Sei sulla pagina 1di 8

S188 Electrophoresis 2009, 30, S188–S195

Celebrating 30 years
Nancy C. Stellwagen Electrophoresis of DNA in agarose gels,
Department of Biochemistry, polyacrylamide gels and in free solution
University of Iowa, Iowa City, IA,
USA
This review describes the electrophoresis of curved and normal DNA molecules in
agarose gels, polyacrylamide gels and in free solution. These studies were undertaken to
Received January 23, 2009 clarify why curved DNA molecules migrate anomalously slowly in polyacrylamide gels
Revised January 23, 2009 but not in agarose gels. Two milestone papers are cited, in which Ferguson plots were
Accepted March 9, 2009
used to estimate the effective pore size of agarose and polyacrylamide gels. Subsequent
studies on the effect of the electric field on agarose and polyacrylamide gel matrices,
DNA interactions with the two gel matrices, and the effect of curvature on the free
solution mobility of DNA are also described. The combined results suggest that the
anomalously slow mobilities observed for curved DNA molecules in polyacrylamide
gels are primarily due to preferential interactions of curved DNAs with the poly-
acrylamide gel matrix; the restrictive pore size of the matrix is of lesser importance. In
free solution, DNA mobilities increase with increasing molecular mass until leveling off
at a plateau value of (3.1770.01)  10 4 cm2/V s in 40 mM Tris-acetate-EDTA buffer at
201C. Curved DNA molecules migrate anomalously slowly in free solution as well as in
polyacrylamide gels, explaining why the Ferguson plots of curved and normal DNAs
containing the same number of base pairs extrapolate to different mobilities at zero
gel concentration.

Keywords:
Agarose gels / Capillary electrophoresis / DNA / Free solution mobility /
Polyacrylamide gels DOI 10.1002/elps.200900052

1 Historical overview fragments of known size [33, 34], it became apparent that
the separation of DNA fragments by molecular mass
The study of DNA electrophoresis began in 1964, when depended on the gel matrix in which the separation was
three groups of investigators [1–5] measured the mobility in carried out [33, 35]. Electron microscopy experiments [36]
free solution using moving boundary methods. They found indicated that the mobilities observed in agarose gels
that the mobility was independent of size for DNA accurately reflected DNA molecular mass, whereas the
molecules larger than 400 base pairs (bp) [5], and varied mobilities observed in polyacrylamide gels did not. More
with ionic strength [3, 5] and the identity and valence of the detailed studies [37, 38], carried out with restriction
cation in the background electrolyte [2, 3]. At about the same fragments ranging from 40 to 4000 bp in size, showed that
time, inspired by the separation of proteins in synthetic gel DNA mobilities decreased monotonically with increasing
matrices [6–10], other investigators began to use similar molecular mass in agarose gels, but that 25% of the same
matrices to separate RNA [11–19] and DNA molecules [15, fragments migrated anomalously slowly in polyacrylamide
20–24] by molecular mass. The separation matrices included gels.
agar [11, 18, 20, 21], agarose [22, 23], polyacrylamide [13, 14, Many investigators then turned their attention to the
19, 24–28] and composite agarose-acrylamide [16, 17] gels. physical properties of DNA that could be responsible for the
As electrophoretic methods were improved by the purifica- anomalously slow mobilities observed for certain restriction
tion of agarose [29–31] and the use of slab gels instead of fragments in polyacrylamide gels. It was soon discovered
tube gels [25, 32], and as the discovery of restriction that the largest mobility anomalies were observed for DNA
enzymes allowed the preparation of monodisperse DNA molecules containing A-tracts, runs of 4–6 contiguous
adenine residues repeated about every 10 bp, in phase with
the helix repeat [39–41]. Other studies, using a variety of
Correspondence: Dr. Nancy C. Stellwagen, Department of
techniques [42–48], showed that DNA molecules that
Biochemistry, University of Iowa, Iowa City, IA 52242, USA
E-mail: nancy-stellwagen@uiowa.edu migrated anomalously slowly in polyacrylamide gels had
Fax: 11-319-335-9570 helix backbones that were curved, not straight. Because

& 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.electrophoresis-journal.com
Electrophoresis 2009, 30, S188–S195 Nucleic acids S189

A N A B N A differences in the two gel media are illustrated in Fig. 1,


where the mobilities of monomers, dimers, trimers and
higher multimers of normal (N) and anomalous (A) 167-bp
DNA fragments are compared. As shown in the left panel,
normal (N) and anomalous (A) multimers of equal size
migrate with equal mobilities in 2% agarose gels, as
expected because of their equal molecular masses. However,
as shown in the right panel, the anomalous multimers (A)
migrate more slowly than their normal counterparts (N) in
polyacrylamide gels containing 5.7%T and 1.5%C. Similar
results are observed for polyacrylamide gels with different
compositions [51–53].
Several obvious questions therefore arise. Do agarose
and polyacrylamide gels, on average, have different pore
sizes and are these differences responsible for the anom-
alously slow mobilities observed in polyacrylamide gels?
2% agarose 5.7%T, 1.5%C polyacrylamide
Does the electric field affect agarose and polyacrylamide gels
Figure 1. Electrophoresis of normal and anomalous DNA differently, and can these differences explain the results? Do
fragments in: (A), 2.0% agarose gels; and (B), 5.7%T, 1.5%C DNA molecules interact with agarose and polyacrylamide
polyacrylamide gels. Monomers of normal (N) and anomalous
(A) DNA restriction fragments containing 167 bp were ligated
gels during electrophoresis and, if so, do curved and normal
(separately) to create multimers of various sizes. From bottom to DNA molecules interact with the two matrices differently?
top, successive bands in each lane of each gel correspond to To answer these questions, we have undertaken detailed
monomers, dimers, trimers, 4-mers, 5-mers and higher multi- studies of DNA electrophoresis in agarose and poly-
mers of the normal (N) and anomalous (A) DNAs. The effective
acrylamide gels. Two of these studies [52, 53], which were
pore radius of the agarose gel is estimated to be 51 nm [49]; the
effective pore radius of the polyacrylamide gel is estimated to be designed to measure apparent gel pore size, are being cited
132 nm [50]. in this issue of Electrophoresis as milestone papers:

curved DNA molecules have shorter end-to-end lengths and These papers have been cited 59 times and 54 times,
larger cross-sectional areas than linear molecules containing respectively, through 2008. The results obtained in the
the same number of bp, curved DNAs would require larger milestone papers and related studies, along with subsequent
pores to migrate through the gel matrix, behaving electro- work on the effect of the electric field on agarose and poly-
phoretically as though they were larger than their true sizes acrylamide gels and interactions of the two gel matrices with
[40, 47]. curved and normal DNA molecules, are described in the
An equally important question is why curved, A-tract- next three sections of this review. The following section
containing DNA molecules migrate anomalously slowly in describes the mobility of DNA in free solution. A brief
polyacrylamide gels, but not in agarose gels. The mobility summary concludes the paper.

& 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.electrophoresis-journal.com
S190 N. C. Stellwagen Electrophoresis 2009, 30, S188–S195

3
3

0.5

Mobility x 104, cm2/Vs


1 1.6
µ x 104, cm2/Vs

0.17
3.1 0.3
0.33

5.1
0.3 0.67
0.1

1.34
9.2
0.03
0.1
0.0 0.4 0.8 1.2 1.6 0 4 8 12
%A %T
Figure 2. Ferguson plots observed for normal DNA molecules in
Figure 3. Ferguson plots observed for normal (D) and anom-
agarose gels. The logarithm of the mobility, extrapolated to zero
alous (o) DNA fragments in polyacrylamide gels. The logarithm
electric field strength at each gel concentration, is plotted as a
of the mobility is plotted as a function of polyacrylamide
function of agarose concentration, %A. The lines were drawn by
concentration, %T, in gels containing 3%C. The normal and
linear regression; the size of each DNA, in kbp, is given beside
anomalous fragments contained one, two, four or eight mono-
each line. Adapted from [52] with permission.
mers, from top to bottom; the size of each pair of fragments, in
kbp, is indicated beside each pair of lines. All lines were drawn
by linear regression. Adapted from [71] with permission.
2 Apparent gel pore size

2.1 Agarose gels models of DNA gel electrophoresis [56, 60] tend to be
twofold smaller than those determined by Ferguson plot
Agarose is an alternating copolymer of 1,3-linked b-D- methods, whereas the gel pore radii measured by NMR [61]
galactose and 1,4-linked 3,6-anhydro-a-L-galactose, infre- or atomic force microscopy [62] are twofold larger. One
quently substituted with carboxylate, pyruvate and/or sulfate could argue that the gel pore radii determined by NMR or
residues [29–33]. Agarose molecules in solution have a atomic force microscopy methods are more accurate,
random coil structure at high temperatures [33–35]. Upon because the values determined by electrophoretic methods
cooling, the agarose chains form helical fiber bundles held represent the subset of pores that are accessible to the
together by noncovalent hydrogen bonds; gelation occurs at migrating DNA molecules. However, the average size of the
still lower temperatures when the fiber bundles become electrophoretically accessible pores in a given gel matrix is
linked together in ‘‘junction zones’’ by the formation of probably more relevant for interpreting gel electrophoresis
additional hydrogen bonds [35–40]. Strand partner exchange experiments.
occurs in the junction zones by hydrogen bond rearrange-
ments [32, 33, 38].
The effective pore size of agarose gels can be estimated 2.2 Polyacrylamide gels
from Ferguson plots (log mobility versus gel concentration
[54]) of DNA molecules of different sizes. Assuming a Polyacrylamide gels are chemically cross-linked gels formed
Gaussian distribution of pore sizes, the median pore radius by the reaction of acrylamide with a bifunctional cross-
of the gel in which the mobility of a given DNA molecule is linking agent such as Bis. The composition of the gel is
reduced to one-half its mobility at zero gel concentration is given by %T, the total w/v concentration of acrylamide plus
equal to the radius of gyration of that DNA [52, 55–57]. cross-linker, and %C, the w/w percentage of cross-linker
Ferguson plots, such as those illustrated in Fig. 2, were included in %T. Polyacrylamide gels are polydisperse in
measured for normal DNA molecules ranging in size from structure, because Bis polymerizes with itself more rapidly
0.5 to 12.2 kbp pairs in the first milestone paper. The than with acrylamide [62–69]. For this reason, polyacryla-
median pore radius of a 1% agarose gel was found to be mide gels consist of highly cross-linked, Bis-rich nodules
100 nm [52], similar to the pore sizes obtained in other linked together by sparsely cross-linked, relatively acryla-
electrophoretic studies [49, 56–59]. However, the estimated mide-rich fibers [63, 68, 69].
gel pore radius depends somewhat on the method used to Because of the structural heterogeneity of the poly-
determine it. Agarose gel pore radii estimated from lattice acrylamide gel matrix, the effective pore size determined by

& 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.electrophoresis-journal.com
Electrophoresis 2009, 30, S188–S195 Nucleic acids S191

electrophoretic methods depends on the size of the analyte. stand for 24 h before use [79]. Hence, the oriented agarose
If proteins are used as the analytes, the apparent pore size gel fibers and fiber bundles become randomized upon
corresponds to the pores in the Bis-rich nodules [70]. Larger standing, presumably by rearrangement of the hydrogen
analytes, such as DNA, do not ‘‘see’’ the small pores in the bonds in the junction zones. DNA mobilities in such
nodules and instead are retarded by migration through the randomized gels are identical to those observed in gels that
acrylamide-rich fibers [70]. In the second milestone paper have not been subjected to pre-electrophoresis [79].
cited above [53], Ferguson plots were measured for DNA Polyacrylamide gels, which are chemically cross-linked,
molecules ranging in size from 123 to 1600 bp in poly- are distorted somewhat by the electric field, but significant
acrylamide gels containing 3.5% to 10.5%T and 3% Bis. orientation of the gel fibers does not occur [77]. For this
Typical examples of the Ferguson plots obtained for multi- reason, DNA mobilities observed in polyacrylamide gels are
mers of 167-bp curved and normal DNA fragments are essentially independent of the electric field strength used for
illustrated in Fig. 3. The Ferguson plots obtained for the electrophoresis [38, 51].
curved multimers are steeper in slope and extrapolate to
lower mobilities at zero gel concentration than the Ferguson
plots of their normal counterparts. Analysis of the Ferguson 4 DNA interactions with agarose and
plots indicates that the effective pore radius ranges from 20 polyacrylamide gel matrices
to 140 nm in gels containing 3.5% to 10.5%T and 0.5% to
10%C [50, 53]. Similar gel pore radii have been observed by 4.1 Agarose gels
scanning and transmission electron microscopy [72–74].
However, NMR relaxation experiments [60] have suggested According to the Ogston–Rodbard–Chrambach theory of gel
that the average pore radius is fivefold smaller, possibly electrophoresis [80, 81], the mobility of a polyelectrolyte in a
because this technique measures an average of the nodule gel matrix is determined by the fractional volume of the gel
and gel fiber pore radii. that is accessible to the migrating macromolecules. Hence,
Ferguson plots are expected to extrapolate to the free
solution mobility of an analyte at zero gel concentration. For
3 Effect of the electric field on agarose and DNA, the Ferguson plots should extrapolate to a common
polyacrylamide gels intercept at zero gel concentration, because the free solution
mobilities of DNA molecules larger than 400 bp are
The mobilities observed for DNA molecules in agarose gels independent of molecular mass [5, 82]. However, as shown
are highly dependent on the electric field applied to the gel for agarose gels in Fig. 2, the intercepts at zero gel
[52, 75], most likely because the electric field disrupts the concentration decrease monotonically with increasing
hydrogen bonds in the junctiones zones, allowing the gel DNA size. Hence, DNA molecules migrating in agarose
fibers and fiber bundles to orient in the electric field [76, 77]. gels appear to be retarded by a molecular mass-dependent
The oriented gel fibers and fiber bundles are very large, mechanism that occurs in addition to sieving. This
ranging up to 22 mm in length [76]. Surprisingly, the gel molecular mass-dependent effect is most likely the transient
fibers and fiber bundles orient in the perpendicular interaction of the DNA molecules with the agarose gel fibers
direction when the electric field is reversed in polarity [77]. during electrophoresis. Such an interaction is not surpris-
The resulting ‘‘flip-flop’’ orientation and reorientation of ing, because DNA is a highly negatively charged polyelec-
agarose fibers and fiber bundles in reversing electric fields trolyte and agarose molecules are known to bind anions [83,
provides a mechanism for creating transient pores in the gel 84]. Other studies have also indicated that DNA molecules
matrix, allowing very large DNA molecules to migrate interact with the agarose gel matrix during electro-
through the gel during pulsed field gel electrophoresis phoresis [review: 85]. If the mobilities observed at zero
[77, 78]. agarose gel concentration are extrapolated linearly to zero
The orientation of the agarose gel fibers and fiber DNA molecular mass, the resulting mobility is
bundles also affects electrophoresis in unidirectional electric (3.070.1)  10 4 cm2/V s in 40 mM Tris-acetate-EDTA
fields. If an agarose gel is pre-electrophoresed in a direction buffer [86], very close to the mobility observed in free
perpendicular to the eventual direction of electrophoresis, solution in the same background electrolyte [82].
linear DNA molecules travel in lanes skewed toward the side
of the gel, as though the pre-electrophoresis had created
pores or channels in that direction [79]. The lanes gradually 4.2 Polyacrylamide gels
straighten out and become aligned in the parallel direction
as electrophoresis is continued, presumably because the gel The Ferguson plots observed for DNA molecules in
fibers and fiber bundles gradually become oriented in the polyacrylamide gels extrapolate to very different mobilities
new field direction. Lanes skewed toward the side of the gel at zero gel concentration, as shown in Fig. 3 [53, 70, 71]. If
are not observed if a gel is pre-electrophoresed in the the mobilities observed at zero gel concentration are
direction in which electrophoresis will be carried out, or if a extrapolated linearly to zero DNA molecular mass, the free
gel oriented in the perpendicular direction is allowed to solution mobility of DNA is calculated to be

& 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.electrophoresis-journal.com
S192 N. C. Stellwagen Electrophoresis 2009, 30, S188–S195

(3.170.1)  10 4 cm2/V s in 40 mM Tris-acetate-EDTA density is essentially constant. The early moving boundary
buffer, equal within experimental error to the value obtained electrophoresis measurements of DNA solutions [5], using
from extrapolation of the Ferguson plots observed in agarose relatively polydisperse DNA samples, were consistent with
gels [86]. Hence, DNA molecules electrophoresed in this assumption. However, more accurate mobility measure-
polyacrylamide gels are also retarded by a molecular mass- ments carried out by capillary electrophoresis, using
dependent mechanism that occurs in addition to sieving, monodisperse DNA samples, have shown that this assump-
most likely transient interactions of the migrating DNA tion is not correct [82, 89]. The mobilities of small DNA
molecules with the gel matrix [50, 53, 86, 87]. Curved DNA molecules increase with increasing molecular mass, as
molecules appear to be more retarded by this mechanism shown by the open circles in Fig. 4, before leveling off and
than normal DNAs containing the same number of bp, reaching a constant plateau value for DNAs larger than
because the Ferguson plots of curved DNAs have steeper 400 bp. Because theoretical studies have suggested that the
slopes and extrapolate to somewhat lower mobilities at zero mobilities of charged rods should increase with increasing
gel concentration (compare the solid and dashed lines in molecular mass [91, 92], the results in Fig. 4 suggest that
Fig. 3). small DNA molecules are rigid and rod-like in solution.
The relative importance of sieving and gel matrix DNAs larger than 400 bp begin to adopt a random coil
interactions to the anomalously slow mobilities observed for conformation and migrate as free-draining coils [5, 82, 89,
curved DNA molecules in polyacrylamide gels can be eval- 93, 94]; the mobility is then dependent only on the effective
uated by measuring the mobilities of curved and normal charge per unit mass. The plateau mobility observed for
DNAs in gels in which the pore size is varied by changing DNA molecules Z400 bp in size in 40 mM Tris-acetate-
%T at constant %C (the usual method of changing gel pore EDTA buffer is (3.1770.01)  10 4 cm2/V s [90], essentially
size) and by changing %C at constant %T. If sieving effects equal to the mobilities calculated by the double extrapolation
are the primary factor contributing to the anomalously slow of the Ferguson plots obtained in agarose and polyacryla-
mobilities, the mobility anomalies should be independent of mide gels to zero gel concentration and zero DNA molecular
the method used to vary the gel pore size. However, if mass. Hence, the double extrapolation procedure leads to
preferential interactions of curved DNA molecules with the reasonably accurate values of the free solution mobility of
polyacrylamide gel fibers are responsible for the anom- DNA.
alously slow mobilities, the mobility anomalies should
correlate with the acrylamide concentration in the gel, not
the apparent gel pore size. When the polyacrylamide gel 5.2 Dependence of free solution mobility on DNA
pore size is decreased by increasing %T at constant %C, the curvature
anomalously slow mobilities of the curved DNA molecules
increase with decreasing gel pore radius [53, 71, 87], as The free solution mobilities of curved DNA molecules (as
though sieving effects were responsible for the mobility well as the mobilities observed in polyacrylamide gels)
anomalies [39, 88]. However, when the gel pore size is varied depend on the extent of curvature of the helix backbone [90].
by changing %C at constant %T, the anomalously slow This effect is illustrated by the closed circles in Fig. 4, which
mobilities are independent of gel pore radius as long as the correspond to 199-bp DNA fragments containing 1–5
apparent gel pore radius is larger than the DNA radius of
gyration [87]. If the DNA radius of gyration is larger than the
apparent gel pore radius, the anomalously slow mobilities of 3.20
the curved DNAs decrease with increasing gel pore radius
until the gel pore radius becomes equal to the DNA radius
Mobility x 10 , cm /Vs

of gyration, after which the mobility anomalies become


2

3.15
constant and independent of gel pore size. Hence, the
4

interaction of curved DNA molecules with the acrylamide-


rich gel fibers appears to be the primary cause of the
anomalously slow mobilities observed for curved DNAs in 3.10
polyacrylamide gels; sieving effects are of secondary
importance.

3.05
0 300 600 900
5 DNA mobility in free solution
Number of Base Pairs

5.1 Dependence of free solution mobility on DNA Figure 4. Free solution mobility of normal (open circle) DNA
molecular mass fragments of various sizes in 40 mM Tris-acetate-EDTA buffer,
plotted as a function of the number of bp in each fragment. The
closed circles indicate the mobilities of curved 199-bp fragments
The free solution mobility of DNA is often assumed to be containing 1–5 A-tracts in a curvature module located in the
independent of molecular mass, because the linear charge center of each fragment. Adapted from [90] with permission.

& 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.electrophoresis-journal.com
Electrophoresis 2009, 30, S188–S195 Nucleic acids S193

A-tracts (from top to bottom) located in a ‘‘curvature monovalent cations [98]. Monovalent cation binding by DNA
module’’ [95] in the center of each fragment. The fragment A-tracts would decrease the effective charge of curved DNAs,
with a single A-tract in the curvature module has the decreasing the observed mobility [90, 96, 97]. However, a
conformation of normal DNA [95] and exhibits a mobility detailed discussion of cation binding by DNA A-tracts is
that falls on the same smooth curve describing the another story for another day.
mobilities of other normal DNA fragments. However,
fragments containing 2–5 A-tracts in the curvature module, The work carried out in the author’s laboratory was
which become increasingly more curved as the number of supported in part by grants GM29690, EB002808 and
A-tracts increases [95, 96], migrate increasingly more slowly GM61009 from the National Institutes of Health and grant
in free solution. The anomalously slow mobilities observed CHE-0748271 from the Analytical and Surface Chemistry
for curved DNAs in free solution are due in part to the Program of the National Science Foundation. The expert tech-
dependence of the frictional factor on molecular shape [96] nical assistance of various coworkers is also gratefully acknowl-
and in part to the preferential binding of monovalent cations edged.
by DNA A-tracts, which decreases the effective net charge
[96, 97]. The author has declared no conflict of interest.

6 Concluding remarks
7 References
The milestone papers cited in this review were the first of a
series of papers designed to better understand why curved [1] Ross, P. D., Biopolymers 1964, 2, 9–14.
DNA molecules migrate anomalously slowly in polyacryla- [2] Ross, P. D., Scruggs, R. L., Biopolymers 1964, 2,
mide gels, but not in agarose gels. The milestone papers 231–236.
showed that the restrictive pore size of the polyacrylamide [3] Ross, P. D., Scruggs, R. L., Biopolymers 1964, 2, 79–89.
gel matrix is not the primary cause of the mobility [4] Costantino, L., Liquori, A. M., Vitagliano, V., Biopoly-
anomalies, because gels with equivalent apparent pore sizes mers 1964, 2, 1–8.
can be cast in both gel matrices. The milestone papers also [5] Olivera, B. M., Baine, P., Davidson, N., Biopolymers
suggested that DNA molecules interact with agarose and 1964, 2, 245–257.
polyacrylamide gels during electrophoresis, because the [6] Smithies, O., Biochem. J. 1955, 61, 629–641.
Ferguson plots do not extrapolate to a common intercept at
[7] Raymond, S., Nakamichi, M., Anal. Biochem. 1962, 3,
zero gel concentration. The free solution mobility of DNA 23–30.
can be recovered if the mobilities observed at zero gel
[8] Ornstein, L., Ann. N. Y. Acad. Sci. 1964, 121, 321–349.
concentration are also extrapolated to zero DNA molecular
[9] Tombs, M. P., Anal. Biochem. 1965, 13, 121–132.
mass.
Subsequent work, also described here, showed that the [10] Fawcett, J. S., Morris, C. J. O. R., Sep. Sci. 1966, 1, 9–26.
anomalously slow mobilities observed for curved DNA [11] Bachvaroff, R., McMaster, P. R. B., Science 1964, 143,
molecules in polyacrylamide gels are correlated with the 1177–1179.
acrylamide concentration in the gel matrix, not the gel pore [12] Richards, E. G., Gratzer, W. B., Nature 1964, 204,
size, because the mobility anomalies are observed even in 878–879.
gels in which the effective gel pore radius is larger than the [13] Richards, E. G., Coll, J. A., Gratzer, W. B., Anal.
DNA radius of gyration. Hence, curved DNA molecules Biochem. 1965, 12, 452–471.
appear to have a greater affinity for the acrylamide-rich gel [14] Richards, E. G., Lecanidou, R., Anal. Biochem. 1971, 58,
fibers than normal DNAs containing the same number of 43–71.
bp. Mobility differences between curved and normal DNA [15] Bishop, D. H. L., Claybrook, J. R., Spiegelman, S.,
molecules containing the same number of bp are not J. Mol. Biol. 1967, 26, 373–387.
observed in agarose gels, most likely because the agarose gel [16] Peacock, A. C., Dingman, C. W., Biochemistry 1968, 7,
fibers are oriented by the electric field, creating transient 668–674.
pores in the gel matrix. Because the orienting fibers and [17] Dahlberg, A. E., Dingman, C. W., Peacock, A. C., J. Mol.
fiber bundles are very large, the transient pores are insen- Biol. 1969, 41, 139–147.
sitive to small differences in DNA conformation. [18] Tsanev, R., Staynov, D., Kokileva, L., Mladenova, I.,
The free solution mobility of DNA increases with Anal. Biochem. 1969, 30, 66–85.
increasing molecular mass until leveling off at a constant [19] Philippsen, P., Zachau, H. G., Biochim. Biophys. Acta
plateau value for fragments larger than 400 bp. Curved 1972, 277, 523–538.
DNA molecules that migrate anomalously slowly in poly- [20] Thorne, G. V., Virology 1966, 29, 234–239.
acrylamide gels also migrate anomalously slowly in free [21] Thorne, G. V., J. Mol. Biol. 1967, 24, 203–211.
solution, most likely because the curved DNAs usually [22] Aaij, C., Borst, P., Biochim. Biophys. Acta 1972, 269,
contain multiple A-tracts and A-tracts are known to bind 192–200.

& 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.electrophoresis-journal.com
S194 N. C. Stellwagen Electrophoresis 2009, 30, S188–S195

[23] Helling, R. B., Goodman, H. M., Boyer, H. W., [56] Slater, G. W., Rousseau, J., Noolandi, J., Turmel, C.,
J. Virology 1974, 14, 1235–1244. Lalande, M., Biopolymers 1988, 27, 509–524.
[24] Maniatis, T., Jeffrey, A., van de Sande, H., Biochemistry [57] Stellwagen, N. C., Holmes, D. L., Electrophoresis 1990,
1975, 14, 3787–3794. 11, 649–652.
[25] Studier, F. W., J. Mol. Biol. 1973, 79, 237–248. [58] Serwer, P., Hayes, S. J., Anal. Biochem. 1986, 158,
[26] Jeppesen, P. G. N., Anal. Biochem. 1974, 58, 195–207. 72–78.
[27] Flint, D. H., Harrinigton, R. E., Biochemistry 1972, 11, [59] Slater, G. W., Treurniet, J. R., J. Chromatogr. A 1997,
4858–4864. 772, 39–48.
[28] Elson, E., Jovin, T. M., Anal. Biochem. 1969, 27, [60] Chui, M. M., Phillips, R. J., McCarthy, M. J., J. Colloid
193–204. Interface Sci. 1995, 174, 336–344.
[29] Hickson, T. G. L., Polson, A., Biochim. Biophys. Acta [61] Maaloum, M., Pernodet, N., Tinland, B., Electrophoresis
1968, 165, 43–58. 1998, 19, 1606–1610.
[30] Duckworth, M., Yaphe, W., Carbohydr. Res. 1971, 16, [62] Hecht, A.-M., Duplessix R., Geissler E., Macromolecules
189–197. 1985, 18, 2167–2173.
[31] Duckworth, M., Yaphe, W., Carbohydr. Res. 1971, 16, [63] Baselga, J., Hernández-Fuentes, I., Piérola, I. F.,
435–445. Llorente, M. A., Macromolecules 1987, 20,
3060–3065.
[32] McDonell, M. W., Simon, M. N., Studier, F. W., J. Mol.
Biol. 1977, 110, 119–146. [64] Baselga, J., Llorente, M. A., Nieto, J. L., Hernández-
[33] Roberts, R. J., Crit. Rev. Biochem. 1976, 4, 123–164. Fuentes, I., Piérola, I. F., Eur. Polym. J. 1988, 24,
161–165.
[34] Modrich, P., Q. Rev. Biophys. 1979, 12, 315–369.
[65] Baselga, J., Llorente, M. A., Hernández-Fuentes, I.,
[35] Zeiger, R. S., Salomon, R., Dingman, C. W., Peacock, A. Piérola, I. F., Eur. Polym. J. 1989, 25, 477–480.
C., Nature New Biol. 1972, 238, 65–69.
[66] Righetti, P. G., Caglio, S., Electrophoresis 1993, 14,
[36] Thomas, M., Davis, R. W., J. Mol. Biol. 1975, 91, 573–582.
315–328.
[67] Weiss, N., Van Vleit, T., Silberberg, A., J. Polym. Sci.:
[37] Stellwagen, N. C., Biochemistry 1983, 22, 6180–6185. Polym. Phys. Ed. 1979, 17, 2220–2240.
[38] Stellwagen, N. C., Biochemistry 1983, 22, 6186–6193.
[68] Joosten, J. G. H., McCarthy, J. L., Pusey, P. M., Macro-
[39] Wu, H.-M., Crothers, D. M., Nature 1984, 308, 509–513. molecules 1991, 24, 6690–6699.
[40] Koo, H.-S., Wu, H.-M., Crothers, D. M., Nature 1986, 320, [69] Heuer, D. M., Saha, S., Archer, L. A., Biopolymers 2003,
501–506. 70, 471–481.
[41] Hagerman, P. J., Proc. Natl. Acad. Sci. USA 1984, 81, [70] Stellwagen, N. C., Electrophoresis 1998, 19, 1542–1547.
4632–4636.
[71] Stellwagen, N. C., Electrophoresis 2006, 27, 1163–1168.
[42] Hagerman, P. J., Biochemistry 1985, 24, 7033–7037.
[72] Rüchel, R., Brager, M. C., Anal. Biochem. 1975, 68,
[43] Levene, S. D., Wu, H.-M., Crothers, D. M., Biochemistry 415–428.
1986, 25, 3988–3995.
[73] Rüchel, R., Steere, R. L., Erbe, E. F., J. Chromatogr. 1978,
[44] Stellwagen, N. C., Biopolymers 1991, 31, 1651–1667. 166, 563–575.
[45] Muzard, G., Théveny, B., Révet, B., EMBO J. 1990, 9, [74] Hsu, T.-P., Cohen, C., Polymer 1989, 174, 336–344.
1289–1298.
[75] Stellwagen, N. C., Biopolymers 1985, 24, 2243–2255.
[46] Griffith, J., Bleyman, M., Rauch, C. A., Kitchen, P. A.,
Englund, P. T., Cell 1986, 46, 717–724. [76] Stellwagen, J., Stellwagen, N. C., Biopolymers 1994, 34,
187–201.
[47] Macdonald, D., Herbert, K., Zhang, X., Polgruto, T., Lu, P.,
J. Mol. Biol. 2001, 306, 1081–1098. [77] Stellwagen, J., Stellwagen, N. C., Biopolymers 1994, 34,
1259–1273.
[48] Barbič, A., Zimmer, D. P., Crothers, D. M., Proc. Natl.
Acad. Sci. USA 2003, 100, 2369–2373. [78] Stellwagen, N. C., Stellwagen, J., Electrophoresis 1993,
14, 355–368.
[49] Stellwagen, N. C., Electrophoresis 1992, 13, 601–603.
[79] Holmes, D. L., Stellwagen, N. C., J. Biomol. Struct. Dyn.
[50] Holmes, D. L., Stellwagen, N. C., Electrophoresis 1991,
1989, 7, 311–327.
12, 612–619.
[80] Ogston, A. G., Trans. Faraday Soc. 1958, 54, 1754–1757.
[51] Stellwagen, A., Stellwagen, N. C., Biopolymers 1990, 30,
309–324. [81] Rodbard, D., Chrambach, A., Proc. Natl. Acad. Sci. USA
1970, 65, 970–977.
[52] Holmes, D. L., Stellwagen, N. C., Electrophoresis 1990,
11, 5–15. [82] Stellwagen, N. C., Gelfi, C., Righetti, P. G., Biopolymers
1997, 42, 687–703.
[53] Holmes, D. L., Stellwagen, N. C., Electrophoresis 1991,
12, 253–263. [83] Piculell, L., Nilsson, S., J. Phys. Chem. 1989, 93, 5596–5601.
[54] Ferguson, K. A., Metabolism 1964, 13, 985–1002. [84] Piculell, L., Nilsson, S., J. Phys. Chem. 1989, 93, 5602–5611.
[55] Slater, G. W., Guo, H. L., Electrophoresis 1996, 17, [85] Stellwagen, N. C., Stellwagen, E., J. Chromatogr. A
1407–1415. 2009, 1216, 1917–1929.

& 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.electrophoresis-journal.com
Electrophoresis 2009, 30, S188–S195 Nucleic acids S195

[86] Strutz, K., Stellwagen, N. C., Electrophoresis 1998, 19, [92] Mohanty, U., Stellwagen, N. C., Biopolymers 1999, 49,
635–642. 209–214.
[87] Stellwagen, N. C., Electrophoresis 1997, 18, 34–44. [93] Stellwagen, E., Stellwagen, N. C., Electrophoresis 2002,
[88] Marini, J. C., Effron, P. N., Goodman, T. C., Singleton, C. K., 23, 2794–2803.
Wells, R. D., Wartell, R. M., Englund, P. T., J. Biol. Chem. [94] Viovy, J.-L., Rev. Modern Phys. 2000, 72, 813–872.
1984, 259, 8974–8979.
[95] Lu, Y. J., Weers, B. D., Stellwagen, N. C., Biophys. J.
[89] Stellwagen, E., Lu, Y. J., Stellwagen, N. C., Biochemistry 2005, 88, 1191–1206.
2003, 42, 11745–11750.
[96] Lu, Y. J., Stellwagen N. C., Biophys. J. 2008, 94, 1710–1725.
[90] Stellwagen, E., Lu, Y. J., Stellwagen, N. C., Nucleic
Acids Res. 2005, 33, 4425–4432. [97] Stellwagen, N. C., Magnúsdóttir, S., Gelfi, C., Righetti, P. G.,
[91] Völkel, A. R., Noolandi, J., J. Chem. Phys. 1995, 102, J. Mol. Biol. 2001, 305, 1025–1033.
5506–5511. [98] Egli, M., Chem. Biol. 2002, 9, 177–186.

Nancy C. Stellwagen received her Ph.D. in physical chemistry from the University of California,
Berkeley, and did post-doctoral studies in the laboratory of Bruno H. Zimm at the University of
California, San Diego. She is currently Adjunct Professor of Biochemistry and Research
Scientist at the University of Iowa, Iowa City, IA. Her laboratory has focused on the
conformation and dynamics of DNA molecules in solution, the mechanism of DNA
electrophoresis in agarose and polyacrylamide gels, and the use of capillary electrophoresis to
study cation binding to DNA.

& 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.electrophoresis-journal.com

Potrebbero piacerti anche