You are on page 1of 7

J. Phys. Chem.

1984, 88, 6661-6667


AI = (h2/2rmlkT)1/2

with h, k , and ml being Planks constant, Boltzmanns constant,


and the mass of a molecule of species 1, respectively, and Iluo is
an integral

6661

Aql loses the contribution from kinetic energy (first term on


right-hand side of eq 41) and measures the difference AZlno of
interactions of a molecule of species 1 with the two pure solvent
environments.
The activity coefficient is represented as follows
n

R T In y1 =
In this integral Nu is the number of solvent molecules in the system
of total volume V, ulu(r)is the potential energy of a molecule of
species 1 and a solvent molecule separated by distance r in the
absence of any other molecules (Le., a vacuum potential energy),
and gluois a weakened radial distribution function. We refer
elsewhere for the meaning of the weakening, or coupling, parameter X.30~31The important point for us here is the correspondence between the superscript on g and the limiting operation
of eq 39: glue, hence Zloo, reflects the mean potential energy of
interaction of a molecule of species 1 with a pure solvent environment (one mole of species 1 added to an infinite amount of
pure solvent). Note that when two different solvents are compared,

E l l j + ( I I u- ZlUo)

j- 1

(44)

The integrals I look just like Ila0in eq 43, but with j replacing
(r (and the obvious meanings for Nj and u,,) and without a superscript on g,,, the weakened radial distribution function for
solute molecules of species 1 and species j (where j can be 1 or
some other solute species). The first sum in eq 44, then, represents
the effect of solute-solute interactions in the solution at the
specified concentrations. These interactions are mediated by, and
so y 1 depends on, the solvent species, which appears implicitly
in g l j . The second term in eq 44 is the effect of the change in
solute-solvent interaction due to the nonzero concentrations of
all solute species.

A Lattice Theory of Polyelectrolyte Adsorption


H. A. Van der Schee and J. Lyklema*
Laboratory for Physical and Colloid Chemistry, Agricultural University, De Dreijen 6, 6703 BC Wageningen,
The Netherlands (Received: May 29, 1984)

The lattice theories for polymer adsorption of Roe and of Scheutjens and Fleer have been extended to include an electrical
term. Equilibrium segment profiles and adsorbed amounts are computed as a function of chain length, solvent quality, electrolyte
concentration, chain charge, surface charge, and adsorption energy. The most conspicuous difference with the adsorption
of uncharged polymers is that, due to the repulsion between chain segments, loop and tail formation is suppressed, so that
flat adsorption ensues with no influence of molecular weight. Addition of electrolytes screens this repulsion but very high
ionic strengths are required to approach the behavior of uncharged polymers. Agreement with experiments is satisfactory.

Introduction
The adsorption of polyelectrolytes is a matter of considerable
practical and theoretical interest. The phenomenon plays a role
in fields as disparate as water purification (adsorption of polyelectrolytes can lead to flocculation of particles, rendering them
more easily filterable), food technology (e.g., as emulsifiers in food
emulsions), pharmacy (also as emulsifiers), medical science (the
polyelectrolyte heparine is an anticoagulant), soil structure, paint
manufacture, etc. For a good understanding it is usually mandatory to have an insight into the conformation of the polyelectrolyte at the phase boundaries involved.
From the theoretical side, very few attempts have been made
to describe polyelectrolyte adsorption. The relatively most elaborate picture has been put forward by H e s ~ e l i n k - ~Although
this theory is able to explain a number of influences on the adsorption at least semiquantitatively, it has a number of limitations,
the most serious one being that a step function is assumed for the
polymer segment distribution in the adsorbate whereas this distribution is just one of the principal parameters sought. As in
Hesselinks model, Silberbergs4model does not contain an ab initio
derivation of the segment profile.

That theories for the conformation of adsorbed polyelectrolytes


have not yet been developed is not surprising, since the underlying
theory for the adsorption of uncharged macromolecules has only
recently become available, mainly through the contributions of
Scheutjens and Fleer.5~6Essentially, this theory is a lattice theory
in which lateral interaction is accounted for in the mean field
approximation. A characteristic feature is that individual conformations can be discriminated so that the contributions of trains,
loops, and tails can be evaluated. The earlier Roe theory is similar
to SF theory in that it is capable of describing the segment profile.
However, it does not distinguish between individual conformations.
Here we present extensions to the SF and Roe theories for the
case where the adsorbed macromolecules are charged. Hence this
is the first ab initio theory for the conformation of adsorbed
polyelectrolytes. The most detailed picture is of course obtained
with the SF theory. However, the Roe theory is relatively more
satisfactory for polyelectrolytes than for uncharged polymers
because with polyelectrolytes the formation of tails is largely
suppressed, virtually ignoring that tail formation is the main actual
shortcoming of Roe theory as compared with the S F approach.
As the Roe procedure offers fewer computational problems, a large
part of our results will be based on this theory.

(1) Hesselink, F. Th. In Adsorption from Solution at the Solid-Liquid


Interface; Rochester, C. H., Parfitt, G. D., Eds.; Academic Press: London,
1983.
(2) Hesselink, F. Th. J. Electroanal. Chem. Interfacial Electrochem. 1972,
37, 317.
(3) Hesselink, F. Th. J. Colloid Interface Sci. 1977, 60, 448.
(4) Silberberg, A. In Ions in Macromolecular and Biological Systems;
Everett, D. H., Vincent, B., Eds.; Scientechnica: Bristol, England, 1978; p
1.

General Theoretical Principle


The liquid phase adjacent to the adsorbent is mimicked by a
lattice of coordination number z. The lattice layers are numbered
( 5 ) Scheutjens, J. M. H. M.; Fleer, G. J. J. Phys. Chem. 1979,817, 1619.
( 6 ) Scheutjens, J. M. H. M.; Fleer, G. J. J. Phys. Chem. 1980, 84, 178.
(7) Roe, R.-J. J. Chem. Phys. 1974, 60, 4192.

0022-3654/84/2088-666l $ O 1.50/0 0 1984 American Chemical Societv

Van der Schee and Lyklema

6662 The Journal of Physical Chemistry, Vol. 88, No. 26, 1984
i = 1, 2, ..., M , 1 representing the surface layer and the layers
beyond M constituting the bulk. Each lattice site can be occupied
by either a polymer segment or solvent molecule. In order to better
imitate real molecules, variations are possible in which polymer
segments occupy more than one site or where the small ions have
a finite volume. Backfolding is allowed, which is a restriction of
the theory, especially in the case of polyelectrolytes, but this
limitation is partly offset by the fact that it is also accepted in
the bulk to which all interfacial quantities are normalized. At
present, we discuss homopolymers only.
In SF theory, the conformation of a molecule is taken as fully
defined by the number of segments the molecule has in any layer
i. A molecule is counted as adsorbed if it has at least one segment
in i = 1. The probability P, of finding a certain conformation
c is given by

P, = w,

i = 1 2 3 4

n p;'
i1

with
w,

xoqx1F-1-q

(2)

In these equations, p I is the weighting factor for individual segments in layer i (referred to p* = p M + 1 = p M + 2 = ... l), also
called free segment probabilities. They represent the probability
of finding a segment in i if this segment were not covalently bound
in the polymer chain. ri is the number of segments that conformation c has in i. The factor w, accounts for the lattice-determined number of steps that can be made along the chain from
a certain site to the next: Xo is the fraction of the number of
possible steps remaining in i, XI is the same for a step from i to
(i + 1) or ( i - l), e.g., for a hexagonal lattice Xo = 0.50 and A,
= 0.25. r is the number of segments per chain and q the number
of steps in i, all of this applying for conformation c. The degeneracy of conformation c is related to w, through

a, = Lz'lw,

(3)

where L is the number of sites per layer.


The free segment probabilitiesp , depend on the available volume
in i, given by 4i/$o.(the superscript 0 refers to the solvent, the
subscript * to the bulk, Le., 4* = 4M+l=
= ...) and on
the interaction energy with its neighbors, written according to
Flory-Huggins as

-2x((4O,) - 4*)
where

(4OA =

W01-+1X04O1 + X14O1+1

(4)

Here, x is the Flory-Huggins interaction parameter. For segments


in the first layer, an adsorption energy parameter xs must be
introduced. This quantity has been defined by Silberberg.*
Finally, for charged macromolecules an electrical energy term
for monovalent ions amounting to -me($i - $.)/kT = -abl -ye)
must be added, where a is the degree of charging and
the
average potential at i. Consequently

PI =

41(40M)-1 exP[-2X((41)

- 4O*)1 exP(~1,Xs)exP[-a(Yl

- Y*)l
(5)

It is inherent in this picture that for the average potential


at i is chosen instead of the potential of mean force. Although
for dissolved polyelectrolytes the difference is not too great because
of compensation of corrections? it is difficult to assess the quality
of the approximation for adsorbed polyelectrolytes. A rigorous
analysis also has to involve chain rigidity which is beyond the
present approach.
In principle, the segment density profile (&) (Le., q5i as a function
of i) can be obtained from (1) and ( 5 ) if $i(C$,) is available.
(8) Silberberg, A. J . Chem. Phys. 1968, 48, 2835.
(9) Fixman, M. J . Chem. Phys. 1979, 70, 4995.

bulk

4 4
01

Figure 1. Pictorial representation of an adsorbate and the potential


distribution. The top shows two molecules, A is adsorbed, B is not. For
A r = 32 and q = 14; for B r = 27 and q = 14. Bottom diagram:
potential profile for a positive surface potential #,, and a positive adsor-

bate. Th? bulk potentials have maxima


value is J/..

J/.
at the plates and their average

Iteration is necessary as the free segment probabilities themselves


depend on this profile. The equilibrium distribution is obtained
by maximizing the canonical partition function with respect to
the numbers of conformation and chains are generated by a matrix
procedure, as described by Scheutjens and Fleer.s,6
If the Roe theory is used, a somewhat different path is followed.
In this case the grand canonical partition function is written (eq
5 of ref 7) and maximized with respect to {C$J. If the adsorbate
is charged, the electrical free energy of each profile has to be
computed and included in the maximization.

Incorporation of Electrical Terms


Most of the computations have been done on the basis of a
model that is consistent with the mean field approach for the
underlying picture of uncharged polymers, Le., the polyelectrolyte
charges are thought to be concentrated in the centers of the lattice
layers and smeared out. The charge on each layer per unit area
is
ui = f m e 4 , / a o

(6)

where a. is the area of the unit cell. Counterions and co-ions are
considered point charges and distributed between these plate
charges according to Poisson-Boltzmann (PB) statistics. As a
rule, the full PB equation is used, Le., the Gouy-Chapman (GC)
approximation is analyzed. As this equation cannot be solved
analytically in our case, some computations have also been done
using the Debye-Hiickel (DH) approximation. It is along the
lines of the model to treat the bulk solution in equilibrium with
the adsorbate also as a repeating array of plate charges; here the
D H approximation is used throughout.
In order to verify the quality of some of the above model
assumptions, some computations have also been done with polyelectrolyte charges not confined to lattice layers but smeared
out over the cells and another one with small ions of finite sizes.
Figure 1 visualizes the model. Let us first consider the bulk
situation. Following Moller et a1.,I0 we introduce averaged
(10) Moller, W. J. H. M.; van Os, G. A. J.; Overbeek, J. Th. G. Trans.
Faraday SOC.1961, 57, 325.

Lattice Theory of Polyelectrolyte Adsorption

The Journal of Physical Chemistry, Vol. 88, No. 26, 1984 6663

stoichiometric concentrations fi+ and E- of the small ions per unit


cell of volume aorO,where ro is the plate distance. Because of
electroneutrality
fi+

- It-

f a&/aoro = 0

a.e

(7)

The sign in front of the term containing 4. is the sign of the


polyelectrolyte. If it is positive, fi+ may be identified as the
stoichiometric salt concentration in the bulk polyelectrolyte; if
the polyelectrolyte is negative, the same applies to fi-. The actual
concentrations n,(x) at any position between two plates is given
by
n*(x) = fi* exp[YJ(x) - ?+I

i, so that it covers the entire bulk phase. At the crests

(8)

where p+ = e$+/kT, etc., $(*) being those values of $(x) where


na(x) = fi+. For the region between the two plates the Poisson
equation reads

In the adsorbate layer, potentials can become so high that


application of the D H approximation is no longer justified. Also,
the additivity and the symmetry with respect to the mid-distances
r0/2 are lost. Because of the high y values, (9) must now be solved
numerically without approximation. There are two boundary
conditions.
(1) At the adsorbent surface (dy/dx),=, = -uoe/ttokT, where
uo is the surface charge density.
(2) At the solution side, i = M , potential and field strength
contain contributions from both the surface adsorbate charge
and from the repeating bulk plane charges. The potential y Mdue
to the bulk charges differs from y* in (18) since in (15 ) only the
sum for x > xM is needed. It reads

yM(bulk side) =
where tois the permittivity of free space and e the relative dielectric
constant. As in the bulk the potentials are usually small, (9) will
be analyzed in the D H approximation. Defining the reciprocal
Debye length aslo

= (fi+

K'

+ fi-)e2/ttokT
+

(11)

&a&/aoro- fi+jj+ - fi-jj-

e(ii+

+ ri-)/kT

Levine and Neale" pointed out that for the space between two
layers (1 1) is valid to lower potentials than for a semiinfinite space
near one plate, where fi+ = E-. The choice of b is equivalent to
choosing a value for the average potential in the bulk. As our
analysis contains only the difference in potential between bulk
and adsorbate, this choice is arbitrary and for sake of convenience
we set b = 0. This quantity does, however, play a role in describing
the Donnan potential between bulk polyelectrolyte and dialyzate
where the equilibrium electrolyte concentration equals 2fi+fi-/(fi+
fi-). Setting b = 0 implies that

+ jj+)

= fi-(l - j j - )

p+ = j j -

I jj*

(13)
(14)

(dashed horizontal line to the right of Figure 1, bottom)


In the DH approximation, the contributions to y(x) at any
position x in the bulk due to all plane charges uI are additive. For
one such layer, say layer i positioned at x = xi, the solution of
(11) is
y(i) =

- bulk side)]

K ~ M

(20)
- K ( $ ~

- 2qM(bulk side)).

with

fi+(

[d~(x)/dxl,i,, = -

so that the total field strength at the solution side equals

d2y(x)/dx2 = K ' ~ ( x ) b]

(19)

and its contributions to the field strength is [dy(x)/dx],=,, =


xyM(bulkside). The contribution of the charges on surface and
adsorbate are given by

(10)

the PB equation becomes

b=

ale
2tto~kT

-[exp(Kro) - 1]-'

ule
2toK k T

-exp[rK(x - Xi)]

where the - and + sign apply to the semiinfinite spaces x > xi


and x < xi, respectively. The total potential at x due to an infinite
array of parallel charges ur separated ro from each other is

Subject to the two boundary conditions, (9) is solved by a


sixth-order Runge-Kutta integration. For details of the mathematical procedure and iteration involved, see the Appendix. The
obtained potentials are substituted in ( 5 ) .
If the analysis is based on the Roe theory, the electrical free
energy Pel of any preset segment profile is needed. Generally, F,,
follows from integration of

where u[ is the charge on plate i during the isothermal-reversible


charging process. Different paths are open to carry out this
integration, e.g., depending on whether or not the small ions are
concomitantly charged and whether or not the bulk polyelectrolyte
is simultaneously charged. The equivalence of the former of these
alternatives has been verified by Casimir in the absence of polyelectrolyte;12the equivalence in the presence of polyelectrolyte
has now also been e~tab1ished.l~In view of the demands of the
numerical procedures we found it most convenient to charge the
adsorbed macromolecule with the bulk polyelectrolyte already
having the final charge, the free energy of which acts as the
reference. Computational details are collected in the Appendix.

Solutions for Low Potentials


Especially if xs is not very high, the amounts of polyelectrolyte
adsorbed remain relatively low and so do the potentials in the
adsorbate. Under that condition the Debye-Hiickel (DH) approximation is also applicable for the adsorbent and analytical
solutions can be given for the free energy. At any x in the
adsorbate the potentials are additive contributions of all plate
charges. Instead of (16) we have now for x between layers i and
(i 1)

Y(X) =
u0e

(16)
ale exp[-K(x - xi)] + exp(-xro) exp[x(x - xi)]
2tco~kT
1 - exp(-nro)

=-

exp(-xx) (22)

The last term is the contribution of the surface charge go on the


(17)

This equation applies to the region between xi and xi + r,, for any
(1 1) Levine, S.; Neale, G. J . Colloid Interface Sci. 1974, 49, 330.

CEOKkT

(12) Casimir, H. B. G. In "Theory of the Stability of Lyophobic Colloids";


Verweij, E. J. W., Overbeek, J. Th. G., Eds.; Elsevier: Amsterdam, 1948; p
63.
(13) van der Schee, H. A. Thesis, Agricultural University of Wageningen,
Netherlands, 1984.

Van der Schee and Lyklema

6664 The Journal of Physical Chemistry, Vol. 88, No. 26, 1984

0-

;-.c.

rl

-81,
-10

10

,
20

1s

Figure 2. Segment concentrationprofile for an uncharged (a = 0) and


fully charged (a = 1) polyelectrolyte: x = 0.5, x s = 2, r = 2000, uo =
0, 6. = lo4, a. = 1 nm2, ro = 1 nm. Hexagonal lattice. The concentration (M) of 1:l electrolyte is indicated. Roes theory.

adsorbent. Unlike in (16) all plane charges are now different,


but the distribution can be computed if (q+) is known, using ( 6 ) .
By the same token, for the potentials at the planes
-

10

15

Figure 3. Effect of chain length on the segment concentration profile:


x = 0.5, x s = 4, a. = 1 nm2, ro = 1 nm,uo = -e/lOao, 1:l electrolyte,

0.01 M. SF theory.
OF,

h l

The ( i in the last term derives from the fact that the first
layer is at r0/2 from the surface.
In the SF picture these potentials are substituted in (5).
If the Roe theory is used, the free energy is needed. Extending
(21) to a multilayer system having in each layer an excess of
charge over that on bulk layers amounting to (ai - a,) dX, if X
is the charging parameter, gives
dFeI = C+j(X)(aj i

US)

dX

20

(24)

a=l

which in view of the proportionality of


to give
FeI = XC$jCuj
i

with X can be integrated

- a*)

(25)

Results
Figure 2 is a typical example of a segment density profile. In
order to exaggerate the qualitative features the plot is made
semilogarithmic. The most conspicuous conclusion is that at low
salt content, where charge effects of the polyelectrolyte count most
heavily, a region of negative adsorption occurs. Significant adsorption takes place only in the layer(s) adjacent to the surface.
In other words, polyelectrolytes adsorb in flat layers because the
mutual repulsion between the charges on the chain inhibits loop
formation.
If electrolyte is added, the profile progressively resembles that
of an uncharged polymer. For a = 0, the part at low i is linear,
Le., the $i(i)distribution is exponential, but at large i @i decays
more slowly to its bulk value 4..
Considering the experimentally established fact that hydrophobic colloids are coagulated by not more than 0.2 M (1:l)
electrolytes, the very high salt concentrations that are needed to
screen the polyelectrolyte charges completely are unexpected.
However, this feature seems well-established; it is also observed
if the theory is modified such as to let each small ion occupy a

J
LA
I

The Journal of Physical Chemistry, Vol. 88, No. 26, 1984 6665

Lattice Theory of Polyelectrolyte Adsorption


0

-2

:1,

r.10

O01. 1

Q.

lo-'

5~10-~

Figure 7. Theoretical adsorption isotherms for polyelectrolytes as a


function of the number of segments r per chain: x = 0.5, xs = 4, a =
1, u0 = -e/lOao, a. = 1 nm2, ro = 1 nm,
M 1:l electrolyte. The
amount adsorbed 0 is expressed in equivalent monolayers.

-8

10

15

20

Figure 5. Comparison between profiles in the Gouy-Chapman and Debye-Hiickel approximation. Poor solvent (x = 2) x, = 5 a = l, r = 100,
otherwise as in Figure 3.

-3t I rl

Figure 8. First layer volume fraction as a function of xo,cff.Drawn


curves: polyelectrolyte-0.05 M 1: 1 electrolyte. Dashed curves: polyelectrolyte4.1 M electrolyte. x = 0.5, a = 1, 6,=
The chain length
r is indicated. For comparison, the corresponding curve for the unch,arged polymer is also given (dotted curve).

tails only

-5

I
5

10

15

2o

25

Figure 6. Segment concentration profile for 4. =


x = 0.5, xs = 4,
r = 1000, a = 1, uo = -e/lOao, a. = 1 nm2,ro = 1 nm. No electrolyte

added. SF theory.
more extended decay for CY = 0 in the S F case. However, for
polyelectrolytes, loops and tails are suppressed and then the adsorbed amounts are almost indistinguishable. The difference in
depth of the profile minimum does not measurably contribute.
Figure 5 illustrates the effect of invoking the DH approximation:
it overestimates the potentials and therefore predicts too deep
minima and too low adsorptions in the first two layers.
An interesting feature is observed at high 4. (Figure 6 ) . In
this case the contribution of the tails to the profile has a maximum
beyond the generally observed minimum. The high potentials at
a short distance from the surface tend to expel tail segments from
that region, but further away, where potentials are low, the chain
can "curl up". This phenomenon can have some relevance for
colloid stability, notwithstanding the fact that there are only few
of them (about one per ten chains) so that on a weight basis they
contribute by less than 0.3% and hence are easily overlooked.
Typical adsorption isotherms are presented in Figure 7. They
are of the high affinity type if r is not too low and, in agreement
with the conclusions arrived at above, the plateau value is then
insensitive to r.
As is the case with the adsorption of the uncharged polymers,
a certain critical value of the adsorption energy parameter xs is
needed to let the polyelectrolyte adsorb at all, but once this critical
value has been surpassed, the adsorption rises sharply with x,.

e,,:

CM-

-- ------__

x*=4

c*.10-'

0 . 7 ~
0.60.5-

---->,=4

--------

.--2s=lo-'

--_----.

a. ( c i a , )

Figure 9. Influence of the surface charge: (-) Gouy-Chapman; (---)

Debye-Hiickel. xs and salt concentrationsc, are indicated.


= :loo, ro = 1 nm, a. = 1 nm2.

x = 0.5, r

Figure 8 illustrates this by representing the volume fraction #J1


in the first layer as a function of the effective adsorption energy
parameter, defined through
(26)
Xs,eff = Xs - CYY 1
Under any condition dl is less than for uncharged polymers. The
rea.son is that the buildup of #J1 is entropically favored by the
formation of loops and tails, but with polyelectrolytes this process
is suppressed.

The Journal of Physical Chemistry, Vol. 88, No. 26, 1984

6666
1 .o

c
0

01

-3

-2

J
0

-1

log ( c , l M )

Figure 10. Adsorption of oligo- and polylysine on silver iodide as a


function of chain length and concentration of HN03. Drawn curves:
theory for x = 0.6, xs = 4.86, I#W = lo4, 01 = constant = 0.78, uo =
0.22e/a0,ro = 0.55, a. = 0.3 nm2.

The influence of the surface charge is demonstrated in Figure


9. The decrease of Oexc with uo is linear and steeper in more dilute
electrolyte. The DH approximation overestimates the generated
potentials and therefore predicts too low adsorptions. However,
if the adsorbed amounts are low (this is the case for xs = 1) the
difference between the GC and D H approximation vanishes.
Increase of c, also makes the DH approximation relatively better
because the potentials are then kept low.

Comparison with Experiment and Further Discussion


Although the present paper is primarily concerned with theory,
a brief comparison of some characteristic results shows that our
conclusions do agree well with experimental trends.
A series of experiments with oligo- and polylysine on silver
iodide and polystyrene latices have been carried out, partly in
cooperation with Dr. B. C. Bonekamp. At low pH, where the
oligo- or polypeptide is fully charged, it was found that the adsorption rises with r if r is low but becomes independent of r at
higher r, in agreement with Figures 3 and 5 . However, at high
pH the adsorbate becomes uncharged and then 19 continues to
increase with r, because then loops and tails can develop.
A similar trend has been observed with the effect of electrolytes:
the higher c,,,,, the higher the adsorbed amount and the more
sensitive it is to r. In agreement with theoretical prediction (Figure
2) electrolytes continue to have an effect even at concentrations
exceeding 0.5 M, a feature that was also observed with the adsorption of poly(styrenesu1fonate) on ~i1ica.l~More examples
of the polylysine work have been reported e l s e ~ h e r e . ~ . ~ ~
Several results of other authors may be quoted that at least
semiquantitatively support our picture. For instance, Horn found
flat adsorption for polyethyleneimines on polystyrene latex at low
pH, where the adsorbate is charged16and Williams et al. reported
the same for carboxymethyl cellulose on barium sulfate; these
authors found thicker adsorbed layers at elevated ionic strengths.
Flat adsorption for charged polyelectrolytes has also been reported
by Mabire et al.ls and by PavloviE and Miller.Ig All told, the
predicted behavior is well-documented.
Also the consequences for colloid stability are entirely in line
with the present picture. A very typical demonstration is that
at low pH, AgI sols undergo charge reversal due to the adsorption
(14) Marra, J.; van der Schee, H. A.; Fleer, G. J.; Lyklema, J. In
Adsorption from Solution; Ottewill, R. H., Rochester, C. H., Smith, A. L.,
Eds.; Academic Press: London, 1983; p 245.
(15) Bonekamp, B. C.; van der Schee, H. A.; Lyklema, J. Croat. Chem.
Acta 1983, 56, 695.
(16) Horn, D. In Polymeric Amines and Ammonium Salts;Goethals, E.
J., Ed.; Pergamon Press: Oxford, 1980; p 333.
(17) Williams, P. A,; Harrop, R.; Phillips, G. D.; Robb, I. D.; Pass, G. In
The Effect of Polymers on Dispersion Properties; Tadros, Th. F., Ed.;
Academic Press: London, 1982; p 361.
(18) Mabire, F.; Audebert, R.; Quivoron, C. J. Colloid Interface Sci. 1984,
97, 120.
(19) PavloviE, 0.;Miller, I. R. J . Polym. Sei., Part C 1971, 34, 181.

Van der Schee and Lyklema


of polylysine, but the critical coagulation concentration remains
very similar to that in the absence of polyelectrolyte, whereas at
high pH no charge reversal should occur although the critical
coagulation concentration rises sharply. In line with our picture,
charged polylysine adsorbs in flat layers and no steric stabilization
ensues, whereas uncharged polylysine does stabilize AgI sols
sterically because now loops and tails can develop.
If we accept that there is a gratifying qualitative agreement
between theory and experiment, the question rises as to how good
the quantitative accordance is. This can only be judged on the
basis of the question how realistic the parameters are that fit
experiment to theory. Before discussing this, it should be realized
that for the time being there are no satisfactory experimental data
available that may be considered discriminative. First, because
of analytical and surface area problems, experimental values of
0 are seldom better than a few tens of percents, all polyelectrolytes
studied so far are heterodisperse copolyelectrolytes and most
adsorbents are heterogeneous. Direct data on the distribution (&)
are all but absent. From the theoretical side, apart from the
inherent limitations of lattice theories, a number of other approximations have been made, of which it is difficult to assess the
quality, e.g., the substitution of the average potential for the
potential of mean force. Of others, such as the constancy of a
and t and the zero volume of the small ions and the assignment
of the charges to plates, rather than assuming a certain spatial
distribution, it was verified that they had a relatively minor influence. Against this background it is too early to seek full
quantitative agreement. However, as a general trendl3-I5 all data
and trends are explainable with our theory, using reasonable
parameter values. By way of illustration Figure 10 is included.
The values for a. and ro correspond well with the volume of the
lysyl group; further that x > 0.5 agrees with the fact that at high
pH, where polylysine is uncharged, phase separation takes place.
In conclusion, our theory predicts a behavior of polyelectrolytes
at interfaces that is a t least in qualitative and semiquantitative
agreement with the facts. Further elaboration of this theory and
designing better experiments are therefore warranted.

Acknowledgment. The authors have benefited greatly from


the help of Mr. J. M. H. M. Scheutjens with physical and
mathematical problems.
Appendix
Calculation of the Potential Profile. The potential profile is
calculated by integration of the Poisson-Boltzmann equation (9)
between the two boundary conditions at the surface and at layer
M . At each of the M planes through the centers of the lattice
layers there is a discontinuity in the field strength amounting to

Because of these discontinuitiesthe integration has to be performed


from layer to layer, with the end values of y and dyldx of the
previous layer as starting values. To this end value of dyldx the
discontinuity according to eq A1 is added. Within each layer the
integration was performed using a sixth-order Runge-Kutta
method.20p21To this end eq 9 is written as a system of two coupled
first-order differential equations

-dY1
= - dY
dx
dx

(A3

with yl defined as y and y 2 as dyldx.


The integration starts at y Mand dy,/dx, which are interrelated
by eq 20. The value dy,/dx must obey the boundary condition
(20) Zonneveld, J. A. Automatic Numerical Integration: Mathematisch
Centrum: Amsterdam, 1964; Mathematical Centre Tracts 8.
(21) Gear, C. W. Numerical Initial Value Problems in Ordinary Differential Equations; Prentice-Hall: Englewood Cliffs, NJ, 1971.
(22) Stigter, D. J. Colloid Interface Sci. 1975, 53, 296.

J. Phys. Chem. 1984,88, 6667-6670


at the surface. To achieve this the value of yu is iterated with
a Newton-type iteration. After each integration step across a
lattice layer, the potential must be examined and, occasionally,
the Newton-step length must be reduced to prevent overflows.
Calculation ofthe Polymer Profile. Using the Scheutjens-Fleer
theory, we followed their line in the evaluation of the polymer

6667

concentration profile, except for the modification that we used


In +01 as the iteration variable to find the zero of +! +O,-l, where
+I is obtained from 40i using (5). As eq 29 of Roes paper shows
the derivatives of the partition function with respect to &, a term
-abi - p)has to be added to the left-hand side of this equation,
which is solved, using a Newton-iteration method.

Single Molecule Gas-Phase Polymerization Kinetics of Vinyl Acetate. Nonsteady


Measurements
H.Reiss* and M. A. Chowdhury
Department of Chemistry and Biochemistry, University of California, Los Angeles, California 90024
(Received: July 23, 1984)

This paper continues our recent work on the kinetics of free-radical chain polymerization in the gas phase. As in the previous
work, we avoid having the involatile polymers condense out of the vapor phase by employing a cloud chamber technique
which allows us to study the kinetics under conditions such that there are so few polymers growing simultaneously that they
cannot encounter one another. We detect the arrival of individual product molecules by having them nucleate macroscopic
drops of monomer liquid. The arrival of product molecules is so discrete that signal averaging must be employed in the
determination of the rate. As in the previous work, the polymer is poly(viny1 acetate). In the present study we employ a
nonsteady technique in which the reaction is initiated by a pulse of photons. Among other things, this approach makes it
possible to determine the degree of polymerization of the polymer capable of participating, along with monomer molecules,
in the binary nucleation process and therefore makes it unnecessary to rely on a difficult nucleation theory. The monomer
supersaturation has been deliberatelychosen to be high, so that high enough rates could be achieved to allow the signal averaging
to be performed by hand. As a result, the polymers responsible for nucleation are small, containing only about six monomer
units.

1. Introduction

In a recent paper a method was introduced, capable of


measuring the kinetics of gas-phase chain polymerization reactions
by detecting the arrival of single polymer molecules. This method
involved the use of an upward thermal diffusion cloud chamber
which maintained monomer vapor in a steady, supersaturated state.
Ultraviolet photons, admitted to the chamber, produced free
radicals and initiated chain polymerization. The polymer chains
grew to a critical size at which they were capable of participating
in condensation nuclei leading to the formation of drops of liquid
monomer. The drops fell through a laser beam, scattered strong
light signals, and were counted. The rate of drop formation was
identical with the rate of production of polymers of the critical
size and was therefore a measure of the kinetics of polymerization.
Reference 1 contains a detailed verification of the validity of
this method and its underlying mechanism and should be consulted
for this purpose.
The critical degree of polymerization (necessary for participation
in a condensation nucleus) depends on the degree of supersaturation of the monomer vapor. By adjusting the supersaturation,
it is possible to tune to the arrival of polymer molecules of a
specified size. The series of events, outlined above, will only occur
faithfully in the manner described, if conditions can be arranged
such that each condensation nucleus contains only one polymer
molecule. Thus, the number of chains growing simultaneously
must be so small that there is little probability of two polymer
molecules encountering one another. This, in fact, is the reason
for hoping that the gas-phase process can be sustained. In other
attempts to study such gas-phase kinetics it has not been possible
to avoid polymer-polymer encounters and abundant condensation
of involatile small polymer molecules. The enormous sensitivity
of the nucleation phenomenon enables one to detect the arrival
( 1 ) Chowdhury, M. A.; Reiss, H.; Squire, D. R.; Stannett, V. Macromolecules 1984,17, 1436.

0022-3654/84/2088-6667$01.50/0

of single molecules and, therefore, to work under conditions (if


they can be established) such that growing polymers cannot encounter one another and escape the vapor phase.
In ref 1 emphasis was placed on ( 1 ) demonstrating that the
observed droplet formation was due to the polymer process just
described and (2) showing that conditions could be established
such that only the single polymer condensation process occurred.
As a somewhat secondary goal, the measured data were used to
estimate the various kinetic constants. As explained in ref 1, the
study was a bootstrap one, in which confirmation of the postulated phenomenon depended on the compatibility of experimental
observations with one another, as well as with theory. This
bootstrap character was mandated by the fact that no other
method of comparable sensitivity was available to provide independent confirmation of the underlying process.
When referred to these goals, the studies of ref 1 may be
considered successful. The monomer involved was vinyl acetate,
and with this monomer, the results of ref 1 did indicate that
conditions for single polymer nucleation had been established.
Furthermore, the above-mentioned compatibilities of experiment
with experiment and of experiment with theory were observed.
The experiments of ref 1 were conducted under steady illumination
(under a steady flux of UV photons). Among other things, the
critical polymer size had to be deduced from an approximate
theory of nucleation.2 The accuracy of that theory could be
assessed for the case of homogeneous nucleation of the monomer
i t ~ e l f but
, ~ one cannot be assured of its quantitative validity in
the case of binary nucleation involving both monomer and polymer.
Nevertheless, its application indicated that, in ref 1, polymers
having degrees of polymerization of the order of 30 were responsible for nucleation. As explained in ref 1 , many improvements can be effected in subsequent studies. Even the degrees
(2) Reiss, H.; Chowdhury, M. A. J . Phys. Chem. 1983,87,4599
(3)Chowdhury, M.A. J . Chem. Phys. 1984,80,4569.

0 1984 American Chemical Society

You might also like