
totally relaxed the scDNA (lane B).
EtBr Binding Studies
The quantification of the uncoiling pattern of
scDNA induced by Al-maltol, Aβ(1–42), and
Aβ(1–16) was studied by measuring EtBr fluores-
cence intensity at 1:1 (w/w) DNA/EtBr concentra-
tions (Fig. 6). EtBr fluorescence of scDNAwas 13.74
(Fig. 7a), whereas it was 5.65 for the DNA+Al-maltol
complex (Fig. 6B). The significant decrease in the
fluorescence intensity in the case of the DNA+
Al-maltol complex might be the result of the relax-
ation of the scDNA. The fluorescence intensity for
the scDNA+Aβ(1–42) complex was 8.81 (Fig. 6C), and
this might be attributed to ψ-DNAconformation. The
fluorescence intensity for the scDNA+Aβ(1–16) com-
plex was 17.81 (Fig. 6D), and this might be the result
of the modified B-form of the scDNA. The Scatchard’s
plot analysis of the values of the EtBr binding to
the scDNA indicated that the average number of
base pairs per bound EtBr molecule was 6.85 for
scDNA, 86.6 for the scDNA+Al complex, 15.7 for
the scDNA+Aβ(1–42) complex, and 4.6 for the
scDNA+Aβ(1–16) complex.
Fig. 2. The interaction of Aβ(1–42) with scDNA. The changes in the ellipticities at 275 () and 220 nm () were
expressed as percentage of its value for control (native B-DNA) and plotted against the molar ratio of Aβ(1–42)/scDNA
(C,D). The native B-DNA was transformed to a limit C-motif at a Aβ(1–42)/scDNA molar ratio lower than 0.1 (A). At a
higher Aβ(1–42)/scDNA ratio (<0.1), the C-DNA was further converted into an asymmetrically condensed ψ-DNA (B).
Notably, the significant increase in the magnitude of negative CD signal (B,D) of scDNAat Aβ(1–42)/scDNA (<0.1) indi-
cates the gradual transformation of C-DNA to ψ-DNA.
Aβand Al Affect Supercoiled DNA 25
Journal of Molecular Neuroscience Volume 22, 2004
Fig. 3. The interaction of Aβ(1–16) with scDNA (25 ×10–6 g). Aβ(1–16) alters the B-DNA conformation of scDNA to
modified B-DNA. Experimental procedures for CD were as mentioned earlier. (a) scDNA; (b) 5 ×10 –7 MAβ(1–16); (c)
1 ×10–5 MAβ(1–16); (d) 0.8 ×10–4 MAβ(1–16).
Fig. 4. Effect of Al-maltol, Aβ(1–42), and Aβ(1–16) bind-
ing on scDNA mobility in 1% agarose gel. scDNA was incu-
bated in the absence and presence of Al-maltol and
Aβpeptides for 12 h and then was separated by 1% agarose
gel electrophoresis. scDNA alone (cesium chloride
purified) showed >90% superhelicity (lane A). Al-maltol
(5 ×10– 3 M) uncoiled the scDNA nearer to fully relaxed
form (lane B), whereas Aβ(1–42) (0.8 ×10–4 M) and Aβ(1–16)
(0.8 ×10–4 M) partially uncoiled the scDNA (lanes C,D).
DNA concentration loaded in all lanes was 1 µg, and elec-
trophoresis was carried out at 4 V/cm at room temperature.
The samples were stained with EtBr.
Figure 5. Sensitivity for chloroquine-induced topoiso-
mer separation. scDNA was incubated with chloroquine in
the absence and presence of Al-maltol and Aβpeptides.
The effective topoisomer separation was observed for scDNA
at 1 µg/mL chloroquine (lane A), whereas Aβ(1–42) and
Aβ(1–16) DNA complexes were relaxed to topoisomers at
much lower chloroquine levels (0.8 µg/mL of sample volume)
(lanes C,D). Chloroquine had no effect on the DNA-Al com-
plex (lane B).
26 Hegde et al.
Journal of Molecular Neuroscience Volume 22, 2004
Melting Profile
The melting temperature study provides insight
on the stability of DNA. Thus, in the next set of exper-
iments, we examined the melting profile of scDNA
in the absence or presence of Al-maltol or Aβ. The
Tmvalue for scDNA was 59.50°C (±0.5), and its melt-
ing profile revealed an unusual monophasic pattern
(Fig. 7a). Al-maltol (at concentrations 10–5 to 10–4 M)
enhanced the Tmto 66.70°C (±0.5) and at higher
(5 ×10 –3 M) concentration did not alter the Tm
(Fig. 7b). These kinds of changes in Tmare typical
for B Ahelical transitions (Champion et al., 1998).
The Tmfor the scDNA+Aβ(1–42) complex showed a
typical biphasic melting pattern with two Tmvalues
(59°C and 88°C) (Fig. 7c). This is the first report study-
ing the melting pattern of ψ(+) DNA induced by
Aβ(1–42). Aβ(1–16) did not alter the Tmat lower con-
centrations (10–6 M), whereas it enhanced the Tmto
71.4°C (±0.5) at higher (10–4 M) concentrations. The
scDNA+Aβ(1–16) complex showed a customary
monophasic melting profile (Fig. 7d). To further
explore the nature of the unusual monophasic melt-
ing profile of scDNA and to see the effect of uncoil-
ing the scDNA on its Tm, we studied the melting
pattern of the scDNA in the presence of chloro-
quine. There was no change in the Tm(59.5°C ±0.5)
in scDNA treated with chloroquine at 1 µg/mL
(topoisomer separation concentration as confirmed
by agarose gel) and 10 µg/mL (complete relaxation
of superhelicity to linear DNA) concentration (data
not shown). These results indicate that the unusual
monophasic melting profile observed in the case of
scDNA might not be the result of supercoiling,
because the same melting pattern was obtained with
linearized scDNA as well.
Gallium Chloride Does Not Affect
scDNA Conformation
To examine whether the effect of Al on scDNA
conformation is specific or a more general effect of
metals that have a similar ionic radius and belong
to the same group (group 13), we studied the effect
of gallium, another group 13 metal having essen-
tially a similar chemical behavior in aqueous solu-
tion as Al. The ionic radii of Al and gallium are
0.57 Å and 0.62 Å, respectively. Thus, we compared
the effect of gallium chloride on scDNA to those of
Al-maltol. These experiments were carried out in a
similar manner (methods used and concentration
range) as those described above for Al-maltol. The
results showed that in contrast to Al-maltol, gallium
chloride did not alter either the conformation or the
superhelicity of scDNA. In addition, gallium chlo-
ride did not change the Tmof the scDNA (data not
shown). At higher concentrations, gallium caused
aggregation of scDNAas observed by the zeroing or
flattening of the CD signal. Taken together, these
results suggest that the effect of Al on DNAconfor-
mation does not represent a general effect of group
13 metals but rather is a specific feature of Al.
Localization of Aβin the Nuclei of CA1 Neurons
in Hippocampal Region of Brain
In previous studies we showed that hippocampal
DNA obtained from AD brain samples has a prefer-
entially left-handed Z-DNA conformation (Anitha
et al., 2002), and our present results show that Aβ
induces similar (?) conformational changes in vitro.
Moreover, it was shown that Aβinduces apoptosis
in cell lines and rabbit brains (Clements et al., 1996;
Selkoe, 1996; Huang et al., 2000). Thus, we hypoth-
esize that Aβis presented in the neuronal nuclei of
affected AD brain areas, where it might induce DNA
conformational changes. These changes might in
turn lead to apoptotic neuronal death. In an attempt
to examine whether Aβ(1–42) is present in neuronal
nuclei in AD brains and explore its relationship to
apoptosis, the presence of Aβ(1–42) immunoreac-
tivity and apoptotic nuclei (TUNEL positive) was
examined in neurons of the hippocampal CA1 region
Fig. 6. EtBr binding pattern. Equal concentrations (w/w)
of DNA and EtBr were used as models to study the effect of
EtBr intensity on Al-maltol and Aβpeptides. EtBr intensity
also provides information on DNA conformation. The
uncoiling of scDNA was quantified by measuring the EtBr
fluorescence intensity of 1:1 (w/w) DNA/EtBr solutions. The
solutions were excited at 535 nm and emission monitored
at 600 nm. The emission intensity values for scDNA were
13.74 (A); scDNA+Al, 5.65 (B); scDNA+Aβ(1–42), 8.81 (C);
and scDNA+Aβ(1–16), 17.81 (D).
Aβand Al Affect Supercoiled DNA 27
Journal of Molecular Neuroscience Volume 22, 2004
of normal and AD brains. The number of Aβand/or
TUNEL-positive cells counted per square millime-
ter of brain section is given in Table 1. Quantifica-
tion of data revealed that in normal brain
hippocampal sections, 25% of the cells were apop-
totic, whereas in the case of AD brain, about 75% of
the cells were apoptotic. When normal brain hip-
pocampal sections were tested for colocalization of
apoptosis (TUNEL positive) and Aβimmunoreac-
tivity, we could not find a TUNEL-positive cell that
was also Aβpositive. But in the AD brain sections,
out of 100 TUNEL-positive cells counted, 50% of cells
were found to be positive also for Aβ. These results
suggest that Aβis deposited in the vicinity of DNA
in the nuclear region of AD cells. In the case of normal
brain sections, Aβ(1–42) immunoreactivity was not
observed in either apoptotic or nonapoptotic hip-
pocampal neurons. Figure 8A presents a confocal
image of a single representative neuron, showing
absence of Aβdeposition in an apoptotic nucleus
from the hippocampal section of normal brain. Figure
8B shows the localization of Aβimmunoreactivity
in the apoptotic nucleus of hippocampal neuron in
the CA1 region.
Discussion
AD is associated with several complex neu-
ropathological events like deposition of Aβ, abnor-
mal phosphorylation of Tau, aggregation of these
proteins into NFTs, oxidative stress, and DNA
damage (Iqbal et al., 1994; Lyras et al., 1997). It is of
interest to mention that Aβ(Gouras et al., 2000; Grant
et al., 2000; present study) and Al (Crapper et al.,
1980; Perl and Brody, 1980) were found to be local-
ized in the chromatin region of nuclei. Recently, our
team first evidenced the presence of left-handed,
rigid Z-DNA in severely affected AD brain and a
B-Z intermediate DNA conformation in moderate
AD. In contrast, normal young and aged brains have
usual/right-handed Watson-Crick DNA conforma-
tion (Anitha et al., 2002). It has also been hypothe-
sized that the prime etiological factors of AD, such
as Aβ, Tau, and Al (a highly debatable etiological
factor), might be playing a role in right- to left-handed
helical change associated with AD. Based on this
DNA conformational change, we hypothesized an
explanation for the unusual phenomenon-like nucle-
osome misassembly, G*-specific DNA oxidation,
Fig. 7. Effect of Al and Aβon melting profile of scDNA. The UV absorbance at 260 nm was recorded for scDNA and
scDNA complexes with Al-maltol and Aβpeptides at different temperatures. The melting curves (10°C/min, 25–95°C)
for scDNA and other complexes in 10–4 MHEPES, pH 7.0, were recorded in a UV spectrophotometer with a thermo-
stat. Tmvalues were determined from hyperchromicity vs temperature plots. Tmfor scDNA was 59.5°C (a).
Al-maltol initially enhanced the Tm(not shown in the graph) but did not alter at higher concentrations (b). scDNA+Aβ(1–42)
showed a biphasic melting pattern (c); Aβ(1–16) enhanced the Tmof scDNA (d). The results are representative of three
independent experiments.
28 Hegde et al.
Journal of Molecular Neuroscience Volume 22, 2004
terminal differentiation, and altered gene expression
associated with AD brain (Anitha et al., 2001, 2002).
There were few reports on the binding ability of Al
to DNA (Karlik et al., 1980; Rao et al., 1993; Rao and
Divaker, 1993); and our team first evidenced that Al
could strongly bind to AT*-specific oligomers (Rajan
et al., 1996), whereas Al could bring structural tran-
sition from Z-Aconformation in GC*-rich oligomers
(Champion et al., 1998). However, there is no report
to date on Aβbinding to DNA. In the present study,
we first evidenced the interaction of Aβand Al with
scDNA. We found that Aβ(1–42) not only binds to
scDNA but also is able to alter the conformation of
DNA. An initial BC transition was observed, which
gradually transformed into ψ-DNA, presumably
reflecting a partial DNA collapse into a ψ-phase
(Zuidam et al., 1999). In ψ-DNA, the DNA molecules
are tightly packed into a toroidal superhelical bundle
whose chiral sense is defined by intrinsic DNAhand-
edness. Specifically, right-handed secondary con-
formations, such as B and C motifs, stabilize a
left-handed tertiary conformation (Reich et al., 1994).
Such a tightly packed left-handed DNA organiza-
tion exhibits negative CD signals whose magnitude
is larger than that characterizing dispersed DNA
molecules, which lack a tertiary structure (Fig. 3b).
The ψ-DNA conformation induced by Aβ(1–42) is
structurally closer to Z-DNA, which was observed
in severely affected AD brain (Anitha et al., 2002).
Studies by Thomas and Thomas (1989) showed
clearly that ψ-DNA, an ordered, twisted, tight-pack-
ing arrangement of the double helix, is structurally
and immunologically closely related to the Z-DNA
family. It is left-handed in conformation like Z-DNA.
This evidently indicates that DNA topological
changes induced by Aβare similar to the changes
seen in AD brain DNA. Al, another strong and highly
debatable etiological factor, could play a key role in
AD pathology by contributing to complex DNAheli-
cal transitions (B A; B C, or B-C-A) of scDNA,
and this complex DNAconformation is energetically
weak and is likely to go into Z-DNA conformation
as reported in AD brain (Anitha et al., 2002). Vast
arrays of small scDNA packets have been found to
be present in animal and human cells and are known
to be involved in gene expression (Bauer et al., 1980).
Al at a physiological concentration (10–6 M) was
reported to irreversibly unwind the scDNA (Rao et
Table 1
Number of Apoptotic and Aβ-Positive Cells in Normal and AD-Affected Human Brain Samples
Normal Ad
Brain region TUNEL AβTUNEL + AβTUNEL AβTUNEL + Aβ
Hippocampus 15 ± 1.0 0 0 100 ± 5.0 75 ± 2.0 50 ±4.0
TUNEL and Aβ-positive cells in hippocampal sections were counted microscopically in square millimeters. The results
presented were obtained from sections of each of three brains for normal and six sections of each of six brains for AD. Values
are expressed as mean ± S.D.
Fig. 8. Localization of Aβin human normal and AD brain samples. Frozen sections of 15 µm were cut from hip-
pocampal brain regions and stained for apoptosis using the fluorescence TUNEL technique, costained for Aβ(1–42) with
a monoclonal antibody, and developed with a DAB reagent. Colocalization of apoptosis and Aβimmunoreactivity in
the vicinity of DNA in a nucleus of hippocampal neuron in the CA1 region was imaged by confocal microscopy. (A)
Normal brain apoptotic nuclei showing absence of Aβimmunoreactivity. (B) AD brain apoptotic nuclei from
hippocampal region showing Aβimmunoreactivity. The indicates Aβimmunoreactivity.
Aβand Al Affect Supercoiled DNA 29
Journal of Molecular Neuroscience Volume 22, 2004
al., 1993). Our study also shows that Al uncoils the
scDNA to a fully relaxed form, whereas Aβ(1–42)
and Aβ(1–16) relax the DNApartially. Thus, the pre-
sent study hearsay for the first time that Al and Aβ
uncoil scDNA besides bringing about helicity
changes. Another interesting feature observed is the
sensitivity of DNA+Aβ(1–42) and DNA+ Aβ(1–16)
complexes to chloroquine, a drug that mimics topoi-
somerase I in action. This observed sensitivity indi-
cates a possible alteration in DNA replication and
gene expression in the cells.
The possible complex role of Al and Aβin mod-
ulating DNAhelicity with relevance to AD is hypoth-
esized in Fig. 9. We propose that neuropathological
factors like Al and Aβmight modulate DNA topol-
ogy in AD brain. In stage I, the complex conforma-
tional changes (ψ-DNA, B A, B C, or B-C-A,
altered B) observed experimentally in scDNA,
induced by Al, Aβ(1–42), and Aβ(1–16), are pre-
sumably the early events in AD pathology. In mod-
erately affected AD brain, the DNA has a B-Z
intermediary conformation and other intermediary
complex conformations might exist (Anitha et al.,
2001, 2002). In stage II, secondary factors such as
oxidative stress, cell shrinkage, ionic imbalance, and
polyamines are likely to play a role in converting
these intermediary complex conformations to rigid,
left-handed Z-DNA. Accordingly, we reported the
presence of left-handed Z-DNAin severely affected
AD brain (Anitha et al., 2002). We propose that DNA
topological changes play a role in AD progression.
These novel observations and further work in this
direction may have important implications for aiding
in our understanding of toxicity of Aβand Al in terms
of their direct role in altering DNAconformation and
its relevance in neurodegeneration occurring in AD.
Acknowledgments
The authors wish to thank Dr. V. Prakash, Direc-
tor, CFTRI (Mysore) for all of his support and encour-
agement. The authors also thank Prof. S. Sharath
Chandra, Centre for Human Genetics (Bangalore)
for his support. M. L. H. thanks the Council for Sci-
entific and Industrial Research (CSIR) for awarding
the Junior Research Fellowship; S. A., K. S. L., and
Fig. 9. Hypothesis. Possible role of Al, Aβ(1–42), and Aβ(1–16) in modulating scDNA topology with relevance to AD.
30 Hegde et al.
Journal of Molecular Neuroscience Volume 22, 2004
M. S. M. are grateful to CSIR for awarding the Senior
Research Fellowships. This work was supported by
a grant from the Ministry of Science, Culture, and
Sports (Israel) and the Department of Biotechnology
(India). We profoundly thank Prof. Surolia, Indian
Institute of Science, Bangalore, India, for allowing
us to use CD facility and The Netherlands Brain Bank
for providing the human brain specimens.
References
Anitha S., Rao K. S. J., Latha K. S., and Viswamitra M. A.
(2002) First evidence to show the topological change
of DNA from B-DNA to Z-DNA conformation in the
hippocampus of Alzheimer’s brain. J. Neuromol. Med.
2, 287–295.
Anitha S., Reuvin S., Rao K. S. J., Latha K. S., and Viswami-
tra M. A. (2001) Alink between apoptotic DNA damage
and DNA topology in Alzheimer’s disease brain: a
hypothesis. Alz. Rep. 4, 121–131.
Balakrishnan R. Parthasarathy R., and Sulkowski E. (1998)
Alzheimer’s β-amyloid peptide: affinity for metal chela-
tors. J. Peptide Res. 51,91–95.
Bauer W. R. Crick F. H. C., and White H. (1980) Super-
coiled DNA. Sci. Am. 16, 243–245.
Braak H. and Braak E. (1991) Neuropathological stageing
of Alzheimer’s-related changes. Acta Neuropathol. 82,
239–259.
Braak H. Braak E., Bohl J., and Reintjes R. (1996) Age, neu-
rofibrillary changes, A beta-amyloid and the onset of
Alzheimer’s disease. Neurosci. Lett. 210, 87–90.
Champion C. S., Kumar D., Rajan M. T., Rao K. S. J.,
and Viswamitra M. A. (1998) Interaction of Co, Mn,
Mg, and Al with d(GCCCATGGC) and d(CCGGGCC
CGG): a spectroscopic study. Cell. Mol. Life Sci. 54, 488–496.
Clements A., Allsop D., Walsh D. M., and Williams C. H.
(1996) Aggregation and metal-binding properties of
mutant forms of the amyloid beta peptide of
Alzheimer’s disease. J. Neurochem. 66, 740–747.
Crapper D. R., Quittkat S., Krisnan S. S., Dalton A. J., and
DeBoni U. (1980) Intranuclear aluminum content in
Alzheimer’s disease, dialysis encephalopathy, and
experimental aluminum encephalopathy. Acta Neu-
ropathol. 50, 19–24.
Finneagan M. M., Retigg S. J., and Orvig C. (1986) Aneu-
tral water soluble aluminum complex of neurological
interest. J. Am. Chem. Soc. 108, 5033–5035.
Gouras G. K., Tsai J., Naslund J., Vincent B., Edgar M.,
Checher F., et al. (2000) Intraneuronal Aβ42 accumu-
lation in human brain. Am. J. Pathol. 156, 15–20.
Grant S. M., Ducatenzeiler A., Szyl M., and Cuello A. C.
(2000) Aβimmunoreactive material is present in all
intracellular compartments in transfected neuronally
differentiated, P19 expressing the human amyloid
β-protein precursor. J. Alzheimer’s Dis. 2, 207–222.
Gray D. M., Taylor T. N., and Lang D. (1978) Dehydrated
circular DNA: circular dichroism of molecules in
ethanol solutions. Biopolymers 17, 1–45.
Hanlon S., Brudno S., Wu T. T., and Wolf B. (1975) Struc-
tural transitions of deoxyribonucleic acid in aqueous
electrolyte solutions: reference spectra of conforma-
tional limits. Biochemistry 14, 1648–1660.
Huang H. M., Ou H. C., and Hsieh S. J. (2000) Antioxi-
dants prevent amyloid peptide-induced apoptosis and
alteration of calcium homeostasis in cultured cortical
neurons. Life Sci. 66, 1879–1892.
Iqbal K., Alonso A. C., Gong C. X., Khatoon S., Singh T. J.,
and Grundke-Iqbal I. (1994) Mechanism of neuro-
fibrillary degeneration in Alzheimer’s disease. Mol.
Neurobiol. 9, 119–123.
Kang J., Lemaire H. G., Unterbeck A., Salbaum J. M., Mas-
ters C. L., Grzeschik K. H., et al. (1987) The precursor
of Alzheimer’s disease amyloid A4 protein resembles
a cell-surface receptor. Nature 325, 733–736.
Karlik S. J., Eichhorn G. L., Lewis P. N., and Crapper D.
R. (1980) Interaction of aluminum species with deoxyri-
bonucleic acid. Biochemistry 19, 5991–6001.
Lyras L., Cairns N. J., Jenner A., Jenner P., and Halliwell
B. (1997) An assessment of oxidative damage to pro-
teins, lipids, and DNA in brain from patients with
Alzheimer’s disease. J. Neurochem. 68, 2061-2069.
Maestre M. F., and Wang J. C. (1971) Circular dichroism
of superhelical DNA. Biopolymers 10, 1021.
Martin R. B. (1991) Aluminum in biological systems, in
Aluminum in Chemistry, Biology, and Medicine, Nicolini,
P., Zatta, P. F., and Corain, B. eds., Raven, New York,
pp. 3–20.
McKhann G., Drachman D., Folstein M., Katzman R., Price
D., and Stadlan E. (1984) Clinical diagnosis of
Alzheimer’s disease: report of the NINCDS-ADRDA
work group under the auspices of Department of
Health and Human Services Task Force on Alzheimer’s
disease. Neurology 34, 939–944.
Mirra S. S. (1997) The CERAD neuropathology protocol
and consensus recommendations for the postmortem
diagnosis of Alzheimer’s disease: a commentary. Neu-
robiol. Aging 18 (Suppl.), S91–94.
Nielsen S. A., Ravid R., Kamphorst W., and Orgensen
O. S. (2003) Apoliprotein E e4 in an autopsy series of
various dementing disorders. J. Alzheimer’s Dis. 5,
119–125.
Perl D. P., and Brody A. R. (1980) Alzheimer’s disease:
X-ray spectrometric evidence of aluminium accumu-
lation in neuro-fibrillary tangle bearing neurons. Sci-
ence 208, 297.
Rajan M. T., Champion C. S., Kumar D., Vishnuvardhan
D., Rao K. S. J., and Viswamitra M. A. (1996) Interac-
tion of Co, Mn, Mg and Al with d(GCGTACGC): a spec-
troscopic study. Mol. Biol. Rep. 22, 47–52.
Rao K. S. J. and Divaker S. (1993) Spectroscopic studies of
Al-DNA interactions Bull Environ. Contam. Toxicol. 50,
92–99.
Rao K. S. J., Anitha S., and Latha K. S. (2000) Aluminium
induced neurodegeneration in hippocampus of aged
rabbits mimics Alzheimer’s disease. Alz. Rep. 3, 83–88.
Rao K. S. J., Letada P., Haverstick D. M., Herman M. M.,
and Savory J. (1998) Modifications to the in situ TUNEL
Aβand Al Affect Supercoiled DNA 31
Journal of Molecular Neuroscience Volume 22, 2004
method for detection of apoptosis in paraffin-embedded
tissue section. Ann. Clin. Lab. Sci. 28, 131–137.
Rao K. S. J., Rao B. S., Vishnuvardhan D., and Prasad K.
V. S. (1993) Alteration in superhelical state of DNAby
aluminium. Biochem. Biophys. Acta 1172, 17–20.
Ravid R. and Winblad B. (1993) Brain banking in
Alzheimer’s disease: factors to match for, pitfalls and
potentials, in Alzheimer’s Disease: Advances in Clinical
and Basic Research, Corain, B., Iqbal, K., Nicolini, M.,
Winblad, B., Wisniewsky, H., and Zatta, P., eds., John
Wiley and Sons, NY, pp. 213–218.
Ravid R., Van Zwieten E. J., and. Swaab D. F. (1992) Brain
banking and the human hypothalamus—factors to match
for, pitfalls and potentials. Prog. Brain Res. 93, 83–95.
Ravid R., Swaab D. F., Van Zwieten E. J., and Salehi A.
(1995) Controls are what makes a brain bank go round,
in Neuropathological Diagnostic Criteria for Brain Bank-
ing, Biomedical and Health Research, Vol. 10, Cruz-
Sanchez, F. F. Cuzner , M. L., and Ravid, R., eds., IOS
Press, Amsterdam, The Netherlands, pp. 4–13.
Ravid R., Swaab D. F., Kamphorst W., and Salehi A. (1998)
Brain banking in aging and dementia research—the
Amsterdam experience, in Progress in Alzheimer’s and
Parkinson’s disease, Fisher, A., Yorshida, M., and Hanin,
I., eds., Plenum Press, New York, pp. 277–286.
Reich Z., Levin Z. S., Gutman S. B., Arad T., and Minsky
A. (1994) Supercoiling-regulated liquid-crystalline
packaging of topologically-constrained, nucleodsome-
free DNA molecules. Biochemistry 33, 14177–14184.
Reisberg B., Ferris S. H., Del Leon M. J., and Crook T.
(1982) The global deterioration scale for assessment of
primary degenerative dementia. Am. J. Psychol. 139,
1136–1139.
Savory J., Rao K. S. J., Huang Y., Letada P. R., and Herman
M. M. (1999) Age related changes in Bcl-2 and Bax ratio,
oxidative stress, redox active iron, and apoptosis asso-
ciated with aluminium induced neurodegeneration:
increased susceptibility with aging, Neurotoxicology 20,
805–818.
Selkoe D. J. (1996) Amyloid beta-protein and the genetics
of Alzheimer’s disease J. Biol. Chem. 271, 18295–18298.
Shin Y. A. and Eichhorn G. L. (1984) Formation of psi(+)
and psi (–) DNA. Biopolymers 23, 325–335.
Thomas T. J. and Thomas T. (1989) Direct evidence for the
presence of left handed conformation in a supramole-
cular assembly of polypeptides. Nucleic Acids Res. 17,
3795–3810.
Zuidam N. J., Barenholz Y., and Minsky A. (1999) Chiral
DNApackaging in DNA-cationic liposome assemblies.
FEBS Lett. 457, 419–422.
... As evident from the results (FIG. 4A-E & TABLE 1), the amino acid residues of Aβ- (1)(2)(3)(4)(5)(6)(7)(8)(9)(10)(11)(12)(13)(14)(15)(16) namely, HIS6, ASP7, HIS13, LYS16 interact with D chain CYT, GUA, and THY of DNA D chain and form 6 hydrogen bonds across the major and minor groove. Most of the docked poses shown in the figure (FIG. ...
... DNA stability and conformation are important in the life cycle of an organism and DNA instability and changes in conformation can affect various reactions including replication, transcription, epigenetic modifications, recombination, and repair. This is postulated to be one of the risk factors for neuronal death in neurodegenerative diseases[14][15][16]. Copper is a physiologically important metal and micronutrient required for several biological functions and plays a role in the central nervous system development. ...
Article
Full-text available
Copper is an essential metal for life that plays a key role in CNS development and acts as a cofactor for numerous enzymes. Low concentrations of copper cause an incomplete development, whereas an excess, is injurious. Redox reactions are the basis of copper toxicity. Copper directly binds to DNA and proteins alter the structures and stability of proteins and DNA. Interactions between Aβ peptides and copper ions may play a role in the promotion of toxic aggregation and produce ROS that may link to Alzheimer's disease. DNA conformation and stability are important in the life cycle of an organism. Any change in the conformation and stability alters the gene expression and is one of the risk factors for neuronal death in neurodegenerative disorders. In the present study, we have attempted to study the binding of copper and Aβ1-16 with the DNA sequence through the Circular dichroism and these results were further supported by the UV-VIS absorbance, fluorescence studies, and in-silico Molecular docking studies
... In fact, Aβ neuronal inclusions have prion-like properties (Olsson et al., 2018) and are able to enter the nucleus through nuclear pores complex across the nuclear envelope, but only Aβ 42 plays a role in gene transcription (Barucker et al., 2014). It has been suggested that Aβ is deposited in the vicinity of DNA in the nuclear region of AD cells (Hegde et al., 2004). Intranuclear neuronal Aβ-labeling was observed in two cases in this study: an elderly S. frontalis and a subadult Z. cavirostris. ...
... While it was initially thought that nuclear amyloid is responsible for neural cell death, time-resolved experiments that correlate nuclear amyloid and neurodegeneration on the single cell level also suggest a cell protective role (von Mikecz, 2014). Even though the mechanism is still unknown and there is no confirmation to date on Aβ binding to DNA, the nuclear localization of Aβ might play a role in bringing about changes in DNA topology, modulating both helicity and superhelicity in super coiled DNA (Hegde et al., 2004). Although the role of nuclear amyloid is still unknown, the Aβ intranuclear expression provides new insights regarding cerebral safeguards for hypoxic-ischemic brain injury from accidents or disease. ...
Article
Full-text available
Hypoxia could be a possible risk factor for neurodegenerative alterations in cetaceans’ brain. Among toothed whales, the beaked whales are particularly cryptic and routinely dive deeper than 1,000 m for about one hour in order to hunt squids and fishes. Samples of frontal cerebral and cerebellar cortex were collected from nine animals, representing six different species of the suborder Odontoceti. Immunohistochemical analysis employed anti-β-amyloid (Aβ) and anti-neurofibrillary tangle (NFT) antibodies. Six of nine (67%) animals showed positive immunolabeling for Aβ and/or NFT. The most striking findings were intranuclear Aβ immunopositivity in cerebral cortical neurons and NFT immunopositivity in cerebellar Purkinje neurons with granulovacuolar degeneration. Aβ plaques were also observed in one elderly animal. Herein, we present immunohistopathological findings classic of Alzheimer's and other neurodegenerative diseases in humans. Our findings could be linked to hypoxic phenomena, as they were more extensive in beaked whales. Despite their adaptations, cetaceans could be vulnerable to sustained and repetitive brain hypoxia.
... Alternatively, they also force intercalate onto the preformed Z-DNA, thereby enabling the Z-to-B transition. The list includes amyloid beta aggregates (Geng et al. 2010), aluminum (Hegde et al. 2004), the popular antimalarial drug chloroquine (Kwakye-Berko and Meshnick 1990), and popular DNA intercalating stain ethidium bromide (EtBr) that is commonly used as a drug for the treatment and prophylaxis of trypanosomiasis in cattle (Fuertes et al. 2006), the antitumor antibiotic elsamicin A, the tetra peptide KWGK peptide (Fuertes et al. 2006), the anticancer compound NC-182 (Fuertes et al. 2006), netropsin and distamycin (Zimmer et al. 1983), histones H1 and H5 (Russell et al. 1983), and a guanine base forming 7-deazaguanine (7daG) with a substitution of N7 atom (Seela and Driller 1989). Although several anticancer drugs have shown a B-DNA or Z-DNA preference, there is no clear evidence supporting the correlation between anticancer activity and B-DNA or Z-DNA binding preference. ...
... However, there is a significant disadvantage to the use of nanoporous alumina materials in medical applications. Although aluminum is a constituent of several medical alloys (e.g., Ti-6Al-4V, ASTM F136), it is currently unknown whether aluminum is a biocompatible material [65]. Nanoporous alumina membrane coated with titanium oxide using atomic layer deposition has proved compatible for biomedical applications [66]. ...
Article
Full-text available
Infectious diseases and nosocomial infections may play a significant role in healthcare issues associated with biomedical materials and devices. Many current polymer materials employed are inadequate for resisting microbial growth. The increase in microbial antibiotic resistance is also a factor in problematic biomedical implants. In this work, the difficulty in diagnosing biomedical device-related infections is reviewed and how this leads to an increase in microbial antibiotic resistance. A conceptualization of device-related infection pathogenesis and current and future treatments is made. Within this conceptualization, we focus specifically on biofilm formation and the role of host immune and antimicrobial therapies. Using this framework, we describe how current and developing preventative strategies target infectious disease. In light of the significant increase in antimicrobial resistance, we also emphasize the need for parallel development of improved treatment strategies. We also review potential production methods for manufacturing specific nanostructured materials with antimicrobial functionality for implantable devices. Specific examples of both preventative and novel treatments and how they align with the improved care with biomedical devices are described.
... [13][14][15] Two such well-documented detrimental effects of Al accumulation would be Parkinson's disease and Alzheimer's disease, for which to date the best solution relies on the philosophical remedy of "prevention is better than cure". [16][17][18][19] Naturally as preventive measures, precise sensing and quantitative estimation of Al remain the rst and foremost step in the ght against Al-related diseases. It is noteworthy here that the permissible Al consumption limit set by the World Health Organization (WHO) is 7 mg per week per kg body weight, while that of the European Food Safety Authority (EFSA) is 1 mg per week per kg body weight. ...
Article
Full-text available
We report herein the development of a new pyridine-pyrazole based bis-bidentate asymmetric chemosensor that shows excellent turn-on chelation-enhanced Al3+-responsive fluorescence. The presence of two 'hard' phenolic hydroxyl groups plays a pivotal role in switching-on the sensing through coordination to the 'hard' Al3+ ion, while the mechanism can be interpreted by the chelation-enhanced fluorescence (CHEF) process. The X-ray single structure show a planar conjugated structure of the ligand, which was further stabilized by extensive H-bonding and π-π stacking. The photophysical studies related to the sensing behavior of the titular ligand toward aluminum was investigated in detail using various spectroscopic techniques like UV-Vis, photoluminescence, fluorescence and time-correlated single-photon count (TCSPC) and time-resolved NMR. The spectroscopic methods also confirm the selective detection of Al3+ ion in the presence of other metal ions. The theoretical calculations using Density Functional Theory (DFT) and the Time Dependent Density Functional Theory (TD-DFT) provide further insight on the mechanistic aspects of the turn-on sensing behavior including the electronic spectra of both the ligand and the complex. Interestingly, the as-synthesized H2DPC-Al complex can also be utilized as a fluorescence-based sensor for various nitroaromatics including picric acid, for which an INHIBIT logic gate can also be constructed. The as synthesized complex was subsequently used as a fluorescent probe for imaging of human breast adenocarcinoma (MCF7) cells using live cell confocal microscopic techniques.
... 11 Studies also reveal that (Aβ ) and tau undergo conformational changes on interaction with Cu invitro. 12,13 Binding of metals and (Aβ ) to DNA results in conformational changes from the B DNA form to an altered B DNA and ψ -DNA, which is the transition that occurs in the pathology of AD. 14 This change alters the normal functions of the cell which enhances the DNA instability, and is thus implicated in the cellular dysfunctions in brain disorders. [15][16][17][18] The Aβ peptide interacts with Cu and can initiate peptide misfolds and aggregation. ...
Article
The neuroscience and neurobiology of gene editing to enhance learning and memory is of paramount interest to the scientific community. The advancements of CRISPR system have created avenues to treat neurological disorders by means of versatile modalities varying from expression to suppression of genes and proteins. Neurodegenerative disorders have also been attributed to non-canonical DNA secondary structures by affecting neuron activity through controlling gene expression, nucleosome shape, transcription, translation, replication, and recombination. Changing DNA regulatory elements which could contribute to the fate and function of neurons are thoroughly discussed in this review. This study presents the ability of CRISPR system to boost learning power and memory, treat or cure genetically-based neurological disorders, and alleviate psychiatric diseases by altering the activity and the irritability of the neurons at the synaptic cleft through DNA manipulation, and also, epigenetic modifications using Cas9. We explore and examine how each different OMIC techniques can come useful when altering DNA sequences. Such insight into the underlying relationship between OMICs and cellular behaviors leads us to better neurological and psychiatric therapeutics by intelligently designing and utilizing the CRISPR/Cas9 technology.
Article
Deoxyribonucleic acid (DNA) is a vital biomacromolecule. Although the right-handed B-DNA type helical structure is the most abundant and extensively studied form of DNA, several noncanonical forms, such as triplex, quadruplex, Z-DNA, A-DNA, and ss-DNA, have been probed from time to time to gain insights into the DNA's function. Z-DNA was recently found to be involved in cancer and several autoimmune diseases. In the present Article, we evaluated the conformational stability of locked-sugar-based Z-DNA via all-atom explicit-solvent molecular dynamics simulations and found that the modified DNA maintained the left-handed conformation even in the absence of counterions, wherein the structural rigidity dominates over the electrostatic repulsion between the complementary strands. The control Z-DNA without counterions, as expected, instantaneously resulted in unfolded states. The remarkable stability of the conformationally locked model system was thoroughly investigated via structural and energetic perspectives and was probably the result of the backbone widening in tandem with enhanced electrostatics between complementary strands. We believe that the design of the proposed modified Z-DNA construct could help understand the otherwise delicate Z-DNA conformation even in salt-deprived conditions. The design could also motivate the medicinal use of short segments of such modified nucleotides and could be utilized in more advanced modeling techniques, such as DNA origami which has gained popularity in recent years.
Article
Eighty-three brains obtained at autopsy from nondemented and demented individuals were examined for extracellular amyloid deposits and intraneuronal neurofibrillary changes. The distribution pattern and packing density of amyloid deposits turned out to be of limited significance for differentiation of neuropathological stages. Neurofibrillary changes occurred in the form of neuritic plaques, neurofibrillary tangles and neuropil threads. The distribution of neuritic plaques varied widely not only within architectonic units but also from one individual to another. Neurofibrillary tangles and neuropil threads, in contrast, exhibited a characteristic distribution pattern permitting the differentiation of six stages. The first two stages were characterized by an either mild or severe alteration of the transentorhinal layer Pre-alpha (transentorhinal stages I-II). The two forms of limbic stages (stages III-IV) were marked by a conspicuous affection of layer Pre-alpha in both transentorhinal region and proper entorhinal cortex. In addition, there was mild involvement of the first Ammon's horn sector. The hallmark of the two isocortical stages (stages V-VI) was the destruction of virtually all isocortical association areas. The investigation showed that recognition of the six stages required qualitative evaluation of only a few key preparations.
Article
In many forms of DNA the double helix itself forms a helix. This supercoiling, which has important biological consequences, is best described and analyzed by means of a simple mathematical model.
Article
The aluminium (Al)-treated aged rabbit has been proposed as an animal model to understand neuronal cell death in Alzheimer's disease (AD) brain. Immunohistochemical studies show here that Al-maltolate treated aged rabbit brain resembles AD neuropathology in terms of neurofibrillary tangles (NFT), oxidative stress, PHF1, and Aβ deposition and the co-localization of above events with apoptosis in neuronal cells in hippocampus region. To our knowledge, this is the first evidence showing Al-maltolate induced neuropathological changes in the hippocampal region of aged rabbit brain mimicing the neuropathology of AD. This model may serve as a useful animal model to understand the mechanism of neurodegeneration in Alzheimer's disease.
Article
Hexammine cobalt(III) chloride (Co(NH3)63+) provokes a B-DNA → Z-DNA → ÂΨ-DNA conformational transition in poly(dG-dC).poly(dG-dC) and poly(dG-m5dC).poly(dG-m5dC). The circular dichroism spectrum of S′-DNA is characterized by a manyfold increase of positive ellipticity in the range of 300-225 nm and the complete absence of a negative peak. In order to ascertain the helical handedness of >P-DNA, we used a recently developed enzyme immunoassay technique. This method consisted of treating the polynucleotides with Co(NH3)63+ to convert them to the Z- or y-DNA forms and immobilizing these conformations on a microtiter plate. The plates were subsequently treated with a monoclonal anti-Z-DNA antibody Z22, alkaline phosphatase conjugated, affinity purified immunoglobulins, and the phosphatase substrate. The enzyme-substrate reaction was monitored by reading the absorbance at 405 nm with a microplate autoreader. The monoclonal anti-Z-DNA antibody had no reactivity to the B-DNA form, but bound strongly to both the Z- and Ψ-DNA forms, showing that Co(NH3)63+-induced ÂF-DNA form of the polynucleotides exists in the left-handed Z-DNA conformation.
Article
: The fibrillogenic properties of Alzheimer's Aβ peptides corresponding to residues 1–40 of the normal human sequence and to two mutant forms containing the replacement Ala21 to Gly or Glu22 to Gln were compared. At pH 7.4 and 37°C the Gln22 peptide was found to aggregate and precipitate from solution faster than the normal Aβ, whereas the Gly21 peptide aggregated much more slowly. Electron microscopy showed that the aggregates all had fibrillar structures. Circular dichroism spectra of these peptides revealed that aggregation of the normal and Gln22 sequences was associated with spectral changes consistent with a transformation from random coil to β sheet, whereas the spectrum of the Gly21 peptide remained almost unchanged during a period in which little or no aggregation occurred. When immobilised by spotting onto nitrocellulose membranes the peptides bound similar amounts of the radioisotope 65Zn2+. Of several competing metal ions, tested at 20× the concentration of Zn2+, Cu2+ displaced >95% of the radioactivity from all three peptides and Ni2+ produced >50% displacement in each case. Some other metal ions tested caused lesser displacement, but Fe2+ and Al3+ were without effect. In a saturation binding assay, a value of 3.2 µM was obtained for the binding of Zn2+ to Aβ but our data provided no evidence for a reported higher affinity site (107 nM). The results suggest that the neuropathology associated with the Gly21 mutation is not due to enhanced fibrillogenic or different metal-binding properties of the peptide and that the binding of zinc to amyloid peptides is not a specific phenomenon.