You are on page 1of 71

Progress in Polymer Science 36 (2011) 1558–1628

Contents lists available at ScienceDirect

Progress in Polymer Science


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

Polymers for enhanced oil recovery: A paradigm for


structure–property relationship in aqueous solution
D.A.Z. Wever a,b,1 , F. Picchioni a,2 , A.A. Broekhuis a,∗
a
Department of Chemical Engineering - Product Technology, University of Groningen, Nijenborgh 4, 9747 AG Groningen, The Netherlands
b
Dutch Polymer Institute DPI, P.O. Box 902, 5600 AX Eindhoven, The Netherlands

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

Article history: Recent developments in the field of water-soluble polymers aimed at enhancing the aque-
Received 2 June 2010 ous solution viscosity are reviewed. Classic and novel associating water-soluble polymers
Received in revised form 3 May 2011
for enhanced oil recovery (EOR) applications are discussed along with their limitations.
Accepted 6 May 2011
Particular emphasis is placed on the structure–property correlations and the synthetic
Available online 27 May 2011
methods. The observed rheological properties are conceptually linked to the polymer chem-
ical structure (1) and topology (2). In addition, the influence of external parameters, e.g.
Keywords:
Water soluble polymers temperature, pH, salt, and surfactant, on the rheological behavior is reviewed. Progress
Enhanced oil recovery (EOR) booked in deeper understanding of the structure–property relationship is thoroughly dis-
Hydrophobically modified polymers cussed. Furthermore, a critical overview of the synthetic methods as well as of the solution
Polyacrylamide properties of these polymers is provided. In this respect the influence of “internal” (i.e.
Associating polymers chemical structure) and “external” (vide supra) factors on these properties provide a con-
Solution rheology ceptual toolbox for the rationalization of the response of water-soluble polymers to external
stimuli. In turn, such rationalization constitutes the basis for the design of new polymeric
structures for EOR applications.
© 2011 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1560
2. Currently used EOR polymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1561
2.1. Polyacrylamide (PAM) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1561
2.2. Partially hydrolyzed polyacrylamide (HPAM) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1561
2.2.1. HPAM chemical structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1561
2.2.2. Rheological properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1563
2.3. Xanthan gum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1565
3. Recent developments in water-soluble thickening polymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1566
3.1. Hydrophobically modified polyacrylamide (HMPAM) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1567
3.1.1. Chemical structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1581
3.1.2. Rheological properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1583
3.2. Hydrophobically modified ethoxylated urethane (HEUR) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1589
3.2.1. Rheological properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1599
3.2.2. Effects of inherent parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1599

∗ Corresponding author. Tel.: +31 050 363 4918; fax: +31 050 363 4479.
E-mail addresses: D.A.Z.Wever@rug.nl (D.A.Z. Wever), F.Picchioni@rug.nl (F. Picchioni), A.A.Broekhuis@rug.nl (A.A. Broekhuis).
1
Tel.: +31 050 363 4461; fax: +31 050 363 4479.
2
Tel.: +31 050 363 4333; fax: +31 050 363 4479.

0079-6700/$ – see front matter © 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.progpolymsci.2011.05.006
D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628 1559

3.2.3. Effects of external parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1600


3.3. Hydrophobically modified alkali swellable emulsion (HASE) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1602
3.3.1. Rheological properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1603
3.3.2. Effect of inherent parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1608
3.3.3. Effect of external parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1609
3.4. Hydrophobically modified cellulose derivatives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1610
3.4.1. Rheological properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1611
3.4.2. Effect of inherent parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1617
3.4.3. Effect of external parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1617
4. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1619
Acknowledgement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1619
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1619

Nomenclature
CMC critical micelle concentration
4-BA 4-butylaniline CS* critical surfactant concentration
AA acrylic acid CTA chain transfer agent
ABS alkylbenzenesulfonates CTAB cetyltrimethyl ammonium bromide
ACVA 4,4 -azobis-4-cyanopentanoic acid CTAT cetyltrimethyl ammonium p-toluenesul-
AGE alkylglycidyl ether fonate
AM acrylamide Da dalton
AMM associative macromonomer DAAM diacetone acrylamide
AMPDAB 4-(2-acrylamido-2- DADMAC N,N-diallyl-N,N-dimethylammonium
methylpropyldimethylammonio) chloride
butanoate DADPMA [(dimethylammonioethoxy)-
AMPDAC 2-(acrylamido-2- dicyanoethenolate]propyl-
methylpropyldimethylammonium) methacrylamide
chloride DAGE (3,3-dialkoxymethyl)propylglycidyl ether
AMPDAE 2-(2-acrylamido-2- DAMA N,N-diallyl-N,N-methylamine chloride
methylpropyldimethylammonio) DAMAPS 3-(N,N-diallyl-N-
ethanoate methylammonoi)propanesulfonate
AMPDAH 6-(2-acrylamido-2- DEmMA substituted methacrylate
methylpropyldimethylammonio) DiC3 AM N,N-dipropylacrylamide
hexanoate DiC6 AM N,N-dihexylacrylamide
AMPDAPS 3-(2-acrylamido-2-methylpropane- DiC8 AM N,N-dioctylacrylamide
dimethylammonio)-1-propanesulfonate DiC10 AM N,N-didecylacrylamide
AMPTAC 2-(acrylamido)-2- DiC12 AM N,N-didodecylacrylamide
methylpropyl]trimethylammonium DiC14 AM N,N-ditetradecylacrylamide
chloride DiC16 AM N,N-dihexadecylacrylamide
APS N-[(1-pyrenylsulfonamido)ethyl]acrylamide DMSO dimethylsulfoxide
BD bromododecane DR drag reduction
BPAM N-(4-butyl)phenylacrylamide DTAB dodecyltrimethylammonium bromide
CAC critical association concentration DTAC dodecyltrimethylammonium chloride
C8 AM octylacrylamide EФAM N-(4-ethyl-phenyl)acrylamide
C10 AM decylacrylamide EA ethyl acrylate
C12 AM dodecylacrylamide EHEC ethyl hydroxyethyl cellulose
C14 AM tetradecylacrylamide EO ethylene oxide
C18 AM octadecylacrylamide EOR enhanced oil recovery
CN AM alkylacrylamide EP16 1,2-epoxyhexadecane
C12 Acl dodecyl ammonium chloride EVE ethyl vinyl ether
C16 TAAc hexadecyltrimethyl ammonium acetate FX-13 2-(N-ethylperfluorosulfoamido)ethyl acry-
C12 TABr dodecyltrimethyl ammonium bromide late
C12 TACl dodecyltrimethyl ammonium chloride FX-14 2-(N-ethylperfluoro-
C16 TACl hexadecyltrimethyl ammonium bromide octane/sulfoamido)ethyl methacrylate
CAC critical association concentration HASE hydrophobically modified alkali swellable
CDMAO cetyldimethylamine oxide emulsion
1560 D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628

H12 MDI bis(4-isocyanateocyclohexyl)methane PDI polydispersity index


HDI hexamethylene di-isocyanate PEO poly(ethylene oxide)
HEC hydroxyethyl cellulose PGSE pulse gradient spin echo
HEUR hydrophobically modified ethoxylated ure- PMAAM N-phenylmethacrylamide
thane ppm parts per million
HHM-HEC hydrophobically–hydrophilically modi- Py pyrene
fied hydroxyethyl cellulose RRF residual resistance factor
HMPAM hydrophobically modified polyacrylamide S-HEC hydrophilically modified hydroxyethyl cel-
HM-CMC hydrophobically modified carboxymethyl lulose
cellulose S-G HEUR step-growth hydrophobically modified
HM-EHEC hydrophobically modified ethyl hydrox- ethoxylated urethane
yethyl cellulose SANS small-angle neutron scattering
HM-HEC hydrophobically modified hydroxyethyl SEC size exclusion chromatography
cellulose SDS (NaC12 S) sodium dodecyl sulfate
HM-HPC hydrophobically modified hydroxypropyl SMR surfactant to micelle ratio
cellulose SOBS sodium octyl benzene sulfonate
HM-polysaccharides hydrophobically modified THF tetrahydrofuran
polysaccharides TMSPMA 3-[tris(trimethylsilyoxy)sily]propyl
HPAM partially hydrolyzed polyacrylamide methacrylate
HTAC hexadecyltrimethylammonium chloride TTAB tetradecyltrimethyl ammonium bromide
IPBC 4-isopropylbenzyl chloride
IPDI isophorone diisocyanate
KDS potassium dodecyl sulfate
LCST lower critical solution temperature
MA maleic anhydride 1. Introduction
MAA methacrylic acid
MAM methacrylamide Enhanced oil recovery (EOR) is a challenging field for
MeEФAM N-methyl-N-(4-ethyl-phenyl)- different scientific disciplines. The importance of this field
acrylamide is highlighted by the number of patents (mainly filed
MO methyleneoxide by multinational companies) involving polymers for EOR.
MWD molecular weight distribution Nevertheless, the limited number of patents filed in the
NaA sodium acrylate last 10 years (less than 25) demonstrates the difficulty
NaAMB sodium 3-acrylamido-3-methylbutanoate of this research field as well as the relative maturity in
NaAMPS sodium 2-acrylamido-2-methylpropane- the scientific and technological concepts linked to rele-
sulfonate vant applications. Given the fact that the easily recoverable
NaC10 S sodium decyl sulfate oil is running out and that much oil remains in the reser-
NaC14 S sodium tetradecyl sulfate voir after conventional methods have been exhausted, the
NaC12 (EO)2 S sodium dodecyl-di(ethyleneoxide)- implementation of EOR is crucial to guarantee a contin-
sulfate uing supply. In addition, alternative energy sources have
NAEA 2-(1-naphthylacetyl)ethyl acrylate not yet proved to be capable of meeting the world energy
NAEAm 2-(1-naphthylacetamido)ethyl acrylamide demand, so that a mixture of different sources, includ-
NaMAMB sodium 3-methacrylamido-3-methylbu- ing oil, is required to meet the world energy demand
tanoate in the near future. According to Thomas [1], approxi-
NBAM N-benzylacrylamide mately 7.0 × 1012 barrels of oil will remain in oilfields after
NIPAM N-isopropylacrylamide conventional methods have been exhausted; this value
NNDAM N,N-dimethyl acrylamide constitutes also the target recovery for EOR. Water-soluble
nm nanometer polymers for EOR applications have been successfully
NMA [(1-naphthyl)methyl]acrylamide implemented, mainly in Chinese oilfields [2,3]. The pur-
NMR nuclear magnetic resonance pose of the water-soluble polymers in this application is to
NP nonylphenol ethoxylates enhance the rheological properties of the displacing fluid.
NPEAM N-phenethylacrylamide The oil production increases with the microscopic sweep
NRET non-radiative energy transfer of the reservoir and the displacement efficiency of the oil
OG n-octyl ␤-d-glucopyranoside [4]. Indeed, the use of water-soluble polymers improves
OOIP original oil in place the water–oil mobility ratio [4], and leads to enhanced oil
OTG n-octyl ␤-d-thioglucopyranoside recovery. However, given the harsh conditions present in
PAA poly(acrylic acid) most oil reservoirs, new problems and limitations arise
PAGE poly(alkylglycidyl ether) with the use of water-soluble polymers. Besides positively
PAM polyacrylamide affecting solution rheology, water-soluble polymers should
withstand high salt concentration, the presence of calcium,
D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628 1561

high temperatures (>70 ◦ C) and long injection times (at 2. Currently used EOR polymers
least 12 months) [4,5]. High salt concentrations reduce the
thickening capability of most ionic water-soluble polymers 2.1. Polyacrylamide (PAM)
while the presence of calcium leads to flocculation [6].
New water-soluble polymers were successfully tested at Polyacrylamide was the first polymer used as thickening
higher temperatures [7,8]. Associative water-soluble poly- agent for aqueous solutions. The thickening capabil-
mers were tested and showed promising results compared ity (increase of the corresponding solution viscosity) of
to traditionally used polymers [9,10]. PAM resides mainly in its high molecular weight, which
Several studies [11–16] demonstrated that the oil is pro- reaches relatively high values (>1 × 106 g/mol). In the gen-
duced faster (compared to water flooding), but also more eral framework of EOR processes, PAM is mainly used
oil can be recovered. A mechanism based on the visco- as the reference “model system” for chemical modifi-
elastic properties of the polymers has been proposed to cation. Many authors have reported different attempts
explain the higher oil recovery. This mechanism is mainly to alter the chemical structure of PAM or to synthe-
supported by indirect evidence and mathematical models size new acrylamide-based copolymers with improved
[11–14]. properties, i.e. shear resistance, brine compatibility and
Independently of the exact displacement mechanism temperature stability [22–25]. The synthesis of the copoly-
and efficiency, the use of water-soluble polymers for EOR mer N,N-dimethyl acrylamide with Na-2-acrylamido-2-
still constitutes a challenging field, at both industrial and methylpropanesulfonate (NNDAM–NaAMPS) was reported
academic levels. Moreover, the broad variety of polymers by Sabhapondit et al. [22,23] and tested for its perfor-
studied for EOR, at least at the academic level, clearly mance in EOR applications. The stability of the polymer at
suggests this research field as paradigmatic for water- high temperature was demonstrated by ageing at 120 ◦ C
soluble polymers in general. Indeed, fundamental scientific for 1 month [22]. By using a sand pack, an improved
studies linked to EOR, particularly those involving the performance in terms of EOR for the NNDAM–NaAMPS
relationship between the polymer structure and the cor- copolymer [23] as compared to unmodified partially
responding properties in aqueous solution (e.g. viscosity), hydrolyzed polyacrylamide, HPAM, was demonstrated.
have a conceptual character, thus providing a broad and In another example, Song et al. [24] reported the syn-
general understanding, valid also for other “application” thesis of starch-graft-poly(acrylamide-co-(2-acrylamido-
fields. The interaction between polymer chains in aque- 2-methylpropanesulfoacid)). The oil recovery rate of the
ous solution as well as the influence of the solute structure subsequent polymer solution was higher compared to
on the corresponding viscosity are of paramount impor- HPAM, and the novel polymer displayed better temper-
tance in EOR applications, but are studied and modeled ature and shear stability. These two examples already
through very general methods and theories. This general define a common research theme in the general field
interest in structure–property relationships is testified by of water-soluble polymers for EOR. That is, a strategic
the constantly increasing amount of scientific literature approach involving the chemical modification of commer-
that is dedicated to this topic [17–20], the last dedicated cial polymers (in this case PAM) to tailor and improve the
review (from 1990 [21]) being already outdated by the corresponding solution properties and eventually EOR per-
most recent findings. As a consequence, the present review formance.
paper concentrates on different water-soluble polymers
for EOR by discussing their (associative) behavior in water 2.2. Partially hydrolyzed polyacrylamide (HPAM)
solution and outlining the most general concepts that
can be learned from the corresponding scientific litera- HPAM, by far the most used polymer in EOR applica-
ture. tions, is a copolymer of PAM and PAA obtained by partial
Section 2 presents some background information on hydrolysis of PAM or by copolymerization of sodium acry-
oil recovery methods, thus defining the subject in terms late with acrylamide [21].
of polymeric systems reviewed in this work. In Section
3 the synthetic methods used for the different polymers 2.2.1. HPAM chemical structure
are discussed, including the effect of the relationship The chemical structure of HPAM is provided in Fig. 1.
between the process conditions employed and the cor- In most cases the degree of hydrolysis of the acrylamide
responding polymer chemical structure. Subsequently, monomers is between 25 and 35% [4,26]. The fact that
the chemical structures are linked to the corresponding a relevant fraction of the monomeric units needs to be
rheological properties in water solution. The sensitiv- hydrolyzed (lower limit of 25%) is probably related to the
ity of the rheological behavior is also elucidated as a
function of external parameters, such as (the presence
of) mechanical shear, solution temperature, electrolyte
type and concentration, pH, ionic strength and sur-
factant type and concentration. Relations between the
polymer structure, i.e. chemical structure and topol-
ogy, are proposed for the different polymers. Finally,
in Section 4 we provide some general conclusions on
the most recent, relevant and generally accepted con-
cepts. Fig. 1. Chemical structure of HPAM.
1562 D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628

formation of the corresponding salt. According to the gen- has been presented as a solution to the salt sensitivity of the
eral theory of polyelectrolyte solutions [27], the presence of HPAM [36,37], although the exact reasons for this behavior
electrostatic charges along a polymer backbone is respon- have not been fully elucidated.
sible for prominent stretching (due to electric repulsion) The addition of monovalent NaCl leads to a reduction in
of the polymeric chains in water and, eventually, results the level of aggregation. However, at higher ionic strengths
in a viscosity increase compared to the uncharged ana- (higher salt concentration) the addition of NaCl leads to
logue. On the other hand, according to a study by Shupe macroscopic flocculation [38]. It has also been demon-
[28] the degree of hydrolysis cannot be too high because strated that multivalent cations can form polyion–metal
the polymer solution will become too sensitive to salinity complexes which affect the viscosity of the resulting solu-
and hardness of the brine (electrolytes present in solution tion [39–41]. A study by Peng and Wu [39] investigated the
have a “shielding effect” on the electrostatic repulsion). dependence of the self-complexation of HPAM on the Ca2+
Indeed, polyelectrolytes, i.e. polymers bearing charges, concentration and the degree of hydrolysis of HPAM. They
show significantly different rheological behavior compared demonstrated that depending on the Ca2+ concentration
to their neutral analogues [29–31]. The thickening capabil- intrachain and interchain complexations take place (Fig. 2)
ity of HPAM lies in its high molecular weight and also in the [39].
electrostatic repulsion between polymer coils and between Besides the salt dependency, other factors influencing
polymeric segments in the same coil [4]. When poly- the viscosity of HPAM solutions are the degree of hydrol-
electrolytes are dissolved in water containing electrolytes ysis, solution temperature, molecular weight and solvent
(salts) a reduction in viscosity is observed [26,32–34]. It quality [35]. Pressure also affects the viscosity of HPAM
has been demonstrated that the specific viscosity of HPAM solutions. According to a study by Cook et al. [42] the
solutions depends on the amount of salt present [35]. This increase in the viscosity of the HPAM solutions cannot
effect is attributed to the shielding effect of the charges solely be accounted for by the increase in viscosity of the
[4,33] leading in turn to a reduction in electrostatic repul- solvent. The intrinsic viscosity and the radius of gyration
sion and consequently to a less significant expansion of the are both invariant with pressure, albeit with a 10% exper-
polymer coils in the solution. This results in a relatively imental uncertainty [42]. In principle, the dimensions of
lower hydrodynamic volume, which is synonymous with a the polymer coils do not change while the solvent volume
lower viscosity [34]. A few decades ago, substitution of one decreases. Therefore the volume fraction of the polymer
or both hydrogens on the amide nitrogen with alkyl groups coil per unit volume of the solvent increases, hence a higher

Fig. 2. Complexation behavior of HPAM under different conditions.


Reproduced with permission from [39] 1999, ACS Publishing.
D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628 1563

Fig. 3. Schematic presentation of behavior of HPAM coils in shear flow.


Reproduced with permission from [46] 1995, ACS Publishing.

viscosity [42]. Another parameter that affects the solution fore the entanglements, related to the solution viscosity,
viscosity of the polymer solution is shear [43]. Under high will be conserved as the shear increases.
shear the HPAM polymer chains are reduced in size due The shear thickening behavior has been attributed to
to chain scission, i.e. fragmentation [44]. This leads to a changes in the molecular conformation involving the for-
reduction in the solution viscosity. mation of additional links between two chains [49]. The
shear thickening behavior is observed both in laboratory
2.2.2. Rheological properties rheometers [45] (in pure water and aqueous salt solu-
HPAM is preferred in EOR applications since it can toler- tions) and in porous media. According to several studies
ate the high mechanical forces present during the flooding the shear thickening behavior in porous media arises due
of a reservoir. In addition, HPAM is a low cost polymer to coil-stretched transitions of the polymer chains. The
and is resistant to bacterial attack [4]. Although the HPAM structure of a porous medium can be seen as alternating
solutions display pseudoplastic behavior [4,26,32,45,46] wide openings and confined throats through which the
(shear thinning) in simple viscometers, it has been demon- polymer coils have to navigate. In the wide openings the
strated that these solutions show pseudodilatant [47,48] polymer chains attain a coil structure. When these coils
characteristics (shear thickening) in porous media as well then have to pass through a narrow throat the polymer
as in viscometers at relatively high shear rates. Research coils are forced to deform and stretch (elongational strain
has demonstrated the presence of a critical shear rate at [32,48,52]) in order to pass. This successive contraction
which the shear thickening behavior arises in viscometers and expansion of the polymer coils leads to pseudodilatant
[32,33,45,46,49,50]. This critical shear rate depends on the behavior of the polymer solutions [48,53,54]. This con-
degree of hydrolysis of the HPAM, the solution concentra- formational change of the macromolecules is reversible
tion, the temperature, the quality of the solvent and also on since it is commonly explained by formation at macro-
the molecular weight of the polymer [33,45]. An increase molecular level of reversible interactions like hydrogen
in the degree of hydrolysis leads to an onset of shear thick- bonding. Indeed, it is believed that hydrogen bonding arises
ening at lower shear rates [45]. By decreasing the average for HPAM solutions between the carboxylic functional-
molecular weight, an increase in the polymer concentra- ities [55]. However, this is contested due to conflicting
tion results in a higher critical shear rate [45,46]. data [49,56] on similar polymeric solutions (e.g. dextran
The aforementioned shear thinning of HPAM solutions solution). Instead, aggregation of hydrophobic bonds has
below a critical shear rate arises due to uncoiling of poly- been proposed [55], albeit in polymethacrylic acid, but this
mer chains and the dissociation of entanglements between has not been confirmed [57]. Hu et al. [46] proposed a
separate polymer coils [4]. Stiffening of the polymer back- schematic presentation depicting the essential behavior of
bone has been suggested as a possible approach to control HPAM solutions in shear flow (see Fig. 3).
the dependency of HPAM polymer solutions on the shear Another behavior that has been identified for HPAM
[51]. A stiff polymer will have a lower mobility and there- solutions which is important for EOR is their negative

Fig. 4. Type I and II of rheopectic behavior of HPAM solutions.


Reproduced with permission from [58] 1995, Springer.
1564 D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628

Fig. 5. Interactions between sodium oleate and HPAM at low (A) and high (B) surfactant concentration.

thixotropic (rheopectic) property, i.e. an increase in vis- mono-valent cations is amplified, i.e. the effect is seen at
cosity with shear-time at a constant shear rate [32,58–61]. lower cation concentrations [61]. This is due to the higher
Researchers have identified two different types of rheopec- screening capability of multivalent cations [61]. In addi-
tic behavior for HPAM solutions (Fig. 4), type I and type tion a reduction of the degree of hydrolysis leads to a less
II [58]. The type I effect is observed at low shearing and prominent rheopectic behavior of type I [60].
consists in a slow viscosity increase with shear-time up Besides salts, also surfactants are able to interact with
to an asymptotic value. The type II effect is seen at high polymer chains in solutions and thus display a relevant
shear rates and is displayed as a steep viscosity increase influence on the corresponding rheological behavior. Xin
after a given shear-time, followed by pronounced viscosity et al. [63] investigated the interaction between the surfac-
oscillation [58]. tant sodium oleate and HPAM and found that the viscosity
of the polymer–surfactant aqueous solution depends on
2.2.2.1. Effect of inherent parameters on rheology. For HPAM the surfactant concentration. At low surfactant concentra-
the degree of hydrolysis has a significant impact on the tion an enhancement of the viscosity is observed due to
rheological properties of the subsequent solution. If the interpolymer cross-linking of the surfactant and HPAM. At
hydrolysis degree of HPAM is too large, insolubility prob- high surfactant concentration the repulsion between the
lems can arise. When too small, a large dependence of micellar aggregates attached to the polymer increases and
the solution viscosity on electrolyte presence is observed. this leads to a decrease in the hydrodynamic volume and
Lewandowska [45] investigated the effect of the degree of thus a decrease in the solution viscosity. The authors [63]
hydrolysis on the solution viscosity in a NaCl solution. If have proposed a model depicting the interactions between
the degree of hydrolysis is increased the zero shear rate HPAM and sodium oleate (a surfactant) at low and high
viscosity and the critical shear rate for the onset of shear concentration of the latter (Fig. 5).
thickening are both reduced. The shear rate region where Interchain cross-linking arises due to hydrogen bond-
the shear thinning behavior is observed is reduced with the ing. At high surfactant concentration repulsion between
increase in the degree of hydrolysis [45,46,62]. the surfactant molecules dominates and this leads to
Another inherent parameter of HPAM is its molecular the collapse of the network between the surfactant and
weight. Increase of the molecular weight will lead to a more HPAM, which in turn results in a reduction of the vis-
pronounced shear thinning behavior. The critical shear rate cosity. Methemitis et al. [64] investigated the interactions
for the onset of shear thickening is also affected by the between SDS and HPAM and concluded that the effect of the
molecular weight in that it is increased with an increase surfactant on the rheological properties of HPAM depends
in the molecular weight [45]. on the pH of the solution and the presence of electrolyte.
When no effort is devoted at altering the pH of the solu-
2.2.2.2. Effect of external parameters on rheology. The influ- tion, a reduction in the solution viscosity is observed up
ence of salt (NaCl) addition on the rheological behavior to the CMC of the surfactant either in water or in salty
of HPAM solutions has also been extensively studied. It water. Above the CMC the solution viscosity of the ternary,
was found that, below the critical shear rate, adding salt system, i.e. water–polymer–surfactant, remains relatively
reduced the extent of shear thinning while above the crit- constant with further addition of SDS in the salty water. In
ical shear rate the amplitude of the shear thickening is pure water the solution viscosity still decreases with fur-
increased [32]. The zero shear rate solution viscosity is also ther addition of SDS. If the pH of the solution is reduced (to a
reduced as the concentration of salt increases [4]. pH of 2.5) the solution viscosity decreases until the CMC of
Adding salt to the solvent will reduce the extent of the the surfactant is reached, after which a significant increase
rheopectic behavior of the polymer solution [59]. Looking is observed with further addition of SDS. The authors have
more closely to the effect of salt, addition of mono-valent hypothesized that at low pH fixation of some protons onto
cations (e.g. NaCl) was found to increase the onset (i.e. the the surfactant micelles occurs [64]. This changes the ion-
critical shear rate value) of rheopectic behavior type I while ization equilibrium of the carboxylic groups leading to an
for type II no changes were observed [58]. When using increase in the surface charge density, which corresponds
multivalent cations (CaCl2 and AlCl3 ) the effect seen with to an increase in solution viscosity.
D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628 1565

Fig. 6. Chemical structure of xanthan gum.

2.3. Xanthan gum tions employing xanthan gum display high viscosity at
low shear rates [74] and thus the disordered conforma-
Xanthan gum is a polysaccharide, which is produced tion predominates at low shear rates. At high shear rates
through fermentation of glucose or fructose by different both conformations display similar rheological behaviors
bacteria [65]. The most efficient xanthan gum producer is [73]. In addition, pseudoplastic behavior is observed for
the Xanthomonas campestris bacterium [65,66]. The chem- the polymer solutions [75]. Unlike HPAM, xanthan gum
ical structure of xanthan gum (Fig. 6) displays the presence displays good resistance to high temperatures. It was
of two glucose units, two mannose units and one glucuronic demonstrated that the solution viscosity of a polymeric
acid unit with a molar ratio of 2.8–2.0–2.0 [67]. The back- solution employing a commercial xanthan gum remained
bone of xanthan gum is similar to cellulose. The side chains relatively constant for more than 2 years at 80 ◦ C [76]. Loss
of the polymer contain charged moieties, i.e. acetate and of solution viscosity occurs at temperature above 100 ◦ C.
pyruvate groups, and the polymer is thus a polyelectrolyte. Several studies [77–80] have investigated the temperature
However the classic polyelectrolyte behavior according to dependence of the apparent viscosity of xanthan gum solu-
which the solution viscosity decreases with the addition of tions. In order to display resistance to temperatures up
salt is not displayed in this case. The thickening capability to 90 ◦ C, the conventional understanding for xanthan gum
of xanthan gum lies in its high molecular weight, which solutions is that the ionic strength of the solution has to be
ranges from 2 to 50 × 106 g/mol [67,68] and in the rigidity relatively high. Another positive property of xanthan gum
of the polymer chains. is its ability to withstand high shear forces. Unlike HPAM
It has been demonstrated that upon addition of salt the solution viscosity does not decrease at relatively high
(mono or divalent) the xanthan gum chains undergo a shear stresses [47]. Especially the ordered structure, i.e. in
cooperative conformational transition from a disordered the presence of salt, can withstand high shear forces [73]
conformation to an ordered and more rigid structure (up to a shear rate of 5000s−1 ).
[69–72] (Fig. 7). A disadvantage of xanthan gum is its susceptibility
The temperature and the ionic strength (the amount to bacterial degradation. It has been demonstrated that
of electrolyte) of electrolyte, of the solution are triggers salt tolerant aerobic and anaerobic microorganisms can
for the conformational transition. When testing at low degrade the xanthan gum chains which leads to the loss
shear, the rheology of the polymer solution is dependent in solution viscosity [81–84]. Biocides are used to suppress
on the conformation with the disordered conformation the growth of the xanthan gum degrading microorganisms.
displaying higher solution viscosities [73]. Polymeric solu- In most cases formaldehyde is the most efficient biocide

Fig. 7. Conformational transition of xanthan gum.


1566 D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628

[83,84]. However the use of biocides to protect the xanthan can increase the viscosity of the aqueous solution contain-
gum renders the low environmental impact of the polymer ing hydrophobically modified polymers significantly [90].
obsolete. Other polymers that posses interesting properties,
Combinations of xanthan gum with surfactants have such as high molecular weight and intrinsic viscosity,
also been studied. It has been demonstrated that the com- have been developed for EOR and are known as “rigid
bination can be beneficial. According to Taugbol et al. [85] rod” water-soluble polymers [93]. One study compared
more than 50% of the residual oil (after a waterflood) can be hydrophobically modified polyacrylamide (HMPAM) with
recovered using xanthan gum and an alkyl propoxy-ethoxy polyacrylamide (PAM) in a simple core flood test and
sulfate (C12–15 (PO)4 -(EO)2 -OSO3 − Na+ ) as the surfactant. demonstrated that the residual resistance factor (RRF) after
The recovery of the residual oil using only a surfactant solu- the polymer flood is much higher for the HMPAM compared
tion was lower. However it has been demonstrated that to PAM [94]. All these modification strategies, together
the combination of xanthan gum and dodecyl-o-xylene with new kinds of water-soluble systems, have been exten-
sulfonate recovered less of the residual oil compared to sively reported in the literature and will be discussed in the
the case where only a surfactant solution was used [86]. next paragraph.
According to the authors a possible explanation is the for- As mentioned earlier, a relatively new class of water-
mation of large micellar aggregates, which have a negative soluble polymers is the one constituted by hydrophobically
effect on the flow performance of the surfactant through associative polymers [87]. The first hydrophobically asso-
the porous media. ciative polymers were synthesized almost fifty years ago
[95,96], albeit for a different purpose than EOR. Indeed, the
research on these types of polymers has been primarily
3. Recent developments in water-soluble fueled by the coating industry [87], where improvement
thickening polymers in the rheology of the coating systems was required. Dur-
ing the 1980s when the oil crisis hit, a lot of research was
The limited number of available commercial poly- performed on EOR. From the many patents [97–102] that
mers currently employed in EOR has been the subject have been filed during those years it is evident that this
of recent developments aimed at improving their perfor- accelerated the development of hydrophobically associa-
mance. Indeed, an alternative concept has been studied in tive polymers for use in EOR applications.
the last four decades, and involves the association between Hydrophobically associative polymers contain, in
hydrophobic groups that are incorporated in the back- most cases, a small number of hydrophobic groups,
bone of the polymers [87]. Through these associations a i.e. 8–18 carbon atoms moieties [103–106], distributed
higher thickening capability can be achieved compared to along the main backbone [20,107,108]. These hydropho-
the traditional polymers [87]. Several different types of bic groups can be distributed randomly or block-like
associating polymers have been studied. These include the [88,91,103,107,109–120], and coupled at one or both
hydrophobically modified polyacrylamide (HMPAM) [88], ends [104,121–131]. Above a given polymer concentration
ethoxylated urethane (HEUR) [89], hydroxyethylcellulose (dependent on the molecular structure) the hydropho-
(HMHEC) [90] and alkali-swellable emulsion (HASE) [91]. bic groups associate, when the polymer is dissolved
Also combinations of associative polymers with surfactants in water, to form hydrophobic micro-domains (intra or
have been developed for EOR [92]. It has been demon- intermolecular liaisons) [88,89,104,105,107,108,132–135].
strated that the addition of small amounts of surfactants These lead to an increase in hydrodynamic volume, which

Fig. 8. Intra- and intermolecular associations.


Reproduced with permission from [132] 1996, ACS Publishing.
D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628 1567

in turn yields a polymer with a much better thickening 3.1. Hydrophobically modified polyacrylamide (HMPAM)
(higher viscosity [107]) capability compared to its non-
associative analogue [88]. Depending on the concentration, HMPAM constitutes the most popular basic structure
intra- or intermolecular associations are formed, which are for the synthesis of new water-soluble polymers for EOR.
schematically illustrated in Fig. 8. Indeed, as given in Table 1, many types have been pub-
When the hydrophobic elements are distributed in a lished.
block-like fashion along the backbone of a water-soluble There are different methods of synthesizing HMPAM
copolymer, the intramolecular associations are stronger such as micellar [88,113,135], homogeneous [88,248,249]
compared to randomly or discretely distributed hydropho- and heterogeneous [88] copolymerization. Polyacrylamide
bic groups [104,121]. is usually prepared via a free radical polymerization in
The temperature dependence of the solution viscos- aqueous solution [88,250]. However, as is evident from
ity is an interesting property of hydrophobically modified the name, HMPAM cannot be synthesized using this tech-
polymers for EOR applications. It has long been accepted nique as the hydrophobic monomer is not soluble in
that increasing the temperature of the polymer solution water. In order to disperse the hydrophobic monomer,
will lead to a reduction in viscosity [103,133,136–140]. it is dissolved using a co-solvent (homogeneous copoly-
This since an increase in temperature implies a decrease merization) or a surfactant (micellar copolymerization) or
of the association strength of the hydrophobes. Increas- dispersed without any additives (heterogeneous copoly-
ing the temperature of the solution leads to a reduction merization) [88,135].
of the solvent viscosity and hence an increase in the Among these different methods, micellar copolymeriza-
mobility of the polymer chains while the solubility of the tion has been mostly studied. It generally involves the use
polymer will increase with temperature. However, many of a hydrophobic monomer (soluble in micelles stabilized
different aqueous systems have been demonstrated to by a surfactant) and a hydrophilic one which is soluble
display an increase in viscosity upon increasing the tem- in the water phase. The first reports on micellar copoly-
perature [141–154]. Indeed, a temperature increase will merization appeared simultaneously some 25 years ago
decrease the solubility of one of the components (lower [251–256]. A schematic presentation of this polymeriza-
critical solution temperature [LCST]-groups) of the poly- tion technique is presented in Fig. 10.
mers. These less soluble components will self-aggregate When synthesizing polymer 17 (Table 1) Chang
with the hydrophobic groups of the polymers, which leads and McCormick [220] noted that the use of micellar
to an increase in viscosity [119]. Several researchers have copolymerization leads to a block-like distribution of the
proposed a concept for thermo-associative polymers based hydrophobes whereas solution copolymerization leads
on the switch, i.e. the transition between low and high to a random distribution. Both polymers exhibit com-
temperature, of the polymers characterized by a lower pletely different rheological properties. The same behavior
critical solution temperature [140,151,152]. The concept was observed when synthesizing polymer 29 (Table 1)
involves a highly water-soluble polymer containing blocks using both (micellar and solution) techniques [110,111].
or side chains of LCST groups. Upon heating of the polymer The most popular surfactants in micellar copolymeriza-
solution, these LCST groups will segregate. A schematic pre- tion are sodium dodecyl sulfate (SDS) [92,103,104,109,
sentation of this behavior has been presented by Hourdet 112,121,134,135,138,173–179,203,229,230,241,257–260]
and coworkers [151] and is depicted in Fig. 9. and cationic hexadecyltrimethylammonium bromide
Above the critical overlap chain concentration this tran- (CTAB) [113,133,220,238]. Many different hydropho-
sition will lead to an increase in the viscosity of the solution bic monomers have been used, such as acrylate or
through intermolecular associations. methacrylate-derivatives, alkyl groups with vary-
Fundamental research on different polymers, in binary ing number of carbons and different topologies
(polymer–water) and ternary (polymer–water–surfactant) [99,100,103,229,261–264], aryl or alkylaryl functionalities
systems, has been performed using different techniques [88,110,111,113,173–177,230,260,265–268], fluorocarbon
which include 13 C NMR [155–158] (solution or solid- containing agents [132,241,242,269–271] and zwitterionic
state), 1 H NMR [109], 23 Na NMR [159,160], 19 F NMR groups [196–199,201–203,234,272–274]. By incorporating
[161], NMR self-diffusion [126,162–165] potentiome- water-soluble spacers between the hydrophilic backbone
try [166–168], static and dynamic laser light scatter- and the hydrophobic group an enhancement of the viscos-
ing [128,130,131,146,163,167–172], UV-spectroscopy for ity can be achieved compared to systems without spacers
polymers bearing chromophores [88,110,111,173–181], [199,241]. According to Hwang and Hogen-Esch [241] the
small-angle neutron scattering (SANS) [182], non-radiative formation of hydrophobic micro-domains is effectively
energy transfer (NRET) studies [178,183–185], size exclu- promoted by increasing the lengths of the spacers.
sion chromatography (SEC) [170,186–189] and surface Parameters affecting the properties of the polymers
tension [131,133,162,163,165]. prepared by micellar copolymerization are the type
Several different associative hydrophobically modified and concentration of the hydrophobic and hydrophilic
polymers have been developed which include polyacry- monomer [109,174,175,258], the molar ratio between
lamides (HMPAM), ethoxylated urethanes (HEUR), alkali the two monomers, the content and type of surfactant
swellable emulsions (HASE), and polysaccharides (HM- [174,275], the content and type of initiator and the tem-
polysaccharides). Their synthesis, rheological behavior and perature of the reaction [135]. Another parameter that has
adsorption on surfaces will be discussed in the following been identified is the molar ratio between the hydrophobic
sections. monomer and surfactant, i.e. the number of hydrophobic
1568
Table 1
Structure of different water-soluble polymers, HMPAM.

Polymer Structure References

Polyelectrolyte

Copolymer of acrylamide (AM) and sodium 2-acrylamido-2-methylpropanesulfonate (NaAMPS, R = 2) [190–193]


or sodium 3-acrylamido-3-methylbutanoate (NaAMB, R = 1) or
(2-acrylamido-2-methylpropyl)dimethylammonium chloride (AMPDAC, R = 3)

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628


Polyelectrolyte

Copolymer of acrylamide (AM) and sodium 3-methacrylamido-3-methylbutanoate (NaMAMB) [194,195]

Polyelectrolyte, zwitterionic monomer

Copolymer of 3-(2-acrylamido-2-methylpropane-dimethylammonio)-1-propanesulfonate [196]


(AMPDAPS) with 4-(2-acrylamido-2-methylpropyldimethylammonio) butanoate (AMPDAB)

Polyelectrolyte, zwitterionic monomer

Copolymer of acrylamide (AM) with [197]


3-(2-acrylamido-2-methylpropane-dimethylammonio)-1-propanesulfonate (AMPDAPS)
Table 1 (Continued)

Polymer Structure References

Polyelectrolyte, zwitterionic monomer

Copolymer of AM with 2-(2-acrylamido-2-methylpropyldimethylammonio) ethanoate (AMPDAE, [198–200]


N = 1) or AMPDAB, N = 3 or 6-(2-acrylamido-2-methylpropyldimethylammonio) hexanoate
(AMPDAH, N = 5)

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628


Polyelectrolyte, zwitterionic monomer

Terpolymer of AM with acrylic acid (AA) and AMPDAPS [201]

Polyelectrolyte, zwitterionic monomer

Terpolymer of AM with sodium acrylate (NaA) and AMPDAB [202]

Polyelectrolyte, zwitterionic monomer

Copolymer of AM with [(dimethylammonioethoxy)dicyanoethenolate]propyl-methacrylamide [203]


(DADPMA)

1569
1570
Polyelectrolyte, zwitterionic backbone

Terpolymer of AM and AMPDAC (R1 = H) with NaAMPS (R2 = 1) or AMPTAC (R1 = CH3 ) with NaAMPS [204–207]
(R2 = 1) or AMPTAC (R1 = CH3 ) with NaAMB (R2 = 2)

Polyelectrolyte, zwitterionic backbone

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628


Terpolymer of AM with AMPDAC and sodium 3-acrylamido-3-methylbutanoate (NaAMB) [208]

Polyelectrolyte, zwitterionic backbone

Copolymer of 2-acrylamido-2-methylpropanesulfonate (NaAMPS) with [209–211]


(2-acrylamido-2-methylpropyl)-dimethylammonium chloride (AMPDAC, R = H) or
[2-(acrylamido)-2-methylpropyl]trimethylammonium chloride (AMPTAC, R = CH3 )

Polyelectrolyte, amphiphilic

Copolymer of AM (R1 = NH2 ) and sulfonate containing monomer (R2 = 1) [115,212,213]


Table 1 (Continued)

Polymer Structure References

Copolymer of NaA (R1 = O− Na+ ) and alkylchains (R2 = 2) or C8F (R2 = 3) or C10F (R2 = 4) or 3-PDCA
(R2 = 5)
Polyelectrolyte, amphiphilic

Copolymer of NaA and [156,214–216]

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628


A, Dodecylacrylamide (C12 AM), R = NH, M = 11, N = 0 or
A, Octadecylacrylamide (C18 AM), R = NH, M = 17, N = 0 or
B, Substituted methacrylate (DEmMA), N = 2, 6 or 25) or
Copolymer of NaAMPS and
B, Substituted methacrylate (DEmMA), N = 2, 6 or 25)

Polyelectrolyte, amphiphilic

Terpolymer of AM (R1 = NH2 ) and AA (R2 = OH) with [175,176,217]

N-[(hexyl)phenyl]acrylamide (R3 = 2, N = 5) or
N-[(decyl)phenyl]acrylamide (R3 = 2, N = 10) or
Terpolymer of AA (R2 = OH) and NaAMPS (R2 = 1) with
PEO chain (R3 = 4) or
Terpolymer of AM (R1 = NH2 ) and NaAMPS (R2 = 1) with
N,N-Dihexylacrylamide (DiHexAM, R = 5 and M = 5)

1571
1572
Polyelectrolyte, amphiphilic

Terpolymer of AA and methacrylamide (MAM) with [185,218]

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628


(DiC6 AM, N = 4) N,N-dihexylacrylamide or
(DiC8 AM, N = 6) N,N-dioctylacrylamide or
(DiC10 AM, N = 8) N,N-didecylacrylamide or
(DiC12 AM, N = 10) N,N-didodecylacrylamide or
(DiC14 AM, N = 12) N,N-ditetradecylacrylamide or
(DiC16 AM, N = 14) N,N-dihexadecylacrylamide

Polyelectrolyte amphiphilic

Terpolymer of maleic anhydride (MA), ethyl vinyl ether (EVE) and 4-butylaniline (4-BA) [219]

Polyelectrolyte, amphiphilic

Copolymer of AM (R1 = NH2 ) and dimethyldodecyl(2-acrylamidoethyl)ammonium bromide (DAMAB, [115,220,221]


R2 = 3, R3 = H and M = 2) or dimethyldodecyl(2-methacrylamidopropyl)ammonium bromide
(DMAMAB, R2 = 3, R3 = CH3 and M = 3) or

Copolymer of AA (R1 = OH) and 2-ethylhexyl (R2 = 1) or n-alkyl (R2 = 2)


Table 1 (Continued)

Polymer Structure References

Polyelectrolyte, amphiphilic

Copolymers of N-isopropylacrylamide (NIPAM, R2 = 1) or AA (R2 = OH) with [221–225]

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628


2-(N-ethylperfluorosulfoamido)ethyl acrylate (FX-13), R1 = H or
2-(N-ethylperfluoro-octane/sulfoamido)ethyl methacrylate (FX-14), R1 = CH3

Polyelectrolyte, zwitterionic monomer and amphiphilic

Copolymer of 3-(N,N-diallyl-N-methylammonio)propane-sulfonate (DAMAPS, R2 = 1) with [226–228]


N,N-diallyl-N,N-dimethylammonium chloride (DADMAC, R1 = CH3 ) or N,N-diallyl-N-methylamine
chloride (DAMA, R1 = H)

Copolymer of DADMAC, (R1 = CH3 ) with


N,N-Diallyl-N-hexylbeznyl-N-methylammonium chloride (R2 = 2) or
N,N-Diallyl-N-octylbeznyl-N-methylammonium chloride (R2 = 3)

Polyelectrolyte, zwitterionic monomer and amphiphilic

Terpolymer of AM and n-decylacrylamide (C10 AM) with [229]

1573
1574
NaA, R = O− Na+ or
NaAMB, R = C7 H17 N2 O2 − Na+ or
NaAMPS, R = C4 H9 NSO3 − Na+

Polyelectrolyte, zwitterionic monomer and amphiphilic

Terpolymer of AM and N-(4-butyl)phenylacrylamide (BPAM) with [230]

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628


NaA, R = O− Na+ or
NaAMB, R = C7 H17 N2 O2 − Na+ or
NaAMPS, R = C4 H9 NSO3 − Na+

Polyelectrolyte, zwitterionic monomer and amphiphilic

1. Terpolymer of sulfur dioxide, N,N-diallyl-N-carboethoxymethylammonium chloride with [231–234]


N,N-diallyl-N-alkylammonium chloride (R2 = 1) or dendritic quadruple-tailed hydrophobic group
(R2 = 2)
Table 1 (Continued)

Polymer Structure References

2. Terpolymer of sulfur dioxide, 3-(N,N-diallylammonio) propanesulfonate with


N,N-diallyl-N-octadecylammonium chloride
3. Terpolymer of sulfur dioxide N,N-diallyl-N-carboethoxymethylammonium chloride with single-,
twin- and triple-tailed hydrophobic groups (A, B or C)
Amphiphilic

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628


Copolymer of AM (R1 = H) and [103,235]

Octylacrylamide (C8 AM, R2 = 1), N = 7 or


Decylacrylamide (C10 AM, R2 = 1), N = 9 or
Dodecylacrylamide (C12 AM, R2 = 1), N = 11 or
t-octylacrylamide (R2 = 2)
n-decylacrylamide (R2 = 3 or 4)
Copolymer of N-isopropylacrylamide (NIPAM, R1 = CH3 ) and
Decylacrylamide (C10 AM, R2 = 1), N = 9 or
Tetradecylacrylamide (C14 AM, R2 = 1), N = 13 or
Octadecylacrylamide (C18 AM, R2 = 1), N = 17

Amphiphilic

Copolymer of AM and [109,138,190,191,236,237]

N-octylacrylamide (OctAM, R = (CH2 )7 –CH3 ) or


Diacetone acrylamide (DAAM, R = 1) or

1575
1576
N-hexylacrylamide (HexAM, R = 2) or
N-methyl-N-hexylacrylamide (MeHexAM, R = 3) or
N,N-dihexylacrylamide (DiHexAM, R = 4 and N = 5) or
di-n-propylacrylamide (DPAM, R = 4 and N = 2) or
di-n-octylacrylamide (DOAM, R = 4 and N = 7) or
N-(4-ethyl-phenyl)acrylamide (EAM, R = 5) or

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628


N-methyl-N-(4-ethyl-phenyl)-acrylamide (MeEAM, R = 6) or
BPAM, R = 7

Amphiphilic

Copolymer of AM and [133,238,239]

N-benzylacrylamide(NBAM, R1 = H and R2 = (CH2 )1 )


N-phenethylacrylamide(NPEAM, R1 = H and R2 = (CH2 )2 )
N-phenylmethacrylamide(PMAAM, R1 = CH3 and R2 = NH)

Amphiphilic

Copolymers of acrylic acid (AA) and [240]

3-[tris(trimethylsilyoxy)sily]propyl methacrylate (TMSPMA)


Table 1 (Continued)

Polymer Structure References

Amphiphilic

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628


Copolymer of AM and [132,241–243]

1,1-Dihydroperfluorobutyl acrylate (2a) or


1,1-Dihydroperfluorooctyl acrylate (2b) or
2-(N-ethylperfluorooctanesulfonamido)ethyl acrylate, FOSA (3a) or
2-(N-ethylperfluorooctanesulfonamido)ethyl methacrylate, FOSM (3b) or
1,1-Dihydroperfluorooctylmono-(ethyleneoxy) acrylate (4a) or
1,1-Dihydroperfluorooctylbis-(ethyleneoxy) acrylate (4b) or
1,1-Dihydroperfluorooctyltris-(ethyleneoxy) acrylate (4c) or
Dodecyl acrylate (5a) or
Dodecylmono-(ethyleneoxy) acrylate (5b) or
Dodecylbis-(ethyleneoxy) acrylate (5c) or
Dodecyltris-(ethyleneoxy) acrylate (5d) or

1577
Poly(propylene oxide) methacrylate (6)
1578
Amphiphilic

1. Copolymer of AM with [104,244,245]

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628


A monosubstituted monomer derived from 4,4 -azobis(4-cyanopentanoic acid, ACVA), R1 = A or
A disubstituted monomer derived from 4,4 -azobis(4-cyanopentanoic acid, ACVA), R1 = B
2. Terpolymer of AM and DHAM with
A monosubstituted monomer derived from 4,4 -azobis(4-cyanopentanoic acid, ACVA), R2 = C

Model polymer

Copolymer of (AA, R1 = H) or (MAA, R1 = CH3 ) with 2-(1-naphthylacetyl)ethyl acrylate (NAEA) or [246,247]


2-(1-naphthylacetamido)ethyl acrylamide (NAEAm)

Model polymer

Copolymer of AM with N-[(1-pyrenylsulfonamido)ethyl]acrylamide (APS) [110,111]


Table 1 (Continued)

Polymer Structure References

Model polymer

Terpolymer of AM and AA with N-[(1-pyrenylsulfonamido)ethyl]acrylamide (APS) [177]

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628


Model polymer

Terpolymer of AM and SA with [(1-naphthyl)methyl]acrylamide (NMA) or APS [178,179]

Model polymer

Copolymer of N-isopropylacrylamide (NIPAM) with N-[4-(1-pyrenyl)butyl]-N-n-octadecylacrylamide [235]

1579
1580 D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628

Fig. 9. Thermal induced micro-domains [151].

Fig. 10. Schematic presentation of micellar copolymerization [88,135].


Reproduced with permission from [88] 1993, ACS Publishing.

monomers per micelle (NH ) [88,109,134,175,258,276]. It another method that is closely related to the micel-
was first thought that the solubilization of the hydrophobic lar copolymerization technique seems promising, but it
monomer in a surfactant micelle, would cause an increase involves using a micelle-forming polymerizable surfac-
in the incorporation rate of the hydrophobic monomer tant [114,115,178,179,220,278–280]. Although identifying
(into the polymer backbone) [109,174]. A NH increase will a correct polymerizable surfactant for the desired molecu-
change the randomness, i.e. the regularity of the num- lar structure is difficult, the technique offers the advantage
ber of hydrophobic units incorporated in the polymer as a that purification (i.e. removal of surfactants as is the case
function of time [109,174]. However two studies [109,277] with micellar copolymerization) of the reaction mixture
demonstrated that by using a disubstituted acrylamide as no longer would be required. Yet another method, which
the hydrophobic monomer no drift in copolymer com- recently has been developed, is template copolymerization.
position was observed. Therefore, according to the study The structure of the copolymer is defined by the template
[109], the previously thought dissolution of the hydropho- that is used. A schematic presentation illustrating the tem-
bic monomer in the surfactant micelle no longer holds plate copolymerization is given in Fig. 11.
true. The observed behavior can be then attributed to The advantage of this technique with respect to the
the difference in polarity between the bulk and micel- other ones is that the block-like distribution of hydrophobic
lar phase, which modifies the reactivity of hydrophobes groups is better controlled. According to several differ-
[109]. Micelle copolymerization can also be used to synthe- ent studies [281–285] a longer sequence distribution of
size a more randomly distributed copolymer. This can be hydrophobic groups can be achieved with this technique.
achieved by using a molar ratio where approximately one The molecular weights of the polymers influence their
hydrophobic unit is solubilized in one micelle [138,275]. behavior in solutions. The molecular weight of HMPAM
Micellar copolymerization remains the most used remains difficult to determine since intramolecular associ-
method for synthesizing HMPAM [135]. Nevertheless, ations still exist even at very low polymer concentrations.

Fig. 11. Schematic presentation of the template copolymerization technique.


D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628 1581

Nevertheless, several different methods have been devised. 19–21. The properties of the subsequent polymer solutions
These include using the Mark–Houwink–Sakurada rela- can be tailored by careful design of the polymer structure
tion for polymers dissolved in water [110,113,114] or and composition. The polymers can be designed to be salt-
unmodified polymers in the exact same conditions as their tolerant or responsive to changes in the salt concentration
modified analogues [103,229,230,286,287]. However, all or ionic strength.
these methods introduce errors in the molecular weight Amphiphilics, 23–28, without electronic charges have
determination and it would be better to find a suitable sol- been studied extensively for their salt-tolerance. Addition
vent in which the polymer is molecularly dispersed [135]. of salt will not affect the viscosity of the polymer solu-
Several authors reported on formamide as a suitable sol- tion since its thickening capability arises from hydrophobic
vent for the molecular weight determination and measured associations and not from electronic interactions.
it by using light scattering experiments [88,109,174]. The polymers 29–33 have been synthesized in order to
Another popular strategy to increase the thickening enable fluorescence studies of hydrophobically associating
ability of polyacrylamide polymers is related to the pres- polymers. This technique allows studying the associating
ence of electrical charges along the backbone. Indeed, behavior of these polymers through their photo-physical
many different polyelectrolytes based on acrylamide have behavior in response to changes in the system. It has
been synthesized. These polyelectrolytes can be divided been demonstrated that polymers of acrylamide containing
into polyampholytes which include zwitterionic backbone pyrene functionalities exhibit increased in excited-state
polymers 9–11, having positive and negative charges in dimer (excimer) formation and viscosity [110,111,179].
separate building blocks in the backbone, and zwitterionic Variation in type of polymers causes different behavior
monomers 3–8 having positive and negative charges in the of the subsequent polymeric solutions. However there are
same building block, and traditional polyelectrolytes 1 and other parameters that also affect the structure and associ-
2. The charges can be induced by controlling the pH of the ations of the polymer solution: the chemical structure, the
subsequent polymer solution when, for example, AA is used synthesis method, the temperature, the type and concen-
as a co-monomer. In all cases it is reported that the solu- tration of salt and the pH (ionic strength). The following
tion viscosity is a function of the chemical structure and discussion summarizes the effect of these parameters.
the solution characteristics (i.e. ionic strength).
From the discussion on the polymers 1–11 one can 3.1.1. Chemical structure
conclude that there are many different advantages of incor- Generally speaking the first macroscopic effect of the
porating electronic charges into water-soluble polymers. polymer chemical structure is observed from the cor-
One such advantage is the ability to control the interaction responding solubility in water. Indeed, the solubility of
between the polymer chains by altering the pH or ionic polymer 3 is dependent on the level of AMPDAPS monomer
strength of the polymer solution. Another is the solubil- incorporated [196]. This has a clear influence on the vis-
ity in water, which makes the corresponding solutions less cosity of the solution [196] as observed for polymer 5 as
sensitive, in terms of viscosity, to the external temperature. function of the AMPDAE and AMPDAH monomers incor-
On the other hand if water-soluble polymers are poration level [199,200]. The same considerations hold for
required whose rheological properties are indepen- polymer 6 as function of the monomer composition (in
dent of the pH, polyelectrolytes bearing only one type this case AMPDAPS and AA) [201] as well as polymer 8
of charge can be used [201]. For applications where as function of the charged monomer (DADPMA) intake
concentrated salt solutions are used (such as EOR), [203]. The same can be said for polymer 19 where phase
polyampholytes are suitable. They have been demon- separation is observed at low incorporation levels of the
strated to display an enhancement in the solution viscosity DAMAPS monomer. These trends are easily understand-
upon addition of low molecular weight electrolytes able on the basis of simple considerations regarding the
[197,201,204,209–211,288,289]. This behavior has been overall polarity of the polymeric chains. However, in con-
attributed to the shielding of intramolecular Coulom- trast to this, polymer (8) is no longer soluble in water
bic attraction rather than the intermolecular interactions. when it contains more than 1 mol% of DADPMA. In a similar
With careful molecular design water-soluble polymers example, McCormick and Johnson [208] observed hydrogel
containing electronic charges can be synthesized exhibit- formation when the incorporation level of charged groups
ing the required rheological behavior. Peiffer and Lundberg surpassed 1 mol% in polymer 10. Similar to polymer 8, poly-
[289] reported that in order to control the physical prop- mer (10) forms hydrogels and is no longer water-soluble
erties of the zwitterionic polymers it is crucial to separate above 1 mol% incorporation of the charged groups. The car-
the oppositely charged monomers using neutral ones. boxylate groups lead to strong ionic interactions which
Derivatives of polyelectrolytes have been investigated prevents the dissolution of polymer 10 [208]. It seems thus
with amphiphilic (hydrophobic) moieties incorporated that the nature of the charge (e.g. carboxylate versus, sul-
along with charges, either positive or negative. These acry- fonate anion in Table 1) exert a clear influence on the
lamide based amphiphilic polyelectrolytes, 12–17, have association behavior and ultimately on the solubility of the
been synthesized by many different research groups in polymer. This phenomenon is never observed in the case
search for better thickening polymers in different appli- of a zwitterionic backbone. Indeed, at low incorporation of
cations. either monomer, AMPDAC or NaAMPS, in the correspond-
Polymers where both charges are incorporated along ing copolymer (polymer series 11) classic polyelectrolyte
with hydrophobic groups have also been studied. These behavior is exhibited by the solution. However, poly-
polymers can be classified as zwitterionic amphiphilics, meric solution employing polymers containing equimolar
1582 D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628

amounts of the AMPDAC and NaAMPS monomers display the co-monomer, does impart associations [237]. However,
high salt tolerance [209]. A similar behavior is observed in contrast to this, increasing the hydrophobicity too much
when, instead of AMPDAC, AMPTAC is used as the sec- will lead to insolubility of the polymers as observed with
ond monomer [211]. Similar to polymer 11, polymer 19 polymer 15 using longer twin-tailed groups than DiC12 AM
exhibits classic polyelectrolyte behavior when the incor- [218].
poration rate of DAMAPS is below 40 mol% [226]. However, The presence of spacers, i.e. small chains linking the
in dilute solutions of polymer 19 (both DADMAC-1 or 2 hydrophobic groups to the backbone, affects the associ-
copolymer) primarily intra-molecular hydrophobic associ- ation behavior of the polymers markedly. Increasing the
ations are present [228]. length of the spacers allows for easier movement of the
The presence of hydrophobic groups along the polymer hydrophobic group in solution which in theory should
backbone results generally in the formation of micro- lead to easier formation of micro-domains and thus inter-
domains. Important parameters for association behavior molecular association is favored. This is demonstrated
of the polymers are the distribution of the hydrophobic with polymer 13 [216] (DemMA series). The association
groups and their hydrophobicity. As mentioned before, behavior of the polymer is not affected by the type of
a block-like distribution of the hydrophobic groups will co-monomer that is used as long as the length of the
lead to stronger associations compared to a random dis- spacer is large enough (in this case a length of 25 EO
tribution. The surfactant to micelle ratio (SMR) during moieties). At intermediate lengths (2 or 6 EO moieties)
polymerization affects the distribution of the hydrophobic this is no longer true as demonstrated with the same
groups. At low SMR the block-like distribution is favored polymer backbone [216]. With NaA as co-monomer the
while at high SMR a random distribution is preferred network formation was favorable at shorter EO spac-
[175]. Polymers 13, 14 (synthesized using a high SMR) ers while just the opposite behavior is observed with
and 17 (using solution polymerization) are classic exam- NaAMPS as the co-monomer. The authors [216] attribute
ples of low level association due to random distribution this difference to the stronger tendency of intermolecular
of the hydrophobic groups [156,175,220] while polymers association for the NaA containing copolymer. The pres-
14 (synthesized with a low SMR), 17 (using micellar poly- ence of a charged group further away from the polymer
merization) and 30 are classic examples of a block-like backbone will disrupt the hydrophobic associations to a
distribution [111,175,220]. Although the SMR has been greater extent than when the charged group is closer to
identified as an important parameter for the observed the polymer backbone [229]. The NaAMPS co-monomer
association behavior, its effect on the composition and places the charged group further away from the backbone
molecular weight of the polymer is minimal as demon- thus leading to a stronger disruption of the hydropho-
strated with polymer 31 [177]. bic group association at similar length EO spacers. This
The presence of hydrophobic groups suppresses the behavior has been identified for the aforementioned poly-
solubility of the polymer. Following this, it is easy to mer 13 and for polymer 20. In addition, charge density
understand that increasing the fraction of the hydrophobic of the polymer affects the association behavior. Using
groups above a certain percentage will lead to solubil- low charge density co-monomers (carboxylate anions)
ity issues, i.e. the polymer is no longer water-soluble. will display stronger association above the CAC due to
This has been demonstrated with polymer 24 where a less interference with the hydrophobic associations, while
water insoluble polymer was obtained using DiC8 AM as sulfonate anions (higher charge density than carboxy-
the co-monomer above an incorporation rate of 1.2 mol% late anions) will display the opposite behavior. Polymers
[237]. bearing high charge density groups are more suscepti-
Increasing the hydrophobicity of the hydrophobic ble to screening effects in the presence of low molecular
groups will impart better thickening capability of the poly- weight electrolytes and will therefore be less responsive
mer as demonstrated with the polymers 12, 15, 17, 18, 23, at high electrolyte concentration or low pH. The poly-
24 and 27 (2a, 2b, 3a–c and 4a–d). It has been demonstrated mer series 21 is a classic example of this [230]. When
that only a fraction of the hydrophobic groups contribute using 3 co-monomers the incorporation level of charged
to the formation of micro-domains which is affected by the species will also affect the association behavior of the poly-
hydrophobicity of the groups and the incorporation rate mer. At low incorporation levels of charged species no
[290]. The hydrophobicity of the groups can be increased suppression of the hydrophobic associations is expected.
by increasing the length of the groups [103,220,237], using However, at high incorporation levels the interference with
twin-tailed hydrophobes instead of single tailed groups the associations is expected to be significant. This was
[185,218] or using fluorocarbons instead of hydrocarbons demonstrated using the polymer series 14. The polymers
[212,223,242,269]. The critical polymer concentration at with 9 or 21 mol% AA display intermolecular hydrophobic
which hydrophobic associations (critical association con- associations at low pH (<5), and in the presence of NaCl.
centration, CAC) arise is reduced compared to classic However, polymers with 37 mol% acrylic acid do not dis-
associating polymers [218]. These trends can be explained play this behavior, not even in de-ionized water and low
by the stronger interactions between the hydrophobic pH [176].
groups due to their increased hydrophobicity. The length A peculiar behavior, in the context of EOR, is the tem-
of the hydrophobic groups is important for association perature stability of polymers. It has been demonstrated
where a short hydrophobe will not lead to association as that by using phenyl containing moieties as co-monomers
demonstrated with polymer 24 using DiC3 AM as the co- increases the stability of polymer 25 (PMAAM) against high
monomer [237]. Increasing its length, by using DiC8 AM as temperatures [239].
D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628 1583

Fig. 12. Schematic model structure of HMPAM.


Reproduced with permission from [291] 1991, ACS Publishing.

3.1.2. Rheological properties original value [88,92,103,109,132–134,136,138,139,242,


Hydrophobic interactions confer HMPAM interesting 257,278,293].
rheological and solution properties. Leibler et al. [291] Some associative polymeric solutions display pseudodi-
proposed the “sticky reptation” model, which explains latant behavior, i.e. increase in viscosity with increasing
the complex viscoelastic behavior [112,257,258,292] of shear rate [109,134,220,258,260,293,294]. According to
HMPAM polymers. The authors proposed a schematic several studies [109,134,220,295,296] this behavior can be
model structure where a hydrophobic containing chain interpreted as a balance between intra- and intermolecular
and the hydrophobic interactions between the chains’ associations. Above a given shear rate the intramolecu-
hydrophobic groups and other chains in the vicinity are lar associations are disrupted and the polymeric chains
contained in an imaginary tube (Fig. 12). are extended, which leads to more intermolecular asso-
According to the model the HMPAM polymer chains ciations [220]. A later study [258] demonstrated that the
form an entangled transient network where entanglements pseudodilatant behavior arises slightly above the crossover
and hydrophobic interactions are present. The distance concentration (C*), which is the overlap concentration of
between entanglements is shorter than the one between the polymer chains. Increasing the polymer concentration
hydrophobic groups. Two different relaxation times, a short will lead to a polymeric solution that does not display pseu-
and a long one, are predicted for the entangled transient dodilatant behavior [109,134,258,297]. The pseudodilatant
network. The short one corresponds to residence time of behavior of the polymeric solutions is followed by the
a hydrophobic group in a hydrophobic micro-domain and pseudoplastic behavior discussed earlier. Although many
the long one corresponds to the time a chain takes to rep- authors have demonstrated the pseudodilatant properties
tate outside its tube. of HMPAM solutions, it has to be mentioned that there are
In the dilute region (at low polymer concentration) differences in the observed behaviors. The most important
intramolecular associations dominate [108]. The hydro- one being that polymers produced by post modification
dynamic volume is reduced and therefore the viscosity only display the pseudodilatant behavior in the presence
of the subsequent polymer solutions. When the polymer of salt [293,294] while polymers produced using the micel-
concentration is increased the solution ideally moves to lar copolymerization technique display the pseudodilatant
the semi-dilute region where intermolecular associations behavior also in water [109,134,220,258,260]. However,
dominate [107,108,293]. This leads then to network- no clear explanation for this observation has been given
like formations (transient network) which substantially yet. As mentioned before, the type of distribution of the
increases the viscosity of the solution [107,108,293]. hydrophobic groups has a pronounced effect on the type of
Many publications discuss the complex nature of the rheology that is subsequently observed in solution. By post
rheology of polymer solutions of hydrophobically mod- modifying the polymers the distribution of the hydropho-
ified polyelectrolytes. It has been demonstrated that all bic groups would be different since little control is present
the types of non-Newtonian rheological behavior exist on which functional group is modified. The distribution of
for these polymer solutions, i.e. pseudoplastic; pseu- the hydrophobic groups is similar to block-like distribution
dodilatant; thixotropic; and rheopectic. Application of obtained with micellar copolymerization.
shear will, in classical pseudoplastic polymeric solu- Thixotropic and rheopectic (anti-thixotropic) behavior,
tions, disrupt the hydrophobic associations, which leads as mentioned before, has been observed. The viscos-
to a reduction in viscosity [88,92,103,109,132]. How- ity against shear rates curves obtained under increasing
ever this process is reversible, i.e. when the shear is and decreasing shear rates are not superimposable
removed the hydrophobic groups will form new associa- [88,103,175,257]. Thixotropic solutions display a peculiar
tion thus returning the viscosity of the solution up to its behavior where the viscosities along the increasing shear
1584 D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628

(7) undergoes a polyanion → polyzwitterion → polycation


transition as the pH decreases [202].
The shear rate for the onset of pseudoplastic behav-
ior and the viscoelastic properties are dependent on the
strength of the hydrophobic micro-domains. The strength
depends on the aforementioned hydrophobicity of the
groups. Increasing the hydrophobicity of the groups, i.e.
increasing the length of the groups [220,267], the length
of the hydrophobic segment [220], using twin-tailed rather
than single tailed groups [185] or fluorocarbons instead of
hydrocarbons [298], result in a reduction of the shear rate
for the onset of pseudoplasticity and an enhancement of
Fig. 13. Counterion mediated interpolymer associations.
the viscoelastic properties. A classic example of a polymer
Reproduced with permission from [214] 2005, ACS Publishing.
whose shear rate for the onset of pseudoplasticity reduces
with increase in the hydrophobicity of the groups is poly-
rate curve are all higher than those along the decreasing mer 15 [185].
shear rate curve, i.e. the viscosity decreases with time at The rheological properties of HMPAM can be, as men-
a constant shear rate. The application of shear disrupts tioned before, explained by the balance between intra- and
the hydrophobic associations, both intra- and interchain, intermolecular hydrophobic associations. The copolymer-
which need time to re-associate in order to recover the ization process affects the distribution of the hydrophobes
enhanced viscosity. In the semi-dilute region interchain (vida supra). With the preference being intermolecular
associations dominate, as mentioned before, and there dis- associations, an enhancement in the solution viscosity is
ruption by shear will lead to a significant loss in viscosity obtained whereas with intramolecular associations the
[257]. However as the viscosity is recovered, the shear exact opposite is observed [110,111]. Polymers 16 and 29
induced reduction in viscosity is not due to polymer degra- are classic examples of polymers with an enhanced solu-
dation [88]. Polymers prepared using the heterogeneous or tion viscosity due to their preference for intermolecular
homogeneous technique do not display this behavior [88]. association.
Rheopectic solutions display the opposite behavior, i.e. The position of the hydrophobic groups along the back-
the viscosities along the increasing shear rate curve are all bone is important [245]. From a chemical point of view
lower than those along decreasing shear rate curve (the the amount of hydrophobic groups present in the polymer
viscosity increases with time at a constant shear rate). This should have an optimum. If the concentration is low, no
behavior involves the instant recovery of the hydrophobic association will arise, while, if too high, solubility issues
associations after the application of shear and the enhance- play a crucial role. However it is demonstrated that plac-
ment of these hydrophobic associations, which explains the ing the hydrophobic group (polymer 24) at the ends of
enhanced viscosity [103,175]. the polymer is the optimal configuration for obtaining the
highest solution viscosity [245]. According to the authors
3.1.2.1. Effect of inherent parameters. As mentioned the placement of hydrophobic groups along the back-
before, spacers between the hydrophobic group and the bone of the polymer leads to a polymer with a much
hydrophilic backbone can enhance the solution viscosity. more compact structure in solution compared to when
Increasing the length of the spacer will lead to a better the hydrophobic groups are placed at both ends. However
enhancement of the solution viscosity. A classic example of when combining the two extremes (telechelic with mul-
such behavior is polymer 13 (NaAMPS as the co-monomer) tisticker) it was demonstrated that the highest thickening
[215]. A subsequent study [214] demonstrated that, capability, i.e. solution viscosity against polymer concen-
depending on the incorporation rate of the DEmMA, a sig- tration can be obtained [104,245]. The critical polymer
nificant increase in solution viscosity is obtained above the concentration for the onset of hydrophobic associations
critical overlap concentration. The authors hypothesized depends on the placement of the hydrophobic group [104].
that the additional increase in solution viscosity is due For a telechelic polymer with hydrophobic groups at both
to the simultaneous interactions of countercations with ends and a polymer with hydrophobic groups at both ends
the EO spacers via coordination and with the polyanion and along the backbone the critical polymer concentra-
via counterion condensation. A schematic presentation tion for the onset of hydrophobic association is the same
illustrating this behavior is given in Fig. 13. whereas for a multisticker polymer with only hydrophobic
It has been demonstrated that the responsiveness of the groups along the backbone the concentration is higher.
polymer solution is more pronounced when a longer spacer The same kind of consideration, namely strong depen-
is used. A study of Kathmann et al. [199] demonstrated that dence of the rheological behavior from the chemical
using three methylene groups as spacers increased the pH structure is applicable for fluorine containing polymers.
responsiveness of the polymer solution compared to when Indeed, polymer 18 displayed shear thickening, which
only one methylene unit is used. The pH responsive behav- according to Chang and McCormick [228] depends on the
ior of the polymers can also be tailored by molecular design molecular structure of the copolymer. Increase in the incor-
of the polymer. The ampholytic polymer 7 represents such poration rate of the hydrophobe from 3 to 10 mol% leads to
a polymer where by smart design different behavior is a more pronounced shear thickening behavior. However
achieved as a result of changes in the pH. The polymer at an incorporation level of 24 mol% the shear thickening
D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628 1585

behavior is significantly suppressed. The shear thickening The presence of salt causes a shielding of the electro-
is attributed to the breakage of intramolecular hydropho- static interactions and thus a collapse of the network with
bic associations by shear. If the hydrophobic associations total loss of the solution viscosity. Polymer 1 is a classic
are too strong (as is the case for the latter polymer), the example of such a polyelectrolyte where phase separation
applied shear cannot break the associations and thus shear is observed in the presence of multivalency cations [193].
thickening is not observed. Polymer 5 displays a similar trend. However, a decrease
The dependency of the solution viscosity on the in the pH of the polymeric solution (containing salts)
(low molecular weight) electrolyte concentration can be improves the solution viscosity [198] due to the increase
adjusted as function of the chemical structure. Indeed, the in hydrodynamic volume caused by the protonation of the
salt tolerance of terpolymer 19, 20 was examined using carboxylic groups at low pH [198–200].
three different derivatives (NaAMPS, NaAMB and NaA) as The effect of salt on polymers bearing charged moi-
the third monomer. Terpolymers incorporating NaA as eties in combination with hydrophobic groups is peculiar.
the third monomer displayed the best salt tolerance of Indeed, an increase in the solution viscosity is observed
the three different terpolymers [229,230]. The apparent upon the addition of different salts. The presence of salt
viscosity (app ) of the polymer solution was significantly screens the electrostatic repulsion thus suppressing the
higher in high salt concentration solution compared to disruption of hydrophobic associations by the charged
the other two terpolymers. The strength of the hydropho- groups. This allows the formation of a stronger network
bic associations depends on the type of groups used. The through hydrophobic associations [225], i.e. intermolecu-
terpolymers incorporating carboxylate anions displayed lar hydrophobic interactions are dominant rather than the
stronger hydrophobic interactions compared to the poly- electrostatic repulsions [221], and thus an enhancement
mers with sulfonate anions. In addition, the placement of of the solution viscosity. Classic examples of this behav-
the anions is crucial for the strength of the hydrophobic ior (vida supra) are the polymers 12 [115], 17 [221] (AA-CN
associations. It was demonstrated that the further away copolymer, N ≥ 12), 18 [225] and 25 [133,238]. Although
from the polymer backbone the anions are placed the not in the same class of polymer, polymer 4 displays the
weaker the hydrophobic associations will be. This behavior same behavior [197]. Another possibility is that the methyl
was attributed by the authors [229] to the interference by groups on the AMDAPS monomer induce hydrophobic
the anions with the hydrophobic associations. In addition behavior of the monomer at higher salt concentration
the rheological properties of the polymer solutions depend thereby displaying the aforementioned behavior.
on the charge density. Polymers with low charge density Maia et al. [121] demonstrated that the method of
lead to less electrostatic interference of the hydrophobic preparing the HMPAM solutions affects their rheological
associations [230]. behavior in the presence of salt (NaCl). When the polymer,
The rheological properties of polymer 15, 22 and 23 (AM–DiC6 AM copolymer), is dissolved in the salt (brine)
24 depend on which hydrophobe is incorporated. Using solution (A, Fig. 14) the classical behavior, i.e. reduction in
disubstituted monomers leads to a more pronounced thick- viscosity with increasing salt concentration, is displayed.
ening effect compared to monosubstituted ones. This is However, if the polymer is dissolved in water first, which
attributed to the proximity of two hydrophobic chains is preceded by the addition of salt, a complete different
(higher density of the hydrophobic domains) that leads solution behavior is observed. Diluting a polymer solu-
to stronger hydrophobic associations [109]. In addition tion (no salt) with water and adding salt afterwards (B,
the length of the hydrophobic block affects the rheologi- Fig. 14) results in a polymer solution whose viscosity passes
cal properties of the polymer. The longer the hydrophobic through a maximum with increasing salt concentration. If
blocks, the stronger the hydrophobic associations will be. the polymer solution (no salt) is diluted with salt water
The onset of shear thickening shifts to lower shear rates (C, Fig. 14) the viscosity of the resulting polymer solu-
as the length of the hydrophobic blocks increases [109]. tion increases with increasing salt concentration. Fig. 14
This behavior is demonstrated using polymer 14 (R3 = 5) presents the three different behaviors. This confirms the
where the thickening capability increases as the hydropho- non-equilibrium character of these solutions, but a better
bic block increases in length [299]. explanation has yet to be provided.
It has been demonstrated that the following surfac-
tants interact with different HMPAM polymers and give,
3.1.2.2. Effect of external parameters. For neutral associa- depending on the surfactant concentration, a higher
tive copolymers it has been demonstrated that upon viscosity [90,92,115,138,173,287,300–302]. The interac-
addition of salt the viscosity of the solution can be enhanced tion between the polymers 12, 13 (both copolymers),
[103,108,203]. This behavior of the neutral copolymers is 17 (both copolymers), 18 (DADMAC-1 and DADMAC-2
attributed to a “salting out effect”, which arises due to a copolymers), 21 (copolymer AM-C10 AM), 22 (terpolymer
change in the solubility of the hydrophobic units [108]. 1, R2 = 2), 23 (copolymers AM-EAM, AM-BPAM and
The solubility of the hydrophobic groups decreases with AM-2 or 3 or 4), 27 (6) and 33 with several different sur-
increasing salt concentration. This leads to the formation of factants (SDS [90,92,120,138,173,212,228,243,301–305],
aggregates, which enhances the interactions between the potassium dodecyl sulfate (KDS) [212], cetyltrimethyl-
polymeric chains [108]. The intermolecular association is ammonium p-toluenesulfonate [CTAT] [305,306],
enhanced leading to a stronger network and thus increased dodecyltrimethylammonium chloride [DTAC] [290,300],
solution viscosity [217]. Polymer 14 is a classic example of hexadecyltrimethylammonium [HTAC] [303,304], CTAB
this behavior [217]. [90,233], trimethyltetradecylammonium bromide [TTAB]
1586 D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628

Fig. 14. Different behavior of HMPAM solution dependent on the preparation method.
Adapted from Maia et al. [121]. Reproduced with permission from [121] 2005, Wiley VCH.

Fig. 15. Schematic presentation of concentration regions of HMPAM with SDS [92].
Reproduced with permissions from [92] 1992 and [132] 1996, ACS Publishing.

[302], cetyldimethylamine oxide [CDMAO] [90], alkyl- increase below or above the cmc of the surfactant [90]. In
benzenesulfonates [ABS] [287], Triton X-100 [302], region II enough surfactant molecules (higher surfactant
n-octyl ␤-d-glucopyranoside [OG] [303] and n-octyl concentration) are present to solubilize the hydrophobic
␤-d-thioglucopyranoside [OTG] [303,304]) have been groups more effectively (mixed micelles). The formation of
investigated. The viscosity of the HMPAM solution with mixed micelles is the onset of a significant increase in vis-
surfactant passes through a maximum, which is always cosity. Most of the mixed micelles incorporate more than
just under the surfactants’ critical micelle concentration one hydrophobic region, i.e. a higher degree of overlap or
(CMC) [173]. The increase in viscosity is attributed to the links between separate chains is achieved. In region III the
formation of mixed micelles of surfactant and hydrophobic surfactant concentration is higher, at or above the cmc,
regions [92,117,120,154,173,300,301,303,304,307–309]. which results in solubilization of each hydrophobic region
According to several authors [92,303,304], when adding in one micelle. This leads to a reduction in viscosity.
surfactant to a HMPAM solution, the surfactant binds to Transition from spherical to rod-like micelles in the
the copolymeric regions, which leads to preferential inter- surfactant phase has been observed upon addition of
chain micro-domain formation instead of intrachain. A potassium bromide and hexanol to respectively the
schematic presentation of this behavior has been provided polymer–CTAB and polymer–CDMAO mixture [90,115].
by Biggs et al. [92] (Fig. 15). The transition is believed to be due to the gain in free energy
The surfactant concentration in region I is low. The upon addition of the chemicals where the hydropho-
surfactant associates in a non-cooperative way with bic groups no longer are exposed to the solvent but
the hydrophobic groups [92]. However, this is depen- instead are present inside the hydrophobic rods [90,115].
dent on the type of surfactant. Two different studies Fig. 16 provides a schematic presentation of the rod-like
[303,304] demonstrated that the surfactants hexade- micelles.
cyltrimethylammonium chloride (HTAC) and SDS bonded This transition to the rod-like conformation leads to
in a non-cooperative way but the surfactants OG and OTG more intermolecular bridging and thus enhanced solution
bonded in a cooperative manner. Winnik et al. [304] con- properties. This is evident from the significantly higher
cluded that ionic surfactants bind by a non-cooperative viscosity of the solutions compared to the viscosity of
mechanism and neutral surfactants by a cooperative one. the individual components [90,115]. Both, rheopectic and
According to another study there are two classes of interac- thixotropic, behaviors have been observed for polymer
tions, which are differentiated by the bridging and viscosity solutions in the presence of surfactant [92].
D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628 1587

Fig. 16. Schematic presentation of rod-like micelles [90,115].

Using polymer 23 (R2 = 2, 3 or 4 series) Gouveia and polymer 4, displays a unique behavior when the temper-
Muller [306] observed a large increase in the solution vis- ature is increased. The intrinsic viscosity of the polymer
cosity upon the addition of CTAT. Adding salt (NaCl) to this solution increased (from 6.5 to 8.5 dl/g) when the tem-
solution leads to a further increase in the solution viscos- perature is increased from 25 to 60 ◦ C [197]. This has
ity of the mixture. According to the authors the worm-like been observed for polymer 14 in the temperature range
micelles increase in size by the addition of salt and at high of 20–40 ◦ C. According to the authors [217] the increase
salt concentration salting out effects arise. can be attributed to the fact that hydrophobic associa-
A peculiar behavior is observed for the interactions tions are endothermic in the investigated temperature
between fluoro- and hydrocarbon copolymers 12 and sur- range, as hypothesized by McCormick et al. [103]. Using
factants. At low incorporation level of the hydrophobic the latter terpolymers 14 L’Alloret et al. [310] observed
groups the interactions with surfactants is selective to sur- a significant increase in the solution viscosity when the
factants with the same chemical nature as the hydrophobic temperature is increased from room temperature to 80 ◦ C.
groups [212]. In addition the authors demonstrated that the thermoth-
Different oil reservoirs possess different temperatures. ickening behavior of the polymeric solution is more
The polymer solutions that are to be used must be able pronounced at lower salt concentration and lower shear
to cope with the different temperatures, i.e. no loss of rates. The latter effect is also observed with polymer 27
viscosity with alteration in temperature. The viscosity (6) [243] and is attributed to the increase in PPO con-
of polymeric solutions employing the polymers 10 (ter- centration in the hydrophobic micro-domains caused by
polymer of AM–NaAMPS–AMPDAC) or 12 (Table 1) is an increase in the mobility of the chains at relatively
minimally dependent on the temperature in the range high temperature. At higher temperatures, the reduction
30–60 ◦ C [204,209,210]. In addition polymer 12 displayed in viscosity is attributed to the loss in connectivity of
good retention of solution viscosity for a prolonged period the network due to changes in the hydrophobic micro-
(40 days) [209]. The viscosity of the polymer solution domains. According to the authors the size of the micro-
containing the zwitterionic (monomer) polyelectrolyte, domains increase at higher temperatures but their number

Fig. 17. Schematic illustration representing the thermally induced conformational changes.
Reproduced with permission from [235] 1991, ACS Publishing.
1588 D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628

Fig. 18. Schematic presentation of the effect of solution pH on the network structure.
Reproduced with permission from [218] 2001, ACS Publishing.

decreases and causes the aforementioned connectivity tions. Zhou et al. [225] demonstrated, using polymer 18
loss. (AA series), that increasing the pH from 4 to 5 and from
The viscosity of the AM–AMPDAC copolymers in the 11 to 12, a significant increase in the solution viscosity
polymer series 1 decreased as expected with increasing is observed while at a pH between 5 and 11 the solu-
temperature [193]. The same behavior is observed for a tion viscosity is lower. A similar behavior was observed
polymeric solution employing polymer 2 [195]. Unlike for polymer 17 (AA series) although the pH where the
polymers 1 and 2, an intriguing property of polymer 4 is increase in solution viscosity is observed is shifted. The
the increase of the intrinsic viscosity with temperature (in increase is observed between a pH of 5–6 and 12–13 [221].
the range 25–60 ◦ C). Another polymeric solution that dis- Although the effect of pH on polyelectrolyte amphiphilics is
plays peculiar behavior with increase in the temperature complex (vida supra), the investigation of the polymer net-
is the solution of terpolymer (9, AM–NaAMPS–AMPDAC), work structures of solutions employing polymer 15 using
which was minimally dependent on the temperature NRET measurements led to the development of a model
[204]. illustrating the effect of the solution pH on the confor-
The LCST concept, discussed in the introduction of this mation of the polymer chains and their associations [218]
chapter, also applies to polymer 33 where the viscosity (see Fig. 18).
of the solution can increase with an increase in temper- Increasing the pH will lead in principle to the transi-
ature. However, one important finding reported by the tion towards more intermolecular associations rather than
authors [235] is the presence of hydrophobic associations intramolecular. The high degree of ionization (at high pH)
below the LCST. An increase of the solution temperature will disrupt the intramolecular hydrophobic associations
(towards the LCST) leads to disruption of these associations. causing a rearrangement of the network with a prefer-
However, only for long or high hydrophobicity groups do ence for intermolecular associations [218]. Amphiphilics
the associations arise below the LCST. This behavior was are affected by the solution pH mainly due to the carboxylic
observed with the amphiphilic polymer 23 (NIPAM-CN AM). groups of the polymers. Polymer 26 is a classic example
Four decades ago, it was proposed that the LCST should showing what the effect of the pH is on the behavior of an
decrease with the polymer hydrophobicity [311]. However, amphiphilic polymer. The solution viscosity passes through
the study by Ringsdorf et al. [235] demonstrated that the a maximum as the pH rises from 4 to 12. This behavior is
proposed trend is not followed when increasing the length attributed to two different effects: neutralization of the car-
of the hydrophobes in polymer 23 (NIPAM series). This boxylic groups leading to (1) intramolecular electrostatic
is attributed to the formation of a micellar structure by repulsion and thus chain extension and (2) intermolec-
the hydrophobes, which prevents them from contributing ular electrostatic repulsion leading to disruption of the
to the LCST suppression. On the basis of this hypothesis, intermolecular associations. The first effects dominate at
the authors have proposed a model representing the ther- low pH, the latter one at high pH [240]. However, in con-
mally induced polymer conformational changes for both, trast to this classic polyelectrolyte behavior, increase in the
23 (NIPAM series) and 33, polymers (see Fig. 17). pH results in the formation of intramolecular hydrophobic
In addition, increasing the temperature from 25 to association for polymer 29. The carboxyl groups repel each
55 ◦ C leads to the preferential formation of intermolecu- other in classic polyelectrolytes to form expanded coils,
lar hydrophobic associations and thus an increase in the however the presence of large hydrophobic groups on the
viscosity. polymer backbone impart the repulsion between the car-
The pH of the polymeric solution also affects the solu- boxylic groups. The polymer forms a much more compact
tion viscosity. The effect on classic polyelectrolyte is easily structure, which is called a hypercoil [312] (Fig. 19).
understood given the charged nature of the polymers. The hydrophobic groups associate to form the interior
Polyanions will have a low viscosity at low pH and high of the hypercoil, which is surrounded by carboxyl groups.
viscosity at high pH while polycations display the oppo- The same behavior has been observed for polyacids bearing
site behavior. Polymer 2 is a classic example of a polyanion large hydrophobic groups [95,96]. McCormick et al. [219]
where the solution viscosity decreases as the pH of the noticed the same behavior for polymer 16. By increasing
solution is reduced. The effect of pH is more pronounced the ionization degree of the polymers containing 50 mol%
on polyelectrolyte amphiphilics where the associations are 4-butylaniline (4-BA) the polymers undergo a transition
a balance between hydrophobic and electrostatic interac- from closed (intramolecular) associations to open (inter-
D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628 1589

Fig. 19. Schematic presentation of a hypercoil. Fig. 20. Schematic presentation of transition to intramolecular associa-
Reproduced with permission from [312] 1986, ACS Publishing. tion.
Reproduced with permission from [179] 1995, ACS Publishing.

molecular) associations. If the 4-BA content is 70 mol%


only intramolecular associations are present at all ioniza- the polymer is different. Two models have been proposed
tion degrees. Unlike polymer 16 and 29 no intermolecular by the two studies [226,227] illustrating the effect of the
hydrophobic associations could be detected for polymer 31. ionic strength and pH on the structure of polymer 19 that
This is attributed to the high number of carboxylate units is obtained (see Fig. 21).
at high pH which will disturb the water structure to such Indeed, for classic polyelectrolytes, increasing the ionic
an extent that the driving force responsible for hydropho- strength of the solution (i.e. increasing the salt concentra-
bic associations is suppressed. At low pH, intramolecular tion) will lead to a decrease in the solution viscosity due to
hydrogen bonding stabilizes intramolecular hydrophobic screening effects. However for polymer 9, which bears both
associations thus excluding intermolecular association. electronic charges, a peculiar behavior is observed where
The transition from closed (intramolecular) to open (inter- the intrinsic viscosity increases with increase in the ionic
molecular) associations has been confirmed using the strength [205,206].
model polymer 32. This transition, which is triggered by
changes in the pH and salt concentration can be viewed 3.2. Hydrophobically modified ethoxylated urethane
as a molecular level event instead of a macroscopic phe- (HEUR)
nomenon [178]. A schematic presentation of this behavior
has been provided by Kramer et al. [179] (see Fig. 20). The interest in HEUR focussed on fundamental research
Although a transition arises for polymer 19 with [313] and commercial applications in paint formula-
changes in the ionic strength and/or pH, the restructuring of tion [314,315], paper coating [314,315] and shampoo

Fig. 21. Schematic presentation illustrating the effect of pH and ionic strength.
Reproduced with permission from [226] 2000 and [227] 2000, ACS Publishing.
1590 D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628

different structures of the corresponding water solutions


and their rheological behavior.
The physical structure that telechelic polymers attain,
when solubilized in an aqueous environment, was eluci-
dated using fluorescence studies. In principle rosette-like
[124,126,318,319,328] aggregates are formed with diam-
eters in the nanometer range resembling star polymers
[340,341]. The formation of these structures starts above
Fig. 22. Chemical structure HM PEO.
the critical micelle concentration (CMC) [124]. The rosette-
like structure comprises an inner core rich in hydrophobic
groups (ranging from 4 to 20 hydrophobic groups) and a
formulation [314,315]. Their use as model compounds can corona composed of the hydrophilic PEO chains (approx-
significantly increase the scientific understanding of asso- imately 5–10 looped chains) [130,318,319,342]. The
ciating polymers. Table 2 presents an overview of the number of hydrophobic groups present in the hydropho-
published systems with the corresponding chemical struc- bic core depends on the type of group used [326]. The
ture. number of hydrophobic groups in the inner core and
HEUR polymers are prepared by chain extension of the structure of the corona are independent of polymer
poly(ethylene oxide) (PEO) oligomer with a narrow molec- concentration [318,319], only the number of rosettes is
ular weight distribution (MWD) using di-isocyanate and affected by the polymer concentration. Although the exact
end-capping with a hydrophobic group [105]. HEUR poly- dimensions vary with the polymer molecular structure,
mers can be categorized into two main classes: step growth the polymers 35, 36, 43 and 44 all were demonstrated to
(S-G) HEUR and uni-HEUR. The differences between the form the aforementioned structures in an aqueous solu-
two classes relates to the synthesis method and the cor- tion. Fig. 23, provided by Xu et al. [343], shows the micelle
responding MWD of the product [336]. The S-G HEUR structures and the effect of polymer concentration on the
involves a procedure where PEO (of a given Mw) is reacted conformation of the polymer (for polymer 36, AT 22-
with a large excess of a di-isocyanate agent to produce 3).
a precursor which contains isocyanate functional groups Several other publications [105,126,329,338] provided
at both ends. This precursor is subsequently reacted with similar schemes illustrating the different concentration
a hydrophobic group to end-cap both ends yielding a regimes.
telechelic polymer. This procedure leads to a HEUR poly- Aside from the independency of the structures on con-
mer displaying a broad MWD. The procedure for uni-HEUR centration, the type of end group is another parameter
involves the direct addition of a mono-isocyanate contain- that does not affect the network structure. According to
ing a hydrophobic group to the PEO. To this a hydrophobic Fonnum et al. [328], only the relaxation time of the net-
group is added. The resulting polymer has a relatively nar- work depends on the end group as observed with polymer
row MWD, which is related to that of the parent PEO. The 45 (1–6).
chemical structure of a hydrophobically modified PEO is The rosettes can be bridged by polymeric chains, at
given in Fig. 22. higher concentration, leading to a network as displayed
HEUR polymers are designed to be end-capped at both in Fig. 23. The higher the tendency towards bridging, the
ends with a hydrophobic group. However it has been more stable the network will be. The bridging is affected
demonstrated that, in practice, a significant fraction of by the polymer molecular structure. Xu et al. [343] have
these polymers is end-capped at only one side [337]. The demonstrated that the comb polymer 36 has a much
polymers usually comprise a mixture of telechelic polymer higher fraction of bridging chains and a lower fraction of
and nonionic surfactant [125]. loop chains when compared to the telechelic analogues.
In most cases long chain alcohols [105,122,322,324,325] When investigating the structures of HEUR polymers,
are used as hydrophobic groups although other types of polymer 38(2) containing only one hydrophobic group, a
hydrophobic groups have been investigated. These include peculiar behavior is observed. A smaller hydrophobicity
fluorocarbon containing hydrophobes [186,329–331], group (C6 F13 ) leads to dimer formation whereas a larger
amine containing hydrophobes [122], alkyl groups hydrophobicity group (C8 F17 ) leads to the formation of
[124,128,170,316,327] and alkoxy groups [129,318,319]. micelles (Fig. 24) [189].
Combinations of telechelic and internal hydrophobes [122] For HEUR polymers hydrophobic associations start
and comb-like structures using a long chain 1,2-diol [125] at very low polymer concentration (as low as 10 ppm).
have been investigated. Although the exact critical concentration for the onset of
As mentioned earlier an isocyanate agent is used hydrophobic association varies, the solutions employing
in the synthesis. Different isocyanate agents have polymers 39, 43 and 44 have a critical concentration in
been used such as isophorone di-isocyanate (IPDI) the same range [125,327]. Most of the HEUR polymers in
[122,125,129,186,318,319,329–331,338], dicyclohexyl- Table 2 contain urethane linkages. A study by Alami et al.
methane di-isocyanate [322,324,325], octadecyl (alkyl) [344] demonstrated that at low polymer concentration
isocyanate [317,323,339] and different polyisocyanates the presence of urethane linkages dramatically influence
[317,339]. the initial association and the clouding of the polymer
Differences in the synthetic procedures and the used in water. HEUR polymers containing urethane linkages
hydrophobic groups clearly result, as discussed below, in will form aggregates sooner than their analogues without
Table 2
Structure of different water-soluble polymers, HEUR.

Polymer Structure References

Polyethylene oxide end capped at both sides with dodecyl groups. [316]

Polyethylene oxide end capped at both sides with different alkane hydrocarbon [89,317]
groups (R) employing different diisocyanates moieties (DI).
Given the amount of different reagents used we refer to US Pat. 4,079,028 [317] for the
complete list of reagents.

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628


Polyethylene oxide end capped at both sides with an hexadecyl group: [318,319]

AT 22-2, y = 4
AT 22-3, y = 6
AT 107, y = 3
AT 67-3, y = 1
IPDU = isophorone diurethane

Polymers based on ethylene oxide using different hydrophobic groups and different [122]
structure. Either end capped (traditional HEUR polymers, 1) or the hydrophobic end
groups are separated by oxyethylene groups around an internal hydrophobe, 2–4.

1591
1592
R = different amine and alcohol containing groups

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628


Polymers based on ethylene oxide using different hydrophobic groups and different [189,320]
diisocyanate groups. The polymers are either end capped at both sides with internal
hydrophobic groups.

R = C6 H13 , C8 H17 , C12 H23 , C14 H29 and C18 H37

Polymers based on ethylene oxide: [125]

Comb-81, w = 1, y = 2 and z = 3
Comb-82, w = 1, y = 2 and z = 6
Comb-83, w = 1, y = 2 and z = 9
IPDU = isophorone diurethane
Table 2 (Continued)

Polymer Structure References

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628


Polymers based on ethylene oxide: [321]

A: ethylene oxide-alkylglycidyl ether) copolymer (EO-AGE)


B: ethylene oxide-3,3-dialkoxymethyl)propyl glycidyl ether) copolymer (EO-DAGE)
C: poly(methyleneoxide-alt(PEO)polyalkyl-glycidyl ether (MO-PEO-PAGE)

Polymers based on ethylene oxide: [124]

AT18-31, n = 230, DI = 1 and R = 4


AT18-13, n = 90, DI = 1 and R = 4
AT18-19, n = 140, DI = 2 and R = 4
AT15-19, n = 140, DI = 2 and R = 5
AT17:1-19, n = 140, DI = 2 and R = 3
DI = diisocyanate

1593
1594
Polymers based on ethylene oxide, end capped at both sides with: [322–325]

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628


Dodecyl alcohol (N = 12) or
Cetyl alcohol (N = 16)

Polymers based on ethylene oxide, end capped at both sides with hexyl (N = 5), octyl [188,326]
(N = 7) or dodecyl (N = 11) using either hexamethylene diisocyanate (HDI) or
bis(4-isocyanateocyclohexyl)methane (H12 MDI) as the diisocyanate group.

1: HDI with hexyl, octyl or dodecyl


2: H12 MDI with hexyl, octyl or dodecyl
or
3: HDU with hexadecyl endgroups

Polymers based on ethylene oxide, end capped at both sides with n-hexadecyl groups. [327]
Table 2 (Continued)

Polymer Structure References

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628


Polymer based on ethylene oxide, end capped at both sides with different alkyl groups: [328]

1: R1 = C17 H35 , C14 H29 or C11 H23


R2 = C6 H12
2: R3 = C16 H33 , C13 H27 , C10 H21 , C8 H17 or CH3
R4 = C13 H22
3: R5 = C9 H19 C6 H4 (OCH2 CH2 )n , n = 2, 4, 7, 10, 15
R6 = C13 H22
4: R7 = C18 H37
R8 = C7 H6 , C6 H12 , C13 H22 or C12 H24
5: R9 = C18 H37
6: R10 = C17 H37 or C10 H21
R11 = C6 H12 or C13 H22

1595
1596
Polymer based on ethylene oxide, end capped either at one side (1) or at both sides (2) [187]
with octadecyl groups (1) or with nonylphenol (2) groups:

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628


n = the number of oxyethylene units
m = the number of hydrophobes per chain
1: n = 182, 331, 531 or 663
m (182) = 1.6, 1.7 or >2.0
m (331) = 1.4, or >2.0
m (531) = 1.2, or >2.0
m (663) = >2.0
2: n = 182, 331, 531 or 663
m (all cases) = 2.0

Polymers based on ethylene oxide end capped at both sides with fluorocarbon [329]
hydrophobes:

C6 F-35K, n = 6 or
C8 F-35K, n = 8

Polymers based on ethylene oxide end capped either at both (1) or one side (2) with [186,330,331]
fluorocarbon hydrophobes.
Table 2 (Continued)

Polymer Structure References

End group: F(CF2 )8 (CH2 )11

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628


Polymers based on ethylene oxide end capped either at both or one side with [182,332–334]
fluorocarbon hydrophobes:

H17 –H17 : 1
F8 –F8 : 2, N = 0
F8 H2 –H2 F8 : 2, N = 2
F8 H10 –H10 F8 : 2, N = 10
H18 –H2 F8 : 3, N = 2
H18 –H10 F8 : 3, N = 10
H1 –H2 F8 : 4

Polymers based on ethylene oxide end capped at both ends with pyrene. [335]

1597
1598 D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628

Fig. 23. Concentration regimes of a telechelic PEO.


Reproduced with permission from [343] 1997, ACS Publishing.

urethane linkages. The polymer series 49 was synthe- ations instead of intermolecular association which occurs
sized with the idea that this would be better model for telechelic polymers.
polymers since no interference by urethane linkages is When using a combination of small spacers between the
present [332]. Indeed, it was already demonstrated by end groups and the PEO backbone, internal hydrophobe
another study [344] that significant changes in associ- phase separation was noted (polymer 37, 2–4). Accord-
ating behavior are observed when telechelic polymers ing to the authors [122] this behavior can be attributed
with an ether versus a urethane bond connecting the to the dominance of intramolecular hydrophobic associ-
hydrophobic groups with the polyethylene oxide chain are ations. To resolve this, the use of SDS was suggested. The
compared. polymer series 38(1) are prepared by step growth poly-
The association behavior of the polymer in solution merization and, as a result, a mixture of components is
depends on the molecular structure of the polymer. When obtained [320]. In principle the strength of the network in
Liu et al. [321] investigated polymer 40 a significantly solution depends on the type of hydrophobic group used
different association mechanism was observed when com- and the length of the PEO chain. The strength of the network
paring comb and telechelic polymers. Comb associative can be increased by increasing the hydrophobicity of the
polymers display a more compact structure in compari- end groups. The hydrophobicity can be increased by using
son to the association of equivalent telechelic polymers. larger sized hydrophobes, using fluorocarbon rather than
The authors [321] hypothesized, similar to the conclusion hydrocarbons, or by increasing the size of the di-isocyanate
of Jimenez-Regalado et al. [245], that comb associative group [124,320,329,330]. An example of such a poly-
polymers favour the formation of intramolecular associ- mer is polymer 41 where increasing the hydrophobicity

Fig. 24. Dimer and micelle formation for polymer 35 (2).


Reproduced with permission from [189] 1998, ACS Publishing.

Fig. 25. Superbridges, superloops and dangling ends.


D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628 1599

of the end group leads to an enhanced network. It was breakage of the bridges as the shear increases [328]. The
demonstrated that the thickening efficiency decreased in microgels reduce in size and thus also the hydrodynamic
the order of octadecyl > pentadecyl > 9-heptadecenyl [124]. volume leading to a lower viscosity of the solution. As
The same trend was observed for polymer 43 (2) where mentioned before, prior to shear thinning a weak pseu-
the thickening efficiency decreased in the order of dode- dodilatant behavior is observed. At moderate shear the
cyl > octyl > hexyl [188] and for polymer 45 (1–6) [328]. microgels are deformed, which entails the stretching of
Polymers 46, 47 and 48 (1 and 2) represent polymers with the bridges connecting the microgels [105,318,319,328].
increased network strength due to the presence of fluoro- The bridges resist breakage by stretching (up to a cer-
carbons instead of hydrocarbons. tain point) and therefore induce a shear thickening
[105,318,319]. A further increase in the shear rate leads
3.2.1. Rheological properties to the aforementioned pseudoplastic behavior. Tam et al.
Several publications presented the rheological proper- [105] investigated the network structure of polymer 36
ties of HEUR polymers in aqueous solutions. In aqueous (AT22-3) in shear flows and proposed a more refined model
solution HEUR polymers form transient networks, which than the one provided by Yekta et al. [338], depicting
follow a Maxwellian behavior with a single characteristic the different structures the network attains under shear
relaxation time [89,124,318,328,338,343,345–349]. This (Fig. 26).
relaxation time was attributed to the residence time of As the shear is increased shorter superbridges are
a hydrophobic end (group) inside a hydrophobic micro- formed but the density of mechanically active, i.e. part
domain. of the polymer network, chains increases. This leads to a
At low polymer concentration, no micelles are present stronger network and thus a higher solution viscosity. A
and the viscosity of the solution equals the viscosity of further increase of shear will lead to the shear thinning
the solute. Increasing the polymer concentration above regions. According to the study [105] three different shear
the CMC leads to micelle formation without a significant thinning regions can be distinguished for polymer 36, i.e.
increase in viscosity. A further increase in the poly- regions A, B and C. Regions 1 and 2 are characterized by
mer concentration leads to more micelles. As mentioned a significant loss in viscosity and the difference between
before, these micelles can be interpreted as star-like poly- the two is a reduced slope (viscosity versus shear stress
mers, which are known to repel each other in good plot) of the latter one. Region 3 is characterized by a large
solvents [89,345–347,350]. However, telechelic micelles drop in the solution viscosity with increasing shear. The
have the possibility of forming bridges between two sep- difference between regions 1–2 and 3 is the mechanism
arate hydrophobic cores, thus leading to cluster formation causing the loss in viscosity. It was demonstrated that in
between them [341]. The viscosity of the polymer solu- regions 1–2 the superbridge network junctions disrupt,
tion now changes due to an increase in the hydrodynamic but they rearrange quickly without a loss in the num-
volume of the micellar clusters formed. Raising the poly- ber of bridges. The reduction in the relaxation time leads
mer concentration leads to networks (microgels) which to the reduction in viscosity for regions 1 and 2 given
are connected through bridges [351,352]. These microgels that the viscosity is dependent on two separate processes
break upon dilution of the polymer solution. Two studies according to the  = G0 equation ( = relaxation time and
from the Annable group [89,350] proposed similar network G0 = plateau modulus) for a Maxwell fluid [105]. The loss in
formation where micelles (or a collection of micelles) are viscosity for region 3 is attributed to fragmentation of the
connected through superbridges and superloops. Super- network.
bridges connect several micelles in a linear fashion whereas
superloops connect several micelles in a non-linear 3.2.2. Effects of inherent parameters
fashion, but the start and end of this superloop rest in Telechelic polymers such as the polymer series 35
the same micelle [89,350]. In addition, dangling ends are display in steady shear a Newtonian behavior up to rela-
present in the microgels, which are chains with only one tively high shear rates [105]. This is in stark contrast with
hydrophobic end associated with other hydrophobic enti- polymers bearing hydrophobic moieties distributed along
ties (Fig. 25). their backbone, where shear thinning predominates. When
By using fluorescence spectroscopy, it has been demon- incorporating the hydrophobic group along the polymer
strated that a critical association concentration (CAC) exists backbone, the thickening capability, i.e. solution viscos-
for telechelic polymers end-capped with alkyl groups ity as a function of the polymer concentration, is lower
[89,128] (polymer 34). Above the CAC the viscosity of the when compared to telechelic polymer [89]. The comb asso-
polymer solution increases more sharply with increasing ciative polymer 40 is an example of a polymer with less
concentration when compared to polymers lacking the pronounced increase in solution viscosity when compared
hydrophobic groups [128]. to a telechelic equivalent. However, it must be mentioned
In HEUR solution the interchange of the bridges con- that although both polymers are quite similar, the molecu-
necting the micellar clusters is fast at low shear rates, lar weight of the comb type is higher than for the telechelic
without breakage by shear [170] and thus no loss of one. In addition, it was demonstrated that the distribu-
solution viscosity. Nevertheless, HEUR polymers also dis- tion of the hydrophobic groups plays a crucial role in
play pseudoplastic and pseudodilatant behavior. Under the observed behavior. When each of the hydrophobic
steady shear, the HEUR polymeric solutions display a strong moieties consist of two hydrocarbon groups the thicken-
pseudoplastic behavior that is preceded by a weak shear ing capability is significantly enhanced. This is similar to
thickening. The pseudoplastic behavior arises due to the the behavior observed for twin-tailed HMPAM polymer
1600 D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628

15 [321]. The hydrophobicity of the groups increases and concentration for the onset of hydrophobic association
therefore a stronger interaction gives rise to the observed is also affected by the molecular weight of the polymer.
enhanced thickening capability. In light of this, the com- According to Kaczmarski and Glass [187] the increase in
parison between a single comb and telechelic polymers of the number of EO units (increase in molecular weight)
polymer 40 is not viable given the low hydrophobicity of promotes the formation of intermolecular associations
the groups in the comb polymer. and thus increasing the network strength leading to an
Internal hydrophobes affect the rheological properties enhanced solution viscosity. At low molecular weights, a
of HEUR solution due to the preference for intramolecular reduction in the solution viscosity is observed due to a
hydrophobic associations. However, there are conflicting reduction in the hydrodynamic volume of the polymer. At
reports on the effects of internal hydrophobes. According high molecular weights, also a reduction in solution vis-
to Fonnum et al. [328] internal hydrophobes do not affect cosity is observed. This is attributed to the decrease in the
the rheology of polymeric solution employing polymer 45. concentration of the hydrophobic groups, which leads to
On the other hand Lundberg et al. [122] concluded that less hydrophobic micro-domains being formed and thus
using HEUR polymer 37 (with an internal hydrophobe in a weaker network. Polymers 42, 43, 45 (1–6) and 47 are
combination with two terminal hydrophobes) resulted in classic examples of polymers whose rheological properties
a more compact structure due to intramolecular hydropho- depend on the molecular weight. The molecular weight is
bic associations rather than intermolecular (vida supra). closely related to the PDI. The PDI also affects the rheo-
This leads to a reduction of the hydrodynamic volume of logical properties of the polymeric solution. A study [325]
the polymer chains, and thus a reduction in the solution compared the thickening capability of the polymers and
viscosity. found significant differences in the viscosity enhancement
The hydrophobicity of the end groups affects the rhe- capability. For instance, when a broad MWD polymer is
ological properties. According to Calvet et al. [330] the used twice as much polymer (concentration) is required
increase in hydrophobicity of the groups leads to an to attain the same viscosity of a narrow MWD polymer.
increase in the lifetime that the hydrophobic end groups However, the storage and loss moduli were similar despite
are present in the hydrophobic core. It has also been the concentration difference. Broad MWD polymers con-
demonstrated that the monofunctionalized polymer 48 (2) tain small polymer chains of low molecular weight and
displays a marked difference in rheology compared to the these polymers favor the formation of intramolecular
difunctionalized polymer 48 (1) [330]. The concentration hydrophobic associations. This leads to a weakened poly-
at which the onset of significant (static) viscosity increase mer network, which reduces the viscosity enhancement
is observed is much lower for the difunctionalized poly- capability.
mer. However, it cannot be concluded that this difference
is only caused by the single end capping since the molecu- 3.2.3. Effects of external parameters
lar weight of the monofunctionalized polymer is half that Interactions between surfactants and HEUR polymers
of the difunctionalized one. in solution have been demonstrated by several authors
The rheological properties of a HEUR solution also [127,353,354]. Depending on the concentration of the
depend on the molecular weight of the polymers. In prin- surfactant, an increase in viscosity of the solution can
ciple, the molecular weight of the HEUR polymer has an be observed. Interactions between polymers 34, 36 (AT
optimal intermediate range where the highest thicken- 22-3), 37 (1), 41, 49 and 50 and the surfactants SDS
ing capability can be obtained [331]. The critical polymer [124,164,182,335,353–357], DTAB [354], nonylphenol

Fig. 26. Model depicting the effect of shear on the network structure.
Reproduced with permission from [338] 1998, ACS Publishing.
D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628 1601

Fig. 27. Different aggregates formed with the addition of surfactant.


Reproduced with permission from [127] 1996, ACS Publishing.

ethoxylates (NP) with different degree of ethoxylation of the polymer chain has been found [164,182,354–356].
[353], polyoxyethylene glycol (C12 E8 , nonionic surfactant) In comparison, cationic surfactants only interact with the
[127] and polyoxyethylene stearyl ether (C18 [EO]20 ) [358] hydrophobic end groups [127,354]. Interactions between
have been investigated. The effect of the surfactant has nonionic surfactants have been demonstrated where a
been summarized by Annable et al. [353] in terms of the large increase in solution viscosity is observed upon addi-
plateau modulus: (1) at low polymer concentration the tion of the surfactant [124,127,353,358,359]. The study by
plateau modulus is significantly raised using a low to Kim et al. [358] investigated the stability of the polymeric
moderate amount of surfactant, (2) using higher surfactant solution employing the HEUR polymer 37(1) with the
concentration leads to a monotonic decrease in the plateau polyoxyethylene stearyl ether (C18 [EO]20 ) surfactant in
modulus, (3) at high polymer concentration no peak in complex formulations. An important observation made,
the modulus is observed and (4) the modulus peak shifts in the context of EOR-applications, was the stability of
towards lower surfactant concentration as the polymer the micellar network in solution in an oil-water emulsion,
concentration is lowered. According to Alami et al. [127] despite the changes in ionic strength (different NaCl
the polymer concentration at which the onset of hydropho- concentrations) and pH.
bic micro-domains arise can be influenced by the addition In order to facilitate the investigation of molecular
of nonionic surfactants. The hydrophobic micro-domains weight of HEUR polymer the hydrophobic associations
form at much lower concentrations than the CAC or CMC have to be suppressed. Cyclodextrins, cyclic oligomers
when nonionic surfactants are present in the polymer solu- consisting of 6, 7 or 8 glucose units (corresponding to ␣-,
tion. The study by Alami et al. [127] provided a schematic ␤-, ␥-CD) connected by ␣-1,4-glucosidic linkages [360],
presentation of the interactions between the SDS surfac- are capable of encapsulating the hydrophobic groups
tant and HEUR polymers in solution. Depending on the thus inhibiting hydrophobic associations. The interaction
concentration of the surfactant and the polymer, different between cyclodextrins and the HEUR polymers 37(1), 44
structures (associated polymer aggregates) are formed and 49 (2, F8 –F8 ) support this. A schematic presentation
(Fig. 27). illustrating the interactions at different stages is given in
The study [127] also shows that it is difficult to dis- Fig. 28.
tinguish a concentration region where only micelles are Depending on the cyclodextrin concentration three
formed when a HEUR polymer with a high molecular different regions can be distinguished [314,361,362]. At
weight is used. The formation of micelle clusters starts close low cyclodextrin concentration encapsulation of the end
to the CMC. groups of the loops occurs making the network more
At low (SDS) surfactant concentration susceptible to shear thinning. At moderate cyclodex-
(csurfactant < cmcsurfactant ) the interaction of the surfac- trin concentration detachment of the bridging chains
tant is mainly with the hydrophobic ends of the polymer between micelle clusters occurs, which weakens the net-
chains, which leads to a stronger network and thus an work and thus results in a lower viscosity. A further
enhanced solution viscosity [124,127,164,182,353–356]. increase in the cyclodextrin concentration leads to the
High surfactant concentration leads to a destabilized complete destruction of the network formation leaving
network as the hydrophobic end groups are solubilized, only individual clusters or polymer chains in solution. The
which results in a reduction in the viscosity. In addition an viscosity of such a polymeric solution equals the solute
interaction between the SDS surfactant and oxyethylene viscosity.
1602 D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628

Fig. 28. Interactions between cyclodextrins and a HEUR polymer.


Reproduced with permission from [361] 2007, ACS Publishing.

3.3. Hydrophobically modified alkali swellable emulsion and dissolution follows. These polymers have not been
(HASE) studied as extensively as the HMPAM or HEUR polymers
[363]. However they represent an alternative to the non-
HASE systems are emulsions at low pH where the ionic HEUR polymers for applications as thickeners [363].
polymers exist as macroscopic particles (non-transparent HASE polymers are terpolymers consisting of methacrylic
solution). As the pH is increased the particles will swell acid, ethyl acrylate and a hydrophobic group [169,364,365].

Fig. 29. Chemical structure of a HASE polymer.


D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628 1603

The HASE polymers are, as mentioned before, insoluble neous. Two distinct regions can be identified, which are
in low pH solutions and require higher pH (above pH 6) a hydrophobe enriched one and a hydrophobe depleted
for solubilization. This, given the salinity of most reser- one. The hydrophobe enriched region will form more
voir water, is beneficial for application in EOR. In most compact structures compared to the other region. By using
cases the ratio between methacrylic acid and ethyl acry- fluorescence studies, Prazeres et al. [374] investigated the
late (i.e. molar ratio x/y) equals 1 and the hydrophobic association level in the polymer series 54 and found that in
group is incorporated at a level of a few mol percentages aqueous solution 58% of the hydrophobic pendants asso-
(%). The HASE polymers can be seen as polyelectrolytes, ciate compared to only 5% in tetrahydrofuran (THF). The
which have been hydrophobically modified by the intro- same authors demonstrated though, that the coil expan-
duction of a few mol% of a hydrophobic group. Comb-like sion caused by the aqueous basic conditions will lead to a
structure where the hydrophobic groups are placed on large portion of the hydrophobic pendants (21 ± 6%) being
short PEO side chains randomly distributed along the back- excluded from associated microdomains. This portion is
bone of the HASE polymer chains are favored [169,363]. much larger when compared to polymer solutions in THF.
The backbone of the HASE polymers possesses moderate Varying the length of the hydrophobe groups and the
high molecular weight and can reach 200,000–250,000 Da oxyethylene spacers will induce changes in the strength of
[169]. The general chemical structure of a HASE polymer the polymer network. Increasing the hydrophobe length,
[106,169,366,367] is given in Fig. 29 while a more general i.e. the hydrophobicity, will lead to stronger associations
overview of all studied systems can be found in Table 3. and thus a stronger network. This behavior was demon-
HASE polymers are synthesized using an emul- strated by Tirtaatmadja et al. [367] which is in line with
sion copolymerization process [181,363,372]. The three findings using polyacrylamides and poly(ethylene oxide).
constituents, methacrylic acid, ethyl acrylate and the Increasing the length of the oxyethylene spacers increases
hydrophobic macromonomer, are added to the reac- the accessibility of the hydrophobes, which in turn leads to
tion mixture along with an initiator. The hydrophobic easier formation of hydrophobic domains [169] and there-
macromonomers act as the surfactant for the polymeriza- fore a stronger network. However, if the length exceeds
tion reaction. These are produced by first ethoxylating a a certain critical length intramolecular associations are
hydrophobic group and subsequently reacting the prod- favored causing a collapse of the network [169].
uct of the ethoxylation with an unsaturated isocyanate
[363,364,367]. The reaction product of the emulsion poly- 3.3.1. Rheological properties
merization is the HASE polymer. The rheological properties of HASE polymers in aqueous
Different types of hydrophobic groups have been used solution are dependent on the solubilization conditions.
which include alkyl groups [106,363,367,369,370], com- When using alkali media, the backbone of the HASE
plex poly(alkylaryl) moieties [91,363] and hydrophobic polymer chain becomes hydrophilic and will dissolve in
groups containing pyrene for investigation of the associ- aqueous solutions. The hydrophobic groups of the HASE
ation structures [365,373–375]. polymer will form associations, either intra- or interchain
The length of the PEO groups to which the hydropho- interactions, which will cause a significant increase in the
bic groups are attached can be manipulated by controlling solution viscosity [369].
the number of CH2 CH2 O moieties [169,363]. In addition the The dissolution of HASE polymers has been investigated
length of the hydrophobic group can be altered by using by several authors [167,168,377]. The data obtained using
longer or shorter hydrophobic agents [367]. The hydropho- potentiometric titration, isothermal titration calorimetry
bicity of the hydrophobic groups can be manipulated [169]. and dynamic light scattering studies, led to the elucidation
The number of hydrophobes per chain can be controlled of the dissolution mechanism of the HASE polymers, either
by the variation of the proportion of the macromonomer in the presence or absence of salt (Fig. 31).
(Fig. 29) in the recipe during the polymerization stage. The size (hydrodynamic radius, Rh ) of the latex par-
These controls offer unique possibilities for tailoring the ticles increase in size as the pH rises (increase in base
chemical structure, and thus the rheological properties of concentration) due to the electrostatic repulsions. For a
the HASE polymer solutions [376]. HASE polymer latex particle in a salt solution the size
The structure in solution is dependent on the molecular increases but to a lesser extent due to shielding of the neg-
structure of the HASE polymer. Giving the molecular struc- ative charges on the polymer backbone by the cations, i.e.
ture of the polymers 51, a comb-like structure is obtained Rh, no salt > Rh, salt
with hydrophobic groups tethered through oxyethylene The electrostatic repulsion forces between the chains
spacers to the polymer backbone [363]. also alter the configuration of the polymer aggregates. The
The structure of HASE polymer in aqueous solution has thickening capability of the HASE polymers is a combina-
been elucidated (see Fig. 30). By using superposition of tion of hydrophobic association and the expansion of the
oscillations on steady shear flow [366] of HASE solutions polymer chains [181,370]. At low polymer concentration,
or in simple salt solution [369] a model was developed. the intramolecular hydrophobic interactions dominate
Above a given polymer concentration, the unimers will while at higher polymer concentration the situation arises
form oligomers consisting of two or more polymer chains. where predominantly intermolecular hydrophobic inter-
When the polymer concentration increases, the size of actions are present [370]. The intermolecular hydrophobic
the aggregates will increase. Using spin-echo (PGSE) NMR interactions give rise to network formation thus enhanc-
Nagashima et al. [370] concluded that the distribution ing the viscosity of the polymer solution [366,367,370].
of hydrophobes among the polymer chains is inhomoge- Knaebel et al. [378] provided a schematic presentation
1604
Table 3
Structure of different water-soluble polymers, HASE.

Polymer Structure References

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628


Polymer based on methacrylic acid (MAA), ethyl acrylate (EA) and a hydrophobic [169]
macromonomer (associative macromonomer, AM).

R = H (1), C8 H17 (2), C12 H25 (3), C16 H33 (4) or C20 H41 (5)
m = 33 (for 2)
m = 35 (for 1, 3–5)
X–Y–Z (1) = 48.76–51.24–0.00 (mol%)
X–Y–Z (2) = 49.00–50.00–1.00 (mol%)
X–Y–Z (3) = 49.05–50.05–0.90 (mol%)
X–Y–Z (4) = 49.06–50.04–0.90 (mol%)
X–Y–Z (5) = 49.06–50.04–0.90 (mol%)

Polymer based on methacrylic acid (MAA), ethyl acrylate (EA) and a hydrophobic [106,169,366,367]
macromonomer (associative macromonomer, AM).

Serie A, N = 19 and m = 0, 10, 15, 35 or 40


Serie B, m = 35 and N = 4, 8, 12, 16 or 20
X–Y–Z (A&B) = 49.06–50.01–0.90 (mol%)
Table 3 (Continued)

Polymer Structure References

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628


Polymer based on methacrylic acid (MAA), ethyl acrylate (EA) and a hydrophobic [363,366–368]
macromonomer (associative macromonomer, AM).

m = 80
X–Y–Z = 40.00–40.00–20.00 (wt.%)

1605
1606
D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628
Polymer based on methacrylic acid (MAA), ethyl acrylate (EA) and a hydrophobic [91,371]
macromonomer (associative macromonomer, AM).

m = 53
X–Y–Z (Refs. [169,369,370]) = 45.00–55.00–<1.00 (mol%)
X–Y–Z (HASE65) = 0.49–0.49–0.021
X–Y–Z (HASE48) = 0.50–0.49–0.011
X–Y–Z (HASE42) = 0.44–0.55–0.011
X–Y–Z (HASE35) = 0.45–0.55–0.006
X–Y–Z (HASE12) = 0.40–0.59–0.002
X–Y–Z (HASE11) = 0.40–0.69–0.0035
X–Y–Z (HASE0) = 0.40–0.59–0.002
D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628 1607

Fig. 30. Microstructure as proposed by Dai et al. [369].


Reproduced with permission from [369] 2000, ACS Publishing.

illustrating the structural transitions that a HASE polymer concentration consist of a unimer with the hydropho-
undergoes when the polymer concentration is increased, bic domain in the center and the hydrophilic backbone
from low to a high concentration, in an alkaline polymer extended in the solution. As the polymer concentration
solution (Fig. 32). is increased a transition to an overlap regime occurs. This
A schematic presentation of the microstructure of the transition takes place above the cross-over concentration
polymer chains at very low polymer concentration in an (Cg ) [378] and leads to a significant rise in the viscosity of
alkaline solution, as proposed by Dai et al. [369], is given in the HASE polymer solution due to the interactions between
Fig. 32. At low polymer concentration a colloidal suspen- the hydrophobic groups. The structure of the network has
sion is present, i.e. no solubilization of the polymer chains. been investigated and a schematic presentation has been
The microstructure of the polymer chains at low polymer provided by Tan et al. [379] (Fig. 33).

Fig. 31. Dissolution mechanism in the absence and presence of salt.


Reproduced with permission from [168] 2002, ACS Publishing.

Fig. 32. Concentration regimes for HASE polymers in an alkaline solution.


Reproduced with permission from [378] 2002, John Wiley and Sons.
1608 D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628

There are a number of different interactions in into a higher steady shear and zero shear solution viscos-
the network of HASE polymers, which are the intra- ity but also to a more pronounced pseudoplastic behavior.
and intermolecular hydrophobic interactions, polymer Using the polymer series 51 the positive effect of the
chain entanglements and electrostatic interactions. Araujo length or size of the hydrophobe on the solution viscosity
et al. [180] demonstrated the presence of two different was demonstrated [367,376]. The highest solution viscosity
hydrophobic micro-domains using various pyrene deriva- was obtained with the longest/largest hydrophobic group
tives as probes. The ethyl acrylate rich micro-domains (C20 ). Using polymer 52 English et al. [371] demonstrated
(Fig. 33) are what Araujo et al. [180] envisaged. that by using more complex hydrophobic groups the net-
Most HASE polymer solutions exhibit a strong pseu- work dynamics of the polymer is no longer represented
doplastic behavior when sheared, which is preceded by a by a single characteristic relaxation time but by rather
weak shear thickening region [106,378]. The latter arises the hindered reptation model put forward by Leibler et al.
due to the disruption of the hydrophobic intermolecu- [291]. The deviation of the classical Maxwellian behav-
lar associations by shear [365]. Thixotropic behavior was ior arises by the coexistence of hydrophobic associations
also noticed by Knaebel et al. [378], similar to polymeric and chain entanglements [371]. Ng et al. [381] investi-
solutions containing HMPAM. Increasing the polymer con- gated the difference in solution viscosity in HASE polymers
centration leads to an increase in the zero-shear viscosity. compared to the unmodified analogue. Fig. 34 schemat-
However, the shear rate at the onset of shear thinning ically represents the differences hypothesized by the
shifts to lower values when the polymer concentration is authors.
increased [378]. The hydrophobic groups will interact with the ethyl
acrylate blocks in the semi-dilute concentration regime.
3.3.2. Effect of inherent parameters These in turn will enhance the intermolecular associations
The effects of length of the PEO spacers, hydrophobic- between different clusters. The higher solution viscosity of
ity of the hydrophobic group, length of the hydrophobic the hydrophobically modified polymers compared to the
groups and the molecular weight of the HASE polymer unmodified analogues is attributed to this enhancement.
on the rheological properties of HASE polymer solution As mentioned before the hydrophobicity of the hydropho-
have been elucidated. As discussed earlier, when the bic groups will affect the rheological properties of the HASE
length of the PEO spacers is increased the accessibility polymer. A classic example of this is the polymer series 54
of the hydrophobic groups for hydrophobic associations [365]. The polymer with the pyrene functionality displayed
increases. This leads to a lower number of larger hydropho- a higher solution viscosity compared to the same polymer
bic micro-domains. However if the length of the PEO without pyrene. However, the number of intermolecular
spacers surpasses a critical length intramolecular associa- hydrophobic associations that arise is dependent on the
tions will be the dominant type of hydrophobic interaction pyrene content with more being formed at lower pyrene
[169]. The strength of the polymer network reduces and contents.
a decrease in the solution viscosity follows. This was The degree of ethoxylation affects the thickening capa-
observed by Tam et al. [380] who demonstrated that bility of HASE polymers. The optimum ethoxylation degree
PEO spacers with a moderate length displayed the high- lies between 5-40 mol% of the ethylenoxide on the
est steady shear and zero shear solution viscosity. The hydrophobic monomer [380]. According to Wu and Shay
more pronounced shear thinning and viscoelastic behav- [376], the balance between hydrophobic associations and
ior arose with the same polymer. Changing the type of electrostatic repulsion can be manipulated by the incor-
hydrophobic groups, i.e. its hydrophobicity or length, also poration level of the methacrylic acid moiety [376]. The
affects the polymer network and thus its rheological prop- maximum zero shear rate viscosity is obtained with an
erties. Increasing the hydrophobicity or the length of incorporation rate of 40 wt%. A classic example of this
the hydrophobic groups will increase the strength of the behavior is observed for polymer 51. In addition it was
hydrophobic associations [169,367,376]. This translates demonstrated that the optimum ethoxylation degree cor-

Fig. 33. Network structure of a HASE polymer.


Reproduced with permission from [379] 2001, John Wiley and Sons.
D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628 1609

Fig. 34. Schematic presentation of the associations in HASE and its unmodified analogue.

responds to the dominant elastic property of the polymer pyrene [181]. At a pH of 5.7 a large number of hydropho-
system at 10 mol% ethyleneoxide [380]. bic micro-domains are present. As the pH of the polymer
Another factor that affects the rheological properties is solution is raised a smaller number of larger hydropho-
the molecular weight and this can be controlled by the bic domains are formed. The authors suggest a transition
amount of chain transfer agent (CTA) used during the emul- from intramolecular to intermolecular hydrophobic asso-
sion polymerization. Increase in molecular weight results ciations. This leads to a significant increase in the solution
in higher zero shear and steady shear solution viscosity viscosity.
[376], unlike HEUR polymers where the exact opposite Adding salt to the polymer solution affected its rheolog-
behavior is observed. In addition the authors observed that ical properties. Increase in the salt concentration resulted
a reduction in molecular weight leads to a system where in a decrease of the solution viscosity with the effect being
the Newtonian plateau is more extended. more pronounced for groups with higher hydrophobicity.
However, the polymer 51 (with C20 H41 as the hydrophobe)
3.3.3. Effect of external parameters displayed a shear thickening behavior at a given salt con-
The effect of pH and ionic strength on the rheology of centration [368]. The first effect, significant reduction in
the HASE polymer solutions has also been investigated. the solution viscosity [379] with the addition of salt, is
At relatively low pH, the HASE polymers are insoluble attributed to the shielding of the charges on the polymer
given the uncharged nature of the carboxylic acid groups backbone leading to the shrinkage of the polymer back-
[181]. Due the presence of methylacrylic acid in the back- bone. The intermolecular hydrophobic associations will be
bone of the HASE polymer alkaline pH will ionize these destroyed due to the shrinkage. The second effect, shear
groups and will lead to swelling of the polymer and the thickening with the addition of more salt, is attributed to
formation of a polyelectrolyte with solubility in water the reformation of the network. The authors [379] have
[169,181,367,368,370]. The swelling of the polymer is a provided a model illustrating the network structures in salt
result of the electrostatic repulsion between the poly- solution under shear deformation (Fig. 35).
electrolyte chains [169,368,379]. As the pH is increased a Several different studies [91,106,375,382–384] have
smaller number of larger hydrophobic micro-domains are been published on the interactions between HASE poly-
formed, which leads to an increase in the solution viscos- mers and different surfactants. Polymers 51 (5) and 53 have
ity [181]. The expansion of the polymer chain is partly been shown to interact with surfactants. Depending on
a result of the dissolution of the particle (present prior the nature of the surfactant different behaviors can arise
to the solubilization) [181]. In order to investigate the [91]. Using a partially water-soluble nonionic surfactant
association mechanisms of HASE polymers the hydropho- (nonylphenol ethoxylate) English et al. [91] demon-
bic micro-domains of polymer 51 (5) were probed using strated that the shear rate range where shear induced

Fig. 35. Model illustrating the network structures of HASE in salt under shear.
Reproduced with permission from [379] 2001, John Wiley and Sons.
1610 D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628

Fig. 36. Interactions between nonionic surfactant and HASE polymers.


Reproduced with permission from [384] 1998, John Wiley and Sons.

structuring of the polymer network occurs can be widened. on the rheological property arises due to the same effect as
The use of water-soluble nonionic surfactant, at high suffi- with nonionic surfactant. Seng et al. [382] investigated the
cient surfactant concentration, resulted in the loss of shear interactions between polymer 51 (5) and the SDS surfac-
induced structuring of the polymer network. It was further tant. Above a given critical surfactant concentration (CS*),
demonstrated that, when using water-soluble surfactants, closed to the CMC of the surfactant, the aforementioned
segregation occurs where two distinct systems exists, one positive effect is lost due to the disruption of the poly-
phase rich in polymer and one rich in surfactant. On the mer network [382]. This leads to a reduction in the number
contrary, by using a partially soluble surfactant no evi- of hydrophobic micro-domains. In addition free cations of
dence of complete segregation was found. A few studies the surfactant molecules screen the anionic charges on the
[382–384]demonstrated the increase in viscosity of the HASE polymer backbone, which reduces the electrostatic
polymeric solution, employing polymer 51 (5), when using repulsion between chains. A model representing the dif-
nonionic surfactant (C12 EO4 ). The increase in the viscos- ferent molecular interactions between a ionic surfactant
ity was noticed up to a surfactant concentration of 0.1 M. and polymer 51 (5) has been provided by Seng et al. [382]
The studies attributed the increase in solution viscosity (Fig. 37).
to the formation of mixed micelles of surfactant and the The interactions of SDS with polymer 51 (5) can be
hydrophobic groups of the HASE polymer chains. At sur- divided into three regions (1, 2 and 3 in Fig. 37). In regions
factant concentrations above 0.01 M stronger interactions 1 and 2 the effects of the polymer–surfactant interactions
arise, thus a strengthening of the polymer network, where on the rheology are limited. However, in region 3 weaken-
the formation of a stiff gel-like structure occurs. At this ing and destruction of the polymer network arise due to on
stage a bilayer is formed by the surfactant molecules, which one hand shielding of the charges (less electrostatic repul-
are stabilized as large vesicles. A schematic presentation, as sion and thus decrease in hydrodynamic volume) and on
provided by Tirtaatmadja et al. [384], of the interactions the other hand the isolation of the hydrophobes in micelles
between the nonionic surfactant, at different surfactant thus suppressing intermolecular hydrophobic associations.
concentration, and the HASE polymer is given in Fig. 36. Both of these effects lead to a lower solution viscosity.
The temperature affects the interactions between the The same interaction behavior is observed with SDS and
surfactant and HASE polymers. At low to moderate (non- polymer 53 [373]. This in marked contrast to the behav-
ionic) surfactant concentration the solution shear viscosity ior of nonionic surfactants discussed earlier. Nevertheless
decreases with increasing temperature. Above a certain the same behavior, loss in solution viscosity at high surfac-
shear rate a substantial decrease in viscosity is observed, tant concentration, has also been reported for a nonionic
which can be related to the characteristic relaxation time surfactant (polyoxyethylene) [106].
associated with the dissociation of hydrophobes from
hydrophobic domains. The rate of dissociation is much 3.4. Hydrophobically modified cellulose derivatives
higher than the rate of association [385]. However this con-
cept, i.e. the relation between characteristic relaxation time The current demand for environmental-friendly
and exit rate of hydrophobic groups, as proposed by Aubry materials is relevant also for EOR applications. Hydropho-
and Moan is controversial and needs verification [383]. Tam bically modified hydroxyethyl cellulose (HM-HEC),
et al. [106] investigated the interactions between polymer hydrophobically–hydrophilically modified hydroxyethyl
51 (5) and the nonionic surfactant (C12 EO23 ). The addition cellulose (HHM-HEC), hydrophobically modified hydrox-
of the nonionic surfactant resulted in the strengthening ypropyl cellulose (HM-HPC) or hydrophobically modified
of the polymer network by the increase of the number of ethyl hydroxyethyl cellulose (HM-EHEC) polymers can
intermolecular hydrophobic associations similar to when meet these demands. The parent cellulose polymer is a
C12 EO4 was used as the nonionic surfactant. According to nonionic polysaccharide, which possesses anti-bacterial
the study [106] the polymer-surfactant interactions are properties [386]. Parent cellulose polymers are hydropho-
entropically driven. bically or/and hydrophilically modified to produce the
Interactions between ionic surfactants and HASE poly- HM-HEC, HHM-HEC, HM-EHEC or HM-HPC polymers. A
mers have also been demonstrated [373,382]. The positive general structure of hydrophobically modified cellulosic
effect, i.e. increase in solution viscosity, of the surfactant polymer is presented in Fig. 38.
D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628 1611

Fig. 37. Molecular interactions between SDS and polymer 43 (5).

A general overview of the reported systems is given in group also affects the strength of the network. Increas-
Table 4. ing the hydrophobicity will increase the strength. Polymer
A synthesis method, described by Miyajima et al. [406], 57 and 60 (R1 = 2) are examples where the polymer with
for HM-HEC polymers involves a one-pot synthesis. A reac- fluorocarbon provide a stronger network than the hydro-
tor is charged with the HEC polymer and isopropyl alcohol carbon (un)modified analogue [391]. The distribution of the
together with an aqueous solution (48%) of sodium hydrox- hydrophobic groups is also of crucial importance. Polymer
ide. This slurry is stirred at room temperature for 30 min 62 represents an example in which, due to inhomoge-
under a nitrogen atmosphere after which the hydrophobic neous distribution of the hydrophobic group, two separate
reagent is introduced. The reaction is then performed for phases of polymers are observed [401]. One containing
several hours (e.g. 8 h) at 80 ◦ C. Methods for the synthe- highly substituted chains and another with less substitu-
sis of different HM-EHEC polymers have been developed tion. The highly substituted chains will be preferentially
and provided by Boström et al. [407] and Landoll [98]. incorporated into the network.
These comprise several steps. First the cellulose is isolated Classical polyelectrolyte behavior can be obtained by
from wood pulp using standard techniques, i.e. using a the incorporation of charged moieties in the backbone
21.5 wt% aqueous sodium hydroxide solution for 30 min. of the polysaccharides. An example of this behavior is
Then hydroxyethyl groups are introduced by reaction with observed with polymer 60 (R1 = 3) where cationic moieties
ethylene oxide at 50 ◦ C for 75 min. The last step involves are incorporated into the backbone of the polysaccharide
the incorporation of the hydrophobic group by reaction at [397].
105 ◦ C for 120 min.
The presence of hydrophobic groups on the polymer
backbone of the polysaccharides will cause the formation 3.4.1. Rheological properties
of associations just like the HMPAM, HEUR and HASE ana- Although different types of hydrophobically modified
logues. The network structure is enhanced due to these polysaccharides have been synthesized their rheological
groups and the hydrophobic associations arise at very low properties in aqueous solution depend on the modifi-
polymer concentration [118,399,405,407–409]. Polymers cation rather than on the type of polysaccharide used.
56, 57, 58, 61 and 63 are classic examples of this behav- The hydrophobic modifications introduced on the differ-
ior with associations commencing at very low polymer ent cellulose derivatives decrease the solubility of the
concentrations. When good solvents are used, e.g. alco- polysaccharide. Several studies have demonstrated the
hols or dioxane, no associations arise [118]. Polymers 56, pseudoplasticity of hydrophobically modified polysaccha-
57, 58 and 63 impart a stronger network when compared rides. Typical shear thinning is observed similar to the
to their unmodified analogues. The type of hydrophobic hydrophobically modified polyacrylamides. However, the
concentration of the polymeric solution has to be high
enough before pseudoplastic behavior is observed. Polymer
59 is a classic example of this behavior [395]. In addi-
tion shear thickening behavior is observed. The presence
of shear thickening only arises within a certain concentra-
tion range of approximately 0.15–0.5 wt%. The zero-shear
solution viscosity of a polymer solution is much more
dependent on the concentration in the lower concentra-
tion regime. Increasing the polymer concentration also
Fig. 38. General chemical structure of hydrophobically modified cellu- decreases the dependency of the zero-shear viscosity on
losic polymer. concentration [410]. Examples of the shear thickening
1612
Table 4
Structure of different water-soluble polymers, HMHEC.

Polymer Structure References

HHM-HEC

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628


The hydrophobic groups are long alkyl chains and the hydrophilic groups are sodium sulfonate [106]
groups.

HHM-HEC, R2 = –n–C18 H37 and R3 = 1


S-HEC, R2 = H and R3 = 1
R-HEC, R2 = –n–C18 H37 and R3 = H

HM-HEC

Commercial product, produced by different suppliers. [386]


Table 4 (Continued)

Polymer Structure References

HM-HEC

The hydrophobic groups are fluorocarbon chains [386–389]

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628


HM-HEC

EP16-HAHEC, R2 = 1 [390]

IPBC-HAHEC, R2 = 2
BD-HAHEC, R2 = 3

HM-HEC

Hydrophobic groups are alkyl chains varying between 8 and 40 carbons in length [391]

1613
1614
HM-HEC

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628


R3 , R4 and R5 = alkyl, aryl, aralkyl, alkaryl, cycloalkyl, alkoxyalkyl and alkoxyaryl where a [392–394]
maximum length of 10 C-atoms (the total length of R3 , R4 and R5 is between 3 and 12 C-atoms.
When R3 , R4 or R5 is an alkoxyalkyl radical two C-atoms separate the O and N-atom.

The N-atom is a component of: pyridine, ␣-methylpyridine, 2,5-dimethylpyridine,


2,4,6-trimethylpyridine, N-methylpiperidine, N-ethyl piperdine, N-methyl morpholine or
N-ethyl morpholine.
R1 = 2, flourocarbon groups
R1 = 3, cationic groups
R1 = 3 and 4, cationic and hydrophobic groups randomly distributed
R1 = 5, cationic hydrophobic groups
Table 4 (Continued)

Polymer Structure References

HM-HPC

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628


The polysaccharide is labeled with pyrene (1) or fluorene (2) groups for fluorescence studies [395,396]

HM-EHEC

The hydrophobic groups are either long alkyl chains or a nonylphenol group [397–399]

1615
1616
HM-HEC-py-1

The hydrophobic groups are an alkyl chain and a pyrene group for fluorescence studies [118,400]

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628


HM-CMC

The hydrophobic groups are amino-terminated poly(N-isopropylacrylamide) [307,401–404]

HM-Alginate

The hydrophobic groups are n-octylamines [405]


D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628 1617

behavior are polymers 55 [386,387] and 59 [410,411] and in solution viscosity is not as pronounced [412]. Solutions
an example of a polymer which does not display shear containing polymer 56 display this behavior.
thickening due to a high concentration is polymer 58 The ratio between the hydrophobic/hydrophilic sub-
[392–394]. stitution significantly affects the solution properties at
Thixotropic behavior is also observed for the hydropho- low polymer concentration [389]. The thickening capa-
bically modified polysaccharides. Similar to other bility is increased when this ratio is high. Polymer 55
hydrophobically modified polymers, such as HMPAM, is an example of a polymer whose thickening capability
hysteresis effects can be seen for polymeric solution increases with a high hydrophobic/hydrophilic substitu-
employing polymer 55 (HHM-HEC, R-HEC) [387], 56 [412], tion [387,388]. Fig. 39 displays the effect of the presence
58 [392–394], 59 (hydrophobic groups are hexadecyl of salt on the polymer network that has been envisaged by
moieties) and 60 (R1 = 2, R1 = 5) [397,399]. The viscosity Kawakami et al. [387].
of the polymeric solution at increasing shear rates is As the concentration of the salt increases, the electro-
higher compared to the viscosity of the same solution static repulsion is reduced. This leads to a reinforcement
when the shear rates decrease. According to Maestro of the hydrophobic associations and thus a stronger poly-
et al. [413] hydrophobically modified polysaccharides dis- mer network. Akiyama et al. [389] agree up to a point with
play thixotropic behavior due to the diffusion controlled the model presented in Fig. 39. However, an expansion
migration of hydrophobes to micelles. is introduced in that the salt effect is also dependent on
the hydrophobic/hydrophilic ratio. When the hydropho-
bic microdomains of the polymer network are saturated,
3.4.2. Effect of inherent parameters
i.e. when there are no free hydrophobes for association,
The presence of hydrophobic groups leads to an increase
the effect (reinforcement of the hydrophobic association)
in the solution viscosity of HM-polysaccharides. Similar
depicted in the model will not occur. The addition of salt
to other hydrophobically modified polymers associations
will lead to a reduction of the electrostatic repulsion but
between the hydrophobic groups arise which lead to a
the hydrophobic associations will not be reinforced given
strong polymer network. The behavior, significant increase
their strength prior to the addition of the salt. On the other
in the zero-shear viscosity, was noticed for solutions
hand if the hydrophobic microdomains are not saturated,
employing either polymer 58 [392–394] or polymer 64
i.e. when there are free hydrophobes, than the model pre-
[405].
sented in Fig. 39 can be held valid.
As mentioned before, the incorporation of charged
The solution hysteresis (as a function of the shear rate)
moieties affects the network structure of HHM-
is affected by the presence of salt. Increasing the salt con-
polysaccharides. Therefore the rheological properties
centration will suppress the tendency of the solution to
are also affected. In principle the thickening capability
display hysteresis, and increasing the salt concentration
of the charged polysaccharide is higher compared to the
beyond a certain limit no hysteresis behavior is observed.
uncharged analogue [386]. The HHM-HEC polymer 55 is an
An example of such a solution is the solution with polymer
example of such a charged polysaccharide where sulfonic
55 (R-HEC) [387].
groups are incorporated into the backbone.
The temperature affects the viscosity of hydrophobi-
cally modified polysaccharides. In most cases the viscosity
3.4.3. Effect of external parameters is reduced as the temperature is increased. This behav-
In principle the presence of salt does not affect the ior has been observed with polymer 55 and 62 (R = 1)
rheological properties of HM-polysaccharides solutions. [402]. In addition, the aggregation of hydrophobes is also
However, it has been demonstrated that the presence of affected by temperature. By using polymer 61, Winnik
salt leads to a stronger hydrophobic network and thus an [414] demonstrated that upon increasing the temperature
increase in the solution viscosity. Polymer 57 is an example the hydrophobic associations are less stable, which leads
of this behavior [391]. When charged moieties are incorpo- to a weaker polymer network. The dependence of the vis-
rated into the backbone of the HM-polysaccharides, screen- cosity on temperature was also observed with polymer 56
ing of the electrostatic repulsion arises in the presence of [415]. The shear thickening and shear thinning remained
electrolyte. In classic polyelectrolytes, this screening leads the same but the viscosity values changed as a function of
to a collapse of the polymer coils and a significant reduction temperature.
in the solution viscosity. However, for the polysaccha- Similar to HMPAM, HEUR and HASE polymers
rides modified with hydrophobic groups (in addition to hydrophobically modified polysaccharides also dis-
the charged groups), this behavior is not observed. Indeed, play interactions with different types surfactants, i.e.
the presence of salt will weaken the electrostatic repulsion. anionic, cations or nonionic. The surfactants that have
However this leads to the enhancement of the hydropho- been demonstrated to interact with hydrophobically
bic associations and thus a strengthening of the polymer modified polysaccharides are listed in Table 5.
network. In addition to the presence of salt, the type of The interaction between HM-HEC and a surfactant leads
salt is also important for the observed behavior. According to different behaviors depending on the concentration of
to Picton and Muller [412], the presence of salts that are the surfactant and the type of surfactant. When using the
able to induce a structuring of the water molecules, will anionic surfactant SDS, the surfactant molecule will inter-
lead to an enhancement of the viscosity due to the increase act with the hydrophobic groups forming mixed micelles.
number of hydrophobic association. However using other The formation of these mixed micelles will increase
types of salts, e.g. potassium thiocyanate, the enhancement the number of hydrophobic microdomains and will
1618 D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628

Fig. 39. Schematic model of the effects of added salts.


Reproduced with permission from [387] 2006, ACS Publishing.

enhance interpolymer associations and thus strengthen The surfactant binding mechanism to the hydrophobes
the polymer network [307,415,420,423]. Increasing the of HM-EHEC depends on the surfactant concentration. It
surfactant concentration further will lead to electrostatic was demonstrated that at low surfactant concentration
repulsions between the mixed micelles and subse- the hydrophobic microdomains act as nucleation sites
quently a drop in the solution viscosity of the ternary onto which the surfactants preferentially absorb [426,428].
(water–polymer–surfactant) system [423]. The viscosity An increase in the surfactant concentration leads to a
of the polymeric solution employing polymer 55, 57, transition from non-cooperative binding to a more cooper-
62 (R=1) or 65 increases and passes through a maxi- ative binding [417,426] similar to the unmodified analogue
mum as the concentration of the anionic SDS surfactant (EHEC) in polymer–surfactant solution [426]. In addi-
is increased [408,416,432]. The reduction in solution tion, the data indicate that the number of hydrophobic
viscosity at elevated surfactant concentration can be microdomains is constant with respect to the surfac-
counteracted by increasing the surfactant aggregation tant concentration. According to Evertsson and Nilsson
number in the mixed micelles. This can be achieved by [417], the strongest interaction, in terms of the highest
either adding screening electrolytes or oppositely charged microviscosity, is achieved when the surfactant binds in
surfactants [433]. A polymer–surfactant solution employ- non-cooperative way at very low adsorption degree of SDS.
ing a surfactant with a lower CMC will reach a maximum The polymer concentration also has an effect on the type
solution viscosity at lower surfactant concentration com- of interaction that is observed between surfactants and
pared to a surfactant with a higher CMC [430]. HM-EHEC. With the addition of surfactant (SDS) to the
The zero-shear viscosity of a polymer solution passes polymeric solution employing polymer 62 (R = 1), Lauten
through a maximum with the addition of a nonionic and Nyström [419] demonstrated that the association com-
surfactant (Cm EOn ), similar to the behavior with an plexes between the polymer and surfactant will grow with
anionic surfactant [415]. Increasing the temperature of time. These complexes will eventually sedimentate due
the polymer–surfactant solution leads at first to a reduc- to their size. This effect is observed at low SDS concen-
tion in the solution viscosity. However, beyond this region tration at both high and low polymer concentration. At
a considerable increase in the viscosity is observed until high SDS concentration the intermolecular associations are
the cloud point temperature of the solution is reached. dissolved and no time dependence is observed for the
This increase in solution viscosity is attributed to the system [419]. By using cloud point experiments, Thures-
interchain association caused by micellar aggregation son et al. [425] demonstrated that the solubility of the
[415].

Table 5
The different surfactants shown to interact with hydrophobically modified polysaccharides.

Surfactant Abbreviation Type Source

Sodium dodecyl sulfate SDS (NaC12 S) Anionic [117,307,408,416–428]


Sodium decyl sulfate NaC10 S Anionic [117]
Sodium tetradecyl sulfate NaC14 S Anionic [424]
Sodium dodecyl-di(ethyleneoxide)-sulfate (NaC12 [EO]2 S) Anionic [424]
Sodium alkanoate CH3 [CH2 ]n COONa, n = 4, 6, 8 and 10 Anionic [427]
Sodium octyl benzene sulfonate SOBS Anionic [427,429]
Cetyltrimethylammonium bromide CTAB Cationic [430]
Tetradecyltrimethyl ammonium bromide TTAB Cationic [420,430]
Dodecylammonium chloride C12 ACl Cationic [424]
Dodecyltrimethylammonium chloride C12 TACl Cationic [424]
Dodecyltrimethylammonium bromide C12 TABr Cationic [424]
Hexadecyltrimethylammonium bromide C16 TABr Cationic [424]
Hexadecyltrimethylammonium acetate C16 TAAc Cationic [424]
Alkyl betainate CH3 [CH2 ]n COOCH2 N[CH3 ]3 , n = 9, 11, 13 Cationic [431]
Octylphenol adduct C8 H17 C6 H4 O[EO]10 H (Triton X-100) Nonionic [418]
Polyoxyethylene alkyl-ether Cm Hn , m = 11, 12 and 13/n = 4 Nonionic [117,415,420]
D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628 1619

hydrophobically modified polymer is suppressed at low chemical/molecular structure. Rheological properties will
SDS concentrations compared to its unmodified analogue. be affected by a combination of external parameters and
HM-EHEC also interacts with different surfactant in a the chemical nature and molecular structure of the poly-
different manner. The addition of ionic surfactants leads in mer. For instance, the rheological properties of an aqueous
most cases to an enhancement of the hydrophobic asso- solution of an amphiphilic polyelectrolyte are similar to
ciations at moderate surfactant concentration. Increasing those of an unmodified analogue without amphiphilic
the surfactant concentration further leads to the disrup- moieties. However, in the presence of salt a markedly dif-
tion of the hydrophobic micro-domains and thus of the ferent behavior is observed. The solution viscosity of the
polymer network. According to Kjonikson et al. [434] the unmodified polymer decreases with increasing salt con-
polymer–surfactant system transits from a heterogeneous centration, whereas the solution viscosity of the aqueous
system at low surfactant concentration to a homogeneous solution containing the amphiphilic polyelectrolyte is not
one at high surfactant concentration. The interactions of affected. Another good example is the thermal perfor-
nonionic surfactants with HM-EHEC, i.e. polymer 62 (R = 1), mance of amphiphilic polymers: the rheological properties
are weaker when compared to ionic surfactants [427]. Nev- of an aqueous solution containing the amphiphilic poly-
ertheless, for both ionic and nonionic surfactants the length mer or its unmodified analogue, i.e. without the NIPAM
of the surfactant chain is crucial for the solution behavior. monomer, are quite similar. However, when exposing both
If this length is increased the disruption of the network, i.e. solutions to higher temperatures significant differences
loss of solution viscosity, is postponed due to an increase arise. The effect of temperature on the solution viscosity of
in the aggregation number of the mixed micelles [427]. an aqueous solution containing the amphiphilic polymer is
One striking behavior is the influence of a spacer (EO- limited, whereas the viscosity of the unmodified analogue
spacer) between the hydrophobic group and the polymer changes significantly.
backbone on the solution viscosity [434] in the presence of Although there are many different water-soluble
surfactants. At high surfactant concentrations, the behav- polymers capable of enhancing the solution viscosity,
ior of the polymer with or without the spacer is similar to it is important to understand their differences and
its unmodified non-hydrophobic analogue. The viscosity of analogies. Different polymers display in general differ-
the polymeric solution employing the HM-EHEC with spac- ences in the agglomeration principles governing their
ers of 4 EO units with no surfactant is much lower compared water behavior. On a molecular level the basic princi-
to the one without spacers. However, the authors [434] ple is indeed quite general: the presence of relatively
correctly mention the positive effect a spacer should have weak inter(macro)molecular interactions (e.g. hydropho-
on the viscosity given the studies of Hwang and Hogen- bic association and hydrogen bonding) factually increases
Esch [241] and Tam et al. [380]. Therefore the authors [434] the molecular weight of the polymer coils. As a con-
attribute the discrepancy to the difference in distribution sequence the solution viscosity increases. However, a
of the hydrophobes along the polymer backbone. careful balance must be observed here since predomi-
The presence of salt does not affect the rheological prop- nantly weak interactions (both in terms of strength and
erties of a solution containing a HM-polysaccharide and a number thereof) do not result in observable rheological dif-
surfactant. At intermediate salt concentration no change ferences while excessively strong ones might compromise
in the solution viscosity is observed. However a further the solubility of the system by leading for example to gel
increase in the salt concentration leads to a solution vis- formation.
cosity which is higher than for a solution without the
surfactant.
Acknowledgement
4. Conclusion
This work is part of the Research Programme of the
This review clearly indicates that the successful design Dutch Polymer Institute DPI, Eindhoven, the Netherlands.
of new water-soluble polymers for a given applica-
tion requires an integral multiscale and multidisciplinary References
approach. Proper definition of the required product proper-
ties is in this case crucial. Knowledge of polymer chemical [1] Thomas S. Enhanced oil recovery—an overview. Oil Gas Sci Technol
architecture (and thus of the synthetic methods used) must 2008;63:9–19.
[2] Han D-K, Yang C-Z, Zhang Z-Q, Lou Z-H, Chang Y-I. Recent
be conceptually linked to the desired product application development of enhanced oil recovery in China. J Pet Sci Eng
requirements. In this case viscosity measurements under 1999;22:181–8.
different shear conditions are of paramount importance [3] Li G, Zhai L, Xu G, Shen Q, Mao H. Current tertiary oil recovery in
China. J Disper Sci Technol 2000;21:367–408.
and should be ideally correlated with the “nature” (i.e. poly-
[4] Lake LW. Enhanced oil recovery. Englewood Cliffs, NJ: Prentice-Hall
mer architecture and overall chemical composition) of the Inc.; 1989.
corresponding water solution. [5] Sorbie KS. Polymer-improved oil recovery. Boca Raton, FL: CRC
Press; 1991.
The influence of external parameters (e.g. pH and tem-
[6] Levitt DB, Pope GA. Selection and screening of polymers for
perature) on the rheological behavior must be coupled enhanced-oil recovery. In: SPE Improved Oil Recovery Symp. 2008.
with an in depth knowledge of the relationship between p. 1–18, 113845.
the chemical structure and architecture of the polymer [7] Gaillard N, Giovannetti B, Favero C. Improved oil recovery using
thermally and chemically protected compositions based on co- and
and the rheological behavior. In this respect, an overall ter-polymers containing acrylamide. In: SPE Improved Oil Recovery
correlation cannot be defined only as a function of the Symp. 2010. p. 1–11, 129756.
1620 D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628

[8] Wu Y, Wang K-S, Hu Z, Bai B, Shuler P, Tang Y. A new method for fast [35] Sukpisan J, Kanatharana J, Sirivat A, Wang Y. The specific viscosity
screening of long-term thermal stability of water soluble polymers of partially hydrolyzed polyacrylamide solutions: effects of degree
for reservoir conformance control. In: SPE Improved Oil Recovery of hydrolysis, molecular weight, solvent quality and temperature.
Symp. 2009. p. 1–11, 124257. J Polym Sci Part B Polym Phys 1998;36:743–53.
[9] Pancharoen M, Thiele MR, Kovscek AR. Inaccessible pore volume of [36] Khune GD, Donaruma LG, Hatch MJ, Kilmer NH, Shepitka JS, Martin
associative polymer floods. In: SPE Improved Oil Recovery Symp. FD. Modified acrylamide polymers for enhanced oil recovery. J Appl
2010. p. 1–15, 129910. Polym Sci 1985;30:875–85.
[10] Buchgraber M, Clemens T, Castanier LM, Kovscek AR. The displace- [37] McCormick CL, Neidlinger HH, Hester RD, Wildman GC. Syn-
ment of viscous oil by associative polymer solutions. In: SPE Ann thetic random and graft copolymers for utilization in enhanced oil
Tech Confr. 2009. p. 1–19, 122400. recovery—synthesis and rheology. In: Shah DO, editor. Surface phe-
[11] Zhang L-J, Yue X-A. Displacement of polymer solution on residual nomena in enhanced oil recovery. 1st ed. New York: Plenum; 1981.
oil trapped in dead ends. J Cent South Univ Tech 2008;15:84–7. p. 741–72.
[12] Zhang L-J, Yue X-A, Guo F. Micro-mechanisms of residual oil mobi- [38] Dautzenberg H. Polyelectrolyte complex formation in highly aggre-
lization by viscoelastic fluids. Pet Sci 2008;5:56–61. gating systems. 1. Effect of salt: polyelectrolyte complex formation
[13] Wang D, Xia H, Liu Z, Anda Q, Yang Q. Study of the mechanism in the presence of NaCl. Macromolecules 1997;30:7810–5.
of polymer solution with visco-elastic behavior increasing micro- [39] Peng S, Wu C. Light scattering study of the formation and structure
scopic oil displacement efficiency and the forming of steady “oil of partially hydrolyzed poly(acrylamide)/calcium(II) complexes.
thread” flow channels. In: SPE Asia Pacific Oil Gas Confr. 2001. p. Macromolecules 1999;32:585–9.
1–9, 68723. [40] Ohmine I, Tanaka T. Salt effects on the phase transition of ionic gels.
[14] Yin H, Wang D, Zhong H. Study on flow behaviors of viscoelastic J Chem Phys 1982;77:5725–9.
polymer solution in micropore with dead end. In: SPE Ann Tech [41] Ben Jar P-Y, Wu YS. Effect of counter-ions on swelling and shrinkage
Confr. 2006. p. 1–10, 101950. of polyacrylamide-based ionic gels. Polymer 1997;38:2557–60.
[15] Wang D, Cheng J, Yang Q, Gong W, Li Q, Chen F. Viscous-elastic [42] Cook Jr RL, King HE, Peiffer DG. High-pressure viscosity of
polymer can increase microscale displacement efficiency in cores. dilute polymer solutions in good solvents. Macromolecules
In: SPE Ann Tech Confr. 2000. p. 1–10, 63227. 1992;25:2928–34.
[16] Xia H, Ju Y, Kong F, Wu J. Effect of elastic behavior of HPAM solutions [43] Martin FD, Hatch MJ, Shepitka JS, Ward JS. Improved water-soluble
on displacement efficiency under mixed wettability conditions. In: polymers for enhanced recovery of oil. In: SPE Ann Tech Confr. 1983.
SPE Ann Tech Confr. 2004. p. 1–8, 90234. p. 1–14, 11786.
[17] Glass JE, editor. Hydrophilic polymers. Performance with environ- [44] Martin FD. Mechanical degradation of polyacrylamide solutions in
mental acceptance. Advances in chemistry series 248. Washington, core plugs from several carbonate reservoirs. SPE Formation Eval
DC: American Chemical Society; 1996. 1986;1:139–50.
[18] Schulz DN, Glass JE, editors. Polymers as rheology modifiers. ACS [45] Lewandowska K. Comparative studies of rheological properties of
symposium series no. 462. Washington, DC: American Chemical polyacrylamide and partially hydrolyzed polyacrylamide solutions.
Society; 1991. J Appl Polym Sci 2007;103:2235–41.
[19] Shalaby SW, McCormick CL, Butler GB, editors. Water soluble [46] Hu Y, Wang SQ, Jamieson AM. Rheological and rheooptical stud-
polymers. Synthesis, solution properties and applications. ACS ies of shear-thickening polyacrylamide solutions. Macromolecules
symposium no. 467. Washington, DC: American Chemical Society; 1995;28:1847–53.
1991. [47] Seright RS, Seheult M, Talashek T. Injectivity characteristics of EOR
[20] Glass JE, editor. Polymers in aqueous media: performance through polymers. SPE Reserv Eval Eng 2009;12:783–92.
association. Advances in chemistry series 223. Washington, DC: [48] Chauveteau G. Molecular interpretation of several different prop-
American Chemical Society; 1989. erties of flow of coiled polymer solutions through porous media
[21] Morgan SE, McCormick CL. Water-soluble polymers in enhanced in oil recovery conditions. In: SPE Ann Tech Confr. 1981. p. 1–14,
oil recovery. Prog Polym Sci 1990;15:103–45. 10060.
[22] Sabhapondit A, Borthakur A, Haque I. Characterization of acry- [49] Choplin L, Sabatié J. Threshold-type shear-thickening in polymeric
lamide polymers for enhanced oil recovery. J Appl Polym Sci solutions. Rheol Acta 1986;25:570–9.
2003;87:1869–78. [50] Ferguson J, Walters K, Wolff C. Shear and extensional flow of poly-
[23] Sabhapondit A, Borthakur A, Haque I. Water soluble acry- acrylamide solutions. Rheol Acta 1990;29:571–9.
lamidomethyl propane sulfonate (AMPS) copolymer as an [51] Shepitka JS, Case CE, Donaruma LG, Hatch MJ, Kilmer NH, Khune GD,
enhanced oil recovery chemical. Energy Fuels 2003;17:683–8. Martin FD, Ward JS, Wilson KV. Partially imidized, water-soluble
[24] Song H, Zhang S-F, Ma X-C, Wang D-Z, Yang J-Z. Synthesis and polymeric amides. 1. Partially imidized polyacrylamide and poly-
application of starch-graft-poly(AM-co-AMPS) by using a complex methacrylamide. J Appl Polym Sci 1983;28:3611–7.
initiation system of CS-APS. Carbohydr Polym 2007;69:189–95. [52] Durst F, Haas R, Interthal W. The nature of flows through porous
[25] Vega I, Sánchez L, D’Accorso N. Synthesis and characterization of media. J Non-Newtonian Fluid Mech 1987;22:169–89.
copolymers with 1,3-oxazolic pendant groups. React Funct Polym [53] James D, MacLaren DR. Laminar-flow of dilute polymer-solutions
2008;68:233–41. through porous media. J Fluid Mech 1975;70:733–52.
[26] Borthakur A, Rahman M, Sarmah A, Subrahmanyam B. Partially [54] Magueur A, Moan M, Chauveteau G. Effect of successive contrac-
hydrolysed polyacrylamide for enhanced oil recovery. Res Ind tions and expansions on the apparent viscosity of dilute polymer
1995;40:90–4. solutions. Chem Eng Commun 1985;36:351–66.
[27] Stokes RJ, Evans DF. Fundamentals of interfacial engineering. New [55] Silberberg A, Mijnlieff PF. Study of reversible gelation of partially
York: Wiley-VCH; 1997. neutralized poly(methacrylic acid) by viscoelastic measurements.
[28] Shupe RD. Chemical stability of polyacrylamide polymers. J Pet J Polym Sci Part A2 Polym Phys 1970;8:1089.
Technol 1981;33:1513–29. [56] Quadrat O. Negative thixotropy in polymer solutions. Adv Colloid
[29] Fuoss RM, Strauss UP. Polyelectrolytes. 2. Poly-4-vinylpyridonium Interface Sci 1985;24:45–75.
chloride and poly-4-vinyl-N-N-butylpyridonium bromide. J Polym [57] Ohya S, Matsuo T. Time-dependent change of viscosity in poly-
Sci 1948;3:246–63. methacrylic acid solution. J Colloid Interface Sci 1979;68:593–5.
[30] Fuoss RM, Strauss UP. Electrostatic interaction of polyelectrolytes [58] Bradna P, Quadrat O, Dupuis D. The influence of salt concentra-
and simple electrolytes. J Polym Sci 1948;3:602–3. tion on negative thixotropy in solutions of partially hydrolyzed
[31] Fuoss RM. Viscosity function for polyelectrolytes. J Polym Sci polyacrylamide. Colloid Polym Sci 1995;273:421–5.
1948;3:603–4. [59] Quadrat O, Bradna P, Dupuis D, Wolff C. Negative thixotropy of solu-
[32] Ait-Kadi A, Carreau PJ, Chauveteau G. Rheological proper- tions of partially hydrolyzed polyacrylamide. Part I. The influence
ties of partially hydrolyzed polyacrylamide solutions. J Rheol of shear rate on time changes of flow characteristics. Colloid Polym
1987;31:537–61. Sci 1992;270:1057–9.
[33] Dupuis D, Lewandowski FY, Steiert P, Wolff C. Shear thickening and [60] Bradna P, Quadrat O, Dupuis D. Negative thixotropy of solu-
time-dependent phenomena: the case of polyacrylamide solutions. tions of partially hydrolyzed polyacrylamide. Part II. The influence
J Non-Newtonian Fluid Mech 1994;54:11–32. of glycerol content and degree of ionization. Colloid Polym Sci
[34] Ellwanger RE, Jaeger DA, Barden RE. Use of the empirical hill equa- 1995;273:642–7.
tion for characterization of the effect of added cations on the [61] Bradna P, Quadrat O, Titkova L, Dupuis D. The influence of multiva-
viscosity of aqueous solutions of partially hydrolyzed polyacry- lent cations on negative thixotropy in aqueous glycerol solutions of
lamide. Polym Bull 1980;3:369–74. partially hydrolyzed polyacrylamide. Acta Polym 1997;48:446–9.
D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628 1621

[62] Ghannam MT, Esmail MN. Rheological properties of aqueous poly- [91] English RJ, Laurer JH, Spontak RJ, Khan SA. Hydrophobically
acrylamide solutions. J Appl Polym Sci 1998;69:1587–97. modified associative polymer solutions: rheology and microstruc-
[63] Xin X, Xu G, Gong H, Bai Y, Tan Y. Interaction between sodium oleate ture in the presence of nonionic surfactants. Ind Eng Chem Res
and partially hydrolyzed polyacrylamide: a rheological study. Col- 2002;41:6425–35.
loids Surf A 2008;326:1–9. [92] Biggs S, Selb J, Candau F. Effect of surfactant on the solution
[64] Methemitis C, Morcellet M, Sabbadin J, Francois J. Interactions properties of hydrophobically modified polyacrylamide. Langmuir
between partially hydrolyzed polyacrylamide and ionic surfac- 1992;8:838–47.
tants. Eur Polym J 1986;22:619–27. [93] Sarkar N, Kershner LD. Rigid rod water-soluble polymers. J Appl
[65] Leela JK, Sharma G. Studies on xanthan production from Xan- Polym Sci 1996;62:393–408.
thomonas campestris. Bioprocess Eng 2000;23:687–9. [94] Maia AMS, Borsali R, Balaban RC. Comparison between a polyacry-
[66] Nasr S, Soudi MR, Haghighi M. Xanthan production by a native lamide and a hydrophobically modified polyacrylamide flood in a
strain of X. campestris and evaluation of application in EOR. Pak sandstone core. Mater Sci Eng C 2009;29:505–9.
J Biol Sci 2007;10:3010–3. [95] Dubin PL, Strauss UP. Hydrophobic bonding in alternating copoly-
[67] García-Ochoa F, Santos VE, Casas JA, Gómez E. Xanthan gum: pro- mers of maleic acid and alkyl vinyl ethers. J Phys Chem
duction, recovery, and properties. Biotechnol Adv 2000;18:549–79. 1970;74:2842–7.
[68] Becker A, Katzen F, Puhler A, Ielpi L. Xanthan gum biosynthesis [96] Dubin PL, Strauss UP. Hydrophobic hypercoiling in copolymers of
and application: a biochemical/genetic perspective. Appl Environ maleic acid and alkyl vinyl ethers. J Phys Chem 1967;71:2757–9.
Microbiol 1998;50:145–52. [97] Schwab FC, Sheppard EW, Chen CSH. Waterflooding process
[69] Morris ER, Rees DA, Young G, Walkinshaw MD, Darke A. employing cationic-nonionic copolymers. US Pat 4,110,232; 1978.
Order–disorder transition for a bacterial polysaccharide in [98] Landoll LM. Hydrophobically modified polymers. US Pat 4,529,523;
solution—role for polysaccharide conformation in recognition 1984.
between Xanthomonas pathogen and its plant host. J Mol Biol [99] Bock J, Pace SJ, Schulz DN. Enhanced oil recovery with hydropho-
1977;110:1–16. bically associating polymers containing N-vinyl pyrrolidone
[70] Dentini M, Crescenzi V, Blasi D. Conformational properties of xan- functionality. US Pat 4,709,759; 1987.
than derivatives in dilute aqueous-solution. Int J Biol Macromol [100] Bock J, Valint PL, Pace SJ. Enhanced oil recovery with hydropho-
1984;6:93–8. bically associating polymers containing sulfonate functionality. US
[71] Norton IT, Goodall DM, Frangou SA, Morris ER, Rees DA. Mechanism Pat 4,702,319; 1987.
and dynamics of conformational ordering in xanthan polysaccha- [101] Bock J, Siano DB, Pace SJ. Enhanced oil recovery with hydrophobi-
ride. J Mol Biol 1984;175:371–94. cally associating polymers. Can Pat 1300362; 1992.
[72] Holzwarth G. Conformation of extracellular polysaccharide of [102] Evani S. Enhanced oil recovery process using a hydrophobic associa-
Xanthomonas-campestris. Biochemistry 1976;15:4333–9. tive composition containing a hydrophilic/hydrophobic polymer.
[73] Chen CSH, Sheppard EW. Conformation and shear stability of xan- US Pat 4,814,096; 1989.
than gum in solution. Polym Eng Sci 1980;20:512–6. [103] McCormick CL, Nonaka T, Johnson CB. Water-soluble copolymers.
[74] Yoshida T, Tanner RD, editors. Bioproducts and bioprocess. Berlin: 27. Synthesis and aqueous solution behavior of associa-
Springer-Verlag; 1993. tive acrylamide/N-alkylacrylamide copolymers. Polymer
[75] Richardson RK, Rossmurphy SB. Nonlinear viscoelasticity of 1988;29:731–9.
polysaccharide solutions. 2. Xanthan polysaccharide solutions. Int [104] Lara-Ceniceros AC, Rivera-Vallejo C, Jimenez-Regalado EJ. Synthe-
J Biol Macromol 1987;9:257–64. sis, characterization and rheological properties of three different
[76] Kierulf C, Sutherland IW. Thermal-stability of xanthan prepara- associative polymers obtained by micellar polymerization. Polym
tions. Carbohydr Polym 1988;9:185–94. Bull 2007;58:425–33.
[77] Lambert F, Rinaudo M. On the thermal-stability of xanthan gum. [105] Tam KC, Jenkins RD, Winnik MA, Bassett DR. A structural model of
Polymer 1985;26:1549–53. hydrophobically modified urethane-ethoxylate (HEUR) associative
[78] Seright RS, Henrici BJ. Xanthan stability at elevated temperatures. polymers in shear flows. Macromolecules 1998;31:4149–59.
SPE Reserv Eng 1990;1:52–60. [106] Tam KC, Seng WP, Jenkins RD, Bassett DR. Rheological and
[79] Wellington SL. Bio-polymer solution viscosity stabilization— microcalorimetric studies of a model alkali-soluble associative
polymer degradation and antioxidant use. Soc Pet Eng J polymer (HASE) in nonionic surfactant solutions. J Polym Sci Part B
1983;23:901–12. Polym Phys 2000;38:2019–32.
[80] Ash SG, Clarke-Sturman AJ, Calvert R, Nisbet TM. Chemical stabil- [107] Feng Y, Billon L, Grassl B, Khoukh A, François J. Hydrophobi-
ity of biopolymer solutions. In: SPE Ann Tech Confr. 1983. p. 1–8, cally associating polyacrylamides and their partially hydrolyzed
12085. derivatives prepared by post modification. 1. Synthesis and char-
[81] Sutherland IW. An enzyme system hydrolysing the polysaccharides acterization. Polymer 2002;43:2055–64.
of Xanthomonas species. J Appl Bacteriol 1982;53:385–93. [108] Argillier J-F, Audibert A, Lecourtier J, Moan M, Rousseau L. Solution
[82] Cadmus MC, Jackson LK, Burton KA, Plattner RD, Slodki ME. and adsorption properties of hydrophobically associating water-
Biodegradation of xanthan gum by Bacillus sp. Appl Environ Micro- soluble polyacrylamides. Colloids Surf A 1996;113:247–57.
biol 1982;44:5–11. [109] Volpert E, Selb J, Candau F. Influence of the hydrophobe structure on
[83] Bragg JR, Maruca SD, Gale WW, Gall LS, Wernau WC, Beck D, composition, microstructure, and rheology in associating polyacry-
Goldman IM, Laskin AI, Naslund LA. Control of xanthan-degrading lamides prepared by micellar copolymerization. Macromolecules
organisms in the loudon pilot: approach, methodology, and results. 1996;29:1452–63.
In: SPE Ann Tech Confr. 1983. p. 1–12, 11989. [110] Ezzell SA, McCormick CL. Water-soluble copolymers. 39. Synthe-
[84] Hou CT, Barnabe N, Greaney K. Biodegradation of xanthan by salt- sis and solution properties of associative acrylamido copolymers
tolerant aerobic microorganisms. J Ind Microbiol 1986;1:31–7. with pyrenesulfonamide fluorescence labels. Macromolecules
[85] Taugbol K, Ly TV, Austad T. Chemical flooding of oil-reservoirs. 3. 1992;25:1881–6.
Dissociative surfactant–polymer interaction with a positive effect [111] Ezzell SA, Hoyle CE, Creed D, McCormick CL. Water-soluble
on oil-recovery. Colloids Surf A 1995;103:83–90. copolymers. 40. Photophysical studies of the solution behavior
[86] Austad T, Taugbol K. Chemical flooding of oil-reservoirs. 2. Disso- of associative pyrenesulfonamide-labeled polyacrylamides. Macro-
ciative surfactant–polymer interaction with a negative effect on molecules 1992;25:1887–95.
oil-recovery. Colloids Surf A 1995;103:73–81. [112] Klucker R, Candau F, Schosseler F. Transient behavior of associating
[87] Taylor KC, Nasr-El-Din HA. Water-soluble hydrophobically associ- copolymers in shear flow. Macromolecules 1995;28:6416–22.
ating polymers for improved oil recovery: a literature review. J Pet [113] Dowling KC, Thomas JK. A novel micellar synthesis and photophys-
Sci Eng 1998;19:265–80. ical characterization of water-soluble acrylamide–styrene block
[88] Hill A, Candau F, Selb J. Properties of hydrophobically associating copolymers. Macromolecules 1990;23:1059–64.
polyacrylamides: influence of the method of synthesis. Macro- [114] Schulz DN, Kaladas JJ, Maurer JJ, Bock J, Pace SJ, Schulz WW. Copoly-
molecules 1993;26:4521–32. mers of acrylamide and surfactant macromonomers: synthesis and
[89] Annable T, Buscall R, Ettelaie R, Whittlestone D. The rheology solution properties. Polymer 1987;28:2110–5.
of solutions of associating polymer: comparison of experimental [115] Peiffer DG. Hydrophobically associating polymers and their inter-
behavior with transient network theory. J Rheol 1993;37:695–726. actions with rod-like micelles. Polymer 1990;31:2353–60.
[90] Panmai S, Prud’homme RK, Peiffer DG. Rheology of hydrophobically [116] Iliopoulos I, Wang TK, Audebert R. Viscometric evidence of inter-
modified polymers with spherical and rod-like surfactant micelles. actions between hydrophobically modified poly(sodium acrylate)
Colloids Surf A 1999;147:3–15. and sodium dodecyl sulfate. Langmuir 1991;7:617–9.
1622 D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628

[117] Senan C, Meadows J, Shone PT, Williams PA. Solution behav- mer solutions: thermothinning versus thermothickening. Macro-
ior of hydrophobically modified sodium polyacrylate. Langmuir molecules 2005;38:8512–21.
1994;10:2471–9. [141] Mortensen K, Brown W, JØrgensen E. Phase behavior
[118] Tanaka R, Meadows J, Williams PA, Phillips GO. Interaction of poly(propylene oxide)–poly(ethylene oxide)–poly(propylene
hydrophobically modified (hydroxyethyl)cellulose with various oxide) triblock copolymer melt and aqueous solutions.
added surfactants. Macromolecules 1992;25:1304–10. Macromolecules 1994;27:5654–66.
[119] Winnik FM. Association of hydrophobic polymers in water: [142] Nishinari K, Hofmann KE, Moritaka H, Kohyama K, Nishinari
fluorescence studies with labeled (hydroxypropyl)celluloses. N. Gel–sol transition of methylcellulose. Macromol Chem Phys
Macromolecules 1989;22:734–42. 1997;198:1217–26.
[120] Durand A, Hourdet D. Synthesis and thermoassociative properties [143] Doelker E. Cellulose derivatives. Adv Polym Sci 1993;107:199–265.
in aqueous solution of graft copolymers containing poly(N- [144] Sarkar N. Thermal gelation properties of methyl and hydroxypropyl
isopropylacrylamide) side chains. Polymer 1999;40:4941–51. methylcellulose. J Appl Polym Sci 1979;24:1073–87.
[121] Maia AMS, Costa M, Borsali R, Garcia RB. Rheological behavior [145] Yoshioka H, Mikami M, Mori Y, Tsuchida E. A synthetic hydro-
and scattering studies of acrylamide-based copolymer solutions. gel with thermoreversible gelation. 1. Preparation and rheological
Macromol Symp 2005;229:217–27. properties. J Macromol Sci Part A Pure Appl Chem 1994;31:
[122] Lundberg DJ, Brown RG, Glass JE, Eley RR. Synthesis, characteri- 113–20.
zation, and solution rheology of model hydrophobically-modified, [146] Nyström B, Walderhaug H, Hansen FK. Dynamic light scat-
water-soluble ethoxylated urethanes. Langmuir 1994;10:3027–34. tering and rheological studies of thermoreversible gelation
[123] Winnik MA, Yekta A. Associative polymers in aqueous solution. Curr of a poly(ethylene oxide)–poly(propylene oxide)–poly(ethylene
Opin Colloid Interface Sci 1997;2:424–36. oxide) triblock copolymer in aqueous solution. Faraday Discuss
[124] Huldén M. Hydrophobically modified urethane-ethoxylate (HEUR) 1995;101:335–44.
associative thickeners. 1. Rheology of aqueous solutions and inter- [147] Wang G, Lindell K, Olofsson G. On the thermal gelling of
actions with surfactants. Colloids Surf A 1994;82:263–77. ethyl(hydroxyethyl)cellulose and sodium dodecyl sulfate. Phase
[125] Xu B, Li L, Zhang K, MacDonald PM, Winnik MA, Jenkins R, Bas- behavior and temperature scanning calorimetric response. Macro-
set D, Wolf D, Nuyken O. Synthesis and characterization of comb molecules 1997;30:105–12.
associative polymers based on poly(ethylene oxide). Langmuir [148] Loyen K, Iliopoulos I, Audebert R, Olsson U. Reversible thermal
1997;13:6896–902. gelation in polymer/surfactant systems. Control of the gelation
[126] Rao B, Uemura Y, Dyke L, Mcdonald PM. Self-diffusion coefficients temperature. Langmuir 1995;11:1053–6.
of hydrophobic ethoxylated urethane associating polymers using [149] Sarrazin-Cartalas A, Iliopoulos I, Audebert R, Olsson U. Associa-
pulsed-gradient spin-echo nuclear magnetic resonance. Macro- tion and thermal gelation in mixtures of hydrophobically modified
molecules 1995;28:531–8. polyelectrolytes and nonionic surfactants. Langmuir 1994;10:
[127] Alami E, Almgren M, Brown W. Interaction of hydrophobically 1421–6.
end-capped poly(ethylene oxide) with nonionic surfactants in [150] Greenhill-Hooper MJ, O’Sullivan TP, Wheele PA. The aggrega-
aqueous solution. Fluorescence and light scattering studies. Macro- tion behavior of octadecylphenylalkoxysulfonates. 1. Temperature
molecules 1996;29:5026–35. dependence of the solution behavior. J Colloid Interface Sci
[128] Alami E, Almgren M, Brown W. Aggregation of hydropho- 1988;124:77–87.
bically end-capped poly(ethylene oxide) in aqueous solu- [151] Bokias G, Hourdet D, Iliopoulos I, Staikos G, Audebert R.
tions. Fluorescence and light-scattering studies. Macromolecules Hydrophobic interactions of poly(N-isopropylacrylamide) with
1996;29:2229–43. hydrophobically modified poly(sodium acrylate) in aqueous solu-
[129] Yekta A, Duhamel J, Brochard P, Adiwidjaja H, Winnik MA. A tion. Macromolecules 1997;30:8293–7.
fluorescent probe study of micelle-like cluster formation in aque- [152] Hourdet D, L’Alloret F, Audebert R. Synthesis of thermoassociative
ous solutions of hydrophobically modified poly(ethylene oxide). copolymers. Polymer 1997;38:2535–47.
Macromolecules 1993;26:1829–36. [153] Hourdet D, L’Alloret F, Audebert R. Reversible thermothickening of
[130] Maechling-Strasser C, Clouet F, Francois J. Hydrophobically aqueous polymer solutions. Polymer 1994;35:2624–30.
end-capped polyethylene-oxide urethanes: 2. Modelling their [154] de Vos S, Möller M, Visscher K, Mijnlieff PF. Synthesis and
association in water. Polymer 1992;33:1021–5. characterization of poly(acrylamide)-graft-poly(ethylene oxide-
[131] Maechling-Strasser C, Francois J, Clouet F. Hydrophobically end- co-propylene oxide). Polymer 1994;35:2644–50.
capped poly(ethylene oxide) urethanes. 1. Characterization and [155] Hutchinson BH, McCormick CL. Water-soluble copolymers. 15.
experimental study of their association in aqueous solution. Poly- Studies of random copolymers of acrylamide with N-substituted
mer 1992;33:627–36. acrylamides by C-13 NMR. Polymer 1986;27:623–6.
[132] Xie X, Hogen-Esch TE. Copolymers of N,N-dimethylacrylamide [156] Magny B, Lafuma F, Iliopoulos I. Determination of microstructure
and 2-(N-ethylperfluorooctanesulfonamido)ethyl acrylate in aque- of hydrophobically modified water-soluble polymers by C-13 NMR.
ous media and in bulk. Synthesis and properties. Macromolecules Polymer 1992;33:3151–4.
1996;29:1734–45. [157] Newman JK, McCormick CL. Water-soluble copolymers. 51.
[133] Abu-Sharkh BF, Yahaya GO, Ali SA, Kazi IW. Solution and interfacial Copolymer compositions of high-molecular-weight functional
behavior of hydrophobically modified water-soluble block copoly- acrylamido water-soluble polymers using direct-polarization
mers of acrylamide and N-phenethylacrylamide. J Appl Polym Sci magic-angle-spinning C-13 nuclear-magnetic-resonance. Polymer
2001;82:467–76. 1994;35:935–8.
[134] Volpert E, Selb J, Candau F. Associating behaviour of poly- [158] McCormick CL, Chen GS, Hutchinson BH. Water-soluble copoly-
acrylamides hydrophobically modified with dehexylacrylamide. mers. 5. Compositional determination of random copolymers
Polymer 1998;39:1025–33. of acrylamide with sulfonated co-monomers by infrared-
[135] Candau F, Selb J. Hydrophobically-modified polyacrylamides pre- spectroscopy and C-13 nuclear magnetic-resonance. J Appl Polym
pared by micellar polymerization. Adv Colloid Interface Sci Sci 1982;27:3103–20.
1999;79:149–72. [159] Newman JK, McCormick CL. Water-soluble copolymers. 53. Na-23
[136] Yahya GO, Ali SA, Alnaafa MA, Hamad EZ. Preparation and viscosity NMR-studies of hydrophobically-modified polyacids—copolymers
behavior of hydrophobically-modified poly(vinyl alcohol) (PVA). J of 2-(1-naphthylacetamido)ethylacrylamide with acrylic-acid and
Appl Polym Sci 1995;57:343–52. methacrylic-acid. Macromolecules 1994;27:5123–8.
[137] Yahya GO, Hamad EZ. Solution behaviour of sodium maleate/1- [160] Newman JK, McCormick CL. Water-soluble copolymers. 52. Na-
alkene copolymers. Polymer 1995;36:3705–10. 23 NMR-studies of ion-binding to anionic polyelectrolytes—poly-
[138] Kopperud HM, Hansen FH, Nyström B. Effect of surfactant and (sodium 2-acrylamido-2-methylpropanesulfonate), poly(sodium
temperature on the rheological properties of aqueous solutions of 3-acrylamido-3-methylbutanoate), poly(sodium acrylate), and
unmodified and hydrophobically modified polyacrylamide. Macro- poly(sodium galacturonate). Macromolecules 1994;27:5114–22.
mol Chem Phys 1998;199:2385–94. [161] Furo I, Iliopoulos I, Stilbs P. Structure and dynamics of associative
[139] Shaikh S, Asrof Ali SK, Hamad EZ, Abu-Sharkh BF. Synthesis water-soluble polymer aggregates as seen by F-19 NMR spec-
and solution properties of poly(acrylamide-styrene) block copoly- troscopy. J Phys Chem B 2000;104:485–94.
mers with high hydrophobic content. Polym Eng Sci 1999;39: [162] Walderhaug H, Hansen FK, Abrahmsen S, Persson K, Stilbs P. Asso-
1962–8. ciative thickeners—NMR self-diffusion and rheology studies of
[140] Hourdet D, Gadgil J, Podhajecka K, Badiger MV, Brûlet A, aqueous-solutions of hydrophobically-modified poly(oxyethylene)
Wadgaonkar PP. Thermoreversible behavior of associating poly- polymers. J Phys Chem 1993;97:8336–42.
D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628 1623

[163] Nystrom B, Walderhaug H, Hansen FK. Dynamic crossover effects [184] Hu YX, Smith GL, Richardson MF, McCormick CL. Water sol-
observed in solutions of a hydrophobically associating water- uble polymers. 74. pH responsive microdomains in labeled
soluble polymer. J Phys Chem 1993;97:7743–52. n-octylamide-substituted poly(sodium maleate-alt-ethyl
[164] Abrahmsén-Alami S, Stilbs P. 1 H NMR self-diffusion and multi- vinyl ethers): synthesis, steady-state fluorescence, and non-
field 2H spin relaxation study of model associative polymer and radiative energy transfer studies. Macromolecules 1997;30:
sodium dodecyl sulfate aggregation in aqueous solution. J Phys 3526–37.
Chem 1994;98:6359–67. [185] Smith GL, McCormick CL. Water-soluble polymers. 78. Viscosity and
[165] Persson K, Abrahmsen S, Stilbs P, Hansen FK, Walderhaug H. The NRET fluorescence studies of pH-responsive twin-tailed associative
association of urethane-polyethyleneoxide (Heur) thickeners. As terpolymers based on acrylic acid and methacrylamide. Macro-
studied by NMR self-diffusion measurements. Colloid Polym Sci molecules 2001;34:918–24.
1992;270:465–9. [186] Cathébras N, Collet A, Viguier M, Berret J-F. Synthesis and linear vis-
[166] McCormick CL, Elliott DL. Water-soluble copolymers. 14. Potentio- coelasticity of fluorinated hydrophobically modified ethoxylated
metric and turbidimetric studies of water-soluble copolymers of urethanes (F-HEUR). Macromolecules 1998;31:1305–11.
acrylamide—comparison of carboxylated and sulfonated copoly- [187] Kaczmarski JP, Glass JE. Synthesis and solution properties
mers. Macromolecules 1986;19:542–7. of hydrophobically-modified ethoxylated urethanes with vari-
[167] Wang C, Tam KC, Jenkins RD, Bassett DR. Potentiometric titration able oxyethylene spacer lengths. Macromolecules 1993;26:
and dynamic light scattering of hydrophobically modified alkali sol- 5149–56.
uble emulsion (HASE) polymer solutions. Phys Chem Chem Phys [188] May R, Kaczmarski JP, Glass JE. Influence of molecular weight dis-
2000;2:1967–72. tributions on HEUR aqueous solution rheology. Macromolecules
[168] Wang C, Tam KC, Jenkins RD. Dissolution behavior of HASE poly- 1996;29:4745–53.
mers in the presence of salt: potentiometric titration, isothermal [189] Zhang HS, Pan J, Hogen-Esch TE. Synthesis and characteriza-
titration calorimetry, and light scattering studies. J Phys Chem B tion of one-ended perfluorocarbon-functionalized derivatives of
2002;106:1195–204. poly(ethylene glycol)s. Macromolecules 1998;31:2815–21.
[169] Dai S, Tam KC, Jenkins RD, Bassett DR. Light scattering of [190] McCormick CL, Hester RD, Morgan SE, Safieddine AM. Water-
dilute hydrophobically modified alkali-soluble emulsion solutions: soluble copolymers. 31. Effects of molecular-parameters, solvation,
effects of hydrophobicity and spacer length of macromonomer. and polymer associations on drag reduction performance. Macro-
Macromolecules 2000;33:7021–8. molecules 1990;23:2132–9.
[170] Chassenieux C, Nicolai T, Durand D. Association of hydropho- [191] McCormick CL, Hester RD, Morgan SE, Safieddine AM. Water-
bically end-capped poly(ethylene oxide). Macromolecules soluble copolymers. 30. Effects of molecular-structure on drag
1997;30:4952–8. reduction efficiency. Macromolecules 1990;23:2124–31.
[171] Prochazka K, Martin TJ, Webber SE, Munk P. Onion-type micelles in [192] McCormick CL, Blackmon KP. Water-soluble copolymers.
aqueous media. Macromolecules 1996;29:6526–30. 21. Copolymers of acrylamide with 2-acrylamido-2-
[172] Prochazka K, Martin TJ, Munk P, Webber SE. Polyelectrolyte methylpropanedimethylammonium chloride: synthesis and
poly(tert-butyl acrylate)-block-poly(2-vinylpyridine) micelles in characterization. Polymer 1986;27:1971–5.
aqueous media. Macromolecules 1996;29:6518–25. [193] McCormick CL, Blackmon KP, Elliot DL. Water-soluble copoly-
[173] Biggs S, Selb J, Candau F. Copolymers of acrylamide/N- mers. 22. Copolymers of acrylamide with 2-acrylamido-2-
alkylacrylamide in aqueous solution: the effects of hydrolysis on methylpropanedimethylammonium chloride: aqueous solution
hydrophobic interactions. Polymer 1993;34:580–91. properties of a polycation. Polymer 1986;27:1976–80.
[174] Biggs S, Hill A, Selb J, Candau F. Copolymerization of acrylamide [194] McCormick CL, Blackmon KP. Water-soluble copolymers. 17.
and a hydrophobic monomer in an aqueous medium: effect of Copolymers of acrylamide with sodium 3-methacrylamido-3-
the surfactant on the copolymer microstructure. J Phys Chem methylbutanoate: synthesis and characterization. Macromolecules
1992;96:1505–11. 1986;19:1512–5.
[175] Branham KD, Davis DL, Middleton JC, McCormick CL. Water-soluble [195] McCormick CL, Elliot DL, Blackmon KP. Water-soluble copolymers.
polymers. 59. Investigation of the effects of polymer microstructure 18. Copolymers of acrylamide with sodium 3-methacrylamido-3-
on the associative behaviour of amphiphilic terpolymers of acry- methylbutanoate: microstructural studies and solution properties.
lamide, acrylic acid and N-[(4-decyl)phenyl]acrylamide. Polymer Macromolecules 1986;19:1516–22.
1994;35:4429–36. [196] Kathmann EE, White LA, McCormick CL. Water-soluble polymers.
[176] Branham KD, Snowden HS, McCormick CL. Water-soluble copoly- 73. Electrolyte- and pH-responsive zwitterionic copolymers
mers. 64. Effects of pH and composition on associative properties of 4-[(2-acrylamido-2-methylpropyl)-dimethylammonio]-
of amphiphilic acrylamide/acryl acid terpolymers. Macromolecules butanoate with 3-[(2-acrylamido-2-methylpropyl)dimethyl-
1996;29:254–62. ammonio]propanesulfonate. Macromolecules 1997;30:5297–304.
[177] Branham KD, Shafer GS, Hoyle CE, McCormick CL. Water- [197] McCormick CL, Salazar LC. Water soluble copolymers. 46.
soluble copolymers. 61. Microstructural investigation of Hydrophilic sulphobetaine copolymers of acrylamide and
pyrenesulfonamide-labeled polyelectrolytes. Variation of label 3-(2-acrylamido-2-methylpropanedimethyl-ammonio)-1-
proximity utilizing micellar polymerization. Macromolecules propanesulphonate. Polymer 1992;33:4617–24.
1995;28:6175–82. [198] Kathmann EE, White LA, McCormick CL. Water soluble polymers.
[178] Kramer MC, Steger JR, Hu Y, McCormick CL. Water-soluble 69. pH and electrolyte responsive copolymers of acrylamide and the
copolymers. 65. Environmentally responsive associations probed zwitterionic monomer 4-(2-acrylamido-2-methylpropyldimethyl-
by nonradiative energy transfer studies of naphthalene and ammonio) butanoate: synthesis and solution behaviour. Polymer
pyrene-labeled poly(acrylamide-co-sodium 11-(acrylamido)- 1997;38:871–8.
undecanoate). Macromolecules 1996;29:1992–7. [199] Kathmann EE, White LA, McCormick CL. Water soluble polymers.
[179] Kramer MC, Welch CG, Steger JR, McCormick CL. Water-soluble 70. Effects of methylene versus propylene spacers in the pH and
copolymers. 63. Rheological and photophysical studies on the asso- electrolyte responsiveness of zwitterionic copolymers incorporat-
ciative properties of pyrene-labeled poly[acrylamide-co-sodium ing carboxybetaine monomers. Polymer 1997;38:879–86.
11-(acrylamido)undecanoate]. Macromolecules 1995;28:5248–54. [200] Kathmann EE, McCormick CL. Water-soluble polymers. 72.
[180] Araujo E, Rharbi Y, Huang X, Winnik MA, Bassett DR, Jenkins RD. Synthesis and solution behavior of responsive copolymers of
Pyrene excimer kinetics in micelle like aggregates in a 0-HASE acrylamide and the zwitterionic monomer 6-(2-acrylamido-2-
associating polymer. Langmuir 2000;16:8664–71. methylpropyldimethylammonio) hexanoate. J Polym Sci Part A
[181] Kumacheva E, Rharbi Y, Winnik MA, Guo L, Tam KC, Jenkins RD. Polym Chem 1997;35:243–53.
Fluorescence studies of an alkaline swellable associative polymer [201] Kathmann EEL, Davis DD, McCormick CL. Water-soluble poly-
in aqueous solution. Langmuir 1997;13:182–6. mers. 60. Synthesis and solution behavior of terpolymers of acrylic
[182] Rufier C, Collet A, Viguier M, Oberdisse J, Mora S. SDS interactions acid, acrylamide and the zwitterionic monomer 3-[(2-acrylamido-
with hydrophobically end-capped poly(ethylene oxide) studied by 2-methylpropyl)dimethylammonio]-1-propanesulfonate. Macro-
13
C NMR and SANS. Macromolecules 2009;42:5226–35. molecules 1994;27:3156–61.
[183] Hu Y, Kramer MC, Boudreaux CJ, McCormick CL. Water-soluble [202] Kathmann EE, McCormick CL. Water-soluble polymers. 71.
copolymers. 62. Nonradiative energy-transfer studies of pH- pH responsive behavior of terpolymers of sodium acrylate,
responsive and salt-responsive associations in hydrophobically- acrylamide, and the zwitterionic monomer 4-(2-acrylamide-2-
modified, hydrolyzed maleic anhydride-ethyl vinyl ether copoly- methylpropanedimethylammonio)butanoate. J Polym Sci Part A
mers. Macromolecules 1995;28:7100–6. Polym Chem 1997;35:231–42.
1624 D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628

[203] Grassl B, Francois J, Billon L. Associating behaviour of polyacry- isopropylacrylamide) on its association in water. Macromolecules
lamide modified with a new hydrophobic zwitterionic monomer. 1998;31:2527–32.
Polym Int 2001;50:1162–9. [223] Zhang Y, Wu C, Fang Q, Zhang YX. A light scattering study of the
[204] McCormick CL, Johnson CB. Water-soluble polymers: 33. aggregation behavior of fluorocarbon modified polyacrylamides in
Ampholytic terpolymers of sodium 2-acrylamido-2-methylpro- water. Macromolecules 1996;29:2494–7.
panesulphonate with 2-acrylamido-2-methylpropanedimethy- [224] Zhang YX, Fang Q, Fu YQ, Da AH, Zhang Y, Wu C, Hogen-Esch TE.
lammonium chloride and acrylamide: Synthesis and aqueous Synthesis and characterization of fluorocarbon-modified poly(N-
solution behaviour. Polymer 1990;31:1100–7. isopropylacrylamide). Polym Int 2000;49:763–74.
[205] Mumick PS, Welch PM, Salazar LC, McCormick CL. Water-soluble [225] Zhou H, Song GQ, Zhang YX, Chen JY, Jiang M, Hogen-Esch TE,
copolymers. 56. Structure and solvation effects of polyampholytes Dieing R, Ma L, Haeussling L. Hydrophobically modified poly-
in drag reduction. Macromolecules 1994;27:323–31. electrolytes. 4. Synthesis and solution properties of fluorocarbon-
[206] Mumick PS. Structurally tailored water soluble polymers for the containing poly(acrylic acid). Macromol Chem Phys 2001;202:
study of drag reduction. PhD Dissertation, Hattiesburg MS: The 3057–64.
University of Southern Mississippi; 1993. [226] Armentrout RS, McCormick CL. Water soluble polymers. 76.
[207] Salazar LC. Synthesis and solution behavior of electrolyte- Electrolyte responsive cyclopolymers with sulfobetaine units
responsive polyampholytes. PhD Dissertation, Hattiesburg MS: The exhibiting polyelectrolyte or polyampholyte behavior in aqueous
University of Southern Mississippi; 1991. media. Macromolecules 2000;33:419–24.
[208] McCormick CL, Johnson CB. Water-soluble polymers. [227] Armentrout RS, McCormick CL. Water-soluble polymers. 77.
XXXIV. Ampholytic terpolymers of sodium 3-acrylamido-3- Amphoteric cyclopolymers with sulfobetaine units: phase behavior
methylbutanoate with 2-acrylamido-2-methylpropanedimethyl- in aqueous media and solubilization of p-cresol in microdomains.
ammonium chloride and acrylamide: synthesis and absorbancy Macromolecules 2000;33:2944–51.
behavior. J Macromol Sci Part A Pure Appl Chem 1990;27:539–47. [228] Chang YH, McCormick CL. Water-soluble copolymers.
[209] McCormick CL, Johnson CB. Water-soluble copolymers. 57. Amphiphilic cyclocopolymers of diallylalkoxybenzyl-
29. Ampholytic copolymers of sodium 2-acrylamido-2- methylammonium chloride and diallyl-dimethylammonium
methylpropanesulfonate with (2-acrylamido-2-methylpropyl)- chloride. Polymer 1994;35:3503–12.
dimethylammonium chloride: solution properties. Macro- [229] McCormick CL, Middleton JC, Cummins DF. Water-soluble copoly-
molecules 1988;21:694–9. mers. 37. Synthesis and characterization of responsive hydrophobi-
[210] McCormick CL, Johnson CB. Water-soluble polymers. 28. cally modified polyelectrolytes. Macromolecules 1992;25:1201–6.
Ampholytic copolymers of sodium 2-acrylamido-2-methyl- [230] McCormick CL, Middleton JC, Grady CE. Water soluble copolymers.
propanesulfonate with (2-acrylamido-2-methylpropyl)- 38. Synthesis and characterization of electrolyte responsive ter-
dimethylammonium chloride: synthesis and characterization. polymers of acrylamide, N-(4-butyl)phenylacrylamide, and sodium
Macromolecules 1988;21:686–93. acrylate, sodium-2-acrylamido-2-methylpropanesulphonate or
[211] McCormick CL, Salazar LC. Water-soluble copolymers. 43. sodiu-acrylamido-3-methylbutanoate. Polymer 1992;33:4184–90.
Ampholytic copolymers of sodium 2-(acrylamido)-2-methyl- [231] Ali SA, Umar Y, bu-Sharkh BF, Al-Muallem HA. Synthesis and com-
propanesulfonate with[2-(acrylamido)-2-methylpropyl]trimethy- parative solution properties of single-, twin-, and triple-tailed
lammonium chloride. Macromolecules 1992;25:1896–900. associating ionic polymers based on diallylammonium salts. J
[212] Petit F, Iliopoulos I, Audebert R, Szonyi S. Associating polyelec- Polym Sci Part A Polym Chem 2006;44:5480–94.
trolytes with perfluoroalkyl side chains: aggregation in aqueous [232] Umar Y, Al-Muallem HA, bu-Sharkh BF, Ali SA. Synthesis and solu-
solution, association with surfactants, and comparison with hydro- tion properties of hydrophobically associating ionic polymers made
genated analogues. Langmuir 1997;13:4229–33. from diallylammonium salts/sulfur dioxide cyclocopolymerization.
[213] Shedge AS, Lele AK, Wadgaonkar PP, Hourdet D, Pcrrin P, Chas- Polymer 2004;45:3651–61.
senieux C, Badiger MV. Hydrophobically modified poly(acrylic [233] Ali SA, Umar Y, Al-Muallem HA, bu-Sharkh BF. Synthesis and vis-
acid) using 3-pentadecylcyclohexylamine: synthesis and rheology. cosity of hydrophobically modified polymers containing dendritic
Macromol Chem Phys 2005;206:464–72. segments. J Appl Polym Sci 2008;109:1781–92.
[214] Tomatsu I, Hashidzume A, Yusa S, Morishima Y. Unique associative [234] Umar Y, bu-Sharkh BF, Ali SA. The effects of charge densities
properties of copolymers of sodium acrylate and oligo(ethylene on the associative properties of a pH-responsive hydrophobi-
oxide) alkyl ether methacrylates in water. Macromolecules cally modified sulfobetaine/sulfur dioxide terpolymer. Polymer
2005;38:7837–44. 2005;46:10709–17.
[215] Noda T, Hashidzume A, Morishima Y. Effects of spacer length [235] Ringsdorf H, Venzmer J, Winnik FM. Fluorescence studies of
on the side-chain micellization in random copolymers of hydrophobically modified poly(N-isopropylacrylamides). Macro-
sodium 2-(acrylamido)-2-methylpropanesulfonate and methacry- molecules 1991;24:1678–86.
lates substituted with ethylene oxide-based surfactant moieties. [236] Wang TK, Iliopoulos I, Audebert R. Aqueous-solution behavior
Macromolecules 2001;34:1308–17. of hydrophobically modified poly(acrylic acid). In: Shalaby SW,
[216] Noda T, Hashidzume A, Morishima Y. Rheological properties of McCormick CL, Butler GB, editors. Water soluble polymers. ACS
transient networks formed from copolymers of sodium acry- symposium series no. 467. Washington, DC: American Chemical
late and methacrylates substituted with amphiphiles: comparison Society; 1991. p. 218–31.
with sodium 2-(acrylamido)-2-methylpropanesulfonate copoly- [237] Xue W, Hamley IW, Castelletto V, Olmsted PD. Synthesis and
mers. Langmuir 2001;17:5984–91. characterization of hydrophobically modified polyacrylamides
[217] Zhong C, Luo P, Ye Z, Chen H. Characterization and solution proper- and some observations on rheological properties. Eur Polym J
ties of a novel water-soluble terpolymer for enhanced oil recovery. 2004;40:47–56.
Polym Bull 2009;62:79–89. [238] Yahaya GO, Ahdab AA, Ali SA, Abu-Sharkh BF, Hamad EZ.
[218] Smith GL, McCormick CL. Water-soluble polymers. 80. Rheo- Solution behavior of hydrophobically associating water-soluble
logical and photophysical studies of pH-responsive terpolymers block copolymers of acrylamide and N-benzylacrylamide. Polymer
containing hydrophobic twin-tailed acrylamide monomers. Macro- 2001;42:3363–72.
molecules 2001;34:5579–86. [239] Reddy GJ, Naidu SV, Reddy AVR. Synthesis and characteriza-
[219] McCormick CL, Hoyle CE, Clark MD. Water-soluble copolymers. tion of phenyl methacrylamide copolymers. Adv Polym Technol
26. Fluorescence probe studies of hydrophobically modified maleic 2006;25:41–50.
acid–ethyl vinyl ether copolymers. Polymer 1992;33:243–7. [240] Zhang HP, Xu K, Ai H, Chen DH, Xv LL, Chen MC. Synthe-
[220] Chang Y, McCormick CL. Water-soluble copolymers. 49. Effect sis, characterization and solution properties of hydrophobically
of the distribution of the hydrophobic cationic monomer modified polyelectrolyte poly(AA-co-TMSPMA). J Solution Chem
dimethyldodecyl(2-acrylamidoethyl)ammonium bromide on the 2008;37:1137–48.
solution behavior of associating acrylamide copolymers. Macro- [241] Hwang FS, Hogen-Esch TE. Effects of water-soluble spac-
molecules 1993;26:6121–6. ers on the hydrophobic association of fluorocarbon-modified
[221] Zhuang DQ, Da JCAH, Zhang YX, Dieing R, Ma L, Haeussling L. poly(acrylamide). Macromolecules 1995;28:3328–35.
Hydrophobically modified polyelectrolytes II: synthesis and char- [242] Zhang YX, Da AH, Butler GB, Hogen-Esch TE. A fluorine-containing
acterization of poly(acrylic acid-co-alkyl acrylate). Polym Adv hydrophobically associating polymer. 1. Synthesis and solution
Technol 2001;12:616–25. properties of copolymers of acrylamide and fluorine-containing
[222] Zhang Y, Mei L, Fang Q, Zhang YX, Jiang M, Wu C. Effect acrylates or methacrylates. J Polym Sci Part A Polym Chem
of incorporating a trace amount of fluorocarbon into poly(N- 1992;30:1383–91.
D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628 1625

[243] Bastiat G, Grassl B, Francois J. Study of sodium dodecyl sul- associating fluorocarbon-containing polymers. Macromolecules
fate/poly(propylene oxide) methacrylate mixed micelles for the 1992;25:4784–91.
synthesis of thermo-associative polymers by micellar polymeriza- [270] Zhang YX, Da AH, Hogen-Esch TE. A fluorocarbon-containing
tion. Polym Int 2002;51:958–65. hydrophobically associating polymer. J Polym Sci Part C Polym Lett
[244] Lara-Ceniceros AC, Rivera-Vallejo C, Jimenez-Regalado EJ. Synthesis 1990;28:213–8.
and characterization of telechelic polymers obtained by micellar [271] Amis EJ, Hu N, Seery TAP, Hogen-Esch TE, Hwang F. Associating
polymerization. Polym Bull 2007;59:499–508. polymers containing fluorocarbon hydrophobic units. In: Glass JE,
[245] Jimenez-Regalado EJ, Cadenas-Pliego G, Perez-Alvarez M, editor. Hydrophilic polymers—performance with environmental
Hernandez-Valdez Y. Study of three different families of water- acceptance. Advances in chemistry series 248. Washington, DC:
soluble copolymers: synthesis, characterization and viscoelastic American Chemical Society; 1996. p. 279–302.
behavior of semidilute solutions of polymers prepared by solution [272] Umar Y, Abu-Sharkh BF, Ali SA. The effects of zwitterionic and
polymerization. Polymer 2004;45:1993–2000. anionic charge densities in polymer chains on the viscosity
[246] McCormick CL, Hoyle CE, Clark MD. Water-soluble copolymers. behavior of a pH-responsive hydrophobically modified ionic poly-
36. Photphysical investigations of water-soluble copolymers mer. J Appl Polym Sci 2005;98:1404–11.
of 2-(1-naphthylacetamide)ethylacrylamide. Macromolecules [273] Fevola MJ, Bridges JK, Kellum MG, Hester RD, McCormick
1991;24:2397–403. CL. pH-responsive polyzwitterions: a comparative study of
[247] McCormick CL, Hoyle CE, Clark MD. Water-soluble copolymers. acrylamide-based polyampholyte terpolymers and polybetaine
35. Photophysical and rheological studies of the copolymer of copolymers. J Appl Polym Sci 2004;94:24–39.
methacrylic acid with 2-(1-naphthylacetyl)ethyl acrylate. Macro- [274] Johnson KM, Poe GD, Lockhead RY, McCormick CL. The synthe-
molecules 1990;23:3124–9. sis of hydrophobically modified water-soluble polyzwitterionic
[248] Emmons WD, Stevens TE. Acrylamide copolymer thickener for copolymers and responsiveness to surfactants in aqueous solution.
aqueous systems. US Pat 4,395,524; 1983. J Macromol Sci Part A Pure Appl Chem 2004;41:587–611.
[249] Emmons WD, Stevens TE. Acrylamid-copolymers suitable as thick- [275] Candau F, Biggs S, Hill A, Selb J. Synthesis, structure and prop-
ening agent for aqueous systems and aqueous compositions erties of hydrophobically associating polymers. Prog Org Coat
containing them. Eur Pat 63018; 1982. 1994;24:11–9.
[250] Kulicke W-M, Kniewske R, Klein J. Preparation, characterization, [276] Branham KD, Middleton JC, McCormick CL. Photophysical and rhe-
solution properties and rheological behaviour of polyacrylamide. ological properties of naphthalene-labeled water-soluble copoly-
Prog Polym Sci 1982;8:373–468. mers polymerized in surfactant solution. Polym Prepr Am Chem
[251] Evani S. Water-dispersible hydrophobic thickening agent. Eur Pat Soc 1991;32(1):106.
57875; 1982. [277] Candau F, Volpert E, Lacik I, Selb J. Free-radical polymerization
[252] Evani S. Water-dispersible hydrophobic thickening agent. US Pat in micellar media: effect of microenvironment. Macromol Symp
4,432,881; 1984. 1996;111:85–94.
[253] Bock J, Siano DB, Kowalik RM, Turner SP. Process for forming [278] Wang G-J, Engberts JBFN. Synthesis of hydrophobically and elec-
acrylamide-alkyl acrylamide copolymers. Eur Pat 115213; 1984. trostatically modified polyacrylamides and their catalytic effects
[254] Turner SP, Siano DB, Bock J. Acrylamide–alkylacrylamide copoly- on the unimolecular decarboxylation of 6-nitrobenzisoxazole-3-
mers. US Pat 4,520,182; 1985. carboxylate anion. Langmuir 1995;11:3856–61.
[255] Turner SP, Siano DB, Bock J. Microemulsion process for producing [279] Vaskova V, Renoux D, Bernard M, Selb J, Candau F. Mechanism of
acrylamide-alkyl acrylamide copolymers. US Pat 4,521,580;1985. copolymerization of acrylamide with a polymerizable surfactant.
[256] Turner SP, Siano DB, Bock J. Micellar process for the production of Polym Adv Technol 1995;6:441–51.
acrylamide-alkyl acrylamide copolymers. US Pat 4,528,348;1985. [280] Renoux D, Selb J, Candau F. Aqueous solution properties of
[257] Vittadello ST, Biggs S. Shear history effects in associative thickener hydrophobically associating copolymers. Prog Colloid Polym Sci
solutions. Macromolecules 1998;31:7691–7. 1994;97:213–7.
[258] Regalado EJ, Selb J, Candau F. Viscoelastic behavior of semidi- [281] Charalambopoulou A, Bokias G, Staikos G. Template copolymerisa-
lute solutions of multisticker polymer chains. Macromolecules tion of N-isopropylacrylamide with a cationic monomer: influence
1999;32:8580–8. of the template on the solution properties of the product. Polymer
[259] Shashkina YA, Zaroslov YD, Smirnov VA, Philippova OE, Khokhlov 2002;43:2637–43.
AR, Pryakhina TA, Churochkina NA. Hydrophobic aggregation in [282] Rainaldi I, Cristallini C, Ciardelli G, Giusti P. Copolymerization
aqueous solutions of hydrophobically modified polyacrylamide in of acrylic acid and 2-hydroxyethyl methacrylate onto poly(N-
the vicinity of overlap concentration. Polymer 2003;44:2289–93. vinylpyrrolidone): template influence on comonomer reactivity.
[260] Ma J-T, Huang R-H, Zhao L, Zhang X. Solution properties of Macromol Chem Phys 2000;201:2424–31.
ionic hydrophobically associating polyacrylamide with an arylalkyl [283] Frisch HL, Xu QH. Copolymerization of styrene and methacrylic-
group. J Appl Polym Sci 2005;97:316–21. acid in the presence of poly(2-vinylpyridine) as the template.
[261] Bock J, Valint PL. Process for preparing hydrophobically associating Macromolecules 1992;25:5145–9.
terpolymers containing sulfonate functionality. US Pat 4,730,028; [284] Jantas R, Polowinski S. Copolymerization of a multiallyl monomer
1988. with styrene. Acta Polym 1989;40:225–8.
[262] Valint PL, Bock J. Hydrophobically associating terpolymers contain- [285] Polowinski S, Janowska G. Thermal copolymerization of acryloni-
ing sulfonate functionality. US Pat 5,089,578; 1992. trile with methacrylate units arranged in matrix. Eur Polym J
[263] Bock J, Siano DB, Turner SP. Hydrophobically associating terpoly- 1975;11:183–5.
mers of acrylamide, salts of acrylic acid and alkyl acrylamide. US [286] Effing JJ, McLennan IJ, Kwak JCT. Associative phase separation
Pat 4,694,046; 1987. observed in a hydrophobically modified poly(acrylamide)/sodium
[264] Schulz DN, Berluche E, Maurer JJ, Bock J. Tetrapolymers of N-vinyl dodecyl sulfate system. J Phys Chem 1994;98:2499–502.
pyrrolidone/acrylamide/salt of acrylic acid/N-alkyl acrylamide. US [287] Effing JJ, McLennan IJ, van Os NM, Kwak JCT. 1 H NMR
Pat 4,663,408; 1987. investigations of the interactions between anionic surfactants
[265] Hill A, Candau F, Selb J. Aqueous solution properties of hydrophobi- and hydrophobically modified poly(acrylamide)s. J Phys Chem
cally associating copolymers. Prog Colloid Polym Sci 1992;84:61–5. 1994;98:12397–402.
[266] Camail M, Margaillan A, Martin I. Copolymers of N-alkyl- and N- [288] McCormick CL, Salazar LC. Water-soluble copolymers. 45.
arylalkylacrylamides with acrylamide: influence of hydrophobic Ampholytic terpolymers of acrylamide with sodium 3-
structure on associative properties. Part 1. Viscometric behaviour acrylamido-3-methylbutanoate and 2-acrylamido-2-methyl-
in dilute solution and drag reduction performance. Polym Int propanetrimethylammonium chloride. J Appl Polym Sci
2009;58:149–54. 1993;48:1115–20.
[267] Camail M, Margaillan A, Martin I. Copolymers of N-alkyl- and N- [289] Peiffer DG, Lundberg RD. Synthesis and viscometric prop-
arylalkylacrylamides with acrylamide: influence of hydrophobic erties of low charge-density ampholytic ionomers. Polymer
structure on associative properties. Part 2. Rheological behaviour 1985;26:1058–68.
in semi-dilute solution. Polym Int 2009;58:155–62. [290] Petit-Agnely F, Iliopoulos I, Zana R. Hydrophobically modified
[268] Zhao Y, Zhou J, Xu X, Liu W, Zhang J, Fan M, Wang J. Synthesis sodium polyacrylates in aqueous solutions: association mechanism
and characterization of a series of modified polyacrylamide. Colloid and characterization of the aggregates by fluorescence probing.
Polym Sci 2009;287:237–41. Langmuir 2000;16:9921–7.
[269] Seery TAP, Yassini M, Hogen-Esch TE, Amis EJ. Static and dynamic [291] Leibler L, Rubinstein M, Colby RH. Dynamics of reversible networks.
light scattering characterization of solutions of hydrophobically Macromolecules 1991;24:4701–7.
1626 D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628

[292] Candau F, Regalado EJ, Selb J. Scaling behavior of the zero shear differential scanning calorimetry (DSC), and turbidity measure-
viscosity of hydrophobically modified poly(acrylamide)s. Macro- ments. J Colloid Interface Sci 1996;179:20–33.
molecules 1998;31:5550–2. [317] Emmons WD, Valley H, Stevens TE. Polyurethane thickeners in latex
[293] Feng Y, Billon L, Grassl B, Bastiat G, Borisov O, François J. compositions. US Pat 4,079,028; 1978.
Hydrophobically associating polyacrylamides and their partially [318] Xu B, Yekta A, Li L, Masoumi Z, Winnik MA. The functional-
hydrolyzed derivatives prepared by post modification. 2. Proper- ity of associative polymer networks: the association behavior of
ties of non-hydrolyzed polymers in pure water and brine. Polymer hydrophobically modified urethane-ethoxylate (HEUR) associative
2005;46:9283–95. polymers in aqueous solution. Colloids Surf A 1996;112:239–50.
[294] Feng Y, Grassl B, Billon L, Khoukh A, Francois J. Effects of [319] Yekta A, Xu B, Duhamel J, Adiwidjaja H, Winnik MA. Fluorescence
NaCl on steady rheological behaviour in aqueous solutions studies of associating polymers in water: determination of the
of hydrophobically modified polyacrylamide and its partially chain end aggregation number and a model for the association
hydrolyzed analogues prepared by post modification. Polym Int process. Macromolecules 1995;28:956–66.
2002;51:939–47. [320] Kaczmarski JP, Glass JE. Synthesis and characterization of step-
[295] Witten Jr TA, Cohen MH. Cross-linking in shear-thickening growth hydrophobically-modified ethoxylated urethane associa-
ionomers. Macromolecules 1985;18:1915–8. tive thickeners. Langmuir 1994;10:3035–42.
[296] Ballard MJ, Buscall R, Waite FA. The theory of shear-thickening [321] Liu F, Frere Y, Francois J. Association properties of poly(ethylene
polymer solutions. Polymer 1988;29:1287–93. oxide) modified by pendant aliphatic groups. Polymer
[297] Maerker JM, Sinton SW. Rheology resulting from shear-induced 2001;42:2969–83.
structure in associating polymer-solutions. J Rheol 1986;30:77–99. [322] Barmar M, Barikani M, Kaffashi B. Synthesis of ethoxylated ure-
[298] Berret JF, Calvet D, Collet A, Viguier M. Fluorocarbon associative thane and modification with cetyl alcohol as thickener. Iran Polym
polymers. Curr Opin Colloid Interface Sci 2003;8:296–306. J 2001;10:331–5.
[299] Kujawa P, udibert-Hayet A, Selb J, Candau F. Rheological properties [323] Barmar M. Study of the effect of PEG length in uni-HEUR thickener
of multisticker associative polyelectrolytes in semidilute aqueous behavior. J Appl Polym Sci 2009;111:1751–4.
solutions. J Polym Sci Part B Polym Phys 2004;42:1640–55. [324] Barmar M, Barikani M, Kaffashi B. Steady shear viscosity study of
[300] Magny B, Iliopoulos I, Zana R, Audebert R. Mixed micelles formed various HEUR models with different hydrophilic and hydrophobic
by cationic surfactants and anionic hydrophobically modified poly- sizes. Colloids Surf A 2005;253:77–82.
electrolytes. Langmuir 1994;10:3180–7. [325] Barmar M, Barikani M, Kaffashi B. The effect of molecular weight
[301] Iliopoulos I, Olsson U. Polyelectrolyte association to micelles and on the behaviour of step-growth hydrophobically modified ethoxy-
bilayers. J Phys Chem 1994;98:1500–5. lated urethane (S-G HEUR) end-capped with dodecyl alcohol. Iran
[302] Chang Y, Lochhead RY, McCormick CL. Water-soluble copolymers. Polym J 2004;13:241–6.
50. Effect of surfactant addition on the solution properties of [326] Vorobyova O, Yekta A, Winnik MA, Lau W. Fluorescent probe studies
amphiphilic copolymers of acrylamide and dimethyldodecyl(2- of the association in an aqueous solution of a hydrophobically mod-
acrylamidoethyl)ammonium bromide. Macromolecules ified poly(ethylene oxide). Macromolecules 1998;31:8998–9007.
1994;27:2145–50. [327] Wang YC, Winnik MA. Onset of aggregation for water-soluble
[303] Winnik FM, Ringsdorf H, Venzmer J. Interactions of surfactants with polymeric associative thickeners—a fluorescence study. Langmuir
hydrophobically modified poly(N-isopropylacrylamides). 1. Fluo- 1990;6:1437–9.
rescence probe studies. Langmuir 1991;7:905–11. [328] Fonnum G, Bakke J, Hansen FK. Associative thickeners. Part 1.
[304] Winnik FM, Ringsdorf H, Venzmer J. Interaction of surfactants with Synthesis, rheology and aggregation behavior. Colloid Polym Sci
hydrophobically modified poly(N-isopropylacrylamides). 2. Fluo- 1993;271:380–9.
rescence label studies. Langmuir 1991;7:912–7. [329] Xu B, Li L, Yekta A, Masoumi Z, Kanagalingam S, Winnik MA, Zhang
[305] Penott-Chang EK, Gouveia L, Fernandez IJ, Muller AJ, az-Barrios A, K, MacDonald PM, Menchen S. Synthesis, characterization, and
Saez A. Rheology of aqueous solutions of hydrophobically modified rheological behavior of polyethylene glycols end-capped with flu-
polyacrylamides and surfactants. Colloids Surf A 2007;295:99–106. orocarbon hydrophobes. Langmuir 1997;13:2447–56.
[306] Gouveia LM, Muller AJ. The effect of NaCl addition on the rheo- [330] Calvet D, Collet A, Viguier M, Berret J-F, Séréro Y. Perfluoroalkyl
logical behavior of cetyltrimethylammonium p-toluenesulfonate end-capped poly(ethylene oxide). Synthesis, characterization,
(CTAT) aqueous solutions and their mixtures with hydropho- and rheological behavior in aqueous solution. Macromolecules
bically modified polyacrylamide aqueous solutions. Rheol Acta 2003;36:449–57.
2009;48:163–75. [331] Séréro Y, Aznar R, Porte G, Berret J-F, Calvet D, Collet A, Viguier M.
[307] Nyström B, Thuresson K, Lindman B. Rheological and dynamic Associating polymers: from “Flowers” to transient networks. Phys
light-scattering studies on aqueous solutions of a hydrophobically Rev Lett 1998;82:5584–7.
modified nonionic cellulose ether and its unmodified analogue. [332] Hartmann P, Collet A, Viguier M. Synthesis and characterization
Langmuir 1995;11:1994–2002. of model fluoroacylated poly(ethylene oxide). J Fluorine Chem
[308] Goddard ED, Leung PS. Studies of gel formation, phase behavior and 1999;95:145–51.
surface tension in mixtures of a hydrophobically modified cationic [333] Zhang HS, Hogen-Esch TE, Boschet F, Margaillan A. Complex
cellulose polymer and surfactant. Colloids Surf 1992;65:211–9. formation of beta-cyclodextrin- and perfluorocarbon-modified
[309] Goddard ED, Leung PS. Interaction of cationic surfactants with water-soluble polymers. Langmuir 1998;14:4972–7.
a hydrophobically modified cationic cellulose polymer. Langmuir [334] Boschet F, Branger C, Margaillan A, Condamine E. Synthesis, charac-
1992;8:1499–500. terisation and aqueous behaviour of a one-ended perfluorocarbon-
[310] L’Alloret F, Hourdet D, Audebert R. Aqueous solution behav- modified poly(ethylene glycol). Polymer 2002;43:5329–34.
ior of new thermoassociative polymers. Colloid Polym Sci [335] Hu YZ, Zhao CL, Winnik MA, Sundararajan PR. Fluorescence studies
1995;273:1163–73. of the interaction of sodium dodecyl-sulfate with hydrophobically
[311] Taylor LD, Cerankowski LD. Preparation of films exhibiting a modified poly(ethylene oxide). Langmuir 1990;6:880–3.
balanced temperature-dependence to permeation by aqueous- [336] Wetzel WH, Chen M, Glass JE. Associative thickeners. An overview
solutions—study of lower consolute behavior. J Polym Sci Part A with an emphasis on synthetic procedures. In: Glass JE, editor.
Polym Chem 1975;13:2551–70. Hydrophilic polymers—performance with environmental accep-
[312] Guillet JE, Rendall WA. Studies of the antenna effect in polymer- tance. Advances in chemistry series 248. Washington, DC:
molecules. 8. Photophysics of water-soluble copolymers of American Chemical Society; 1996. p. 163–79.
1-naphthylmethyl methacrylate and acrylic-acid. Macromolecules [337] Yekta A, Nivaggioli T, Kanagalingam S, Xu B, Masoumi Z, Winnik
1986;19:224–30. MA. Urethane-coupled poly(ethylene glycol) polymers containing
[313] Jenkins RD. The fundamental thickening mechanism of associative hydrophobic end groups: NMR characterization as a step toward
polymers in latex systems, a rheological study. PhD Dissertation, determining aggregation numbers in aqueous solutions. In: Glass
Bethlehem PA: Lehigh University; 1990. JE, editor. Hydrophilic polymers—performance with environmen-
[314] Liao D, Dai S, Tam KC. Rheological properties of hydrophobic tal acceptance. Advances in chemistry series 248. Washington, DC:
ethoxylated urethane (HEUR) in the presence of methylated b- American Chemical Society; 1996. p. 363–76.
cyclodextrin. Polymer 2004;45:8339–48. [338] Yekta A, Duhamel J, Adiwidjaja H, Brochard P, Winnik MA. Asso-
[315] Glass JE, Karunasena A. Associative thickeners: from nonsense to ciation structure of telechelic associative thickeners in water.
reality. Polym Mater Sci Eng 1989;61:145–52. Langmuir 1993;9:881–3.
[316] AbrahmsenAlami S, Alami E, Francois J. The lyotropic cubic phase of [339] Singer W, Tenneck NJ, Driscoll AE. Method of improving viscosity
model associative polymers: small-angle X-ray scattering (SAXS), stability of aqueous compositions. US Pat 3,770,684; 1973.
D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628 1627

[340] Daoud M, Cotton JP. Star shaped polymers, a model for the confor- [365] Prazeres TJV, Duhamel J, Olesen K, Shay G. Correlations between
mation and its concentration dependence. J Phys 1983;43:531–8. the viscoelastic behavior of pyrene-labeled associative polymers
[341] Witten TA, Pincus PA. Colloid stabilization by long grafted poly- and the associations of their fluorescent hydrophobes. J Phys Chem
mers. Macromolecules 1986;19:2509–13. B 2005;109:17406–16.
[342] Semenov AN, Joanny J-F, Khokhlov AR. Associating poly- [366] Tirtaatmadja V, Tam KC, Jenkins RD. Superposition of oscillations
mers: equilibrium and linear viscoelasticity. Macromolecules on steady share flow as a technique for investigating the structure
1995;28:1066–75. of associative polymers. Macromolecules 1997;30:1426–33.
[343] Xu B, Yekta A, Winnik MA, Sadeghy-Dalivand K, James DF, Jenkins R, [367] Tirtaatmadja V, Tam KC, Jenkins RD. Rheological properties of
Basset D. Viscoelastic properties in water of comb associative poly- model alkali-soluble associative (HASE) polymers: effect of varying
mers based on poly(ethylene oxide). Langmuir 1997;13:6903–11. hydrophobe chain length. Macromolecules 1997;30:3271–82.
[344] Alami E, AbrahmsenAlami S, Vasilescu M, Almgren M. A comparison [368] Tam KC, Guo L, Jenkins RD, Bassett DR. Viscoelastic properties of
between hydrophobically end-gapped poly(ethylene oxide) with hydrophobically modified alkali-soluble emulsion in salt solutions.
ether and urethane bonds. J Colloid Interface Sci 1997;193:152–62. Polymer 1999;40:6369–79.
[345] Tanaka F, Edwards SF. Viscoelastic properties of physically [369] Dai S, Tam KC, Jenkins RD. Microstructure of dilute hydrophobically
crosslinked networks. Part 1. Non-linear stationary viscoelasticity. modified alkali soluble emulsion in aqueous salt solution. Macro-
J Non-Newtonian Fluid Mech 1992;43:247–71. molecules 2000;33:404–11.
[346] Tanaka F, Edwards SF. Viscoelastic properties of physically [370] Nagashima K, Strashko V, MacDonald PM, Jenkins RD, Bassett DR.
crosslinked networks. Part 2. Dynamic mechanical moduli. J Non- Diffusion of model hydrophobic alkali-swellable emulsion associa-
Newtonian Fluid Mech 1992;43:273–88. tive thickeners. Macromolecules 2000;33:9329–39.
[347] Tanaka F, Edwards SF. Viscoelastic properties of physically [371] English RJ, Gulati HS, Jenkins RD, Khan SA. Solution rheology of
crosslinked networks. Part 3. Time-dependent phenomena. J Non- a hydrophobically modified alkali-soluble associative polymer. J
Newtonian Fluid Mech 1992;43:289–309. Rheol 1997;41:427–44.
[348] Tanaka F, Edwards SF. Viscoelastic properties of physically cross- [372] Jenkins RD, Bassett DR, Shay GD. Polymers containing
linked networks. Transient network theory. Macromolecules macromonomers. US Pat 5,292,843; 1994.
1992;25:1516–23. [373] Siu H, Duhamel J. Comparison of the association level of a pyrene-
[349] Groot RD, Agterof WGM. Dynamic viscoelastic modulus of labeled associative polymer obtained from an analysis based on
associative polymer networks: off-lattice simulations, theory two different models. J Phys Chem B 2005;109:1770–80.
and comparison to experiments. Macromolecules 1995;28: [374] Prazeres TJV, Beingessner R, Duhamel J, Olesen K, Shay G, Bas-
6284–95. sett DR. Characterization of the association level of pyrene-labeled
[350] Annable T, Buscall R, Ettelaie R. Network formation and its con- HASEs by fluorescence. Macromolecules 2001;34:7876–84.
sequences for the physical behaviour of associating polymers in [375] Siu H, Duhamel J. Associations between a pyrene-labeled
solution. Colloids Surf A 1996;112:97–116. hydrophobically modified alkali swellable emulsion copolymer and
[351] Milner ST, Witten TA. Bridging attraction by telechelic polymers. sodium dodecyl sulfate probed by fluorescence, surface tension,
Macromolecules 1992;25:5495–503. and viscometry. Macromolecules 2006;39:1144–55.
[352] Witten TA. Associating polymers and shear thickening. J Phys [376] Wu W, Shay GD. Tailoring HASE rheology through polymer design:
1988;49:1055–63. effects of hydrophobe size, acid content, and molecular weight. J
[353] Annable T, Buscall R, Ettelaie R, Shepherd P, Whittlestone D. Influ- Coat Technol Res 2005;2:423–33.
ence of surfactants on the rheology of associating polymers in [377] Siddiq M, Tam KC, Jenkins RD. Dissolution behaviour of model
solution. Langmuir 1994;10:1060–70. alkali-soluble emulsion polymers: effects of molecular weights and
[354] Zhang K, Xu B, Winnik MA, MacDonald PM. Surfactant interactions ionic strength. Colloid Polym Sci 1999;277:1172–8.
with HEUR associating polymers. J Phys Chem 1996;100:9834–41. [378] Knaebel A, Skouri R, Munch JP, Candau SJ. Structural and rheologi-
[355] Binana-Limbele W, Clouet F, Francois J. Hydrophobically end- cal properties of hydrophobically modified alkali-soluble emulsion
capped poly(ethylene oxide) urethanes. Part 3. Effect of sodium solutions. J Polym Sci Part B Polym Phys 2002;40:1985–94.
dodecyl sulfate on their association in aqueous solution. Colloid [379] Tan H, Tam KC, Jenkins RD. Network structure of a model HASE
Polym Sci 1993;271:748–58. polymer in semidilute salt solutions. J Appl Polym Sci 2001;79:
[356] Persson K, Wang G, Olofsson G. Self-diffusion, thermal effects and 1486–96.
viscosity of a monodisperse associative polymer. Self-association [380] Tam KC, Farmer ML, Jenkins RD, Bassett DR. Rheological proper-
and interaction with surfactants. J Chem Soc Faraday Trans ties of hydrophobically modified alkali-soluble polymers—effects
1994;90:3555–62. of ethylene-oxide chain length. J Polym Sci Part B Polym Phys
[357] Persson K, Bales BL. EPR study of an associative polymer in solu- 1998;36:2275–90.
tion. Determination of aggregation number and interactions with [381] Ng WK, Tam KC, Jenkins RD. Rheological properties of methacrylic
surfactants. J Chem Soc Faraday Trans 1995;91:2863–70. acid/ethyl acrylate co-polymer: comparison between an
[358] Kim D-H, Kom J-W, Oh S-G, Kim J, Han S-H, Chung DJ, Suh unmodified and hydrophobically modified system. Polymer
K-D. Effects of nonionic surfactant on the rheological prop- 2001;42:249–59.
erty of associative polymers in complex formulations. Polymer [382] Seng WP, Tam KC, Jenkins RD. Rheological properties of model
2007;48:3817–21. alkali-soluble associative (HASE) polymer in ionic and non-ionic
[359] Lundberg DJ, Ma Z, Alahapperuna K, Glass JE. Surfactant influences surfactant solutions. Colloids Surf A 1999;154:365–82.
on hydrophobically modified thickener rheology. In: Schulz DN, [383] Tirtaatmadja V, Tam KC, Jenkins RD. Effects of temperature on the
Glass JE, editors. Polymers as rheology modifiers. ACS symposium flow dynamics of a model HASE associative polymer in nonionic
series no. 462. Washington, DC: American Chemical Society; 1991. surfactant solutions. Langmuir 1999;15:7537–45.
p. 234–53. [384] Tirtaatmadja V, Tam KC, Jenkins RD. Effect of a nonionic surfactant
[360] Harada A. Design and construction of supramolecular architec- on the flow dynamics of a model HASE associative polymer. AIChE
tures consisting of cyclodextrins and polymers. Metal complex J 1998;44:2756–65.
catalysts supercritical fluid polymerization supramolecular archi- [385] Aubry T, Moan M. Influence of a nonionic surfactant on the rheol-
tecture. Adv Polym Sci 1997;133:141–91. ogy of a hydrophobically associating water soluble polymer. J Rheol
[361] Liao D, Dai S, Tam KC. Rheological properties of a telechelic associa- 1996;40:441–8.
tive polymer in the presence of a- and methylated b-cyclodextrins. [386] Kjoniksen AL, Beheshti N, Kotlar HK, Zhu KZ, Nystrom B. Modified
J Phys Chem 2007;111:371–8. polysaccharides for use in enhanced oil recovery applications. Eur
[362] Liao D, Dai S, Tam KC. Interaction between fluorocarbon end- Polym J 2008;44:959–67.
capped poly(ethylene oxide) and cyclodextrins. Macromolecules [387] Kawakami K, Ihara T, Nishioka T, Kitsuki T, Suzuki Y. Salt toler-
2007;40:2936–45. ance of an aqueous solution of a novel amphiphilic polysaccharide
[363] English RJ, Raghavan SR, Jenkins RD, Khan SA. Associative polymers derivative. Langmuir 2006;22:3337–43.
bearing n-alkyl hydrophobes: rheological evidence for micro-gel [388] Ihara T, Nishioka T, Kamitani H, Kitsuki T. Solution properties of a
behavior. J Rheol 1999;43:1175–94. novel polysaccharide derivative. Chem Lett 2004;33:1094–5.
[364] Jenkins RD, DeLong LM, Bassett DR. Influence of alkali-soluble asso- [389] Akiyama E, Yamamoto T, Yago Y, Hotta H, Ihara T, Kitsuki T. Thicken-
ciative emulsion polymer architecture on rheology. In: Glass JE, ing properties and emulsification mechanisms of new derivatives
editor. Hydrophilic polymers—performance with environmental of polysaccharide in aqueous solution. 2. The effect of the substitu-
acceptance. Advances in chemistry series 248. Washington, DC: tion ratio of hydrophobic/hydrophilic moieties. J Colloid Interface
American Chemical Society; 1996. p. 425–47. Sci 2007;311:438–46.
1628 D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 1558–1628

[390] Karlberg M, Thuresson K, Piculell L, Lindman B. Mixed solutions of [415] Zhao GQ, Khin CC, Chen SB, Chen BH. Nonionic surfactant
hydrophobically modified graft and block copolymers. Colloids Surf and temperature effects on the viscosity of hydrophobically
A 2004;236:159–64. modified hydroxyethyl cellulose solutions. J Phys Chem B
[391] Hwang FS, Hogenesch TE. Fluorocarbon-modified water-soluble 2005;109:14198–204.
cellulose derivatives. Macromolecules 1993;26:3156–60. [416] Dualeh AJ, Steiner CA. Hydrophobic microphase formation in
[392] Dai SS, Ye L, Huang RH. A study on the solution behavior of IPBC- surfactant solutions containing an amphiphilic graft copolymer.
hydrophobically-modified hydroxyethyl cellulose. J Appl Polym Sci Macromolecules 1990;23:251–5.
2006;100:2824–31. [417] Evertsson H, Nilsson S. Microstructures formed in aqueous solu-
[393] Ye L, Li Q, Huang RH. Study on the rheological behavior of tions of a hydrophobically modified nonionic cellulose derivative
the hydrophobically modified hydroxyethyl cellulose with 1,2- and sodium dodecyl sulfate: a fluorescence probe investigation.
epoxyhexadecane. J Appl Polym Sci 2006;101:2953–9. Carbohydr Polym 1999;40:293–8.
[394] Li Q, Ye L, Cai Y, Huang RH. Study of rheological behavior of [418] Kaczmarski JP, Tarng MR, Ma ZY, Glass JE. Surfactant and salin-
hydrophobically modified hydroxyethyl cellulose. J Appl Polym Sci ity influences on associative thickener aqueous solution rheology.
2006;100:3346–52. Colloids Surf A 1999;147:39–53.
[395] Gonzalez JM, Muller AJ, Torres MF, Saez AE. The role of shear [419] Lauten RA, Nystrom B. Time dependent association phenom-
and elongation in the flow of solutions of semi-flexible polymers ena in dilute aqueous mixtures of a hydrophobically modified
through porous media. Rheol Acta 2005;44:396–405. cellulose derivative and an anionic surfactant. Colloids Surf, A
[396] Shaw KG, Leipold DP. New cellulosic polymers for rheology control 2003;219:45–53.
of latex paints. J Coat Technol 1985;57:63–72. [420] Maestro A, Gonzalez C, Gutierrez JM. Interaction of surfactants with
[397] Kästner U, Hoffmann H, Donges R, Ehrler R. Interactions between thickeners used in waterborne paints: a rheological study. J Colloid
modified hydroxyethyl cellulose (HEC) and surfactants. Colloids Interface Sci 2005;288:597–605.
Surf A 1996;112:209–25. [421] Nilsson S, Goldraich M, Lindman B, Talmon Y. Novel organized
[398] Stone FW, Rutherford JM Jr. Quaternary nitrogen-containing cellu- structures in mixtures of a hydrophobically modified poly-
lose ethers. US Pat 3,472,840; 1969. mer and two oppositely charged surfactants. Langmuir 2000;16:
[399] Kästner U, Hoffmann H, Donges R, Ehrler R. Hydrophobically and 6825–32.
cationically modified hydroxyethyl cellulose and their interactions [422] Nilsson S, Thuresson K, Lindman B, Nystrom B. Associations in mix-
with surfactants. Colloids Surf A 1994;82:279–97. tures of hydrophobically modified polymer and surfactant in dilute
[400] Winnik FM, Winnik MA, Tazuke S, Ober CK. Synthesis and char- and semidilute aqueous solutions. A rheology and PFG NMR self-
acterization of pyrene-labeled (hydroxypropyl)cellulose and its diffusion investigation. Macromolecules 2000;33:9641–9.
fluorescence in solution. Macromolecules 1987;20:38–44. [423] Patruyo LG, Muller AJ, Saez AE. Shear and extensional rheology of
[401] Karlson L, Joabsson F, Thuresson K. Phase behavior and rheology solutions of modified hydroxyethyl celluloses and sodium dodecyl
in water and in model paint formulations thickened with HM- sulfate. Polymer 2002;43:6481–93.
EHEC: influence of the chemical structure and the distribution of [424] Piculell L, Egermayer M, Sjostrom J. Rheology of mixed solutions of
hydrophobic tails. Carbohydr Polym 2000;41:25–35. an associating polymer with a surfactant. Why are different surfac-
[402] Badiger MV, Lutz A, Wolf BA. Interrelation between the ther- tants different? Langmuir 2003;19:3643–9.
modynamic and viscometric behaviour of aqueous solutions of [425] Thuresson K, Nystrom B, Wang G, Lindman B. Effect of surfactant on
hydrophobically modified ethyl hydroxyethyl cellulose. Polymer structural and thermodynamic properties of aqueous-solutions of
2000;41:1377–84. hydrophobically-modified ethyl(hydroxyethyl)cellulose. Langmuir
[403] Um SU, Poptoshev E, Pugh RJ. Aqueous solutions of ethyl (hydrox- 1995;11:3730–6.
yethyl) cellulose and hydrophobic modified ethyl (hydroxyethyl) [426] Thuresson K, Soderman O, Hansson P, Wang G. Binding of SDS to
cellulose polymer: dynamic surface tension measurements. J Col- ethyl(hydroxyethyl)cellulose. Effect of hydrophobic modification
loid Interface Sci 1997;193:41–9. of the polymer. J Phys Chem 1996;100:4909–18.
[404] Zou S, Zhang WK, Zhang X, Jiang BZ. Study on polymer micelles [427] Thuresson K, Lindman B, Nystrom B. Effect of hydrophobic modi-
of hydrophobically modified ethyl hydroxyethyl cellulose using fication of a nonionic cellulose derivative on the interaction with
single-molecule force spectroscopy. Langmuir 2001;17:4799–808. surfactants. Rheology. J Phys Chem B 1997;101:6450–9.
[405] Nishikawa K, Yekta A, Pham HH, Winnik MA, Sau AC. Fluores- [428] Thuresson K, Lindman B. Effect of hydrophobic modification of
cence studies of hydrophobically modified hydroxyethylcellulose a nonionic cellulose derivative on the interaction with surfac-
(HMHEC) and pyrene labeled HMHEC. Langmuir 1998;14:7119–29. tants. Phase behavior and association. J Phys Chem B 1997;101:
[406] Miyajima T, Kitsuki T, Kita K, Kamitani H, Yamaki K. Polysaccha- 6460–8.
ride derivative, and preparation process and use thereof. US Pat [429] Joabsson F, Rosen O, Thuresson K, Piculell L, Lindman B. Phase
5,891,450; 1999. behavior of a “Clouding” nonionic polymer in water. Effects of
[407] Boström P, Ingvarsson I, Sundberg K. Water soluble nonionic cellu- hydrophobic modification and added surfactant on phase compo-
lose ethers and their use in paints. US Pat 5,140,099; 1992. sitions. J Phys Chem B 1998;102:2954–9.
[408] Tanaka R, Meadows J, Phillips GO, Williams PA. Viscometric and [430] Panmai S, Prud’homme RK, Peiffer DG, Jockusch S, Turro NJ. Interac-
spectroscopic studies on the solution behavior of hydrophobically tions between hydrophobically modified polymers and surfactants:
modified cellulosic polymers. Carbohydr Polym 1990;12:443–59. a fluorescence study. Langmuir 2002;18:3860–4.
[409] Landoll LM. Modified nonionic cellulose ethers. US Pat 4,228,277; [431] Karlberg M, Stjerndahl M, Lundberg D, Piculell L. Mixed solutions
1980. of an associating polymer with a cleavable surfactant. Langmuir
[410] Zhao GQ, Chen SB. Nonlinear rheology of aqueous solutions of 2005;21:9756–63.
hydrophobically modified hydroxyethyl cellulose with nonionic [432] Bu HT, Kjoniksen AL, Knudsen KD, Nystrom B. Effects of surfac-
surfactant. J Colloid Interface Sci 2007;316:858–66. tant and temperature on rheological and structural properties of
[411] Maestro A, Gonzalez C, Gutierrez JM. Thickening mechanism in semidilute aqueous solutions of unmodified and hydrophobically
associative polymers. Macromol Symp 2002;187:919–27. modified alginate. Langmuir 2005;21:10923–30.
[412] Picton L, Muller G. Rheological properties of modified cellulosic [433] Nilsson S, Thuresson K, Hansson P, Lindman B. Mixed solutions
polymers in semi-dilute regime: effect of salinity and temperature. of surfactant and hydrophobically modified polymer. Controlling
Prog Colloid Polym Sci 1996;102:26–31. viscosity with micellar size. J Phys Chem B 1998;102:7099–105.
[413] Maestro A, Gonzalez C, Gutierrez JM. Shear thinning and thixotropy [434] Kjoniksen AL, Nilsson S, Thuresson K, Lindman B, Nys-
of HMHEC and HEC water solutions. J Rheol 2002;46:1445–57. trom B. Effect of surfactant on dynamic and viscoelastic
[414] Winnik FM. Effect of temperature on aqueous-solutions of properties of aqueous solutions of hydrophobically modi-
pyrene-labeled (hydroxypropyl)cellulose. Macromolecules fied ethyl(hydroxyethyl)cellulose, with and without spacer.
1987;20:2745–50. Macromolecules 2000;33:877–86.

You might also like