Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2011 Feb 2.
Published in final edited form as: J Phys Condens Matter. 2010 Feb 2;22(7):72202. doi: 10.1088/0953-8984/22/7/072202

Fluctuations and interactions of semi-flexible polyelectrolytes in columnar assemblies

D J Lee 1,*, S Leikin 2,, A Wynveen 3
PMCID: PMC2844736  NIHMSID: NIHMS183660  PMID: 20352061

Abstract

We have developed a statistical theory for columnar aggregates of semi-flexible polyelectrolytes. The applicability of previous, simplified theories was limited to polyelectrolytes with unrealistically high effective charge and, hence, with strongly suppressed thermal undulations. To avoid this problem, we utilized more consistent approximations for short-range image-charge forces and steric confinement, resulting in new predictions for polyelectrolytes with more practically important, lower effective linear charge densities. In the present paper, we focus on aggregates of wormlike chains with uniform surface charge density, although the same basic ideas may also be applied to structured polyelectrolytes. We find that undulations effectively extend the range of electrostatic interactions between polyelectrolytes upon decreasing aggregate density, in qualitative agreement with previous theories. However, in contrast to previous theories, we demonstrate that steric confinement provides the dominant rather than a negligible contribution at higher aggregate densities and significant quantitative corrections at lower densities, resulting in osmotic pressure isotherms that drastically differ from previous predictions.

1. Introduction

Semi-flexible polyelectrolytes play an important role in living systems and they are utilized in many industrial applications [1,2,3]. Their interactions are often studied in condensed columnar assemblies; most notably, examples include DNA [4,5,6,7], actin [4,8,9,10,11], xanthan [5,12] and some viruses such as TMV [4,13,14]. Aggregation of semi-flexible polyelectrolytes into columnar assemblies is also observed in nature, for instance, DNA in viral capsids [15,16,17] and actin filaments in microvilli [18]. The aggregation may be brought about by condensing agents [4,19,20] or osmotic pressure [6]. Intermolecular interactions determine physical properties of the aggregates [21] and play a crucial role in such phenomena as packing of meters of DNA inside a cell [22] or ejection of DNA from viruses [17].

It has long been noted that the interplay between interaction forces and undulations may be important for the physics of assemblies of semi-flexible polymers [23,24,25]. An early model of DNA aggregates approximated this interplay by treating a DNA molecule as a flexible chain confined within a hard cylinder with a radius dependent on the strength of the interaction [26], but this approximation did not fit the experimental data [27]. In Ref. [27] a more sophisticated model of a Gaussian chain confined by a harmonic potential was considered. The Gaussian chain model affords a relatively simple analytical solution, but the wormlike chain (WLC) model, based on elastic rod theory, provides a better description of the conformation of semi-flexible polyelectrolytes [28,29]. Theories of wormlike chains confined in a harmonic potential have been described in, for instance, Refs. [30,31,32,33,34]. In Ref. [33] a variational approximation was developed that combined the confined WLC model [30,32] with a more consistent description of the exponentially decaying electrostatic interaction between polyelectrolytes.

The latter polyelectrolyte theories focused primarily on the interplay between electrostatic forces and undulations at large surface separations, neglecting or not consistently incorporating shorter range image-charge forces and hard-core collisions between polyelectrolyte molecules [35]. However, because undulations bring semi-flexible polyelectrolytes into close contact, short-range forces are likely not to be negligible. Here, we demonstrate that these forces may qualitatively alter physical properties of aggregates regardless of the average surface separation.

We construct a variational theory for quasi-harmonic undulations, combining the ideas proposed in Refs. [30] and [33] with a more consistent treatment of electrostatics and hard-core collisions, at large undulation amplitudes, and a self-consistent calculation of the undulation amplitude based on the Gibbs-Bogoliubov inequality. Within this theory, we calculate pertinent pair correlation functions for the undulations and osmotic pressures of polyelectrolyte aggregates. In contrast to commonly held views, we find that undulations play a crucial role, not only in highly hydrated aggregates at large surface separations between the molecules, but also in dense columnar assemblies at surface separations comparable to, or smaller than, the Debye screening length. In the present paper we discuss the simpler case of interactions between cylindrical wormlike chains with uniform surface charge density. We will describe the more complex interplay between undulations and interactions for structured, helical polyelectrolytes in a separate study.

2. Model

Consider a columnar assembly of uniformly charged, semi-flexible polyelectrolytes that exhibit small displacements rj,l(z) from a hexagonal lattice parallel to the z axis (Fig. 1B). The indices j and l label the molecules in the x and y directions, correspondingly (Fig. 1B). Provided that the displacement rj,l(z) is confined within one lattice spacing (i.e., molecules do not entangle) and the tilt angle ψj,l of each molecule is small (|ψj,l|<<1), the energy cost of the undulations may be approximated by

δE{rj,l(z)}=δEb{rj,l(z)}+δEel{rj,l(z)}+δEst{rj,l(z)}, (1)

where δEb {rj,l(z)}, δEel {rj,l(z)}, and δEst {rj,l(z)} are the energetic costs of bending, the energy of electrostatic interactions between molecules, and the energy attributable to steric (hard-core) collisions between the molecules, respectively. For the purpose of the present study, we neglect entanglement and sharp bending of the molecules, which may introduce additional corrections to the free energy of polyelectrolyte aggregates [36].

Fig.1.

Fig.1

Pictures denoting: a.). Of part of an undulating molecule with respect to the z-axis; the orientation of its mean major axis. The curved dotted line represents the fluctuating position of the molecular axis and ψ(z) the tilt angle for one molecule is the angle the molecular axis makes with the z-axis. b.) A crossection of the undulating hexagonal lattice. The large dots are mean positions of the major axis of each molecule, which form the heaxagonal lattice with spacing R0. The small dots indicate the actual positions of each molecular axis. The deviation of an axis from its mean value for a molecule sitting at site {j,l} is given by rj,l (S). For each nearest neighbour about {j,l} is numbered from 1 to 6 with the index i. Then, each interaxial separation can be expressed as R0 + δrj,l,i (z), where for instance the change in intexaxial separation δ rj,l,1(z) of a molecule with respect to its nearest neibour i = 1 may be expressed as δrj,l,i(z)=Rj1,l+1Rj,l+rj1,l+1(z)rj,l(z)R0, where Rj,l are the position vectors of each mean axis in the lattice.

We describe the cost of bending within the wormlike chain approximation [28,29] as

δEb{rj,l(z)}=j,l0Ldzlpb2(d2rj,l(z)dz2)2, (2)

where L is the total length and lpb is the persistence length of the molecules.

2.1 Electrostatic energy

We calculate the electrostatic cost of undulations by pairwise summation of interactions of each molecule with its six nearest neighbors [21],

δEel{rj,l(z)}=12j,li=160Ldz[ucyl(R0,δrj,l(z))ucyl(R0,0)], (3)

where [37]

ucyl(R0,δr)2ξeff2lB{K0[κ(R0+δr)]j=(Kj[κ(R0+δr)])2Ij(κa)Kj(κa)}; (4)
ξeff=lBl~cκaK1(κa); (5)

ξeff is a dimensionless effective Manning parameter that can be calculated, e.g., as described in Ref. [38]; lB is the Bjerrum length (7 Å in water); l~c is the effective distance per unit charge along the molecules, which accounts for partial charge neutralization by bound/condensed counterions (e.g., within the Manning model l~c=lB, when the distance per unit fixed/bare charge lc is less than lB); a is the molecule radius (we assume that water-impenetrable core of the molecule has a low dielectric constant); κ is the inverse screening length in the assembly; In(x), Kn(x), I'n(x), and K'n(x) are the modified Bessel functions and their derivatives, respectively; the index i labels the six nearest neighbors of each molecule (j,l); and δrj,l,i(z) is the change in the interaxial distance between the molecule (j,l) and its neighbors due to undulations. The second term in Eq. (4) describes the contribution of image-charge forces associated with dielectric cores of the molecules. For simplicity, hereafter all energies and free energies are dimensionless, measured in the units of the thermal energy kBT.

Note that κ is generally larger than the inverse Debye screening length κD in the surrounding electrolyte solution outside the aggregate, because of the accumulation of additional counterions within the assembly [21]. The relationship between κ and κD can be approximated within the cylindrical cell model, in which the hexagonal Wigner-Seitz cell of each molecule is replaced with a cylinder of the same volume. In this model [39]

κ=κDcosh(eφskBT), (6)
tanh(eφskBT)=2ξeff[I0(κRs)K1(κRs)+I1(κRs)K0(κRs)][I1(κRs)K1(κRs)I1(κa)K1(κa)], (7)

where φs is the electrostatic potential at the surface of the effective cylindrical Wigner-Seitz cell and

Rs=R032π. (8)

is the radius of this cell. Note that the present, simplified form of Eqs. (6)–(8) is appropriate only for 1:1 electrolyte solutions; more general expressions can be found in Ref. [39].

2.2 Steric confinement (Helfrich-Harbich model)

To model steric confinement, we utilize the approximation proposed by Helfrich and Harbich [30,26]. Specifically, we replace δEst {rj,l(z)} with an effective Gaussian confinement potential

δEsteff{r(z)}=α020Ldz[r(z)2], (9)

in which the parameter α0 is selected to provide the same mean square displacement of uncharged molecules. Since the maximum displacement is ~ R0-2a, we may expect [30]

r(z)2=μ(R02a)2. (10)

Here μ<1 is a dimensionless constant, e.g., simulations for a confined wormlike chain suggest that μ ≈ 1/2 [40].

Calculation of <r(z)2> with the energy functional given by Eq. (9) and a comparison with Eq. (10) then yields

α0=1223μ43(R02a)83(lpb)13. (11)

2.3 Variational approximation

To calculate the free energy and correlation functions for the undulations, we utilize a variational approximation, in which we replace δE{rj,l(z)} with an effective energy,

δEeff{rj,l(z)}=12j,l0Ldz[lpb(d2rj,l(z)dz2)2+αrj,l(z)2], (12)

where α is a variational parameter. We then define the following free energy functional

FundlnδZeff+δE{rj,l(z)}δEeff{rj,l(z)}eff, (13)

where

δZeff=j,lDrj,l(z)exp[δEeff{rj,l(z)}] (14)

is the partition function for the effective energy; and < >eff indicates averaging with the statistical weight of exp [-δEeff {rj,l (z)}], i.e.,

f{rj,l(z)}eff1δZeffj,lDrj,l(z)f{rj,l(z)}exp[δEeff{rj,l(z)}] (15)

for any functional frj,l(z). Here, ∫ Drj,l(z) corresponds to path integration [41] over both components of the displacement vector rj,l(z).

The Gibbs-Bogoliubov inequality [42] states that the right hand side of Eq. (13) is always larger or equal to the exact free energy at any α. Therefore, we approximate the undulation free energy by the value of Fund minimized with respect to α.

As any model with unlimited fluctuations, this approximation effectively allows the cores of the molecules to overlap (e.g., at δrj,l,i (z) < 2aR0). Within this unphysical overlap region, δEel {rj,l (z)} is not defined; it has to be put into the model artificially. In previously used models [33,34,25], the expression for δEel {rj,l (z)} derived outside of the overlap was simply extended into the overlap region. As long as the contribution of this region to the free energy and relevant correlation functions is small, the unphysical overlap does not present a problem. However, this may not always be the case.

To avoid the latter problem, we define δEel {rj,l (z)} in such a way that it does not generate any artificial forces within the unphysical overlap region or at the boundary of this region. Specifically, δEel {rj,l (z)} remains constant throughout this region and equal to its value at the region boundary, i.e., we replace Eq. (3) with

δEel{rj,l(z)}12j,li=160Ldz[δuj,l,i(R0,z)], (16)

where

δuj,l,i(R0,z)={ucyl(R0,δrj,l,i(z))ucyl(R0,0),δrj,l,i(z)R02aucyl(R0,R02a)ucyl(R0,0),δrj,l,i(z)>R02aucyl(R0,2aR0)ucyl(R0,0),δrj,l,i(z)<2aR0}. (17)

In addition, provided that 2κa >> 1 and δr/R0<<1, we may approximate Eq. (4) with

ucyl(R0,δr)ξeff22πexp[κ(R0+δr)]lB(κR0)12+ξeff2πΩ0exp[2κ(R0+δr)]lB(κR0) (18)

where

Ω0=ΣjIj(κa)Kj(κa). (19)

Equations (12)–(14) combined with (1), (2), (9), (11), (16) and (17) fully define the model we use for calculating all relevant correlation functions and the contribution of undulations to the aggregate free energy.

2.4 Osmotic pressure

The effective osmotic pressure of the aggregate is given by

Π=Π0+Πund, (20)

where

Π0=κD2kBT4πlB[cosh(eφskBT)1] (21)

is the osmotic pressure of an aggregate of straight molecules [43], and

Πund=kBT(FundVw)κ=kBTNLR03(FundR0)κ (22)

is the contribution from undulations. In Eq. (22), the kBT prefactor provides dimensionality to the dimensionless undulation free energy Fund, and Fund is calculated from Eq. (13) and by minimization with respect to α. The partial derivative of Fund is taken at fixed κ to account for the averaging over the counter-ion degrees of freedom.

3. Results

3.1 Pair correlation functions

Within our model, the pair correlation function for undulations is given by (see Appendix)

rj,l(z)rm,n(z)=δj,mδl,nG(zz), (23)

where

G(zz)=d22cos(zz2λBπ4)exp(zz2λB), (24)
d=(4α3lpb)18 (25)

is the rms undulation amplitude, and

λB=(2d2lpb)13 (26)

is the undulation correlation length (first introduced by Odijk [44,45]).

Note that our approximation for the electrostatic energy is valid only at small fluctuations of the tilt angle ψ of each molecule with respect to the vertical position (Fig. 1A). The mean square amplitude of the latter fluctuations is given by (Appendix)

ψ2=12(dr(z)dz)2=(d4lpb)23, (27)

indicating that the present theory is valid at

(d4lpb)131. (28)

It may be possible to account for undulations with larger tilt angles utilizing the ideas developed in Ref. [46], but such a theory would be substantially more complicated.

3.2 Free energy and amplitude of undulations

As shown in the Appendix, the calculation of Fundl from Eqs. (1)–(3), (9), (11), (12)–(18), and (25) yields

FundNL1253d23(lpb)13[3+d83μ43(R02a)83]+3ξeff22πexp(κR0)lB(κR0)12γ(κ(R02a),κd),+3ξeff2πΩ0exp(2κR0)lB(κR0)γ(2κ(R02a),2κd) (29)

where

γ(x,y)=1+cosh(x)[1erf(xy2)]+12ey22[erf(xy2y2)+erf(xy2+y2)]. (30)

We expressed Fund as a function of the rms undulation amplitude d rather than as a function of the variational parameter α, by utilizing Eq. (25), thus establishing a relationship between d and α. Minimization of Fund with respect to α is equivalent to minimization with respect to d. From this minimization we find

lB223κ2d83(lpb)13[1d83μ43(R02a)83]=32πξeff2exp(κR0)(κR0)12η(κ(R02a),κd),+12πξeff2Ω0exp(2κR0)(κR0)η(2κ(R02a),2κd) (31)

where

η(x,y)=12ey22[erf(12(xy+y))+erf(12(xyy))]2y2πsinh(x)exp(x22y2). (32)

One may now calculate the undulation amplitude by solving Eqs. (31) and (32) with respect to d. For uncharged molecules (ξeff=0), we recover d2rj,l2=μ(R02a)2 (c.f., Eq. (10)). Because η(x,y) is positively defined, electrostatic repulsion between charged molecules reduces the undulation amplitude as one might expect.

Figure 2 illustrates the dependence of the undulation amplitude d on the average surface-to-surface separation between the molecules (R0−2a) at different linear charge densities ξ=lBl~c (Fig. 2A) and different bending persistence lengths lpb (Fig. 2B). The results are represented in dimensionless form, in which they are nearly independent of the specific values of κD and a.

Fig.2.

Fig.2

Polyelectrolyte undulations in aggregates. A. Dependence of the rms undulation amplitude d on average surface separation between the molecules (R0–2a) at different values of the renormalized Manning parameter ξ=lBl~c; calculated with κ self-consistently determined from Eqs. (6)–(8) (solid lines) and with κ = κD (dotted lines). Although κ rapidly increases with decreasing surface separation (inset), the effect of this change in κ on the undulation amplitude is relatively minor even at large ξ. Line colors in the inset correspond to the same ξ as in the main panel. B. Undulation amplitude calculated within our model (Eq. (31), solid lines) and within the previously proposed [33] approximation for highly charged polyelectrolytes (Eq. (33), dashed lines). The following parameters were used for all calculations: 1/κD = 7 Å, a = 10.5 Å (2κDa=3). These values and κDlpb=71(lpb500) correspond to DNA in physiological saline.

As follows from Eq. (31), d depends explicitly on ξ and lpb through a single parameter, ξ2(lpb)13. Although d also depends implicitly on ξ through κ (Fig. 2A, inset), the latter effect is weak for moderately charged molecules (ξ < 2), as shown by the dotted-line curve in Fig. 2A. For such molecules, increasing the charge density (Fig. 2A) and increasing the persistence length (Fig. 2B) have similar effects on the undulation amplitude.

3.3 Highly charged polyelectrolytes

Formally, Eqs. (29)–(32) can be simplified for highly charged molecules in the case of strong electrostatic confinement (d2<<μ(R02a)2,κd2<<R02a) and negligible image-charge forces. In this case, γ(x,y)η(x,y)ey22 and

κlB32πξeff2(4κlpb)13(κd)83exp(κR0+κ2d22)(κR0)12. (33)

This equation is similar to the one reported in [33], except the left hand side in our case is larger by a factor of 3/2 and we do not assume κ ≈ κD. The numerical discrepancy appears to be related to the difference between the ad hoc approximation for the free energy in Ref. [33] and the more rigorous approximation based on the Gibbs-Bogoliubov inequality utilized in the present work.

Eq. (33) works reasonably well at ξ >> 1 (Fig. 2B). At ξ ~ 1, Eq. (33) fails in two ways: (i) At surface separations up to ~ 1/κD, the undulation amplitude is determined primarily by steric confinement even for very rigid polyelectrolytes. At these distances, Eq. (33) predicts unphysically large d, exceeding the undulation amplitude for uncharged molecules (which is restricted by steric collisions). As a result, it predicts a large unphysical enhancement of the mean electrostatic force. (ii) At large separations, Eq. (33) strongly underestimates d for flexible polyelectrolytes; it does not consistently account for the steric confinement, extending the exponential electrostatic interaction into the region where the molecules overlap (see also the discussion of Eqs. (16) and (17)). Numerical comparison of Eqs. (31) and (32) with Eq. (33) suggests that neither steric confinement nor image forces can be neglected for flexible polyelectrolytes with ξ ~ 1 at any separation.

3.4 Osmotic pressure

The aggregate osmotic pressure is given by Eqs. (20)–(22). After substitution of Eqs. (29) and (30) into Eq. (22) we find that the contribution of undulations to the osmotic pressure is described by

Πund=Πundst+Πundel, (34)

where

Πundst24333kBTd2μ43(lpb)13R0(R02a)113 (35)

results from the steric collisions between the molecules and

Πundels3kBTκξeff2R0lB[2πexp(κR0)(κR0)12χ(κ(R02a),κd)+2πΩ0exp(2κR)(κR0)χ(2κ(R02a),2κd)], (36)

with

χ(x,y)=1+ex[1erf(xy2)]+12ey22[erf(xy2y2)+erf(xy2+y2)], (37)

is the electrostatic component of the undulation osmotic pressure.

As illustrated in Fig. 3, undulations provide the dominant contribution to the osmotic pressure at small and large separations, regardless of polyelectrolyte charge and rigidity. At small separations, via steric collisions, the entropic cost associated with the elimination of undulations diverges as the separation between molecules approaches zero, while the electrostatic interaction energy remains finite. At large separations, undulations extend and enhance the short-range electrostatic repulsion by bringing parts of the molecules much closer together. Although, in this case, undst is small compared with undels, steric collisions still play an important role in limiting electrostatic interactions, reducing the undulation amplitude and limiting the enhancement of the electrostatic interaction (through the erf-functions in Eq. (37)).

Fig 3.

Fig 3

Components of osmotic pressure in polyelectrolyte aggregates. Note that undulations are important both at large and small surface separations, shifting the whole osmotic pressure curve (Π) compared to the osmotic pressure in aggregates of straight, infinitely rigid rods (=Π0). The calculations were performed with the same parameter values as in Fig. 2. Note that κD2kBTlB12MPa=120 bar at κD1lB7.

4. Discussion

The present study suggests that undulations significantly contribute to interactions between polyelectrolytes at all surface separations, even in well-ordered columnar aggregates (Fig. 3).

The undulations become negligible only for rigid, highly charged (ξ >>1) polyelectrolytes and even then only at surface separations close to 1/κD, which are not common conditions since counterion condensation and binding reduce ξ in polyelectrolytes with a high density of fixed surface charges. Indeed, ξ ≤ 1 for all polyelectrolytes in the limit of infinitely dilute salt [47]. Large ξ is possible only at finite salt concentrations in the absence of counterions that may strongly bind to the polyelectrolyte [48].

The effect of enhancement of electrostatic interactions by undulations at large separations was well recognized before us [27,33,25]. However, the contribution from undulations at intermediate and small separations was not. We now find that, depending on the polyelectrolyte charge and rigidity, undulations may significantly shift the whole osmotic pressure curve. Figs. 3A and 3B illustrate this effect for uniformly charged worm-like chains with DNA-like diameter and rigidity at physiological ionic strength. The value of ξ = 0.5 corresponds to 87% neutralization of DNA charge by counterions, a value that models the binding (or condensation) of Ca2+, Mg2+ and other divalent and polyvalent counterions under physiological conditions; ξ = 2 corresponds to 50% neutralization of DNA charge, which approximates counterion condensation in monovalent salt solutions at physiological ionic strength [21,48,49].

These findings are in sharp contrast with the commonly held view [27] that undulations are important only in hydrated aggregates at large separations. Our model differs from previous theories in several aspects, e.g., it accounts for image-charge forces due to water-impermeable cores of polyelectrolyte molecules. However, its most important distinction, primarily responsible for the discrepancy, is a more consistent treatment of steric collisions between the molecules. Previous theories [27,33,34, 25] effectively presumed a purely electrostatic suppression of otherwise unlimited undulations, resulting in an infinite undulation amplitude for uncharged molecules and a finite, nonzero undulation amplitude at zero surface separation between charged molecules. Our model accounts for steric confinement of undulations by hard-core collisions, introducing the corresponding steric confinement energy based on the idea proposed by Helfrich and Harbich for uncharged worm-like chains [30]. Like the previous theories, our model is based on a quasi-Gaussian description of undulations, but it captures the limiting cases of uncharged molecules and small surface separations (Fig. 2).

In our calculations, we have assumed local hexagonal packing of the molecules in a columnar assembly. We should bear in mind, however, that undulations may induce a melting (deconfinement) phase transition at d = cR0 (the Lindeman criterion), where the constant c has been estimated to be c ~ 0.15 – 0.48 [50,34,25]. A theory for this transition was proposed in Ref. [24]. It included an estimate of steric interactions, but it was based on an oversimplified description of the hexagonal phase and electrostatic interactions. Perhaps, our work in conjunction with Ref. [24] may form the basis of a more accurate theory [51].

We believe that the present study makes an important step towards a better understanding of the physics of interactions between uniformly charged worm-like chains in a continuum electrolyte solution with point-like, weakly coupled salt ions. At the same time, it is worthwhile to keep in mind the limitations of such a description for real polyelectrolytes. This description is likely to be reasonable, at least qualitatively, when the distance scales for neglected effects (e.g., ion size) are smaller than the pertinent characteristic lengths in the theory, e.g., the effective screening length and surface-surface separation [52].

Eqs. (34)–(37) provide a recipe for understanding the physics of osmotic pressures in polyelectrolyte aggregates. The osmotic pressure is not just a commonly used method for studying such aggregates (see, e.g., [21,53,54] and refs. therein) but also a functionally important property of cartilage and other natural and engineered systems [55,56,57]. Eqs. (34)–(37) predict osmotic pressures (Fig. 3) that are qualitatively and semi-quantitatively consistent with the measurements [58,59], but we urge caution in more detailed quantitative comparisons with experiments. Eqs. (34)–(37) have not been designed for and may not be sufficiently accurate for analysis of specific experimental data. For the latter purpose, the approximations of uniform surface charge density and hexagonal packing may not be appropriate; and these equations may have to be modified to account for specific details of the system under study.

One factor that has to be taken into account for practical applications is the polyelectrolyte structure, particularly when the distance between regularly organized surface charges exceeds the effective screening length inside aggregates (3–10 Å in typical experiments). In the latter case, the uniform surface charge approximation may, e.g., severely underestimate image-charge forces and completely miss crucially important effects associated with mutual alignment of the molecules [21]. We will illustrate these effects and ways to modify the model to account for them in a separate study of DNA, for which the surface charge periodicity defined by the 34 Å helical pitch is much larger than the screening length, resulting in rather nontrivial effects of undulations on helix-specific interactions.

Similarly, the extent and patterns of counterion binding/condensation may need to be treated at a more sophisticated level than simply by employing an effective Manning parameter, e.g., long-range correlations between multi- and polyvalent ions at the polyelectrolyte surface may need to be explicitly incorporated into the theory, when such counterions are present in the solution [60]. Work in this direction is currently in progress.

5. Conclusions

1. In contrast to previous models, the present theory describes the intermolecular interactions within an assembly of undulating, semi-flexible polyelectrolytes for both the limiting cases of uncharged and highly-charged molecules as well as for polyelectrolytes with more practically important, intermediate linear charge densities.

2. In the limit of small surface separations, hard-core collisions provide the dominant interaction between undulating, uniformly-charged, semi-flexible polyelectrolytes confined in a columnar aggregate, regardless of the polyelectrolyte charge density.

3. At effective linear charge densities comparable to or smaller than one elementary charge per Bjerrum length, the steric confinement due to hard-core collisions significantly affects aggregate osmotic pressures at all interaxial separations.

Acknowledgements

This work was supported in part by the Max-Planck Society (Lee), the Intramural Research Program of NICHD, NIH (Leikin), and the Alexander von Humboldt Foundation (Wynveen).

Appendix

Correlation functions

After representing rj,l (z) by Fourier series,

rj,l(z)=kr~j,leikz, (A.1)

the expression for the effective free energy (Eq. (12)) can be rewritten as

δEeff{rj,l(k)}=L2j,lk(lpbk4+α)r~j,l(k)r~j,l(k). (A.2)

Applying the equipartition theorem [1],

r~j,l(k)r~n,m(k)=2δj,nδl,mL(lpbk4+α), (A.3)

we then find

(rj,l(z)rj,l(z))2=2G(0)2G(zz), (A.4)

where (in the limit of L→∞)

G(zz)=1πdkexp(ik(zz))lpbk4+α. (A.5)

Contour integration of Eq. (A.5) in the top half on the complex plane yields

G(zz)=(λB)3lpbexp(zz2λB)cos(zz2λBπ4), (A.6)

with the correlation length for bending fluctuations,

λB=(lpbα)14. (A.7)

Since G (zz′) exponentially decreases at ZZλB, the centerline of each rod is effectively confined within a cylinder with the radius

d=G(0). (A.8)

From Eqs. (A.6)–(A.8) we recover Eqs. (25) (26) of the main text. Note that this approximation may also be used at finite L provided that L>>λB.

From similar calculations we find another useful correlation function

drj,l(z)dzdrj,l(z)dz=λBlpbexp(zz2λB)cos(zz2λB+π4), (A.9)

The special case of this correlation function at z = z′ defines the mean square fluctuations of the tilt angle ψ given by Eq. (27) of the main text.

Variational free energy

To calculate ln(δZeff), we use that

1NLαln(δZeff)=r2eff2=d22. (A.10)

After integration, we find

1NLln(δZeff)=(2d2lpb)13, (A.11)

where the integration constant was selected from the assumption that Fund=0 at Fund=0atlpb.

To calculate δE{rj,l(z)}δEeff{rj,l(z)}eff, we utilize that for any ergodic random process rj,l (z) [2],

f{rj,l(z)}efff~(rj,l(z))dzeff=Lf~(rj,l(z))eff=Lf~(r)P(r)dr (A.12)

whrere f~(r)is an arbitrary function of r. The probability density P(r) for a Gaussian random process is given by [2]

P(r)=14π2dueiureiur(z)eff=1πer2r2=1πer2d2, (A.13)

where we took into account that r(z) is a 2D vector. From Eqs.(2) (9) (12) (13) (16), we find

1NLδE{rj,l(z)}δEeff{rj,l(z)}eff=α0α2d2+3δuj,l,i(R0,z)eff, (A.14)

where δuj,l,i (R0, z) is given by Eq. (17) of the main text.

Since δuj,l,i(R0,z)eff does not depend on j, l, or i, it is to calculate δuj,l,3(R0,z)eff (see Fig. 1b). At small xj,l and yj,l displacements of the molecules (j,l) and (j+1,l), i.e., |xj+1,l - xj,l|/R0 << 1 and |yj+1,l - yj,l|/R0 << 1,

δrj,l,3(z)(xj+1,lxj,l), (A.15)

where we neglected the second order terms with respect to xj+1,lxj,l and yj+1,lyj,l. After substituting Eqs. (17), (A.14), (A.15) into Eqs. (A.12), (A.13), we then find

δuj,l,3(R0,z)eff1d2π×[ucyl(R0,2aR0)2aR0exp(δx22d2)d(δx)]+2aR0R02aucyl(R0,δx)exp(δx22d2)d(δx)[ucyl(R0,R02a)R02aexp(δx22d2)d(δx)], (A.16)

where we changed the integration variables from xj+1,l and xj,l to δx= xj+1,lxj,l and xj+1,l+xj,l and integrated out xj+1,l+xj,l.

At κa >> 1, substitution of Eq. (18) into Eq. (A.16) yields

δuj,l,3(R0,z)effξeff22πexp(κR0)lB(κR0)12g(κ(R02a),κd)+ξeff2πΩ0exp(2κR0)lB(κR0)g(2κ(R02a),2κd) (A.17)

where g(x,y) is defined by Eq. (30). Finally, after substituting Eqs. (A.11), (A.14), and (A.17) into Eq. (13), we arrive at Eq. (29) of the main text.

Footnotes

1

L. D. Landau, E. M. Lifshitz, Statistical Physics (Pergamon, Oxford, 1969)

2

Feynman, R. P., and Hibbs, A. R., Quantum Mechanics and Path Integrals (McGraw-Hill, New York, 1965)

References

  • 1.Knaapila M, Almasy L, Garamus VM, Pearson C, Pradhan S, Petty MC, Scherf U, Burrows HD, Monkman AP. J. Phys. Chem. B. 2006;110:10248. doi: 10.1021/jp0560563. and references contained therin. [DOI] [PubMed] [Google Scholar]
  • 2.Guilleaume B, Blaul J, Wittemann M, Rehahn M, Ballauff M. J. Phys: Condens. Matter. 2000;12:A245. doi: 10.1140/epje/i2001-10105-8. [DOI] [PubMed] [Google Scholar]
  • 3.Zaroslov YD, Gordeliy VI, Kuklin AI, Islamov AH, Philippova OE, Khokhlov AR, Wegner G. Macromolecules. 2002;35:4466. [Google Scholar]
  • 4.Tang JX, Ito T, Tao T, Traub P, Janmey PA. Biochemisty. 1997;36:12600. doi: 10.1021/bi9711386. [DOI] [PubMed] [Google Scholar]
  • 5.Livolant F, Bouligand Y. J. Phys. France. 1986;47:1813. [Google Scholar]
  • 6.Podgornik R, Rau DC, Parsegian VA. Macromolecules. 1989;22:1780. [Google Scholar]
  • 7.Rau DC, Lee B, Parsegian VA. Proc. Natl. Acad. Sci. USA. 1984;81:2621. doi: 10.1073/pnas.81.9.2621. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Tang JX, Janmey PA. J. Bio. Chem. 1996;271:8556. doi: 10.1074/jbc.271.15.8556. [DOI] [PubMed] [Google Scholar]
  • 9.Kwon HJ, Kakugo A, Shikinaka K, Osada Y, Gong JP. Biomacromolecules. 2005;6:3005. doi: 10.1021/bm050320g. [DOI] [PubMed] [Google Scholar]
  • 10.Angelini TE, Liang H, Wriggers W, Wong GCL. Proc. Natl. Acad. Sci. USA. 2003;100:8634. doi: 10.1073/pnas.1533355100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Angelini TE, Liang H, Wriggers W, Wong GCL. Eur. Phys. J. 2005;16:389. doi: 10.1140/epje/i2004-10097-9. [DOI] [PubMed] [Google Scholar]
  • 12.Rau DC, Parsegian VA. Science. 1990;249:1278. doi: 10.1126/science.2144663. [DOI] [PubMed] [Google Scholar]
  • 13.Nedoluzhko A, Douglas T. J. Inorg. Biochem. 2001;84:233. doi: 10.1016/s0162-0134(01)00174-x. [DOI] [PubMed] [Google Scholar]
  • 14.Parsegian VA, Brenner SL. Nature. 1976;259:632. doi: 10.1038/259632a0. [DOI] [PubMed] [Google Scholar]
  • 15.Klimenko SM, Tikchonenko TI, Andreev VM. J. Mol. Biol. 1967;23:523. doi: 10.1016/s0022-2836(67)80122-0. [DOI] [PubMed] [Google Scholar]
  • 16.Earnshaw WC, Harrison SC. Nature. 1977;268:598. doi: 10.1038/268598a0. [DOI] [PubMed] [Google Scholar]
  • 17.Tzlil S, Kindt JT, Gelbart WM, Ben-Shaul A. Biophys. J. 2003;84:1616. doi: 10.1016/S0006-3495(03)74971-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Glenny JR, Kaulfus P, Matsudaira P, Weber K. J. Bio. Chem. 1981;256:9283. [PubMed] [Google Scholar]
  • 19.Widom J, Baldwin RL. J. Mol. Biol. 1980;144:431. doi: 10.1016/0022-2836(80)90330-7. [DOI] [PubMed] [Google Scholar]
  • 20.Todd BA, Parsegian VA, Shirahata A, Thomas TJ, Rau DC. Biophys. J. 2008;94:4775. doi: 10.1529/biophysj.107.127332. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Kornyshev AA, Lee DJ, Leikin S, Wynveen A. Rev. Mod. Phys. 2007;79:943. [Google Scholar]
  • 22.Bloomfield VA. Curr. Opin. Struct. Biol. 1996;6:334. doi: 10.1016/s0959-440x(96)80052-2. [DOI] [PubMed] [Google Scholar]
  • 23.Selinger JV, Bruinsma RF. Phys. Rev. A. 1991;43:2910. doi: 10.1103/physreva.43.2910. [DOI] [PubMed] [Google Scholar]
  • 24.Selinger JV, Bruinsma RF. Phys. Rev. A. 1991;43:2922. doi: 10.1103/physreva.43.2922. [DOI] [PubMed] [Google Scholar]
  • 25.Strey HH, Parsegian VA, Podgornik R. Phys. Rev. E. 1999;59:999. [Google Scholar]
  • 26.Stigter D. Cell Biochem. Biophys. 1987;11:139. doi: 10.1007/BF02797120. [DOI] [PubMed] [Google Scholar]
  • 27.Podgornik R, Parsegian VA. Macromolecules. 1990;23:2265. [Google Scholar]
  • 28.Kratky O, Porod G. Rec. Trav. Chim. Pays-Bas. 1949;68:1106. [Google Scholar]
  • 29.Fixman M, Kovac J. J. Chem. Phys. 1973;58:1564. [Google Scholar]
  • 30.Helfrich W, Harbich W. Chem. Scr. 1985;25:32. [Google Scholar]
  • 31.Kleinert H. J. Math. Phys. 1986;27:3003. [Google Scholar]
  • 32.Burkhardt TW. J. Phys. A: Math. Gen. 1995;28:L629. [Google Scholar]
  • 33.Odijk T. Biophys. Chem. 1993;46:69. [Google Scholar]
  • 34.Jain S, Nelson DR. Macromolecules. 1996;29:8523. [Google Scholar]
  • 35.In Ref. [25] short-range hydration forces were included in the harmonic approximation, but image-charge interaction and steric interactions were still neglected.
  • 36.Odijk T. Macromolecules. 1993;26:6897. [Google Scholar]
  • 37.Kornyshev AA, Leikin S. J. Chem. Phys. 1997;107:3656. [Google Scholar]
  • 38.Philip JR, Wooding RA. J. Chem. Phys. 1970;52:953. [Google Scholar]
  • 39.Cherstvy AG, Kornyshev AA, Leikin S. J. Phys. Chem. B. 2002;106:13362. doi: 10.1021/jp0380475. [DOI] [PubMed] [Google Scholar]
  • 40.For purely steric interactions this approximation leads to a free energy Fs=ckBTLD23(lpb)13, where D is the effective diameter of the cylinder. From simulations [M. Dijkstra, D. Frenkel, H. N. W. Lekkerkerker, Physica A 193, 374 (1993)], the value of the constant is c = 2.46 ± 0.07. If we suppose that the confinement diameter is D = 2(R0 – 2a) and μ = 1/2, we find c≃ 2.51, in good agreement with the simulation results.
  • 41.Feynman RP. Statistical Mechanics (A set of lectures) Westview Press; New York: 1998. [Google Scholar]
  • 42.Bogoliubov NN. Dokl. Akad. Nauk USSR. 1958;119:244. [Google Scholar]
  • 43.Katchalsky A. Pure Appl. Chem. 1971;26:327. [Google Scholar]
  • 44.Odijk T. Macromolecules. 1983;16:1340. [Google Scholar]
  • 45.Odijk T. Macromolecules. 1986;19:2313. [Google Scholar]
  • 46.Miklavic SJ. Phil. Trans. R. Soc. London A. 1994;348:209. [Google Scholar]
  • 47.Manning GS. J. Chem. Phys. 1969;51:924. [Google Scholar]
  • 48.Netz RR, Orland H. Eur. Phys. J. E. 2003;11:301. doi: 10.1140/epje/i2002-10159-0. [DOI] [PubMed] [Google Scholar]
  • 49.Hsiao C, Tannenbaum M, Van Deusen H, Hershkovitz E, Perng G, Tannenbaum A, Williams LD. In: Nucleic Acid Metal Ion Interactions. Hud N, editor. The Royal Society of Chemistry; London: 2008. pp. 1–35. [Google Scholar]
  • 50.Odijk T. Europhys. Lett. 1993;24:177. [Google Scholar]
  • 51.For instance, an additional term, β (drj, l (z) / dz)2, may be introduced into our effective theory (Eq. (12)) to describe nematic aggregates in which we may no longer assume that ψ1. Here β is the order parameter for nematic ordering [24].
  • 52.At small surface-surface separations, factors like the discreteness of the solvent may become important.
  • 53.Parsegian VA, Rand RP, Fuller NL, Rau DC. Methods Enzymol. 1986;127:400. doi: 10.1016/0076-6879(86)27032-9. [DOI] [PubMed] [Google Scholar]
  • 54.Yethiraj A. J. Phys. Chem. B. 2009;113:1539. doi: 10.1021/jp8069964. [DOI] [PubMed] [Google Scholar]
  • 55.Basser PJ, Schneiderman R, Bank RA, Wachtel E, Maroudas A. Arch. Biochem. Biophys. 1998;15:207. doi: 10.1006/abbi.1997.0507. Arch Biochem Biophys. 1998 Mar 15;351(2):207-19. [DOI] [PubMed] [Google Scholar]
  • 56.Elliott GF, Hodson SA. Rep. Progr. Phys. 1998;61:1325. [Google Scholar]
  • 57.Ateshian GA, Soltz MA, Mauck RL, Basalo IM, Hung CT, Lai WM. Transport in Porous Media. 2003;50:5. [Google Scholar]
  • 58.Rau DC, Parsegian VA. Biophys. J. 1992;61:246–260. doi: 10.1016/S0006-3495(92)81832-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Podgornik R, Strey HH, Gawrisch K, Rau DC, Rupprecht A, Parsegian VA. Proc Natl Acad Sci USA. 1996;93:4261. doi: 10.1073/pnas.93.9.4261. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Ossawa F. Polyelectrolytes. Marcel Dekker; New York: 1971. [Google Scholar]

RESOURCES