Separation of Double-Stranded and Single-Stranded DNA in Polymer Solutions: I. Mobility and Separation Mechanism

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

1962 Electrophoresis 1999, 20, 1962±1977

Christoph Heller Separation of double-stranded and single-stranded


Abteilung Lehrach, DNA in polymer solutions:
Max-Planck-Institut für I. Mobility and separation mechanism
molekulare Genetik,
Berlin-Dahlem, Germany We have studied the separation of single-stranded and double-stranded DNA in a
matrix of entangled, linear poly-N,N-dimethylacrylamide. Our results give better insight
into the mechanisms involved during separations in polymer solutions. The depend-
ence of different parameters on DNA size, electric field, pore size and the polymer
chain length are evaluated and compared to theoretical predictions. Striking differ-
ences between experimental data and predicted scaling laws are found. Our data
should help to optimize DNA separation in capillary electrophoresis and to improve
existing models for DNA separation in porous matrices.

Keywords: Separation matrix / Polymer solution / DNA separation / Electrophoresis theory /


Capillary electrophoresis EL 3522

1 Introduction matrix, fundamental studies on DNA separation in this


polymer have been undertaken recently [8]. There, we
For DNA separation and other applications, capillary elec- found hints that the separation of ssDNA gives superior
trophoresis (CE) has many advantages over slab gel resolution compared to the same DNA fragments in the
electrophoresis, in particular high speed, high sensitivity double-stranded form (in a matrix of the same pore size).
and the possibility for automation. It also offers full com- This is in agreement with an earlier study on linear poly-
patibility to existing biochemistry, which is not always the acrylamide [9]. The present study is a continuation of our
case in alternative techniques, such as MALDI-MS. earlier work and is aimed at describing this and other
Instead of cross-linked gels, so-called ªentangledº (ªsemi- effects in more detail and getting a better insight into the
diluteº) polymer solutions are now widely in use as sepa- underlying mechanisms.
ration matrix (see [1] for review). Separation in unentan-
gled (ªdiluteº) polymer solutions has also been described
The properties of entangled polymer solutions have been
[2], but is probably irrelevant for most molecular biology
reviewed in several articles (see, e.g. [8, 10]). One impor-
applications, as the resolution is rather low. Therefore, in
tant parameter is the ªpore sizeº or ªmesh sizeº of such a
our study, we concentrate on entangled polymer solu-
network. To estimate this value, we can regard a polymer
tions.
chain segment between two entanglement points as an
independent subunit, which can undergo a so-called ªran-
1.1 Polymers used in CE dom walkº. The volume, which this chain segment can
take, is called a ªblobº and is described by the ªblob sizeº,
Most of the polymers used in CE are modified polysac- xb. Viovy and Duke [11] argued that this ªblob sizeº could
charides or synthetic polymers such as derivatives of be used as a ªpore sizeº of the network. It can be calcu-
acrylamide. One of these is poly-N,N-dimethylacrylamide lated as follows [12]:
(pDMA), which has ideal properties as a separation
matrix, due to its low to medium viscosity and its self-coat- xb = 1.43 Rp (c/c*)±(1+a)/3a (c > c*) (1)
ing ability. Solutions of pDMA are successfully used for
DNA separation, both in capillary electrophoresis [3±6] as with c* being the overlap threshold, i.e., the concentration
well as in microchips [7]. To better characterize this c above which the polymers become entangled, and Rp
being the radius of gyration of the polymer. The parame-
ter a is the exponent of the empirical Mark-Houwink equa-
Correspondence: Dr. Christoph Heller, Max-Planck-Institut für tion, which describes the relationship between the intrin-
molekulare Genetik, Ihnestraûe 73, D-14195 Berlin, Germany
sic viscosity [h] and the molecular mass of the
E-mail: heller@mpimg-berlin-dahlem.mpg.de
Fax: +49-30-8413-1380 polymer:[h] % K Mva. Here, Mv is the ªviscosity average
molecular massº, which is different, but often close to the
weight average molecular mass, Mw [13]. The Mark-Hou-
Abbreviations: BRF, biased reptation with fluctuation; BRM,
biased reptation model; pDMA, poly-N,N-dimethylacrylamide; wink coefficients K and a have been determined for many
TBE, Tris-boric acid-EDTA buffer polymer-solvent systems (see, e.g. [14]). For flexible mol-

 WILEY-VCH Verlag GmbH, 69451 Weinheim, 1999 0173-0835/99/1010-1962 $17.50+.50/0


Electrophoresis 1999, 20, 1962±1977 Separation of double-stranded and single-stranded DNA 1963

ecules, the value of a varies between 0.5 (polymer in an short DNA molecule as a so-called ªequivalent sphereº.
ideal solvent) and 0.8 for so-called ªswollenº chains. Therefore, in the context of DNA separation in entangled
polymer solutions we can only talk about ªOgston-likeº
The so-called Flory-Fox theory relates the radius of gyra- sieving mechanisms. Recently, the Ogston model has
tion to the intrinsic viscosity: [h] = F Rp3/M, with F being come under radical reconsideration [22], and an improved
the Flory constant. Following Thielking and Kulicke [15], model is under development.
this theory, which has been developed for the unper-
turbed state, can also be applied to dissolved polymers in Obviously, sieving models are not valid for the situation
good solvents by introducing expansion coefficients. This when DNA molecules are larger than the pores of the gel
modified Flory viscosity constant takes into account (Rg > x, with x being the pore size of the separation
excluded volume effects by: F = F0 (1±2.63 b + 2.86 b2) matrix), and a number of models based on the so-called
with b = (2a±1)/3 and F0 being 2.5 ´ 1023 mol±1 for reptation concept were developed for this case. In such
uncharged polymers [13, 16]. For ideal chains (a = 0.5; no models, the DNA chain is assumed to thread its way
excluded volume), b becomes zero and we regain F = through a ªtubeº defined by the fibers (for a rigid mesh) or
F0. Following Cottet et al. [16], Eq. (1) can be rewritten to: the ªblobsº (for a flexible network) surrounding it. One of
the models to which experimental data fit best, the
xb % 1.43 (K/63/2F)±1/3a c±(a+1)/3a (1.5/2 p NA)(a+1)/3a ªbiased reptation with fluctuationsº (BRF) model [23], is
(c > c*) (2) an improvement of the original ªbiased reptation modelº
(BRM) [24]. Here, we follow the notation by Viovy [25],
with xb in cm and c in g/mL and NA being the Avogadro who has recently further improved the BRF by taking into
number. Note that Eq. (2) is the more general form of Eq. account the electrohydrodynamic forces acting on the
(11.12) in [10] and Eq. (6) in [17], respectively, with a whole section of DNA in a pore. This replaces the so-
slightly different prefactor. However, note that the equa- called ªlocal forceº idea, which assumes that each seg-
tions mentioned above are based on scaling laws and the ment of the chain (Kuhn segment) is submitted to an elec-
prefactors cannot be given exactly. Equation (2) means tric force.
that the ªpore sizeº of an entangled polymer solution does
not depend on the degree of polymerization but only on Very briefly, in the BRF, two different cases are consid-
its concentration, c, and the nature of the polymer ered: first, when the pore size x of the matrix is larger than
(through K and a). the Kuhn length bD of DNA (xk = x/bD > 1) and second,
when the pore size is smaller (so-called ªtight gelº or ªper-

CE and CEC
sistent chainº situation, xk < 1). The Kuhn length (twice
1.2 Mechanism of DNA movement
the persistence length) is a natural parameter and a
The mechanism of DNA movement in a porous matrix measure for the flexibility of a polymer. Here we have
(gels and entangled polymer solutions) under the influ- introduced a dimensionless parameter, xk, which is the
ence of an electric field has been the subject of intensive pore size normalized to the Kuhn length. In the first case,
research and reviewed in many articles (e.g., [10, 18, which can also be regarded as the ªflexible chainº situa-
19]). In the earliest model, the so-called Ogston-model tion, the BRF describes the electrophoretic mobility of a
[20, 21], the gel is assumed to act as a sieve with a distri- polyelectrolyte as [25]:
bution of pore sizes and the separation is regarded as a
kind of filtration, driven by the electric field. This leads to a m/m0 = xk2 [(1/3Nk)2 + (2 ek/(5+2 a ek xk2))2]1/2 (4)
mobility m proportional to exp(±conc.):
where a represents the ratio between the free mobility
m/m0 = exp[±KR c] (3) and the limiting in-gel mobility of DNA in strong fields and
was introduced phenomenologically to account for the
where m0 is the mobility in pure solvent (free solution mo- saturation of mobility at strong electric fields. In experi-
bility), KR is the retardation coefficient (proportional to (Rg ments and computer simulations it is found to be of the
+ r)2 with r being the radius of the gel fibers and Rg the order of 3 [25]. In Eq. (4), we again introduced two dimen-
radius of gyration of the DNA molecule), and c is the gel sionless parameters, i.e., the DNA size (curvilinear
concentration. The Ogston model was originally devel- length, N) also normalized to the Kuhn length: Nk = N/bD
oped for the movement of spherical objects in a random and the ªscaled electric fieldº ek = hbD2m0E/kT, with h
array of cylindrical obstacles, where the pore size follows being the ªlocal viscosityº (viscosity of pure buffer without
Poisson statistics. In entangled polymer solutions, how- polymer), E the electric field, k the Boltzman constant,
ever, the ªporesº are probably more uniform. Second, it is and T the absolute temperature. Thus, all dimensionless
not clear if one can model a quickly rotating rigid rod-like parameters depend on the properties of the DNA (i.e., the
1964 C. Heller Electrophoresis 1999, 20, 1962±1977

Kuhn length) and not on the properties of the separation The contribution of constraint release to the mobility of
matrix. We have chosen this notation, as the derived DNA is given by:
equations then directly give us the pore size dependence.
Theoretical curves obtained with this equation [25] fit the mCR/m0 % (xb/bD) (Rp/xb)±5 (7)
experimental data of dsDNA mobility in agarose gels [26].
Assuming that the biased reptation and the constraint
Equation (4) gives the well-known behavior described by release are independent processes, the net DNA mobility
all reptation models: For small molecules (below a critical is then given by the sum of both mobilities: mtot = mBRF +
size, Nk*), the first term dominates and the mobility is mCR. Note that the constraint release mobility is depend-
inversely proportional to the DNA size (ªreptation without ent on both polymer concentration and the chain length,
orientationº), but for large DNA and/or high electric fields, but independent of DNA size and applied electric field.
the second term dominates, which means that the mobili- This means that for efficient separation to be achieved in
ty becomes independent of size (plateau mobility) and entangled polymer solution, the constraint release mobili-
separation fails (ªreptation with orientationº). For the criti- ty should be smaller than the BRF component. For large
cal size, the model predicts, that Nk* = ek±1, i.e., independ- pores, this condition can be rewritten to: c/c* > (NkbD/
ent of pore size. This prediction, which is supported by Rp)1/3 [11].
experimental results in gels [26] is the major difference to
earlier models, which predict a xk±1 [23] and a xk±4 de-
pendence [27], respectively. Also, the revised model pre- 1.3 Network dynamics
dicts a mobility scaling with xk2 in the reptation with orien-
The influence of network dynamics on size-based electro-
tation regime, whereas earlier models predict an
phoretic separation has been studied in more detail by
exponent of 3 [23] and 6 [25], respectively.
Cottet et al. [16]. Estimating the reptation time of the
In tight gels (xk < 1), reptation still occurs below a critical matrix polymer, they found that a threshold value exists
size: below which the electrophoretic mobility of the analyte is
influenced by the polymer motion. Above this value (i.e.,
m/m0 ~ 1/3 Nk Nk < Nk* (5) for long chains or at high concentrations), the polymer
reptation ceases to be a parameter influencing the sepa-
but for large DNA (Nk > Nk*), the existence of three differ- ration, and mobility vs. size plots will fall onto the same
ent regimes with the following plateau mobilities are pre- curve. The approach of comparing the lifetime of obsta-
dicted [23, 25]: cles to the renewal time of the analyte is consistent with
the concept of constraint release and allows further
m/m0 ~ ek xk3/2 ek < xk3/2 (6a) insight into the complex behavior of electrophoretically
driven polyelectrolytes in polymer solutions. It is also
m/m0 ~ (ek xk6)2/5 xk±1 > ek > xk3/2 (6b) possible that the polymer chains are actively dragged by
the moving DNA (ªnetwork ruptureº) and indeed such
and: effects have been observed in videomicroscopic studies,
even in semidilute solutions [28, 29]. They can occur dur-
m/m0 ~ ek2 xk4 ek > xk±1 (6c) ing a so-called U-conformation, when a DNA molecule
becomes hooked around a polymer chain. In contrast to
For the critical size, the predictions are Nk* = ek±3/2, ek2/5
gels, often the whole conformation is then seen to move
xk±12/5 and ek±2 xk±4, respectively. The last Eq. (6c) repre-
downfield. This behavior has some similarity to the ªen-
sents the regime when x 5 bD and length fluctuations are
tanglement couplingº process, which has been studied
frozen. In this case we recover the prediction of the origi-
both experimentally and theoretically in dilute solutions [2,
nal BRM. The second regime (Eq. 6b) has been observed
30]. However, to date there is no theoretical treatment of
by Mitnik et al. [28] for dsDNA in entangled hydroxypro-
the combination or reptative movement and polymer
pylcellulose solutions.
dragging, which might describe the DNA movement in
entangled polymer solutions.
Polymer solutions differ from gels in that the physical
entanglements between the chains have a finite lifetime
and we have to take into account the dynamics of the 2 Material and methods
matrix itself. Viovy and Duke [11] have adopted the physi-
cal concept of ªconstraint releaseº (CR) to describe this 2.1 Polymers
phenomenon. It is based on the fact that polymers them- pDMA is synthesized from the monomer, dimethylacryl-
selves move by a random curvilinear diffusion (reptation). amide, by radical polymerization. The chain length can be
Electrophoresis 1999, 20, 1962±1977 Separation of double-stranded and single-stranded DNA 1965

influenced by changing the temperature or by adding radi- 3 Results


cal traps (ªchain transfer reagentsº), such as isopropanol.
The preparation follows the protocol by Chiari et al. [3] as 3.1 Characterization of the separation matrix
described before [8]: N,N-dimethylacrylamide (Polyscien- We have produced the following three pDMA prepara-
ces, Warrington, PA, USA) was dissolved in water to a tions of different chain length, which have been character-
final concentration of 10% w/w and isopropanol was ized by size exclusion chromatography: (i) ªShort chainº
added in various amounts. After degassing (with vacuum pDMA, synthesized in presence of 2% isopropanol at
and ultrasound), 1 mL TEMED per mL solution and 1 mL/ 50oC. Mv = 169 039, Mw = 216 774, polydispersity (Mw/
mL of a 40% ammonium persulfate solution was added. Mn): 4.39, [h] = 139.07 mL/g. (ii) ªMedium chainº pDMA,
After careful mixture, the solution was allowed to react for synthesized in presence of 1% isopropanol at 25oC. Mv =
1 h at 25oC or 50oC, followed by incubation at 25oC over- 477 488, Mw = 603 846, polydispersity (Mw/Mn): 5.29, [h]
night. The reaction product was extensively dialyzed = 326.71 mL/g. (iii) ªLong chainº pDMA, synthesized with-
against water using 25 000 MWCO dialysis membranes out isopropanol at 25oC. Mv = 816 574, Mw = 1 146 319,
(Spectrum, Los Angeles, CA, USA) and lyophilized. polydispersity (Mw/Mn): 5.35, [h] = 424.06 mL/g. Note that
the same polymer preparations were characterized before
by viscosity measurements (Table 1 of [8]), with molecu-
2.2 Electrophoresis lar masses of 490 kDa, 934 kDa and 1536 kDa, respec-
tively. We assume that the measurements with size
For all separations the ABI 310 capillary electrophoresis
exclusion chromatography are more precise than the esti-
device (Perkin Elmer ± ABI, Foster City, CA, USA) was
mations based on intrinsic viscosity alone and therefore
used, equipped with a 47 cm long (36 cm to the detection
we shall use the values given above.
window) 375/100 mm OD/ID fused-silica capillary (Polymi-
cro Technologies, Phoenix, AZ, USA). 0.5 ´ TBE buffer
(44.5 mM Tris, 44.5 mM boric acid, 1.25 mM EDTA, pH 8.3, As the published Mark-Houwink coefficients for pDMA are
condictivity 625 mS) was used as background electrolyte disputable [8], we determined them again by a least-
square fit (not shown) of [h] vs. Mv, using the values
for the polymer (which was only in the capillary) and as
electrode buffer. For denaturing conditions, 4 M urea (final above. We obtained K = 0.023 mL/g and a = 0.72 (r2 =
concentration) was added to the polymer and electro- 0.98), which means that he exponent a is between the
phoresis took place at 50oC; for nondenaturing conditions values given in the literature for 25oC (K = 0.023 mL/g, a
the temperature was set to 25oC (and urea was omitted). = 0.81) and for 40oC (K = 0.02 mL/g, a = 0.65) [14]. The
For the free mobility measurements, the capillary was entanglement threshold of polymer solutions is inversely
dynamically coated by flushing a 2% pDMA solution proportional to the intrinsic viscosity [12]: c* % 1.5/[h].
through the capillary followed by rinsing with buffer (0.5 ´ Taking into account the relation between the viscosity
TBE). This coating proved to be stable for at least ten average molecular mass and the intrinsic viscosity (see
injections. Section 1), we determined c* for all three polymer prepa-
rations by using the molecular masses shown above and
the newly determined Mark-Houwink parameters: c* %
2.3 Samples (1.5/K)Mv±a (data not shown). As a second method, the

As DNA sample we used Sizer 50±500 (Pharmacia, Table 1. Free mobility of ssDNA and dsDNA measured
Uppsala, Sweden), Low Range DNA Standard (pBR322- in capillary electrophoresis with 0% pDMA at
HinfI digest), 100 bp ruler and 500 bp ruler (Bio-Rad, different temperature with and without urea
Hercules, CA, USA), all of them being labeled with fluo-
Buffer ssDNa dsDNA
rescein. For denaturing conditions, 8 mL formamide
(mm/s)/(V/cm) (mm/s)/(V/cm)
(Amresco, Solon, OH, USA) were added to 2 mL DNA,
heated to 90oC for 2 min and chilled on ice before injec- 0.5 ´ TBE at 25oC 3.95 4.37c)
tion. For nondenaturing conditions, water was added 0.5 ´ TBE at 50oC 5.25 5.89
instead of formamide, without any heating. For unknown 0.5 ´ TBE + 4 M urea
reasons, under denaturing conditions, double peaks were at 25oC 3.58a) 4.01
0.5 ´ TBE + 4 M urea
observed for all fragments of the 100 bp and 500 bp ruler.
at 50oC 4.64b) 5.24
Comparative runs revealed that in all cases, the first peak
of the first five doublets in the 100 bp ruler comigrates a) For comparison, see [36]
with the respective fragment of the 50 bp ruler. Therefore, b) In agreement with data from Fig. 3
the faster one of each double peak was taken for subse- c) In agreement with data from [8, 35]
quent analysis. Other conditions as in Fig. 1
1966 C. Heller Electrophoresis 1999, 20, 1962±1977

Figure 1. Dependence of the electrophoretic mobility of linear DNA fragments on size, separated in
pDMA of different molecular mass and at different concentrations. (A) dsDNA in short-chain pDMA;
(B) dsDNA in medium-chain pDMA; (C) dsDNA in long-chain pDMA; (D) ssDNA in short-chain pDMA;
(E) ssDNA in medium-chain pDMA and (F) ssDNA in long-chain pDMA. Conditions: 0.5 ´ TBE at
210 V/cm in a 47 cm long (36 cm to the detector) fused-silica (100 mm ID) capillary. (A)±(C) Separa-
tion at 25oC, (D)±(F) separation in 4 M urea and at 50oC. Zones I±IV (separated by dotted lines) corre-
spond to the different regimes explained in Section 3.2. The straight lines in (E) indicate the finding of
the critical size, N*; the straight line in (F) has a slope of ±1.
Electrophoresis 1999, 20, 1962±1977 Separation of double-stranded and single-stranded DNA 1967

threshold was also determined by measuring the viscosity regime IV) and smaller ones (70±250 bp, regime II) show
of the polymer solution at different concentrations and less difference in mobility. Regime II is most expressed at
finding the point of departure from linearity on the double higher polymer concentration (5 and 10%), whereas at
logarithmic specific viscosity vs. concentration plot (not low concentration, there is a smooth transition from
shown) [31]. Also, by using the independently measured regime I to III. This is in full agreement with earlier data
intrinsic viscosity values, we have previously determined [8]; however, the border between regimes III and IV has
the corresponding c* values of the different pDMA prepa-
rations (see Table 1 in [8]). Taking together the three
independent data sets, we get c* = 0.013  0.003 g/mL,
c* = 0.007  0.002 g/mL and c* = 0.004  0.001 g/mL for
ªshort chainº, ªmedium chainº and ªlong chainº pDMA, re-
spectively.

Using Eq. (2) and taking into account the new Mark-Hou-
wink parameters, we obtain ªmesh sizesº of about 3, 5, 11
and 19 nm, for 10%, 5%, 2% and 1% pDMA solutions, re-
spectively, independent of the chain length. Note that the
Mark-Houwink parameters were determined for a temper-
ature of 25oC. Under denaturing conditions (i.e., at 50oC
and in presence of urea), we might therefore obtain
slightly different values for the overlap threshold and the
mesh size. However, in any case ± as pointed out in Sec-
tion 1 ± the above calculations are based on scaling laws
only and precise prefactors can not be given. In our setup
it is difficult to measure the electroosmotic flow (EOF), but
under similar conditions, the residual EOF in a 0.01%
pDMA (98 kDa) solution was reported to be 0.2 ´ 10±4
cm2/Vs [5]. For our experiments, using longer pDMA
chains and a hundred times higher polymer concentra-
tion, we can therefore safely assume that the residual
EOF in our separation matrix is negligible.

3.2 Mobility and separation regimes


We have undertaken a systematic study of the mobility of
double-stranded and single-stranded DNA in pDMA prep-
arations of different molecular masses and at different
polymer concentrations. Figure 1 shows the dependence
of the electrophoretic mobility on DNA size (dsDNA and
ssDNA) at an electric field strength of 210 V/cm and at
polymer concentrations of 1%, 2%, 5% and 10% in ªshort
chainº, ªmedium chainº and ªlong chainº pDMA (with ªlong
chainº pDMA, we did not perform experiments at 10%,
due to the rather high viscosity, > 10 000 mm2/s). In all
cases, we can observe the usual sigmoidal curve on the
double logarithmic mobility vs. size plot, in agreement
with separation of dsDNA in agarose gels [26, 32], ssDNA
in polyacrylamide gels [33] or dsDNA in entangled poly-
mer solutions [17, 28]. Figure 2. Plot of mobility vs. the inverse of DNA size for
separation in medium chain pDMA (data from Figs. 1B
For dsDNA (Fig. 1A±C), we can define the following and E). (A) dsDNA; separation at 25oC; (B) ssDNA; sepa-
regimes (according to the notation in our previous work ration in 4 M urea and at 50oC. Other conditions as in Fig.
[8]): medium-sized molecules (250±900 bp) are well re- 1. The zones I±IV (separated by dotted lines) correspond
solved (regime III), whereas larger ones (> 900 bp; to the different regimes explained in Section 3.2.
1968 C. Heller Electrophoresis 1999, 20, 1962±1977

been slightly shifted to larger DNA sizes. A possible


explanation for this might be that in earlier studies, inter-
calators were used as fluorescent markers instead of a
covalently bound dye. Such intercalators can change the
properties of DNA (charge, flexibility and length), there-
fore leading to slightly different results. Also, note that the
limits are difficult to define as smooth transitions occur be-
tween the different regimes. In this new series of experi-
ments we did not use small DNA; therefore in this case,
regime I (< 70 bp) can not be well observed. For better
illustration, we refer our readers to Fig. 7 of [8].

In our previous work [8], we could identify the different


regimes as follows: Regime I corresponds to an Ogston-
type sieving mechanism, whereas regime III and IV indi-
cate unoriented reptation and reptation with orientation,
respectively. Regime II (plateau regime on the left hand
side) is a newly discovered regime and presumably corre-
sponds to the situation when the radius of gyration of
DNA, Rg, is larger than the pore size (so that sieving is Figure 3. Semilogarithmic plot of the electrophoretic mo-
impossible), but at the same time the DNA is too stiff to bility vs. pDMA concentration (ªFerguson plotº) for three
effectively reptate through the separation matrix [8]. This different ssDNA fragments, each one selected from one
is the case when the DNA is shorter than one Kuhn of the three regimes of Fig. 1E (medium-chain pDMA,
length, and can occur in the ªtight gelº situation (see Sec- 603 kDa). Conditions as in Fig. 1. The straight line repre-
tion 4). Presumably, the DNA molecules assume a ªrod- sents a fit of the data for the 75 b fragment to the following
likeº conformation and size-based separation is sup- relation: Y = 4.60 exp[±0.122c]; r2 = 0.995.
pressed in this regime.
dsDNA in different entangled polymer solutions [8, 17,
28], the mobility does not become totally independent of
In the case of ssDNA (denaturing conditions, Fig. 1D±F),
size.
the situation is more complex. We can observe again that
plateau mobility (regime IV) starts to occur above a cer-
tain size, but this critical size increases with decreasing In order to verify if zone I corresponds to an Ogston-like
polymer concentration. At the left-hand side of the plot, separation, we picked one representative fragment from
we cannot observe a plateau regime, but instead a direct each zone of Fig. 1E and plotted the data in a semiloga-
transition from zone I to III. Again, the point of transition rithmic mobility vs. concentration plot (so-called Fergu-
increases with increasing polymer concentration. As in son-plot [34], Fig. 3). If we assume that the radius of gyra-
the case of dsDNA, the slope in regime III decreases with tion of the DNA is larger than the fiber radius, Eq. (3)
increasing polymer concentration but a slope of ±1, as predicts that the Ogston-type separation should lead to a
would be demanded by the reptation model, is not straight line with negative slope. Indeed, we find that the
reached. However, considering the rather strong electric 75 b fragment fulfills this condition, whereas the other
field (210 V/cm) used here, we cannot expect to observe fragments do not. The straight line intercepts the y-axis at
ªtrueº, i.e. unbiased reptation, but ± in analogy to the a value of 4.6 ´ 10±4 cm2/Vs, which agrees well with the
dsDNA separation ± we can again assume that in this experimental data of free solution mobility (see below and
regime the separations is still based on a reptation mech- Table 1). The same behaviour has been verified for the
anism. To verify this assumption, we have replotted ± as separations in ªshort chainº and ªlong chainº pDMA (not
an example ± a subset of the data of Figs. 1B and 1E as shown).
linear mobility vs. inverse of size plot (Fig. 2A and B). If
reptation is the dominant mechanism, the mobility is pro-
3.3 Mobility in free solution
portional to 1/size and the data points will fall on a straight
line. Indeed, in Fig. 2 we observe such a linear relation- In Fig. 1, we also observe that the mobility of the small
ship within the domain of regime III, both for ssDNA as ssDNA molecules (i.e.; at the left-hand side of the plots,
well as for dsDNA. In zone IV (Fig. 1), the mobility levels Fig. 1D±F) is generally higher than the mobility of the cor-
off, which indicates that it corresponds to the reptation responding dsDNA standards (Fig. 1A±C). This is in
with orientation regime. As observed for the separation of agreement with the findings by van der Schans et al. [9]
Electrophoresis 1999, 20, 1962±1977 Separation of double-stranded and single-stranded DNA 1969

rithmic mobility vs. size plot, indicating that a higher reso-


lution might be achieved when using denaturing condi-

Figure 4. Mobility difference per base or per base pair, in


((mm/s)/(V/cm))/b or ((mm/s)/(V/cm))/bp, respectively, for
ssDNA and dsDNA in 5% pDMA (1146 kDa). The curves
are a guide to the eye only. Data are taken from Figs. 1C
and 1F.

and can be attributed to a lower viscosity due to the high-


er temperature (50oC vs. 25oC). Table 1 gives the free
solution mobilities (measured in 0% pDMA) for both
cases, i.e., ssDNA separated in denaturing conditions
(50oC and 4 M urea) and dsDNA in nondenaturing condi-
tions (25oC without urea). In order to obtain the influence
of urea and temperature alone, we also measured the
free solution mobilities under the ªwrongº conditions, i.e.,
ssDNA in nondenaturing conditions and dsDNA in dena-
turing conditions. The result for dsDNA agrees well with
previously evaluated data [8, 35], but ssDNA is about two
times faster than reported in the literature [36]. Note, how-
ever, that in our case, the buffer ionic strength was about
five times higher. Generally, we find that under the same
conditions, the mobility of ssDNA is about 0.9 times the
mobility of dsDNA. Additionally, the presence of urea
slows down the mobility by about the same factor. How-
ever, this is more than compensated for by a 1.3-fold
increase in mobility due to the higher temperature under
denaturing conditions. Figure 5. (A) Dependence of the electrophoretic mobility
of linear DNA fragments on the pore size, separated in
medium chain pDMA (603 kDa). Conditions as in Fig. 1.
3.4 Mobility difference The lines represent a fit of the data to the following rela-
tion: 500 b ssDNA: Y = ±1.23 + 3.3 log(pore size), (r2 =
However, more striking than the mobility difference for
0.997); 5000 b ssDNA: Y = ±0.64 + 2.0 log(pore size),
small molecules is the fact that the large single-stranded (r2 = 0.999); 500 bp dsDNA: Y = ±0.27 + 2.2 log(pore
molecules (i.e., those nearly assuming a plateau mobility, size), (r2 = 0.998); 5000 bp dsDNA: Y = ±0.354 + 1.98
at the right-hand side of the plots, regime IV) migrate con- log(pore size), (r2 = 0.997). (B) Same data as in (A), but
siderably slower than their nondenatured counterparts. presented in a double logarithmic plot. The straight line
This leads to steeper slopes in zone III of the double loga- has a slope of 1.5, the dashed line has a slope of 1.
1970 C. Heller Electrophoresis 1999, 20, 1962±1977

tions instead of nondenaturing conditions, due to larger polymer concentrations and different chain lengths (not
spacing between the peaks. This effect is better displayed shown). This striking difference and the differences
in Fig. 4: We have taken the difference in mobility (nor- explained above led us to the assumption that a different
malized to the size) between neighboring fragments, plot- separation regime might be valid for ssDNA compared to
ted vs. the size of the smaller fragment of each ªpairº for dsDNA. As ssDNA has a much smaller Kuhn length than
the data obtained in 5% pDMA (long chain), as an exam- duplex DNA, it might be possible that here we might not
ple. We clearly see that the mobility differences (and have the situation of a ªtight gelº. In the following, we fur-
therefore the peak distances on the electropherograms) ther investigate this possibility.
obtained under denaturing conditions are considerably
higher than under nondenaturing conditions (for frag-
ments < 400 b). Similar results are obtained for different 3.5 Influence of pore size

To gain more insight into the separation mechanism, we


have investigated the influence of the pore size on the
mobility. Again, we only present the data from the experi-
ments in medium sized pDMA, as we could not find signif-
icant differences in the data sets obtained in ªsmall chainº
or ªlong chainº pDMA. We picked the 500 b (and 500 bp,
respectively) fragment as a representative for DNA sep-
arated by nonoriented reptation, and the 5000 b (5000
bp) fragment as an example for an ªorientedª molecule.
Figure 5 shows a semilogarithmic plot of the dependence
of mobility on the pore size for dsDNA and ssDNA. The
data can be fitted to logarithmic functions (see legend).
Interestingly, the slopes for both dsDNA fragments are
similar, whereas they change considerably from ssDNA in
the unoriented regime to ssDNA in the oriented regime.
We have replotted the same data in a double logarithmic
mobility vs. pore size plot in order to verify if the data
could be described by a power law. Clearly, the data do
not fit a straight line; only for small pore sizes (concentra-
tions of 5 and 10%) will they approximately fall on a line
with a slope of 1.5 for dsDNA and 1 for ssDNA.

3.6 Influence of chain length

As pointed out in Section 1, the length of the polymer


chains might influence the network dynamics and there-
fore the separation. To illustrate this effect, we have
regrouped part of the data from Fig. 1 into new plots pre-
senting the DNA mobility in matrices of the same concen-
tration but of different chain length (Fig. 6). Under native
conditions (Fig. 6A), the respective curves obtained are
identical (within experimental error), both at low (1%) and
at high (5%) concentration. Following the argument by
Cottet et al. (see Fig. 2 of [16]), this indicates that even for
short chain pDMA, at 1% the ªthreshold valueº is already
reached, above which the reptation of the polymer ceases
to affect the separation. The situation is slightly different
Figure 6. Dependence of the electrophoretic mobility of under denaturing conditions (Fig. 6B): in 1% and in 5%
linear DNA fragments on the polymer chain length at a pDMA, the mobility of the larger fragments (zone III and
concentration of 1% and 5% pDMA. (A) dsDNA; separa- IV) slightly decreases with increasing chain length, indi-
tion at 25oC; (B) ssDNA; separation in 4 M urea and 50oC. cating that network dynamics may be influencing the sep-
Other conditions as in Fig. 1. aration.
Electrophoresis 1999, 20, 1962±1977 Separation of double-stranded and single-stranded DNA 1971

Figure 7. (A) Dependence of the electrophoretic mobility on the size of linear ssDNA fragments sep-
arated in 2% pDMA (1146 kDa) at different electric fields. Other conditions as in Fig. 1. The straight
line is for illustration only and has a slope of ±1. (B) Dependence of the electrophoretic mobility on the
size of linear ssDNA fragments separated in 5% pDMA (1146 kDa) at different electric fields. Other
conditions as in Fig. 1. The straight line illustrates a slope of ±1. (C) Dependence of the electropho-
retic mobility of linear ssDNA fragments on the electric field in 2% pDMA (1146 kDa). Data are from
Fig. 5A. The straight line has a slope of 0.4. (D) Dependence of the electrophoretic mobility of linear
ssDNA fragments on the electric field in 5% pDMA (1146 kDa). Data are from Fig. 5B. The straight
lines (for illustration only) have a slope of 0.4 and 1, respectively.

3.7 Influence of electric field on the mobility of dsDNA has been studied before [8]; we
will therefore restrict ourselves to the mobility of ssDNA.
Besides the polymer concentration and the pore size, the Figures 7A and B show again double logarithmic plots of
electric field strength is one of the major factors influenc- mobility vs. size under denaturing conditions, obtained in
ing the mobility of DNA in a porous matrix. As most elec- 2% long chain pDMA and 5% long chain pDMA, respec-
trophoresis models are strictly valid for zero or low electric tively, but at different electric fields. Again, the result is
field only, it is important to measure the field dependence similar to that obtained for dsDNA in pDMA [8] and in
of a separation method. The influence of the electric field other matrices [28]. However, in the example shown here
1972 C. Heller Electrophoresis 1999, 20, 1962±1977

3.8 Critical size, N*


We have also evaluated the crossover point N*, which
marks the transition from reptation in Gaussian conforma-
tion to oriented conformations. This parameter is impor-
tant because it indicates the point where separation starts
to fail. In DNA sequencing, for example, this can limit ±
among other factors ± the number of bases that can be
read. N* can be estimated graphically by finding the inter-
section between the average straight line with a steep
slope formed by the data in regime III of the double loga-
rithmic mobility vs. size plot and the nearly horizontal line
connecting the data points from the plateau mobility (for
example, see data for 5% pDMA in Fig. 1E). Figure 8A
shows a double logarithmic plot of N* vs. pore size for
ssDNA in all three polymer preparations. With the excep-
tion of 1% short-chain pDMA (which is probably only a
weakly entangled polymer solution), the data can be fitted
to straight lines with a slope of 0.4. As already mentioned
above, for dsDNA the ªcritical sizeº N* is constant over
about one order of magnitude of pore sizes. The cross-
over point has also been determined in dependence on
the electric field. In Fig. 8B we present the data obtained
from separations in 5% long-chain pDMA (Fig. 7B): the
data satisfy a power law of ±0.5.

3.9 Comparison to theoretical predictions


For small molecules (regime I), the separation of ssDNA
and dsDNA can be explained by a sieving mechanism
Figure 8. Dependence of the critical size, N* on the pore
(see Fig. 3, and Fig. 8 of [8]). However, for larger DNA
size. Conditions as in Fig. 1. The straight line is horizon-
(regime III and IV) the situation is more difficult. As the
tal, the dashed line (for illustration only) has a slope of
0.4. (B) Dependence of the critical size, N*, of ssDNA on theoretical treatment of the BRF is rather complex, we
the electric field strength in 5% pDMA (1146 kDa). Condi- have compiled the major scaling predictions in Table 2
tions as in Fig. 7. The straight line (for illustration only) (again, we follow [25]). dsDNA has a persistence length
has a slope of ±0.5. of about 45 nm [37, 38], which means a Kuhn length of
about 90 nm, corresponding to about 300 bp. Taking into
account the pore sizes calculated above for the pDMA
(Fig. 7A), a slope of ±1 could not be reached in 2% solutions used here, we were confident that, in this case,
pDMA. However, at higher concentration (5% pDMA; Fig. we were working under ªtight gelº conditions. The fact that
7B) and low electric field, we nearly approach ideal repta- we can observe a plateau zone at the left-hand side of the
tion conditions in the size range between 200 and 700 b. mobility vs. size plot (regime II) supports this assumption
A subset of the data from Figs. 7A and B was replotted in (see also Fig. 9 and Section 4). Therefore, we have
order to demonstrate the dependence of the mobility on compared our dsDNA data with the predictions for persis-
the electric field (Figs. 7C and D). Again, the results are in tent chains (lower part of Table 2). On the other hand, the
agreement to those obtained under nondenaturing condi- persistence length of ssDNA was recently determined to
tions [8]: At low electric field and small DNA size, the mo- be about 4 nm (in low ionic strength, [39]), corresponding
bility remains constant (but dependent on the size), to a Kuhn length of 8 nm or about 20 nucleotides. This is
whereas larger molecules show an electric field depend- in the same order of magnitude as the ªpore sizeº of a 5±
ence. It is difficult to find scaling laws over this range of 10% pDMA solution, but given the fact that the prefactors
electric field strengths, but for large molecules in 2% long- in Eq. (2) cannot be given exactly and that the flexibility of
chain pDMA the mobilities fall on a line with a slope of DNA is greater in higher ionic strength buffers (persis-
0.4. In 5% pDMA and at strong electric field, there is a tence length of ssDNA in 0.5 ´ TBE is about 2 nm [36]),
deviation to a steeper slope, close to unity. we think that we do not have a ªtight gelº situation here.
Electrophoresis 1999, 20, 1962±1977 Separation of double-stranded and single-stranded DNA 1973

Table 2. Scaling predictions of the BRF for different parameters, according to [25], in comparison to experimental results
for ssDNA and desDNA in entangled pDMA solutions
DNA size Electric field Pore size
Predicted Observed Predicted Observed Predicted Observed
Flexible chain ssDNA Flexible chain ssDNA Flexible chain ssDNA

m/m0 unoriented Nk±1 N±1 a) Independent Independent xk2 x1.5 0.2 d)

m/m0 oriented Independent Nearly indep. ek E0.4±1 b) xk2 x1.5 0.2 d)
 0.4  0.1
N* ± ± ek±1 E±0.5 0.1 Independent x

Persist. chain dsDNA Persist. chain dsDNA Persist. chain dsDNA



m/m0 unoriented Nk±1 N±1 a) Independent Independent Independent x1.5 0.2 d)

m/m0 oriented Independent Nearly indep. ek; ek2/5; ek2 E0.15±0.4 b, c) xk3/2; xk12/5; xk4 x1 0.2 d)
N* ± ± ek±1; ek±2/5; ek±2 n.d. xk±3/2; xk±12/5; xk±4 Independent
a) For small electric fields and small pore size
b) For large DNA
c) See Fig. 9C of [8]
d) Only for high polymer concentration, but better fitted by a logarithmic function (see Fig. 4)

aration, i.e., a separation inversely proportional to the size


and a plateau mobility above a critical size (sigmoidal
shape of the double logarithmic mobility vs. size plot).
However, other predictions, like the dependence of mobil-
ity and critical size on pore size, do not agree with the
results. Concerning the electric field, the independence of
the mobility in the unoriented reptation regime is con-
firmed; however, in the oriented regime, the electric field
dependence is more complex. For the critical size under
denaturing conditions, we find a dependence on the
square root of the inverse electric field whereas the BRF
predicts N* proportional to E±1. Our results are in contrast
to findings from Dovichi©s group, who indeed reported a
dependence of N* nearly inversely proportional to the
electric field in 5% non-cross-linked polyacrylamide (see
Figs. 10 and 11 of [40]).

Figure 9. Diagram of the different regimes existing in the


electrophoretic separation of DNA in polymer solutions.
4 Discussion
The dotted line indicates the border between (A) the ªper-
sistent chainº situation and (B) the ªflexible chainº situa- In ªclassicalº gel electrophoresis, dsDNA and ssDNA are
tion. The dashed line indicates the predicted critical size
usually separated in different matrices. The pore size of
(for better clarity, we have only indicated one of the three
the gel is adapted to the flexibility of the DNA, i.e., we use
different possibilities existing in ªtight gelsº). The arrows
indicate the situation for ssDNA and dsDNA in polymer agarose with large pores to separate the rather stiff
solutions of different concentration. For explanation of the dsDNA and cross-linked polyacrylamide (small pores) for
different regimes, see Section 3.2. the more flexible ssDNA. In both cases, we can assume
that the average pore size is larger than the Kuhn length
Also, we cannot observe a zone II as exists in dsDNA of the DNA and we have a ªflexible chainº situation. In
separations and therefore we have compared our ssDNA capillary electrophoresis, often the same matrix is used
data to the predictions for flexible chains (upper part of for both types of DNA. In most cases, we shall then have
Table 2). two different situations: Native DNA with a large Kuhn
length has to thread through rather small pores (ªtight
As can be seen from Table 2, the BRF can only correctly gelº), whereas for denatured and ssDNA, the ªflexible
predict the major observation in electrophoretic DNA sep- chainº situation remains.
1974 C. Heller Electrophoresis 1999, 20, 1962±1977

4.1 ssDNA sis, and we reach regime IV (Rg 4 bD > x), again with a
possible, but less than good separation.
This leads to different regimes for the mobility in depend-
ence on size, which is explained in Fig. 9: For ssDNA, the
Kuhn length is smaller than the pore size (except maybe 4.3 Separation mechanism
for very high polymer concentrations), i.e., bD < x, or xk >
As outlined above, we have attributed the separation in
1. For short molecules (i.e., smaller than the pore size, Rg
zone I to a sieving mechanism, due to the fact that the
< x), the DNA undergoes separation by a sieving mech- analytes are smaller than the estimated pore size of the
anism (regime I). With increasing size, the point will come
matrix. The ªclassicalº sieving theory applied in electro-
where the radius of gyration is bigger than the pore size
phoresis is the so-called Ogston model and, in fact, the
and the DNA will start to reptate (regime III, Rg > x > bD).
electrophoretic behavior of short DNA in our experiments
Naturally, this first crossover point shifts to larger DNA
can be explained using this model (see Fig. 8 of [8] and
with increasing pore size, which indeed can be observed
Fig. 3 in this study). However, this is somewhat surpris-
in Figs. 1D±F. The radius of gyration of a polymer can be
ing, because under our conditions the main assumptions
described by the so-called Kratky-Porod formula and has
of the Ogston model are not valid: the Ogston model was
been estimated by Viovy et al. [10] for both ssDNA and
originally developed for the movement of spherical
dsDNA (see Fig. 4 of [10]). For ssDNA, a size of about 40
objects in a random, fixed array of fibers, where the pore
b, 100 b and 400 b would give a radius of gyration of 3, 5
size follows Poisson statistics. In entangled polymer solu-
and 11 nm, respectively corresponding to the estimated
tions, the ªporesº are probably more uniform and the
pore sizes of the 10%, 5% and 2% pDMA solutions (see
obstacles themselves might move (ªconstraint releaseº).
above). The crossover points from regime I to regime III
Second, a short DNA molecule is rod-like and it is ques-
for ssDNA (see dotted lines in Fig. 1D±F), do not fully
tionable to what extent it can be regarded as a so-called
match these values, but come rather close. By further
ªequivalent sphereº. Therefore, in the context of electro-
increasing the size, the DNA will become oriented and we
phoretic separation of short DNA molecules in entangled
reach regime IV (Rg 4 x > bD). According to the reptation
polymer solutions we can only talk about ªOgston-likeº
model, here the mobility should become independent of
sieving mechanisms and more theoretical work might be
size; however, in entangled polymer solutions we often
necessary to clarify this point.
observe a slight size dependence in this regime. This
means that ± in contrast to gels ± the separation does not
The separation of larger DNA molecules, i.e., larger than
totally fail for large DNA.
the pore size, is classically explained reptation models. In
Table 2, we have compiled the predictions of the latest
4.2 dsDNA reptation model, the biased BRF, taking into account elec-
trohydrodynamic forces. The BRF correctly predicts the
For small dsDNA, we have Rg < x < bD and the DNA is major observation in electrophoretic DNA separation, i.e.
sieved (regime I). As in ssDNA, with increasing size the the inversely proportional dependence of the mobility on
DNA will become larger than the pore size; however, as size (see Fig. 2) and the plateau mobility for large DNA
the molecules are rather stiff, their radius of gyration can and high electric field. However, the model dramatically
be smaller than their Kuhn length. This leads to the spe- fails when we look at the predicted dependencies on the
cial situation of bD > Rg > x (regime II), where the mole- pore size. Neither for ssDNA nor for dsDNA could the pre-
cules are too large to be separated by sieving, but too stiff dicted dependence of the mobility on the pore size be
to effectively reptate. Presumably, they adopt a rod-like verified. Even more important, the dependence of the crit-
conformation and travel through the matrix at the same ical size, N*, is predicted to be independent of pore size
speed. The crossover from regime I to regime II should for flexible chains, but increases with pore size in our
also be dependent on the polymer concentration, which experiments. In tight gels, it should decrease with pore
does not fully agree with our observations. However, the size (in all three predicted cases; see Eq. 6A±C), but it
crossover point is difficult to determine and the fact that was found to be constant within experimental error. In the
for large pore sizes, zone II becomes small (Fig. 9), original version of the BRF [23] and in the BRM [27], N*
explains that for low polymer concentration, we have a was even predicted to scale with the pore size with a neg-
smoother transition from zone I to zone II (Fig. 1A±C). ative exponent, i.e. ±1 and ±4, respectively.
The crossover from regime II to III (Rg > bD > x), however,
is given by the Kuhn length and the observed transition This wrong prediction of the pore size dependence seems
point at about 250 bp agrees well with the estimated Kuhn to be a major weakness of reptation models and only the
length of 300 bp. As in ssDNA, by further increasing the latest improvements have brought it closer to reality. For
size, the molecules become oriented during electrophore- instance, in gel electrophoresis, N* has been found to be
Electrophoresis 1999, 20, 1962±1977 Separation of double-stranded and single-stranded DNA 1975

nearly independent of pore size for dsDNA in agarose cal size, more experiments at different field strengths are
gels (see Fig. 3 of [26]) and also for ssDNA in polyacryl- necessary. It is possible that extrapolations to low or zero
amide gels (see Fig. 2 in [33]), which can now be electric field could then reveal and asymptotic trend to dif-
explained by the most recent version of the BRF [25]. ferent scaling laws.
However, for DNA separation in polymer solutions, it still
cannot account for the pore size dependence. The wrong For flexible chains, there is also a discrepancy between
theoretical scaling of N* with pore size has led to the the prediction and the observation concerning the electric
wrong assumption that the read length in DNA sequenc- field dependence of the mobility in the oriented regime
ing could be increased with increasing gel concentration, and of the critical size, N*. In both cases, the scaling is
in contrast to the practical findings (e.g., [41]). Based on less strong than predicted. Only in highly entangled poly-
the findings of the BRF model, for example, Slater et al. mer solutions and at rather strong electric fields do we
[42] have predicted that the read length in gel electropho- approach a linear scaling for the mobility of oriented
retic DNA sequencing could be increased by increasing ssDNA on the electric field (Fig. 7D). Similarly, for dsDNA,
simultaneously the electric field and the gel concentration. a slope of 0.4 could only be found for strong electric fields
If we apply their rule (Eq. 25 of [42]) to our experimental (see Fig. 9C of [8]). Again, what has been said above
data in polymer solutions, we cannot confirm this predic- about the possible sources of the discrepancies remains
tion. In 2% pDMA at 105 V/cm, we find a higher critical valid. Moreover, it is difficult to determine precise scalings
size (about 1100 bases) than in 160 V/cm (1.6 ´ the elec- over a range of field intensities of only one order of magni-
tric field) and 5% pDMA (1.62 = 2.5 ´ the polymer concen- tude.
tration).
4.4 Separation matrix
The reason why the severe failure of reptation models
concerning the pore size dependence has been over- When working with polymer solutions, we also have to
looked might be due to the fact that, in gels, the effective take into account the flexible nature of the network itself.
pore size is difficult to estimate. In fact, the relation be- The polymer chains are not linked to each other but can
tween the pore size (and the pore size distribution) and move themselves. This leads to a renewal of the obsta-
the concentration of agarose gels and polyacrylamide cles and can be described by the concept of ªconstraint
gels, was the subject of intense discussion (see e.g., releaseº [11]. The constraint release mobility adds to the
Table 1 in [43]). In entangled polymer solutions, we can electrophoretic mobility and increases with shorter chain
assume that the pore size is more uniform and that it can length and lower polymer concentration (see Eq. 7). How-
correctly be described by the blob size. Still, in entangled ever, in our experiments with dsDNA in entangled pDMA
polymer solutions, the assignement of the pore size to a solutions we could not observe an effect on the separa-
solution with known concentration depends heavily on the tion with increasing chain length (Fig. 6A), indicating that
correct estimation of Mark-Houwink parameter a. Accord- the threshold value, above which the reptation of the poly-
ing to Eq. (2), the pore size scales with c±(a+1)/3a. In our mer ceases to affect the separation, is already reached.
case, this exponent is ±0.79. However, even if this expo- Therefore, we assume that constraint release does not
nent were not correct, it should only vary between ±0.75 play a major role under these conditions.
(for a = 0.8; swollen chain) and ±1 (a = 0.5; ideal solvent).
This could not explain the striking differences between However, when using the same matrix for the separation
the predicted and the observed pore size dependence of of ssDNA, a slight effect on the electrophoretic mobility
the mobility and the critical size. can be observed with increasing chain length. This could
be explained as follows: The more flexible ssDNA has a
The observed difference could be due to other factors. longer renewal time (biased reptation time) than dsDNA.
For instance, it is not certain if the blob size corresponds The reptation time of the matrix polymers, however,
to the effective pore size of the network. In other words, remains the same. According to Cottet et al. [16], the net-
the DNA could ªseeº pores which are different from the work dynamics has an influence on the separation when
pores derived by the theory of polymer solutions. Also, as the reptation time of the polymer is smaller than the
the applied electric fields are rather high, the possibility of renewal time of the analyte, which could be the case here.
ªhernia formationº [44] is increased. ªHerniasº or ªloopsº Another explanation could be given using the concept of
are a sidewise escape of the DNA from the ªtubeº, and constraint release: The Kuhn length bD for ssDNA is much
reptation models become invalid. However, they should smaller than for dsDNA, so that the first term in Eq. (7)
only occur in flexible chains. Most models were originally becomes larger. This in turn means a larger contribution
developed for low electric field strengths. To gain more of constraint release to the total mobility than compared
insight into the pore size dependence of mobility and criti- to the case under nondenaturing conditions.
1976 C. Heller Electrophoresis 1999, 20, 1962±1977

Note, however, that the polymer chains can also be I would like to thank R. Reinhard for providing the capil-
actively dragged by the moving DNA, a process which ± lary electrophoresis instrument and R. Steinkamp, V.
in contrast to constraint release ± should be dependent Egelhofer and J. Schacherl for their help with the charac-
on the DNA size and the electric field. The more flexible terization of the polymer solutions and with the DNA sepa-
ssDNA molecules probably get hooked around matrix rations. J.-L. Viovy is thanked for providing his manuscript
fibers more frequently than do the rather stiff duplex DNA prior to publication and for helpful discussions and one of
chains. Thus, a network rupture would be more likely. the referees is thanked for very helpful comments. I am
Such effects could also explain the fact that even large very grateful to M. Millequant, Labo PCM, ESPCI, Paris
molecules are still separated and their mobility does not (France) for performing the size exclusion chromatogra-
fully level off. So far, no model exists that takes into phy analysis. This work was supported by a grant from
account such ªdraggingº effects in entangled polymer sol- the German Ministry of Research (BMBF), FKZ 0311040
utions. We feel that the electrophoretic separation of DNA and by a grant from the Commission of the European
in entangled polymer solutions is not fully understood and Union, CT 97-2627.
more experimental and theoretical work would be needed
to reveal these phenomena. Received February 1, 1999

Under denaturing conditions, the mobility of large DNA is


lower than under native conditions. Consequently, DNA 5 References
up to the critical size is better separated and we can say [1] Heller, C., in: Heller, C. (Ed.), Analysis of Nucleic Acids by
that the system has a better selectivity under denaturing Capillary Electrophoresis, Vieweg Verlagsgesellschaft,
conditions than under nondenaturing conditions (Fig. 4). Wiesbaden 1997, pp. 3±23.
Therefore, if high resolution is necessary (i.e., one or two [2] Barron, A. E., Blanch, H. W., Soane, D. S., Electrophoresis
base resolution) the DNA should be separated under 1994, 15, 597±615.
denaturing conditions. This is especially interesting in [3] Chiari, M., Riva, S., Gelain, A., Vitale, A., Turati, E., J. Chro-
applications such as genotyping and fingerprinting and, in matogr. A 1997, 781, 347±355.
fact, such conditions are successfully applied to different [4] Rosenblum, B. B., Oaks, F., Menchen, S., Johnson, B.,
Nucleic Acids Res. 1997, 25, 3925±3929.
biological applications (e.g., [6]). We also found that for
[5] Madabhushi, R. S., Electrophoresis 1998, 19, 224±230.
ssDNA in pDMA solutions, the critical size, i.e., the point
[6] Wenz, H. M., Robertson, J. M., Menchen, S. M., Oaks, F.,
beyond which separation fails, is increased with decreas-
Demorest, D. M., Scheibler, D., Rosenblum, B. B., Wike, C.,
ing polymer concentration. This effect could probably be Gilbert, D. A., Efcavitch, J. W., Genome Res. 1998, 8,
used for longer read lengths in DNA sequencing. How- 69±80.
ever, both a higher resolution and longer read length can [7] Waters, L. C., Jacobson, S. C., Kroutchinina, N., Khandur-
only be achieved if the peak width does not increase ina, J., Foote, R. S., Ramsey, J. M., Anal. Chem. 1998, 70,
simultaneously with the selectivity or the critical size. The 5172±5176.
dependence of the peak width on different experimental [8] Heller, C., Electrophoresis 1998, 19, 3114±3127.
conditions is the subject of a further study [45]. [9] van der Schans, M. J., Kuypers, A. W. H. M., Kloosterman,
A. D., Janssen, H. J. T., Everaerts, F. M., J. Chromatogr. A
1997, 772, 255±264.
[10] Viovy, J.-L., Heller, C., in: Righetti, P. G. (Ed.), Capillary
4.5 Conclusion Electrophoresis: An Analytical Tool in Biotechnology Series,
CRC Press, Boca Raton 1996, pp. 477±508.
In this study, we have used pDMA as the separation [11] Viovy, J.-L., Duke, T., Electrophoresis 1993, 14, 322±329.
matrix. The results obtained for dsDNA are very similar to [12] Broseta, D., Leibler, L., Lapp, A., Strazielle, C., Europhys.
those obtained in other matrices such as hydroxypropyl- Lett. 1986, 2, 733±737.
cellulose [28] or dextran [17]. Therefore, we can assume [13] Cowie, J. M. G., Polymers: Chemistry and Physics of Mod-
that the findings for ssDNA might also be valid in other ern Materials, Blackie Academic, London 1991.
entangled polymer matrices. We preferred pDMA [14] Kurata, M., Tsunashima, Y., in: Brandrup, J., Immergut, E.
H. (Eds.), Polymer Handbook, John Wiley and Sons, New
because of its unique properties such as its self-coating
York 1989, pp. VII/1±VII/46.
ability and its low viscosity [5, 8]. Very recently, pDMA
[15] Thielking, H., Kulicke, M. W., Anal. Chem. 1996, 68,
has also become commercially available, but only as a 1169±1173.
100 kDa preparation, which is considerably shorter than [16] Cottet, H., Gareil, P., Viovy, J.-L., Electrophoresis 1998, 19,
the preparations used here. The work presented here 2151±2162.
should help to optimize DNA separation in capillary elec- [17] Heller, C., Electrophoresis 1998, 19, 1691±1698.
trophoresis and to improve existing models for DNA sepa- [18] Slater, G. W., Mayer, P., Drouin, G., Methods Enzymol
ration in porous matrices. 1996, 270, 272±295.
Electrophoresis 1999, 20, 1962±1977 Separation of double-stranded and single-stranded DNA 1977

[19] Slater, G. W., in: Heller, C. (Ed.), Analysis of Nucleic Acids [32] Heller, C., Viovy, J.-L., Biopolymers 1995, 35, 485±492.
by Capillary Electrophoresis, Vieweg, Wiesbaden 1997, pp. [33] Heller, C., Beck, S., Nucleic Acids Res. 1992, 20,
24±66. 2447±2452.
[20] Ogston, A. G., Trans. Faraday Soc. 1958, 54, 1754±1757. [34] Ferguson, K. A., Metabolism 1964, 13, 985±1002.
[21] Morris, C. J. O. R., Morris, P., Biochem. J. 1971, 124, [35] Stellwagen, N. C., Gelfi, C., Righetti, P. G., Biopolymers
517±528. 1997, 42, 687±703.
[22] Slater, G. W., Guo, H. L., Electrophoresis 1995, 16, 11±15. [36] Pluen, A., Tinland, B., Sturm, J., Weill, G., Electrophoresis
[23] Duke, T., Viovy, J.-L., Semenov, A. N., Biopolymers 1994, 1998, 19, 1548±1559.
34, 239±247. [37] Hagerman, P. J., Ann. Rev. Biophys. Biophys. Chem. 1988,
[24] Slater, G. W., Noolandi, J., Biopolymers 1986, 25, 431±454. 17, 265±268.
[25] Viovy, J.-L., Rev. Mod. Phys. 1999, in press. [38] Hagerman, P. J., Ramadevi, V. A., J. Mol. Biol. 1990, 212,
[26] Heller, C., Duke, T., Viovy, J.-L., Biopolymers 1994, 34, 351±362.
249±259. [39] Tinland, B., Pluen, A., Sturm, J., Weill, G., Macromolecules
[27] Slater, G. W., Rousseau, J., Noolandi, J., Turmel, C., 1997, 30, 5763±5765.
Lalande, M., Biopolymers 1988, 27, 509±524. [40] Yan, J. Y., Best, N., Zhang, J. Z., Ren, H., Jiang, R., Hou,
[28] Mitnik, L., SalomØ, L., Viovy, J. L., Heller, C., J. Chromatogr. J., Dovichi, N. J., Electrophoresis 1996, 17, 1037±1045.
A 1995, 710, 309±321. [41] Ansorge, W., Barker, R., J. Biochem. Biophys. Methods
[29] Carlsson, C., Larsson, A., Jonsson, M., in: Heller, C. (Ed.), 1984, 9, 33±47.
Analysis of Nucleic Acids by Capillary Electrophoresis, [42] Slater, G. W., Mayer, P., Drouin, G., Electrophoresis 1993,
Vieweg, Wiesbaden 1997, pp. 67±89. 14, 961±966.
[30] Hubert, S. J., Slater, G. W., Viovy, J.-L., Macromolecules [43] Stellwagen, N. C., Electrophoresis 1997, 18, 34±44.
1995, 29, 1006±1009. [44] Deutsch, J. M., Science 1988, 240, 922±924.
[31] Grossman, P. D., Soane, D. S., J. Chromatogr. 1991, 559, [45] Heller, C., Electrophoresis 1999, 20, 1978±1986.
257±266.

You might also like