Sei sulla pagina 1di 8

Article

pubs.acs.org/Macromolecules

Collapse-to-Swelling Transitions in pH- and Thermoresponsive


Microgels in Aqueous Dispersions: The Thermodynamic Theory
Alexey A. Polotsky,† Felix A. Plamper,‡ and Oleg V. Borisov†,§,⊥,*

Institute of Macromolecular Compounds of the Russian Academy of Sciences, 31 Bolshoy pr., 199004 St.-Petersburg, Russia

Physikalische Chemie II, RWTH Aachen University, 52056 Aachen, Germany
§
St.Petersburg National Research University of Information Technologies, Mechanics and Optics, 197101, Kronverkskiy pr., 49,
St.Petersburg, Russia

Institut des Sciences Analytiques et de Physico-Chimie pour l’Environnement et les Matériaux, UMR 5254 CNRS/UPPA, Pau,
France

ABSTRACT: We present a theory of a conformational


collapse-to-swelling transition that occurs in aqueous dis-
persions of multiresponsive (pH- and thermoresponsive)
microgels upon variation of ionic strength, temperature, or
pH. Our theory is based on osmotic balance arguments and
explicitly accounts for ionization equilibrium inside microgel
partices. The theory predicts complex patterns in the
dependence of the microgel particle dimensions on the control parameters: An increase in temperature leads to worsening of
the solvent quality for the gel forming LCST-polymers and to concomitant decrease in the dimensions of the gel particles. This
collapse of the gel particles provoked by an increase in temperature occurs either smoothly (at high or low ionic strength), or
may exhibit a jump-wise character at intermediate ionic strength. The theory further predicts that the degree of swelling of
microgel particles varies nonmonotonously and exhibits a maximum as a function of salt concentration at a pH close to the pK.
This nonmonotonous variation of the particle dimensions occurs continuously at temperatures below or slightly above LCST
(good or marginal poor solvent strength conditions, respectively), whereas at higher temperatures the jump-wise swelling of the
gel particles is followed by either continuous or jump-wise collapse induced by progressive increase in the salt concentration. A
decrease/increase in pH leads to deswelling of the weak polyacid/polybase gel particles, which occurs smoothly at temperatures
below LCST, but may exhibit a discontinuity above LCST. These theoretical predictions can be used for design of smart stimuli-
responsive microgels.

I. INTRODUCTION circumstances are known where the macroscopic size of gels is


The ability of ionic polymers (polyelectrolytes) to change their detrimental. E.g., the characteristic switching time is propor-
conformations under the action of external triggers, such as tional to the surface of the gel.5 Therefore, gels in the
ionic strength and pH, is exploited in many applications ranging nanometer to micrometer scale can provide beneficial switching
from oil recovery and paper-making to nanotechnology and kinetics and rapid uptake/release of guest molecules. These gels
nanomedicine. A remarkable progress in understanding the are termed as microgels6−8 and will be regarded within this
properties of linear1,2 and branched3 polyelectrolytes has been investigation.
achieved in the past few years. In presence of additional attractive (solvophobic) inter-
There are important distinctions in properties of strong actions, discontinuous volume transitions in the ionic hydrogels
(quenched) and weak (annealing) polyelectrolytes. In a strong can appear. This means that large effects in the swelling can be
polyelectrolyte the fraction α of charged monomer units is fixed observed when only small changes in the solvent strength are
chemically and does not depend on the environmental applied.9−11 The ion pair formation in less polar medium may
conditions. In a weak polyelectrolyte the fraction α of charged provide additional driving force for the discontinuous gel
monomer units coincides with the degree of ionization and is collapse.12−15
governed by the local pH which may noticeably differ from the
These examples indicate that polyelectrolytes and the gels
pH of the buffer.4 The latter is especially true for branched
polyelectrolytes3 and ionic gels. thereof are interesting species. A subtle balance between
Hereby, polyelectrolyte gels can be regarded as a special class attracting and repellant interactions appears to offer this rich
of branched ionic polymers. Because of osmotic pressure of the behavior. Therefore, the use of weak polyelectrolytes with
confined counterions, they exhibit often tremendous swelling in
solvents of high dielectric constant like water, while still having Received: July 4, 2013
a defined macroscopic shape. The swelling is especially Revised: September 18, 2013
pronounced when it comes to highly charged gels. Generally, Published: October 24, 2013

© 2013 American Chemical Society 8702 dx.doi.org/10.1021/ma401402e | Macromolecules 2013, 46, 8702−8709
Macromolecules Article

tunable charge and hydrophilicity might envision further occurring in pH-sensitive polyelectrolyte brushes upon
stimuli-responsiveness for future applications. variation of salt concentration. Furthermore, combined effects
There are literature reports on the stimuli-responsive of added salt and tunable hydrophobicity of the monomer units
behavior of the microgels formed by poly(acrylic acid) (PAA) on conformations of pH-sensitive star-branched polyelectro-
or different kinds of copolymers including monomer units of lytes were recently investigated.39,40 It was demonstrated, that
acrylic acid.16,17 The swelling behavior of these weak salt-induced re-entrant swelling-to-collapse intramolecular
polyelectrolyte microgel particles can be efficiently controlled conformational transitions in star-branched hydrophobic
by pH: a decrease in pH causes a decrease in the degree of polyions occur continuously either at high or at low ionic
ionization of the microgel particles, drop in the intragel osmotic strength but may exhibit features of a first-order phase
pressure of the counterions and concomitant collapse of the transition (i.e., intramolecular microphase separation) at
microgel. intermediate salt concentration.
In contrast to pure PAA, poly(N,N-dimethylaminoethyl The aim of the present paper is to investigate volume
methacrylate) (PDMAEMA) and especially copolymers of N- transitions occurring in dilute aqueous dispersions of multi-
isopropylacrylamide (NIPAM) and its derivatives exhibit responsive (pH- and thermoresponsive) microgel particles
additional thermosensitive features.18−21 That is, the polymer upon variation of external parameters, such as temperature,
is soluble in water under ambient conditions but the solubility ionic strength and pH. The latter two affect the strength of
decreases upon an increase in temperature. Hence, aqueous Coulomb interactions between charged monomer units,
solutions of such polymers exhibit lower critical solution whereas temperature controls the strength of short-range
temperature (LCST): Below the LCST the polymer is soluble excluded volume interactions (solubility of the polymer in the
at any concentration, while above LCST macroscopic phase absence of charges). In addition, hydrostatic pressure might be
separation into dilute and concentrated phases (precipitate) considered as additional parameter, which affects the strength
may occur.22 Regarding microgels, collapse of the microgel of excluded volume interactions and thus solubility of the
particles dispersed in aqueous media occurs upon an increase in monomer units of the microgel in water.41−43 As we
temperature above LCST which can be termed in this case as demonstrate below, a competition of short-range attractive
volume phase transition temparature (VPTT).6 Remarkably, and long-range repulsive interactions with ionization equili-
the microgel dispersions may exhibit colloidal stability even at brium inside the gel leads to complex patterns in the
temperature above VPTT. dependence of the microgel particles dimensions on these
We have already revealed23,24 an interplay between three control parameters even in the absence of any polymer-
protonation and thermoresponsive behavior in case of linear or ion-specific interactions in the system. Hence, our
and branched PDMAEMA. Dependent on the pH, PDMAEMA conclusions are relevant for a wide class of thermosensitive
turns water-insoluble upon heating. Furthermore, microgels of weak polyelectrolyte microgels, and enables one to rationalize
PDMAEMA show an excellent pH-responsiveness. Charged stimuli-responsive behavior of e.g. microgels of PNIPAM with
thermosensitive microgels of PDMAEMA have been pre- pH-sensitive ionic comonomers or PDMAEMA in aqueous
pared.25,26 Otherwise, cationic or anionic monomers can be dispersions.
copolymerized with monomers (like NIPAM) yielding
thermosensitive polymers.18,19,27 For both weak and strong II. MODEL AND FREE ENERGY
polyelectrolytes, the swelling of the gel particles can be Our approach is based on the analysis of the Gibbs free energy
manipulated by changing the strength of electrostatic of a swollen microgel particle (calculated per subchain) which
interactions by, e.g., addition of salt ions.28 For weak can be presented in a mean-field approximation as
polyelectrolytes pH can be considered as additional control
parameter since its variation enables a tuning of the charge of F = Fconf + Fex . vol + Fion (1)
the gel particles. In addition, issues of colloidal stability of the
The first term in eq 1 accounts for conformational free energy
microgel dispersion might arise.29
of a stretched subchain comprising N monomer units and
Concluding, the combination of pH and thermosensitive
extended up to the size R
properties of, e.g., PDMAEMA makes the stimuli-responsive
behavior of the microgel more sophisticated, especially when it 3R2
comes to the salt dependence. Indeed, such complex patterns as Fconf /kBT =
re-entrant swelling-to-collapse transitions provoked by increas- 2Nb2 (2)
ing salt concentration were predicted theoretically30,31 and where kB is the Boltzmann constant and T is the temperature.
observed experimentally32−34 for pH-sensitive polyelectrolyte Here we assume that subchains are extended with respect to
brushes and for star-branched polyelectrolytes.35,36 The non- ideal (Gaussian) dimensions, exhibit Gaussian elasticity and are
monotonous salt dependence of the degree of swelling of pH- intrinsically flexible, that is, the statistical segment length is on
sensitive branched polyions (or polyelectrolyte brushes) was the order of the monomer unit length b. The condition of the
explained by dual effect of added salt on the differential subchain extension, R ≥ bN1/2, may be violated in the microgel
intramolecular osmotic pressure: the initial increase in salt particles collapsed in poor solvent. The corresponding
concentration provokes an increase in the osmotic pressure due correction term which accounts for conformational entropy
increasing ionization of the monomer units, whereas further losses in contracted subchains could be introduced following
increase in salt concentration leads to a leveling off of the the lines of ref 44. This correction, however, is negligible
concentrations of small ions inside the star (or in the brush) compared to other contributions to the free energy in the
and in the solution and to a concomitant decrease in the collapsed state and does not qualitatively affect the results of
differential osmotic pressure. More recent studies based on our analysis.
more advanced theoretical approaches37,38 have enabled a The subchain dimension R is related to the polymer volume
detailed analysis on the internal structural rearrangements fraction c in the microgel as
8703 dx.doi.org/10.1021/ma401402e | Macromolecules 2013, 46, 8702−8709
Macromolecules Article

qNb3 QNb3 larger than its characteristic size V1/3 reduced by the Bjerrum
c= = length lB = e2/ϵkBT (here ε is the dielectric constant of the
2R3 V (3)
medium). Since for the microgel particle V scales proportion-
where q is the number of subchains emanating from a junction ally to Q and the dimensions of a subchain in the low salt limit
point, Q is the number of subchains and V is the volume of the can be estimated as ∼ α1/2Nb, the microgel particles comprising
microgel particle. Q ≫ (α−1/2b/lB)3/2 subchains can be considered as almost
The free energy of nonelectrostatic (excluded volume) electroneutral. Hence, the local electroneutrality condition
interactions can be presented in the virial approximation as applies with a good accuracy to sufficiently large microgel
particles, with the typical particle size on the order of 100 nm
Fex . vol /kBT = N (vc + wc 2) (4) or larger. It is essential, however, that volume fraction occupied
where v and w are the second and the third virial coefficients, by the gel particles in the solution is sufficiently small to ensure
respectively. The second virial coefficient (the excluded volume quasi-constant values of chemical potentials of all mobile ion
parameter) v can be tuned by variation in the temperature. For species.
most of the water-soluble polymers (including PNIPAM and The degree of ionization α of subchains in the microgel
PDMAEMA), v is a decreasing function of the temperature: depends on the local pH inside the gel, which in turn depends
vT≤LCST ≥ 0 and vT≥LCST ≤ 0 correspond to good and poor on the pH in the buffer and on the excess local electrostatic
solvent conditions, respectively, whereas vT=LCST = 0 corre- potential in the gel. By combining the mass action law with the
sponds to the Θ-point. In the vicinity of LCST the second virial local electroneutrality condition the degree of ionization α of
coefficient is expected to vary approximately linearly as a the gel can be expressed as3
function of temperature, v ∼ (LCST − T). An increase in ⎛ αc ⎞ 2
pressure at constant temperature is known to improve solubility α 1 − αb αc
= 1+⎜ ⎟ −
of polymer in water,41 that is, v should be an increasing function 1 − α αb ⎝ cs ⎠ cs (6)
of pressure. On the contrary, the third virial coefficient is
virtually independent of temperature, w = const. It is known, The degree of ionization αb of the respective nonpolymeric
that solubility of nonionic polymers, like e.g. PNIPAM in water ionizable groups in bulk solution is related to pH of the buffer
is affected by salt concentration,45 that is, the second virial solution as
coefficient v may exhibit a dependence on the ionic strength. In
αb = (1 + 10±(pH − pK ))−1 (7)
our analysis we, however, disregard this dependence because of
the primary effect of salt concentration on the strength of ionic where pK = −log K, and K is the ionization constant of a single
interactions acting between charged monomer units of a pH- ionizable group and signs “+” and “−” refer to cationic and
sensitive microgel. We also disregard possible dependence of v anionic microgels, respectively.
on the pH of the solution, which may take place in the case of, It is convenient to characterize the extent of swelling of the
e.g., hydrogen bond formation between nonionized monomer microgel particles by the dimensionless parameter
units.
The ionic contribution to the free energy is specified in the R ⎛ c ⎞1/3
= ⎜ θ ⎟ ≡ β1/3
framework of the local electroneutrality approximation, i.e., by Rθ ⎝c⎠
assuming that the charge density of ionized monomer units
inside the gel is locally compensated by excess density of where
mobile counterions. With the account of equilibrium exchange
R θ = N1/2bw1/8q1/4 2−1/8
of mobile ions between microgel volume and the surrounding
solution (where concentrations of all mobile ion species are and
assumed to be fixed), the ionic contribution to the Gibbs free
energy is specified as3 (q/2)1/4
cθ =
⎧ ⎡ ⎤⎫ N1/2(2w)3/8
⎪ ⎛ αc ⎞ 2 ⎪
Fion(c)/kBT = N ⎨ln(1 − α) − ⎢ 1 + ⎜ ⎟ − 1⎥⎥⎬
cs ⎢
are respectively the subchain dimensions and the volume
⎪ c
⎣ ⎝ cs ⎠ ⎦⎪ fraction of polymer in the microgel at the Θ-temperature (i.e.,
⎩ ⎭
at v = 0) in the absence of ionic interactions. Assuming that in a
(5) dry state the volume fraction of polymer in the microgel equals
where cs is the overall concentration (volume fraction) of unity, we can express the degree of swelling as R/Rdry = β1/3·
monovalent ion species (including H+ and OH− ions and added N1/6(2w)1/8(q/2)−1/12.
monovalent salt) in the solution. Below we characterize the temperature-controlled solvent
The local electroneutrality approximation is justified for strength using the dimensionless parameter τ/cθ, where τ = −v/
dilute dispersions of ionic microgels with or without added 2w. An increase in temperature leads to a decrease in v and to
electrolyte provided that the size of the gel particles is an increase in τ. Under poor solvent conditions, v ≤ 0, τ
sufficiently large to ensure that the major fraction of mobile coincides with the volume fraction of polymer in the precipitate
counterions is retained by the Coulomb attraction inside the (or in a large polymer globule48).
microgel particles. The solution of the corresponding Poisson− The average number N of monomer units in a subchain is
Boltzmann equation for a branched polyion (e.g., a microgel inversely proportional to the degree of cross-linking and can be
particle) placed into an electroneutral Wigner−Seitz cell46,47 considered as a variable architectural parameter of the gel. An
proves, that with the accuracy of logarithmic factors, the increase in N is manifested in a decrease in cθ. Below we shall
elecrtroneutrality condition is fulfilled if the bare charge of the analyze how the pH-/thermo-responsive properties of the
microgel particle, αNQ, (measured in elementary charges e) is microgels depend on the degree of cross-linking.
8704 dx.doi.org/10.1021/ma401402e | Macromolecules 2013, 46, 8702−8709
Macromolecules Article

Figure 1. Dependences of the reduced size of the microgel particle R/Rθ (a) and degree of ionization α (b) on reduced temperature τ/cθ for N = 20,
q = 4, αb = 0.6 and different salt concentrations.

The equilibrium degree of swelling of the microgel particle contraction of the gel occurs continuously upon a decrease in
with respect to the Θ-dimensions, R/Rθ  β1/3 obeys the the solvent strength. For weak (pH-sensitive) polyelectrolyte
following equation, which results from minimization of the gels, the degree of ionization of the subchains depends on the
Gibbs free energy determined by eqs 1, 4, 5 and 6: polymer concentration inside the gel and on the ionic strength
⎡ ⎤ in the surrounding solution, eq 6. Therefore, the temperature-
8/3 τ cs 3⎢ ⎛ αc θ ⎞ 2 ⎥ induced collapse transition in such microgels acquires novel
β + β− β 1+⎜ ⎟ − 1⎥ = 1
cθ 2wcθ3 ⎢⎣ ⎝ csβ ⎠ ⎦
features.
(8) Figure 1a illustrates how the character of the temperature-
induced collapse transition changes upon variations in salt
where the degree of ionization α is given by
concentration cs. The microgel gets contracted continuously as
⎛ αc ⎞ 2 a function of decreasing solvent strength (increasing temper-
α 1 − αb αc
= 1 + ⎜ θ⎟ − θ ature) at low or high salt concentrations, but exhibits a jump-
1 − α αb ⎝ csβ ⎠ csβ (9) wise collapse below the Θ-point (above LCST) in the
intermediate range of salt concentrations. This discontinuity
The free energy of the microgel (per subchain) can be is also reflected in the temperature dependences of the degree
expressed as of ionization, presented in Figure 1b. At any salt concentration
F cs an increase in temperature (in τ) leads to compactization of
2wcθ 2NkBT the microgel particles and thus provokes a concomitant
decrease in degree of ionization α. Indeed, α is a monotonously
⎡ ⎤
5 2/3 1 cs ⎢ ⎛ αc θ ⎞ 2 ⎥
decreasing function of polymer concentration, eq 6, since
= β − − β 1 + ⎜ ⎟ − 1 repulsive electrostatic interactions of charged monomer units
2 2β 2 wcθ 3 ⎢⎣ ⎝ csβ ⎠ ⎥
⎦ hinder ionization. Therefore, the degree of ionization in the
(10)
collapsed state is much smaller than that in the swollen state.
In the first order transition point the polymer density and the In Figure 2 the critical value of the bulk degree of ionization,
ionization degree in the coexisting collapsed and swollen states αb,crit(cs), is presented as a function of salt concentration. At
(denoted as state 1 and state 2) are determined from the each given salt concentration cs the temperature-induced
condition contraction of the microgel particles occurs continuously, if
F(β1 , α1) = F(β2 , α2) (11) the bulk degree of ionization αb ≤ αb,crit(cs). On the contrary, at

solved together with eqs 8, 9. The critical line in (αb, cs) plane
separating the ranges of continuous and discontinuous
temperature induced collapse transitions is obtained from the
condition
dτ(β)/dβ = d2τ(β)/dβ 2 = 0 (12)

III. RESULTS
A. Temperature-Induced Swelling-to-Collapse Tran-
sition. An increase in temperature leads to a decrease in
solubility (enhancing hydrophobicity) of polymer chains
forming the microgel and, as a result, to deswelling of the gel
particles. This deswelling is, however, opposed by excess
osmotic pressure of the mobile ions inside the gel. For a
microgel with permanent charges (strong polyelectrolytes) in
salt-free solution the competition of increasingly strong short- Figure 2. Critical value of the bulk degree of ionization αb,crit vs salt
range monomer−monomer attractions with the osmotic concentration calculated for different values of N, as indicated in the
pressure of confined counterions results in a jump-wise collapse Figure. At αb ≥ αb,crit the temperature-induced collapse of the microgel
transition, whereas at sufficiently high ionic strength the occurs as a jump-wise first order phase transition.

8705 dx.doi.org/10.1021/ma401402e | Macromolecules 2013, 46, 8702−8709


Macromolecules Article

Figure 3. Dependences of reduced size of the microgel particle R/Rθ (a) and degree of ionization α (b) on reduced salt concentration cs/cθ for N =
20, q = 4, αb = 0.6 and different temperatures. Solid points (circles) and dashed lines indicate the position of the first order phase transition in the
cases τ/cθ = 2, 3, and 3.25.

αb ≥ αb,crit(cs) an increase in temperature triggers a jump-wise strated by Figure 3a, a continuous increase in the microgel
collapse of the microgel. The critical value of αb,crit(cs) depends particle size caused by addition of small amounts of salt can be
nonmonotonously on salt concentration cs, that is, decreases at interrupted by a jump-wise swelling transition. Upon further
low salt concentrations, passes through a minimum and then increase in salt concentration the gel shrinks. Depending on
increases at high salt concentrations. Hence, at fixed pH (fixed temperature, this contraction either occurs continuously or
αb) an increase in temperature provokes progressive deswelling involves a jump-wise collapse transition.
of the microgel particles either at small or at high salt As one can see in Figure 3b, the degree of ionization α of the
concentration, whereas discontinuous temperature-induced gel particles is a continuous and smoothly increasing function
volume transition can be expected in a certain range of of salt concentration under good or nearly Θ-solvent
intermediate salt concentrations. conditions, T ≈ LCST, as well as under very poor solvent
Remarkably, a decrease in N (equivalent to an increase in the conditions. Under moderately poor solvent conditions the
degree of cross-linking) leads to shrinkage of the region of degree of ionization exhibits discontinuities which are coupled
jump-wise collapse transition toward higher values of αb and to to abrupt swelling and collapse of the gel particles upon an
a decrease in the magnitude of the jump in the size of the gel increase in salt concentration. Moreover, in a narrow range of
particles. At given (buffered) value of pH, the range of salt temperatures, the degree of ionization exhibits a weak
concentrations where discontinuous collapse may be observed maximum as a function of the ionic strength. These peculiarities
experimentally narrows upon a decrease in N and for strongly in the dependence of the degree of ionization of the gel
cross-linked gel temperature-induced deswelling is expected to particles on salt concentration arise due to coupling between
occur smoothly at any salt concentration. ionization and local polymer density, that is expressed by eq 6.
B. Salt-Induced Swelling-to-Collapse Transition. The The phase diagram of the microgel solution in (temperature,
conformational response of the microgel particles to the salt concentration) coordinates is presented in Figure 4. The
variation in the ionic strength of the solution at fixed
temperature and pH is more complex: The initial increase in
the salt concentration promotes ionization and, consequently,
provokes an increase in the osmotic pressure inside the gel. As a
result, swelling of the microgel particles occurs. As soon at the
degree of ionization of the microgel reaches (upon an increase
in salt concentration) its maximal possible value α ≈ αb, (at
given pH in the buffer) then a decrease in the differential
osmotic pressure caused by added salt governs deswelling of the
gel. As a result, the dependence of the size of the gel particles
on salt concentration acquires nonmonotonous character. A
similar effect has been predicted earlier for pH-sensitive
polyelectrolyte brushes30,31 and stars36 where stretching of
the brush-forming chains or of the star arms is governed by the Figure 4. Phase diagram of the microgel solution in the temperature−
salt concentration coordinates calculated for αb = 0.6 and different
excess osmotic pressure exerted by entrapped counterions. degrees of cross-linking (the values of N indicated at the lines). The
The dependence of the reduced size of the microgel particle lines are terminated by the critical points.
on salt concentration cs at different temperatures is presented in
Figure 3. As one can see from the Figure, under good solvent
conditions (T ≤ LCST) as well as in the vicinity of the Θ-point τtr(cs) lines in the Figure 4 correspond to the conditions of
(T ≈ LCST) the nonmonotonous but continuous variation in coexistence between collapsed and swollen states for the
the dimensions of the gel (swelling-contraction) occurs upon microgels with different degrees of cross-linking. Crossing of
an increase in salt concentration. On the contrary, under the coexistence lines upon variation in temperature (at constant
moderately poor solvent conditions (T ≥ LCST) one or two ionic strength) or upon variation in the ionic strength (at
sharp conformational transitions accompanied by abrupt constant temperature) implies jump-wise transitions between
changes in the dimensions of the microgel particles occur swollen and collapsed states of the microgel. Each of the first-
upon progressive increase in salt concentration. As demon- order transition lines ends in two critical points corresponding
8706 dx.doi.org/10.1021/ma401402e | Macromolecules 2013, 46, 8702−8709
Macromolecules Article

Figure 5. Dependences of the reduced size of the microgel particles R/Rθ (a) and degree of ionization α (b) on the pH-controlled bulk degree of
ionization αb for N = 20, cs/cθ = 0.1 and different temperatures, as indicated at the curves. The dotted line in panel b corresponds to α = αb. Solid
points (circles) and dashed lines indicate the position of the pH-induced first order phase transition in the cases τ/cθ = 3 and 4.

to the lower and the upper critical salt concentrations. The pH-triggered transition from collapsed to swollen states occurs
temperature-induced contraction of the gel particles occurs with a jump in the gel size, as the first order phase transition
continuously, without a jump in the size, if the salt and it is accompanied by a concomitant jump-wise increase in
concentration in solution is kept below the lower or above the degree of ionization α.
the upper critical salt concentration. The difference between
the former and the latter diminishes upon an increase in the IV. DISCUSSION AND CONCLUSIONS
degree of cross-linking.
Remarkably, for weakly cross-linked microgels the depend- The presented above theory suggests that aqueous dispersions
ence of the transition temperature τtr on salt concentration cs is of multiresponsive (pH- and thermoresponsive) microgels
a nonmonotonous function, that is, exhibits a maximum. The exhibit a complex response to variations in temperature, ionic
presence of the maximum in the τtr(cs) dependence, Figure 4, strength of the solution and pH. Our conclusions are based on
for weakly cross-linked microgels, indicates that in a certain the thermodynamic analysis of tunable balance between short-
temperature range above LCST a progressive increase in the range solvophilic/solvophobic interactions and excess osmotic
salt concentration provokes two successive jump-wise con- pressure exerted by the ions partitioned between the microgel
formational transitions in microgel particles: The first one and the solution with the account of ionization equilibrium for
(which occurs at lower salt concentration) is the transition the monomer units of the gel-forming subchains. Assuming that
from the collapsed weakly ionized state to the swollen strongly inside the microgel the charge of the ionized monomer units is
ionized state, whereas the second one (which occurs at higher approximately compensated by excess concentration of mobile
salt concentration) is the transition from the swollen to the counterions, we were able to derive analytical equation for the
collapsed state of the gel with concomitant drop in the degree degree of swelling of the microgel as a function of external
of ionization α. (For strongly cross-linked microgels the latter parameters (solvent strength, salt concentration, and pH in the
transition is continuous). buffer).
C. pH-Induced Collapse-to-Swelling Transition. The Analysis of the temperature dependence of the degree of
swelling of pH-sensitive microgel particles can be naturally swelling indicates that an increase in temperature (decrease in
governed by changing pH of the solution that affects the degree the solvent strength) triggers the gel contraction. This
of ionization of the microgel particles. As follows from eq 6, the contraction occurs continuously at low and high ionic strength
degree of ionization α of the microgel increases/decreases whereas at intermediate salt concentration a jump-wise
monotonously as a function of the bulk pH for anionic/cationic temperature-induced shrinkage of the microgel particles is
gel. Importantly, at any pH in the buffer, α remains smaller than expected. The range of salt concentration corresponding to the
αb which is directly related to the buffer pH by eq 7. discontinuous temperature-induced collapse of the gel becomes
Figure 5b demonstrates that under good solvent conditions wider upon a decrease in the degree of cross-linking (an
(T ≤ LCST) the degree of ionization α in a swollen microgel increase in N).
closely follows increasing αb, which is reflected in continuous Furthermore, similarly to pH-sensitive polyelectrolyte
increase in the size of the gel particles, Figure 5a. The difference brushes30,31 or star-branched polyelectrolytes,35,36 the microgel
between α and αb becomes more pronounced upon an increase particles are expected to exhibit nonmonotonous dependence
in temperature, particularly above LCST. The higher the of their dimensions on salt concentration: an increase in the
temperature, the larger the polymer concentration in the microgel size at low ionic strength is followed by a decrease at
microgel, the more significant is the deviation of α in the high ionic strength. The appearance of a maximum in the
collapsed (at small αb) gel from αb. On the contrary, the particle size dependence on salt concentration is explained by
difference between α and αb in the swollen gel is small at any the fact that the ionization inside the microgel is suppressed at
temperature at large αb. low ionic strength but grows upon moderate increase in salt
The swelling of the microgel particles caused by an increase concentration. The theory-predicted variation of the microgel
in αb (governed by the pH) occurs smoothly at temperatures particle size as a function of salt concentration is smooth under
below and around LCST (good, Θ, and marginal poor solvent good, Θ, or marginal poor solvent conditions (below or around
conditions). Upon an increase in temperature the pH-triggered LCST), whereas under moderately poor solvent conditions
swelling transition becomes more sharp (occurs in a narrower (above LCST) successive swelling and collapse of the microgel
pH range). At high temperature (sufficiently above LCST) the may occur as a jump-wise first-order phase transition.
8707 dx.doi.org/10.1021/ma401402e | Macromolecules 2013, 46, 8702−8709
Macromolecules Article

The universality of predictions of our theory concerning concentration results in progressive repartitioning of the chains
complex behavior of the microgel upon variation in temper- between the two phases and concomitant smooth variation of
ature or/and salt concentration and character of the stimuli- the experimentally measurable dimensions of the stars or
induced volume transitions is assured by the account of only brushes. On the contrary, under the same conditions the cross-
main thermodynamic forces (contributions to the free energy) linked pH- and thermoresponsive microgel particles are
acting in the system. Additionally, ion binding to the chains, expected to exhibit discontinuous volume transitions man-
variation in local dielectric constant upon gel collapse, etc., may ifested in abrupt variation in the particles dimensions and in
influence quantitatively the shape of the swelling curves. coexistence of collapsed and swollen microgel particles in
However, corresponding analysis requires application of more solution in the vicinity of the transition point.
sophisticated numerical schemes for the free energy mini- Finally, we remark that an interplay of short-range
mization60 and involves unavoidably empirical adjustable monomer−monomer attraction with long-range Coulomb
parameters. At the same time, phenomenology of any particular repulsion may result (at low ion concentration inside the gel)
experimental system can be even richer due to polymer- or ion- in formation of nanoscale collapsed hydrophobic domains
specific effects possibly coming into play. For example, similarly along subchains,64 association of subchains into bundles,65,66 or
to polyelectrolyte brushes and stars, reswelling of cationic other types of the domain structures.67,68 As has been discussed
hydrogels collapsed at moderate salt concentration may occur in ref 66, because of branched topology, multiple morphologies
at very high ionic strength (more than one mol per liter) if I− or of microdomain structures may appear in partially collapsed
Br− are used as counterions.61,62 This behavior was explained hydrophobic polyions. The formation of locally collapsed
theoretically using modified Donnan theory in terms of volume hydrophobic domains may lead to smoothening of abrupt first-
of the ions in ref 63, but it is not related to the pH-sensitive order volume transitions in hydrogels. A more refined analysis
properties of the gel, i.e., could be expected for strong of these effects based on Monte Carlo or molecular dynamics
polyelectrolyte gels as well. The comprehensive analysis of such simulations is necessary.
ion-specific effects is, however, beyond capabilities of the mean- The predictions of our theory can be checked, e.g., by
field theoretical approaches. dynamic light scattering or viscosimetry in dilute aqueous
Furthermore, in the present study we focused only on the dispersion of multiresponsive microgels. Indeed, the aqueous
influence of monovalent salt on the swelling of the microgels. It dispersions of the microgels often convey colloidal stability (or
is known that multivalent counterions provoke much stronger exhibit very slow coagulation) even in the collapsed state due to
collapse of polyelectrolyte brushes or branched polyelectrolytes residual charges of the particles. In addition, the particle
as compared to that induced by monovalent salt ions.3,49−52 concentration can be low, while still allowing efficient light
Similar effects of the ion valence should be expected for scattering.
microgels as well and can be qualitatively explained by
preferential localization and lower osmotic pressure exerted
by multivalent ions inside polyelectrolyte brushes or stars (or,
■ AUTHOR INFORMATION
Corresponding Author
equally, in microgels). At the same time, in the case of *E-mail: oleg.borisov@univ-pau.fr.
multivalent counterions quantitative analysis points to the
importance of ion−ion or/and ion−polymer correlations, Notes
which leads to stronger collapse as compared to prediction of The authors declare no competing financial interest.
the theory based on the osmotic balance arguments.50,52 A
quantitative analysis of the ion−ion correlations in strongly
charged polyelectrolyte gels based on the liquid-state integral
■ ACKNOWLEDGMENTS
This work has been partially supported by the Russian
equation theory has been presented in the recent work.53 The Foundation for Basic Research (Grant 11-03-00969a), by
contribution of ion−ion correlations to swelling of polyelec- Department of Chemistry and Material Science of the Russian
trolyte brushes54 or gels is, however, negligible in the case of Academy of Sciences, and by German Research Foundation
monovalent ions considered here. (DFG) within SFB 985.


At this point it is worth mentioning that a possibility of
discontinuous collapse transitions provoked by decreasing REFERENCES
solvent strength has been suggested earlier for star-branched
polyelectrolytes and polyelectrolyte brushes on the basis of (1) Barrat, J. L.; Joanny, J. F. In Advances in Chemical Physics;
Prigogine, I., Rice, S. A., Eds.; Wiley: New York, 1996.
theories which exploited box-like models.54−57 More accurate (2) Dobrynin, A. V.; Rubinstein, M. Prog. Polym. Sci. 2005, 30, 1049.
analysis with the account of nonuniform distribution of (3) Borisov, O. V.; Zhulina, E. B.; Leermakers, F. A. M.; Ballauff, M.;
polymer density and nonequal stretching of chains in Müller, A. H. E. Adv. Polym. Sci. 2011, 241, 1.
polyelectrolyte stars or in brushes proved that under poor (4) Plamper, F. A.; Becker, H.; Lanzendörfer, M.; Patel, M.;
solvent conditions coexistence of swollen and collapsed Wittemann, A.; Ballauff, M.; Müller, A. H. E. Macromol. Chem. Phys.
conformations occurs in polyelectrolyte stars or brushes on 2005, 206, 1813−1825.
the level of individual chains.58,59 More specifically, a decrease (5) Tanaka, T.; Fillmore, D. J. J. Chem. Phys. 1979, 70, 1214−1218.
in the solvent strength or/and an increase in salt concentration (6) Pelton, R. Adv. Colloid Interface Sci. 2000, 85, 1−33.
provokes microphase segregation in a star (or in a brush). This (7) Thorne, J. B.; Vine, G. J.; Snowden, M. J. Colloid Polym. Sci. 2011,
289, 625−646.
segregation is manifested in the formation of dense central
(8) Pich, A.; Richtering, W. Adv. Polym. Sci. 2011, 234, 1−37.
region formed by less extended chains and sparse outer region (9) Tanaka, T.; Fillmore, D.; Sun, S.-T.; Nishio, I.; Swislow, G.; Shah,
formed by more extended chains. In the case of pH-sensitive A. Phys. Rev. Lett. 1980, 45, 1636−1639.
polyelectrolytes the degree of ionization of the chains in the (10) Khokhlov, A. R.; Starodubtzev, S. G.; Vasilevskaya, V. V. Adv.
dense phase is smaller than that of the chain forming the Polym. Sci. 1993, 109, 123.
extended phase.23,39,40 The variation of temperature or/and salt (11) Schneider, S.; Linse, P. Macromolecules 2004, 37, 3850−3856.

8708 dx.doi.org/10.1021/ma401402e | Macromolecules 2013, 46, 8702−8709


Macromolecules Article

(12) Klooster, N. T. M.; Van der Touw, F.; Mandel, M. (49) Zhulina, E. B.; Borisov, O. V.; Birshtein, T. M. Macromolecules
Macromolecules 1984, 17, 2087−2093. 1999, 32, 8189−8196.
(13) Khokhlov, A. R.; Kramarenko, E.Yu. Macromol.Theory Simul. (50) Mei, Y.; Lauterbach, K.; Hoffmann, M.; Borisov, O. V.; Ballauff,
1994, 3, 45. M.; Jusufi, A. Phys. Rev. Lett. 2006, 97, 158301.
(14) Khokhlov, A. R.; Kramarenko, E.Yu. Macromolecules 1996, 29, (51) Plamper, F. A.; Walther, A.; Müller, A. H. E.; Ballauff, M. Nano
681. Lett. 2007, 7, 167−171.
(15) Starodoubtsev, S. G.; Khokhlov, A. R.; Sokolov, E. L.; Chu, B. (52) Jusufi, A.; Borisov, O. V.; Ballauff, M. Polymer 2013, 54, 2028−
Macromolecules 1995, 28, 3930−3936. 2035.
(16) Bromberg, L.; Temchenko, M.; Alakhov, V.; Hatton, T. A. (53) Sing, C. E.; Zwanikken, J. W.; Olvera de la Cruz, M.
Langmuir 2005, 21, 1590−1598. Macromolecules 2013, 46, 5053−5065.
(17) Bysell, H.; Malmstren, M. Langmuir 2006, 22, 5476−5484. (54) Borisov, O. V.; Birshtein, T. M.; Zhulina, E. B. J. Phys. II (Fr.)
(18) Hu, X.; Tong, Z.; Lyon, L. A. Colloid Polym. Sci. 2011, 289, 1991, 1, 521−526.
333−339. (55) Ross, R.; Pincus, P. Macromolecules 1992, 25, 2177.
(19) Kleinen, J.; Klee, A.; Richtering, W. Langmuir 2010, 26, 11258− (56) Borisov, O. V.; Birshtein, T. M.; Zhulina, E. B. Prog. Colloid
11265. Polym. Sci. 1992, 90, 177−181.
(20) Echeverria, C.; Lopez, D.; Mijangos, C. Macromolecules 2009, (57) Polotsky, A. A.; Zhulina, E. B.; Birshtein, T. M.; Borisov, O. V.
42, 9118−9123. Macromol. Symp. 2009, 278, 24−31.
(21) Pinheiro, J. P.; Moura, L.; Fokkink, R.; Farinha, J. P. S. Langmuir (58) Misra, S.; Mattice, W. L.; Napper, D. H. Macromolecules 1994,
27, 7090−7098.
2012, 28, 5802−5809.
(59) Pryamitsyn, V. A.; Leermakers, F. A. M.; Fleer, G. J.; Zhulina, E.
(22) Aseyev, V. O.; Tenhu, H.; Winnik, F. M. Adv. Polym. Sci. 2011,
B. Macromolecules 1996, 29, 8260.
242, 29−89.
(60) Hua, J.; Mitra, M. K.; Muthukumar, M. J. Chem. Phys. 2012, 136,
(23) Plamper, F. A.; Ruppel, M.; Schmalz, A.; Borisov, O. V.; Ballauff,
134901.
M.; Müller, A. H. E. Macromolecules 2007, 40, 8361−8366. (61) Mei, Y.; Ballauff, M. Eur. Phys. J. E 2005, 16, 341−349.
(24) Steinschulte, A. A.; Schulte, B.; Erberich, B.; Borisov, O. V.; (62) Plamper, F. A.; Schmalz, A.; Penott-Chang, E.; Drechsler, M.;
Plamper, F. A. ACS Macro Lett. 2012, 1, 504−507. Jusufi, A.; Ballauff, M.; Müller, A. H. E. Macromolecules 2007, 40,
(25) Emileh, A.; Vasheghani-Farahani, E.; Imani, M. Macromol. Symp. 5689−5697.
2007, 255, 1−7. (63) Jha, P. K.; Zwanikken, J. W.; Olvera de la Cruz, M. Soft Matter
(26) Hu, L.; Chu, L.-Y.; Yang, M.; Wang, H.-D.; Niu, C. H. J. Colloid 2012, 8, 9519−9522.
Interface Sci. 2007, 311, 110−117. (64) Dobrynin, A. V.; Rubinstein, M.; Obukhov, S. P. Macromolecules
(27) Nur, H.; Pinkrah, V. T.; Mitchell, J. C.; Benee, L. S.; Snowden, 1996, 29, 2974.
M. J. Adv. Colloid Interface Sci. 2010, 158, 15−20. (65) Sandberg, D. J.; Carillo, J. M. Y.; Dobrynin, A. V. Langmuir
(28) Edgecombe, S.; Schneider, S.; Linse, P. Macromolecules 2004, 37, 2007, 23, 12716−12728.
10089−10100. (66) Kosovan, P.; Kuldova, J.; Limpouchova, Z.; Prochazka, K.;
(29) Hou, Y.; Yu, C.; Liu, G.; Ngai, T.; Zhang, G. J. Phys. Chem. B Zhulina, E. B.; Borisov, O. V. Soft Matter 2010, 6, 1872.
2010, 114, 3799−3803. (67) Tagliazucchi, M.; Olvera de la Cruz, M.; Szleifer, I. Proc. Natl.
(30) Zhulina, E. B.; Birshtein, T. M.; Borisov, O. V. Macromolecules Acad. Sci. U.S.A. 2010, 107, 2693.
1995, 28, 1491. (68) Longo, G. S.; Olvera de la Cruz, M.; Szleifer, I. Macromolecules
(31) Lyatskaya, Yu. V.; Leermakers, F. A. M.; Fleer, G. J.; Zhulina, E. 2011, 44, 147−158.
B.; Birshtein, T. M. Macromolecules 1995, 28, 3562.
(32) Currie, E. P. K.; Sieval, A. B.; Fleer, G. J.; Cohen Stuart, M. A.
Langmuir 2000, 16, 8324−8333.
(33) Guo, X.; Ballauff, M. Phys. Rev. E 2001, 64, 015406.
(34) Konradi, R.; Rühe, J. Macromolecules 2005, 38, 4345−4354.
(35) Borisov, O. V.; Zhulina, E. B. Eur. Phys. J. B 1998, 4, 205−217.
(36) Klein Wolterink, J.; van Male, J.; Cohen Stuart, M. A.; Koopal,
L. K.; Zhulina, E. B.; Borisov, O. V. Macromolecules 2002, 35, 9176−
9190.
(37) Cong, P.; Genzer, J.; Szleifer, I. Phys. Rev. Lett. 2007, 98, 018302.
(38) Zhulina, E. B.; Borisov, O. V. Langmuir 2011, 27, 10615−10633.
(39) Polotsky, A. A.; Zhulina, E. B.; Birshtein, T. M.; Borisov, O. V.
Soft Matter 2012, 8, 9446−9459.
(40) Rud, O. V.; Mercurieva, A. A.; Leermakers, F. A. M.; Birshtein,
T. M. Macromolecules 2012, 45, 145.
(41) Shibayama, M.; Isono, K.; Okabe, S.; Karino, T.; Nagao, M.
Macromolecules 2004, 37, 2909−2918.
(42) Puhse, M.; Keerl, M.; Scherzinger, C.; Richtering, W.; Winter, R.
Polymer 2010, 51, 3653−3659.
(43) Grobelny, S.; Hofmann, C. H.; Erlkamp, M.; Plamper, F. A.;
Richtering, W.; Winter, R. Soft Matter 2013, 25, 5862−5866.
(44) Borisov, O. V.; Birshtein, T. M.; Zhulina, E. B. Polym. Sci. USSR
1988, 30, 772.
(45) Zhang, Y.; Furyk, S.; Bergbreiter, D. E.; Cremer, P. S. J. Am.
Chem. Soc. 2005, 127, 14505−14510.
(46) Borisov, O. V. J. Phys. II (Fr.) 1996, 6, 1−19.
(47) Kramarenko, E.Yu.; Khokhlov, A. R.; Yoshikawa, K. Macro-
molecules 1997, 30, 3383.
(48) Grosberg, A. Y.; Khokhlov, A. R. Statistical Physics of
Macromoecules; AIS Press: New York, 1994.

8709 dx.doi.org/10.1021/ma401402e | Macromolecules 2013, 46, 8702−8709

Potrebbero piacerti anche