Adsorption of Polyelectrolytes at Oppositely Charged Surfaces

15
Adsorption of Hydrophobic Polyelectrolytes at Oppositely Charged Surfaces Andrey V. Dobrynin* and Michael Rubinstein Department of Chemistry, University of North Carolina, Chapel Hill, North Carolina 27599-3290 Received August 20, 2001 ABSTRACT: We develop a scaling theory of adsorption of necklace-like hydrophobic polyelectrolytes at an oppositely charged hydrophilic and hydrophobic surfaces. At low surface charge densities we predict a two-dimensional adsorbed layer with thickness determined by the balance between electrostatic attraction to the charged surface and chain entropy. At high surface charge densities we expect a 3-dimensional layer with polymer density profile determined either by the balance between two-body monomer-monomer attraction or by electrostatic attraction to the surface and three-body monomer- monomer repulsion. These different stabilizing mechanisms result in the nonmonotonic dependence of the layer thickness on the surface charge density. For adsorption of polyelectrolyte chains from salt solutions the screening of the electrostatic repulsion between adsorbed polyelectrolyte chains results in large overcompensation of the surface charge for two-dimensional adsorbed layers. The polymer surface coverage for this regime increases with increasing salt concentration. The opposite trend is predicted for 3-D adsorbed layers where the polyelectrolyte surface excess decreases with increasing salt concentration. 1. Introduction Polyelectrolytes are polymers with ionizable groups that in solutions with high dielectric constant (e.g., in aqueous solutions) dissociate, releasing counterions into solutions and leaving charged monomers on polymer backbones. 1,2 The intramolecular electrostatic repulsion between charged monomers forces chains to adopt elongated conformations. The behavior of polyelectro- lytes in solutions strongly depends on the solvent quality for polymer backbone. In a good or Θ-solvent the chain size can be estimated from the balance of the chain elasticity and electrostatic repulsion between charged monomers. 3-6 Polyelectrolyte size increases with in- creasing fraction of charged monomers on polymer chains. The situation is qualitatively different for poly- electrolytes in a poor solvent. The shape of these polymers is determined by the interplay between elec- trostatic and polymer-solvent interactions. Polyelec- trolyte chains in a poor solvent adopt a necklace-like structure of beads connected by narrow strings 7-14 as a compromise between electrostatic and surface energies. Larger polymeric globules break up into smaller ones with increasing fraction of charged monomers on the chain. The surface energy of the polymer-solvent interface in equilibrium globules is balanced by the electrostatic repulsion between charged monomers. The necklace-like structure was shown to manifest itself in unique scaling laws in semidilute polyelectrolyte solutions for the concentration range where the correla- tion length Œ of a solution is of the order of the string length between two neighboring beads. 9 In this concen- tration range the correlation length Œ decreases with polymer concentration c as Œ c -1/3 , viscosity is con- centration-independent, and self-diffusion coefficient D increases with concentration D c 1/3 in both unen- tangled and entangled solutions. Another interesting feature of the pearl necklace structure is the sawtooth force-extension curve predicted for hydrophobic poly- electrolytes under the applied external force. 15-17 This type of force-extension curve formally corresponds to the abrupt transitions between necklaces with different number of beads resulting in a series of abrupt length jumps with increasing applied force. The next step toward a complete picture of polyelec- trolytes in a poor solvent is to provide a theoretical framework for understanding their behavior at surfaces and interfaces. The latest progress in the theory of polyelectrolyte adsorption 18-22 provides an unique op- portunity to apply these results to adsorption of necklace- like molecules. In the present paper we develop a theory of adsorption of polyelectrolytes from a poor solvent at an oppositely charged hydrophilic and hydrophobic surfaces and calculate the adsorption diagrams of these systems as functions of the surface charge density and the salt concentration. The paper is organized as follows. In the next section we briefly review the necklace model of polyelectrolytes in a poor solvent and estimate the free energy due to confinement of a necklace molecule into a slit between parallel plates. Sections 3-5 discuss the adsorption of hydrophobic polyelectrolytes at hydrophilic surfaces when polymer adsorption is caused exclusively by electrostatic interactions between adsorbing surface and oppositely charged polyelectrolytes. In sections 3 and 4 we apply the concept of the strongly correlated Wigner liquid 21-26 to describe the structure of dilute and semi- dilute adsorbed layers formed by necklace-like chains. In section 5 we use a mean-field approach 21-23,27-29 to calculate polymer density profile and surface coverage for concentrated adsorbed layers of overlapping beads. In section 6 we discuss the effects of van der Waals attractions between adsorbing surface and polymer backbones on its adsorption and calculate the adsorption diagram for the case of adsorption of hydrophobic polyelectrolytes at hydrophobic oppositely charged sur- faces. ² Permanent address: Institute of Materials Science and De- partment of Physics, University of Connecticut, Storrs, CT 06269- 3136. 2754 Macromolecules 2002, 35, 2754-2768 10.1021/ma0114943 CCC: $22.00 © 2002 American Chemical Society Published on Web 02/22/2002

Transcript of Adsorption of Polyelectrolytes at Oppositely Charged Surfaces

Adsorption of Hydrophobic Polyelectrolytes at Oppositely ChargedSurfaces

Andrey V. Dobrynin*,† and Michael Rubinstein

Department of Chemistry, University of North Carolina, Chapel Hill, North Carolina 27599-3290

Received August 20, 2001

ABSTRACT: We develop a scaling theory of adsorption of necklace-like hydrophobic polyelectrolytes atan oppositely charged hydrophilic and hydrophobic surfaces. At low surface charge densities we predicta two-dimensional adsorbed layer with thickness determined by the balance between electrostaticattraction to the charged surface and chain entropy. At high surface charge densities we expect a3-dimensional layer with polymer density profile determined either by the balance between two-bodymonomer-monomer attraction or by electrostatic attraction to the surface and three-body monomer-monomer repulsion. These different stabilizing mechanisms result in the nonmonotonic dependence ofthe layer thickness on the surface charge density. For adsorption of polyelectrolyte chains from saltsolutions the screening of the electrostatic repulsion between adsorbed polyelectrolyte chains results inlarge overcompensation of the surface charge for two-dimensional adsorbed layers. The polymer surfacecoverage for this regime increases with increasing salt concentration. The opposite trend is predicted for3-D adsorbed layers where the polyelectrolyte surface excess decreases with increasing salt concentration.

1. IntroductionPolyelectrolytes are polymers with ionizable groups

that in solutions with high dielectric constant (e.g., inaqueous solutions) dissociate, releasing counterions intosolutions and leaving charged monomers on polymerbackbones.1,2 The intramolecular electrostatic repulsionbetween charged monomers forces chains to adoptelongated conformations. The behavior of polyelectro-lytes in solutions strongly depends on the solvent qualityfor polymer backbone. In a good or Θ-solvent the chainsize can be estimated from the balance of the chainelasticity and electrostatic repulsion between chargedmonomers.3-6 Polyelectrolyte size increases with in-creasing fraction of charged monomers on polymerchains. The situation is qualitatively different for poly-electrolytes in a poor solvent. The shape of thesepolymers is determined by the interplay between elec-trostatic and polymer-solvent interactions. Polyelec-trolyte chains in a poor solvent adopt a necklace-likestructure of beads connected by narrow strings7-14 as acompromise between electrostatic and surface energies.Larger polymeric globules break up into smaller oneswith increasing fraction of charged monomers on thechain. The surface energy of the polymer-solventinterface in equilibrium globules is balanced by theelectrostatic repulsion between charged monomers.

The necklace-like structure was shown to manifestitself in unique scaling laws in semidilute polyelectrolytesolutions for the concentration range where the correla-tion length ê of a solution is of the order of the stringlength between two neighboring beads.9 In this concen-tration range the correlation length ê decreases withpolymer concentration c as ê ∼ c-1/3, viscosity is con-centration-independent, and self-diffusion coefficient Dincreases with concentration D ∼ c1/3 in both unen-tangled and entangled solutions. Another interesting

feature of the pearl necklace structure is the sawtoothforce-extension curve predicted for hydrophobic poly-electrolytes under the applied external force.15-17 Thistype of force-extension curve formally corresponds tothe abrupt transitions between necklaces with differentnumber of beads resulting in a series of abrupt lengthjumps with increasing applied force.

The next step toward a complete picture of polyelec-trolytes in a poor solvent is to provide a theoreticalframework for understanding their behavior at surfacesand interfaces. The latest progress in the theory ofpolyelectrolyte adsorption18-22 provides an unique op-portunity to apply these results to adsorption of necklace-like molecules. In the present paper we develop a theoryof adsorption of polyelectrolytes from a poor solvent atan oppositely charged hydrophilic and hydrophobicsurfaces and calculate the adsorption diagrams of thesesystems as functions of the surface charge density andthe salt concentration.

The paper is organized as follows. In the next sectionwe briefly review the necklace model of polyelectrolytesin a poor solvent and estimate the free energy due toconfinement of a necklace molecule into a slit betweenparallel plates. Sections 3-5 discuss the adsorption ofhydrophobic polyelectrolytes at hydrophilic surfaceswhen polymer adsorption is caused exclusively byelectrostatic interactions between adsorbing surface andoppositely charged polyelectrolytes. In sections 3 and 4we apply the concept of the strongly correlated Wignerliquid21-26 to describe the structure of dilute and semi-dilute adsorbed layers formed by necklace-like chains.In section 5 we use a mean-field approach21-23,27-29 tocalculate polymer density profile and surface coveragefor concentrated adsorbed layers of overlapping beads.In section 6 we discuss the effects of van der Waalsattractions between adsorbing surface and polymerbackbones on its adsorption and calculate the adsorptiondiagram for the case of adsorption of hydrophobicpolyelectrolytes at hydrophobic oppositely charged sur-faces.

† Permanent address: Institute of Materials Science and De-partment of Physics, University of Connecticut, Storrs, CT 06269-3136.

2754 Macromolecules 2002, 35, 2754-2768

10.1021/ma0114943 CCC: $22.00 © 2002 American Chemical SocietyPublished on Web 02/22/2002

2. Necklace Model of a Polyelectrolyte Chain ina Poor Solvent7,9

A polyelectrolyte chain with degree of polymerizationN, fraction of charged monomers f, and bond length ain a poor solvent forms a necklace-like globule of beadsconnected by narrow strings7,9 (see Figure 1). Themonomer density F inside these globules is determinedby the balance between the two-body attraction -τa3FN,where the reduced temperature is given by

and Θ is the theta temperature for a polymer backbonein a given solvent, and the three-body repulsion a6F2N.The resulting density inside the globules is

The shape of a globule is determine by the competitionbetween electrostatic repulsion between charged mono-mers and the surface energy of the globule. The size ofthe beads

is determined by the Rayleigh’s stability condition30

where

is the number of monomers in a bead (mb ≈ FDb3), lB )

e2/(kTε) is the Bjerrum length, and u is the ratio of theBjerrum length lB (the length scale at which electro-static interaction between two elementary charges e inthe media with dielectric constant ε is of the order ofthe thermal energy kT) to the bond size a. The diameterof the strings is of the order of the thermal blob size

The length lstr of a string connecting two neighboringbeads can be estimated by balancing the electrostaticrepulsion between two closest beads kTlBf 2mb

2/lstr andthe surface energy of the string kTlstr/êT. The equilib-rium distance between beads is

The mass of a string between neighboring beads mstr ≈FlstrêT

2 is much smaller than the mass of a bead mb

In this case the number of beads Nbead on the chain isapproximately equal to N/mb ≈ uf 2N/τ. Since most ofthe necklace length is due to the strings (lstr > > Db),the length of the necklace can be estimated as thenumber of beads Nbead per chain times the length of astring lstr between neighboring beads

Within this approximation the total free energy of thepolyelectrolyte chain in a poor solvent is

where the first term is the electrostatic and surfaceenergies of the beads, the second one is the electrostaticrepulsion between beads which is of the order of thesurface energy of the strings, and the last one is thefree energy of a polymer backbone in a poor solvent. Theelectrostatic repulsion between beads (the second termon the right-hand side of eq 10) is smaller than theelectrostatic energy of the beads (the first term) and canbe neglected as long as the length of the string lstrbetween two neighboring beads is larger than the beadsize Db.

The transverse size of the polyelectrolyte chain in apoor solvent can be estimated by considering theelectrostatic energy change of any three consecutivebeads due to conformational bent by an angle θ from astraight alignment (see Figure 2). The variation of theelectrostatic repulsion between beads due to suchdeformation is

Thus, the mean-square value of the bending angle is⟨θ2⟩ ≈ êT/lstr. The mean-square transverse size ⟨R⊥

2⟩ ofthe polyelectrolyte chain with a sequence of bendingangles θi is

The transverse size of a polyelectrolyte chain in a poorsolvent is smaller than that in a Θ-solvent axN be-cause τ3 > uf 2.

Figure 1. Polyelectrolyte chain in a dilute salt-free solution.See text for definition of length scales. Figure 2. Necklace in a slit.

τ ) (Θ - T)/Θ (1)

F ≈ τ/a3 (2)

Db ≈ a(uf 2)-1/3 (3)

lB f 2mb2/Db ≈ Db

2êT-2 (4)

mb ≈ τuf 2

(5)

êT ≈ a/τ (6)

lstr ≈ a( τuf 2)1/2 ≈ amb

1/2 (7)

mstr

mb≈ (uf 2

τ3 )1/2

, 1 (8)

Lnec ≈ Nbeadlstr ≈ a(uf 2

τ )1/2

N ≈ a Nmb

1/2(9)

Wnec

kT≈ N

lB f 2mb

Db+

lB f 2N2

Lnec- Nτ2 (10)

Wnec

kT≈ N(uf 2)1/3τ - Nτ2 (11)

δWel

kT≈ lB(fmb)

2

lstr ( 1

x4 cos2 θ + sin2 θ- 1

2) ≈

lB(fmb)2

lstrθ2 ≈ lstr

êTθ2 (12)

⟨R⊥2⟩ ≈ lstr

2 ∑i)1

Nbead-2

⟨θi2⟩ ≈ N

mb

lstrêT ≈ a2N(uf 2

τ3 )1/2

(13)

Macromolecules, Vol. 35, No. 7, 2002 Adsorption of Hydrophobic Polyelectrolytes 2755

The confinement energy of a polyelectrolyte chain in

a poor solvent with transverse size x⟨R⊥2⟩ in a slit of

size D (see Figure 2) can be estimated as

3. Dilute Two-Dimensional Wigner LiquidConsider adsorption of polyelectrolyte chains at an

oppositely charged surface with number density ofsurface charge σ from a dilute polymer solution withpolymer concentration Fbulk and with salt concentrationFsalt. The presence of the univalent salt at concentrationFsalt leads to an exponential screening of the electrostaticinteractions at the length scales of the order of theDebye screening length

In Figure 3 we sketch the adsorption diagram ofpolyelectrolyte chains at an oppositely charged surfacewith number density of surface charges σ-Debye lengthrD plane.

The structure of the adsorbed polyelectrolyte layer atlow surface charge densities can be described using theconcept of the strongly correlated Wigner liquid.24-26 Inthe framework of this approach the polyelectrolytechains are considered to be localized at the centers ofthe Wigner-Seitz cells of size R. The total electrostaticenergy of an adsorbed polyelectrolyte chain includes theelectrostatic attraction of the chain to the oppositelycharged surface

and repulsion from other adsorbed polyelectrolytesdistributed with effective surface charge density fN/R2

starting at distance R from a given polyion

The total electrostatic energy of the adsorbed layer with

the surface area S is the sum of the contributions fromall chains

(The factor 1/2 in front of the repulsive term Wrep isadded to avoid the double counting of the electrostaticrepulsion between adsorbed chains.) The dependence ofthe Wigner-Seitz cell size R on the salt concentrationis derived by minimizing the electrostatic energy U withrespect to the cell size R. This minimization leads tothe following equation for the cell size R

Low Salt Regime (Regime IL in Figure 3). At low saltconcentrations (rD > R) we can expand the terms on theleft-hand side of eq 19 in the power series of the ratioof the Wigner-Seitz cell size R to the Debye radius rD.Within the zeroth approximation the size of the Wigner-Seitz cell

is independent of the Debye screening length. To obtainthe surface overcharging δσ, one has to keep the first-order term in the powers of R/rD on the left-hand sideof eq 19. Within this approximation the overchargingof the surface by adsorbed polyelectrolyte chains is19,22

It follows from this equation that the effective surfacecharge density δσ at the crossover between dilute low(IL) and dilute high (IH) salt regimes (at rD ≈ R ≈xfN/σ (see the crossover line in Figure 3) is of theorder of the bare surface charge density σ, but has anopposite sign.

From the equilibrium cell size R0 one can estimatethe adsorption energy of the polyelectrolyte chain withthe center of mass located at the distance z (z , R0)from the surface as the energy of the electrostaticattraction between a charged disk of radius R0 and acharge of valency fN

The polyelectrolyte chains are strongly attracted to thesurface with the binding energy

as long as the surface charge density σ is larger thanthe adsorption threshold value σWC

Figure 3. Adsorption diagram of polyelectrolyte chains in asalt solution as a function of surface charge density σ andDebye radius rD. The thin solid line separates low and highsalt regimes.

Fconf

kT≈ ⟨R⊥

2⟩

D2≈ (uf 2

τ3 )1/2a2ND2

(14)

rD ) (8πlBFsalt)-1/2 (15)

Watt

kT≈ -lB fNσ∫0

∞dr exp(- r

rD) ≈ -lB fNσrD (16)

Wrep

kT≈ lB(fN)2

R2 ∫R

∞dr exp(- r

rD) ≈

lB(fN)2rD

R2exp(- R

rD) (17)

UkT

≈ SR2(12 Wrep

kT+

Watt

kT ) ≈

SlBrD fN(12 fNR4

exp(- RrD

) - σR2) (18)

fN

R2exp(- R

rD)(1 + 1

4RrD

) ≈ σ (19)

R ≈ R0 ≈ xfN/σ, low salt rD > xfN/σ (20)

δσ ≈fN

R2- σ ≈ σ

R0

rD, low salt rD > xfN/σ (21)

Wads(z) ≈ -kTlB(fN)3/2σ1/2(1 - zR0

),low salt rD > xfN/σ (22)

|Wads(0)| ≈ kTlB(fN)3/2σ1/2 ≈ (σ/σWC)1/2 . kT (23)

σWC ≈ 1lB

2(fN)3(24)

2756 Dobrynin and Rubinstein Macromolecules, Vol. 35, No. 7, 2002

The chains are localized within a layer of thickness

inside which (for z < D) the change of the adsorptionenergy Wads(0) - Wads(z) (eq 22) is of the order of thethermal energy kT.

The strong electrostatic attraction of polyelectrolytechains to the charged surface starts to deform themwhen the layer thickness D (eq 25) becomes comparable

with the transverse size of the chains x⟨R⊥2⟩ (see eq

13). This occurs at the surface charge density

At higher surface charge densities σ > σdef the thicknessof the chains in the direction perpendicular to thesurface is determined by balancing the confinemententropy of a polyelectrolyte chain in a poor solvent (eq14) inside the layer of the thickness D with its electro-static attraction Wads(D) (see eq 22) to the chargedsurface. This gives the equilibrium thickness of thechain

that is independent of the degree of polymerization N.High Salt Regime (Regime IH in Figure 3). If the

Debye radius rD becomes smaller than xfN/σ, theadsorbed polyelectrolyte chains interact only with thepart of the surface within the Debye screening lengthrD. The Wigner-Seitz cell size R that minimizes theelectrostatic energy eq 18 is

Thus, up to logarithmic prefactor the cell size R isproportional to the Debye radius rD.19,21 The effectivesurface charge density in this high salt concentrationregime is

and can be much larger than the bare surface chargedensity σ.

As the salt concentration increases, the Debye radiusdecreases and the absolute value of the attractive energyof the polyelectrolyte chain to the charged surfacelBfNσrD decreases as well. Polyelectrolyte chains desorbfrom the charged surface when this energy becomes ofthe order of the thermal energy kT. At this value of theadsorption energy the concentration of polyelectrolytechains at the surface is of the order of polymer concen-tration in the bulk. In dilute high salt (regime IH)desorption occurs if the Debye radius rD is of the orderof the thickness of the adsorbed layer eq 25. The line

defines the desorption threshold for nonoverlapping(dilute) adsorbed polyelectrolyte chains (left boundaryof regime IH in Figure 3).

4. Semidilute Two-Dimensional Wigner Liquid4.1. String-Controlled Regime (Regime II). The

adsorbed polyelectrolytes begin to overlap when the sizeof the Wigner-Seitz cell R is of the order of the chainsize Lnec (eq 9). At low salt concentrations (rD > xfN/σ)the chains begin to overlap at the surface charge density

For higher surface charge densities (σ > σ*) the ad-sorbed polyelectrolytes are arranged in a two-dimen-sional semidilute polyelectrolyte solution with the dis-tance between chains ê (see Figure 4). Polyelectrolytechains are strongly stretched by intrachain electrostaticrepulsion on length scales smaller than the cell size ê.The number of monomers inside the cell gê is

To calculate the electrostatic energy of the adsorbedlayer, we will once again divide the electrostatic con-tribution to the total electrostatic energy of the adsorbedchains into the repulsive and the attractive ones. Theelectrostatic energy of the adsorbed layer can be writtenby analogy with eq 18 by substituting gê for N and ê forR, which leads to

Minimizing the total electrostatic energy U of theadsorbed layer with respect to the Wigner-Seitz cellsize ê, we obtain an equation that defines the cell sizeas a function of the Debye radius rD and the surfacecharge density σ

At low salt concentrations (regime IIL) the size of thecell

is inversely proportional to the surface charge densityσ, and the surface overcharging δσ in this regime is

Figure 4. Schematic sketch of the adsorbed layer in semidi-lute string-controlled 2-D regime.

σ* ≈ τua2fN

, low salt rD > xfN/σ (31)

gê ≈ NêLnec

≈ êτ1/2a-1u-1/2f -1 (32)

UkT

≈ SlBrD fgê( fgê

2ê4exp(- ê

rD) - σ

ê2) (33)

τ1/2

u1/2êa(1 + ê2rD

) exp(- êrD

) ≈ σ (34)

ê ≈ τ1/2

u1/2aσ, low salt rD > τ1/2/(u1/2aσ) (35)

D ≈ R0

lB(fN)3/2σ1/2≈ (lB fNσ)-1, low salt rD > xfN/σ

(25)

σdef ≈ τ3/4

a2u5/4f 3/2N3/2(26)

D ≈ a1/3u-1/6τ-1/2σ-1/3 (27)

R ≈ rD ln( fN

σrD2), high salt rD < xfN/σ (28)

δσ ≈fN

R2≈

fN

(rD ln( fN

σrD2))2

≈fN

rD2,

high salt rD < xfN/σ (29)

rDdes ≈ λfN ≈ (lB fNσ)-1, high salt rD < xfN/σ (30)

Macromolecules, Vol. 35, No. 7, 2002 Adsorption of Hydrophobic Polyelectrolytes 2757

Surface overcharging δσ is inversely proportional to theDebye radius rD and is independent of the surfacecharge density σ.

For high salt concentrations (regime IIH) the solutionof eq 34 is

The surface overcharging by the adsorbed polyelectro-lyte chains at high salt concentrations

has the same functional form with logarithmic accuracyas eq 36 for low salt concentrations. However, it isimportant to point out that at low salt concentrationsthe surface overcharging |δσ| is much smaller than thebare surface charge density σ, while at higher saltconcentrations |δσ| > σ.

The thickness of the adsorbed layer D in regime II isstill given by eq 27 and decreases with increasingsurface charge density as σ-1/3. This decrease continueswhile the thickness of the polyelectrolyte layer is largerthan the bead size Db (eq 3) or as long as the surfacecharge density σ is smaller than

For higher surface charge densities σ > σsat the thick-ness of the adsorbed layer D saturates at Db, and thebeads keep their spherical shape because the electro-static attraction to the charged surface is weaker thanthe electrostatic self-energy of the beads.

Flexible hydrophobic polyelectrolyte chains with thepersistence length proportional to the Debye screeninglength9 rD desorb when the electrostatic attractionbetween the charged surface and a section of size rD isof the order of thermal energy31 kT

The desorption transition occurs at the Debye screeninglength32 (lower left boundary of the regime IIH in Figure3)

The crossover to a rigid chain behavior takes place whenthe Debye screening length rD is of the order of the chainsize Lnec. For larger values of the Debye radius the chaincan be viewed as a charged rod of length Lnec. Thedesorption of such rod occurs along the line in theadsorption diagram (left boundary of the regime IH inFigure 3) given by eq 30.

This string-controlled regime (regime II) continuesuntil the Wigner-Seitz cell size ê is larger than the

distance between neighboring beads lstr (eq 7). The cellsize ê is comparable with the string length lstr at lowsalt concentrations for the surface charge density (bound-ary between regimes IIL and IIIL in Figure 3)

and at higher salt concentrations for the value of theDebye radius (boundary between regimes IIH and IIIHin Figure 3)

4.2. Bead-Controlled Regime (Regime III). Forsmaller cell sizes (ê < lstr) there will be one bead pereach correlation cell, and the adsorbed layer can beeffectively viewed as 2-D dilute Wigner liquid of chargedbeads (see Figure 5). The electrostatic energy per beadin the dilute Wigner liquid of beads is the sum of theelectrostatic attraction of the bead with the center ofmass located at distance Db from the surface to thecharged background

and its electrostatic repulsion from other beads

Minimization of the total electrostatic energy S(Watt +Wrep/2)/êb

2 of the adsorbed layer with respect to the cellsize êb results in the following equation

At low salt concentrations (regime IIIL) the solution ofthis equation for the equilibrium cell size êb is

For this range of salt concentrations the overchargingof the surface by polyelectrolyte chains is

At high salt concentrations (regime IIIH) the solutionof eq 46 can be approximated by

Figure 5. Schematic sketch of the adsorbed layer in semidi-lute bead-controlled 2-D regime.

σbead ≈f

a2, low salt rD > τ1/2/(u1/2aσ) (42)

rDbead ≈ lstr ≈ a( τ

uf 2)1/2, high salt rD < τ1/2/(u1/2aσ)

(43)

Watt ) - kTlBrDσfmb (44)

Wrep ) kTlB(fmb)2rDêb

-2 exp(- êb/rD) (45)

fmb

êb2

exp(-êb

rD)(1 + 1

4êb

rD) ≈ σ (46)

êb ≈ xfmb

σ≈ x τ

ufσ, low salt rD > xτ/(ufσ) (47)

δσ ≈ σêb

rD, low salt rD > xτ/(ufσ) (48)

δσ ≈ σ( êrD

) ≈ τ1/2

u1/2rDa, low salt rD > τ1/2/(u1/2aσ) (36)

ê ≈ rD ln( τ1/2

σrDau1/2), high salt rD < τ1/2/(u1/2aσ) (37)

δσ ≈ fgê

ê2- σ ≈ τ1/2

u1/2arD ln( τ1/2

σrDau1/2)≈ τ1/2

u1/2rDa,

high salt rD < τ1/2/(u1/2aσ) (38)

σsat ≈u1/2f 2

a2τ3/2(39)

|Wads| ≈ kTlBσfgê rD ≈ kT (40)

rDdes ≈ 1

u1/4τ1/4σ1/2, for rD > lstr (41)

2758 Dobrynin and Rubinstein Macromolecules, Vol. 35, No. 7, 2002

Thus, up to logarithmic prefactor, the cell size êb isproportional to the Debye radius rD.19,21

The beads begin to overlap when the Wigner-Seitzcell size êb is of the order of the beads size Db. This takesplace at low salt concentrations for surface chargedensity σ equal to (boundary between IIIL and IVL)

and at high salt concentrations for the value of theDebye radius rD

Polyelectrolyte chains desorb when the electrostaticattraction of a bead to the charged surface

is of the order of the thermal energy kT. The adsorption/desorption threshold takes place at the Debye radius(left boundary of regime IIIH in Figure 3)

This adsorption/desorption line crosses the line rD ≈ Dbat the surface charge density

At this surface charge density the attraction of a beadto the surface with area Db

2 and charge σwDb2 is of the

order of kT.At high salt concentrations with Debye radius rD

smaller than the bead size Db the electrostatic interac-tions between monomers are negligible in comparisonwith nonelectrostatic ones (see for details ref 35). Eachpolyelectrolyte chain forms a spherical globule with sizeRglob ≈ aN1/3τ-1/3. The solution of globules with concen-tration c higher than cdil

is thermodynamically unstable with respect to phaseseparation into polymer concentrated and dilute phases.36

In the present paper we consider only the case of saltconcentrations for which rD is larger than the bead sizeDb.

5. Three-Dimensional Adsorbed LayerAt higher surface charge densities (σ > σ*bead) the

polyelectrolyte chains form a concentrated polymer layernear the charged surface. In this regime the distributionof the polymer density can be described within theframework of the mean-field approximation assumingthat the electrostatic interactions of a polymer witheffective external potential æ(z) created by other chainsand the surface dominates over the electrostatic self-energy of the chain. In this approximation the polymer

density F(z) and the small ion density FR(z) depend onlyon the distance z from the charged surface. The densityprofile of the salt ions in the electrostatic potential æ(z)satisfies the Boltzmann distribution

The relation between the electrostatic potential æ(z) andlocal polymer density F(z) can be derived by assumingthat the adsorbed layer is built of blobs of size ê(z). Thenumber of monomers g(z) in a blob is determined fromthe fact that these blobs are space-filling in the adsorbedlayer g(z) ≈ F(z)ê3(z) and the statistics of the chain insidea blob is Gaussian ê2(z) ≈ a2g(z). These blobs are multi-valent sections of polyions with the valency fg(z). Eachblob interacts with the external electrostatic potentialæ(z) with the energy of the order of thermal energy kT.

This leads to the following relation between the mono-mer concentration F(z) and the electrostatic potentialæ(z) at the distance z from the surface

However, polyelectrolytes are already compressed bymonomer-monomer attractive interactions in a poorsolvent. The external electrostatic potential æ(z) willfurther compress the polyelectrolyte chains only whenthe interaction with external potential fæ(z) is strongerthan monomer-monomer attraction τ2.

The electrostatic potential æ(z) satisfies the linearizedPoisson-Boltzmann equation

together with the boundary condition at the chargedsurface

In eq 59 we have identified two different cases depend-ing on the surface charge density σ: (i) layer withuniform polymer density for fæ(z) < τ2; (ii) nonuniformlayer density for fæ(z) > τ2.

5.1. Layer with Uniform Density (Regime IV). Inthe case when fæ(0) , τ2 the polymer density F(z) inthe adsorbed layer is almost constant and is equal tothat inside beads τ/a3. The layer is built by the thermalblobs of almost constant size êT ≈ aτ-1 because electro-static attraction of charged monomers to a charged sur-face is weaker than two-body attraction between sec-tions of chains in a poor solvent. The solution of thelinearized Poisson-Boltzmann equation (59) in thisregime is

The thickness D of the polymeric layer can be foundfrom the boundary condition for the electrostatic poten-tial

êb ≈ rD ln( τufrD

2σ) , high salt rD < xτ/(ufσ) (49)

σ*bead ≈ τa2 ( f

u)1/3

, low salt rD > xτ/(ufσ) (50)

r*bead ≈ Db ≈ a(uf 2)-1/3, high salt rD < xτ/(ufσ)(51)

|Wads| ≈ kTlBrDσfmb (52)

rDdes ≈

f

σaτ, for σsat < σ < σ*bead (53)

σw ≈u1/3f 5/3

a2τ(54)

cdil ≈ τa-3 exp(- τ4/3N2/3) (55)

F((z) ) Fsalt exp(-æ(z)) (56)

fg(z) æ(z) ≈ 1 (57)

F(z) ≈ g(z)/ê3(z) ≈ a-3g(z)1/2 ≈ a-3(fæ(z))1/2 (58)

d2æ(z)

dz2) 4πlB(F-(z) - F+(z) + fF(z))

≈ æ(z)

rD2

+4πlB f

a3 {τ, for fæ(z) < τ2

xfæ(z), for fæ(z) > τ2(59)

dæ(z)dz |z)0

) -4πlBσ (60)

æ(z) ≈ ufτrD2

a2sinh2(D - z

2rD) (61)

Macromolecules, Vol. 35, No. 7, 2002 Adsorption of Hydrophobic Polyelectrolytes 2759

In the limit of low salt concentrations D/rD , 1 (regimeIVL) this equation has a simple solution

Thus, the layer thickness in this regime scales linearlywith the surface charge density σ. This can be under-stood by realizing that at low salt concentrations thecharge of the adsorbed layer compensates the charge ofthe surface. The monomer concentration in the adsorbedlayer is constant, and therefore the effective surfacecharge density of the adsorbed layer is proportional toits thickness D.

The surface undercharging by adsorbed polyelectro-lytes is obtained by multiplying the polymeric chargedensity fτa-3 in the adsorbed layer by the layer thick-ness D

The addition of salt decreases the polymer adsorbedamount, because salt ions also take part in screeningof the surface charge.

At high salt concentrations (regime IVH) the thick-ness of the adsorbed layer (solution of eq 62)

is proportional to the Debye radius rD (up to logarithmiccorrections).

However, the mean-field description of the polyelec-trolyte adsorbed layer presented above is incorrect atdistances z from the surface larger than D - êT withinthe thermal blob êT of the outer edge of the adsorbedlayer. At these distances the fluctuations of the polymerdensity become larger than the average polymer density.This layer can be considered as a two-dimensional meltof thermal blobs of size êT. The additional decrease inpolymer surface coverage due to this strongly fluctuat-ing outer layer of thermal blobs can be estimated bycomparing the attraction of the blobs to the chargedsurface with electrostatic repulsion between them. Theeffective surface charge density that the last layer ofthickness êT feels is of the order of fa-2. The attractionenergy of a blob to the charged background with thissurface charge density is

The repulsion of the blob from all other blobs in thislayer with the effective surface charge density δσêT is

At equilibrium the total electrostatic energy per blobin the adsorbed layer Watt + Wrep is of the order of thethermal energy -kT. The surface undercharging due tothis last layer of thermal blobs is

The total undercharging of the adsorbing surface bypolyelectrolyte chains is

Thus, chains adsorbed in the 3-D layer undercom-pensate the surface charge, and polymer surface cover-age always decreases with increasing the salt concen-tration. Polyelectrolyte chains in a poor solvent demon-strate qualitatively different behavior from 3-D ad-sorbed polyelectrolyte layers in Θ and good solvents forthe polymer backbone.22,23 Polyelectrolytes in Θ andgood solvents show nonmonotonic dependences of thesurface overcharging on salt concentrations. At low ionicstrengths the surface overcharging increases with in-creasing salt concentration while at higher ionic strengthsit decreases with salt concentration.

As we show in Appendix the uniform adsorbed layeris stable well below σ*bead. However, it has higher freeenergy per unit area than the free energy of the Wignerliquid of beads. Thus, at the surface charge density σ ≈σ*bead there is a first-order phase transition between theWigner liquid of beads and uniform adsorbed layer.

5.2. Nonuniform Layer Density (Regime V). Theelectrostatic attraction of charged monomers to thecharged surface fæ(0) becomes of the order of two-bodymonomer-monomer attraction τ2 at the surface chargedensities σ of the order of

and at the value of the Debye radius obtained from eqs61 and 65 (including logarithm)

For higher surface charge densities (σ > σθ) or smallervalues of the Debye radius (rD < rD

θ) the electrostaticpotential at the charged surface æ(0) is larger than f -1τ2.There are two different regions in the adsorbed layer(see Figure 6a). The structure of the adsorbed layer closeto the charged surface is similar to the one for adsorp-tion from a Θ-solvent where polymer density profile isdetermined by the balance between electrostatic attrac-tion of charged monomers to the charged surface andthree-body monomer-monomer repulsion,21,23 while thepolymer density in the outer layer stays almost constantτ/a3 and is determined by the balance between two-bodyattraction (poor solvent) and three-body repulsion. Thesolution of the Poisson-Boltzmann equation in this caseis

where h is the thickness of the inner layer. The thick-

fτrD

a3sinh( D

2rD) cosh( D

2rD) ≈ σ (62)

D ≈ rD arcsinh( a3σfτrD

) ≈ a3σfτ

, low salt rD > a3σfτ

(63)

δσD ≈fτD

a3- σ ≈ fτrD

a3arcsinh( a3σ

fτrD) - σ ≈ - a6σ3

f 2τ2rD(64)

D ≈ rD ln(a3σfτrD

) ≈ rD, high salt rD < a3σfτ

(65)

Watt ≈ -kTlB f êT fgT

a2(66)

Wrep ≈ kTlB(δσêT+ f /a2) fgTrD (67)

δσêT≈ - 1

lB fgTrD≈ - τ2

lB frD(68)

δσ ≈ δσêT+ δσD ≈ - τ2

lB frD- a6σ3

f 2τ2rD2

(69)

σθ ≈ τ3/2

a2u1/2, low salt rD > a3σ

fτ(70)

rDθ ≈ τ2

ufσa, high salt rD < a3σ

fτ(71)

æ(z) ≈ {u2f 3rD4

a4sinh4 (D - z

4rD), for z < h

ufτrD2

a2sinh2 (D - z

2rD), for z > h

(72)

2760 Dobrynin and Rubinstein Macromolecules, Vol. 35, No. 7, 2002

ness of the adsorbed layer D ) Dout + h, where Dout isthe thickness of the outer layer, can be obtained fromthe boundary condition eq 60

For low salt concentrations (regime VL) the thicknessof the adsorbed layer

increases with the surface charge density as σ1/3. At highsalt concentrations (regime VH) the layer thickness Dis of the order of the Debye radius rD

At the crossover between the inner and outer layers(z ≈ h) the value of the electrostatic potential æ(h) is ofthe order of τ2f -1. Taking this into account, we can writethe expression for the thickness of the outer layer Doutas follows (see eq 61)

The solutions of this equation for low (Dout < rD, regimeVL) and high (Dout ≈ rD, regime VH) salt concentrationsare

The inner region of the adsorbed layer of thickness his built of the blobs with gradually increasing sizes

As the surface charge density increases, the size of thefirst blob ê(0) decreases. At the surface charge density

the size of this blob ê(0) is of the order of the distancebetween charged monomers a/xf and there is onecharged monomer per such blob. This is the maximumpossible screening of the surface by polyelectrolytes.Further increase of polymer density near the surface isunfavorable due to the high cost of the short-rangethree-body repulsive interactions.

5.3. Screening by Small Ions (Regime VI). Surfacecounterions start to dominate the screening of thesurface charge for the higher surface charge densitiesσ > σion.21,23 In this range of surface charge densitiesthere are three different regions in the adsorbed layer(see Figure 6b). Close to the charged surface the polymerdensity is of the order of a-3f 1/2. The polymer densityF(z) is proportional to a-3xfæ(z) at intermediate lengthscales. It is constant (F ≈ a-3τ) in the outer part of theadsorbed layer. The thickness of the adsorbed layer Dat low salt concentrations (regime VIL) saturates at

while at high salt concentrations it is of the order ofthe Debye radius rD up to logarithmic corrections.21,23

However, the intermediate regime (regime V) withsurface charge densities σθ < σ < σion exists only if thereis more than one charged monomer in the thermal blobêT or for the effective temperatures τ < xf. For highereffective temperatures τ > xf the system directlycrosses over into the counterion-controlled regime whenthe Gouy-Chapman length due to surface charge 1/lBσbecomes of the order of the layer thickness D given byeqs 63 and 65 for low and high salt regimes, respec-tively. The thickness of the adsorbed layer saturates at

in this “screening by small ions” regime at low saltconcentrations.

6. Adsorption Diagram in the Case of Strongvan der Waals Interactions

In sections 3-5 above we have neglected the effect ofthe van der Waals interaction between the charged

Figure 6. Polymer density profile and layer structure in the3-D adsorbed layer: (a, top) regime V; (b, bottom) regime VI.

Dout ≈ {ax τuf 2

, low salt rD > ax τuf 2

rD ln( τa2

uf 2rD2), high salt rD < ax τ

uf 2

(77)

ê(z) ≈ 1F(z)a2

≈ a

xfæ(z)≈ a3

uf 2rD2

sinh-2(D - z4rD

) (78)

σion ≈ { f 3/4

a2u1/2, low salt rD > au-1/2f -3/4

1urDa

, high salt rD < au-1/2f -3/4

(79)

D ≈ au-1/2f -3/4, low salt rD > au-1/2f -3/4 (80)

D ≈ axufτ

, low salt rD > a3σfτ

, τ > xf (81)

uf 3rD3

a3sinh3( D

4rD) cosh( D

4rD) ≈ σa2 (73)

D ≈ a5/3u-1/3f -1σ1/3,

for σθ < σ < σion, low salt rD > D (74)

D ≈ rD ln( σa5

uf 3rD3)

for σθ < σ < σion, high salt rD ≈ D (75)

sinh2 (Dout

2rD) ≈ τa2

uf 2rD2

(76)

Macromolecules, Vol. 35, No. 7, 2002 Adsorption of Hydrophobic Polyelectrolytes 2761

surface and the polymer backbone. The effective inter-action between the adsorbing surface and a monomerin a solvent is37

where Aeff is the effective Hamaker constant for polymersurface interaction.

In regime IV of the adsorption diagram (Figure 3) the3-D adsorbed layer is built of the blobs with constantsize êT ≈ a/τ. Let us estimate the van der Waalsattraction of the blob closest to the charged surface

This attractive interaction is controlled by the low limitof the integral, proving that only monomers in contactwith an adsorbing surface contribute to this type ofinteractions. Taking this into account, we can rewritethe last equation as the energy per contact kTεVW (thesticking energy of the monomer to the adsorbing sur-face) times the number of contacts ncont per blob. Sincethe statistics of the chain on length scales smaller thanthe blob size êT is Gaussian, the number of contacts ncontof a blob with the number of monomers gT ≈ êT

2/a2 in itwith adsorbing surface is xgT ≈ τ-1.

Thus, the sticking energy per contact kTεVW is propor-tional to Aeff.

The van der Waals attraction can be neglected if it isweaker than electrostatic attraction of the first layer ofthermal blobs

However, in regime IV the structure of the adsorbedlayer is controlled by the monomer-monomer interac-tions rather than electrostatic attraction to a chargedsurface, and it will not be affected by the van der Waalsattraction as long as it is weaker than kT per thermalblob (|W(êT)| < kT). This leads to the following inequalityfor the sticking energy

In regimes V and VI of the adsorption diagram(Figure 3) the energy of the closest to the surface blobsis always of the order of the thermal energy kT. In thiscase the sticking energy εVW should be

in order to safely neglect the effects of the van der Waalsattraction to the surface. The number of monomers inthese blobs g(0) is always smaller than that in thethermal blob gT, because these blobs are compressed byelectrostatic attraction to the charged surface. Thus, ifthe sticking energy εVW is smaller than τ, we can neglectthe effect of the van der Waals interactions in the entire3-D region of the adsorption diagram (Figure 3).

In the case of the 2-D adsorbed layers (σ < σ*bead) theadsorbed polyelectrolyte chains have a necklace-likestructure. The beads on the necklace will undergowetting/dewetting transition as the strength of the vander Waals interactions of the polymer backbone withadsorbing surface increases. This transition occurs if thesurface energy gain due to increase the number ofcontacts between polymer backbone and the adsorbingsurface is higher than the surface energy loss due toincrease in the area of the polymer-solvent interface.The energy difference per unit area of the spreadingbead is

where the first term describes the energy difference perunit area between the polymer-surface interface γ -kTεVWτ/a2 and the surface-solvent interface γ, and thelast one is the surface energy per unit area of polymer-solvent interface. The beads spread if the energy dif-ference δγ is negative. This occurs for the values of thesticking energy εVW larger than the effective tempera-ture τ.

Now let us discus how the adsorption diagram willchange if the sticking energy εVW is larger than theeffective temperature τ. Our results for this case aresummarized in Figure 7.

6.1. Three-Dimensional Adsorbed Layer. We be-gin our discussion of the adsorption diagram from 3-Dadsorbed layers. In this case the strong van der Waalsattraction of the polymer backbone to the adsorbingsurface controls the layer structure near the adsorbingsurface. In the vicinity of the charged surface polyelec-trolytes form de Gennes’ carpet38 with polymer densityprofile given by the following equation

where êvw is the size of the closest to the surface blob(adsorption blob). The size of this blob can be found fromthe condition that its van der Waals interaction withadsorbing surface is of the order of the thermal energy

W(z) ≈ -kT a3

z3Aeff (82)

W(êT) ≈ êT2∫a

êTF(z) W(z) dz ≈

-kTAeff

τ (1 - aêT

) ≈ -kTAeff

τ(83)

W(êT) ≈ -kTAeff

τ≈ -kTAeffxgT ≈ -kTεVWxgT (84)

|W(êT)| < kTfgT|æ(0)| (85)

εVW < gT-1/2 ≈ τ (86)

εVW < g-1/2(0) (87)

Figure 7. Adsorption diagram of polyelectrolyte chains in asalt solution as a function of surface charge density σ andDebye radius rD in the case of strong van der Waals interac-tions.

δγ = -kTεVWxgT

êT2

+ kT

êT2

=kT

a2(-εVWτ + τ2) (88)

F(z) =1

a2(êvw + z)(89)

2762 Dobrynin and Rubinstein Macromolecules, Vol. 35, No. 7, 2002

kT or εVWg(0)1/2 ≈ 1. This leads to the following expres-sion for the blob size

In regime IV of the adsorption diagram (see Figure 7)this de Gennes’ carpet ends at the distance z from acharged surface where the polymer density F(z) becomescomparable with polymer density τ/a3 inside thermalblob êT (see Figure 8). This takes place at the distancez of the order of the size êT of the thermal blob. Thepolymer excess within the layer of thickness êT resultsin the screening of the surface charge by the amountfa-2 ln(εVW/τ) leading to the effective surface chargedensity

that the rest of the adsorbed polyelectrolytes “feel”.However, in the 3-D adsorbed regimes the surfacecharge density σ is always higher than f/a2. Thus, thepresence of van der Waals interactions does not signifi-cantly affect the properties of the adsorbed layer.

A similar analysis of the effect of the van der Waalsinteractions can be carried out in regimes V and VI ofthe adsorption diagram (see Figure 7). In both of theseregimes the van der Waals attractions will perturb thestructure of the adsorbed layer at length scales z smallerthan the size of the first blob ê(0) (see eq 78 for the blobsize ê(0) in regime V and ê(0) ≈ af -1/2 in regime VI). Atthese length scales the polymer density profile will begiven by eq 89 while at the distances z > ê(0) thestructure of the adsorbed layer will remain the sameas without van der Waals interactions (see section 5).

However, in the case of strong van der Waals interac-tions (εVW > τ) adsorbed polyelectrolytes form 3-Dadsorbed layer even at the surface charge densities σbelow σ*bead (regime III in Figure 7). The structure ofthe adsorbed layer at these surface charge densities issimilar to the one in regime IV discussed above. At thelength scales z < êT the adsorbed chains form deGennes’s carpet with polymer density given by eq 89.At length scales larger than the thermal blob size êTthe polymer density is constant and is equal to τa-3.The thickness of this layer of constant density can beobtained by substituting the effective surface chargedensity σeff (eq 91) into eq 63

The outer layer of thickness D disappears completely

at the surface charge density σ ≈ fa-2 ln(εVW/τ). At thisvalue of the surface charge density the thickness of theadsorbed layer is of the order of the thermal blob sizeêT. As the surface charge density σ decreases further (σ< fa-2 ln(εVW/τ)) the thickness of the adsorbed layer Dreaches the size of the adsorption blob êvw exponentiallyfast39 and saturates there at lower surface chargedensities. In fact, the system crosses over into 2-Dsemidilute adsorbed layer at surface charge density σ≈ fa-2 at low salt concentrations. However, for highersalt concentrations rD < aεVW

2/f 2u, at which the elec-trostatic interactions per monomer in the adsorbed layerf 2urD/a are weaker than average monomer-surfaceinteractions εVW

2, the adsorbed polyelectrolytes form amonolayer of overlapping chains with thickness êvw,maximizing the number of polymer surface contacts,even for surface charge densities σ lower than fa-2.

6.2. Semidilute Two-Dimensional Wigner Liquid(Regime II). For lower surface charge densities (σ <fa-2) adsorbed polyelectrolytes arrange themselves intoa two-dimensional semidilute polyelectrolyte solutionwith the distance between chains ê and chain thicknessof the order of the adsorption blob size êvw controlledby the van der Waals attraction of the polymer backboneto the adsorbing surface (see Figure 9). The intrachainelectrostatic repulsion leads to chain stretching alongthe adsorbing surface on length scales larger thanelectrostatic blob3,4,6 size De (see Figure 9). The confor-mation of a chain inside the electrostatic blob is almostunperturbed by electrostatic interactions and by two-body monomer-monomer attraction with the numberof monomers in it being ge ≈ (De/a)2. The size of theelectrostatic blob De containing ge monomers can befound by comparing the electrostatic energy of a blobe2ge

2f 2/εDe with the thermal energy kT. This leads tothe electrostatic blob size

and the number of monomers in it

At the length scales smaller than ê the polyelectrolyteconfiguration is that of a fully extended array of gê/geelectrostatic blobs of contour length ê

To calculate the total free energy of the adsorbedlayer, we will once again divide the electrostatic con-tribution to the total free energy into the repulsive andattractive ones and add to it the contribution from theshort-range van der Waals interactions. With these

Figure 8. Polymer density profile and layer structure in the3-D adsorbed layer in the case of strong van der Waalsinteractions (regime IV).

Figure 9. Schematic sketch of the adsorbed layer in semidi-lute 2-D regime in the case of strong van der Waals interac-tions.

De ≈ a(uf 2)-1/3 (93)

ge ≈ (uf 2)-2/3 (94)

ê ≈ gê

geDe ≈ agê(uf 2)1/3 (95)

êvw ≈ a/εVW (90)

σeff ≈ σ -f

a2ln(εVW

τ ) (91)

D ≈ rD arcsinh( a3

fτrD (σ -f

a2ln(εVW

τ ))) (92)

Macromolecules, Vol. 35, No. 7, 2002 Adsorption of Hydrophobic Polyelectrolytes 2763

modifications the free energy of the adsorbed layer is

Minimizing this free energy F with respect to theWigner-Seitz cell size ê and keeping in mind that gê ≈ê/(u1/3f 2/3a), we obtain the equation that defines the cellsize as a function of the Debye radius rD

It follows from the last equation that the cell size canbe obtained by balancing the electrostatic repulsionbetween chains either with their attraction to theoppositely charged surface (the first term on the right-hand side of eq 97) or with their nonelectrostaticattraction to the adsorbing surface (the second term onthe right-hand side of eq 97). The crossover between twostabilizing mechanisms takes place across the line

At low salt concentrations (rD > rDVW) the size of the cell

is inversely proportional to the surface charge densityσ. The surface overcharging δσ in this regime

is inversely proportional to the Debye radius rD and isindependent of the surface charge density σ.

At higher salt concentrations (rD < rDVW) the solution

of eq 97 is

In this regime the distance between chains ê scaleslinearly with the Debye screening length rD. Thus, thedistance between chains decreases as salt is added tothe solution. The 2-D electrostatic blobs start to overlapat the salt concentrations for which the Wigner-Seitzcell size ê is of the order of the electrostatic blob sizeDe. This occurs for the following value of the Debyeradius rD

At this salt concentration (rD ≈ rD3d) the polymer sur-

face coverage Γ

reaches its maximum possible value and stays un-

changed for higher salt concentrations (in regime III ofFigure 7 for the interval of the surface charge densitiesσ < fa-2).

The surface overcharging by adsorbed polyelectrolytesin the interval of salt concentrations (rD

3d < rD < rDVW) is

and it increases with increasing the salt concentration.As the surface charge density σ decreases in the

regime of low salt concentrations (rD > rDVW) or as the

salt concentration decreases in the regime of high saltconcentrations (rD < rD

VW), the size of the Wigner-Seitzcell ê (see eqs 98 and 101) increases and reaches thesize L ≈ aN(uf 2)1/3 of the polyelectrolyte chain at thesurface charge density

and at the value of the Debye radius

These values of the surface charge density σ*VW and ofthe Debye radius r*D define the boundary between 2-Dsemidilute (II) and dilute (I) regimes.

6.3. Dilute Two-Dimensional Wigner Liquid (Re-gime I). The total free energy of the adsorbed layer ina dilute regime is the sum of the contributions from allchains due to electrostatic and short-range van derWaals interactions

The dependence of the Wigner-Seitz cell size R on thesalt concentration is derived by minimizing the total freeenergy with respect to R. The equilibrium cell size isthe solution of the following equation

At low salt concentrations rD > rDVW the size R of the

Wigner-Seitz cell has very weak dependence on theDebye screening length rD. Thus, on the scaling level,we can assume that at low salt concentrations

At low salt concentrations we can expand the left-handside of eq 108 in the power series of R/rD. After thisexpansion eq 108 reduces to

Within this approximation the overcharging of thecharged surface by adsorbed polyelectrolyte chains is

FkT

≈ SlBrD fgê( fgê

2ê4exp (-

ê

rD) - σ

ê2) - SεVW2

ê2(96)

f 1/3

u1/3êa (1 +ê

2rD) exp(-

ê

rD) ≈ σ +

εVW2

lBrD f(97)

rDVW ≈ εVW

2

lBσf(98)

ê ≈f 1/3

u1/3aσ, rD > rD

VW (99)

δσ ≈ εVW2

lBrD f, rD > rD

VW (100)

ê ≈ rD

u2/3f 4/3

εVW2

, rD < rDVW (101)

rD3d ≈ a

εVW2

uf 2, rD < rD

VW (102)

Γ(rD3d) ≈ gê

ê2≈ εVW

2

lBrD3df 2

≈ 1a2

, rD e rD3d (103)

δσ ≈ fgê

ê2- σ ≈ εVW

2

lBrD f(104)

σ*VW ≈ 1a2u2/3f 1/3N

, rD > rDVW (105)

r*D ≈ aNεVW

2

u1/3f 2/3, rD < rD

VW (106)

FkT

≈ SlB rD fN (12 fN

R4exp(- R

rD) - σ

R2) - SεVW2 N

R2

(107)

fN

R2exp(- R

rD)(1 + 1

4RrD

) ≈ σ +εVW

2

lBrD f(108)

R ≈ xfN/σ, rD > rDVW (109)

fN

R2 (1 - 34

RrD

) ≈ σ +εVW

2

lBrD f(110)

2764 Dobrynin and Rubinstein Macromolecules, Vol. 35, No. 7, 2002

When the Debye radius rD becomes smaller than rDVW

the electrostatic repulsion between adsorbed polyelec-trolyte chains is balanced by their van der Waalsattraction to adsorbing surface. The Wigner-Seitz cellsize R that minimizes the total free energy eq 107 is

Thus, the cell size R is proportional to the square rootof the Debye radius rD. The effective surface chargedensity in this high salt concentrations regime is

It has the same functional form as in the semidiluteregime (see eq 104) and can be much larger than thebare surface charge density σ.

7. ConclusionWe have presented a scaling analysis of the adsorp-

tion of hydrophobic polyelectrolytes at the oppositelycharged surfaces from dilute polymer solutions. In thecase of hydrophilic surfaces, when the sticking energyεVW per monomer due to van der Waals interactions withcharged surface is smaller than the effective tempera-ture τ (εVW < τ), the structure of the adsorbed layer inthe dilute (regime I in Figure 3) and semidilute string-controlled (regime II in Figure 3) regimes is qualitativelysimilar to that for adsorption of polyelectrolytes from aΘ-solvent for polymer backbone.22 In these regimes theadsorbed necklaces form a two-dimensional stronglycorrelated Wigner liquid. At low salt concentrations thesize of the corresponding Wigner-Seitz cells is deter-mined by the electroneutrality condition of these cellswhile at higher salt concentrations it is proportional tothe Debye radius rD (up to logarithmic corrections). Thechains are localized inside the layer of thickness D ≈(lBfNσ)-1 (see Figure 10) for surface charge densitiessmaller than σdef and the value of the Debye radius rD

larger than the transverse size of the necklace x⟨R⊥2⟩.

For higher surface charge densities σ > σdef and smaller

values of the Debye radius rD < x⟨R⊥2⟩ the thickness of

an adsorbed chain D is determined by the balance ofthe energy gain due to electrostatic attraction and theconfinement entropy loss due to chain localization. Thethickness of the adsorbed chain D decreases withincreasing surface charge density as σ-1/3 (see Figure10) and is independent of the degree of polymerizationN. The thickness of the adsorbed layer D ≈ Db saturatesat the surface charge density σ ≈ σsat.

In the interval of the surface charge densities wherethe size of the Wigner-Seitz cell is smaller than thebare length of the string lstr between two neighboringbeads but larger than the bead size Db (regime III inFigure 3) there is on average one bead per eachcorrelation cell. In this regime the adsorbed layerresembles 2-D Wigner liquid of charged colloidal par-ticles for which the Wigner-Seitz cell size is inverselyproportional to square root of the surface charge densityσ-1/2 at low salt concentrations.

The crossover to the 3-D adsorbed layer occurs at thesurface charge density σ of the order of σ*bead at whichbeads begin to overlap. For higher surface chargedensities in the interval σ*bead < σ < σθ (regime IV inFigure 3) the thickness of the adsorbed layer growslinearly with the surface charge density σ (see Figure10). The monomer density in the adsorbed layer isconstant and equal to τa-3.

If the surface charge density increases further (σθ <σ < σion) (regime V in Figure 3), the equilibrium densityprofile in the adsorbed layer is determined by thebalance between electrostatic attraction of chargedmonomers to the surface and short-range monomer-monomer repulsion in the layers close to the surface andby the balance of the two-body attraction and three-bodyrepulsion further away from it. On length scales z < hthe polyelectrolytes form a self-similar carpet withpolymer density decaying as (D - z)2 (see Figure 6a).At the intermediate length scales (h < z < D) themonomer density in the adsorbed layer is constant andis equal to that inside a polymeric globule a-3τ.

For very high surface charge densities (σ > σion)(regime VI in Figure 3) the surface counterions domi-nate the screening of the surface potential near the wall.There are three different regions in the adsorbed layer(see Figure 6b). Close to the charged surface counterionsdominate the surface screening, and the monomerdensity is almost constant F ≈ a-3f 1/2 with one elemen-tary charge per correlation blob. Further away from thecharged surface the density profile in the two remaininglayers is similar to the one described above.

The thickness of the 3-D adsorbed layer in the lowsalt concentration regime increases linearly with thesurface charge density σ in the regime σ*bead < σ < σθand grows as σ1/3 for σθ < σ < σion. At very high surfacecharge densities (σ > σion) layer thickness saturates atau-1/2f -3/4. In the high salt concentration regimes thethickness of the adsorbed layer is proportional to theDebye screening length rD.

For comparison, Figure 11 shows the dependence ofthe layer thickness of the adsorbed layer in the case ofstrong van der Waals interactions (hydrophobic surfaces(εVW > τ)). The thickness of the adsorbed chains staysconstant equal to the size of the adsorption blob êvw inthe 2-D dilute (regime I in Figure 7) and semidilute(regime II in Figure 7) regimes. Then for the surfacecharge densities σ higher than a-2f it increases linearlywith the surface charge density. Finally, at high surfacecharge densities σ > σ*bead the surface charge densitydependence of the layer thickness is the same as for

δσ ≈fN

R2- σ ≈ εVW

2

lBrD f, rD > rD

VW (111)

R ≈ xrD lBNf

εVW, rD < rD

VW (112)

δσ ≈fN

R2- σ ≈ εVW

2

lBrD f, rD < rD

VW (113)

Figure 10. Dependence of the thickness of the adsorbed layerD on the surface charge density σ in salt-free solution(logarithmic scales).

Macromolecules, Vol. 35, No. 7, 2002 Adsorption of Hydrophobic Polyelectrolytes 2765

polyelectrolyte adsorption at hydrophilic surface (seeFigure 10).

It is useful to estimate the typical σ values for theboundaries between different regimes in the adsorptiondiagram (Figures 10 and 11). Consider polyelectrolytechains consisting of N ≈ 102 Kuhn segments with sizea ≈ 4A, τ = 0.1, and fraction f ≈ 0.1 of these segmentscharged, adsorbing from an aqueous solution with theBjerrum length lB ) 7A. In this case the boundariesbetween regimes in the adsorption diagram are σWC ≈10-5 A-2, σ* ≈ 4 × 10-4 A-2, σ*bead ≈ 2 × 10-3 A-2, andσion ≈ 10-2 A-2.

The polymer adsorbed amount

can either increase or decrease with addition of saltdepending on the surface charge density. For the surfacecharge densities σ below σ*bead in the regions where theadsorbed polyelectrolytes form 2-dimensional adsorbedlayers, the polymer surface coverage grows with in-creasing salt concentration (see Figure 12). The additionof salt screens electrostatic repulsion between adsorbedpolyelectrolytes, making the increase of the adsorbedamount energetically favorable. This increase of polymersurface coverage continues until the adsorption/desorp-tion threshold where the coverage decreases to zero.

Qualitatively different dependence of the surfacecoverage on salt concentration is predicted for surfacecharge densities σ > σ*bead (see Figure 12). In this rangeof surface charge densities the adsorbed polyelectrolytesform a 3-D adsorbed layer. There are two differenteffects that contribute to the decrease of the polymersurface excess with addition of salt. Inside the 3-D

adsorbed layer the addition of salt decreases the poly-mer adsorbed amount, because salt ions are also par-ticipate in screening of the surface charge. Additionaldecrease in the polymer surface excess comes from theedge of the adsorbed layer.

In the case of polyelectrolyte adsorption at hydropho-bic surfaces the polymer surface coverage increases withincreasing salt concentration in the dilute, semidilute,and 3-D regimes for surface charge densities σ < fa-2

(see regimes I, II, and III (for σ < fa-2) in Figure 7). Inthese regimes it is inversely proportional to the Debyeradius rD (see Figure 13). This growth in polymeradsorbed amount continues until rD g rD

3d. At highersalt concentrations (rD < rD

3d) the surface coveragesaturates at a-2. At high surface charge densities σ >f/a2 the polymer surface coverage decreases with saltconcentration in the same way as in the 3-D adsorbedlayer without van der Waals interactions discussed inthe previous paragraph.

The linear dependence of the polymer adsorbedamount on the square root of salt concentration (Γ ∼ rD

∼ xFsalt) was reported by Papanhuyzen et al.41 foradsorption of poly(styrenesulfonate) on crystals of poly-(oxymethylene) and by Marra et al.42 for adsorption ofpoly(styrenesulfonate) on weakly charged silica at pH≈ 2. A similar salt dependence of the polymer adsorbedamount was also reported by Yim et al.,43 who studiedthe adsorption of poly(styrenesulfonate) to the air/waterinterface. The adsorption of poly(styrenesulfonate) atthe air/water interface is due to a strong surfaceattraction arising from the low surface energy of poly-styrene in comparison with that of water. At low saltconcentrations the adsorbed chains form a single denselayer with almost flat chain configurations. This depen-dence of the polymer adsorbed amount on salt concen-tration is in a good qualitative agreement with ourprediction for polyelectrolyte adsorption on hydrophobicsurfaces (see regimes I, II, and III of the adsorptiondiagram Figure 7 where adsorption of polyelectrolytechains is due to strong affinity of polymer backbone tothe adsorbing surface).

Thus, our theory of adsorption of hydrophobic poly-electrolytes at charged surfaces gives a qualitativelyreasonable interpretation of the experimental results.Of course, our theory does not cover all aspects ofpolyelectrolyte adsorption. We leave aside the questionof the effect of the difference in the dielectric constantsof solvent and adsorbing substrate on polyelectrolyteadsorption. However, the difference in the dielectricconstants of the adsorbing substrate and the solvent can

Figure 11. Dependence of the thickness of the adsorbed layerD on the surface charge density σ in salt-free solution in thecase of strong van der Waals interactions (logarithmic scales).

Figure 12. Dependence of polymer surface coverage Γ on saltconcentration (reciprocal Debye radius) for polyelectrolyteadsorption at hydrophilic surface for three different surfacecharge densities σ1 < σ* , σsat < σ2 < f/a2, σ*bead < σ3 < σθ(logarithmic scales).

Γ ≈ σ + δσf

(114)

Figure 13. Dependence of polymer surface coverage Γ on saltconcentration (reciprocal Debye radius) for polyelectrolyteadsorption at hydrophobic surface for two different surfacecharge densities σ1 < σ*VW, σ*bead < σ2 < σθ (logarithmic scales).

2766 Dobrynin and Rubinstein Macromolecules, Vol. 35, No. 7, 2002

only lead to some quantitative corrections and will notchange the qualitative picture of polyelectrolyte adsorp-tion described in the present paper. The detailedanalysis of this effect requires special consideration andwill be addressed in a separate publication.

Acknowledgment. Authors are grateful to theNational Science Foundation for the financial supportunder Grants ECS-0103307 and DMR-0102267, and tothe donors of the Petroleum Research Fund, adminis-trated by the American Chemical Society, for financialsupport under Grant 37018-AC7.

AppendixConsider the fluctuations h(x,y) of the thickness D of

the adsorbed layer in the direction normal to polymer-solvent interface. For slowly varying fluctuations of thecharged polymer-solvent interface the additional freeenergy of the undulating interface with respect to thatof the flat one is

where the first term describes the surface free energyof undulating interface due to increase of its surfacearea and the second one is the energy of electrostaticinteraction between the charges dq(xi,yi) ) Ffh(xi,yi) dxidyi at the polymer-solvent interface. The chargedq(xi,yi) is positive for h(xi,yi) > 0 and negative for h(xi,yi)< 0, leading to the effective attraction between hills(h(xi,yi) > 0) and valleys (h(xi, yi) < 0) that flatters theinterface. Introducing the Fourier transformation h(q)

where qb is a two-dimensional vector with coordinatesqx and qy, one can rewrite the expression for the freeenergy of undulating interface with electrostatic inter-actions as follows

where we introduced q2 ) qx2 + qy

2. The probability offluctuations h(qb) is proportional to exp(-δF({h(qb)})/kT).In the Gaussian approximation the mean-square valueof the interface fluctuations h(qb) is

where we used the following relation 2πlBF2f 2êT2 ≈

a-3uf 2 ≈ Db-3. The correlation function ⟨h(qb) h(-qb)⟩ has

the maximum at the length scales 2π/q of the order of

the bead size Db. The mean-square roughness of theinterface about its flat configuration is given by

where we choose an upper limit of the integral to beêT

-1 due to the finite thickness of the polymer-solventinterface proportional to êT. Thus, the typical fluctua-tions of the thickness of the adsorbed layer D are of theorder of the thermal blob size êT.

The analysis presented above shows that the solutioneq 61 of the linearized Poisson-Boltzmann eq 59 foradsorbed layer with uniform density is stable signifi-cantly below the bead overlap surface charge densityσ*bead up to the surface charge density σ ≈ f/a2 wherethe thickness of the adsorbed layer D becomes compa-rable with the interface roughness êT.

We also have to check that this solution has higherenergy than the Wigner liquid of beads. The free energyper unit area of the adsorbed layer with uniform densityis

where the first term is the surface energy of polymer-solvent interface and the second one is the energy ofelectrostatic interactions within the adsorbed layer. Thesurface energy of the polymer-solvent interface (thefirst term on the right-hand side of eq 120) dominatesover the energy of electrostatic interactions (the secondone on the right-hand side of eq 120) for the surfacecharge densities σ below σ*bead.

The free energy per unit area of the Wigner liquid ofbeads also has two major contributions one is due toself-energy of the beads that includes beads surface andelectrostatic energies (at equilibrium these two termsare of the same order of magnitude) and the electrostaticattraction of a bead to the oppositely charged back-ground

In the interval of the surface charge densities σ < σ*beadthe first term on the right had side of eq 121 is biggerthan the second one. Thus, the adsorbed layer withstructure of Wigner liquid of beads has lower free energyper unit area than the adsorbed layer with uniformpolymer density as long as the cell size êb is bigger thanthe bead size Db. This is true for the surface chargedensities σ < σ*bead. At the surface charge density σ ≈σ*bead there is a first-order phase transition between theWigner liquid of beads and uniform adsorbed layer.

References and Notes

(1) Polyelectrolytes; Hara, M., Ed.; Marcel Dekker: New York,1993.

(2) Schmitz, K. S. Macroions in Solution and Colloidal Suspen-sion; VCH: Weinheim, 1993.

(3) de Gennes, P.-G.; Pincus, P.; Velasco, R. M.; Brochard, F. J.Phys. (Paris) 1976, 37, 1461.

δF({h(x,y)})kT

≈ 12êT

2∫dx dy (hx2(x,y) + hy

2(x,y)) +

lBF2f 2

2 ∫dx dy ∫dx1 dy1 ×

h(x,y) h(x1,y1) exp(-x(x - x1)

2 + (y - y1)2

rD)

x(x - x1)2 + (y - y1)

2(115)

h(qb) ) ∫dx dy h(x,y) exp(-i(qxx + qyy)) (116)

δF({h(qb)})kT

12êT

2 ∫dqx dqy(q2 +2πlBF2f 2êT

2

xq2 + rD-2 )h(qb) h(-qb) (117)

⟨h(qb) h(-qb)⟩ ≈ êT2(q2 + 1

xq2 + rD-2Db

3)-1(118)

⟨h(x,y)2⟩ )êT

2

(2π)2 ∫dqx dqy (q2 + 1

xq2 + rD-2Db

3)-1 ≈

êT2 ln(Db

êT) (119)

F3d

S≈ kT ( τ2

a2-

lBa3σ3

fτ ) (120)

Fbeads

S≈ kT ( τ2

a2

Db2

êb2

-lB mb fσ

êb ) (121)

Macromolecules, Vol. 35, No. 7, 2002 Adsorption of Hydrophobic Polyelectrolytes 2767

(4) de Gennes, P.-G. Scaling Concepts in Polymer Physics; CornellUniversity Press: Ithaca, NY, 1979.

(5) Khokhlov, A. R.; Khachaturian, K. A. Polymer 1982, 23,1742.

(6) Dobrynin, A. V.; Colby, R. H.; Rubinstein, M. Macromolecules1995, 28, 1859.

(7) Dobrynin, A. V.; Rubinstein, M.; Obukhov, S. P. Macromol-ecules 1996, 29, 2974.

(8) Solis, F. J.; Olvera de la Cruz, M. Macromolecules 1998, 31,5502.

(9) Dobrynin, A. V.; Rubinstein, M. Macromolecules 1999, 32,915.

(10) Micka, U.; Holm, C.; Kremer, K. Langmuir 1999, 15, 4033.(11) Lyulin, A. V.; Dunweg, B.; Borisov, O. V.; Darinskii, A. A.

Macromolecules 1999, 32, 3264.(12) Chodanowski, P.; Stoll, S. J. Chem. Phys. 1999, 111, 6069.(13) Micka, U.; Kremer, K. Europhys. Lett. 2000, 49, 189.(14) Castelnovo, M.; Sens, P.; Joanny, J.-F Eur. Phys. J. E 2000,

1, 115.(15) Balazs, A. C.; Singh, C.; Zhulina, E.; Pickett, G.; Chern, S.

S.; Lyatskaya, Y. Prog. Surf. Sci. 1997, 55, 181.(16) Vilgis, T. A.; Johner, A.; Joanny, J.-F. Eur. Phys. J. E 2000,

2, 289.(17) Tamashiro, M. N.; Schiessel, H. Macromolecules 2000, 33,

5263.(18) Netz, R. R.; Joanny, J.-F. Macromolecules 1999, 32, 9026.(19) Nguyen, T. T.; Grosberg, A. Yu.; Shklovskii, B. I. Phys. Rev.

Lett. 2000, 85, 1588. J. Chem. Phys. 2000, 113, 1110.(20) Joanny, J.-F.; Castelnovo, M.; Netz, R. J. Phys.: Condens.

Matter 2000, 12, A1.(21) Dobrynin, A. V.; Deshkovskii, A.; Rubinstein, M. Phys. Rev.

Lett. 2000, 84, 3101.(22) Dobrynin, A. V.; Deshkovskii, A.; Rubinstein, M. Macromol-

ecules 2001, 34, 3421.(23) Dobrynin, A. V. J. Chem. Phys. 2001, 114, 8145.(24) Rouzina, I.; Bloomfield, V. J. Phys. Chem. 1996, 100, 9977.(25) Perel, V. I.; Shklovskii, B. I. Physica A 1999, 274, 446.(26) Shklovskii, B. I. Phys. Rev. Lett. 1999, 82, 3268.(27) Varoqui, R. J. Phys. II 1993, 3, 1097.(28) Borukhov, I.; Andelman, D.; Orland, H. Europhys. Lett. 1995,

32, 499.(29) Joanny, J.-F. Eur. Phys. J. B 1999, 9, 117.(30) Lord Rayleigh Philos. Mag. 1882, 14, 182.(31) This definition of the adsorption/desorption transition cor-

responds to the ground-state dominance approximation. Inthis approximation a polymer chain considered to be adsorbedif there is discrete level in the chain energetic spectrum wellseparated from its continuous spectrum. See for review:Grosberg, A. Yu.; Khokhlov, A. R. Statistical Physics ofMacromolecules; AIP Press: Melville, NY, 1994

(32) For the Odijk-Skolnick-Fixman model of the electrostaticpersistence length in salt solutions33,34 the electrostaticpersistence length lp is due to intrachain electrostatic repul-sion and is proportional to the square of the Debye radiusrD. Considering a necklace as the polymer chain with theeffective bond length equal to the length of a string lstrconnecting two neighboring beads carrying the charge mbf,we can rewrite the Odijk-Skolnick-Fixman expression for

polyelectrolyte persistence length33,34 lp ≈ [lB(mb f)2rD2]/lstr

2

≈ rD2/êT, for rD > lstr. The crossover to a rigid chain behavior

occurs when the electrostatic persistent length lp is of theorder of the chain size Lnec. This takes place at the value ofthe Debye radius rD proportional to the transversal chain size

x⟨R⊥2⟩. For larger values of the Debye radius the chain can

be viewed as a charged rod of length Lnec. Polyelectrolytechains desorb when the electrostatic attraction between thecharged surface and a section of size lp is of the order ofthermal energy kT: |Wads| ≈ kTlBσfgêlp ≈ kT. The desorptiontransition takes place at the Debye screening length rD

des ≈(a/u1/2τ3/2σ)1/3, for rD > lstr.

(33) Odijk, T. J. Polym. Sci., Polym. Phys. Ed. 1977, 15, 477.(34) Skolnick, J.; Fixman, M. Macromolecules 1977, 10, 944.(35) The electrostatic part of the globular free energy has two

terms. One is associated with the electrostatic interactionbetween charged monomers within the globule, and thesecond one is due to formation of the electrical double layernear the globular surface. The bulk contribution may beestimated as the electrostatic self-energy of a bead of size rDand carrying a charge fFrD

3 multiplied by the number of suchbeads in the globule volume N/(FrD

3): W elglob ≈ kTN(τuf 2rD

2)/a2. This contribution to the free energy is smaller than onedue to short-range monomer-monomer interaction kTNτ2 bythe factor (rD

2êT/Db3) and can be neglected. The energy of the

electrical double layer can be estimated as the electrostaticenergy of the capacitor of thickness rD with the surface chargedensity fFrD and the surface area Rglob

2: W elsurf ≈ kTlBrD

(fFrD)2Rglob2 ≈ kTRglob

2(uf 2τ2rD3)/a5 ≈ kTRglob

2(τ2/a2)(rD/Db)3.This surface contribution to the globular free energy isweaker than the surface energy term due to polymer/solventinterface kTRglob

2τ2/a2 for the values of the Debye radiussmaller than the bead size Db and can be omitted from theglobular free energy as well.

(36) Grosberg, A. Yu.; Khokhlov, A. R. Statistical Physics ofMacromolecules; AIP Press: New York, 1994.

(37) Israelachvilli, J. N. Intermolecular and Surface Forses; Cor-nell University Press: Ithaca, NY, 1985.

(38) de Gennes, P. G. Macromolecules 1981, 14, 1637.(39) The exponential dependence of the layer thickness on the

surface charge density D ≈ êvw exp(σa2/f) is due to theneutralization condition of the surface charge by adsorbedpolyelectrolytes σ ≈ f∫0

D dz/[a2(êvw + z)] ) fa-2 ln(D/êvw).(40) The two-body monomer-monomer attraction between two

adsorption blobs with gvw ≈ εVW-2 monomers -τa3gvw

2/êvw3

≈ -τxgvw ≈ -τ/εVW is always weaker than that of three-body contacts a6gvw

3/êvw6 ≈ 1 for adsorption at hydrophobic

surfaces εVW > τ.(41) Papenhuyzen, J.; Fleer, G. J.; Bijsterbosh, B. H. J. Colloid

Interface Sci. 1985, 104, 530.(42) Marra , J.; van der Schee, H. A.; Fleer, G. J.; Lyklema, J. In

Adsorption in Solutions; Ottiwel, R., Rochester, C. H., Smith,A. L., Eds.; Academic Press: New York, 1983; p 245.

(43) Yim, H.; Kent, M.; Matheson, A.; Ivkov, R.; Satija, S.;Majewski, J.; Smith, G. S. Macromolecules 2000, 33, 6126.

MA0114943

2768 Dobrynin and Rubinstein Macromolecules, Vol. 35, No. 7, 2002