Advances in Colloid and Interface Science, 34 (1991) 477-535 Elsevier Science Publishers B.V., Amsterdam

477

Adsorption of Ions, Polyelectrolytes and Proteins M.A. Cohen Stuart, G. J. Fleer, J. Lyklema, W. Norde, and J.M.H.M. Scheutjens Department of Physical and Colloid Chemistry Dreijenplein

6

6703 HB Wageningen The Ne~erl~ds 1 IIltroduction ........................................................

.478

2 Electrosorption of ions and small molecules: chemical and electrical fntwractions.......................................................... 479 488 3 Lattice theory of electroso~tion ..................................... a) l&tiCe Variant Of the Gauy-ChapmanStem model ................ b) Volume filing and chemical interactions .......................... c ) Theoretical results ................................................ 4 Polyelectrolyte adsorption model .................................... a) Conformations b) Theoretical

.................................................... resufts................................................50

Strom2 nolvelectrolvtes ........................................... Weak Dolvelectrolvtes ............................................ 5 Polyelcctrolyte adsorption experiments .............................. a) Introductory b) Adsorption

remarks.............................................50 of strong ~iye~trolgtes

of weak polyelectrolgtes ............................... d) Metastablestates..................................................51 e) Adsorption of polyampholytes ..................................... 6 Protein adsorption ................................................... remarks .............................................

b) Adsorbed amount ................................................. c ) Co-adsorption of small tons ...................................... d) Energy and entropy of protein adsorption ........................ ...................................................... A&nowledgement ............................................................. References.

0001~8666/91/$20.65

0

1991-

492 492 499 499 1 501 504 508 8

............................ .509

c) Adsorption

a) Intr~~to~

.488

Elsevier Science Publishers B.V.

515 8 519 521 521 523 .527 -528 531 531

478

1 Introduction According interactions (i)

to

present-day

insights,

there

are

four

types

of

that play a role in colloid science:

Van der Waals

forces,

that are always

attractive

for particles

of the

same nature (ii)

Electrostatic

and attractive (iii)

Solvent

(iv)

Osmotic

interactions,

for opposite structure-based

adsorption,

and

are operative

entropic

been

polymer

interactions

are

may

known

elements

development

and larger polymers

two decades

experiments.

because

configurational considerable and

and electrostatic progress

relatively

effects

simple

and

adsorption

polyelectrolytes,

the state of the art. In describing

systematic

state

are strongly

we shall review

monograph

theory

has by a

that the

of Verwey

and

of

more

can

adsorption. already

Overbeekz).

be

the

complex

is far from complete.

polyelectrolyte

description

interrelated.

has been made in the area of ions, small

of

first

can have a

supported

we may

of

is much more complicated

and proteins

ion

the

the situation

polyampholytes

a

amounts

and is now

understanding

in small

small amounts

are involved,

At present,

the

of interactions

are understood. macromolecules

Although

which

of which

state. The existence

colloids

over the past

For charged

molecules

(i) and (ii)

and Overbeekz).

for more than a century:

destabilize

of appropriate

essential

macromolecular

and Verwey

a new

effect. As far as uncharged

developed

number

to

in which only factors

and Landaul)

is still in its inceptive

of type (iv) is already stabilizing

related

or repulsive.

theory for systems

Derjaguin

structure

understanding added

interactions

is now almost fifty years old and known as the DLVO-theory,

its founders

Solvent

for the same charge sign

short range forces

that may be attractive

The basic stability after

that are repulsive

signs

theoretical

molecules

In the present the electrostatic

many found

By treating

elements in

the

these

way, we hope to pay a tribute to these authors.

like paper, effects

recur

of

admirable effects

in a

479

Given the variation

of phenomena

in polyelectrolyte

and protein

adsorption, it is mandatory to systemize the treatment and we shall do so by adopting the following hierarchy: (1) The structure

of “simple” double layers and the adsorption

of low

molecular weight molecules and ions. In order to avoid unrealistically high ion concentrations

close to a highly charged surface, the well-

known concept of a (single) Stern-layer, exclusion

accounting

for the volume

between surface and ions besides specific adsorption,

is

very useful (section 2). (2) A self-consistent

field model of “simple” double layers which also

takes into account the mutual volume exclusion between the ions (and solvent). To that end, the multilayer Stem-model

introduced by

Bdhmer et al.31will be discussed (section 3). (3) Extension

of

this

incorporating

the

model

to

polyelectrolyte

configurational

properties

adsorption of

flexible

by chain

molecules (section 4). (4) Experimental verification of the theory for (strong and weak) simple polyelectrolytes experimental

(section

5).

In this

data on polyampholytes

section

also

some

recent

(for which the theory, so far,

has not yet been applied) will be given. (5) Adsorption

of proteins, for which the internal coherence (structural

stability) plays an important role. An a priory theory for these systems is not yet available, experimental

hence the approach

is essentially

based

on

evidence (section 6).

2 Electrosorption of ions and small molecules: chemical and electrical interactions The

distinction

between

electrical

and

chemical

contributions

pervades the entire domain of colloid and interface science. In its most elementary including

form,

it is reflected

charge-determining

in the adsorption (or

of simple

potential-determining)

ions

ions, and

counterions. Several features of colloid stability, the distinction between interaction

at constant potential

and at constant charge being one of

them, cannot be understood without properly recognizing

the chemical

and electrical contributions to the Gibbs energy of double layer formation.

480

Resolving

this

issue

was

one

theory. Also for our purpose At the

outset

“chemical”. suffices hence,

of the main

this distinction

it is necessary

to define

Since we will deal mainly

to identify

potentials

achievements deserves

of the DLVO

due attention.

the terms

“electrical”

with mean field approximations,

of mean force with average

potentials

it and,

use

(1)

I& = pi + viev for the relation the

and

between

chemical

potential

the electrochemical ).ti, and

potential

the

w. In eq. (1). vi is the

elementary

charge.

pi and w depend

potential

valency

In an inhomogeneous on the position

fii of component

(macroscopic

or

mean)

of component

system

z, but

i,

electric

i, and

e the

(such as a double

layer)

iii is constant

throughout

the

system. Equation

(1) implies

an element

of definition

to the Gibbs energy of an ionic species is by definition

chemical.

if the chemical

contribution

bonding,

hydrophobic

Henceforth,

form

adsorption symbol

by virtue

Au: while is that

tenets

realizing

Au:

specific hydrogen

. ..) is nonzero.

of the

terms may be included point

we shall call any adsorption

in colloid

chemical

that

science

attraction

of charge-determining

Au:

for by vrev

(such as Van der Waals interaction,

effects,

It is one of the basic layers

in that any contribution

i that is not accounted

ions

that electrical or (Gibbs)

to a surface.

non-configurational

double

energy

We

use

(or local)

of the

entropy

in it, thus making it a Gibbs energy. The important

does

not contain

the

configurational

entropy

of the

adsorptive. The charge-determining electrical

field

reflected (i)

in a number

chemical (ii)

is just

surface

Addition

continue

potential

v(O),

equal

to the

This

is

against

the

immediately

including

ions

and solution.

at the surface

adsorption. to

adsorb,

thereby

till this potential

difference

The driving

in the

gradually

has risen

chemical

so far

potential

force for the adsorption

is the

part Au; of this difference.

Indifferent

screen

their

of observations,

the surface

viev(O)

between

ions accumulate from

Charge-determining

increasing that

resulting

the

electrolytes

electrostatic

of such

(of which repulsion

electrolytes

allows

no ions adsorb between

specifically:

charge-determining

the adsorption

of the latter

Auy20) ions. ions to

481

proceed

further.

In other words,

surface

charge density

at given surface

potential

o(0) due to charge-determining

w(O), the

ions becomes

higher. There

is no distinction

ions and specifically

of principle

adsorbing

between

charge-determining

ions. Only if it is known from model

considerations that a certain ion type has a special affinity for the surface (e.g., Ag+ for silver halides, and H+ for oxides) does it make sense to identify them as “surface ions”. For protons on oxides Au: is of the order of

12-20

kT at room

whereas

temperature4).

Au: for

the

strongly

specifically binding Cdz+- ion is lo-16 kT51, which is not much lower. On the other hand, for weakly specifically adsorbing ions, like alkali ions on AgI, Au; is only a few kT-unit&I. It is now only one step towards small organic molecules. They may be charged or uncharged. For charged surfactants on hydrophobic surfaces, each CH2-group contributes an amount of order kT to Au;, whereas the electrical term for vi = 1 is about kT for each 25 mV of potential at the location where the charged group finds itself in the adsorbed state. The electrical and chemical contributions other,

depending

on the

indifferent

electrolyte

favourable

and adverse

screened. For long polyelectrolytes,

charge

may counteract signs.

Studying

may be diagnostic

organic

electrostatic molecules,

the picture

such

also contribute

polymers

to the adsorption

influence

because

as uncharged more

the driving force for the adsorption

both are and

because,

conformational

behaviour.

of

between

polymers

complicated

addition to chemical and electrical interactions, effects

the

in discriminating

contributions,

becomes

or reinforce each

in

entropy

For uncharged

is always of a chemical

nature. It is worth noting that a rather low chemical adsorption energy per segment (often a value well below 1 kT suffices) may lead to a strong binding of the chain as a whole. This is why linear uncharged polymers tend to adsorb tenaciously

on a variety of surfaces.

uncharged, electrolyte that is indifferent little effect on this adsorption the polymer).

However,

If the surface is

with respect to the surface has

(except when salt affects the solvency of

especially

for highly

charged

surfaces,

the

counterions neutralizing the surface charge compete for space and upon addition

of electrolyte

the adsorption

of polymer may decrease.

further discussion we refer to sections 4 and 5.

For a

482 Polyelectrolytes

have,

apart

from the chemical

contribution,

an

electrostatic part in their Gibbs energy of adsorption. With an uncharged surface,

this

electrostatic

polyelectrolyte.

effect

opposes

the

accumulation

of the

In the case of charged surfaces, there is an additional

electrostatic component in the adsorption energy, which may be positive or negative,

depending

screen all electrostatic

on the charge interactions,

signs.

Indifferent

electrolytes

both the intermolecular

repulsion

and the segment-surface interaction, so that they have at least two - often - effects.

antagonistic

These

situations

are more fully dealt with in

sections 4 and 5. In order to interpret more quantitatively the structure of the double layer and the effect of small molecules, the Gouy-Chapman picture 7. 81is usually inadequate, for at least two reasons: it considers the ions as point charges without volume (giving unrealistically high ion concentrations in regions

of high

adsorption. reasonably

A

and

potentials), simple

extension

it does

not

account

proposed

by

Sterns)

for

specific

works

in a number of cases, despite its relative simplicity.

quite In this

model, a chemical contribution is assigned to the adsorption energy of an ion and the volume exclusion between adsorbed ions and the surface is incorporated. Below we briefly treat this Stern-theory. A further extension to the more general

case where

also the mutual

overlap

of ions is

accounted for in a self-consistent way will be given in section 3. The basic premise of the Stern-model and ion volume

are considered

important

is that specific interactions only in the solution

layer

adjacent to the surface. The reasoning is that chemical interactions are short-ranged distance

and that usually the ion concentrations

from

the surface.

The potential

parallel

are low at some to the surface

is

considered to be smeared out, which is typical for a mean-field model. We denote the potential in the middle of this first layer, usually called the inner Helmholtz-layer.

as ~(1). Beyond the Stern-layer,

the outer Helmholtz-plane, Poisson-Boltzmann

that is, beyond

the double layer is assumed

to obey the

equation, i.e., its composition is formulated in terms

of the Gouy-Chapman theory. The

adsorption

isotherm

of a neutral

species

approximation, be written as a Langmuir equation:

i can, as a first

483

where $I = &(l) is the volume fraction of component i in the Stern-layer and $p is the bulk solution

volume

fraction.

If desired,

$i may be

identified with the fraction of the surface occupied by i. We prefer to use volume fractions in order to maintain consistency with the more general multilayer Stern-model to be discussed in section 3. Equation

(2) can conveniently

be derived

for a two-component

solution (solvent and one solute) using a lattice theorylo). It accounts for volume exclusion in the Stern-layer through the factor 1 - oi, but not for that in the solution: in order to achieve that, ‘pt should be replaced by @/(l - $).

Moreover, it is not generally valid for a system of more than

two components

(e.g., several types of ions and one or more neutral

solutes) because all these components occupy volume. For big monomers the statistics could be modified by letting more than one site match with one monomer, leading to a more complicated equation. In the Langmuir-picture,

Au; contains the adsorption energy and any

local entropy of adsorption: obviously the configurational entropy of the adsorbate is included in the term +I/( 1 - $1). If there is lateral interaction, an

additional

coordination parameter Moreover,

contribution number

Z&w@I could be added,

in the surface

that is positive

layer

for repulsion

where

Z, is the

and w a pair interaction

and negative

for charged species an electrostatic

for attraction.

term vieyr(1) enters. In

this way,

which

is a special

equationllv

form

of the Frumkin-Fowler-Guggenheim

(FFG)

12). Note that this equation does not explicitly account for

dipole orientation effects. Unlike eq. (2). which is explicit, equation (3) is an implicit equation in $1: not only does oi occur in the exponent but v( 1) is also a function of $1, and a suitable relation between the two has to be established. The FFG equation recurs as a special case in the multilayer Stern-model, see eq. (17) of section 3. The adsorption

energy AU: is then written as -x&T

where xS is a dimensionless adsorption energy parameter, and instead of w the conventional Flory-Huggins parameter x, which is a measure of the solute-solvent interaction, occurs. Moreover, the volume fraction oi in the lateral interaction term is replaced by the difference between a contact

484

fraction

in the surface layer and in the bulk solution.

The complete

solution in the multilayer Stern-model involves such an FFG-equation for each lattice layer in the system and not just for the ilrst layer. For all layers except the first, the adsorption energy term Au; is usually zero. For further details we refer to section 3. We shall now give a few experimental

examples

of low molecular

weight electrosorption. The purposes are to illustrate some of the trends discussed above, and to establish a bridge towards the following sections. Most examples given below are for AgI as the sorbent which, because of its hydrophobic surface and controllable surface charge and potential, is an excellent model system.

Figure 1. Surface charge on dispersed AgI as a function of pAg, in the presence of different concentrations of n-butanol13). Indifferent electrolyte 10-l M XN03. Temperature 25°C.

Figure 1 illustrates the influence of n-butanol on the electrical double layer on AgIlsl. The surface charge o(O) due to Ag+ or I- ions is plotted as a function of pAg I -log[Ag+l,

which quantity is directly related to the

surface potential y(O): if pAg increases by one unit, ~(01 becomes more negative electrolyte

by an amount concentration

of 58 mV (at ZOOC). In this experiment,

the

is high and kept constant, so that the diffuse

part of the double layer is compressed

and the Stern-layer,

where the

butanol is adsorbed, gives the dominating contribution. The picture shows that uncharged adsorbates can also screen the surface charge, though this is a more subtle process than with electrolytes. Butanol has basically two effects:

485 (i)

the lowering

of the dielectric

partial displacement

constant

in the Stern-layer

and the

of the counterions reduce the capacity 0(0)/w(O) of

this layer, which shows up as a smaller slope of the o(O)-pAg curve, and (ii) the replacement

of oriented water dipoles adjacent

to the surface

causes a shift in the point of zero charge (pzc), i.e., the pAg-value where o(0) = 0. If the curves were plotted as o(0) - w(O), the curves would all intersect each other at the pzc. If plotted against pAg, as in fig. 1, the curves are shifted along the horizontal axis and a common intersection point (tip) is found at a pAg above the pzc. In this tip the replacement of water in the Stern-layer by butanol has no consequence for the surface charge. The amount

of adsorbed

butanol

(rb) is too small to be detected

analytically, let alone the minor change of rb with pAg. However, it follows straightforwardly from the Gibbs adsorption equation that rb as a function of o(O) or, for that matter,

as a function

of pAg, passes

through

a

maximum at the tip. It is interesting to note that a very similar behaviour was found for n-butanol on mercuryl4). The model hydrophobic

interpretation

considers

the adsorption

bonding139 1s). Hence, on a butanol-covered

as driven by AgI-surface the

hydroxyl groups point towards the solution. A supporting observation is that

due

to

counter(cat)ions increases

this

butanol

adsorption

the

lyotropic

sequence

for

reverses. On bare hydrophobic AgI the specific binding

in the order Li+ < K+ < Rb+.

whereas

on butanol-covered

hydrophilized AgI it is the other way aroundIs).

PAg

Figure 2. Surface charge on dlspersed AgI as a function of pAg. in the presence of tetrabutyl ammonium (TBA+) nitratel6). Indifferent electrolyte O.lM KNOs. temperature 22oc.

486

Common

intersection

since been found including

for a variety

charged

ammonium function

ones.

(TBA+)

ammonium

points such as those observed

ions.

of other low molecular

Figure

nitratela).

Similar

As for n-butanol,

of o(0) or pAg passes

ions it was possible because

adsorbing surface

cations.

The

is positive,

point,

stronger

competition

positive

direction.

decreases

with increasing

sight

between

is rather

region,

I- and TBA+

for polyampholytes

and

increasing specifically

at low pAg, Au&.

where

adsorption

surface of the

be explained

sites.

experience

cation

by the

In addition,

where

attraction

the

To the right

sometimes

between

which

dipole less

adsorbate

TBA+ on

competition orientation

5 and 6, similar examples

proteins,

as a

methods.

with

o(O). Also this phenomenon,

for surface

if the electrostatic

ions

TBA+ ions, and o(0) shifts in the

may

effects may play a role. In sections found

adsorbs

the

negative

unexpected,

tetraalkyl

of these

by the

contribution

of the voluminous

for other

negative

the I- ions on the crowded

In this

tetrabutyl

at the tip. For some

of I- is facilitated

that TBA+

have

adsorptives,

for

by direct analytical more

points to a chemical

of the intersection

first

fact

apply

a maximum

becomes

adsorption

weight

example

adsorption

to verify this maximum

the

an

curves

the

through

To the left of the tip. o(0) [TBA+]

2 gives

for butanol

will be given adsorption

and

is

surface

is

increased. Koopall7)

has demonstrated

also be obtained When

the

adsorbed

carrier in trains

layer

electrophoretic electrolyte,

electrolyte

Having return

by train

this

to the statement

distinguish

between

established

the

is

that

above

on the segment chemical

of the

the data

and

with

electrostatic

diagnostic

species

we

of the tool

Figures

l9). They have in common cationic

the

distribution.

the influence

contributions.

of

of carrier

of small molecules

additional

and chemical

segments

the coverage

this information

that studying

a useful

of such trendsls.

adsorption

only

in low concentrations

useful information

electrostatic

and 4 are illustrations coverage

Combining

obtained

made

concentration

low

high,

on AgI.

By comparing

both play a role in the electrosorption

electrolyte

alcohol

he was able to determine

segments.

way

is

by this method.

layer thickness,

in

for polyvinyl

concentration

monomers,

then provided

interactions

particularly

are detected

with those of simple the first

that curves like those in figs. 1 and 2 can

for polymers,

decreases

to 3

that at with

487

Figure 3. Influence of the indifferent electrolyte concentration on the adsorption of tetrabutyl ammonium (TESA+) ions on AgIl*). The pAg was fixed at 11.5. temperature 22OC.

r

Figure 4. Influence of the indifferent electrolyte concentration on the adsorption of dodecylpyridinium chloride (DPC) on kaoIinitel** 19). The pH was fixed at 7. temperature 2ooc.

increasing salt concentration: then there is again a common intersection point, above which the salt effect is the other way around. Below the tip, the competition

between the cation of the electrolyte

and the organic

cation (ion exchange) leads to a smaller adsorption of the latter. At the tip, the electrophoretic mobility of the covered particles is zero: here the particles with their adsorbates are uncharged

(isoelectric). Beyond this

point, the screening by the salt makes the accumulation

of the organic

ions easier. The fact that the adsorption continues to increase beyond the tip points to a chemical driving force. Such an adsorption causes charge accumulation,

which

would

inhibit

further

adsorption

unless

these

488

charges higher

are screened. electrolyte

In

the

This screening

next

chapter

we

predicted

by rhe multilayer

3 Lattice

theory

Very model

Bohmer

were assigned just

volume

polymeric

were

avoided

(Debye-Htickel next

a of the segments

to vary with the distance could be modelled,

is not

types of

a mean field and ions

more consistent The theory only

u-eat

leaving

to assign

of Bijhmer

in those

the

approximation) surface.

case

model

where

and only electrical

to show sow- chemical

low

cases

where

by

the ionic success

However,

to all the components these demands.

molecular

aspects

weight

till section

each lattice

interactions

interactions

E

In this way weak is determined

layer

filling

in

In this

ions

is behaving

play a role. Then

and the volume

it

and

4. We separate

into two parts. First we deal with the lattice version

Gouy-Chaprnan-Stern

the

permittivity

new trends.

a volume

oi

or by

bater221,

have had considerable

et al.31 satisfies

the polyelectrolyte

relation

The fact that small ions have

data and predicting

the system.

as point

a linear

from the surface.

drawback

the

Very high counterion

to the

equilibrium.

a serious

is deffnitely

Stern-layer

lattice

(several

through

of .which the charge

experimental

the discussion

also

account

and the dielectric

in interpreting

we

into

by using

of a Stern-layer

are low, and these models

section,

species)

theory73).

concentrations

moiecuies,

takes

of this mode120. 211, polyelectrolytes

a local association-dissociation no volume

are

a multicomponent

which

and the potential

of ionination

polye!ectrolytes

trends

of all the components

molecules,

near the surface

the charge

were allowed

similar

a fixed charge and the small ions were considered

the incorporation degree

that

like in the Gouy-Chapman

concentrations between

find

et al31 presented

In earlier versions

charges,

shall

adsorption

and excluded

small ions, uncharged approach.

of

Stern-model.

polyelectrolyte

interactions

in the presence

of electrosorption

recently,

for

is more effective

concentrations.

of the like a

we proceed

constraint

are

incorporated. a) Lattice

variant of the Gouy-Chapman-Stem

The molecules one could

assign

are assumed solvent

model

to be arranged

molecules

on a lattice.

(H201, small

In principle,

ions (e.g., OH-. H30+,

489

Na+, Cl-, etc.),

and small uncharged

solutes

one lattice

site each.

However, since ions in aqueous solution are strongly hydrated, one could also take lattice cells with a volume equal to that of an (average) hydrated ion and assume an equal volume for clusters of water molecules. As water is a highly structured

solvent, it is not an unrealistic

assumption

to

consider the water molecules to be clustered. Lattice cells of this size are also considered to occupy one polyelectrolyte segment (section 4).

Figure 5.

(a) Lattice with a solution of solvent (open circles), ions, and uncharged solutes (black circles) in a sobent, next to a charged surface. lb) Potential profile v(z) and field strength profile E(z) in an electrical double layer according to the multilayer Stem-model. The field strength changes at each charged plane (z = 0. 1, . ..) due to the plane charge density C(Z). and at each boundary between the layers (z = l/2. l/3. . ..) due to a change in the dielectric permittivlty ELz).The potential drops linearly in each half-layer because there is no space charge.

Like in the Gouy-Chapman theory 7.8). the potential is only a function of the distance from the surface: lateral fluctuations

are neglected

by

adopting a mean-field approach. The charge of the ions and segments is assumed

to be located

on planes in the middle of each lattice layer

parallel to the surface, thus forming an assembly of Stern-layers

(see

figure 5). The layers are numbered from z = 0 (the surface) to z = M (in the bulk solution). The interplane separation e equals the diameter of a segment or ion. In between the midplanes there is no charge so that the electric

displacement

permittivity

D(z),

which

is the product

E(Z) and the electric field strength

of the

dielectric

E(z). is constant. The

490

quantity

E(Z) is the

permittivity

relative

dielectric

Q, of free space. According

constant

in layer

to standard

z times

D(z) = &(z)E(z) = ia z’=O

(4)

The field strength

E(z) changes

I, 2, . ..). because

of the presence

boundaries

(z=i,g,

discontinuously

E(Z) is taken

the various denoted

of charges.

...) there are changes

in each layer leads to another imation,

the

electrostatics:

components.

because

dielectric

to be a linear

at the mid-planes However,

a different

permittivity

combination

(z = 0,

also at the layer composition

E. As a first approx-

of the permittivities

If in layer z the volume fraction

of component

of i is

by o‘(z), we have (5)

e(z) = &4$(Z)

i where

Ei is the permittivity

extends

over solvent,

densities equals

v&z)e,

charge

uncharged

are functions

component

of the pure solutes

of the volume

component

The charge

charge

of each unit i

and ai(z) the degree of ionization

i in layer z, and e is the elementary

density

summation

and all ions. The plane

fractions.

where vi is the valency

i. The

charge.

Then

of

the plane

o(z) is given by:

o(z) = Cviai(z)eoi(z)

/ a,

(6)

i Here,

as is the surface

the surface zero

volume

area of a lattice

charge density fraction

site. Equation

is that

for i = S, where

S is the solid

$3(z) = 1 for z I 0 and @s(z) = 0 for z 2 1. Generally, species,

VI = -1 for negative

ones. The degree equilibrium polyacid).

which

(where constant surface

also for

adsorbent:

vi = 0 for uncharged

and vi = +l

can be unity (e.g., for H30+, concentration.

for positive OH-, and salt

For example,

of a weak acid residue A (or a segment

can be represented

the

of a weak

by the reaction

H+A-

W = H20

and

H = H30+),

Kd = (a~/(1 - aA)lIHI/(Wl. the

(or segments)

of the local counterion

dissociation

HA+W$

ions

of ionization

ions) or a function

(6) applies

o(O) by putting z = 0. In this case the only non-

ratio

(H)/(W)

should

is determined

by

In the concentration be replaced

by the

the

dissociation

gradient ratio

near the

of contact

491

fractions

(I$H(Z))/($W(Z)).

The contact fraction

(@i(z)) is the average

volume fraction of component i around a site in layer z. To this average, three layers (z - 1, z. and z + 1) contribute. Hence, (@l(z)) is given by (@i(Z)) = Q$,(z - l)+ Q$i(z) + h.#z + 1)

(7)

where ho is the fraction nearest neighbours within layer z. and X1 that with each of the layers z - 1 and z + 1. For example, in a hexagonal lattice b = 2x1 = 6/12, and in a cubic lattice Xo = 4x1 = 4/6. Obviously, Xo + 2X1= 1. For the local degree of dissociation we arrive at

ol*wl = 1+K~‘(oH(z))/(ow(Z)) Note

that

&

concentration.

is

a dimensionless

(8) quantity

The normally used dissociation

moles/P, corresponds

to K = Q( lOsa,dN&-1

rather

than

a molar

constant K. which is in . The conversion

factor is

the molar concentration of pure solvent. It is straightforward

to compute the electrical

potential

from the

field strength. The potential on each plane at z = 1, 2, 3, . .. is related to that on the previous plane by ~(z+l)=+{E(z)+E(z++)}

(9)

where I is the lattice layer thickness. We assume that all molecules obey Boltzmann statistics. Accordingly, the concentration proportional

of ions, solvent molecules, and other monomers are -u*(z)/kT , where ul(z) is the energy

to a Boltzmann factor e

(with respect to the bulk solution) of species i in layer z. In the bulk solution, up = u,(m) s 0. If the interactions nature,

ui(z) is simply the electrostatic

are only of an electrostatic

energy u:‘(z)

of monomer

i in

layer z:

uf(z) =vp,bdeW)

(10)

As before, the volume fraction of component denoted

as $.

The

volume

components is then obviously

fraction

i in the bulk solution is

distribution

of the monomeric

492

(11) If there

are

components solved u,(z)

only

present,

numerically. gives

equation

immediately

potential

In order

iteratively

until equation

uncharged total

density is

Nevertheless, accounted

and

for

magnitude are

present,

equations

simple

and

condition

positive,

hence

ui(z)

example

is not and

ui(z)

w(O)

equation

are

adjusted

finite

(12),

equation

lateral

below,

of

the

monomers

is

However, packing obeyed.

of molecules

layers

restricts

is easily

is the

high concentrations

factor of positive ions

not

and monovalent

concentrations

for all

than the original

(and association)

thickness

of negative

it follows

equation. no

sophisticated

so that unrealistically

in the

is zero

are inhomogeneously

restricted:

more

since dissociation

that

molecules

eq.

is already

the

above,

with the Gouy-Chapman

molecules

error

neutrality

energies

potential with

(10) is obeyed for all i and z.

(10) and (11). The weighting

exp(ev(z)/kT)

charge

(4) and (9) lead to the potential

the potential

If only uncharged

the

(7)

and plane

be consistent

of the potential

ions do not occur.

using equations

o(0) or surface

incorporated model

to

should

the model

Gouy-Chapman

can be

according

the permittivity

Only the charged

of

distributions

charge

in full agreement

constraint

fraction

equations

this,

polymeric

set which

the surface

in the

molecules.

distributed, the

that,

a closed

no

distribution

to achieve

It is clear

form

can then be calculated

Finally,

provided

This

4-11

and

an initial guess for the set of potentials

(5) and (6) provide

distributions.

is given.

interactions

the volume

(11). Any dissociation

on each plane, (10).

equations

For example,

and (8). Equations density

electrostatic

of

salt ions

obtained

from

ions in layer z is

is exp(-ev(z)/kT).

From

the

that half of the ions in the bulk solution (11)

leads

to

the

well-known

is

relation

%aIt (z) = &3Ii cosh(ev(z) / kT]. The volume fraction of uncharged molecules is 1 - okIt in any layer z. The total volume fraction in layer z is then 1 + o~gIt(cosh{e~(z)/kT) (according b)

- 11, which

to the Debye-Htickel

Volume filling

approaches

unity

approximation)

only

at low otaIt

or

at low v(z).

and chemicaI interactions

The next step is the complete of the ions and molecules.

incorporation

We require

of the hard core volume

that each lattice

layer is just

fully

493

occupied.

In terms of volume fractions the packing constraint

can be

expressed as: &(z)=l i

(12)

for all z. This boundary condition is imposed by a lateral pressure which affects the potential energy of all monomers in a given layer in the same way.

More specifically,

Boltzmann

statistics

are maintained

but the

potential energy ul(z) is incremented by a term u’(z), which is positive or negative since the local pressure can be higher or lower than in the bulk solution. The effect of such a term is that the Boltzmann multiplied

by a factor exp(-u’(z)/kT)

which normalizes

factors are

the sum of the

volume fractions to unity and at the same time induces a redistribution of the molecules

because

of the accompanying

change

potential profile. Note that no extra independent

in the electric

parameter is needed:

equation (12) fully determines the value of u’(z). Like the self-consistency of ion distribution the

profile

and electric potential, all molecule distributions

of u’(z)

become

self-consistent.

In the

example

exp(u’(z)/kT] would have the value 1 + Q!&(cosh(e~(z)/kT)

and

above,

- 11 for any z,

but the potential profile v(z) would be different from the simplified case u’(z) = 0. Obviously, electrical

the potential

contribution,

energies

may contain,

also other interactions.

apart

from

the

Let us first consider a

binary mixture of solvent 1 and uncharged solute 2, where the molecules are attracted by the surface so that their energy levels in the first layer are lower than in the bulk solution by amounts uf and uz, respectively. In this case u:‘(z)=0

for any z. If there are no other contact interactions

than try, ui(z) = 0 for all layers except the first. This implies @i(z) = @ for z 2 2, according to eq. (11). For z = 1 we have in this simple case l+(l) = u’(1) + UQ From

eq.

(13)

(11)

($2 / I$) exp(-Au; / kT) difference

between

we

can

where

a solute

now

eliminate

Au; = uz - ut and a solvent

is the

u’( 1):

$,(l)/@,(l) =

adsorption

If we define

a

dimensionless adsorption energy parameter xS as x, = - Au; / kT (which

is

positive if the solute adsorbs preferentially with eq. (12):

molecule.

energy

from the solvent), we have

For

dilute

solutions

Langmuir

(I$; -C-C1). this

equation,

It is straightforward l+(z) = u’(z) + The adsorption interactions For

the

model

energy contribution

can be used.

is identical

the

simple

is usually zero for z t 2, but long range

The electrical

contact energy

term is given

interactions, up

to XIJand to the contact

i. This contact

are defined

with

energy of a molecule

Generally,

+ ...

up

any

fraction

by eq. (10).

suitable

through

of a solute

i in the layer z with other molecules

molecule

to

other interactions.

One way is to express

The contact

proportional levels

be included. neighbour

parameter@). segment)

to incorporate

u?(z) + u;‘(z)+ q(z)

could

nearest

result

see eq. (2) in section 2.

solution

Flory-Huggins

molecule

x-

(or polymer

(or segments)

of type j is

(@I(z)) of molecules

j around

fraction

is formulated

in eq. (7). Since the energy

respect

to the bulk

solution,

i with all other components

the total

contact

j is given by

u~t(zl=kT~Xlj[(~j(z))-O~} 1 If the solvent expressing u:(z)

is monomeric,

it in the

and up(z),

solvent

underlining

Since the solvent

it is always profile

@l(z)

possible and

to eliminate

the

potential

that u’(z) is not an independent

is uncharged,

u’(z) by energies

variable.

u1e1 = 0 . For the ratio @i(z)/@1 (z) we find

from eqs. (11) and (15):

(17)

where equation

Au:(z)

= u:(z)

applies

simultaneous the volume

as before,

for any component

equations constraint

It is instructive the special

- u:(z)

and

Aup

= uft(z) - $(z).

i and any layer z. The whole

may be solved

numerically,

taking

This set of

into account

(12). to compare

eq. (17) with the FFG-equation

case of a two-component

system

(e.g., solvent

(3). For

and one type of

495

salt) where the adsorption is restricted to a monolayer, eq. (17) reduces to nearly the same form as eq. (3). Note, however, that the volume of the co-ions is neglected in the FFG-picture. Hence, if @I( 1) = 1 - o2( 1). where component

2 is the counter-ion. and @ = 1, the factors containing the

volume fractions are the same in eqs. (3) and (17). The term Au: in eq. (3) is identified term u:‘(l)

as the difference between u; and u:. The electrostatic

is the same in both models. For the contact term At$(l)

may write,

using

(@l(l))

XIX + 2~{(@2(1)) - $},

+ (@2(l))

we

= 1 - hl and eq. (16). A$(l)/kT=

where x = ~12 This contribution

closely resembles

the lateral interaction term in the FFG-equation: the constant hlx may be incorporated

in Auf,

and w may be replaced by 2x/Z, where Z is the

lattice coordination number. Since Z, = hoZ, the term &w& in eq. (3) can be written as 2koxoi. The full model gives, for small @,

essentially the

same result because in that case (@2(l)) = I&z(l). The illustrations given above show that the multilayer Stern-model is easily adapted to different types of interactions, which may be long range as well as short range. Before proceeding to the application of the model to polyelectrolyte shown

in section

adsorption (section 4), we compare some of the trends 2 (figures

l-4)

with the predictions

of the model

calculations. c)

Theoretical results

d(nr’

Figure 6. Comparison between the multilayer Stern-model and the GouyChapman model. Potential decay in the electrical double layer, for three concentrations of monovalent ions and ~(0) = 100 mV. The full curves were computed from the Poisson-Boltzmann equation, the various symbols correspond to computations with the multi-layer Stemmodel, using various lattice spacings C. All x and xS parameters are zero and E = 60 E, for all components.

In fig. 6 we show the potential profiles v(z) for a charged wall in the presence of indifferent electrolyte at three different salt concentrations cs, for an intermediate surface potential of 100 mV. The distance d in fig. 6 equals zl. The curves in fig. 6 were calculated using the Gouy-Chapman theory7.s)

based on the Poisson-Boltzmann

equation,

the points were

obtained from the multilayer model using different values for I. It is seen

496

that

for the given

surface

potential

(100 mV) the results

virtually

coincide, which means that volume exclusion plays almost no role. For higher pronounced,

surface

potentials

and the multilayer

the accumulation

of ions is more

model leads to a slower decay of v(z)

close to the surface and, consequently, a lower surface charge than the single Stern-layer

model. Further away from the surface (a few lattice

layers) the profile becomes exponential, with a decay length equal to the Debye-length. just like in the Gouy-Chapman

picture. These trends are

fully consistent with Monte Carlo simulations by Camey and Torrie24).

Figure 7. Potential decay in the electrical double layer in the absence (dotted curve) and presence [full curve) of a specifically adsorbing monomer A, according to the multilayer Stern-mode13). Parameters: - I+I(O)= 400 mV. c, = 0.1 M. x$” =lO, other xS and all x parameters zero, L = 0.6 nm, components, hexagonal

d (nm)

The effect of a neutral monomeric adsorbate on v(z) at 0.1 M salt is illustrated in fig. 7. The adsorbate is assumed to adsorb strongly (xs = 10). Notwithstanding displacing

the high surface potential

counterions

(400 mV) it is capable

from layer 1 so that the potential

of

decays more

slowly in layers 1 and 2 than it does in the absence of the adsorbate. Beyond layer 2 the potential decay is exponential again, with the same decay

length

(1 nm for

10-l M salt) as in the absence

of specific

adsorption. In fig. 8 we present the surface charge-surface

potential

relation

(‘titration curve’) for a similar case: a charged surface in the presence of 0.1

M indifferent

electrolyte

and

a neutral

specifically

adsorbing

adsorbate. It is seen that at given ~(0) the adsorbate reduces the amount of charge, simply because of the fact that it reduces the screening by counterions

close to the surface. Since o(O) is plotted as a function of

w(O). the two curves intersect at the p.z.c. When comparing this result with the data given in fig. 1 we note that the shape of the curves is somewhat

different. This is because the ions in this example have no

chemical

affinity

with

the

surface.

Assigning

them

a

non-zero

Au,” (and/or Au_“) would produce a shift in the pzc and would make the

497

curves asymmetric (unless Au: = Au_“), see fig. 9 below. Related to this is the question of the partial polarization the layer of water adjacent to the solid

phase,

Adsorption

which

contributes

to the

establishment

of the

of molecules in the first layer changes this polarization

pzc. but

this feature has not yet been included in the model.

m

100

-100 mV

J, (0)

-1.0, 05

Figure 8. Theoretical titration curves 40) - 1+40)In the absence and presence of a specifically adsorbing monomer A. The surface charge is given as the dimensionless quantity a&0)/e, the number of elementaly charges per lattice site. Parameters: X* = 3, ux = 2 where x = SOlvent or iOk). &A = 2h,. L = 0.31 nm. other parameters as in fig. 7.

I

I

.

,

U(O) e

-0.6

-0.6

-0.4

-0.2

0.0 0

- 200

-400

-600

-800 mv

tiC0)

Our next adsorbing

example

cation

(fig.

9) illustrates

on the charge

potential

Figure 9. Theoretical titration curves o(O) - ~(0) in the absence and presence of a specifically adsorbing catton A:. Parameters: x:=5, ! = 0.31 nni, other parameters as in fig. 7.

the effects relation,

of a strongly again

at 0.1 M

indifferent electrolyte. In this case the cation was modeled as As, i.e., as a trimer with a charge of +1/s for each segment. The segmental adsorption energy xf

was chosen to be 5, so that this cation, with a rather diffuse

charge and a total adsorption energy of -15 kT, somewhat resembles the TBA+

ion in figs.

2 and 3. The

conformational

properties,

treated

according to section 4b, are different but in this case the electrostatic effects dominate. Four titration curves o(O) - ~(0) are shown, for bulk solution volume fractions (I: of 0. 10-7. 10-6, and 10-S M. respectively.

It is clear that

498

now there is a shift in the p.z.c. towards

the positive

side and a common

intersection

v(O). Below

this tip the organic

cation

point

at a strongly

promotes

increases

co-adsorption

-o(O),

qualitatively

beyond

along

is again

affinity

for the surface.

considerably

as found

the same

curves

of negative

charge-determining

the tip the reverse

the same

explained

negative

related

applies.

All these

in the experiment

lines.

The

different

ions

and

trends

are

(fig. 2) and

shape

can be

of the o(O) - ~(0)

to the fact that the small ions have no chemical Even a small Au: or Au_” would high electrolyte

at this relatively

affect

the results

concentration.

Figure 10.

Theoretical adsorption isotherms of a monomeric cation A+ on a surface with surface potential ~(0) = -100 mV. at three concen-

trations of indifferent electrolyte. The adsorbed amount 13: is given in monolayers. Parameters: xt =15, EA = 5 ~0.C= 0.31 nm. other parameters as in fig. 7.

Our last example When A+ adsorbs between

of this section

onto a negative

adsorbate

and surface,

is again for a monomeric

surface,

but a repulsive

adsorbate

molecules.

Indifferent

may have

a reducing

or an enhancing

seen

in our

isotherms

final

of

logarithmically around

(at

for

competition

with

(including

to be screened the theoretical 4 is striking. to

monomeric

mV

salt

monolayers. so that

point

the

Above

that is

electrolyte

are

The

That

our

coverage,

reduced

plotted

semi-

to

at

repulsion

of A+: below

due

is

intersect

lateral

the

electrostatic point,

the

so that there is nothing

has no effect.

whereas

This

adsorption

In the intersection

that these figures picture

and therefore

curves

in fig. 10 with the experimental adsorptives,

neighbouring

10) where

adsorption

counterions.

It may be added units.

(fig.

potential)

enhances

both

cation A+. interaction

on the adsorption.

A+) is just isoelectric,

and indifferent

oligomeric

section

adsorption

adsorbed

screens

effect

surface

one between

concentrations.

salt

indifferent

curves

electrolyte

of this

-100

three

important

intersection

apply

A+

r = 0.12

becomes

surface

example

there is an attractive

The similarity

ones in figs. 3 and

(and also figs.

figures

nevertheless

of

1 and 2)

8 and

10 are

applies

so well

for is

499

explained

by the fact that in practice the electrostatic

effects, notably

those in the first layer, dominate. Having description

established

that the multilayer

of the electrosorption

turn our attention

Stern-model

gives a fair

of ions and small molecules, we now

to the extension

of this model

to polyelectrolyte

adsorption. 4 Polyelectrolyte adsorption model a) Conformations As compared previous

to the monomer

section,

a theory

electrosorption for

model discussed

in the

adsorption

again

polyelectrolyte

is

considerably more complicated because the conformational entropy of the flexible

chains has to be accounted

for. Earlier models had to make

drastic assumptions in order to cope with the complex interplay of long range electrostatics,

short range chemical interactions

and conforma-

tional effects. The first model is due to Hesselink25.26) and is an extension of an early polymer adsorption model of Hoeve27). The latter is based upon an assumed However,

exponential

Hesselink

assumption

form

calculated

of the segment the electrostatic

that the charge in the adsorbed

density

distribution.

free energy under the layer is homogeneously

distributed. As a result, very dilute and extremely thick adsorption layers were

predicted.

In more

recent

models,

which

have

been

briefly

discussed in the introduction of section 3, the segment density profile is not

predetermined

and

very

thin

layers

are

obtained

if the

salt

can deaf with polymers

without

the

properties

are

concentration is not too high. The multilayer introduction

Stern-model

of extra parameters.

The configurational

treated along the lines of the Scheutjens-Fleer

theoryzs-so) where all the

possible conformations are taken into account by assigning each segment of a chain a weighting factor e-urWKT that depends on the local environment

according

to eq. (15).

For polyelectrolytes,

contains an electrostatic contribution, given by eq. (10).

ui(z)

again

500

Figure 11. Schematic illustration of a lattice with a solution next to a solid adsorbent. The solution sites are filled with solvent (open circles). small ions, and three polyelectrolyte chains (of five segments each) in specified conformations c. d and e. Conformations c and d are adsorbed and have r”(l) = 3 and @(I) = 2 seg-ments. respectively. in the first layer. Conformation e is free (nonadsorbed). with f(4) = 1 and 175) = 4.

Here we do not treat publications39

28-30).

is a modification

the

model

in

detail

and

refer

to

earlier

Th e principles can be illustrated with fig. 11, which

of fig. 5a. In this figure three chain molecules

are

indicated in specified conformations (labeled c, d and e). Each conformation

may have a different energy, depending on the

position (layer number) of the segments. The energy of a chain is the sum of the energies ui(z) of its segments, taking into account their spatial distribution. (hence,

If ri is the number of segments of a chain of component i

x-i-1 is the number

conformation

of bonds),

the Boltzmann

factor Gy of a

c is then the product of the x-i Boltzmann

factors of the

segments, multiplied by a product of ri-1 bond weighting factors ho or 11. By summation over all possible conformations, the volume fraction profile of the polyelectrolyte component i has r:(z)

is obtained.

If conformation

c of (polymeric)

segments in layer z, the volume fraction profile

becomes:

(18) This equation also applies to monomers (ri = 1): in this case r:(z)

= 1 if

the monomer is in layer z, GF is just the simple Boltzmann factor e-u*(zl/KT, and the summation over c involves only one term, so that eq. (18) reduces to eq. (11). For chain molecules more complicated.

the expression

For example, the conformation

for Gf is

labeled c in fig. 11.

501

having

five segments

(rl = 5). two “parallel”

bonds,

and two “perpen-

dicular” bonds, has an energy u: = 3u,(l) + 2u,(2). In this case rF( 1) = 3, r:(2) = 2, and Gf = A$: exp(-ur / kT). Similarly, for conformations d and e in fig. 11 we have the Boltzmann factors Gy = hoh: exp(-up / kT) and GF = h&exp(-u:

/kT), where uf =2u,(l)+u,(2)+~,(3)+~,(4)

and u: = u,(4)+

4ui (5), respectively. The summation in eq. (18) is most easily performed using a matrix formalism. We refer to refs. 3 and 28-30 for more details of the theory and the numerical theoretical

procedures.

In the next six figures we give some

results obtained with the multilayer

Stem-model.

In a few

cases the data were obtained with an earlier version of the mode12o-22)but under such conditions that the difference, if any, is very small. b) Theoretical

results

Strong nolvelectrolvtes

1

In figures strong

10

5

z

15

20

Figure 12. Semilogarithmic segment density profiles of a strong polyelectrolyte adsorbing on an uncharged surface, at various salt concentrations3g). For comparison, the profile for an uncharged polymer is also given. Parameters: $b = 104. r = 2000, f = 0.71 IXIL~~= 2. x = 0.5. E = 80 q, for all components. hexagonal latice.

12 and 13 we show some results for the adsorption

polyelectrolytes.

Fig.

12 gives

semilogarithmic

of

concentration

profiles near an uncharged surface, at various salt concentrations cs and at a constant

polyelectrolyte

bulk solution volume

fraction

of 10-a. The

figure was computed with the model of Evers et al.22) but the results of the multilayer Stem-model

are virtually the same: only at the higher salt

502

concentrations

some minor

differences

occur. The most conspicuous

feature of fig. 12 is the minimum at low ionic strength, which originates from the high potential generated by the adsorbed molecuies,

repelling

other chains, that are consequently depleted. Under these conditions. the adsorbed

amount

conformations,

is low and the adsorbed

chains

adopt

very

flat

with only a few short loops and hardly any tails. As a

result, the chain length dependence is also very weak. With increasing cs the adsorbed amount (and the chain length dependence) increases and the minimum in the profile becomes shallower. ~though

the minimum at

low cs in fig. 12 is rather pronounced on the logarithmic scale used for o(z), it could probably not be detected experimentally, because of the low absolute values of the concentrations involved.

1.5 1

1.0

CS

2.0 M

Figure 13. The adsorbed amount e”p [in equivalent monolayers) of a strong polyelectrolyte (v = -1) as a function of the salt concentration c, at three surface charge densities ~(0)~). The abscissa axis scale is linear in 4,. Parameters: r = 500. ~~=t~,L=O.6nm,~~=1,~=0.5 for polymer-solvent or polymer-ion interaction. other xs and x parameters zero. Ed = 20~~. E = 80&, for the other components.

In fig. 13 we present adsorbed amounts as a function of indifferent salt concentration, for three different values of the (fixed] surface charge. The abscissa scale in this figure is linear in u’cs. The polymer is assumed to have a weak chemical (specific) interaction with the surface and the surface charge is taken opposite to that of the polymer. Small ions have only electrostatic interactions with the surface, At low ionic strength, the adsorption is nearly proportional to the surface charge. If the surface is uncharged, the adsorption is almost zero. and when the surface charge is increased, the amount of adsorbed polymer adjusts itself such that the surface charge is slightly overcompensated.

The extent of overcompen-

sation depends on xs. At high ionic strength the adsorption increases by nearly the same amount for all surface charge levels, the slope being slightly higher for lower surface charge. This implies that the adsorbed amount changes by only a small factor when the surface charge is high, but shows a dramatic relative increase when the surface is near to neutral.

503

Of course, the enhancement of adsorption by salt is due to the screening of lateral electrostatic repulsions in the polymer layer. We note that electrostatic

interactions

between

colloidal

particles

(with bare, smooth walls) are usually already entirely screened at salt concentrations

of about 0.1 M, whereas with adsorbed ‘polyelectrolytes

screening effects still play a role up to 2M. In colloid stability only longrange interactions are important, and since in 0.1 M salt the Debye length K-1 is only becomes

1 nm the electrostatic

repulsion

between

small with respect to the long-range

However,

the

conformation

(and

colloids

then

Van der Waals forces.

adsorption)

of polyelectrolytes

is

determined by interactions between charges in close proximity, and salt affects these interactions even if ~-1 is considerably below 1 nm. This is also the reason why a strong polyelectrolyte

does not reach the same

adsorption properties as the corresponding uncharged polymer would do, not even at very high ionic strength,

although

with increasing

cs its

behaviour tends towards that of the uncharged polymer. Since in most problems

of screening

due to electrolyte

the basic

length scale is the Debye length ~-1, which varies as ~2.~. it might be anticipated that the adsorbed amount of the polyelectrolyte also increases linearly

with

accordingly.

c or, for that matter, Indeed

the curves

with dc,.

Fig.

are (approximately)

13 was linear

plotted

in dc, for

higher salt concentrations, but at low cs the dependence is not so simple. In that range the charge compensation

between surface and adsorbed

polyelectrolyte dominates, and 0: depends more weakly on k,. One uncharged

might

compare

polymers.

polyelectrolyte

In the latter

adsorption

with

case two parameters

that

govern

of the

adsorption behaviour: xs for the segment-surface interaction and x for the segment-segment

interaction

in the solvent.

One could

visualize

a

polyelectrolyte as an uncharged polymer with an overall xs-parameter that contains

a chemical

part (xs as defined before) plus an electrostatic

contribution that depends on surface charge and ionic strength. If surface and polyelectrolyte

are oppositely

charged.

this electrostatic

part is

positive, favouring the adsorption. It decreases with increasing cs. In this case, salt opposes the adsorption of the polyelectrolyte. Similarly,

one

could

imagine

an

effective

X-parameter.

Its

electrostatic part is always negative because of the mutual electrostatic repulsion between the segments, and it also decreases with increasing cs. Hence, the electrostatic part of x tends to decrease the adsorbed amount.

504

At low cS the overall

x is low

polyelectrolytes

show

the same

trends

low adsorbed

amounts,

flat conformations

solvents:

dependence.

With

electrolytes

o(O), see fig. negative,

adsorption

interesting

situation

adsorption

energy

electrical

energy

adsorption

the adsorbed

energy

below.

14,

of

of a segment, vev(1).

Rewriting

Hiickel

approximation

o(o)cr =

e

This equation from

the surface

poly-

chemical

xS + hl~,

An

to

the

by a repulsive

out at xsc. the minimum

the conformational

this in terms of the surface

more

decreases.

contribution

is balanced

comes

to compensate

linear in

charge

transition:

entropy

loss

xsc = xS + hrx -

charge by using the Debye-

one obtains:

(x,+$x-x,)

(19)

can be used to estimate

the critical

good

and

is approximately

of the chain, we have an adsorption/desorption vev(1).

in (extremely)

and weak chain length

polyelectrolyte

the

If that balance

needed

adsorbing

x increases

By making

when

so that

polymers. amount

a negative

arises

negative),

as polymers

more like uncharged above,

13 and fig.

the

often

c S the effective

increasing

behave

As discussed

(and

surface

charge

the chemical

o(O)“’

where

contribution

xS + Xix

the polyelectrolyte

is just

fully desorbed. Of

course,

such

a

balance

between

chemical

interactions

could also occur with small molecules,

will

to a sharp

it lead

interaction effect.

transition.

for the many

It is

sensitive

this

probes

monomers

cooperativity of

This

the

which

makes

interactions

that

electrical

but only for polymers

is because

in a chain

and

the

leads that

each

same

polymers of

segment

to a cooperative their

are

very

monomers

experiences. Weak

nolvelectrolvtes It is possible

by including now

becomes

degree

to extend

a degree

following

relation

polyelectrolyte

of dissociation

va(z)e~(z),

of dissociation

eq. (19) to the case of weak a(z) into the electrical

as in eq. (10). Rewriting

ab in the bulk solution,

for the critical

just desorbs:

surface

a(z)

polyelectrolytes energy,

which

in terms

of the

Evers et al.22) derived

charge

~$0)~~ where

the

the weak

505

(xs+~lmc)

O(“)cr =Kcab+(l-ab)

(20)

(X,+h,X-&)

la5*

Numerical

results

for the adsorption

of a weak

polyelectrolyte

(pK = 4) as a function of o(O) obtained by Evers et al.221 are shown in fig. 14. The adsorption is indeed linear in o(O), with a slope that decreases if the polyelectrolyte molecules,

charge becomes less. For sufficiently charged macro-

there is a clear adsorption/desorption

transition

at o(O)Cr

which shifts towards more negative values as the negative polymer charge decreases. It is found that o(0)Cr as it comes out of the numerical results is well described by eq. (20).

a uncharged 0.5 -

1

2

3

4

5

6

PH

7

I3

Figure 15. Degree of dissociation a( 1) on the surface and ab in solution for a weak polyacid. as a function of pH3). The quantity all) was computed for three surface charge densities o(O). Parameters: c, = 0.1 M. pK = 5, other parameters as in fig. 13.

That the distinction between a(l) and oh is not meaningless is borne out by fig. 15 where a(11 is plotted as a function of pH for three values of surface

charge

density

a(O). For comparison,

ab is also shown. At

506

pH > 4.5, dissociation to the high density significantly expulsion adjusts a(l)

at the surface of segments.

enhance

the

is less than in the bulk solution,

However,

dissociation

of H+ from the surface

region.

itself such as to effectively

a strongly at low

due

The local degree

neutralize

surface

to the

can

strong

of dissociation

the surface charge:

at low pH

is higher than ab, the more so as o(O) increases.

2

4

Figure 16. Adsorbed amount of a weak polyacid as a function of pH for three fixed surface charges3). Parameters as in fig. 15.

8

6 PH

The amount has

effect

of the

is shown

positive

charge. of the

polymer

and

adsorption

chemical small

Three

we

surface

neutral. affinity

the

same

its adsorption adsorbing

amount

with

have

trends At

carries

to three different

as in fig. surface

increasing

charged 13: a low

charge

pH < 2 the

segments

competition

effects

driven by the

have to compete

Therefore surface

attraction

were

segments

already

between

a is

with

we now see a charge:

to find

discussed

as the towards

a place

to

in section

2.

the salt ions and the surface

this effect.

on positively

part of fig. 16 is at intermediate charged

below the pK of the dissociating as

for polymer

and

polymer

goes up, more and more small ions are attracted it harder

i.e. it

a constant

a strongly

is now entirely

charged.

adsorbed

is again weak,

surface

to the

Therefore,

is highly

on the

corresponding

proportional

the surface

enhance

understood

density

pH > 6, we

conformation.

The most interesting adsorption

At

to see

We may add that specific will further

are shown,

xs. However,

making

Similar

The

polymer

in adsorbed charge

charge

charge.

charge.

is nearly

flat

the surface, adsorb.

curves

expect

which

ions when

decrease

(negative)

surface

corresponding essentially

polyelectrolyte

in fig. 16. Here the polyelectrolyte

a pH-dependent

values

be

positive

pH,

due

follows.

As

surfaces

has a maximum

group of the polymer. the

chain

acquires

pH. Here the at a pH slightly

This maximum some

charge,

can the

507

attraction

to the surface is no longer purely chemical.

An additional

electrostatic contribution builds up, so that the adsorption is enhanced. Moreover, the polyelectrolyte

can now compete more efficiently with the

small ions in neutralizing the surface charge. Soon, however, the lateral repulsion

between

charges

on the chain reverses

the trend and the

adsorption decreases strongly until the charge on the polymer reaches it maximum value.

2

3

4

5

6

7

8

as in fig. parameters

15.

PH

If the pH is increased from 4 to 6. major structural changes occur as one would expect: strong reductions in size and mass of tails and loops and a concomitant increase in train fraction with increasing pH (fig. 17). The trends displayed in fig. 17 are easily understood from the discussion given above:

flat adsorbed

conformation

at low pH where the polymer is only weakly charged. The

chains at high pH and a more extended

transition occurs around or just below the pK, and coincides more or less with the maximum in fig. 16.

508

5 Polyelectrolyte adsorption experiments a)

Introductory In this

section

electrolytes

with

comparison have

remarks

particular,

this

means

will

largely

polymers

like

electrolytes predicts

that

every

exclude

we

will

monomer results

true

section.

systems

which

on

with

with

it should

properties,

versa.

an ionizable

simple,

strongly

referring obvious

group.

with

branched either

be emphasized

In

flexible

oligomers,

and with copolymers,

It is not always

essentially

or vice

concentrate carries

poly-

For this

unit

DNA,

Finally,

equilibrium

free energy.

for adsorbed

of the model

obtained

(double-stranded)

(proteins).

select

as those

like polyethyleneimine

or biological lowest

where

data

in the preceding

we must

characteristics

homopolymers

experimental

discussed

to be meaningful

the same

This

we compare

the theory

stiff poly-

synthetic

that the theory

to the

(stable)

that real systems

state

of

can reach

this state and, if they do, how rapidly this occurs. We might even suspect that

a polyelectrolyte

metastable

has

a good

chance

to be strongly

trapped

in a

state by many strong ionic bonds. We will return to this point

at the end of this section.

In fact, with proteins

(section

6) this feature

is

even more salient. The

adsorption

described

by

experimental

of simple,

no

more

parameters.

number

of techniques.

internal

degrees

addressed. accepted when

The

and

of long

contributions

when range

that

The normal The

masses however, the

of multiple

structure

conditions

is usually mass

as

are readily have

question

already

measured

so large leads

of by a

a number

of structure

anchoring

well

a function

must

of be

to the generally-

which leaves room for a lot of variation

(and polymer

the repulsion

between

(as in polyelectrolytes),

primary

structure)

monomer

units

a flat

structure

from loops and tails is most favourable,

of zero net monomer/monomer formation

molecules adsorbed

Adsorbed

possibility

experimental

For example,

the

Polymers,

of freedom

‘loop-train-tail’

small

than

interaction

with

whereas

(polymer

change.

is very strong minor

in the case

in a O-solvent)

the

of long loops and tails is much more likely. averaged

spatial

to the interface

shape

scatteringsz)

of

this

distribution

is usually

profile

has

of monomer

called been

units

the segment studied

and. for some cases, by neutron33)

by

(or segments)

density means

proillesr). of

neutron

or X-ray reflection34).

An

509

alternative is to study the thickness of the adsorbed layer, operationally defined

in terms of a particular

optical

techniques

escence3611, hydrodynamic sedimentation and pores, (surface

experiment.

(reflectometry 351, total

We mention internal

the use of

reflection

techniques 311 (determination

fluor-

of diffusion or

coefficients of colloidal particles, flow through capillaries electrokinetics),

force

balance,

and

‘disjoining

osmotic

pressure

pressure’

techniques311

of concentrated

colloidal

dispersions). It may be added that when people speak of the “average thickness” of an adsorbed polymer layer, they are referring to a quantity that is not unique in the sense that different types of experiment produce different “thicknesses”. In particular by hydrodynamic and electrokinetic techniques a “hydrodynamic thickness” is measured, determined by extended loops

and

tails.

“ellipsometric

This

thickness

thickness”

which

is usually reacts

much

on the

higher

index

than

the

of refraction

difference between polymer and solution and hence essentially reflects the inner parts of the adsorbates. b) Adsorption

of strong polyelectrolytes

1.0 h...

0.15 74 0.5

0

1

kg/ml M

v,

NoCl

2

3

4

780

:.. Sta.*

kg/

mol

5

dhm)

Figure 18. Segment densityprofilesfor poIystyrenesulphonateof two diiIerentmolecular weightson positively ifuUcurves)and negativeIy(dottedcurves)chargedpolystyrenelatex as obtainedby smalIangleneutronscattering37). A suitable model polyelectrolyte

is polystyrene

sulphonate

(PSS). It

contains strong negative sulphonate groups. It can be prepared from anionically defined Segment

polymerized narrow density

styrene

fractions profiles

negative polystyrene

and is commercially

over a large

range

for PSS adsorbed

available

of molecular

both on positive

in wellweights. and on

latex particles were measured by means of small

510

angle neutron is confined

scatterings71. These profiles show that most of the polymer

in a thin layer, even at moderate

trends shown in this figure are consistent computations,

ionic strength (fig. 18). The

with those predicted by model

see fig. 12. At the salt concentrations

used in the experi-

ment, the minima in the theoretical profiles are insignificant

or absent.

Figure 19. Adsorbed amount of polystyrene sulphonate of three different molecular weights as a function of the square root of the concentration NaCI. The adsorbent is po~yo~ethylene3s~ [full curves) or silica at pH 23g) (dashed curve).

When PSS is adsorbed the lateral interactions Such

systems

on (almost) uncharged

were studied

by Papenhuyzen

crystals of polyowethylene who

adsorbed

uncharged.

PSS on silica

Both authors

et al.381 who used single

at pH 2, where

amounts

concentration

(fig. 19). curves for

this

find that the adsorption

the adsorbed

curve

the effects of

(POM) as the substrate, and by Marra et al.39)

manner with ionic strength, the extent depending

theoretical

surfaces,

in the polymer layer show up in their purest form.

oxide

is virtually

increases

in a regular

on molecular

weight. If

I- are plotted against the square root of the salt are obtained

o(0) = 0 in fig.

is negligible

whereas

that are very similar to the

13. In the

at high

absence

electrolyte,

f

experimental

error, linearly with &,. The slope of the linear parts at high

es increases

with increasing

molecular

weight:

longer loops and tails (when the electrostatic Papenhuyzen

cs it increases,

of added

long chains

interactions

et al.381 have shown that the molecular

is well predicted

by the model. For a further discussion

within can form

are screened).

weight dependence of this aspect we

refer to figures 24 and 25, below. The very fact that adsorption must be due to a non-zero

does occur on these uncharged

chemical

adsorption

surfaces

energy (xs). The resuhs

obtained

with SiO2 (pH 2) and POM are very similar, from which we

deduce that xs is nearly the same for both substrates.

I0 mm-’ r c

0.5 -L--

/’

P

.A

/exp-

.@--

_-‘)A

0.5

AI

_A’

t 1

-rM

them ,A .--

,

I

1F3

mm*

I

10

10-l M

c*

Figure 20. Comparison between experimental adsorbed amount I? (in mg mW2) for polylysine on negatively charged Agl and theoretical amount 9; (in equivalent monolayers), as a function of the salt concentration2** *O). The polylysine (307 amino acid residues) was adsorbed from NaBr-solutions at pH = 6 and pAg = 11. The parameters used in the model are: r = 300, Qi = 10T4, up = 0.71, o(0) = 0.2 e/a, with a, = 0.3 nm , B= 4.85,~=0.6,e=8Oc, iraIl cubic lattice components, Go = 4/6).

Another interesting model polyelectrolyte is polylysine (PL), which is a polybase that can also be obtained in relatively well-defined with a narrow molecular

weight distribution.

The intrinsic

fractions pK of the

primary amine group (- 10.8) is such that the molecule is fully positively charged up to pH-values of about 8. The L-stereoisomer

(poly-L-lysine)

can develop a helical secondary structure but only at pH 2 10.5. Helices are not found with the racemic random poly-Dllysine. homopolyaminoacids.

polylysine

becomes

The

very

small,

becomes

adsorption

insoluble

of PL onto

Just like most when

the charge

silver-iodide

(AgI)

crystals20.401and onto silica and glass 411has been carefully studied by Van der Schee et al.201and Bonekamp et al. 4o+*ll, respectively. Both poly-L and poly-Dllysine relevant enhanced

were used but significant differences were not found in the

pH range

(2 8). As with the neutral

by the addition

surfaces,

of salt in the case of Agl

adsorption

is

(fig. 20). For

comparison, theoretical curves computed by Van der Schee et al. are also shown in fig. 20, with a log~i~mic

scale for es; the agreement is semi-

quantitative. In the case of silica, however, there is only a minor salt effect with a weak maximum around 10-l-10-2 M, rather than a

monotonous

increase (fig. 21). This is probably due to specific adsorption of (sodium) counterions on the silica, which compete for surface sites with the iysine residues, We then have two opposing tendencies: (if increased screening of lateral

interactions,

leading

to enhanced

adsorption:

(ii) increased

screening of attractive polymer surface interactions, leading to a weaker

512

binding that

the

polymer

of polymer.

Since

outcome

depends

the latter on the

and the substrate.

I

0.0 -

is a specific

chemical

In order to model

one would have to assign a chemical

1

effect

adsorption

I

natures

one, we expect of the

these results

salt,

the

theoretically,

energy to the small ion.

I’

mg m-2

A-&

lA

7.51 c.3

/

0.6 -

Ol.-

O---L

/

0

\

0

clz-

o----c

0

10-3

I.0 10-2

10-lM 0

c. Figure 21. Effect of the NaCl concentration on the adsorption of polylysine (307 amino acid resfdues) on silica [Aerosil 0x50). at three different pH values41).

(0)

Figure 22. Adsorption of polylysine (190 amino acid residues) on silica (Aerosol 0x50) from 0.01 M NaBr. as a function of the surface charge4f).

Figure 23. Adsorption of pofyacryhc acid (M = 2000 kg/mol) on hematite as a function of the surface charge, in the presence of 1 mM NaC104. The data were computed22) from the ex ertments of Gebhardt and Fuerstenau42 P The surface charge of hematite has been measured by potentiometrtc titration in the absence of polyacrylic acid.

The effects respectively. substrate

of the surface

charge

that can carry an appreciable

the polymer.

can be seen from figs. 22 and 23,

The data in fig. 22 are for polylysine Charge

force for adsorption.

compensation

charge

adsorbed

of opposite

is thus an additional,

onto silica,

a

sign to that of strong

driving

This is borne out by the results, which show that the

513

adsorbed amount at relatively low ionic strength (lea-lo-2 entirely electrostatically

M) is nearly

determined. Similar conclusions were drawn by

Evers et al.22) from data provided by Gebhardt and Fuerstenat@i,

for

polyacrylic acid adsorbed on hematite (fig. 23). The curves of figs. 22 and 23 display the linear behaviour predicted by the model calculations

(fig.

14) and show that the agreement between theory and experiment is quite satisfactory. The extrapolation of the linear section of fig. 22 to I = 0 gives a positive value for the critical surface charge density o(0)Cr, compare eq. (19). This points to a non-zero chemical component

to the adsorption

energy. A similar conclusion applies to fig. 23, where the charge signs are opposite.

r

6050-

Figure 24. Adsorption of polystyrene sulphonate on polyoqnnethylene crystals, as a function of molecular weight. at three NaCl concentrations38). M (kg/mole)

The effect of molecular weight M on the adsorption of PSS on POM and on silica, respectively,

is shown in figs. 24 and 25. At low ionic

strength the adsorbed molecules adopt very flat conformations and there is nearly no molecular weight dependence. As the ionic strength goes up, the slope of the F versus

log M curve increases

as well. This is in

agreement with the discussion given in section 3 (after fig. 13) about the influence

of salt on the overall (effective)

concluded

that the negative electrostatic

the solvent molecular

poorer with increasing

weight dependence

X-parameter.

contribution

There

it was

effectively makes

cs. For uncharged

polymers,

the

of P increases with increasing x. In poor

solvents F continues to increase with M as the chains become longer, but in good solvents F(M) reaches a plateau at high M. For PSS, we do not find such a plateau. Apparently, water is a rather poor solvent for PSS

514

when

the

charge

on

the

polymer

is strongly

screened

by

a high

salt

concentration.

M f kglmole)

Figure 25. Molecular weight dependence for the adsorption of polystyrene sulphonate on activated Cab-0-Sil. at pH 2 and various concentrations of NaCl (a) and MgCl,(b)3g). In the taken,

original

an attempt

weight dependence

paper@. was

39) from which

figures

24

made to compare quantitatively

of r with theoretical

computations.

and

25 were

the molecular These authors

used an early version of the polyelectrolyte adsorption theory, based upon an extension

of the Roe-theory431 instead

Scheutjens-Fleer

of the more sophisticated

mode12s-30). They concluded that the agreement was

satisfactory,

but that their model predicts too weak a molecular weight

dependence

at high M. This is probably due to an artefact in the Roe-

theory, where tails are neglected. It is very likely that a better agreement between theory and experiment would be obtained if the conformational statistics

are better accounted

for, e.g., by using the Seheutjens-Fleer model. So far, such a full analysis has not been made.

515

/ k

60~x10‘5monomole

6Oc

50

40

30

/g

/

0’

.P’O 4%

/

A0

_l-n

0

31

Figure 26. Adsorption isotherms of polystyrene sulphonate on polyoxymethylene crystals from 2.18 M NaCl. for three different molecular wetghts381. 200

100

Qk

may be concluded

It

experimental

trends

neutral

substrates

to what

adsorbing more

indicated effects

is

availabless).

macromolecules interpenetrate. have

rounded

could

adsorb

in mutually

shape. The mean-field

those

theory,

are these

for well-fractionated

(by chain some

explanation degradation)

evidence

is that ‘blobs’

the

adsorbed

which

to adsorption

of monomers,

than

for such

inhomogeneous:

repelling

that this leads with

on the

of sulphonation)

strongly

the

However,

isotherms

One possible

suggestion44)

expect

in common

degree

of the supplier:

laterally

is

One might more

than is common substrates.

Another

layer

adsorption

by the mean-field

(as to the

by the specifications

theory.

of PSS adsorption

are less homodisperse

heterogeneous

polyelectrolyte

which

is predicted

that in general

by recent

feature

in fig. 26, where

on well-defined

is that the PSS samples

examples

described

are much more ‘rounded’

polymers and/or

the above

One unexplained

is shown

In contrast

isotherms

from

are excellently

there is one exception. given.

3@l PPm

the do not

isotherms

i.e.,

a more

theory is of course unable to take this into

account. c) Adsorption

of weak polyelectrolytes

One

known

of the best

acid (PAA). It is a simple

model

polyacid,

weak

polyelectrolytes

with an intrinsic

is polyacrylic

dissociation

constant

516

K of pK = 4.6. Therefore, and almost side

chains,

so that

precipitation) candidate

at low charge charge.

(--NR:)

independent, polymer

densities

such

between

which

groups

surface

a suitable

on the effects

carried

to have

so that

show up most clearly.

(or even

It is therefore

concentrates

zero

of

out by Blaakmeer

(PS) lances with fully dissociating

in order charge,

virtually

no hydrophobic

as hypercoiling

are absent.

study

used polystyrene

surface

positive

charge

effects

Such a study was recently

These authors

positive

is variable

pH range. Also it contains

complicating

for a fundamental

the polymer et alas).

its dissociation

100% in a suitable

a constant,

the effects

We denote

i.e..

pH

of a variable

this positively

charged

latex as PS+.

- O.Wl

[KNOJ

M

zoappm IPW

In

section

adsorbing

on

intersect might

a negative the same

indeed

near to the plateau adsorption

model

affect

As

taken

the surface

to happen

adsorption

isotherms

at different

ionic

plus adsorbate

two adsorption

of the isotherm

which

a consequence,

isotherms

adsorbed

which

show a similar amount.

equilibrium

concentration.

The

that in the saturated

polydisperse

nature of the PAA used.

point

it is situated

difference

We

for PAA on PS+

but only by a very

experimentally

intersection

However,

strengths,

on a positive

by indifferent

is indeed

A+

point. This point is very

implies

screening

for

is isoelectric.

with poly A- adsorbing

seem to have an intersection

I only slightly,

calculations

(relative)

that

layer the surface charge is overcompensated,

amount.

should

saw

surface,

In fig. 27 we present

latex which

small

10) we

at the point where

expect

surface.

3 (fig.

(+160 mC/m2) at pH = 4.045).

at about

electrolyte found.

the same

at a much

is probably

The

related

lower to the

517

Above

we found

polyelectrolyte polyelectrolytes can

adjust

effectively complex

that

adsorption (figs.

their

the salt

dependence

13, 19 and 20). The reason

degree

of dissociation

the surface charge is nearly

concentration

is only weak, and much smaller

neutral.

such

for weak

than for strong

is that the weak groups as to compensate

more

(fig. 15). As a result, the substrate-adsorbate In the overall

balance

of all the interactions,

salt then has only a minor effect.

Jo

0

2

4

6

8

Figure 28. Comparison of the experimental adsorbed amount r (In mg/m*. left-hand scale) for a weak polyacid with the theoretical 0: (in equivalent monolayers. right-hand scale). as a function of PHIL). The experimental points are for polyacrylic acid (71 kg/mol) onto a cationic latex with ~(01 = 160 mC/m2 at an equilibrium concentration of 100 ppm in lo- 1 M KNOX. The full curve was computed with the model of Bdhmer et a1..3) using $I; = 104. pK = 4.25, o(O) = 160 mC/m2, cs = 0.1 M. L =0.6nm. r=500, x:=2, andX=0,5forthe polymer-solvent and polymerion interaction. The other xS and X-parameters are zero, and E = 80 ~0 except Ed which is 20 ~0.

10

w

The central fig.

28, where

function

and most striking the

adsorbed

of pH, at constant

there is a clear maximum

of Blaakmeer’s

result

amount

of PAA

ionic strength

onto

work is given in

PS+ is plotted

(10-l M KNOa).

It is seen that

at a pH below pK. Such a maximum

had been

predicted

four years earlier by Evers et al.22) and later by Bijhmer

see

16. A

fig.

maximum is given

qualitative

in fig. 28, where

Stem-model.

The agreement

the experimental one. Nevertheless, observation positively

explanation

was given in the discussion the curve

occurrence

was calculated

is about

Blaakmeer’s

of this interesting

system both polymer

the

study provides

TiO246) the effect

although

one pH-unit

pattern.

below

a

comparison

the multilayer the position

of

the theoretical

the first well-documented

In studies

was never

with

et al.3).

of such

of fig. 16. A quantitative

is quite satisfactory,

maximum

charged

for

as a

of PAA adsorption

discovered

and surface have a pH dependent

since

charge.

on

in that

518

d) Metastable

states

It has been noted more than once polyelectrolytes

do not necessarily

surrounding adsorbed

solution.

amount

For

changes

in surface

PL/AgI

system.

polyelectrolytes strong

macromolecules conditions finally

More

study

increasing

of

and not by chain

attachment

polymer/surface

is more

adsorbed

at pH

relaxation

processes

extremely

slow;

for

this

consistent et also)

3 onto

is

determined with

the

data.

for the adsorption

a carboxylic

in adsorbed

PS latex

polyelectrolyte

in the case studied

surface.

it was

by

the

provided found

by a

that

but decreased

would

predict

the

by

with

the

‘frozen’

conditions

Studies

the very

that the

determined

during

of adsorption

of polyvinyl-4-pyridine provide layers

they found

The

than by those

of molecules by

between

dictated was

entirely

A picture

that

by the

be sufficiently

rather

on coverage,

contacts

than

on the

viewpoint

Theory

rather observed

formed

conformations

in the that the

followed

might

processes

weight.

length.

positions

by Pefferkom

bonds

surface

et a1.49). where

molecular

and

reversibly

ionic

did not depend

coverage

kinetics

that

charged

have

evidence

number

conformations

route taken,

they first touch the surface

by Denoyel

of adsorption

with

not entirely

relaxation

then

to changes

Also Van der Schee4s)

seem

inhibit

would

imposed.

calorimetric much

would

under which

enthalpy

were

with adsorbed

et a1.47) found

on the experimental

and an oppositely

to efficiently

reversibly

Meadows

conditions.

charge

It

respond

example,

depends

on the final experimental

that systems

47. 48)

evidence

do occur,

a relaxation

that

but are

time

of 37

hours. We may conclude experimental should

results

be aware

from this that it is always with

equilibrium

theories,

of the fact that relaxation

very useful but when

to compare

doing

times may be widely

so one different

from one case to the next and that experimental

time scales

be chosen

in the first part of this

section found

accordingly.

illustrate

that it is perfectly

in experiment

very well.

The examples and those

discussed possible

obtained

to do so since

with equilibrium

may have to the trends

theories

agree

519

e) Adsorption

of polyampholytes

So far we have dealt exclusively with simple polyelectrolytes

which

carry only one kind of charged group. Restricting ourselves to this kind of system helped us understand the role of electrostatic interactions in the adsorption

of such macromolecules.

Proteins constitute a group of with at least two new complicating factors: (i)

charged macromolecules

they have important intramolecular attractions which give the molecule a certain

degree

of rigidity,

(ii) they carry both positive

charges, i.e., they are polyampholytes. between

simple polyelectrolytes

and negative

In an attempt to reduce the gap

and proteins we discuss here another

interesting

study carried out by Blaakmeer

ampholytic

polyaminoacids

were adsorbed

et a1.511, where onto PS latices.

synthetic both with

positive (PS+) and negative (PS-) surface groups.

r 0.4

-

.n’”

.2-w R \

0.2 -

a; ‘9

0

2

4

6

8

m

10

Figure 29. The adsorbed amount of the oligomer (Lys-Glu-GlyI4 onto negatively charged latex (-25 mC/m2) as a function of the pH. at an equilibrium concentration of 200 ppm51). The ionic strength (KC1 + HCl or KC1 + NaOH) was 0.1 M.

As polyampholytes, copolymers of glutamic acid and lysine were used. One polymer was a random copolymer of 60% glutamic acid and 40% lysine, with a molecular weight of about 20 kg/mol. The other adsorptive was a specially synthesized tetramer of L-lysyl-L-glutamyl-glycine, homodisperse

i.e., a

oligomer with twelve amino acid residues, of which four

carried a carboxylic

and four an amino group. This oligomer

(with an

isoionic point at 8.4) is soluble in water at any pH between 1 and 12, but the random copolymer appeared to be insoluble between pH 4.1 and 5.7, i.e., around its isoelectric point (iep) which is at pH = 4.9. If electrostatic effects would play the dominant

role one might expect

adsorption behaviour for the polyampholyte.

the following

At the negative side of the

i.e.p. adsorption onto PS+ would be preferred, whereas the reverse would

520

occur on PS-. However, both

sides

adsorbed

since the absolute

of the iep, lateral amount

becomes

less

adsorbed

amount

on each

repulsion side, just

at a pH where

it is highly

polymer

and surface

range

where

about

the same for the negatively

they attract

should

dictate

as P for

Hence

decrease

of the

polyelectrolyte a maximum

in the pH range

repel each other, and a minor decrease would

be more

a lowering

a weak

charged.

at or near the iep, a strong

where

pattern

value of the charge goes up on would

be the pattern

expected.

and positively

or less symmetrical

If xs would

charged

with

in the

segments,

respect

be that

to reversal

of

the surface charge. The adsorption

of the oligomer

onto PS- follows

(fig. 29) and the data for the random pH 9.7, respectively, to anticipate maxima

complex.

against

proteins

usually

to the iep of the protein

In this case, the explanation molecule

trend

at low pH and at

that for flexible

at a pH corresponding

of the protein

the expected

taken

with it (fig. 30). It is interesting

6 by mentioning

are also found

resilience

1 6 -

are also consistent

section

plus adsorbent

copolymer

involves

conformational

the internal

changes,

mg mm2 /

P!?

I-

Figure 30. The adsorbed amount of a Lys-Glu polypeptide onto negatively (-25 mC/m2) and positively (+90 mC/m2) charged latex as a function of the pH. at an equilibrium concentration of 200 ppm in 0.1 M KC151). The polypeptide (20 kg/mole) was a random copolymer of 40 mol % glutamic acid and 60 mole % lysine. It is insoluble in water around pH 5.

I I

&

\ \

q

w 08-

ps+ . T_A-

I ,I

I

a

\ \-*

.

.,.,..A

:

AA. 4

0

a

12

w

However, is lower even

at pH 7 the adsorbed

on PS+,

regions.

and -NRg difference

despite

Clearly,

adsorption. between

amount

of the polyampholyte

than on PS-, and at pH > 11 the adsorption the negative

As a possible

-SO,

and -NH;

explanation

is found effect

electric

charge

stronger

is considered

the ionic

Evidence

experimentsdl).

to be chemical

pH-

drives

than that between

size of the latter cation.

in conductometric

that

on PS+ or absent

in these

field which

we suggest

is perhaps

due to the larger

nonelectrostatic

polyampholyte

it is not only the overall

is small

the

bond -COO-

for such a

Of course

this

in our terminology,

521

corresponding

to a different xs. Another contributing

factor might be a

strong competition by the OH--anion for surface sites, especially at high pH. The matter is, however, not yet completely understood: certain asymmetry

in solution

properties

point also plays a role. Such assymmetry

with respect

perhaps a

to the isoionic

does indeed show up in the

viscosity of the copolymer solution. The study discussed above is only a first, but promising, step towards a better

understanding

polyampholytes.

of the interfacial

behaviour

The gap between simple polyampholytes

of

(synthetic)

and proteins is

still very great and statistical theories have so far not been invoked to interpret protein adsorption. The main reason is the difficulty of handling the multitude of specific internal interactions, some of them persisting in off-equilibrium

states. We shall therefore discuss proteins at interfaces

mainly from an experimental point of view. 6 Protein adsorption a) Introductory Proteins

remarks

are copolymers

containing

some twenty different

amino acid

monomers. The amino acids are linked to each other in a polypeptide chain, as illustrated in fig. 31. Two of the three bonds in the peptide unit are free to rotate, whereas the C-N bond is fixed because of its partial double bond character. The side groups R, R’, . .. are positioned in the trans configuration, so that rotation around the bonds in the main chain is minimally

hindered.

Furthermore, negatively

The side groups

vary greatly

some of the side groups charged,

so

that

the

in hydrophobicity.

are positively

protein

and others are

molecule

is

a

complex

polyampholyte. As compared

to synthetic

polyelectrolytes,

globular

aqueous environment adopt a compact three-dimensional

proteins

in

structure. This

structure is determined by various types of interaction occurring inside the protein aqueous

molecule,

environment.

but also between

the protein

The main structure-determining

molecule

and its

factors are: (i)

dehydration of hydrophobic side groups, which favours the formation of a compact structure, and (ii) the tendency to maximize the conformational

522

Figure 31. Structure of a peptideunit in a polypeptide chain. Rotation is possible around two of the three bonds, the middle C-N bond is fixed in the trans configuration. of the

entropy internal other

structures minor,

globular The

Gibbs

per

including

amino

acid

residue reasons

of the external

conditions,

In

investigating

metastable

structures,

considered.

All told,

the synthetic

of the

native

this

protein in many

ways

structure

the

that

proteins

discussed

are more

surfaces. sight

of proteins from the

from

of the different

data

aqueous

sides,

determined

by: (il changes

the protein

molecule,

the sorbent

surface,

seem between

sections,

which

protein

proteins

be

than and so

ions),

molecules.

a variety the

techniques.

of solid

relative

at

generally is mainly

surface

and

the protein

and

of charged

and (iii) structure

was

Although

it is now

between

of the

problem

and interfaces

a redistribution The

studies

of the adsorbent

interactions

involves

weight

onto system

conflicting,

in the hydration

(ii) Coulomb

low molecular

adsorbing

of the

using various

sometimes

that the interaction

(including

always

complicated

in the previous

solution

complexity

established

in the

of freezing

behaviour.

In view

approached first

purely

being one of

must

Over the past fifteen years there have been numerous adsorption

for

1 kT

by alteration

possibility state,

the

of a few

less than

of an adsorbent stable

native stable.

(taking

in the range

is so low

adsorption,

off the absolutely

the

marginally

can be easily changed

the introduction

of

and some

that

with much

amount

formation

These,

reason

is typically

conformations

polyelectrolytes

is their adsorption

the

This corresponds

and

thermodynamic them.

are

as the reference)

of protein.

the

thermodynamically

of stabilization

structure per gram

is usually

counteracts and P-sheets.

a-helices

contributions

structure

energy

which

molecule,

opposing

protein

unfolded Joules

protein

groups

rearrangements

importance

of these

523

contributions obviously depends on the nature of the system and in some systems

specific

attraction

may occur. Special

attention

deserve

the

structural alterations because this is a typical protein feature. Their Gibbs energy change ApsG (where the subscript ps stands for protein structure) is positive, because otherwise the molecule would spontaneously

pass

from its native state to another one. However, the rise in entropy (TAP&S) due to the enhanced randomization

to a large extent compensates

the

unfavourable ApsH due to structure breakdown, hence ApsG may be small and can then easily be surmounted by factors (i) or (ii). The occurrence of such conformational

changes

adsorption

of structurally

behaviour

allows

us to differentiate “soft” molecules

“rigid” ones (ApsG large). In comparison

between

(ApsG small) and

with polyampholytes

feature is that for soft proteins part of the three-dimensional can break

down, which

gives rise to an increase

the

the new structure

in conformational

entropy: rigid molecules rather behave like bulky monomers. Below some general features of protein adsorption will be discussed, after

which

the

various

observations

will

be

integrated

into

a

thermodynamic analysis. b) Adsorbed amount As for most polymers, protein adsorption isotherms more often than not show a high affinity character, i.e., the initial part of the isotherm coincides with the f-axis,

after which a (pseudo-)plateau

is reached.

However, protein adsorption isotherms having a finite initial slope are not rarities. When the adsorption isotherm is not of the high affinity type, the corresponding character.

desorption

isotherm

With polydisperse

usually

polymers

does show a high affinity

such irreversibility

has been

explained by the adsorption preference of the larger moleculessal. Upon increasing

the polymer

concentration

in solution,

the high molecular

weight fraction in the adsorbed layer becomes gradually enriched. If, after establishment molecules

of equilibrium,

the system

is diluted,

remain at the surface until concentrations

reached which are below the limit of detectability.

the adsorbed in solution

are

With proteins that

tend to aggregate in solution preferential adsorption of such aggregates over monomers has been founds3, s4. 55). which can explain the hysteresis. However, also in homodisperse protein systems real hysteresis has been observedss. 57). It implies that the adsorbing protein molecule differs from

524

the

desorbing

desorbed before

one;

protein

more

molecule

adsorption.

other

sources,

This

to

creation

of entropy58).

content%

is more

increase

for instance

system

induce

the findings

specifically,

by acid,

Thus

the

positive

of Gibbs

than

base

that

the

protein

be supplied added

by

to the

hysteresis

of hysteresis

implies

is in line with

secondary

a larger conformational

will return to this point in subsection

of the

of the

must

have a decreased

5% 60,611 and, consequently,

energy

or electrolyte,

the occurrence

proteins

Gibbs

energy

Furthermore,

desorption.

that desorbed

that

structure

entropy. We

d.

Figure 32. of the therms

Plateau values

adsorption isoas a function of

pH, for human serum albumin from 0.05 M

electrolyte solution on various surfaces: positive and negative polystyrene latex (PS), neutral polvoxymethylene cry&is (POM). negative silica (SiOz) and positive hematite (a-Fez@) 68. 85).

l-

0

n

I



4

I

I

I

I

I

5

6

7

6

9 PH

As

a rule,

plateaus. value

protein

For many proteins

at the isoelectric

32 such

adsorption

Tmax(pH)

isotherms

curves

for serum

It must

area of the adsorbent

be realized

electrostatic

material

in

the

repulsion both

hydrophobicity,

with

protein

between such

the specific uncertainty

collected

charged ample

to

their

of the

On the other

spontaneously

with

protein,

hand, the

evidence

that

than

molecules.

to

All

charge

curves

rather

the structural

(PS) the

than

are very

density

are strongly

and

affected

by protein-sorbent

alterations

native

lateral

Tmax(pH)

As the adsorbents surface

dissolved

polystyrene

point is due to structural rather

adsorbed

a maximum.

respect

are

regarding

molecule the

In fig.

surfaces

leads to a proportional

it is clear that the r ma(pH)

characteristics

interactions. place

on various

on negatively

al.67. 6s) ha ve

in fig. 32 show

different, by the

albumin

in l-ma, at either side of the isoelectric

rearrangements curves

et

attain a maximum

complex62-66).

that the uncertainty

in the value for rmax. For albumin Norde

Imz.

these plateau values,

surface

decrease

well-established

point of the protein-sorbent

shown.

surfaces,

develop

protein,

do not take hence

the

525

adsorbate is necessary to induce them. The fact that albumin also adsorbs at the hydrophilic su,rfaces of hematite (a-FeaOs) under

conditions

of electrostatic

repulsion,

and silica ($02).

implies

that very

even little

interaction between albumin and surface suffices to realize this structural change, after which the Gibbs energy of the adsorbed protein is so much lowered that high affinity binding ensues. As this dominant structural change is a protein property, the influence of the nature of the surface is suppressed. 1.5 mg I+ c

r mclr

0.5

PH

33. Plateau values of the adsorption isotherms as a function of pH. for ribonuclease from 0.01 M electrolyte on various surfaces 6g* 70]. For the abbreviattons of the adsorbents see fig. 32. Figure

Another category of proteins shows a different Pmax(pH) patternsas 70). This is illustrated

in fig. 33 for the adsorption

of ribonuclease

different surfaces. At the hydrophobic PS surface Pma(pH) independent

of the charge on the protein molecule

at

is essentially

and of the charge

contrast between the protein and the surface. The plateau values account for a complete monolayer of native molecules. This suggests that no or only minor structural changes in the adsorbing ribonuclease

molecules

occur. The behaviour is in agreement with the relatively strong internal coherence ribonuclease uncharged

of

the does

ribonuclease not adsorb

polyoxymethylene

molecule.

For

the

same

on the less hydrophobic, (POM)

surface,

and

also

reason,

essentially not

on the

hydrophilic surface of a-FezOs. unless it is electrostatically attracted. The relative importance of the contributions

of hydration changes,

Coulomb interaction, and structural rearrangements

for protein adsorp-

526

tion has been systematically studied experimentally by Arai and Norde71). They used a series of proteins that are similar in size and shape, i.e., lysozyme

(LSZ), ribonuclease

(RNase), myoglobin

(MGB), and a-lactal-

bumin (aLA). Unfolding induced by heat721 and by denaturing agents731 reveal that the structural order

LSZ, RNase,

summarized

stability of these

MGB, aLA.

Plateau

in fig. 34. The charge

proteins decreases in the

values

of the isotherms

of the proteins

are

is qualitatively

indicated by plus and minus signs. PS+;

5-_+32mV

PS -; c=-69mV

2 - mg m-2

rmar

r mar;

aLA1 l-

t&B? LsZ++RNase+

aFetOs+;

55+20mv

aFe20, - ; c_-_-47mv

I

Figure 34. Adsorption of lysozyme (LSZ). ribonuclease (RNase), myoglobin (MGB) and alactalbumin (c&A) on hydrophobic polystyrene (PSI and hydrophilic hematitie (a-FezO3) surfaces71). The electrolyte concentration was 0.05 M. An indication of the surface charge is given through the c-potential of the bare surface, and of the protein charge by + and signs.

At the hydrophobic PS surfaces all proteins adsorb. As a trend, the electrostatic interaction between the sorbent and the protein is reflected in Pmax. However, both on positive and negative PS, aLA adsorbs more strongly

than would

be expected

on the basis

of the electrostatic

interaction. This observation is in line with the relatively “soft” nature of this protein. At the hydrophilic a-Fez03

surfaces the adsorption of LSZ

and RNase is dominated by electrostatic interaction but again aLA behaves differently, for the same reason.

527

All of this fits into the picture that proteins that have a high structure stability

behave

as “hard” particles.

Their adsorption

is governed

by

electrostatic interaction and (partial) dehydration of the sorbent and the protein. Proteins having a low structure stability (“soft” proteins) possess an additional driving force for adsorption which is related to structural rearrangements. Consequently, soft proteins may adsorb spontaneously on a hydrophilic surface even under electrostatically unfavourable conditions. In competition from a mixture containing all four proteins, the least stable

aLA

tends

to

adsorb

preferentially74).

It follows

that the

contribution from structural rearrangements is an important factor in the competition,

resulting

in preferential

adsorption

of the “soft” proteins

over the “hard” ones. c) Co-adsorption of small ions

r/r_

Even

though

adsorbing

protein

Figure 35. The change of the electrokinetic charge upon adsorption of human plasma albumin onto hematite, as a function of the relative adsorption76). The charge effect results from the transfer of low molecular weight ions between the solution and the adsorbed layer. The electrolyte concentration was 0.01 M.

molecules

may

change

their

structure, they usually form a compact adsorbed layerso). Any ensuing adverse

build-up

of electrostatic

repulsion

can be avoided

by the

incorporation of small ions from solutions7* 75). The ion transfer between the solution and the adsorbed layer can be experimentally comparing

the electrokinetic

estimated by

charge density oek of the protein-covered

sorbent on the one hand and those of the bare sorbent and the dissolved protein

molecules

on the other 67. 70). Figure 35, taken from work by

Koutsoukos et a1.76),shows Aado&, the change of o& upon adsorption, as a function of the relative adsorption T/Fma for human serum albumin on negatively charged a-Fe203. The dependence of Aado& on pH points to an increasing uptake of positive ions with increasing negative charge on the protein and the sorbent surface. Furthermore, the linear relation between

and PiPmax [for Pifmax > 0.3) indicates

Aado&

adsorbed

protein

incorporated

molecule

is accompanied

that each additionally

by the same amount

of

charge.

Since the co-adsorption of ions reduces the charge antagonism, it is electrostatically

favourable. However, the chemical effect of transferring

low molecular weight ions from the aqueous solution into the non-aqueous protein

layer is unfavourable

and, hence, opposes the overall protein

adsorption process. Indeed, maximum affinity for protein adsorption has been found under circumstances

where the charge of the protein itself

just matches the charge on the sorbent surface75). d) Energy and entropy of protein a~sorp~n The

affinity

of the protein

adsorption

process

is, at constant

temperature T and pressure p, determined by the change in the Gibbs energy G AadG = AadH - TA,,S

(2 1)

or, if written per molecule of protein Aadg = Aadh - TA,os where H is the enthalpy and S the entropy and g, h and s stand for the same

quantities

per

thermodynamically spontaneously

molecule,

For

adsorption

to be

if the energy of activation is low. In the literature many

examples are encountered of extremely

respectively.

possible, AadG should be negative. The process occurs where protein adsorption is treated in terms

simple models. For instance,

various

authors77-s@

have

analyzed the isotherms as if obeying Langmuir premises, i.e., they assume, either or not tacitly. that the adsorption is reversible, site-bound, and that lateral interaction

and internal structural changes can be disregarded.

Obviously,

all of this is at variance

conclusion

derived from such interpretations,

of values to A,dG,

must be considered

with reality.

Consequently,

any

including the assignment

with suspicion.

For the same

reasons and for lack of definition of the notion “isosteric”, applying the Clausius-Clapeyron

equation

in order to analyze the influence

of the

temperature on the adsorbed amount is questionable. Yet, whatever the adsorption mechanism, Le Chatelier’s principle remains valid: the positive value for X/X

for the adsorption

of serum albumin on PS surfaces,

529

measured

at constant

protein

concentration

therms,

implies

an endothermic

sorbent

systems

AadH was determined

71. 76).

process,

It was found that this quantity

adsorption

is driven by entropy

In their thermodynamic and ribonuclease unravelled

determined

(which involves ments

sorbent

conditions

high

above,

groups

and rearrangeIn some

that TAadS should also be positive

and that,

the

TAP&S is one contribution

from hydrophobic

of charged

contribution),

force.

changes

enthalpy

subprocesses:

in hydration).

as to be the sole driving

derives

of serumalbumin

adsorption

redistribution

(including

at least under the given conditions, stated

overall

and a chemical

structure

AadH > 0, implying

67.

that the

Norde and Lyklemasl)

from the most important

surface,

an electrical

in the protein

confirming

of the adsorption

charged PS surfaces,

calorimetrically

of the

protein-

by microcalorimetry5s.

increase.

AadH in terms of its contributions dehydration

Aa,jH > 0. For various

directly

is often positive,

analysis

on negatively

the

in the initial parts of the iso-

process

One other

dehydration

is entropically

to T&S,

driven. As

but it can never be so important

of the surfaces,

entropic

term

but there are more

contributions. The

Gibbs

substance

energy

may

model substance) Gibbs

involved

be estimated between

energy

of dehydration For instance,

PS surface the estimatessl) AdhH = -1.4 25W).

Taking

becomes

firmly

more

are A&G

effect -TAdhS = -15.4 an adsorbed Such

at hydrophobic

of hydration

of a

substance

(or a

phase. Obviously,

negative

with

increasing

at 25OC of hydrophobic

mJ m-2 and a small enthalpy

amount

the

= -16.8 mJ m-2, which results from a

mJ m-2 (1 mJ m-2 corresponds

molecule.

of that

for the dehydration

with a molar mass of 50 kg/mol, protein

the state

partitioning

water and a non-aqueous

hydrophobicity. large entropy

in changing from

to about

0.25

of 1.5 mg m-2 for a protein this would

a large

value

surfaces,

even

yield

explains under

-T&hS

why

=

change

kT nm-2 at molecule

-226

kT per

all proteins

unfavourable

adsorb

electrostatic

conditions. The Gibbs energy change groups

contains

an electric

and Agch. respectively. accumulation

of net charge

electrical

depend double

from the redistribution

and a chemical

As ion incorporation

does not attain large values. and A&,

resulting

in the protein Its value

on the composition

(medium) compensates sorbent

of charged

contribution,

AgeI

for unfavourable

contact

region,

Agel

and sign, as well as those for Ahe and the dielectric

layers before and after adsorption.

constants

respectively.

of the

530

The ated

chemical

from

aqueous

model

studies

media.

positive

For

on the transfer

most

spontaneous

the chemical protein

development

contribution An

adsorbed

do not unfold indicates

that

than

potential

way

to non-

quantity

is

solution.

the

of a loose

constant

observation

structure,

effect

avoid

is the formation

The general

chemical

opposes

to

for water and small ions. In

layer the dielectric

the exposure

this

to ion incorporation

into such a loose the

aqueous

ions

alternative

layer, that is freely permeable

to that of the bulk

unfavourable

weight

can be estim-

negative values for both Aheh and Aseh

adsorption.

such a highly hydrated proteins

of ions from

molecular

of a high electrostatic

loopy adsorbed

layer,

low

as a result of compensating

82, 831, so that

close

Ag ch of ion incorporation

contribution

that globular

but form a compact

of ion

of hydrophobic

is relatively

incorporation parts

is less

of the protein

to

water as would occur upon unfolding. Adsorption

involves

a change

this, in turn, may induce the protein, molecule, they

which

are still

stabilized

shielded is

from

secondary

intramolecular

if

structures

of the protein to a reduction for the protein

spectrometry human and

Norde

molecule.

of secondary

SiO2.

method, adsorption, desorption

The which

structural desorbed

reduction

with possible

component.

from the a-helix

that

state corresponds structure.

structural

structure,

i.e.

the subscript

surfaces

ps

dichroism content

of

of a-Fe203

to the desorption

change

up to a certain

a reduction

data811 indeed

on circular

to be insensitive

the

groups

A loss of

conformational

in the a-helix

from the hydrophilic

restructuring

to a random

cause

protein

Based

expanded

side

sheets.

of calorimetric

a reduction

step. The helix breakdown

50% in the adsorbed

and pleated

ApsG < 0). As before,

appeared

suggests

acid

is

the sum of

a more

amino

(and tertiary)

of the

solution

whereas

may therefore

Analysis

parts of

of structural

in

which leads to an increased

et al.851 inferred

serum albumin

type

favours

between

as helices

ApsH > 0 and APsS > 0 (yielding stands

This

interaction

and

in such a way that

structure

interactions

interaction

structures,

Hydrophobic

surface

compact

interaction

of the protein

are in the interior

with water.

the

hydrophobic

hydrophobic

of such secondary

to the sorbent

determining

Hydrophobic

supports

rearrangements. environment

contact

probable

structure

structure.

points

exposed

by intramolecular

the other

entropy

structural

in an aqueous

may become

rearrangement

in the environment

occurs

extent

from 65% in the native

upon

during protein

the to

to ca. 90 amino acids transferred Based

on the assumption

of four

531

possible conformations

per peptide unit in the random structures41 and

one in the helix, the resulting increase in conformational

entropy, A+,

amounts to k ln4 per amino acid, or, for the entire protein molecule, 90 k ln4 = 125 k, so that TApss = +125 kT per protein molecule. Calorimetry under the same conditions gave Aadh values of +80 kT and +55 kT for serum albumin adsorption enthalpies

on a-Fe203

is driven

and SiO2. respectively.

by an entropy

increase.

It follows that the

These

positive

are sums of positive and negative contributions.

overall

When this

algebraic sum is less than Apsh, the entropy rise TApss may be enough to overcome Aadh, as in this example. The fact that Aadh c Apsh implies that the entropy rise due to structural rearrangements is not the sole driving force. However, it may be concluded that an increase in conformational entropy

plays an important

role in the adsorption

of the soft serum

albumin molecule on a hydrophilic surface with the same charge sign. Overviewing

all the features

involved

in protein

adsorption,

it is

obvious that basically the same configurational, chemical and electrostatic features occur as with polyelectrolytes or monomeric adsorptives. In principle, therefore,

the theoretical

picture

to proteins.

However,

can be extended

developed

for polyelectrolytes

in practice

this challenging

prospect is still far in the future because of the great number of specific interactions

in proteins,

a complication

that is most dramatically

re-

flected in the positive adsorption entropy, which is found in some cases. This positive contribution

is at least partially

attributable

to intrinsic

structural alterations. Acknowledgement The assistance of Marcel Biihmer in obtaining results with the multilayer Stern-model is gratefully acknowledged. References 1.

B.V. Dejaguin

and L.D. Landau, Acta Physicochim. U.R.S.S. 14 (1941)

633. 2.

E.J.W. Verwey

and J.Th.G.

Overbeek,

Theory

of the Stability

Lyophobic Colloids, Elsevier, (Amsterdam, New York) 1948.

of

532

3.

M.R. Bohmer, O.A. Evers, and J.M.H.M. Scheutjens, Macromolecules

4.

L.G.J. Fokkink, A. de Keizer, and J. LykIema, J. CoIIoid Interface Sci.

23 (1990) 2288. 127 (1989) 116. 5.

L.G.J. Fokkink, A. de Keizer, and J. LykIema, J. Colloid Interface Sci.

6.

J. Lyklema, Discuss. Faraday Sot. 42 (1966) 81.

135 (1990)

118.

7.

G. Gouy, J. Phys. (4) 9 (1910). 457; Ann. Phys. (9) 7 (1917) 129.

8.

D.L. Chapman, Phil. Mag. (6) 25 (1913) 475.

9.

0. Stem, 2. Elektrochem. 30 (1924) 508.

10. T. Hill, Introduction to Statistical Thermodynamics,

Addison-Wesley,

Reading (Mass, USA), London (1960). Ch. 7. 11. A.N. Frumkin, 2. Physik. Chem. I I6 (1925) 466. 12. R.H. Fowler

and E.A. Guggenheim,

Statistical

Thermodynamics,

Cambridge Univ. Press, Cambridge (U.K.) (1939) 426. 13. B.H. Bijsterbosch and J. Lyklema, J. Colloid Interface Sci. 20 (1965) 665. 14. E. Blomgren, J. O’M. Bockris, and C. Jesch, J. Phys. Chem. 65 (1961) 2000. 15. B. Vincent, B.H. Bijsterbosch, and J. LykIema, J. Colloid Interface Sci. 37 (1971) 171. 16. A. de Keizer and J. LykIema, J. CoIIoid Sci. 75 (1980) 171. 17. L.K. KoopaI and J. Lyklema, Faraday Discuss. Chem. Sot. 59 (1975) 230. 18. A. de Keizer, M.R. BGhmer, T. Mehrian, and L.K. KoopaI. Colloids Surfaces 51 (1990). 339. 19. T. Mehrian, A. de Keizer and J. Lyklema, paper presented at the 8th Int. Symp. on Surfactants in Solution, Gainesville, Fl. USA (1990). 20. H.A. van der Schee and J. Lyklema, J. Phys. Chem. 88 (1984) 6661. 21. J. Papenhuyzen,

H.A. van der Schee and G.J. Fleer, J. Colloid

Interface Sci. 104 (1985) 540. 22. O.A. Evers, G.J. Fleer, J.M.H.M. Scheutjens and J. Lyklema, J. CoIloid Interface Sci. I 11 (1986) 446. 23. P.J. Flory,

Principles

of Polymer

Chemistry,

Cornell

Univ. Press,

Ithaca, N.Y., (1953). 24. S.L. Cameyand G.M. Tonie. Adv. Chem. Phys. 56 141 (1984). 25. F.Th. HesseIink. J. Electroanal. Chem. 37 (1972) 317 . 26. F.Th. Hesselink. J. Colloid Interface Sci. 60 (1977) 448 .

533

27. C.A.J. Hoeve, J. Chem. Phys. 44 (1966) 1505: J. Polym. Sci. C30 (1968) 361. 28. J.M.H.M. Scheutjens and G.J. Fleer, J. Phys. Chem. 83 (1979) 1619. 29. J.M.H.M. Scheutjens and G.J. Fleer, J. Phys. Chem. 84 (1980) 178. 30. J.M.H.M. Scheutjens and G.J. Fleer, Macromolecules

18 (1985) 1882.

31. M.A. Cohen Stuart, T. Cosgrove, and B. Vincent, Adv. Colloid Interface Sci. 24 (1986) 143. 32. T.Cosgrove, T.L. Crowley, B. Vincent, K.G. Barnett, and Th.F. Tadros, Faraday Discuss. Chem. Sot. 16 (1982) 101. 33. F.D. Blum, B.R. Sinka, and F.C. Schwab,

Macromolecules 23 (1990)

3592. 34. M.J. Grundy, R.M. Richardson, S.J. Roses, J. Penfold and R.C. Ward, Thin Solid Films 159 (1988) 43. 35. P. Schaaf, P. Dejardin, and A. Schmitt, Langmuir 3 (1987) 1131. 36. I. Caucheteux. H. Hex-vet. R. Jerome, and F. Rondelez, J. Chem. Sot. Faraday Trans. 86 (1990) 1369. 37. T. Cosgrove, T.M. Obey, and B. Vincent. J. Colloid Interface Sci. II 1 (1986) 409. 38. J. Papenhuyzen, G.J. Fleer and B.H. Bijsterbosch, J. Colloid Interface Sci. 104 (1985) 530. 39. J. Marra.

H.A. van der Schee,

G.J. Fleer,

and J. Lyklema,

in

Adsorption from Solution, R. Ottewill, C.H. Rochester and A.L. Smith, eds., Acad. Press, (1983) p. 245. 40. B.C. Bonekamp, H.A. van der Schee and J. LykIema. Croat. Chem. Acta 56 (1983) 695. 41. B.C. Bonekamp and J. Lyklema, J. Colloid Interface Sci. 113

(1986)

67; B.C. Bonekamp. Ph.D. Thesis, Wageningen Agricultural University, The Netherlands

(1984).

42. J.E. Gebhart and D.W. Fuerstenau. Colloids Surfaces 7 (1983) 221. 43. R.J. Roe, J. Chem. Phys. 60 (1974) 4192. 44. M.A. Cohen Stuart, J. Phys. France 49 (1988) 1001. 45. J. Blaakmeer, Macromolecules

M.R. Bohmer,

M.A. Cohen Stuart

and G.J. Fleer,

23 (1990) 2301.

46. A. Foissy. A. El Attar, and J.M. Lamarche, J. Colloid Interface Sci. 96 (1983) 275. 47. J. Meadows, P.A. Williams, M.J. Garvey, R.A. Harrop, and G.O. Phillips, Colloids Surfaces 32 (1988) 275.

534

48. H.A. van der Schee, Thesis, Wageningen Agricultural University, The Netherlands 49.

(1984).

R. Denoyel, G. Dinand, F. Lafuma, and R. Audebert, J. Colloid Interface Sci. (1990), in press.

50. E. Pefferkorn

and H. Elaissari, J. Colloid Interface Sci. (1990). in

press. 51. J. Blaakmeer, M.A. Cohen Stuart, and G.J. Fleer, J. Colloid Interface Sci. (1990). in press. 52. M.A. Cohen Stuart, G.J. Fleer, and J.M.H.M. Scheutjens, J. Pol. Sci., Pol Phys. Ed. 18 (1980) 559. 53. R.L.J. Zsom, J. Colloid Interface Sci. 111 (1986) 434. 54. H.G.W. Lensen, D. Bargeman,

P. Bergveld,

C.A. Smolders,

and J.

Feyen, J. Colloid Interface Sci. 99 (1984) 1. 55. E. Brynda, M. Houska, and F. Lednicky, J. Colloid Interface Sci. 113 (1986)

164.

56. W. Norde, F. MacRitchie, Interface Sci. 112

G. Nowicka and J. Lyklema, J. Colloid

(1986) 1027.

57. H.P. Jennissen, Makromol. Chem. Macromol. Symp. 17 (1988) 111. 58. H.P. Jennissen,

in Surface and Interfacial

Aspects

of Biomedical

Polymers, Vol. 2, J.D. Andrade, ed., Plenum Press, New York, (1985) 295. 59. B.M.C. Chan and J.L. Brash, J. Colloid Interface Sci. 84 (1981) 263. 60. M.E. Soderquist and A.G. Walton, J. Colloid Interface Sci. 75 (1980) 386. 61. H. Sato, T. Tomiyama,

H. Morimoto, and A. Nakajima,. ACS Symp.

Series 343 (1987) 76. 62. P.G. Koutsoukos,

C.A. Mumme-Young,

W. Norde and J. Lyklema,

Colloids and Surfaces 5 (1982) 93. 63. P. Bagchi and S.M. Bimbaum, J. Colloid Interface Sci. 83 (1981) 460. 64. H. Shirahama, K. Takeda, and T. Suzawa, J. Colloid Interface Sci. 109 (1986) 552. 65. A.V. Elgersma,

R.L.J. Zsom. W. Norde. and J. Lyklema, J. Colloid

Interface Sci. (1990). in press. 66. A.V. Elgersma,

R.L.J. Zsom, W. Norde, and J. Lyklema,

Colloids

Surfaces 54 (1991), in press. 67. W. Norde and J. Lyklema, J. Colloid Interface Sci. 66 (1978) 257; 266: 277, 285; 295. 68. W. Norde, Adv. Colloid Interface Sci. 25 (1986) 267.

535

69. W. Norde, in Surfactants

in Solution, Vol. 5, K.L. Mittal, and P.

Bothorel, eds., Plenum Press, New York, 1986, p. 1027. 70. W. Norde, Colloids and Surfaces 10 (1984) 24. 7 1. T. Arai and W. Norde, Colloids Surfaces 51 (1990) 1. 72. P.L. Privalov, Adv. Protein Chem. 33 (1979) 167. 73. F. Ahmad and C.C. Birgelow, J. Biol. Chem. 257 (1982) 12935. 74. T. Arai and W. Norde, Colloids Surfaces 51 (1990) 17. 75. J.G.W.M. Fraaije, Ph.D. Thesis, Wageningen Agricultural University, The Netherlands

(1987).

76. P.G. Koutsoukos, W. Norde, and J. Lyklema, J. Colloid Interface Sci. 95 (1983) 385. 77. J.L. Brash and Q.M. Samak, J. Colloid Interface Sci. 65 (1978) 495. 78. A. Schmitt, R. Varoqui, S. Uniyal, and J.L. Brash, J. Colloid Interface Sci. 92 (1983) 25. 79. J.D. Aptel, J. C. Voegel, and A. Schmitt, Colloids Surfaces 29 (1988) 359. 80. B.R. Young, W.G. Pitt, and S.L. Cooper, J. Colloid Interface Sci. 124 (1988) 28. 81. W. Norde and J. Lyklema, J. Colloid Interface Sci. 71 (1979) 350. 82. M.H. Abraham, J. Chem. Sot. Faraday Trans. I 69 (1973) 1375. 83. M.H. Abraham, Nasehzadeh,

E.Ah-Sing,

and A.F. Danil de Namor, T. Hill, A.

and R.A. Schultz, J. Chem. Sot. Faraday Trans. I 74

(1978) 359. 84. T.E.

Creighton.

“Proteins,

Structures

and Molecular

1983, Ch. 5. 85. W. Norde and J. Lyklema, Colloids Surfaces 38 (1989) 1.

Properties”,

Adsorption of ions, polyelectrolytes and proteins.

Advances in Colloid and Interface Science, 34 (1991) 477-535 Elsevier Science Publishers B.V., Amsterdam 477 Adsorption of Ions, Polyelectrolytes an...
3MB Sizes 0 Downloads 0 Views