You are on page 1of 7

Journal of Colloid and Interface Science 527 (2018) 195–201

Contents lists available at ScienceDirect

Journal of Colloid and Interface Science


journal homepage: www.elsevier.com/locate/jcis

Regular Article

Effect of Ca2+ ion concentration on adsorption of poly(carboxylate


ether)-based (PCE) superplasticizer on mica
Bo Wu a,b, Byong-Wa Chun c, Le Gu a, Tonya L. Kuhl b,⇑
a
School of Mechatronics Engineering, Harbin Institute of Technology, Harbin 150001, China
b
Department of Chemical Engineering and Materials Science, University of California at Davis, Davis, CA 95616, United States
c
GCP Applied Technologies, Cambridge, MA 02140, United States

g r a p h i c a l a b s t r a c t

a r t i c l e i n f o a b s t r a c t

Article history: Hypothesis: Poly(carboxylate ether)-based (PCE) superplasticizers consist of a carboxylic acid backbone
Received 19 March 2018 and grafted poly(ethylene glycol) (PEG) side chains. Ca2+ ion bridging mechanism is commonly purported
Revised 8 May 2018 to control PCE’s adsorption on negatively charged cement particle surfaces in cement suspension, thus
Accepted 8 May 2018
PCE was expected to adsorb on negatively charged surfaces in synthetic pore solutions via Ca2+/ACOO
Available online 9 May 2018
interactions.
Experiments: Adsorption behaviors of a commercial PCE on negatively charged mica were studied in
Keywords:
aqueous electrolyte solutions by a surface forces apparatus.
Surface forces apparatus (SFA)
Poly(carboxylate ether) (PCE)
Findings: Direct force measurements indicated that the PCE adsorbed onto mica from 0.1 M K2SO4 due to
Superplasticizer K+ ion chelation by the ether oxygen units ACH2CH2OA on the PEG chains, but surprisingly did not adsorb
Poly(ethylene glycol) (PEG) from either 0.1 M K2SO4 with saturated Ca(OH)2 or 0.1 M Ca(NO3)2. The adsorption in K2SO4 was weak,
Polymer adsorption enabling the adsorbed PCE layers to be squeezed out under modest compression. Upon separating the
Steric force surfaces, the PCE immediately achieved an identical re-adsorption. In high-calcium conditions, the PCE
Ion chelation was highly positively charged due to Ca2+ ion chelation by PEG chains and backbone carboxylic groups
Cement ACOO, and mica also underwent charge reversal due to electrostatic adsorption/binding of Ca2+ ions.
Suspension
Consequently, the interaction between mica and PCE was electrostatically repulsive and no PCE adsorp-
Dispersant
tion occurred. These findings can be explained by the complex interplay of ion chelation by PEG chains,
electrostatic binding and screening interactions with charged surfaces in the presence of monovalent and
divalent counterions, and ultimately charge reversal of both the charged surfaces and polyelectrolyte in
high divalent ion conditions.
Ó 2018 Elsevier Inc. All rights reserved.

⇑ Corresponding author.
E-mail address: tlkuhl@ucdavis.edu (T.L. Kuhl).

https://doi.org/10.1016/j.jcis.2018.05.016
0021-9797/Ó 2018 Elsevier Inc. All rights reserved.
196 B. Wu et al. / Journal of Colloid and Interface Science 527 (2018) 195–201

1. Introduction have been widely employed to measure the interaction forces


between polyelectrolyte adsorption layers in aqueous solutions
Polyelectrolytes are extensively used as dispersing agents to [25–29]. Uchikawa et al. [30] used AFM for the first time to demon-
improve the rheological properties of colloidal suspensions in a strate the existence of steric repulsive force from the adsorbed
great variety of industries including paints, ceramic slurry process- superplasticizer on cement particle surfaces. Since this pioneering
ing, cement and concrete placement, cosmetics and pharmaceuti- work, measurements of the steric force between PCE adsorption
cal formulations [1,2]. In the case of cement, polyelectrolyte layers have been carried out by AFM and SFA. Ferrari et al. [31]
additives enable improved flowability at reduced water to cement used an AFM silicon tip to detect the interaction force with calcium
ratios [3–5]. Lower water content in cement suspension con- silicate and ettringite surfaces in Milli-Q water with PCE. The mea-
tributes to higher mechanical strength and chemical resistance sured force was very small, suggesting little PCE’s adsorption under
[6–9]. A variety of water-reducing admixtures, also called super- these conditions. In contrast, significant repulsion between mica
plasticizers, are incorporated into high performance cement as surfaces was measured in aqueous PCE solution under both acidic
essential components to improve effective flowability of hydrating and alkaline conditions by SFA [32]. Adhesion was observed upon
cement paste. The mechanism for improved workability is thought separation in low PCE concentration conditions, which was
to rely on adsorption of superplasticizers onto cement particle sur- explained by the bridging of the PCE molecules between the two
faces through their charged backbones. The adsorbed polymer lay- mica surfaces through hydrogen bonding. Kauppi et al. [1] mea-
ers prevent the agglomeration of cement particles by electrostatic sured the interaction forces between MgO surfaces in KCl solution
and/or steric repulsions and release entrapped water to reduce the with anionic acrylic ester-ethylene oxide (AAE-EO) copolymer at
viscosity of cement suspension [10]. pH 10 using AFM. Analysis of the force-distance curves revealed
Comb-like (or bottle-brush) polymers have a linear backbone that the measured force had a long-range electrostatic component
with grafted side chains [11–13]. Steric repulsion between bulky and a short-range steric component, suggesting that the acrylic
side chains can cause both the backbone and side chains to highly backbone adsorbed onto the MgO surface with the PEG chains pro-
extend [14,15]. This unusual molecular configuration makes this truding into the solution in a coiled conformation. Ferrari and
class of polymers attractive for a wide range of applications such coworkers [8] also employed AFM to study the interactions
as cement dispersion [16], boundary friction modification [17], between a silicon nitride tip and multiple substrates (calcite,
biomimetic lubrication [18] and protein adsorption [19]. Comb- quartz, mica and MgO) in Milli-Q water, K2SO4 and KOH solutions
like poly(carboxylate ether)-based (PCE) superplasticizers are the with PCE, respectively. Their force measurements showed that the
most common type of superplasticizer in aqueous systems [16]. addition of PCE eliminated attractive jump-into-contact behavior
They consist of a negatively charged carboxylic backbone and and induced repulsive force between the tip and substrate in all
grafted side chains mainly composed of poly(ethylene glycol) cases. Along with zeta potential measurements and adsorption iso-
(PEO or PEG), as shown in Fig. 1. PEG side chains are used because therms, they concluded that the PCE did not adsorb on negatively
of their high osmotic repulsion in water and insensitivity to ionic charged substrates (calcite, quartz and mica) but adsorbed on the
strength [20]. In a hydrating cement system, the anionic backbones positively charged silicon nitride tip and MgO in these aqueous
are electrostatically attracted onto positively charged cement par- monovalent electrolyte conditions. However, reports on establish-
ticle surfaces (e.g., tricalcium aluminate ðC 3 AÞ and ettringite ðAFtÞ) ing PCE’s adsorption through divalent cation bridging are very lim-
and are also thought to be bridged to negatively charged particle ited in the literature. Flatt and coworkers [2] investigated the
surfaces (e.g., tricalcium silicate ðC 3 SÞ and calcium silicate hydrate interactions between C  S  H surfaces in 5 mM Ca(OH)2 solution
ðC  S  HÞ) through electrostatic and entropic Ca2þ =  COO inter- with PCE at different ionic strengths (adding 10 or 100 mM NaCl)
actions [16,21], with the side chains extending away from the par- by AFM, but the adsorption of PCE on CASAH could not be con-
ticle surfaces into the cement pore solution. Consequently, the firmed from the measured interaction forces. Similarly, direct mea-
cement particles are dispersed primarily by the steric hindrance surements of adsorption of polyelectrolyte or comb-like polymer
produced by the side chains. PCE polymers exhibit superior dis- due to divalent counterion bridging are very sparse, though this
persing ability compared to conventional dispersants such as lig- mechanism is purported to explain phenomenological changes in
nosulfonate and naphthalene or melamine formaldehyde such systems.
condensates based superplasticizers [10]. Large bodies of research The goal of the present study was to determine the adsorption
on zeta potential [5,6,8,10], adsorption isotherm [4–10] and rheo- behavior of a commercial PCE on negatively charged mica in syn-

logical properties [4,5,10] of model cement suspensions and/or thetic pore solution containing abundant Ca2+, K+, SO2 4 and OH

actual cement pastes have indicated the significance of added ions that mimics actual cement suspension conditions. Direct force
PCE dispersants, where enhanced flow properties were ascribed measurements at the molecular level by SFA showed that the PCE
to strong steric repulsion of the side chains. polymer adsorbed onto mica from 0.1 M K2SO4 solution but did not
Atomic force microscope (AFM) [22,23] and surface forces appa- adsorb from either the synthetic pore solution or 0.1 M Ca(NO3)2
ratus (SFA) [24], two techniques for direct force measurement, solution. This work is relevant to understanding and improving
PCEs as cement/concrete admixtures, not only in terms of their
action as dispersing agents, but also their interaction with aggre-
gate minerals such as alminosilicate clays. More broadly, the work
increases our understanding of the properties of adsorbed poly-
electrolyte in high-salt conditions in the absence and presence of
divalent counterions.

2. Experimental

2.1. Materials

A commercial cement dispersant PCE polymer was utilized as


Fig. 1. Molecular structural schematic of the PCE polymer. received from GCP Applied Technologies (Cambridge, USA): 55 wt
B. Wu et al. / Journal of Colloid and Interface Science 527 (2018) 195–201 197

% aqueous solution, pH 6–7. Its molecular configuration was comb- and separation, respectively [33]. Jump-into-contact and jump-out
like, consisting of a carboxylic backbone and grafted PEG side behaviors were frequently observed in the absence of PCE’s adsorp-
chains (Fig. 1). On the backbone, the ratio n/m of carboxylic groups tion. All the SFA experiments were conducted at 25 °C with droplets
(COO ) to PEG chains was approximately 5:1. The side chain PEG of the various solutions injected between the surfaces. To minimize
had about 45 repeating ether oxygen units CH2 CH2 O. The total evaporation, a small volume of water was placed at the bottom of
molecular weight (MW) was MW  13,361 g/mol (Mn  4,878 the SFA chamber. In most cases, experiments were done by first
g/mol), yielding a polydispersity index (PDI) of 2.7. Potassium sul- measuring the interaction in pure electrolyte solution and then
fate (K2SO4, 99.99% trace metals basis) and calcium hydroxide (Ca exchanging the solution with the desired PCE solution.
(OH)2, 99.995% trace metals basis) were purchased from SIGMA-
ALDRICH (MO, USA). Calcium nitrate hydrate (Ca(NO3)24H2O, 3. Results
99.995% metals basis) was purchased from Alfa Aesar (MA, USA).
3.1. Interaction between mica surfaces in K2SO4 with PCE
2.2. Preparation of aqueous PCE polymer solutions
The interaction force as a function of distance between mica
All aqueous PCE polymer solutions were prepared at room tem- surfaces in 0.1 M K2SO4 solution is shown in Fig. 2a. No long-
perature. Aqueous 0.1 M K2SO4 and 0.1 M Ca(NO3)2 solutions were range electrostatic repulsion was observed due to the high elec-
directly prepared with Milli-Q water (17.8 MX cm). A low- trolyte concentration. The Debye length j1 was only 5.6 Å under
calcium K2SO4 solution was prepared by adding a small quantity this condition, calculated by:
of Ca(OH)2 (about 4 mg) into 20 mL of 0.1 M K2SO4 solution, fol-
lowed by sonication for at least 1 h. After sonication, this low- j1 ¼ 0:176½K 2 SO4 0:5 ð2Þ
calcium K2SO4 solution was filtered through a 0.02 lm aluminum where [K 2 SO4 ] is the concentration in mol/L [33]. Upon close
matrix anotop filter (Whatman, England). 0.1 M K2SO4 solution approach, the surfaces jumped into hard contact from about 40 Å
with saturated Ca(OH)2 was prepared by the similar procedure: due to the attractive van der Waals force of the mica substrates.
adding about 20 mg Ca(OH)2 into 20 mL of 0.1 M K2SO4 solution, Upon separation, the surfaces jumped out of adhesive contact with
followed by sonication for more than 1 h. The filtered synthetic F ad
 3mN=m.
R
pore solution pH was 12. All the aqueous PCE electrolyte solu-
With addition of 0.1 wt% PCE polymer, a repulsive force was
tions with desired PCE concentration (0.1 wt% or 1 wt%) were
detected at a separation of about 100 Å. With further approach,
obtained by direct addition of the PCE polymer to the electrolyte
the repulsion increased exponentially, as shown in the semi-
solution without any further processing. In particular, these fresh
logarithmic graph in Fig. 2b. The characteristic decay length was
aqueous PCE electrolyte solutions were clear, with no precipitation
about 20 Å, much larger than the Debye length (5.6 Å). These mea-
observed.
surements clearly demonstrate that the PCE adsorbed onto the
negatively charged mica from 0.1 M K2SO4 solution. The repulsion
2.3. Force-distance measurements was not electrostatic but due to osmotic pressure from the
adsorbed PCE layers. The onset of steric force at 100 Å indicates
A SFA was used to measure the interaction forces between mica that the two adsorption layers began to interact at this separation.
surfaces in aqueous electrolytes in the absence and presence of the Assuming that the polymer layers were symmetric, the thickness L
PCE polymer. The SFA technique directly measures the interaction of one uncompressed PCE adsorption layer was about 50 Å. With
forces between two macroscopic surfaces at the molecular level further compression, the surfaces jumped into the same hard wall
and has been extensively described elsewhere [24,33,34]. Briefly, contact as in the absence of PCE, but from a smaller separation of
back silvered (550 Å) mica sheets of uniform thickness were glued approximately 25 Å. These force profiles demonstrate that the
onto curved cylindrical silica disks and then mounted in a cross- adsorbed PCE layers were squeezed out from the gap at a relatively
cylinder configuration in the SFA chamber. This configuration F
small compression Rsq  9 mN : At lower compressions, the adsorp-
m
was locally equivalent to a sphere-on-flat geometry as the surface
tion layers were not squeezed out, but hysteresis was observed.
separation D was far smaller than the surface curvature. The abso-
Hysteresis between the approach and separation of the surfaces
lute separation D between the two mica surfaces was directly
was due to structural rearrangement and slow relaxation of the
determined by multiple beam interferometry (MBI) based on inter-
adsorbed polymer films during compression/decompression.
ference fringes of equal chromatic order (FECO) [35], which were
produced by passing white light through the opposing surfaces.
3.2. Interactions between mica surfaces in K2SO4 with PCE at different
The FECO fringe wavelengths were measured using a spectrometer,
calcium concentrations
enabling the distance D to be determined with a resolution up to 1
Å. ‘‘Zero” distance was defined as the contact of two bare mica sur-
The series of profiles in Fig. 3 show the interaction forces
faces in air. The interaction force F ðDÞ between the surfaces was
between mica surfaces in 0.1 M K2SO4 solution with different
measured from the deflection of a double cantilever spring sup-
amounts of Ca(OH)2 in the absence and presence of PCE polymer.
porting the lower surface. To enable quantitative comparison
With addition of a small quantity of Ca(OH)2, as shown in Fig. 3a,
between different experiments, the force F ðDÞ was normalized by
the interaction was similar to that in pure 0.1 M K2SO4 solution:
the geometric mean radius R of the contact position according to
van der Waals attraction at small distances with jumps into and
the Derjaguin approximation:
out of adhesive contact. The addition of 1 wt% PCE caused a similar
F ðDÞ ¼ 2pREðDÞ ð1Þ repulsion due to PCE’s adsorption as in the absence of calcium. The
repulsion was exponential with almost the same characteristic
whereEðDÞ is the interaction energy per unit area between two flat decay length, as indicated by the corresponding semi-logarithmic
surfaces. The measured radius R was about 1.5 cm in the reported graph in Fig. 3b. The adsorption layers also exhibited hysteresis
experiments. The Derjaguin approximation is valid when D  R. and squeeze-out behaviors at low and high compressions, respec-
When @FðDÞ
@D
is greater than the stiffness of the spring, there will be tively. By comparing these force profiles with those obtained in 0.1
a mechanical instability that can induce the lower surface to jump M K2SO4 with 0.1 wt% PCE, the repulsion due to the adsorbed PCE
either toward or away from the upper surface during approach was significantly reduced, though the onset of repulsion arose at
198 B. Wu et al. / Journal of Colloid and Interface Science 527 (2018) 195–201

Fig. 2. Interaction force profiles between mica surfaces in 0.1 M K2SO4 in the absence and presence of 0.1 wt% PCE: (a) linear-linear plot; (b) linear-log plot of (a). Solid arrows
designate jump-into or jump-out-of contact. Dashed arrows indicate direction of approach or separation.

Fig. 3. Interaction force profiles between mica surfaces in 0.1 M K2SO4 with different amounts of Ca(OH)2 in the absence and presence of 1 wt% PCE: (a) with a small quantity
of Ca(OH)2, linear-linear plot; (b) linear-log plot of (a); (c) with saturated Ca(OH)2. Solid arrows designate jump-into or jump-out-of contact. Dashed arrows indicate direction
of approach or separation.
B. Wu et al. / Journal of Colloid and Interface Science 527 (2018) 195–201 199

almost the same separation (around 100 Å). In particular, the hydration repulsion due to strong electrostatic binding/adsorption
F
repulsive force at the critical squeeze-out state, Rsq , was about half. of Ca2+ ions on mica [37]. In particular, the negatively charged mica
Additionally, the critical squeeze-out separation was slightly substrates underwent charge reversal in 0.1 M Ca(NO3)2 solution
shifted to larger separation. The decrease in repulsion and increase due to calcium condensation at this concentration [38,39]. Thus,
in squeeze-out/jump-into-contact separation imply a higher com- negatively charged PCE would electrostatically adsorb onto the
pressibility and a lower adsorption strength. The addition of a now positively charged mica through its carboxylic groups ACOO.
small amount of Ca(OH)2, therefore, weakened the adsorption However, the force profiles without and with 1 wt% PCE almost
strength of PCE on mica compared to the pure K2SO4 solution. overlap, indicating that very little PCE adsorbed onto mica from
Overall, the adsorption layers established in only K2SO4 solution this calcium-ion-rich solution.
and with low calcium shared the same interaction characteristics,
suggesting that the adsorption mechanism was the same.
4. Discussion
To mimic cement pore solution conditions, force measurements
were further performed in 0.1 M K2SO4 solution with saturated Ca
4.1. PCE’s adsorption on mica from 0.1 M K2SO4
(OH)2 in the absence and presence of PCE. Fig. 3c shows that the
force-distance profiles obtained in the synthetic pore solution with
PEG is sometimes called ‘‘a poor chemist’s crown” for its ability
1 wt% PCE almost overlap those in the same solution without PCE.
to chelate alkali metal ions [40,41]. The chelation mechanism is a
Thus, almost no PCE adsorbed onto mica from the calcium-
complexation interaction between the alkali metal ion and the lone
saturated K2SO4 solution. The differences in adhesion were due
pair oxygen electrons in the ether oxygen unit CH2 CH2 O [42,43].
to changes in the effective charge density of mica surface under
It has been reported that six ether oxygen atoms chelate one K+ ion
these electrolyte solution conditions. In addition, differences in
to form a crown ether ring [44]. For example, Zhivkova et al. [45]
the orientation of the mica substrates (alignment of the crystallo-
found that PEO behaved as a positively charged polyelectrolyte in
graphic planes) can also affect the measured adhesion [36]. Com-
aqueous solution due to cation chelation. Thus, the PCE in our stud-
parison of the force profiles in Fig. 3a and c clearly shows that
ies behaved as a bifunctional polyelectrolyte with negatively
increasing the concentration of Ca(OH)2 weakened and decreased
charged carboxylic groups ACOO along the backbone and posi-
PCE’s adsorption on mica in 0.1 M K2SO4 solution.
tively charged PEG arms due to K þ ion chelation. The high elec-
trolyte concentration ensured a high level of K+ ion chelation by
the PEG chains. As a result, the PCE was sufficiently positively
3.3. Interaction between mica surfaces in Ca(NO3)2 with PCE
charged to adsorb onto negatively charged mica surfaces. This is
further supported by the thickness of the adsorbed PCE layer.
It was very surprising that PCE adsorbed onto mica from 0.1 M
The Flory radius RF of the PEG side chain was about 35 Å, calculated
K2SO4 solution but did not after adding saturated Ca(OH)2. To con-
by:
firm that calcium inhibited PCE’s adsorption, rather than the previ-
ously suggested bridging mechanism that induces PCE’s adsorption
RF ¼ aðMw =M w;monomer Þ0:6 ð3Þ
on negatively charged surfaces via Ca2+/ACOO interactions,
experiments were conducted in 0.1 M Ca(NO3)2 solution without
where a is the monomer (CH2 CH2 O) size (3.5 Å), Mw is the
and with 1 wt% PCE, respectively. The measured force profiles are
molecular weight (2000 g/mol) of a PEG side chain and
shown in Fig. 4. In the pure Ca(NO3)2 solution, a significant,
Mw;monomer is the molecular weight (44 g/mol) of a monomer
short-range repulsion was observed before the surfaces jumped
(CH2 CH2 O). The thickness L of the adsorbed PCE layer was about
into contact from about 30 Å. This repulsion was consistent with
50 Å, suggesting that some PEG chains adsorbed on mica while
others extended into solution. A range betweenRF and 2RF is
expected for a weakly adsorbed polymer [33].
The surprising squeeze-out behavior at a relatively low com-
pression (9 mN/m) indicates that PCE’s adsorption on mica due
to K þ ion chelation was weak. An earlier study by Chai et al. [43]
showed that the adsorption layers of PEO (M w ¼ 170; 000g=mol,
Mw
Mn
¼ 1:02, concentration = 150 lg/mL) on mica in 0.1 M KNO3 were
not squeezed out. Compared to this study with high molecular
weight PEO, the PCE here has 5–6 relatively short PEG side chains
(M w  2; 000g=mol per chain). The adsorption strength scales with
the molecular weight and the molecular weight also dictates the
total positive charge chelation of the polymer [46]. Further, the
negatively charged carboxylic backbone should also reduce the
binding strength to mica because of electrostatic repulsion. As a
result, the PCE had a much lower binding strength and could be
squeezed out at modest compression. Such squeeze-out phe-
nomenon is relatively rare, but has been observed with low MW
PEG chains adsorbing to lipid bilayers in KNO3 solutions [44,47].
Similarly, it has also been reported that hyaluronic acid layers
adsorbing on mica can be squeezed out with compression in CaCl2
Hepes buffer solution [48]. In this case, the weak adsorption was
attributed to the mediation of a water molecule, which decreased
the strength of the Ca2+ ion bridging interaction between mica
Fig. 4. Interaction force profiles between mica surfaces in 0.1 M Ca(NO3)2 in the
and COO groups in hyaluraonic acid. Additionally, shear was also
absence and presence of 1 wt% PCE. Solid arrows designate jump-into or jump-out- reported to induce abrupt removal of adsorbed polyelectrolyte and
of contact. Dashed arrow indicates direction of approach. neutral polymer brushes from contacting mica surfaces [49,50].
200 B. Wu et al. / Journal of Colloid and Interface Science 527 (2018) 195–201

Another unusual phenomenon observed here was that force- was abolished due to electrostatic repulsion between both the
distance profiles of multiple successive approach-separation runs effectively positively charged mica surface and PCE.
mostly overlap. When the PCE polymer was excluded from the The situation in practical cement suspensions is much more
gap at high compression, attractive van der Waals force brought complicated. High concentrations of ions including Ca2+ are
the surfaces into direct contact. Upon separating the surfaces sev- released upon cement hydration. Both positively and negatively
eral hundred angstroms, the polymer immediately achieved an charged particles are present and PCE may adsorb onto both types
almost identical re-adsorption. In addition, the reproducibility of of surfaces depending on the charge density of the surface, rate of
the force profiles showing compression-decompression hysteresis hydration and local effective salt solution conditions. Given the
also indicates that the PCE adsorption layers almost immediately effort to keep the water to cement ratio as low as possible, non-
recovered to their uncompressed state once the surfaces were sep- equilibrium effects that limit the rate of dissociation may also be
arated more than 2L. Significant time-dependent effects in succes- very important in real applications. Additionally, there is a
sive compression-decompression cycles between monodisperse dynamic balance between two driving forces of PCE’s adsorption
PEO (M w ¼ 160; 000g=mol) adsorption layers on mica in 0.1 M to cement particles: enthalpic interactions of carboxylic acid with
KNO3 were observed by Klein and Luckham [51]. Pettersson et al. calcium and entropic contributions due to the release of water
[52] found that the molar percentage of charged backbone seg- and counterions from carboxylic acid and calcium on the surfaces.
ments had an important influence on the hysteresis between com- The current study implies that cement dispersants may also func-
pression and decompression of PEG bottle brush polymers. The tion as a lubricant between the aluminosilicate clay surfaces in the
reproducibility of the profiles in this work suggests that the pore solution under high calcium concentration.
adsorbed PCE molecules were retained in the gap between the sur-
faces during compression until the compression reached the critical
5. Conclusions
squeeze-out level. A structurally similar polyelectrolyte PLL–g–PEG
also exhibited this retention characteristic during solvent exchange
Adsorption behaviors of a commercial PCE on negatively
and rinsing [53,54]. Although the squeezing out and relatively mod-
charged mica in 0.1 M K2SO4 solution with different concentrations
est adsorption strength of PCE to negatively charged surfaces are
of Ca(OH)2 were directly investigated by SFA. The PCE weakly
undesirable behaviors when considering PCE as a long-term
adsorbed onto mica from the K2SO4 solution due to K+ ion chelation
lubricant and dispersion modifier, the rapid re-adsorption likely
by PEG chain segments (ether oxygen units ACH2CH2OA) and the
mitigates this effect in rheological flow conditions, which is of pri-
adsorption layers were squeezed out with modest compression.
mary importance in cement placement and consolidation.
Adding a small quantity of Ca(OH)2 further weakened the adsorp-
tion strength and increasing Ca(OH)2 to saturated solution condi-
tions abolished the adsorption. The squeeze-out phenomenon in
4.2. Ca2+ ion effect on PCE’s adsorption in 0.1 M K2SO4
K2SO4 solution is relatively rare, which increases our understanding
of polyelectrolyte adsorption behaviors in high-salt conditions. Ca2+
One adsorption mechanism of PCE is thought to rely on Ca2+ ion
ions are commonly thought to anchor the backbone carboxylic
bridging between the backbone carboxylic groups COO and
groups ACOO to negatively charged cement particle surfaces to
negatively charged particle surfaces [16,21]. Ca2+ ion bridging has
form robust PCE adsorption layers for cement suspension stabiliza-
also been suggested as the mechanism for the adsorption of
tion [16,21]. Therefore, PCE was expected to adsorb on mica in syn-
ACOO-containing sodium polyacrylate [55] and hyaluronic acid
thetic pore solution conditions through Ca2+/ACOO interactions.
[48] on negatively charged mica. In contrast, the high-resolution
Surprisingly, however, the effect of Ca2+ ion on the adsorption
force measurements by SFA here clearly reveal that the expected
was inverse to the hypothesized bridging mechanism. Compared
adsorption of PCE on mica by Ca2+ ion bridging did not occur and
to previous force-distance measurements in Milli-Q water [8,31],
that PCE’s adsorption in 0.1 M K2SO4 was significantly weakened
monovalent electrolyte aqueous solutions [1,8,30,32] and low-
after adding a small quantity of Ca(OH)2.
calcium aqueous solutions [2] by AFM and SFA, the present mea-
As with monovalent K+ ions, divalent Ca2+ ions can be also che-
surements give us new understanding of the interactions between
lated by the PEG chain segments (ether oxygen units ACH2CH2OA)
PCE and negatively charged surfaces in high-calcium aqueous solu-
[40,56,57]. Thus, the Ca2+/ACOO and Ca2+/ACH2CH2OA complexa-
tions at high ionic strength and pH, which have not been previously
tion interactions could potentially induce PCE aggregation, thereby
explored. It is precisely these conditions where our understanding
reducing the effective amount of PCE available to adsorb. However,
of polyelectrolyte absorption behavior is limited. The surprising
the hydrodynamic diameter determined by dynamic light scatter-
effect of Ca2+ ions can be explained by charge reversal that hap-
ing (Zetasizer Nano (UK) at 25 °C) at the concentration of 1 wt% in
pened to both the mica surface due to strong electrostatic adsorp-
Milli-Q water was 9.3 nm and was only slightly decreased to 7.3
tion/binding of Ca2+ ions and the PCE due to Ca2+ ion chelation by
nm and 6.9 nm in 0.1 M K2SO4 solution and 0.1 M Ca(NO3)2 solution
the ether oxygen units ACH2CH2OA on the PEG side chains and
due to charge screening, respectively. Therefore, PCE molecules did
backbone carboxylic groups ACOO. What is less appreciated in
not aggregate in the experimental solution conditions.
the literature is that divalent ion bridging mechanisms should only
Nevertheless, the presence of calcium changed the charging
function over a finite concentration range and will depend on the
behaviors of both PCE and the mica surface. With addition of a
charge density of the particles, architecture of the polymer disper-
small amount of Ca(OH)2, the electrostatic adsorption/binding of
sant and solution conditions. Correspondingly, future studies of
Ca2+ ions can partially neutralize some of the negatively charged
PCE’s adsorption behaviors as a function of calcium concentration
sites on the mica surface as well as the negatively charged car-
will further inform their use as dispersants under high ionic
boxylic groups along the PCE backbone. As a result, the electro-
strength conditions.
static attraction between PCE and mica was weakened in the
low-calcium conditions, thereby decreasing the adsorption
strength. At high calcium concentrations, the PCE became more Acknowledgements
positively charged due to Ca2+ ion chelation by PEG chain units
ACH2CH2OA and carboxylic groups ACOO. Meanwhile, the mica This work was supported by GCP Applied Technologies,
was subjected to charge reversal due to strong electrostatic Cambridge, Massachusetts, USA and the National Natural Science
adsorption/binding of Ca2+ ions [38,39]. Consequently, adsorption Foundation of China (51675120, U1537214 and U1637206). The
B. Wu et al. / Journal of Colloid and Interface Science 527 (2018) 195–201 201

overseas study of author Bo Wu at the University of California at [27] J. Yu, J. Mao, G. Yuan, S. Satija, Z. Jiang, W. Chen, M. Tirrell, Structure of
polyelectrolyte brushes in the presence of multivalent counterions,
Davis was sponsored by the non-profit China Scholarship Council.
Macromolecules 49 (2016) 5609–5617.
[28] J. Faivre, B.R. Shrestha, J. Burdynska, G. Xie, F. Moldovan, T. Delair, S. Benayoun,
L. David, K. Matyjaszewski, X. Banquy, Wear protection without surface
References modification using a synergistic mixture of molecular brushes and linear
polymers, ACS Nano 11 (2017) 1762–1769.
[1] A. Kauppi, K.M. Andersson, L. Bergström, Probing the effect of superplasticizer [29] U. Raviv, S. Glasson, N. Kampf, J.-F. Gohy, Lubrication by charged polymers,
adsorption on the surface forces using the colloidal probe AFM technique, Cem. Nature 425 (2003) 163–165.
Concr. Res. 35 (2005) 133–140. [30] H. Uchikawa, S. Hanehara, D. Sawaki, The role of steric repulsive force in the
[2] R.J. Flatt, I. Schober, E. Raphael, C. Plassard, E. Lesniewska, Conformation of dispersion of cement particles in fresh paste prepared with organic admixture,
adsorbed comb copolymer dispersants, Langmuir 25 (2008) 845–855. Cem. Concr. Res. 27 (1997) 37–50.
[3] P.-C. Aitcin, C. Jolicoeur, J.G. MacGregor, Superplasticizers: how they work and [31] L. Ferrari, L. Bernard, F. Deschner, J. Kaufmann, F. Winnefeld, J. Plank,
why they occasionally don’t, Concr. Int. 16 (1994) 45–52. Characterization of polycarboxylate-ether based superplasticizer on cement
[4] E. Sakai, A. Ishida, A. Ohta, New trends in the development of chemical clinker surfaces, J. Am. Ceram. Soc. 95 (2012) 2189–2195.
admixtures in Japan, J. Adv. Concr. Technol. 4 (2006) 211–223. [32] L. Zhang, Q. Lu, Z. Xu, Q. Liu, H. Zeng, Effect of polycarboxylate ether comb-type
[5] Y. Li, C. Yang, Y. Zhang, J. Zheng, H. Guo, M. Lu, Study on dispersion, adsorption polymer on viscosity and interfacial properties of kaolinite clay suspensions, J.
and flow retaining behaviors of cement mortars with TPEG-type polyether Colloid Interface Sci. 378 (2012) 222–231.
kind polycarboxylate superplasticizers, Constr. Build. Mater. 64 (2014) 324– [33] J.N. Israelachvili, Intermolecular and Surface Forces, third ed., Academic press,
332. London, 2010.
[6] K. Yoshioka, E.-I. Tazawa, K. Kawai, T. Enohata, Adsorption characteristics of [34] J. Israelachvili, Direct measurements of forces between surfaces in liquids at
superplasticizers on cement component minerals, Cem. Concr. Res. 32 (2002) the molecular level, Proc. Natl. Acad. Sci. 84 (1987) 4722–4724.
1507–1513. [35] J. Israelachvili, Thin film studies using multiple-beam interferometry, J. Colloid
[7] H. Hommer, Interaction of polycarboxylate ether with silica fume, J. Eur. Interface Sci. 44 (1973) 259–272.
Ceram. Soc. 29 (2009) 1847–1853. [36] P.M. Mcguiggan, J.N. Israelachvili, Measurements of the effect of angular lattice
[8] L. Ferrari, J. Kaufmann, F. Winnefeld, J. Plank, Interaction of cement model mismatch on the adhesion energy between 2 mica surfaces in water, Mater.
systems with superplasticizers investigated by atomic force microscopy, zeta Res. Soc. 138 (1989) 349–360.
potential, and adsorption measurements, J. Colloid Interface Sci. 347 (2010) [37] R. Kjellander, S. Marčelja, R. Pashley, J. Quirk, A theoretical and experimental
15–24. study of forces between charged mica surfaces in aqueous CaCl2 solutions, J.
[9] T. Sowoidnich, T. Rachowski, C. Rößler, A. Völkel, H.-M. Ludwig, Calcium Chem. Phys. 92 (1990) 4399–4407.
complexation and cluster formation as principal modes of action of polymers [38] D.F. Parsons, B.W. Ninham, Charge reversal of surfaces in divalent electrolytes:
used as superplasticizer in cement systems, Cem. Concr. Res. 73 (2015) 42–50. the role of ionic dispersion interactions, Langmuir 26 (2010) 6430–6436.
[10] O. Akhlaghi, Y.Z. Menceloglu, O. Akbulut, Poly(carboxylate ether)-based [39] P. Kékicheff, S. Marc̆elja, T. Senden, V. Shubin, Charge reversal seen in electrical
superplasticizer achieves workability retention in calcium aluminate cement, double layer interaction of surfaces immersed in 2:1 calcium electrolyte, J.
Sci. Rep., 7, 2017. Chem. Phys. 99 (1993) 6098–6113.
[11] S.S. Sheiko, B.S. Sumerlin, K. Matyjaszewski, Cylindrical molecular brushes: [40] F. Bailey, R. Callard, Some properties of poly(ethylene oxide)1 in aqueous
synthesis, characterization, and properties, Prog. Polym. Sci. 33 (2008) 759– solution, J. Appl. Polym. Sci. 1 (1959) 56–62.
785. [41] D. Balasubramanian, B. Chandani, Poly(ethylene glycol): a poor chemist’s
[12] J. Plank, E. Sakai, C. Miao, C. Yu, J. Hong, Chemical admixtures—chemistry, crown, J. Chem. Educ. 60 (1983) 77–78.
applications and their impact on concrete microstructure and durability, Cem. [42] L. Chai, J. Klein, Role of ion ligands in the attachment of poly(ethylene oxide) to
Concr. Res. 78 (2015) 81–99. a charged surface, J. Am. Chem. Soc. 127 (2005) 1104–1105.
[13] C.M. Bates, F.S. Bates, 50th anniversary perspective: block polymers pure [43] L. Chai, R. Goldberg, N. Kampf, J. Klein, Selective adsorption of poly(ethylene
potential, Macromolecules 50 (2016) 3–22. oxide) onto a charged surface mediated by alkali metal ions, Langmuir 24
[14] S. Rathgeber, T. Pakula, A. Wilk, K. Matyjaszewski, K.L. Beers, On the shape of (2008) 1570–1576.
bottle-brush macromolecules: systematic variation of architectural [44] T.L. Kuhl, A.D. Berman, S.W. Hui, J.N. Israelachvili, Part 2. Crossover from
parameters, J. Chem. Phys. 122 (2005) 124904. depletion attraction to adsorption: polyethylene glycol induced electrostatic
[15] B. Zhang, F. Gröhn, J.S. Pedersen, K. Fischer, M. Schmidt, Conformation of repulsion between lipid bilayers, Macromolecules 31 (1998) 8258–8263.
cylindrical brushes in solution: effect of side chain length, Macromolecules 39 [45] I. Zhivkova, A. Zhivkov, D. Stoychev, Electrostatic behavior of polyethylene
(2006) 8440–8450. oxide, Eur. Polym. J. 34 (1998) 531–538.
[16] E. Sakai, K. Yamada, A. Ohta, Molecular structure and dispersion-adsorption [46] P.G. Degennes, Scaling theory of polymer adsorption, Journal de Physique 37
mechanisms of comb-type superplasticizers used in Japan, J. Adv. Concr. (1976) 1445–1452.
Technol. 1 (2003) 16–25. [47] T.L. Kuhl, A.D. Berman, S.W. Hui, J.N. Israelachvili, Part 1. Direct measurement
[17] S. Lee, M. Müller, M. Ratoi-Salagean, J. Vörös, S. Pasche, S.M. De Paul, H.A. of depletion attraction and thin film viscosity between lipid bilayers in
Spikes, M. Textor, N.D. Spencer, Boundary lubrication of oxide surfaces by poly aqueous polyethylene glycol solutions, Macromolecules 31 (1998) 8250–8257.
(Llysine)gpoly(ethylene glycol) (PLLgPEG) in aqueous media, Tribo. [48] R. Tadmor, N. Chen, J. Israelachvili, Normal and shear forces between mica and
Lett. 15 (2003) 231–239. model membrane surfaces with adsorbed hyaluronan, Macromolecules 36
[18] X. Banquy, J. Burdyńska, D.W. Lee, K. Matyjaszewski, J. Israelachvili, (2003) 9519–9526.
Bioinspired bottle-brush polymer exhibits low friction and amontons-like [49] U. Raviv, S. Giasson, N. Kampf, J.-F. Gohy, R. Jérôme, J. Klein, Normal and
behavior, J. Am. Chem. Soc. 136 (2014) 6199–6202. frictional forces between surfaces bearing polyelectrolyte brushes, Langmuir
[19] Z. Zhang, H. Ma, D.B. Hausner, A. Chilkoti, T.P. Beebe, Pretreatment of 24 (2008) 8678–8687.
amphiphilic comb polymer surfaces dramatically affects protein adsorption, [50] J. Klein, E. Kumacheva, D. Perahia, L. Fetters, Shear forces between sliding
Biomacromolecules 6 (2005) 3388–3396. surfaces coated with polymer brushes: the high friction regime, Acta
[20] S. Pasche, S.M. De Paul, J. Vörös, N.D. Spencer, M. Textor, Poly(llysine) Polymerica 49 (1998) 617–625.
graftpoly(ethylene glycol) assembled monolayers on niobium oxide [51] J. Klein, P.F. Luckham, Forces between two adsorbed poly(ethylene oxide)
surfaces: a quantitative study of the influence of polymer interfacial layers in a good aqueous solvent in the range 0–150 nm, Macromolecules 17
architecture on resistance to protein adsorption by ToF-SIMS and in situ (1984) 1041–1048.
OWLS, Langmuir 19 (2003) 9216–9225. [52] T. Pettersson, A. Naderi, R. Makuška, P.M. Claesson, Lubrication properties of
[21] J. Plank, B. Sachsenhauser, J. De Reese, Experimental determination of the bottle-brush polyelectrolytes: an AFM study on the effect of side chain and
thermodynamic parameters affecting the adsorption behavior and dispersion charge density, Langmuir 24 (2008) 3336–3347.
effectiveness of PCE superplasticizers, Cem. Concr. Res. 40 (2010) 699–709. [53] M.T. Müller, X. Yan, S. Lee, S.S. Perry, N.D. Spencer, Lubrication properties of a
[22] G. Binnig, C.F. Quate, C. Gerber, Atomic force microscope, Phys. Rev. Lett. 56 brushlike copolymer as a function of the amount of solvent absorbed within
(1986) 930–933. the brush, Macromolecules 38 (2005) 5706–5713.
[23] W.A. Ducker, T.J. Senden, R.M. Pashley, Direct measurement of colloidal forces [54] E. Guzman, F. Ortega, M.G. Prolongo, V.M. Starov, R.G. Rubio, Influence of the
using an atomic force microscope, Nature 353 (1991) 239–241. molecular architecture on the adsorption onto solid surfaces: comb-like
[24] J. Israelachvili, Y. Min, M. Akbulut, A. Alig, G. Carver, W. Greene, K. Kristiansen, polymers, Phys. Chem. Chem. Phys. 13 (2011) 16416–16423.
E. Meyer, N. Pesika, K. Rosenberg, Recent advances in the surface forces [55] J.M. Berg, P.M. Claesson, R.D. Neuman, Interactions between mica surfaces in
apparatus (SFA) technique, Rep. Prog. Phys. 73 (2010) 036601. sodium polyacrylate solutions containing calcium ions, J. Colloid Interface Sci.
[25] E. Spruijt, M.A. Cohen Stuart, J. van der Gucht, Dynamic force spectroscopy of 161 (1993) 182–189.
oppositely charged polyelectrolyte brushes, Macromolecules 43 (2010) 1543– [56] M.J. Reddy, P.P. Chu, Effect of Mg2+ on PEO morphology and conductivity, Solid
1550. State Ionics 149 (2002) 115–123.
[26] Q. Lu, D.S. Hwang, Y. Liu, H. Zeng, Molecular interactions of mussel protective [57] N. Rakkapao, V. Vao-soongnern, Y. Masubuchi, H. Watanabe, Miscibility of
coating protein, mcfp-1, from Mytilus californianus, Biomaterials 33 (2012) chitosan/poly(ethylene oxide) blends and effect of doping alkali and alkali
1903–1911. earth metal ions on chitosan/PEO interaction, Polymer 52 (2011) 2618–2627.

You might also like