ArticlePDF Available

Abstract and Figures

DNA microarray technology is a powerful tool for getting an overview of gene expression in biological samples. Although the successful use of microarray-based expression analysis was demonstrated in a number of applications, the main problem with this approach is the fact that expression levels deduced from hybridization experiments do not necessarily correlate with RNA concentrations. Moreover oligonucleotide probes corresponding to the same gene can give different hybridization signals. Apart from cross-hybridizations and differential splicing, this could be due to secondary structures of probes or targets. In addition, for low-copy genes, hybridization equilibrium may be reached after hybridization times much longer than the one commonly used (overnight, i.e., 15 h). Thus, hybridization signals could depend on kinetic properties of the probe, which may vary between different oligonucleotide probes immobilized on the same microarray. To validate this hypothesis, on-chip hybridization kinetics and duplex thermostability analysis were performed using oligonucleotide microarrays containing 50-mer probes corresponding to 10 mouse genes. We demonstrate that differences in hybridization kinetics between the probes exist and can influence the interpretation of expression data. In addition, we show that using on-chip hybridization kinetics, quantification of targets is feasible using calibration curves.
Content may be subject to copyright.
Vol. 44 ı No. 1 ı 2008 www.biotechniques.com ı BioTechniques ı 109
Research Reports
INTRODUCTION
The ability to quantify the expression
levels of all genes in a given tissue or
cell with a single assay is an exciting
promise. Two technologies are widely
used: quantitative reverse transcription
PCR (RT-PCR) and DNA microarray.
While quantitative RT-PCR is easily
applicable to a few hundred genes
(1,2), DNA microarray technology
has proven very useful for studying
gene expression variations between
many biological samples at the whole
transcriptome scale. Although the
successful application of microarray-
based expression analysis was demon-
strated in a number of applications,
the main problem with this approach
is the fact that hybridization signals
do not necessarily correlate with target
concentrations (3–5).
To get reliable microarray hybrid-
ization data, each probe should be
specific for a unique transcript, and all
probes on the same microarray should
have similar thermodynamic stability.
In addition, hybridization signals should
be maximal in order to reach the highest
sensitivity. During probe design, these
requirements can be considered using
a theoretical sequence-based approach
for prediction of duplex stability
and hairpin formation. However, all
predictions are based on hybridization
models deduced from experiments
in solution and not from experiments
in which probes are immobilized on
DNA microarrays. Even fully comple-
mentary oligonucleotide targets
hybridized to immobilized probes
of the same size might give different
hybridization signals at equilibrium.
Evidently, hybridization with complex
cDNA targets will worsen the situation
because of stable secondary structures
of the targets themselves or interac-
tions between targets. A possibility of
competition between the targets can
also complicate the hybridization data
(6–8). Since expression levels can vary
considerably from one gene to another,
there will be high variations in the
concentration of the different cDNAs
hybridized to the microarray. According
to standard protocols, hybridization is
usually performed overnight (15–18
h) (6). For low-copy genes, reaching
equilibrium might require hybrid-
ization times considerably longer.
Consequently for some probes, the
measured fluorescent signals will not
correspond to a plateau. There are
several important consequences to this.
First, as it is theoretically expected and
experimentally proven (9,10), in non-
equilibrium conditions, the proportion
of cross-hybridized targets is higher as
compared with equilibrium conditions.
Moreover, other factors including
thermodynamic stability of on-chip
formed duplexes and the presence
of secondary structures can change
the kinetic properties of probe-target
combinations.
If different oligonucleotide probes
corresponding to the same cDNA, even
of the same length and similar base
composition, show different kinetic
properties, this will have an impact on
hybridization signals and sensitivity.
In addition, for low-copy genes at
typically non-equilibrium conditions,
different expression results will be
obtained with different probes corre-
sponding to the same gene if the kinetic
properties of the probes are different.
Thus, depending on gene expression
levels and kinetic properties of the
On-chip hybridization kinetics for
optimization of gene expression experiments
Elena Khomyakova
1
, Mikheil A. Livshits
2
, Marie-Caroline Steinhauser
3
, Luce Dauphinot
3
,
Sylvia Cohen-Kaminsky
4
, Jean Rossier
3
, Françoise Soussaline
1
, and Marie-Claude Potier
3
BioTechniques 44:109-117 (January 2008)
doi 10.2144/000112622
DNA microarray technology is a powerful tool for getting an overview of gene expression in biological samples. Although the success-
ful use of microarray-based expression analysis was demonstrated in a number of applications, the main problem with this approach is
the fact that expression levels deduced from hybridization experiments do not necessarily correlate with RNA concentrations. Moreover,
oligonucleotide probes corresponding to the same gene can give different hybridization signals. Apart from cross-hybridizations and
differential splicing, this could be due to secondary structures of probes or targets. In addition, for low-copy genes, hybridization
equilibrium may be reached after hybridization times much longer than the one commonly used (overnight, i.e., 15 h). Thus, hybridiza-
tion signals could depend on kinetic properties of the probe, which may vary between different oligonucleotide probes immobilized on
the same microarray. To validate this hypothesis, on-chip hybridization kinetics and duplex thermostability analysis were performed
using oligonucleotide microarrays containing 50-mer probes corresponding to 10 mouse genes. We demonstrate that differences in
hybridization kinetics between the probes exist and can influence the interpretation of expression data. In addition, we show that using
on-chip hybridization kinetics, quantification of targets is feasible using calibration curves.
1
IMSTAR, Paris, France,
2
EIMB, Moscow, Russia,
3
CNRS UMR 7637, Paris, France, and
4
CNRS UMR 8078, Le Plessis-Robinson, France
110 ı BioTechniques ı www.biotechniques.com Vol. 44 ı No. 1 ı 2008
Research Reports
probes, contradictory results might be
obtained.
Indeed this is what we have
observed when studying mouse gene
expression in various tissues using
DNA microarrays probing neuro-
transmission genes, represented by
two or three 50-mer probes each. Out
of the 280 genes represented on the
microarray, 29 corresponding to 14
genes were selected for detailed experi-
ments. We performed complex analysis
of hybridization including on-chip
analysis of duplex thermostability and
measurement of on-chip hybridization
kinetics at different concentrations of
targets. Antisense Texas Red-labeled
oligonucleotide targets fully comple-
mentary to immobilized probes were
used. Based on these experiments,
we have analyzed some of the causes
for the observed variations in cDNA
hybridization to different probes and
propose the recommendations for
the improvement of the reliability of
expression data, particularly for low-
copy genes.
MATERIALS AND METHODS
Oligonucleotides
Twenty-nine oligonucleotide probes
corresponding to 14 mouse genes
involved in neurotransmission were
purchased from Eurogentec (Liège,
Belgium) as 5end amino-modified 50-
mer deoxyribonucleotides (Table 1).
Oligonucleotide targets complementary
to the 29 oligonucleotide probes were
purchased from Eurogentec as well.
Hybridization of cDNAs to the
NeuroTrans Oligoarrays
Total RNA was extracted from
thymus or cerebral cortex from B10D2
mice using the RNeasy kit (Qiagen,
Valencia, CA, USA). The quality of
RNA was checked on the Agilent 2100
Bioanalyzer (Agilent Technologies,
Santa Clara, CA, USA). Fifteen micro-
grams RNA were labeled with Cy5 by
reverse transcription into cDNAs using
7.5 μM random hexanucleotides, 10 μM
dithiothreitol (DTT), 125 μM dNTP, 25
μM dUTP-cyanine (Amersham plc,
Little Chalfont, Buckinghamshire, UK),
and 400 U of Superscript II (Invitrogen,
Carlsbad, CA, USA). Reference RNA
(equimolar mixture of thymus, thymo-
cytes, and cerebral cortex RNAs) was
labeled with Cy3. Labeled targets
were purified on QIAquick columns
(Qiagen). Cy3 and Cy5 cDNAs were
hybridized to NeuroTrans NT280M
containing oligonucleotides for 280
genes involved in neurotransmission
(two to three oligonucleotides per
gene; GPL4746 on GEO; www.ncbi.
nlm.nih.gov/geo). Hybridization was
performed at 42°C in 50% formamide,
4× standard saline citrate (SSC), 0.1%
sodium dodecyl sulfate (SDS), and
5× Denhart’s overnight. Slides were
washed for 2 min in 2× SSC, 0.1% SDS,
for 1 min in 0.2× SSC, for 1 min in 0.1×
SSC, and scanned on the ScanArray
Lite (Packard Biochip, Billerica, MA,
USA) using the same procedure for
each slide. Fluorescence on each spot
was extracted using the optical scan
array (OSA) image analysis software
package (www.imstarsa.com). Data are
Table 1. Probe Sets Corresponding to 14 Genes Selected for the Kinetic Experiments
Figure 1. The differences between hybridiza-
tion signals for probe pairs corresponding to
the same gene. (A) Expression levels (signal
from a spot divided by the total signal from the
microarray) after hybridizing cDNA targets from
mouse cerebral cortex or thymus overnight. (B)
Signal ratios for pairs of probes to mutual genes
in cerebral cortex and in thymus. (C) Signal ra-
tios between cerebral cortex and thymus for each
probe, deduced from panel A. Values are from
three independent microarray experiments.
C
A
B
Vol. 44 ı No. 1 ı 2008 www.biotechniques.com ı BioTechniques ı 111
Research Reports
accessible on GEO database under the
number GSE7574.
On-Chip Hybridization Kinetic
Experiments
Microarrays specially designed for
kinetic experiments were used. Amino-
modified oligonucleotide probes (Table 1)
were dissolved in 150 mM sodium
phosphate buffer, pH 8.3, at 12.5 μM
and spotted on activated glass slides
according to patent no. WO03068712
using the BioRobotics MicroGrid II
spotter (Genomic Solutions, Ann Arbor,
MI, USA). Each oligonucleotide probe
was spotted in quadruplet. Distance
between spots was 400 μm.
Oligonucleotide targets were labeled
at the 3 end using terminal transferase.
One hundred picomoles of an oligonu-
cleotide were mixed with buffer, 5 mM
CoCl
2
, 0.05 mM Texas Red-5-ddATP
(PerkinElmer, Waltham, MA, USA),
400 U terminal transferase (Roche,
Basel, Switzerland), and incubated for
2 h at 37°C. Reaction was stopped by
adding 2 μL 0.2 M EDTA, pH 8.0, on
ice. Labeled oligonucleotides were
purified on DyeEx columns (Qiagen)
and precipitated in 10 volumes of
acetone and 2% LiClO
4
overnight at
-20°C. After a 10-min centrifugation,
the pellet was rinsed once with acetone,
centrifuged again for 5 min, and
dissolved in water at 12.5 μM final
concentration. The oligonucleotide
concentration was adjusted after quanti-
fication using a ND-1000 spectropho-
tometer (NanoDrop Technologies,
Wilmington, DE, USA) at 260 nm.
Oligonucleotide labeling efficiency was
0.65 as estimated from the absorbance
ratio measured at 260 nm (DNA) and at
612 nm (Texas Red).
Labeled oligonucleotide targets
(equimolar mixture) were deposited
on the microarray and covered with
Gene Frame (Abgene, Epsom, UK).
The volume of the hybridization
chamber was 25 μL (1 cm × 1 cm).
Hybridization was carried out at 60°C
in 2.5× SSC, 0.2% SDS at different
target concentrations from 0.05 to 1
nM for kinetic experiments and 10 nM
for the thermostability measurement.
For registration of on-chip kinetics
and melting curves, OSA instrument
(IMSTAR, Paris, France) equipped
with a thermo-controlled microarray
holder was used (Reference 11 and
patent no. WO2004059302). Curves
were recorded by measuring the
fluorescence intensity on all the spots
of the microarray after scanning,
either at a fixed temperature (60°C
for kinetic curves) or at 2°C intervals
starting from 20° to 95°C (2°C/25 min
at lower temperatures to 2°C/200 s at
higher temperatures for equilibrium
melting curves). Each experiment was
started by the denaturation of probes
and targets within the hybridization
chamber at 95°C for 5 min. For kinetic
experiments, after on-chip denatur-
ation, the temperature was decreased to
60°C followed by the immediate start
of hybridization signals registration at
60°C.
Images were captured and analyzed
using OSA image analysis software
package. The positions of the micro-
array spots were determined using
Figure 2. On-chip melting curves for the 21 probes. Each probe pair is represented on a separate
graph. a.u., arbitrary units.
112 ı BioTechniques ı www.biotechniques.com Vol. 44 ı No. 1 ı 2008
Research Reports
an integrated algorithm of automated
spot finding on the hybridized micro-
array images. For kinetics and melting
experiments, the hybridization signal
S and local background fluorescence
intensity B corresponding to each
spot was calculated as described in
Reference 11. To compensate for the
temperature dependence of Texas
Red and its bleaching after several
measurements, the intensity I(T) of
each spot was calculated as described
in Reference 12:
The number of binding sites (N
binding
site
) involved in duplex formation on the
area of a microarray spot A
spot
[cm
2
]
was estimated from the maximum value
of the fluorescence intensity I
max
(upper plateau of the melting curve),
the concentration of oligonucleotide
targets in solution (C
target
), the thickness
of hybridization chamber (l = 500 μm),
and spot area A
spot
[cm
2
] as follows:.
N
binding sites
= 10
3
I
max
A
spot
l
û
c
target
The maximum value I
max
= 0.5 of
hybridization signal registered in the
experiment with 10 nM oligonucleotide
target at 20°C corresponds to 5 amol (5
× 10
-18
) of binding sites. Thus, the target
quantity exceeds at least 10
2
- to 10
3
-fold
the quantities of immobilized oligo-
nucleotide probes, thus guaranteeing
the constancy of target concentration
during hybridization and melting. The
invariability of background intensity at
a fixed temperature during the kinetics
measurements is an additional proof of
the constancy of the target concentration
during hybridization (data not shown).
RESULTS
Selection of Probes Giving Different
Hybridization Signals for the Same
Gene
Out of 22 hybridization experiments
performed on the NeuroTrans oligoarray
containing 50-mer probes with cDNAs
from mouse forebrain or cerebellum,
14 genes showed signal differences of
at least 10-fold between the two probes
corresponding to the same gene in a
minimum of four experiments (data not
shown). Table 1 shows the list of the
genes. We selected for detailed analysis
10 of these genes for which we obtained
sufficient fluorescent signals in hybrid-
ization kinetics (Table 1, not in gray).
For these 10 genes, signal intensities
were determined from six additional
hybridization experiments performed
with either cerebral cortex or thymus
cDNAs on NeuroTrans oligoarrays.
Figure 1A shows the hybridization
signals for each probe normalized to
the total signal on the oligoarray. Figure
1B gives the signal ratios for the pairs
of probes corresponding to the same
gene. There are 11 ratios for 10 genes,
because one gene was represented by
three probes instead of two. As expected,
the relative expression levels in cerebral
cortex and thymus were contradictory
depending on the oligonucleotide probe
(Figure 1C). To analyze the impact
of hybridization kinetics on gene
expression data, we performed complex
analysis, including on-chip study of
duplex thermostability and on-chip
hybridization kinetics at different target
concentrations. Antisense Texas Red-
labeled oligonucleotide targets fully
complementary to immobilized probes
were used for on-chip experiments in the
same experimental conditions (buffer
composition and temperature) as with
hybridization of cDNA.
Theoretical Model of On-Chip
Hybridization
The on-chip hybridization is a
bimolecular reaction between 50-mers
single-stranded oligonucelotide probes
(P) immobilized to the microarray
surface and the antisense target (T)
molecules, either oligonucleotide or
cDNA, present in the hybridization
solution.
P + T J
The kinetic equation of the corre-
sponding reaction describes the
increase of the quantity of target
molecules hybridized to probes on a
spot and the corresponding increase
of the fluorescent hybridization signal
(13):
[Eq. 1]
Figure 3. An example of on-chip hybridization
kinetics (probe pair 22/23). (A) Normalized
fluorescent signals obtained after 9 h hybridiza-
tion of targets complementary to probes 22 and
23. (B) Fluorescence ratios between probes 22
and 23 during hybridization kinetics deduced
from panel A. (C) Dependence of the hybridiza-
tion time (τ) on target concentration: example for
probes 22 and 23. 1/τ, hybridization rate.
A
B
C
Vol. 44 ı No. 1 ı 2008 www.biotechniques.com ı BioTechniques ı 113
Research Reports
Here, p is the surface density of the
immobilized probe, τ is the hybrid-
ization characteristic time, determined
by the effective rate constants of
association (k
ass
) and dissociation (k
diss
)
and the target concentration (c):
1/τ = k
diss
+k
ass
c = k
diss
(1+Kc)
[Eq. 2]
K = k
ass
/k
diss
is the equilibrium
binding constant for the considered
probe-target. Equation 2 is true also
in case of diffusion-limited kinetics
in spite of a different interpretation of
effective rate constants (14).
It is well known that the forward rate
constants k
ass
are very similar for perfect
match (perf) and for mismatch (mism)
duplexes, while reverse rate constants
k
diss
may differ considerably (9,15,16).
Typically, k
diss
(perf) << k
diss
(mism),
indicating that perfect duplexes are much
more stable than mismatch duplexes
formed upon cross-hybridization (16,17).
Thus, the equilibration time τ is longer
for perfect duplexes than for mismatch
duplexes (see Equation 2 and Reference
9). It was shown both theoretically
(10,18) and experimentally (19) that at
early stages of hybridization the highest
concentration species dominate (either
specific or non-specific), while at later
stages the highest affinity species displace
non-specific binding. Considerable
amount of cross-hybridization (non-
specific binding) is present under non-
equilibrium conditions (9,20).
The possible formation of secondary
structures in the probes and in the targets
can influence both the rate and the extent
of hybridization. The kinetic effects of
secondary structures are documented in
a number of studies (21,22).
To model the on-chip hybridization
kinetics, we used oligonucleotide
targets that were strictly comple-
mentary to the oligonucleotide probes
immobilized on the NeuroTrans micro-
array. The linear dependence (Eq. 2)
of the hybridization rate (1/τ) on target
concentration c allows to measure k
diss
and K and then to deduce τ at the cDNA
target concentrations used during gene
expression experiment.
Thermodynamic Analysis of On-
chip Formed Duplexes
Thermodynamic stability of the
hybridized duplexes was estimated
in a separate experimental study.
Oligonucleotides listed in Table 1 were
spotted onto custom microarrays. These
microarrays were hybridized with Texas
Red-labeled 50-mer oligonucelotide
targets complementary to immobilized
probes until equilibrium was reached.
Rather high oligonucleotide target
concentration (10 nM) was chosen
to ensure fast attainment of hybrid-
ization equilibrium. Hybridization was
performed in the same buffer as the
one used for cDNA hybridization. On-
chip melting of the duplexes was then
recorded. Normalized melting curves
are shown in Figure 2.
Equilibrium melting temperatures
(T
m
) obtained by fitting the theoretical
equation for the occupation level (12):
Figure 4. Simulation of hybridization kinetics for 10 probe sets corresponding to 10 different
genes. The simulation is based on extrapolated oligonucleotide kinetics data. Hybridization signal is
represented by the fraction of occupied probe molecules in a spot.
114 ı BioTechniques ı www.biotechniques.com Vol. 44 ı No. 1 ı 2008
Research Reports
to experimentally obtained melting
curves are represented in Table 2.
However, the similarity of melting
curves observed in the thermodynamic
experiments does not guarantee the
identity of hybridization kinetics. An
example using the experimental data for
the probe pair 22-23 is represented in
Figure 3. Figure 3A shows the difference
in on-chip normalized kinetic curves
for probes 22 and 23 recorded at target
concentrations 0.05, 0.1, 0.25, and 1
nM. The dependence of hybridization
signal ratios between probes 22 and 23
on time is represented in Figure 3B. At
high target concentrations (1 nM), the
equilibration time was short, and the
final value of the ratio between probe 22
and probe 23 was close to 1 (Figure 3B).
For lower target concentrations (0.1 and
0.25 nM), because of the difference in
hybridization kinetics between probes 22
and 23, at the beginning of hybridization,
the signal ratio was high, especially at
concentration 0.1 nM. However, when
approaching equilibrium, the ratio
decreased to smaller values close to 1.
At the lowest target concentration (0.05
nM) even after 50,000 s of hybridization,
the ratio 22/23 was only approaching 1
(Figure 3B).
On-chip Hybridization Kinetics
For a detailed study of the kinetic
effects, we have measured the on-chip
hybridization kinetics of Texas Red-
labeled 50-mer oligonucleotide targets
at four concentrations: 0.05, 0.1, 0.25,
and 1 nM. Results of fitting (K and
k
diss
values) are listed in Table 2. These
experiments allowed the calculation
of τ as well as the modeling of hybrid-
ization kinetic curves at any target
concentration.
Simulation of Kinetics Curves from
the Theoretical Model
As an example we simulated the
kinetic curves for 0.01 and 0.005 nM
target concentrations, using Equation
1, and K and k
diss
values listed in
Table 2. Evidently, such estimation
of kinetic parameters K and k
diss
has a
limitation coming from experimental
errors that are more essential at lower
target concentrations. Figure 4 shows
the curves obtained for all the probes
tested. Figure 5 indicates the simulated
hybridization signal ratios after 15
h of hybridization (panel A) and at
saturation (panel B) for the analyzed
probe pairs at four concentrations of
oligonucleotide targets: 0.05, 0.01,
0.005, and 0.001 nM. For slow kinetics,
duplex hybridization ratios are different
depending on the hybridization time.
DISCUSSION
The discrepancy between hybrid-
ization signals obtained from different
probes corresponding to the same
gene is a problem for gene expression
analysis. Except for the well-known
alternative splicing and stable secondary
conformation of cDNA targets, some
Figure 5. Comparison of pair ratios for
simulated hybridization signals after 15 h
of hybridization (A) and at equilibrium (B).
Melting temperature (T
m
) deduced from on-chip
experiments.
Probes T
m
(°C) k
ass
K
3 60.8 1.64 × 10
6
3.23 × 10
10
4 62.1 1.45 × 10
6
2.67 × 10
10
5 59.3 9.74 × 10
5
3.97 × 10
10
6 56.4 4.96 × 10
5
1.46 × 10
11
7 58.3 7.93 × 10
5
1.01 × 10
10
8 55.7 1.55 × 10
6
1.23 × 10
10
9 57.1 1.19 × 10
6
1.03 × 10
10
12 57.8 9.37 × 10
5
3.25 × 10
10
13 56.7 1.71 × 10
6
2.63 × 10
10
14 62.3 1.65 × 10
6
4.95 × 10
10
15 64.4 1.05 × 10
6
2.17 × 10
10
18 59.3 1.10 × 10
6
5.24 × 10
10
19 59.6 1.72 × 10
6
4.85 × 10
10
20 58.4 1.53 × 10
6
2.62 × 10
10
21 64.6 9.44 × 10
5
1.17 × 10
10
22 68.7 1.30 × 10
6
4.46 × 10
10
23 67.0 8.23 × 10
5
9.90 × 10
10
26 62.3 7.51 × 10
5
1.39 × 10
10
27 56.6 1.35 × 10
6
1.63 × 10
10
28 65.6 1.14 × 10
6
2.58 × 10
10
29 63.7 7.79 × 10
5
8.70 × 10
9
T
m
, melting temperature; k
ass
, rate constant of association; K, equilibrium binding constant.
Table 2. Kinetics Parameters and T
m
Deduced from On-Chip Experiments
A
B
Vol. 44 ı No. 1 ı 2008 www.biotechniques.com ı BioTechniques ı 115
Research Reports
other factors could influence micro-
array expression results. To analyze
these factors we performed thermo-
stability study and analysis of on-chip
hybridization kinetics at four concen-
trations of complementary oligonucel-
otide targets: 1, 0.25, 0.1, and 0.05 nM.
Theoretical analysis applied to our on-
chip kinetics data allowed estimation of
kinetic behavior at target concentrations
corresponding to cDNA concentrations
in gene expression experiments.
The probes selected for the study
showed high hybridization signals as
well as a significant difference between
probes corresponding to the same gene
(Figure 1). Figure 1B indicates at least
a 2-fold difference in the hybridization
signals for each pair of probes in
cerebral cortex and/or thymus. When
the difference is the same for both
tissues, it is most probably caused by
different secondary structures of the
probes or/and of the targets. Such a case
makes no confusion in the interpretation
of the expression data. However, several
probe pair ratios were found to differ
in cerebral cortex and thymus (probes
3-4, 7-8, 7-9, 14-15, 18-19, 20-21, 22-
23, 28-29) (Figure 1B). This entailed
a difference in expression level ratios
between cerebral cortex and thymus
depending on the probe (Figure 1C).
This difference could be due to the
presence or absence of alternative
transcripts in cerebral cortex or thymus
that would correspond to one probe or to
the other. Indeed probes from pairs 19-
18, 20-21, 22-23, and 29-28 map either
to different exons or to alternative 5 or
3 untranslated regions (UTRs) in two
databases (www.ensembl.org; genome.
ewha.ac.kr). However, for pairs 3-4, 14-
15, and 27-28, the two probes mapped
to the same transcript. Alternatively, the
observed tissue-dependent variations
in signal ratios between two probes
corresponding to the same gene could
result from differences in hybridization
kinetics at different cDNA target concen-
trations. In addition, kinetic properties
of the probes could be different within
the same microarray. Thus, if a gene is
highly expressed, hybridization signals
at the end of the experiment may reach
saturation levels. In this condition, signal
ratio between oligonucleotide pairs will
only depend on the equilibrium binding
constant K of on-chip duplex formation
(see above). However, for low-copy
genes, hybridization signals might be
rather far from saturation, and thus
hybridization signal ratios will depend
on the hybridization time and target
concentration.
To illustrate this, we performed
on-chip kinetics analysis at different
target concentrations and modeled
the behavior of the duplexes. We first
analyzed the temperature stability
of on-chip formed duplexes. The
experimental data (Figure 2) demon-
strate that thermodynamic stability
of the duplexes, including also the
enthalpy ΔH of duplex formation,
reflected by the shape of the melting
curves, is similar for all immobilized
probes tested. The minor differences
observed cannot explain the difference
in hybridization signals of oligonucle-
otide probes belonging to the same
gene. In addition, the experimental
data confirmed the correct design of
oligonucleotide probes and the absence
of degradation of either probes or
targets. However, the similarity of
melting curves does not guarantee
the identity of hybridization kinetics.
Indeed, probes 22 and 23 have similar
melting curves but different hybridiza-
tions curves (Figure 3, A and B). This
suggests that signal ratios between the
two probes depend on target concen-
tration and hybridization time.
Based on the kinetic data obtained at
higher concentrations of targets using
Equations 1 and 2, it was possible to
model the kinetics behavior of the
duplexes at target concentrations corre-
sponding to real cDNA hybridization
experiments. DNA target concentration
in biological samples was estimated
using the data of Reference 23.
Approximately 10 μg RNA are typically
used for microarray experiments. This
amount corresponds to RNA quantities
contained in 10
6
cells. The copy number
of single genes in a cell varies from
1 to 10
4
, and the efficiency of reverse
transcription is about 20% (24). Thus,
the expected concentration of cDNA
in 20 μL hybridization solution is 10
-14
to 10
-10
M. Unfortunately, the on-chip
kinetic measurements at concentrations
lower than 5 × 10
-11
M (0.05 nM) are
hardly feasible because the fluorescent
hybridization signals are too low as
compared with the fluorescence of the
hybridization solution. Thus, according
to Equation 2, using the linear depen-
dence of 1/τ value measured at concen-
trations 0.05, 0.1, and 0.25 nM, we can
estimate the K and k
diss
values for each
probe-target pair (Figure 4 and Table
2). Based on these data, we were able to
extrapolate the kinetics behavior at low
target concentrations after 15 h hybrid-
ization and at saturation (Figure 5, A
and B). Figure 4 shows that 15 h (54,000
s) hybridization time was not always
sufficient for reaching at least 70% of
maximal hybridization signal, especially
at 0.005 nM target concentration (probes
6, 12 14, 18, 23, and 28). Obviously, for
target concentration below 0.005 nM,
the hybridization kinetics would be even
slower. Thus, one can expect that after
hybridization times used in classical
experiments (roughly 15 h), the majority
of hybridization signals is rather far
from saturation level.
While at equilibrium (saturated
signals) a decrease in target concen-
tration leads to an increase of the
absolute value of the hybridization signal
ratio (Figure 5B), at non-equilibrium
conditions (15 h of hybridization), and
at low concentrations, some probe pairs
(5-6, 12-13, 18-19, and 22-23) had
ratios inverted as compared with the one
obtained at equilibrium. Moreover, two
pairs (5-6 and 28-29) had lower ratios
after 15 h as compared with equilibrium
conditions. Since probes 6 and 28 are
characterized by slow hybridization
kinetics (Figure 4), 15 h is definitively
not enough for reaching equilibrium.
In pairs of probes addressed to
mutual genes, the highest hybridization
signals were obtained with probes 4, 6,
8, 13, 14, 19, 20, 23, 27, and 29 (Figure
1A). Only some of these probes (14,
19, 20, and 27) showed hybridization
signals that differed more than 4-fold
for cDNA targets from cerebral cortex
and thymus (Figure 1C). Simulation of
kinetics indicated that for these probes,
a decrease in target concentration is
expected to induce an increase of the
hybridization signal ratio between
probes within the same pair (Figure 5A).
Indeed for probe 14, which corresponds
to a gene that is expressed more in
cerebral cortex than in thymus (Figure
1A), a decrease in target concentration
induced a decrease in hybridization
signal (Figure 5A). However, for probes
116 ı BioTechniques ı www.biotechniques.com Vol. 44 ı No. 1 ı 2008
Research Reports
19 and 20, a decrease in target concen-
tration in cerebral cortex as compared
with thymus induced an increase in
hybridization signal, which is opposite
to what was predicted. Thus for these
two probes, the kinetics model did not
comply with the expression data. The
most likely explanation may be the
presence of splicing variants in cerebral
cortex but not in thymus.
The tissue-dependent alternative
splicing could be proposed as an alter-
native explanation for the low hybrid-
ization signal for probe 22 as well.
However, since a very low hybridization
signal was observed for probe 22 both
in cerebral cortex and thymus, the most
probable explanation seems to be the
hairpin formation in cDNA fragment
corresponding to probe 22. Stable
secondary structure in cDNA could also
explain the hybridization results obtained
for probes 5, 6, 9, 12, 26, and 28.
Finally, for probe pair 3-4, there was
no difference in expression levels of the
corresponding gene between cerebral
cortex and thymus (Figure 1A). Thus,
the comparatively lower intensity of
hybridization signal corresponding to
probe 3 for both thymus and cerebral
cortex could also be the consequence of
cDNA secondary structure formation.
Based on our results and to improve
microarray gene expression results, we
can now make several recommenda-
tions. First, one should apply the best
experimental conditions for reaching
equilibrium for the majority of cDNAs
hybridized on the microarray. One way
would be to increase the hybridization
time at least up to 24 h. Another solution,
which seems to be more appropriate,
would be to decrease the volume of the
hybridization solution. Using micro-
fluidic stations will greatly accelerate
the hybridization process and essentially
improve the expression data analysis
(6). Secondly, fragmentation of cDNA
targets will result in destabilization of
the secondary conformations. Also, more
attention should be brought to oligonu-
cleotide probe design, including minimi-
zation of hairpin formation. It should be
noted, too, that the kinetic properties of
oligonucleotides are difficult to predict a
priori. Performing experimental studies
of on-chip kinetics for all the immobilized
probes is rather complicated. Thus, the
optimal solution could be to study on the
same microarray several oligonucelotide
probes (at least 3, ideally 5–10) comple-
mentary to different regions of the gene
and to select for expression analysis only
the probes showing the highest hybrid-
ization signals. This should be helpful in
avoiding the misinterpretation coming
from slow-kinetics probes as well as in
excluding the influence of alternative
splicing and of secondary structures of
DNA targets on the hybridization results.
Finally, we demonstrate that
following the linear dependence
between the characteristic hybridization
kinetics and the target concentration, it
is possible, using calibration curves, to
deduce the target copy numbers from
on-chip hybridization experiments.
ACKNOWLEDGMENTS
The authors wish to thank E.
Tatarinova for very helpful discussion, C.
Mikonio, D. Le Clerre, and F. Richard for
spotting the oligoarrays, and S. Ursulet
and E. Dreval for technical help.
COMPETING INTERESTS
STATEMENT
The authors declare no competing
interests
REFERENCES
1. Dolganov, G.M., P.G. Woodruff, A.A.
Novikov, Y. Zhang, R.E. Ferrando, R.
Szubin, and J.V. Fahy. 2001. A novel method
of gene transcript profiling in airway biopsy
homogenates reveals increased expression of
a Na+-K+-Cl- cotransporter (NKCC1) in asth-
matic subjects. Genome Res. 11:1473-1483.
2. Czechowski, T., R.P. Bari, M. Stitt, W.-R.
Scheible, and M.K. Udvardi. 2004. Real-time
RT-PCR profiling of over 1400 Arabidopsis
transcription factors: unprecedented sensitivity
reveals novel root- and shoot-specific genes.
Plant J. 38:366-379.
3. Hekstra, D., A.R. Taussig, M. Magnasco,
and F. Naef. 2003. Absolute mRNA concen-
trations from sequence-specific calibration of
oligonucleotide arrays. Nucleic Acids Res.
31:1962-1968.
4. Held, G.A., G. Grinstein, and Y. Tu. 2006.
Relationship between gene expression and
observed intensities in DNA microarrays—a
modeling study. Nucleic Acids Res. 24:E70.
5. Kuo, W.P., F. Liu, J. Trimarchi, C. Punzo,
M. Lombardi, J. Sarang, M.E. Whipple, M.
Maysuria, et al. 2006. A sequence-oriented
comparison of gene expression measurements
across different hybridization-based technolo-
gies. Nat. Biotechnol. 24:832-840.
6. Bishop, J., S. Blair, and A. Chagovetz. 2006.
A competitive kinetic model of nucleic acid
surface hybridization in the presence of point
mutants. Biophys. J. 90:831-840.
7. Horne, M.T., D.J. Fish, and A.S. Benight.
2006. Statistical thermodynamics and kinet-
ics of DNA multiplex hybridization reactions.
Biophys. J. 91:4133-4153.
8. Fish, D.J., M.T. Horne, R.P. Searles, G.P.
Brewood, and A.S. Benight. 2007. Multiplex
SNP discrimination. Biophys. J. 92:L89-L91.
9. Dai, H., M. Meyer, S. Stepaniants, M.
Ziman, and R. Stoughton. 2002. Use of hy-
bridization kinetics for differentiating specific
from non-specific binding to oligonucleotide
microarrays. Nucleic Acids Res. 30:E86.
10. Bhanot, G., Y. Louzoun, J. Zhu, and C.
DeLisi. 2003. The importance of thermody-
namic equilibrium for high throughput gene
expression arrays. Biophys. J. 84:124-135.
11. Khomyakova, E.B., E.V. Dreval, M. Tran-
Dang, M.-C. Potier, and F.P. Soussaline.
2004. Innovative instrumentation for micro-
array scanning and analysis: application for
characterization of oligonucleotide duplexes
behavior. Cell. Mol. Biol. 50:217-248.
12. Fotin, A.V., A.L. Drobyshev, D.Y.
Proudnikov, A.N. Perov, and A.D.
Mirzabekov. 1998. Parallel thermodynamic
analysis of duplexes on oligodeoxyribonucleo-
tide microchips. Nucleic Acids Res. 15:1515-
1521.
13. Lauffenderberger, D.A. and J.J. Linderman.
1996. Receptors. Oxford University Press,
New York, NY.
14. Sorokin, N.V., V.R. Chechetkin, S.V.
Pan’kov, O.G. Somova, M.A. Livshits,
M.Y. Donnikov, A.Y. Turygin, V.E. Barsky,
and A.S. Zasedatelev. 2006. Kinetics of hy-
bridization on surface oligonucleotide micro-
chips: theory, experiment, and comparison
with hybridization on gel-based microchips. J.
Biomol. Struct. Dyn. 24:57-66.
15. Livshits, M.A. and A.D. Mirzabekov. 1996.
Theoretical analysis of the kinetics of DNA
hybridization with gel-immobilized oligos.
Biophys. J. 71:2795-2801.
16. Wang, S., A.E. Friedman, and E.T. Kool.
1995. Origins of high sequence selectivity:
a stopped-flow kinetics study of DNA/RNA
hybridization by duplex- and triplex-forming
oligos. Biochemistry 34:9774-9784.
17. Ikuta, S., K. Takagi, R.B. Wallace, and K.
Itakura. 1987. Dissociation kinetics of 19
base paired oligo-DNA duplexes contain-
ing different single mismatched base pairs.
Nucleic Acids Res. 15:797-811.
18. Zhang, Y., D.A. Hammer, and D.J. Graves.
2005. Competitive hybridization kinetics re-
veals unexpected behavior patterns. Biophys.
J. 89:2950-2959.
19. Bishop, J., C. Wilson, A.M. Chagovetz, and
S. Blair. 2007. Competitive displacement of
DNA during surface hybridization. Biophys. J.
92:L10-L12.
20. Sartor, M., J. Schwanekamp, D. Halbleib, I.
Mohamed, S. Karyala, M. Medvedovic, and
C.R. Tomlinson. 2004. Microarray results im-
prove significantly as hybridization approaches
equilibrium. BioTechniques 36:790-796.
21. Sekar, M.M.A., W. Bloch, and P.M. St. John.
2005. Comparative study of sequence-depen-
Vol. 44 ı No. 1 ı 2008 www.biotechniques.com ı BioTechniques ı 117
Research Reports
dent hybridization kinetics in solution and on
microspheres. Nucleic Acids Res. 33:366-375.
22. Gao, Y., L.K. Wolf, and R.M. Georgiadis.
2006. Secondary structure effects on DNA
hybridization kinetics: a solution versus sur-
face comparison. Nucleic Acids Res. 34:3370-
3377.
23. Carter, M.G., A.A. Sharov, V. VanBuren,
D.B. Dudekula, C.E. Carmack, C. Nelson,
and M.S. Ko. 2005. Transcript copy number
estimation using a mouse whole-genome oli-
gonucleotide microarray. Genome Biol. 6:
R61.
24. Tsuzuki, K., B. Lambolez, J. Rossier, and
S. Ozawa. 2001. Absolute quantification of
AMPA receptor subunit mRNAs in single hip-
pocampal neurons. J. Neurochem. 77:1650-
1659.
Received 29 June 2007; accepted
18 September 2007.
M.-C.S.’s present address is Max Planck
Institute of Molecular Plant Physiology, Am
Muehlenberg 1, 14476 Golm, Germany.
Address correspondence to Elena
Khomyakova, IMSTAR, 60 rue Notre Dame
des Champs, 75006 Paris, France. e-mail:
elena.khomiakova@imstar.fr; or Marie-
Claude Potier, CNRS UMR 7637, ESPCI 10
rue Vauquelin, 75005 Paris, France. e-mail:
marie-claude.potier@espci.fr
To purchase reprints of this article, contact:
Reprints@BioTechniques.com
... Results from northern hybridization studies in which the GDH-synthesized RNAs were used as probes suggested their involvement in the reprogramming of the abundance of mRNAs, and in the coordinate regulation of catabolism and anabolism [18]. It is not yet possible to detect early signs of unfavorable drug metabolism by analysis of the target mRNAs [5] because cellular expression levels vary considerably from one gene to another [24] and the genetic code-based nucleic acid probes utilized in the hybridization are unable to discriminate and/or integrate hybridization signals with reference to mRNA concentrations. Hairpins, loops and other secondary structural conformations in the mRNAs are the major culprits because they tend to limit the optimal hybridization of probes/primers to target sites in the mRNA. ...
... The results of the Blast 2 sequences alignment between fragments 3a and 3b displayed 4 gaps and 51 mismatches which introduced the flexibility that enabled the GDH-synthesized RNA to interact fully with its target mRNA to give maximum hybridization signals despite the impediment created by mRNA secondary structures. Because the genetic code-based probe 3b matched the target mRNA perfectly, it was rigidly unable to flip around mRNA secondary structures with the consequence that its hybridization signals were lower than expected and substantially inaccurate [24]. The mRNA encoding the translational factor was regulated independent of that encoding the AOX2a in the 4NTPs and ATP treatments, but they were at normal abundance in the control, GTP, combined ATP plus UTP, combined GTP plus CTP, and UTP treatments thus further suggesting that these treatments elicited unfavorable pharmacokinetics that induced catalytic RNA synthesis by GDH. ...
... This was due to probe #5a being an eleven times overlapping repeat, it had higher probability to hybridize to the target mRNA by overcoming mRNA secondary structures. Hairpins, loops and other intramolecular base pairing in mRNA are some of the major factors that minimize the stability of duplexes in northern hybridization [24]. The eleven overlapping repeats permitted fragment #5a to flip around secondary structures in the mRNA and to achieve optimal hybridization signal/sensitivity as compared with fragment #5b. ...
Article
Full-text available
Glutamate dehydrogenase (GHD) synthesizes some RNAs that regulate mRNA abundance in response to the environment. The connection of gene expression and drug metabolism by the GDH-synthesized RNA has not been dem-onstrated experimentally. The regulation of the mRNAs encoding the drug-metabolizing enzymes was studied by northern hybridization using the GDH-synthesized RNAs as probes. The mRNAs encoding cytochrome P-450 reductase, UDP-glucosyltransferase, alternative oxidase, and ABC-transporters were upregulated by the administered ATP+UTP+GTP. Also superoxide dismutase and GSH S-transferase were upregulated by administered ATP. The untreated control, GTP, and UTP did not upregulate any of the mRNAs. The mRNAs encoding the enzymes were coordinately regulated at the molecular level. All the enzymes are also active in drug detoxication in mammals. Photometric assays of enzyme activi-ties confirmed that the enzymes were present at levels proportional to their respective encoding mRNAs as detected by the GDH-synthesized RNA probes. Genetic code-based nucleic acid probes were partially accurate in detecting the mRNAs encoding the enzymes. Therefore, GDH-synthesized RNAs are important genetic metabolic probes for the screening of mRNAs encoding the drug metabolizing enzymes. Nucleoside triphosphates and analogs are antihypertensives, antineo-plastics, antiarrhythmics, antimetabolites, antiviral agents etc and they induce GDH isomerization.
... Variations in signal intensities of probes corresponding to different regions of the same mRNA target have previously been observed [29,30], and highly sequence dependent313233. The hybridization efficiency between a probe and its targets is determined by the balance between the binding strength of the probe-target duplex and the formation of probe-probe dimers and secondary structures in either probes or targets [34,35]. The duplex melting temperature is generally considered as one of the most popular measures in the evaluation of microarray probes. ...
Article
Full-text available
Background We performed a Nimblegen intra-platform microarray comparison by assessing two categories of flax target probes (short 25-mers oligonucleotides and long 60-mers oligonucleotides) in identical conditions of target production, design, labelling, hybridization, image analyses, and data filtering. We compared technical parameters of array hybridizations, precision and accuracy as well as specific gene expression profiles. Results Comparison of the hybridization quality, precision and accuracy of expression measurements, as well as an interpretation of differential gene expression in flax tissues were performed. Both array types yielded reproducible, accurate and comparable data that are coherent for expression measurements and identification of differentially expressed genes. 60-mers arrays gave higher hybridization efficiencies and therefore were more sensitive allowing the detection of a higher number of unigenes involved in the same biological process and/or belonging to the same multigene family. Conclusion The two flax arrays provide a good resolution of expressed functions; however the 60-mers arrays are more sensitive and provide a more in-depth coverage of candidate genes potentially involved in different biological processes.
... where k a (M 21 s 21 ) is the association rate constant and k d (s 21 ) is the dissociation rate constant describing the formation and dissociation of the complex. The amount of complex can be rewritten as [15] LR ½ Š~R tot : ...
Article
Full-text available
The interaction of the epidermal growth factor (EGF) with its receptor (EGFR) is known to be complex, and the common over-expression of EGF receptor family members in a multitude of tumors makes it important to decipher this interaction and the following signaling pathways. We have investigated the affinity and kinetics of 125I-EGF binding to EGFR in four human tumor cell lines, each using four culturing conditions, in real time by use of LigandTracer®. Highly repeatable and precise measurements show that the overall apparent affinity of the 125I-EGF – EGFR interaction is greatly dependent on cell line at normal culturing conditions, ranging from KD≈200 pM on SKBR3 cells to KD≈8 nM on A431 cells. The 125I-EGF – EGFR binding curves (irrespective of cell line) have strong signs of multiple simultaneous interactions. Furthermore, for the cell lines A431 and SKOV3, gefitinib treatment increases the 125I-EGF - EGFR affinity, in particular when the cells are starved. The 125I-EGF - EGFR interaction on cell line U343 is sensitive to starvation while as on SKBR3 it is insensitive to gefitinib and starvation. The intriguing pattern of the binding characteristics proves that the cellular context is important when deciphering how EGF interacts with EGFR. From a general perspective, care is advisable when generalizing ligand-receptor interaction results across multiple cell-lines.
Article
The real-time monitoring of the hybridization signal, giving access to the reaction kinetics, can widen the results of a microarray experiment. Nevertheless, the presence of a strong fluorescent mix often degrade the experimental sensitivity, limiting the interest of this technique: the implementation of an evanescent wave excitation scheme can represent in this case a real advantage. In this paper, we propose high refractive index waveguides fabricated by sol–gel process for evanescent wave microarray applications. The influence of the sol composition and annealing parameters on materials microstructure are carefully studied to obtain good optical properties and high refractive index (n = 1.8–2.1) using thermal treatments below 250 °C. Monomode TiO2 planar waveguides chelated by acetic acid are used as substrates for waveguide-based microarray in real-time experiments, demonstrating a significant increase of the signal to noise ratio.
Article
To develop a simple micro-platform for gene mutation detection, we used a DNA hybridization approach in a combined microchannel. A glass groove immobilized with oligonucleotide probes by UV-crosslinking, and encased in a transparent heat-shrinkable polythene tube. After heat treatment, the polythene tube stretched tightly over the glass groove and was designated as a combined capillary chip (CCC). The CCC assay was optimized with 10 μL reaction solutions shuttling back and forth at 50 μL min−1 for 10 min, give a detection minimum of 0.1 nM target DNA sequences. In hepatitis B virus (HBV) YMDD mutations detection, 61.3% (92/150) was of a single genotype and it was authenticated with sequence analysis. The other 38.7% (58/150) was mixed genotype detected, but 18.9% (11/58) has a missed diagnosis in sequence analysis. It proved a higher sensitivity than sequence analysis. The CCC assay has a simple fabricating process, simple structure and higher specificity in gene mutation detection. These features are promising for clinical gene mutation analysis.
Article
Thermal denaturation, or melting, measurements are a classic technique for analysis of thermodynamics of nucleic base driven associations in solution, as well as of interactions between nucleic acids and small molecule ligands such as drugs or carcinogens. Performed on surface-immobilized DNA films, this well-established technique can help understand how energetics of surface hybridization relate to those in solution, as well as provide high-throughput platforms for screening of small molecule ligands. Here we describe methods for measuring DNA melting transitions at solid/liquid interfaces with focus on the role of immobilization chemistry, including a common "immobilization-through-self-assembly" approach that is effective at moderate temperatures, and a thermo-stable approach based on polymer-supported DNA monolayers that can be used at elevated temperatures. We also discuss conditions necessary for reversible measurements, as signified by superimposition of the association (cooling) and dissociation (heating) transitions of immobilized DNA strands.
Article
New applications of microarray technology requires new approaches to microarray reader development. Miniaturization of the biochips, necessity of massive parallel microarray-based analysis of various biological samples as well as increasing application of different real-time on-chip reactions for research and diagnostics demand improvement of microarray reader instrumentation. We represent the optical scanning array (OSA) reader system, allowing fast capture of high-resolution multifluorescence images and equipped with thermocontrolled hybridization chamber which is required for registration of a wide range of on-chip real-time processes. We demonstrate the application of OSA microarray reader for optimisation of hybridization and washing conditions for microarray-based detection of Q493X mutation of CFTR gene. It was shown that detailed analysis of on-chip thermodynamic and kinetic behaviour of oligonucleotide/oligonucleotide and ssPCR/oligonucleotide duplexes is very helpful for optimisation of experimental procedure.
Article
The competition measurement using simultaneous incubation of labeled and unlabeled Ligand is a common method to assess the specificity of a biomolecular interaction. In this paper we show that invalid assumptions about the interactions may lead to improper experimental setups which in turn can result in inaccurate conclusions about the specificity. To improve understanding of competition measurements, simulations in MATLAB as well as real-time interaction analysis using LigandTracer have been performed. We show that use of a concentration of unlabeled Ligand of at least 10 × K(D) is necessary for assay accuracy. Increasing the incubation time to assure equilibrium, adding a pre-incubation phase, and a general understanding of the reversibility of an interaction may also improve the reliability of the measurement and the conclusions drawn about specificity. These findings may lower the risk of false negative results as well as reducing the amount of reagent needed.
Article
Full-text available
The dissociation kinetics of 19 base paired oligonucleotide-DNA duplexcontaining a various single mismatched base pair are studied on dried agarose gels. The kinetics of the dissociation are first order under our experimental conditions. The incorporation of a single mismatched base pair destabilizes the DNA duplexes to some extent, the amount depending on the nature of the mismatched base pair. G-T and G-A mismatches slightly destabilize a duplex, while A-A, T-T, C-T and C-A mismatches significantly destabilize it. The activation energy for the overall dissociation processes for these oligonucleotide-DNA duplexes containing 19 base pairs is 52 ± 2 Kcalmol−1 as determined from the slope of Arrhenius plot.
Article
Full-text available
Comprehensive and systematic analysis of airway gene expression represents a strategy for addressing the multiple, complex, and largely untested hypotheses that exist for disease mechanisms, including asthma. Here, we report a novel real-time PCR-based method specifically designed for quantification of multiple low-abundance transcripts using as little as 2.5 fg of total RNA per gene. This method of gene expression profiling has the same specificity and sensitivity as RT-PCR and a throughput level comparable to low-density DNA microarray hybridization. In this two-step method, multiplex RT-PCR is successfully combined with individual gene quantification via real-time PCR on generated cDNA product. Using this method, we measured the expression of 75 genes in bronchial biopsies from asthmatic versus healthy subjects and found expected increases in expression levels of Th2 cytokines and their receptors in asthma. Surprisingly, we also found increased gene expression of NKCC1--a Na+-K+-Cl- cotransporter. Using immunohistochemical method, we confirmed increased protein expression for NKCC1 in the asthmatic subject with restricted localization to goblet cells. These data validate the new transcriptional profiling method and implicate NKCC1 in the pathophysiology of mucus hypersecretion in asthma. Potential applications for this method include transcriptional profiling in limited numbers of laser captured cells and validation of DNA microarray data in clinical specimens.
Article
Hybridization kinetics were found to be significantly different for specific and non‐specific binding of labeled cRNA to surface‐bound oligonucleotides on microarrays. We show direct evidence that in a complex sample specific binding takes longer to reach hybridization equilibrium than the non‐ specific binding. We find that this property can be used to estimate and to correct for the hybridization contributed by non‐specific binding. Useful applications are illustrated including the selection of superior oligonucleotides, and the reduction of false positives in exon identification.
Article
α-Amino-3-hydroxy-5-methylisoxazole-4-propionate (AMPA) receptor subunit (GluR1–4) mRNAs expressed by single neurons in rat hippocampal cultures were quantified by single-cell RT-PCR using an internal standard RNA after whole-cell patch-clamp recording. The internal standard RNA, derived from GluR2 with a single nucleotide substitution, was reverse-transcribed and PCR-amplified with the same efficiency as GluR1–4 mRNAs. The mean mRNA numbers harvested in vitro from pyramidal-like neurons on day 9 were 1150 ± 324 molecules of GluR1, 1080 ± 273 molecules of GluR2, 100 ± 20 molecules of GluR3, and 50 ± 10 molecules of GluR4 (mean ± SEM, n = 12). In a non-pyramidal neuronal population that expresses AMPA receptors characterized by high Ca2+ permeability, the numbers of GluR1, GluR3 and GluR4 mRNA molecules harvested per cell were 354 ± 64, 25 ± 17 and 168 ± 36, respectively (n = 8). The GluR2 mRNA was not detected in this cell type. The calculated ratio of AMPAR mRNA molecules per total mRNA molecules was 1/240 in pyramidal-like neurons (1/500 for GluR2), being in the range obtained with total RNA from rat forebrain and cerebellum (1/170 and 1/380, respectively). Finally, our results indicated that the proportion of GluR1–4 mRNA located in neurites reached ∼60% in pyramidal-like neurons. However, we found no evidence of preferential subcellular distribution of a given subunit.
Article
Stopped-flow UV kinetics and thermal denaturation experiments are used to examine the origins of high sequence selectivity and binding affinity of circular triplex-forming oligonucleotides with single-stranded DNA/RNA targets. These 34-nt probes are hybridized to a series of 12-nt target sequences which are fully complementary or which contain a single mismatch. Also studied for comparison are standard 12-nt Watson-Crick DNA or RNA complements. Several novel findings are described: (1) Circular triplex-forming oligomers bind targets with very high thermodynamic selectivity (up to 8-10 kcal/mol against a single-nucleotide mismatch), while linear strands show only 2-3 kcal/mol selectivity. (2) Rates for triplex formation by circular ligands are much greater than other reported triplex formation modes and are nearly the same as for Watson-Crick duplex formation. (3) DNA-DNA and RNA-RNA hybridization rates are similar for both duplex and triplex formation. (4) For both modes of binding, hybridization rates do not vary when a mismatch is introduced into the target, and, therefore, binding selectivity is reflected in large variations in dissociation, rather than association rates. Finally, (5) binding selectivity of circular ligands becomes significantly greater as pH is lowered; results indicate that the high sequence selectivity of the circular DNA ligand is due in large part to the special stability of the protonated C+G-C triad relative to unprotonated mismatched triads. The results are useful in the understanding of properties of nucleic acid complexes in general and give insight into optimum design for synthetic DNA-binding ligands.
Article
A new method of DNA sequencing by hybridization using a microchip containing a set of immobilized oligonucleotides is being developed. A theoretical analysis is presented of the kinetics of DNA hybridization with deoxynucleotide molecules chemically tethered in a polyacrylamide gel layer. The analysis has shown that long-term evolution of the spatial distribution and of the amount of DNA bound in a hybridization cell is governed by "retarded diffusion," i.e., diffusion of the DNA interrupted by repeated association and dissociation with immobile oligonucleotide molecules. Retarded diffusion determines the characteristic time of establishing a final equilibrium state in a cell, i.e., the state with the maximum quantity and a uniform distribution of bound DNA. In the case of cells with the most stable, perfect duplexes, the characteristic time of retarded diffusion (which is proportional to the equilibrium binding constant and to the concentration of binding sites) can be longer than the duration of the real hybridization procedure. This conclusion is indirectly confirmed by the observation of nonuniform fluorescence of labeled DNA in perfect-match hybridization cells (brighter at the edges). For optimal discrimination of perfect duplexes from duplexes with mismatches the hybridization process should be brought to equilibrium under low-temperature nonsaturation conditions for all cells. The kinetic differences between perfect and nonperfect duplexes in the gel allow further improvement in the discrimination through additional washing at low temperature after hybridization.
Article
A microchip method has been developed for massive and parallel thermodynamic analyses of DNA duplexes. Fluorescently labeled oligonucleotides were hybridized with oligonucleotides immobilized in the 100 × 100 × 20 µm gel pads of the microchips. The equilibrium melting curves for all microchip duplexes were measured in real time in parallel for all microchip duplexes. Thermodynamic data for perfect and mismatched duplexes that were obtained using the microchip method directly correlated with data obtained in solution. Fluorescent labels or longer linkers between the gel and the oligonucleotides appeared to have no significant effect on duplex stability. Extending the immobilized oligonucleotides with a four-base mixture from the 3′-end or one or two universal bases (5-nitroindole) from the 3′- and/or 5′-end increased the stabilities of their duplexes. These extensions were applied to increase the stabilities of the duplexes formed with short oligonucleotides in microchips, to significantly lessen the differences in melting curves of the AT- and GC-rich duplexes, and to improve discrimination of perfect duplexes from those containing poorly recognized terminal mismatches. This study explored a way to increase the efficiency of sequencing by hybridization on oligonucleotide microchips.
Article
alpha-Amino-3-hydroxy-5-methylisoxazole-4-propionate (AMPA) receptor subunit (GluR1-4) mRNAs expressed by single neurons in rat hippocampal cultures were quantified by single-cell RT-PCR using an internal standard RNA after whole-cell patch-clamp recording. The internal standard RNA, derived from GluR2 with a single nucleotide substitution, was reverse-transcribed and PCR-amplified with the same efficiency as GluR1-4 mRNAs. The mean mRNA numbers harvested in vitro from pyramidal-like neurons on day 9 were 1150 +/- 324 molecules of GluR1, 1080 +/- 273 molecules of GluR2, 100 +/- 20 molecules of GluR3, and 50 +/- 10 molecules of GluR4 (mean +/- SEM, n = 12). In a non-pyramidal neuronal population that expresses AMPA receptors characterized by high Ca(2+) permeability, the numbers of GluR1, GluR3 and GluR4 mRNA molecules harvested per cell were 354 +/- 64, 25 +/- 17 and 168 +/- 36, respectively (n = 8). The GluR2 mRNA was not detected in this cell type. The calculated ratio of AMPAR mRNA molecules per total mRNA molecules was 1/240 in pyramidal-like neurons (1/500 for GluR2), being in the range obtained with total RNA from rat forebrain and cerebellum (1/170 and 1/380, respectively). Finally, our results indicated that the proportion of GluR1-4 mRNA located in neurites reached approximately 60% in pyramidal-like neurons. However, we found no evidence of preferential subcellular distribution of a given subunit.
Article
Hybridization kinetics were found to be significantly different for specific and non-specific binding of labeled cRNA to surface-bound oligonucleotides on microarrays. We show direct evidence that in a complex sample specific binding takes longer to reach hybridization equilibrium than the non- specific binding. We find that this property can be used to estimate and to correct for the hybridization contributed by non-specific binding. Useful applications are illustrated including the selection of superior oligonucleotides, and the reduction of false positives in exon identification.