Sei sulla pagina 1di 71

Progress in Polymer Science 36 (2011) 15581628

Contents lists available at ScienceDirect

Progress in Polymer Science


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

Polymers for enhanced oil recovery: A paradigm for structureproperty relationship in aqueous solution
D.A.Z. Wever a,b,1 , F. Picchioni a,2 , A.A. Broekhuis a,
a b

Department of Chemical Engineering - Product Technology, University of Groningen, Nijenborgh 4, 9747 AG Groningen, The Netherlands 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
Recent developments in the eld of water-soluble polymers aimed at enhancing the aqueous solution viscosity are reviewed. Classic and novel associating water-soluble polymers for enhanced oil recovery (EOR) applications are discussed along with their limitations. Particular emphasis is placed on the structureproperty correlations and the synthetic methods. The observed rheological properties are conceptually linked to the polymer chemical structure (1) and topology (2). In addition, the inuence of external parameters, e.g. temperature, pH, salt, and surfactant, on the rheological behavior is reviewed. Progress booked in deeper understanding of the structureproperty relationship is thoroughly discussed. Furthermore, a critical overview of the synthetic methods as well as of the solution properties of these polymers is provided. In this respect the inuence of internal (i.e. chemical structure) and external (vide supra) factors on these properties provide a conceptual 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.

Article history: Received 2 June 2010 Received in revised form 3 May 2011 Accepted 6 May 2011 Available online 27 May 2011 Keywords: Water soluble polymers Enhanced oil recovery (EOR) Hydrophobically modied polymers Polyacrylamide Associating polymers Solution rheology

Contents 1. 2. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1560 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 Recent developments in water-soluble thickening polymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1566 3.1. Hydrophobically modied polyacrylamide (HMPAM) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1567 3.1.1. Chemical structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1581 3.1.2. Rheological properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1583 3.2. Hydrophobically modied ethoxylated urethane (HEUR) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1589 3.2.1. Rheological properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1599 3.2.2. Effects of inherent parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1599

3.

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) 15581628

1559

4.

3.2.3. Effects of external parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1600 3.3. Hydrophobically modied 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 modied 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 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1619 Acknowledgement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1619 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1619

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

1560

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

bis(4-isocyanateocyclohexyl)methane hexamethylene di-isocyanate hydroxyethyl cellulose hydrophobically modied ethoxylated urethane HHM-HEC hydrophobicallyhydrophilically modied hydroxyethyl cellulose HMPAM hydrophobically modied polyacrylamide HM-CMC hydrophobically modied carboxymethyl cellulose HM-EHEC hydrophobically modied ethyl hydroxyethyl cellulose HM-HEC hydrophobically modied hydroxyethyl cellulose HM-HPC hydrophobically modied hydroxypropyl cellulose HM-polysaccharides hydrophobically modied polysaccharides HPAM partially hydrolyzed polyacrylamide HTAC hexadecyltrimethylammonium chloride 4-isopropylbenzyl chloride IPBC IPDI isophorone diisocyanate KDS potassium dodecyl sulfate lower critical solution temperature LCST MA maleic anhydride MAA methacrylic acid MAM methacrylamide MeEAM N-methyl-N-(4-ethyl-phenyl)acrylamide MO methyleneoxide MWD molecular weight distribution sodium acrylate NaA NaAMB sodium 3-acrylamido-3-methylbutanoate NaAMPS sodium 2-acrylamido-2-methylpropanesulfonate NaC10 S sodium decyl sulfate NaC14 S sodium tetradecyl sulfate dodecyl-di(ethyleneoxide)NaC12 (EO)2 S sodium sulfate NAEA 2-(1-naphthylacetyl)ethyl acrylate NAEAm 2-(1-naphthylacetamido)ethyl acrylamide NaMAMB sodium 3-methacrylamido-3-methylbutanoate NBAM N-benzylacrylamide NIPAM N-isopropylacrylamide NNDAM N,N-dimethyl acrylamide nanometer nm NMA [(1-naphthyl)methyl]acrylamide nuclear magnetic resonance NMR NP nonylphenol ethoxylates NPEAM N-phenethylacrylamide non-radiative energy transfer NRET OG n-octyl -d-glucopyranoside original oil in place OOIP OTG n-octyl -d-thioglucopyranoside poly(acrylic acid) PAA PAGE poly(alkylglycidyl ether) polyacrylamide PAM

H12 MDI HDI HEC HEUR

polydispersity index poly(ethylene oxide) pulse gradient spin echo N-phenylmethacrylamide parts per million pyrene residual resistance factor hydrophilically modied hydroxyethyl cellulose S-G HEUR step-growth hydrophobically modied ethoxylated urethane SANS small-angle neutron scattering SEC size exclusion chromatography SDS (NaC12 S) sodium dodecyl sulfate SMR surfactant to micelle ratio SOBS sodium octyl benzene sulfonate THF tetrahydrofuran TMSPMA 3-[tris(trimethylsilyoxy)sily]propyl methacrylate TTAB tetradecyltrimethyl ammonium bromide

PDI PEO PGSE PMAAM ppm Py RRF S-HEC

1. Introduction Enhanced oil recovery (EOR) is a challenging eld for different scientic disciplines. The importance of this eld is highlighted by the number of patents (mainly led by multinational companies) involving polymers for EOR. Nevertheless, the limited number of patents led in the last 10 years (less than 25) demonstrates the difculty of this research eld as well as the relative maturity in the scientic and technological concepts linked to relevant applications. Given the fact that the easily recoverable oil is running out and that much oil remains in the reservoir after conventional methods have been exhausted, the implementation of EOR is crucial to guarantee a continuing supply. In addition, alternative energy sources have not yet proved to be capable of meeting the world energy demand, so that a mixture of different sources, including oil, is required to meet the world energy demand in the near future. According to Thomas [1], approximately 7.0 1012 barrels of oil will remain in oilelds after conventional methods have been exhausted; this value constitutes also the target recovery for EOR. Water-soluble polymers for EOR applications have been successfully implemented, mainly in Chinese oilelds [2,3]. The purpose of the water-soluble polymers in this application is to enhance the rheological properties of the displacing uid. The oil production increases with the microscopic sweep of the reservoir and the displacement efciency of the oil [4]. Indeed, the use of water-soluble polymers improves the wateroil mobility ratio [4], and leads to enhanced oil recovery. However, given the harsh conditions present in most oil reservoirs, new problems and limitations arise with the use of water-soluble polymers. Besides positively 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) 15581628

1561

high temperatures (>70 C) and long injection times (at least 12 months) [4,5]. High salt concentrations reduce the thickening capability of most ionic water-soluble polymers while the presence of calcium leads to occulation [6]. New water-soluble polymers were successfully tested at higher temperatures [7,8]. Associative water-soluble polymers were tested and showed promising results compared to traditionally used polymers [9,10]. Several studies [1116] demonstrated that the oil is produced faster (compared to water ooding), but also more oil can be recovered. A mechanism based on the viscoelastic properties of the polymers has been proposed to explain the higher oil recovery. This mechanism is mainly supported by indirect evidence and mathematical models [1114]. Independently of the exact displacement mechanism and efciency, the use of water-soluble polymers for EOR still constitutes a challenging eld, at both industrial and academic levels. Moreover, the broad variety of polymers studied for EOR, at least at the academic level, clearly suggests this research eld as paradigmatic for watersoluble polymers in general. Indeed, fundamental scientic studies linked to EOR, particularly those involving the relationship between the polymer structure and the corresponding properties in aqueous solution (e.g. viscosity), have a conceptual character, thus providing a broad and general understanding, valid also for other application elds. The interaction between polymer chains in aqueous solution as well as the inuence of the solute structure on the corresponding viscosity are of paramount importance in EOR applications, but are studied and modeled through very general methods and theories. This general interest in structureproperty relationships is testied by the constantly increasing amount of scientic literature that is dedicated to this topic [1720], the last dedicated review (from 1990 [21]) being already outdated by the most recent ndings. As a consequence, the present review paper concentrates on different water-soluble polymers for EOR by discussing their (associative) behavior in water solution and outlining the most general concepts that can be learned from the corresponding scientic literature. Section 2 presents some background information on oil recovery methods, thus dening the subject in terms of polymeric systems reviewed in this work. In Section 3 the synthetic methods used for the different polymers are discussed, including the effect of the relationship between the process conditions employed and the corresponding polymer chemical structure. Subsequently, the chemical structures are linked to the corresponding rheological properties in water solution. The sensitivity 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 surfactant type and concentration. Relations between the polymer structure, i.e. chemical structure and topology, are proposed for the different polymers. Finally, in Section 4 we provide some general conclusions on the most recent, relevant and generally accepted concepts.

2. Currently used EOR polymers 2.1. Polyacrylamide (PAM) Polyacrylamide was the rst polymer used as thickening agent for aqueous solutions. The thickening capability (increase of the corresponding solution viscosity) of PAM resides mainly in its high molecular weight, which reaches relatively high values (>1 106 g/mol). In the general framework of EOR processes, PAM is mainly used as the reference model system for chemical modication. Many authors have reported different attempts to alter the chemical structure of PAM or to synthesize new acrylamide-based copolymers with improved properties, i.e. shear resistance, brine compatibility and temperature stability [2225]. The synthesis of the copolymer N,N-dimethyl acrylamide with Na-2-acrylamido-2methylpropanesulfonate (NNDAMNaAMPS) was reported by Sabhapondit et al. [22,23] and tested for its performance in EOR applications. The stability of the polymer at high temperature was demonstrated by ageing at 120 C for 1 month [22]. By using a sand pack, an improved performance in terms of EOR for the NNDAMNaAMPS copolymer [23] as compared to unmodied partially hydrolyzed polyacrylamide, HPAM, was demonstrated. In another example, Song et al. [24] reported the synthesis of starch-graft-poly(acrylamide-co-(2-acrylamido2-methylpropanesulfoacid)). The oil recovery rate of the subsequent polymer solution was higher compared to HPAM, and the novel polymer displayed better temperature and shear stability. These two examples already dene a common research theme in the general eld of water-soluble polymers for EOR. That is, a strategic approach involving the chemical modication of commercial polymers (in this case PAM) to tailor and improve the corresponding solution properties and eventually EOR performance. 2.2. Partially hydrolyzed polyacrylamide (HPAM) HPAM, by far the most used polymer in EOR applications, is a copolymer of PAM and PAA obtained by partial hydrolysis of PAM or by copolymerization of sodium acrylate with acrylamide [21]. 2.2.1. HPAM chemical structure The chemical structure of HPAM is provided in Fig. 1. In most cases the degree of hydrolysis of the acrylamide monomers is between 25 and 35% [4,26]. The fact that a relevant fraction of the monomeric units needs to be hydrolyzed (lower limit of 25%) is probably related to the

Fig. 1. Chemical structure of HPAM.

1562

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

formation of the corresponding salt. According to the general theory of polyelectrolyte solutions [27], the presence of electrostatic charges along a polymer backbone is responsible for prominent stretching (due to electric repulsion) of the polymeric chains in water and, eventually, results in a viscosity increase compared to the uncharged analogue. On the other hand, according to a study by Shupe [28] the degree of hydrolysis cannot be too high because the polymer solution will become too sensitive to salinity and hardness of the brine (electrolytes present in solution have a shielding effect on the electrostatic repulsion). Indeed, polyelectrolytes, i.e. polymers bearing charges, show signicantly different rheological behavior compared to their neutral analogues [2931]. The thickening capability of HPAM lies in its high molecular weight and also in the electrostatic repulsion between polymer coils and between polymeric segments in the same coil [4]. When polyelectrolytes are dissolved in water containing electrolytes (salts) a reduction in viscosity is observed [26,3234]. It has been demonstrated that the specic viscosity of HPAM solutions depends on the amount of salt present [35]. This effect is attributed to the shielding effect of the charges [4,33] leading in turn to a reduction in electrostatic repulsion and consequently to a less signicant expansion of the polymer coils in the solution. This results in a relatively lower hydrodynamic volume, which is synonymous with a lower viscosity [34]. A few decades ago, substitution of one or both hydrogens on the amide nitrogen with alkyl groups

has been presented as a solution to the salt sensitivity of the HPAM [36,37], although the exact reasons for this behavior have not been fully elucidated. The addition of monovalent NaCl leads to a reduction in the level of aggregation. However, at higher ionic strengths (higher salt concentration) the addition of NaCl leads to macroscopic occulation [38]. It has also been demonstrated that multivalent cations can form polyionmetal complexes which affect the viscosity of the resulting solution [3941]. A study by Peng and Wu [39] investigated the dependence of the self-complexation of HPAM on the Ca2+ concentration and the degree of hydrolysis of HPAM. They demonstrated that depending on the Ca2+ concentration intrachain and interchain complexations take place (Fig. 2) [39]. Besides the salt dependency, other factors inuencing the viscosity of HPAM solutions are the degree of hydrolysis, solution temperature, molecular weight and solvent quality [35]. Pressure also affects the viscosity of HPAM solutions. According to a study by Cook et al. [42] the increase in the viscosity of the HPAM solutions cannot solely be accounted for by the increase in viscosity of the solvent. The intrinsic viscosity and the radius of gyration are both invariant with pressure, albeit with a 10% experimental uncertainty [42]. In principle, the dimensions of the polymer coils do not change while the solvent volume decreases. Therefore the volume fraction of the polymer 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) 15581628

1563

Fig. 3. Schematic presentation of behavior of HPAM coils in shear ow. Reproduced with permission from [46] 1995, ACS Publishing.

viscosity [42]. Another parameter that affects the solution viscosity of the polymer solution is shear [43]. Under high shear the HPAM polymer chains are reduced in size due to chain scission, i.e. fragmentation [44]. This leads to a reduction in the solution viscosity. 2.2.2. Rheological properties HPAM is preferred in EOR applications since it can tolerate the high mechanical forces present during the ooding of a reservoir. In addition, HPAM is a low cost polymer and is resistant to bacterial attack [4]. Although the HPAM solutions display pseudoplastic behavior [4,26,32,45,46] (shear thinning) in simple viscometers, it has been demonstrated that these solutions show pseudodilatant [47,48] characteristics (shear thickening) in porous media as well as in viscometers at relatively high shear rates. Research has demonstrated the presence of a critical shear rate at which the shear thickening behavior arises in viscometers [32,33,45,46,49,50]. This critical shear rate depends on the degree of hydrolysis of the HPAM, the solution concentration, the temperature, the quality of the solvent and also on the molecular weight of the polymer [33,45]. An increase in the degree of hydrolysis leads to an onset of shear thickening at lower shear rates [45]. By decreasing the average molecular weight, an increase in the polymer concentration results in a higher critical shear rate [45,46]. The aforementioned shear thinning of HPAM solutions below a critical shear rate arises due to uncoiling of polymer chains and the dissociation of entanglements between separate polymer coils [4]. Stiffening of the polymer backbone has been suggested as a possible approach to control the dependency of HPAM polymer solutions on the shear [51]. A stiff polymer will have a lower mobility and there-

fore the entanglements, related to the solution viscosity, will be conserved as the shear increases. The shear thickening behavior has been attributed to changes in the molecular conformation involving the formation of additional links between two chains [49]. The shear thickening behavior is observed both in laboratory rheometers [45] (in pure water and aqueous salt solutions) and in porous media. According to several studies the shear thickening behavior in porous media arises due to coil-stretched transitions of the polymer chains. The structure of a porous medium can be seen as alternating wide openings and conned throats through which the polymer coils have to navigate. In the wide openings the polymer chains attain a coil structure. When these coils then have to pass through a narrow throat the polymer coils are forced to deform and stretch (elongational strain [32,48,52]) in order to pass. This successive contraction and expansion of the polymer coils leads to pseudodilatant behavior of the polymer solutions [48,53,54]. This conformational change of the macromolecules is reversible since it is commonly explained by formation at macromolecular level of reversible interactions like hydrogen bonding. Indeed, it is believed that hydrogen bonding arises for HPAM solutions between the carboxylic functionalities [55]. However, this is contested due to conicting data [49,56] on similar polymeric solutions (e.g. dextran solution). Instead, aggregation of hydrophobic bonds has been proposed [55], albeit in polymethacrylic acid, but this has not been conrmed [57]. Hu et al. [46] proposed a schematic presentation depicting the essential behavior of HPAM solutions in shear ow (see Fig. 3). Another behavior that has been identied for HPAM 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) 15581628

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

thixotropic (rheopectic) property, i.e. an increase in viscosity with shear-time at a constant shear rate [32,5861]. Researchers have identied two different types of rheopectic behavior for HPAM solutions (Fig. 4), type I and type II [58]. The type I effect is observed at low shearing and consists in a slow viscosity increase with shear-time up to an asymptotic value. The type II effect is seen at high shear rates and is displayed as a steep viscosity increase after a given shear-time, followed by pronounced viscosity oscillation [58]. 2.2.2.1. Effect of inherent parameters on rheology. For HPAM the degree of hydrolysis has a signicant impact on the rheological properties of the subsequent solution. If the hydrolysis degree of HPAM is too large, insolubility problems can arise. When too small, a large dependence of the solution viscosity on electrolyte presence is observed. Lewandowska [45] investigated the effect of the degree of hydrolysis on the solution viscosity in a NaCl solution. If the degree of hydrolysis is increased the zero shear rate viscosity and the critical shear rate for the onset of shear thickening are both reduced. The shear rate region where the shear thinning behavior is observed is reduced with the increase in the degree of hydrolysis [45,46,62]. Another inherent parameter of HPAM is its molecular weight. Increase of the molecular weight will lead to a more pronounced shear thinning behavior. The critical shear rate for the onset of shear thickening is also affected by the molecular weight in that it is increased with an increase in the molecular weight [45]. 2.2.2.2. Effect of external parameters on rheology. The inuence of salt (NaCl) addition on the rheological behavior of HPAM solutions has also been extensively studied. It was found that, below the critical shear rate, adding salt reduced the extent of shear thinning while above the critical shear rate the amplitude of the shear thickening is increased [32]. The zero shear rate solution viscosity is also reduced as the concentration of salt increases [4]. Adding salt to the solvent will reduce the extent of the rheopectic behavior of the polymer solution [59]. Looking more closely to the effect of salt, addition of mono-valent cations (e.g. NaCl) was found to increase the onset (i.e. the critical shear rate value) of rheopectic behavior type I while for type II no changes were observed [58]. When using multivalent cations (CaCl2 and AlCl3 ) the effect seen with

mono-valent cations is amplied, i.e. the effect is seen at lower cation concentrations [61]. This is due to the higher screening capability of multivalent cations [61]. In addition a reduction of the degree of hydrolysis leads to a less prominent rheopectic behavior of type I [60]. Besides salts, also surfactants are able to interact with polymer chains in solutions and thus display a relevant inuence on the corresponding rheological behavior. Xin et al. [63] investigated the interaction between the surfactant sodium oleate and HPAM and found that the viscosity of the polymersurfactant aqueous solution depends on the surfactant concentration. At low surfactant concentration an enhancement of the viscosity is observed due to interpolymer cross-linking of the surfactant and HPAM. At high surfactant concentration the repulsion between the micellar aggregates attached to the polymer increases and this leads to a decrease in the hydrodynamic volume and thus a decrease in the solution viscosity. The authors [63] have proposed a model depicting the interactions between HPAM and sodium oleate (a surfactant) at low and high concentration of the latter (Fig. 5). Interchain cross-linking arises due to hydrogen bonding. At high surfactant concentration repulsion between the surfactant molecules dominates and this leads to the collapse of the network between the surfactant and HPAM, which in turn results in a reduction of the viscosity. Methemitis et al. [64] investigated the interactions between SDS and HPAM and concluded that the effect of the surfactant on the rheological properties of HPAM depends on the pH of the solution and the presence of electrolyte. When no effort is devoted at altering the pH of the solution, a reduction in the solution viscosity is observed up to the CMC of the surfactant either in water or in salty water. Above the CMC the solution viscosity of the ternary, system, i.e. waterpolymersurfactant, remains relatively constant with further addition of SDS in the salty water. In pure water the solution viscosity still decreases with further addition of SDS. If the pH of the solution is reduced (to a pH of 2.5) the solution viscosity decreases until the CMC of the surfactant is reached, after which a signicant increase is observed with further addition of SDS. The authors have hypothesized that at low pH xation of some protons onto the surfactant micelles occurs [64]. This changes the ionization equilibrium of the carboxylic groups leading to an increase in the surface charge density, which corresponds to an increase in solution viscosity.

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

1565

Fig. 6. Chemical structure of xanthan gum.

2.3. Xanthan gum Xanthan gum is a polysaccharide, which is produced through fermentation of glucose or fructose by different bacteria [65]. The most efcient xanthan gum producer is the Xanthomonas campestris bacterium [65,66]. The chemical structure of xanthan gum (Fig. 6) displays the presence of two glucose units, two mannose units and one glucuronic acid unit with a molar ratio of 2.82.02.0 [67]. The backbone of xanthan gum is similar to cellulose. The side chains of the polymer contain charged moieties, i.e. acetate and pyruvate groups, and the polymer is thus a polyelectrolyte. However the classic polyelectrolyte behavior according to which the solution viscosity decreases with the addition of salt is not displayed in this case. The thickening capability of xanthan gum lies in its high molecular weight, which ranges from 2 to 50 106 g/mol [67,68] and in the rigidity of the polymer chains. It has been demonstrated that upon addition of salt (mono or divalent) the xanthan gum chains undergo a cooperative conformational transition from a disordered conformation to an ordered and more rigid structure [6972] (Fig. 7). The temperature and the ionic strength (the amount of electrolyte) of electrolyte, of the solution are triggers for the conformational transition. When testing at low shear, the rheology of the polymer solution is dependent on the conformation with the disordered conformation displaying higher solution viscosities [73]. Polymeric solu-

tions employing xanthan gum display high viscosity at low shear rates [74] and thus the disordered conformation predominates at low shear rates. At high shear rates both conformations display similar rheological behaviors [73]. In addition, pseudoplastic behavior is observed for the polymer solutions [75]. Unlike HPAM, xanthan gum displays good resistance to high temperatures. It was demonstrated that the solution viscosity of a polymeric solution employing a commercial xanthan gum remained relatively constant for more than 2 years at 80 C [76]. Loss of solution viscosity occurs at temperature above 100 C. Several studies [7780] have investigated the temperature dependence of the apparent viscosity of xanthan gum solutions. In order to display resistance to temperatures up to 90 C, the conventional understanding for xanthan gum solutions is that the ionic strength of the solution has to be relatively high. Another positive property of xanthan gum is its ability to withstand high shear forces. Unlike HPAM the solution viscosity does not decrease at relatively high shear stresses [47]. Especially the ordered structure, i.e. in the presence of salt, can withstand high shear forces [73] (up to a shear rate of 5000s1 ). A disadvantage of xanthan gum is its susceptibility to bacterial degradation. It has been demonstrated that salt tolerant aerobic and anaerobic microorganisms can degrade the xanthan gum chains which leads to the loss in solution viscosity [8184]. Biocides are used to suppress the growth of the xanthan gum degrading microorganisms. In most cases formaldehyde is the most efcient biocide

Fig. 7. Conformational transition of xanthan gum.

1566

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

[83,84]. However the use of biocides to protect the xanthan gum renders the low environmental impact of the polymer obsolete. Combinations of xanthan gum with surfactants have also been studied. It has been demonstrated that the combination can be benecial. According to Taugbol et al. [85] more than 50% of the residual oil (after a waterood) can be recovered using xanthan gum and an alkyl propoxy-ethoxy sulfate (C1215 (PO)4 -(EO)2 -OSO3 Na+ ) as the surfactant. The recovery of the residual oil using only a surfactant solution was lower. However it has been demonstrated that the combination of xanthan gum and dodecyl-o-xylene sulfonate recovered less of the residual oil compared to the case where only a surfactant solution was used [86]. According to the authors a possible explanation is the formation of large micellar aggregates, which have a negative effect on the ow performance of the surfactant through the porous media.

3. Recent developments in water-soluble thickening polymers The limited number of available commercial polymers currently employed in EOR has been the subject of recent developments aimed at improving their performance. Indeed, an alternative concept has been studied in the last four decades, and involves the association between hydrophobic groups that are incorporated in the backbone of the polymers [87]. Through these associations a higher thickening capability can be achieved compared to the traditional polymers [87]. Several different types of associating polymers have been studied. These include the hydrophobically modied polyacrylamide (HMPAM) [88], ethoxylated urethane (HEUR) [89], hydroxyethylcellulose (HMHEC) [90] and alkali-swellable emulsion (HASE) [91]. Also combinations of associative polymers with surfactants have been developed for EOR [92]. It has been demonstrated that the addition of small amounts of surfactants

can increase the viscosity of the aqueous solution containing hydrophobically modied polymers signicantly [90]. Other polymers that posses interesting properties, such as high molecular weight and intrinsic viscosity, have been developed for EOR and are known as rigid rod water-soluble polymers [93]. One study compared hydrophobically modied polyacrylamide (HMPAM) with polyacrylamide (PAM) in a simple core ood test and demonstrated that the residual resistance factor (RRF) after the polymer ood is much higher for the HMPAM compared to PAM [94]. All these modication strategies, together with new kinds of water-soluble systems, have been extensively reported in the literature and will be discussed in the next paragraph. As mentioned earlier, a relatively new class of watersoluble polymers is the one constituted by hydrophobically associative polymers [87]. The rst hydrophobically associative polymers were synthesized almost fty years ago [95,96], albeit for a different purpose than EOR. Indeed, the research on these types of polymers has been primarily fueled by the coating industry [87], where improvement in the rheology of the coating systems was required. During the 1980s when the oil crisis hit, a lot of research was performed on EOR. From the many patents [97102] that have been led during those years it is evident that this accelerated the development of hydrophobically associative polymers for use in EOR applications. Hydrophobically associative polymers contain, in most cases, a small number of hydrophobic groups, i.e. 818 carbon atoms moieties [103106], distributed along the main backbone [20,107,108]. These hydrophobic groups can be distributed randomly or block-like [88,91,103,107,109120], and coupled at one or both ends [104,121131]. Above a given polymer concentration (dependent on the molecular structure) the hydrophobic groups associate, when the polymer is dissolved in water, to form hydrophobic micro-domains (intra or intermolecular liaisons) [88,89,104,105,107,108,132135]. 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) 15581628

1567

in turn yields a polymer with a much better thickening (higher viscosity [107]) capability compared to its nonassociative analogue [88]. Depending on the concentration, intra- or intermolecular associations are formed, which are schematically illustrated in Fig. 8. When the hydrophobic elements are distributed in a block-like fashion along the backbone of a water-soluble copolymer, the intramolecular associations are stronger compared to randomly or discretely distributed hydrophobic groups [104,121]. The temperature dependence of the solution viscosity is an interesting property of hydrophobically modied polymers for EOR applications. It has long been accepted that increasing the temperature of the polymer solution will lead to a reduction in viscosity [103,133,136140]. This since an increase in temperature implies a decrease of the association strength of the hydrophobes. Increasing the temperature of the solution leads to a reduction of the solvent viscosity and hence an increase in the mobility of the polymer chains while the solubility of the polymer will increase with temperature. However, many different aqueous systems have been demonstrated to display an increase in viscosity upon increasing the temperature [141154]. Indeed, a temperature increase will decrease the solubility of one of the components (lower critical solution temperature [LCST]-groups) of the polymers. These less soluble components will self-aggregate with the hydrophobic groups of the polymers, which leads to an increase in viscosity [119]. Several researchers have proposed a concept for thermo-associative polymers based on the switch, i.e. the transition between low and high temperature, of the polymers characterized by a lower critical solution temperature [140,151,152]. The concept involves a highly water-soluble polymer containing blocks or side chains of LCST groups. Upon heating of the polymer solution, these LCST groups will segregate. A schematic presentation of this behavior has been presented by Hourdet and coworkers [151] and is depicted in Fig. 9. Above the critical overlap chain concentration this transition will lead to an increase in the viscosity of the solution through intermolecular associations. Fundamental research on different polymers, in binary (polymerwater) and ternary (polymerwatersurfactant) systems, has been performed using different techniques which include 13 C NMR [155158] (solution or solidstate), 1 H NMR [109], 23 Na NMR [159,160], 19 F NMR [161], NMR self-diffusion [126,162165] potentiometry [166168], static and dynamic laser light scattering [128,130,131,146,163,167172], UV-spectroscopy for polymers bearing chromophores [88,110,111,173181], small-angle neutron scattering (SANS) [182], non-radiative energy transfer (NRET) studies [178,183185], size exclusion chromatography (SEC) [170,186189] and surface tension [131,133,162,163,165]. Several different associative hydrophobically modied polymers have been developed which include polyacrylamides (HMPAM), ethoxylated urethanes (HEUR), alkali swellable emulsions (HASE), and polysaccharides (HMpolysaccharides). Their synthesis, rheological behavior and adsorption on surfaces will be discussed in the following sections.

3.1. Hydrophobically modied polyacrylamide (HMPAM) HMPAM constitutes the most popular basic structure for the synthesis of new water-soluble polymers for EOR. Indeed, as given in Table 1, many types have been published. There are different methods of synthesizing HMPAM such as micellar [88,113,135], homogeneous [88,248,249] and heterogeneous [88] copolymerization. Polyacrylamide is usually prepared via a free radical polymerization in aqueous solution [88,250]. However, as is evident from the name, HMPAM cannot be synthesized using this technique as the hydrophobic monomer is not soluble in water. In order to disperse the hydrophobic monomer, it is dissolved using a co-solvent (homogeneous copolymerization) or a surfactant (micellar copolymerization) or dispersed without any additives (heterogeneous copolymerization) [88,135]. Among these different methods, micellar copolymerization has been mostly studied. It generally involves the use of a hydrophobic monomer (soluble in micelles stabilized by a surfactant) and a hydrophilic one which is soluble in the water phase. The rst reports on micellar copolymerization appeared simultaneously some 25 years ago [251256]. A schematic presentation of this polymerization technique is presented in Fig. 10. When synthesizing polymer 17 (Table 1) Chang and McCormick [220] noted that the use of micellar copolymerization leads to a block-like distribution of the hydrophobes whereas solution copolymerization leads to a random distribution. Both polymers exhibit completely different rheological properties. The same behavior was observed when synthesizing polymer 29 (Table 1) using both (micellar and solution) techniques [110,111]. The most popular surfactants in micellar copolymerization are sodium dodecyl sulfate (SDS) [92,103,104,109, 112,121,134,135,138,173179,203,229,230,241,257260] and cationic hexadecyltrimethylammonium bromide (CTAB) [113,133,220,238]. Many different hydrophobic monomers have been used, such as acrylate or methacrylate-derivatives, alkyl groups with varying number of carbons and different topologies [99,100,103,229,261264], aryl or alkylaryl functionalities [88,110,111,113,173177,230,260,265268], uorocarbon containing agents [132,241,242,269271] and zwitterionic groups [196199,201203,234,272274]. By incorporating water-soluble spacers between the hydrophilic backbone and the hydrophobic group an enhancement of the viscosity can be achieved compared to systems without spacers [199,241]. According to Hwang and Hogen-Esch [241] the formation of hydrophobic micro-domains is effectively promoted by increasing the lengths of the spacers. Parameters affecting the properties of the polymers prepared by micellar copolymerization are the type and concentration of the hydrophobic and hydrophilic monomer [109,174,175,258], the molar ratio between the two monomers, the content and type of surfactant [174,275], the content and type of initiator and the temperature of the reaction [135]. Another parameter that has been identied is the molar ratio between the hydrophobic monomer and surfactant, i.e. the number of hydrophobic

1568

Table 1 Structure of different water-soluble polymers, HMPAM. Polymer Polyelectrolyte Structure References

Copolymer of acrylamide (AM) and sodium 2-acrylamido-2-methylpropanesulfonate (NaAMPS, R = 2) or sodium 3-acrylamido-3-methylbutanoate (NaAMB, R = 1) or (2-acrylamido-2-methylpropyl)dimethylammonium chloride (AMPDAC, R = 3)

[190193]

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

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

[196]

Polyelectrolyte, zwitterionic monomer

Copolymer of acrylamide (AM) with 3-(2-acrylamido-2-methylpropane-dimethylammonio)-1-propanesulfonate (AMPDAPS)

[197]

Table 1 (Continued) Polymer Polyelectrolyte, zwitterionic monomer Structure References

Copolymer of AM with 2-(2-acrylamido-2-methylpropyldimethylammonio) ethanoate (AMPDAE, N = 1) or AMPDAB, N = 3 or 6-(2-acrylamido-2-methylpropyldimethylammonio) hexanoate (AMPDAH, N = 5)

[198200]

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

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 (DADPMA)

[203]

1569

1570

Polyelectrolyte, zwitterionic backbone

Terpolymer of AM and AMPDAC (R1 = H) with NaAMPS (R2 = 1) or AMPTAC (R1 = CH3 ) with NaAMPS (R2 = 1) or AMPTAC (R1 = CH3 ) with NaAMB (R2 = 2)

[204207]

Polyelectrolyte, zwitterionic backbone D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

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

[208]

Polyelectrolyte, zwitterionic backbone

Copolymer of 2-acrylamido-2-methylpropanesulfonate (NaAMPS) with (2-acrylamido-2-methylpropyl)-dimethylammonium chloride (AMPDAC, R = H) or [2-(acrylamido)-2-methylpropyl]trimethylammonium chloride (AMPTAC, R = CH3 )

[209211]

Polyelectrolyte, amphiphilic

Copolymer of AM (R1 = NH2 ) and sulfonate containing monomer (R2 = 1)

[115,212,213]

Table 1 (Continued) Polymer 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 Structure References

Copolymer of NaA and

[156,214216]

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

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) 15581628

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

[115,220,221]

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

Table 1 (Continued) Polymer Polyelectrolyte, amphiphilic Structure References

Copolymers of N-isopropylacrylamide (NIPAM, R2 = 1) or AA (R2 = OH) with 2-(N-ethylperuorosulfoamido)ethyl acrylate (FX-13), R1 = H or 2-(N-ethylperuoro-octane/sulfoamido)ethyl methacrylate (FX-14), R1 = CH3

[221225]

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

Polyelectrolyte, zwitterionic monomer and amphiphilic

Copolymer of 3-(N,N-diallyl-N-methylammonio)propane-sulfonate (DAMAPS, R2 = 1) with N,N-diallyl-N,N-dimethylammonium chloride (DADMAC, R1 = CH3 ) or N,N-diallyl-N-methylamine chloride (DAMA, R1 = H)

[226228]

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) 15581628

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

[231234]

Table 1 (Continued) Polymer 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 Structure References

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

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 (E AM, R = 5) or N-methyl-N-(4-ethyl-phenyl)-acrylamide (MeE AM, R = 6) or BPAM, R = 7 Amphiphilic

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

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 Amphiphilic Structure References

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

Copolymer of AM and

[132,241243]

1,1-Dihydroperuorobutyl acrylate (2a) or 1,1-Dihydroperuorooctyl acrylate (2b) or 2-(N-ethylperuorooctanesulfonamido)ethyl acrylate, FOSA (3a) or 2-(N-ethylperuorooctanesulfonamido)ethyl methacrylate, FOSM (3b) or 1,1-Dihydroperuorooctylmono-(ethyleneoxy) acrylate (4a) or 1,1-Dihydroperuorooctylbis-(ethyleneoxy) acrylate (4b) or 1,1-Dihydroperuorooctyltris-(ethyleneoxy) acrylate (4c) or Dodecyl acrylate (5a) or Dodecylmono-(ethyleneoxy) acrylate (5b) or Dodecylbis-(ethyleneoxy) acrylate (5c) or Dodecyltris-(ethyleneoxy) acrylate (5d) or Poly(propylene oxide) methacrylate (6)

1577

1578

Amphiphilic

1. Copolymer of AM with

[104,244,245]

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

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

[246,247]

Model polymer

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

[110,111]

Table 1 (Continued) Polymer Model polymer Structure References

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) 15581628

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) 15581628

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 was rst thought that the solubilization of the hydrophobic monomer in a surfactant micelle, would cause an increase in the incorporation rate of the hydrophobic monomer (into the polymer backbone) [109,174]. A NH increase will change the randomness, i.e. the regularity of the number of hydrophobic units incorporated in the polymer as a function of time [109,174]. However two studies [109,277] demonstrated that by using a disubstituted acrylamide as the hydrophobic monomer no drift in copolymer composition was observed. Therefore, according to the study [109], the previously thought dissolution of the hydrophobic monomer in the surfactant micelle no longer holds true. The observed behavior can be then attributed to the difference in polarity between the bulk and micellar phase, which modies the reactivity of hydrophobes [109]. Micelle copolymerization can also be used to synthesize a more randomly distributed copolymer. This can be achieved by using a molar ratio where approximately one hydrophobic unit is solubilized in one micelle [138,275]. Micellar copolymerization remains the most used method for synthesizing HMPAM [135]. Nevertheless,

another method that is closely related to the micellar copolymerization technique seems promising, but it involves using a micelle-forming polymerizable surfactant [114,115,178,179,220,278280]. Although identifying a correct polymerizable surfactant for the desired molecular structure is difcult, the technique offers the advantage that purication (i.e. removal of surfactants as is the case with micellar copolymerization) of the reaction mixture no longer would be required. Yet another method, which recently has been developed, is template copolymerization. The structure of the copolymer is dened by the template that is used. A schematic presentation illustrating the template copolymerization is given in Fig. 11. The advantage of this technique with respect to the other ones is that the block-like distribution of hydrophobic groups is better controlled. According to several different studies [281285] a longer sequence distribution of hydrophobic groups can be achieved with this technique. The molecular weights of the polymers inuence their behavior in solutions. The molecular weight of HMPAM remains difcult to determine since intramolecular associations 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) 15581628

1581

Nevertheless, several different methods have been devised. These include using the MarkHouwinkSakurada relation for polymers dissolved in water [110,113,114] or unmodied polymers in the exact same conditions as their modied analogues [103,229,230,286,287]. However, all these methods introduce errors in the molecular weight determination and it would be better to nd a suitable solvent in which the polymer is molecularly dispersed [135]. Several authors reported on formamide as a suitable solvent for the molecular weight determination and measured it by using light scattering experiments [88,109,174]. Another popular strategy to increase the thickening ability of polyacrylamide polymers is related to the presence of electrical charges along the backbone. Indeed, many different polyelectrolytes based on acrylamide have been synthesized. These polyelectrolytes can be divided into polyampholytes which include zwitterionic backbone polymers 911, having positive and negative charges in separate building blocks in the backbone, and zwitterionic monomers 38 having positive and negative charges in the same building block, and traditional polyelectrolytes 1 and 2. The charges can be induced by controlling the pH of the subsequent polymer solution when, for example, AA is used as a co-monomer. In all cases it is reported that the solution viscosity is a function of the chemical structure and the solution characteristics (i.e. ionic strength). From the discussion on the polymers 111 one can conclude that there are many different advantages of incorporating electronic charges into water-soluble polymers. One such advantage is the ability to control the interaction between the polymer chains by altering the pH or ionic strength of the polymer solution. Another is the solubility in water, which makes the corresponding solutions less sensitive, in terms of viscosity, to the external temperature. On the other hand if water-soluble polymers are required whose rheological properties are independent of the pH, polyelectrolytes bearing only one type of charge can be used [201]. For applications where concentrated salt solutions are used (such as EOR), polyampholytes are suitable. They have been demonstrated to display an enhancement in the solution viscosity upon addition of low molecular weight electrolytes [197,201,204,209211,288,289]. This behavior has been attributed to the shielding of intramolecular Coulombic attraction rather than the intermolecular interactions. With careful molecular design water-soluble polymers containing electronic charges can be synthesized exhibiting the required rheological behavior. Peiffer and Lundberg [289] reported that in order to control the physical properties of the zwitterionic polymers it is crucial to separate the oppositely charged monomers using neutral ones. Derivatives of polyelectrolytes have been investigated with amphiphilic (hydrophobic) moieties incorporated along with charges, either positive or negative. These acrylamide based amphiphilic polyelectrolytes, 1217, have been synthesized by many different research groups in search for better thickening polymers in different applications. Polymers where both charges are incorporated along with hydrophobic groups have also been studied. These polymers can be classied as zwitterionic amphiphilics,

1921. The properties of the subsequent polymer solutions can be tailored by careful design of the polymer structure and composition. The polymers can be designed to be salttolerant or responsive to changes in the salt concentration or ionic strength. Amphiphilics, 2328, without electronic charges have been studied extensively for their salt-tolerance. Addition of salt will not affect the viscosity of the polymer solution since its thickening capability arises from hydrophobic associations and not from electronic interactions. The polymers 2933 have been synthesized in order to enable uorescence studies of hydrophobically associating polymers. This technique allows studying the associating behavior of these polymers through their photo-physical behavior in response to changes in the system. It has been demonstrated that polymers of acrylamide containing pyrene functionalities exhibit increased in excited-state dimer (excimer) formation and viscosity [110,111,179]. Variation in type of polymers causes different behavior of the subsequent polymeric solutions. However there are other parameters that also affect the structure and associations of the polymer solution: the chemical structure, the synthesis method, the temperature, the type and concentration of salt and the pH (ionic strength). The following discussion summarizes the effect of these parameters. 3.1.1. Chemical structure Generally speaking the rst macroscopic effect of the polymer chemical structure is observed from the corresponding solubility in water. Indeed, the solubility of polymer 3 is dependent on the level of AMPDAPS monomer incorporated [196]. This has a clear inuence on the viscosity of the solution [196] as observed for polymer 5 as function of the AMPDAE and AMPDAH monomers incorporation level [199,200]. The same considerations hold for polymer 6 as function of the monomer composition (in this case AMPDAPS and AA) [201] as well as polymer 8 as function of the charged monomer (DADPMA) intake [203]. The same can be said for polymer 19 where phase separation is observed at low incorporation levels of the DAMAPS monomer. These trends are easily understandable on the basis of simple considerations regarding the overall polarity of the polymeric chains. However, in contrast to this, polymer (8) is no longer soluble in water when it contains more than 1 mol% of DADPMA. In a similar example, McCormick and Johnson [208] observed hydrogel formation when the incorporation level of charged groups surpassed 1 mol% in polymer 10. Similar to polymer 8, polymer (10) forms hydrogels and is no longer water-soluble above 1 mol% incorporation of the charged groups. The carboxylate groups lead to strong ionic interactions which prevents the dissolution of polymer 10 [208]. It seems thus that the nature of the charge (e.g. carboxylate versus, sulfonate anion in Table 1) exert a clear inuence on the association behavior and ultimately on the solubility of the polymer. This phenomenon is never observed in the case of a zwitterionic backbone. Indeed, at low incorporation of either monomer, AMPDAC or NaAMPS, in the corresponding copolymer (polymer series 11) classic polyelectrolyte behavior is exhibited by the solution. However, polymeric solution employing polymers containing equimolar

1582

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

amounts of the AMPDAC and NaAMPS monomers display high salt tolerance [209]. A similar behavior is observed when, instead of AMPDAC, AMPTAC is used as the second monomer [211]. Similar to polymer 11, polymer 19 exhibits classic polyelectrolyte behavior when the incorporation rate of DAMAPS is below 40 mol% [226]. However, in dilute solutions of polymer 19 (both DADMAC-1 or 2 copolymer) primarily intra-molecular hydrophobic associations are present [228]. The presence of hydrophobic groups along the polymer backbone results generally in the formation of microdomains. Important parameters for association behavior of the polymers are the distribution of the hydrophobic groups and their hydrophobicity. As mentioned before, a block-like distribution of the hydrophobic groups will lead to stronger associations compared to a random distribution. The surfactant to micelle ratio (SMR) during polymerization affects the distribution of the hydrophobic groups. At low SMR the block-like distribution is favored while at high SMR a random distribution is preferred [175]. Polymers 13, 14 (synthesized using a high SMR) and 17 (using solution polymerization) are classic examples of low level association due to random distribution of the hydrophobic groups [156,175,220] while polymers 14 (synthesized with a low SMR), 17 (using micellar polymerization) and 30 are classic examples of a block-like distribution [111,175,220]. Although the SMR has been identied as an important parameter for the observed association behavior, its effect on the composition and molecular weight of the polymer is minimal as demonstrated with polymer 31 [177]. The presence of hydrophobic groups suppresses the solubility of the polymer. Following this, it is easy to understand that increasing the fraction of the hydrophobic groups above a certain percentage will lead to solubility issues, i.e. the polymer is no longer water-soluble. This has been demonstrated with polymer 24 where a water insoluble polymer was obtained using DiC8 AM as the co-monomer above an incorporation rate of 1.2 mol% [237]. Increasing the hydrophobicity of the hydrophobic groups will impart better thickening capability of the polymer as demonstrated with the polymers 12, 15, 17, 18, 23, 24 and 27 (2a, 2b, 3ac and 4ad). It has been demonstrated that only a fraction of the hydrophobic groups contribute to the formation of micro-domains which is affected by the hydrophobicity of the groups and the incorporation rate [290]. The hydrophobicity of the groups can be increased by increasing the length of the groups [103,220,237], using twin-tailed hydrophobes instead of single tailed groups [185,218] or using uorocarbons instead of hydrocarbons [212,223,242,269]. The critical polymer concentration at which hydrophobic associations (critical association concentration, CAC) arise is reduced compared to classic associating polymers [218]. These trends can be explained by the stronger interactions between the hydrophobic groups due to their increased hydrophobicity. The length of the hydrophobic groups is important for association where a short hydrophobe will not lead to association as demonstrated with polymer 24 using DiC3 AM as the comonomer [237]. Increasing its length, by using DiC8 AM as

the co-monomer, does impart associations [237]. However, in contrast to this, increasing the hydrophobicity too much will lead to insolubility of the polymers as observed with polymer 15 using longer twin-tailed groups than DiC12 AM [218]. The presence of spacers, i.e. small chains linking the hydrophobic groups to the backbone, affects the association behavior of the polymers markedly. Increasing the length of the spacers allows for easier movement of the hydrophobic group in solution which in theory should lead to easier formation of micro-domains and thus intermolecular association is favored. This is demonstrated with polymer 13 [216] (DemMA series). The association behavior of the polymer is not affected by the type of co-monomer that is used as long as the length of the spacer is large enough (in this case a length of 25 EO moieties). At intermediate lengths (2 or 6 EO moieties) this is no longer true as demonstrated with the same polymer backbone [216]. With NaA as co-monomer the network formation was favorable at shorter EO spacers while just the opposite behavior is observed with NaAMPS as the co-monomer. The authors [216] attribute this difference to the stronger tendency of intermolecular association for the NaA containing copolymer. The presence of a charged group further away from the polymer backbone will disrupt the hydrophobic associations to a greater extent than when the charged group is closer to the polymer backbone [229]. The NaAMPS co-monomer places the charged group further away from the backbone thus leading to a stronger disruption of the hydrophobic group association at similar length EO spacers. This behavior has been identied for the aforementioned polymer 13 and for polymer 20. In addition, charge density of the polymer affects the association behavior. Using low charge density co-monomers (carboxylate anions) will display stronger association above the CAC due to less interference with the hydrophobic associations, while sulfonate anions (higher charge density than carboxylate anions) will display the opposite behavior. Polymers bearing high charge density groups are more susceptible to screening effects in the presence of low molecular weight electrolytes and will therefore be less responsive at high electrolyte concentration or low pH. The polymer series 21 is a classic example of this [230]. When using 3 co-monomers the incorporation level of charged species will also affect the association behavior of the polymer. At low incorporation levels of charged species no suppression of the hydrophobic associations is expected. However, at high incorporation levels the interference with the associations is expected to be signicant. This was demonstrated using the polymer series 14. The polymers with 9 or 21 mol% AA display intermolecular hydrophobic associations at low pH (<5), and in the presence of NaCl. However, polymers with 37 mol% acrylic acid do not display this behavior, not even in de-ionized water and low pH [176]. A peculiar behavior, in the context of EOR, is the temperature stability of polymers. It has been demonstrated that by using phenyl containing moieties as co-monomers increases the stability of polymer 25 (PMAAM) against high temperatures [239].

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

1583

Fig. 12. Schematic model structure of HMPAM. Reproduced with permission from [291] 1991, ACS Publishing.

3.1.2. Rheological properties Hydrophobic interactions confer HMPAM interesting rheological and solution properties. Leibler et al. [291] proposed the sticky reptation model, which explains the complex viscoelastic behavior [112,257,258,292] of HMPAM polymers. The authors proposed a schematic model structure where a hydrophobic containing chain and the hydrophobic interactions between the chains hydrophobic groups and other chains in the vicinity are contained in an imaginary tube (Fig. 12). According to the model the HMPAM polymer chains form an entangled transient network where entanglements and hydrophobic interactions are present. The distance between entanglements is shorter than the one between hydrophobic groups. Two different relaxation times, a short and a long one, are predicted for the entangled transient network. The short one corresponds to residence time of a hydrophobic group in a hydrophobic micro-domain and the long one corresponds to the time a chain takes to reptate outside its tube. In the dilute region (at low polymer concentration) intramolecular associations dominate [108]. The hydrodynamic volume is reduced and therefore the viscosity of the subsequent polymer solutions. When the polymer concentration is increased the solution ideally moves to the semi-dilute region where intermolecular associations dominate [107,108,293]. This leads then to networklike formations (transient network) which substantially increases the viscosity of the solution [107,108,293]. Many publications discuss the complex nature of the rheology of polymer solutions of hydrophobically modied polyelectrolytes. It has been demonstrated that all the types of non-Newtonian rheological behavior exist for these polymer solutions, i.e. pseudoplastic; pseudodilatant; thixotropic; and rheopectic. Application of shear will, in classical pseudoplastic polymeric solutions, disrupt the hydrophobic associations, which leads to a reduction in viscosity [88,92,103,109,132]. However this process is reversible, i.e. when the shear is removed the hydrophobic groups will form new association thus returning the viscosity of the solution up to its

original value [88,92,103,109,132134,136,138,139,242, 257,278,293]. Some associative polymeric solutions display pseudodilatant behavior, i.e. increase in viscosity with increasing shear rate [109,134,220,258,260,293,294]. According to several studies [109,134,220,295,296] this behavior can be interpreted as a balance between intra- and intermolecular associations. Above a given shear rate the intramolecular associations are disrupted and the polymeric chains are extended, which leads to more intermolecular associations [220]. A later study [258] demonstrated that the pseudodilatant behavior arises slightly above the crossover concentration (C*), which is the overlap concentration of the polymer chains. Increasing the polymer concentration will lead to a polymeric solution that does not display pseudodilatant behavior [109,134,258,297]. The pseudodilatant behavior of the polymeric solutions is followed by the pseudoplastic behavior discussed earlier. Although many authors have demonstrated the pseudodilatant properties of HMPAM solutions, it has to be mentioned that there are differences in the observed behaviors. The most important one being that polymers produced by post modication only display the pseudodilatant behavior in the presence of salt [293,294] while polymers produced using the micellar copolymerization technique display the pseudodilatant behavior also in water [109,134,220,258,260]. However, no clear explanation for this observation has been given yet. As mentioned before, the type of distribution of the hydrophobic groups has a pronounced effect on the type of rheology that is subsequently observed in solution. By post modifying the polymers the distribution of the hydrophobic groups would be different since little control is present on which functional group is modied. The distribution of the hydrophobic groups is similar to block-like distribution obtained with micellar copolymerization. Thixotropic and rheopectic (anti-thixotropic) behavior, as mentioned before, has been observed. The viscosity against shear rates curves obtained under increasing and decreasing shear rates are not superimposable [88,103,175,257]. Thixotropic solutions display a peculiar behavior where the viscosities along the increasing shear

1584

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

Fig. 13. Counterion mediated interpolymer associations. Reproduced with permission from [214] 2005, ACS Publishing.

rate curve are all higher than those along the decreasing shear rate curve, i.e. the viscosity decreases with time at a constant shear rate. The application of shear disrupts the hydrophobic associations, both intra- and interchain, which need time to re-associate in order to recover the enhanced viscosity. In the semi-dilute region interchain associations dominate, as mentioned before, and there disruption by shear will lead to a signicant loss in viscosity [257]. However as the viscosity is recovered, the shear induced reduction in viscosity is not due to polymer degradation [88]. Polymers prepared using the heterogeneous or homogeneous technique do not display this behavior [88]. Rheopectic solutions display the opposite behavior, i.e. the viscosities along the increasing shear rate curve are all lower than those along decreasing shear rate curve (the viscosity increases with time at a constant shear rate). This behavior involves the instant recovery of the hydrophobic associations after the application of shear and the enhancement of these hydrophobic associations, which explains the enhanced viscosity [103,175]. 3.1.2.1. Effect of inherent parameters. As mentioned before, spacers between the hydrophobic group and the hydrophilic backbone can enhance the solution viscosity. Increasing the length of the spacer will lead to a better enhancement of the solution viscosity. A classic example of such behavior is polymer 13 (NaAMPS as the co-monomer) [215]. A subsequent study [214] demonstrated that, depending on the incorporation rate of the DEmMA, a signicant increase in solution viscosity is obtained above the critical overlap concentration. The authors hypothesized that the additional increase in solution viscosity is due to the simultaneous interactions of countercations with the EO spacers via coordination and with the polyanion via counterion condensation. A schematic presentation illustrating this behavior is given in Fig. 13. It has been demonstrated that the responsiveness of the polymer solution is more pronounced when a longer spacer is used. A study of Kathmann et al. [199] demonstrated that using three methylene groups as spacers increased the pH responsiveness of the polymer solution compared to when only one methylene unit is used. The pH responsive behavior of the polymers can also be tailored by molecular design of the polymer. The ampholytic polymer 7 represents such a polymer where by smart design different behavior is achieved as a result of changes in the pH. The polymer

(7) undergoes a polyanion polyzwitterion polycation transition as the pH decreases [202]. The shear rate for the onset of pseudoplastic behavior 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 uorocarbons instead of hydrocarbons [298], result in a reduction of the shear rate for the onset of pseudoplasticity and an enhancement of the viscoelastic properties. A classic example of a polymer whose shear rate for the onset of pseudoplasticity reduces with increase in the hydrophobicity of the groups is polymer 15 [185]. The rheological properties of HMPAM can be, as mentioned before, explained by the balance between intra- and intermolecular hydrophobic associations. The copolymerization process affects the distribution of the hydrophobes (vida supra). With the preference being intermolecular associations, an enhancement in the solution viscosity is obtained whereas with intramolecular associations the exact opposite is observed [110,111]. Polymers 16 and 29 are classic examples of polymers with an enhanced solution viscosity due to their preference for intermolecular association. The position of the hydrophobic groups along the backbone is important [245]. From a chemical point of view the amount of hydrophobic groups present in the polymer should have an optimum. If the concentration is low, no association will arise, while, if too high, solubility issues play a crucial role. However it is demonstrated that placing the hydrophobic group (polymer 24) at the ends of the polymer is the optimal conguration for obtaining the highest solution viscosity [245]. According to the authors the placement of hydrophobic groups along the backbone of the polymer leads to a polymer with a much more compact structure in solution compared to when the hydrophobic groups are placed at both ends. However when combining the two extremes (telechelic with multisticker) it was demonstrated that the highest thickening capability, i.e. solution viscosity against polymer concentration can be obtained [104,245]. The critical polymer concentration for the onset of hydrophobic associations depends on the placement of the hydrophobic group [104]. For a telechelic polymer with hydrophobic groups at both ends and a polymer with hydrophobic groups at both ends and along the backbone the critical polymer concentration for the onset of hydrophobic association is the same whereas for a multisticker polymer with only hydrophobic groups along the backbone the concentration is higher. The same kind of consideration, namely strong dependence of the rheological behavior from the chemical structure is applicable for uorine containing polymers. Indeed, polymer 18 displayed shear thickening, which according to Chang and McCormick [228] depends on the molecular structure of the copolymer. Increase in the incorporation rate of the hydrophobe from 3 to 10 mol% leads to a more pronounced shear thickening behavior. However at an incorporation level of 24 mol% the shear thickening

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

1585

behavior is signicantly suppressed. The shear thickening is attributed to the breakage of intramolecular hydrophobic associations by shear. If the hydrophobic associations are too strong (as is the case for the latter polymer), the applied shear cannot break the associations and thus shear thickening is not observed. The dependency of the solution viscosity on the (low molecular weight) electrolyte concentration can be adjusted as function of the chemical structure. Indeed, the salt tolerance of terpolymer 19, 20 was examined using three different derivatives (NaAMPS, NaAMB and NaA) as the third monomer. Terpolymers incorporating NaA as the third monomer displayed the best salt tolerance of the three different terpolymers [229,230]. The apparent viscosity ( app ) of the polymer solution was signicantly higher in high salt concentration solution compared to the other two terpolymers. The strength of the hydrophobic associations depends on the type of groups used. The terpolymers incorporating carboxylate anions displayed stronger hydrophobic interactions compared to the polymers with sulfonate anions. In addition, the placement of the anions is crucial for the strength of the hydrophobic associations. It was demonstrated that the further away from the polymer backbone the anions are placed the weaker the hydrophobic associations will be. This behavior was attributed by the authors [229] to the interference by the anions with the hydrophobic associations. In addition the rheological properties of the polymer solutions depend on the charge density. Polymers with low charge density lead to less electrostatic interference of the hydrophobic associations [230]. The rheological properties of polymer 15, 22 and 24 depend on which hydrophobe is incorporated. Using disubstituted monomers leads to a more pronounced thickening effect compared to monosubstituted ones. This is attributed to the proximity of two hydrophobic chains (higher density of the hydrophobic domains) that leads to stronger hydrophobic associations [109]. In addition the length of the hydrophobic block affects the rheological properties of the polymer. The longer the hydrophobic blocks, the stronger the hydrophobic associations will be. The onset of shear thickening shifts to lower shear rates as the length of the hydrophobic blocks increases [109]. This behavior is demonstrated using polymer 14 (R3 = 5) where the thickening capability increases as the hydrophobic block increases in length [299].

3.1.2.2. Effect of external parameters. For neutral associative copolymers it has been demonstrated that upon addition of salt the viscosity of the solution can be enhanced [103,108,203]. This behavior of the neutral copolymers is attributed to a salting out effect, which arises due to a change in the solubility of the hydrophobic units [108]. The solubility of the hydrophobic groups decreases with increasing salt concentration. This leads to the formation of aggregates, which enhances the interactions between the polymeric chains [108]. The intermolecular association is enhanced leading to a stronger network and thus increased solution viscosity [217]. Polymer 14 is a classic example of this behavior [217].

The presence of salt causes a shielding of the electrostatic interactions and thus a collapse of the network with total loss of the solution viscosity. Polymer 1 is a classic example of such a polyelectrolyte where phase separation is observed in the presence of multivalency cations [193]. Polymer 5 displays a similar trend. However, a decrease in the pH of the polymeric solution (containing salts) improves the solution viscosity [198] due to the increase in hydrodynamic volume caused by the protonation of the carboxylic groups at low pH [198200]. The effect of salt on polymers bearing charged moieties in combination with hydrophobic groups is peculiar. Indeed, an increase in the solution viscosity is observed upon the addition of different salts. The presence of salt screens the electrostatic repulsion thus suppressing the disruption of hydrophobic associations by the charged groups. This allows the formation of a stronger network through hydrophobic associations [225], i.e. intermolecular hydrophobic interactions are dominant rather than the electrostatic repulsions [221], and thus an enhancement of the solution viscosity. Classic examples of this behavior (vida supra) are the polymers 12 [115], 17 [221] (AA-CN copolymer, N 12), 18 [225] and 25 [133,238]. Although not in the same class of polymer, polymer 4 displays the same behavior [197]. Another possibility is that the methyl groups on the AMDAPS monomer induce hydrophobic behavior of the monomer at higher salt concentration thereby displaying the aforementioned behavior. Maia et al. [121] demonstrated that the method of preparing the HMPAM solutions affects their rheological behavior in the presence of salt (NaCl). When the polymer, 23 (AMDiC6 AM copolymer), is dissolved in the salt (brine) solution (A, Fig. 14) the classical behavior, i.e. reduction in viscosity with increasing salt concentration, is displayed. However, if the polymer is dissolved in water rst, which is preceded by the addition of salt, a complete different solution behavior is observed. Diluting a polymer solution (no salt) with water and adding salt afterwards (B, Fig. 14) results in a polymer solution whose viscosity passes through a maximum with increasing salt concentration. If the polymer solution (no salt) is diluted with salt water (C, Fig. 14) the viscosity of the resulting polymer solution increases with increasing salt concentration. Fig. 14 presents the three different behaviors. This conrms the non-equilibrium character of these solutions, but a better explanation has yet to be provided. It has been demonstrated that the following surfactants interact with different HMPAM polymers and give, depending on the surfactant concentration, a higher viscosity [90,92,115,138,173,287,300302]. The interaction between the polymers 12, 13 (both copolymers), 17 (both copolymers), 18 (DADMAC-1 and DADMAC-2 copolymers), 21 (copolymer AM-C10 AM), 22 (terpolymer 1, R2 = 2), 23 (copolymers AM-E AM, AM-BPAM and AM-2 or 3 or 4), 27 (6) and 33 with several different surfactants (SDS [90,92,120,138,173,212,228,243,301305], potassium dodecyl sulfate (KDS) [212], cetyltrimethylp-toluenesulfonate [CTAT] [305,306], ammonium dodecyltrimethylammonium chloride [DTAC] [290,300], hexadecyltrimethylammonium [HTAC] [303,304], CTAB [90,233], trimethyltetradecylammonium bromide [TTAB]

1586

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

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], alkylbenzenesulfonates [ABS] [287], Triton X-100 [302], n-octyl -d-glucopyranoside [OG] [303] and n-octyl -d-thioglucopyranoside [OTG] [303,304]) have been investigated. The viscosity of the HMPAM solution with surfactant passes through a maximum, which is always just under the surfactants critical micelle concentration (CMC) [173]. The increase in viscosity is attributed to the formation of mixed micelles of surfactant and hydrophobic regions [92,117,120,154,173,300,301,303,304,307309]. According to several authors [92,303,304], when adding surfactant to a HMPAM solution, the surfactant binds to the copolymeric regions, which leads to preferential interchain micro-domain formation instead of intrachain. A schematic presentation of this behavior has been provided by Biggs et al. [92] (Fig. 15). The surfactant concentration in region I is low. The surfactant associates in a non-cooperative way with the hydrophobic groups [92]. However, this is dependent on the type of surfactant. Two different studies [303,304] demonstrated that the surfactants hexadecyltrimethylammonium chloride (HTAC) and SDS bonded in a non-cooperative way but the surfactants OG and OTG bonded in a cooperative manner. Winnik et al. [304] concluded that ionic surfactants bind by a non-cooperative mechanism and neutral surfactants by a cooperative one. According to another study there are two classes of interactions, which are differentiated by the bridging and viscosity

increase below or above the cmc of the surfactant [90]. In region II enough surfactant molecules (higher surfactant concentration) are present to solubilize the hydrophobic groups more effectively (mixed micelles). The formation of mixed micelles is the onset of a signicant increase in viscosity. Most of the mixed micelles incorporate more than one hydrophobic region, i.e. a higher degree of overlap or links between separate chains is achieved. In region III the surfactant concentration is higher, at or above the cmc, which results in solubilization of each hydrophobic region in one micelle. This leads to a reduction in viscosity. Transition from spherical to rod-like micelles in the surfactant phase has been observed upon addition of potassium bromide and hexanol to respectively the polymerCTAB and polymerCDMAO mixture [90,115]. The transition is believed to be due to the gain in free energy upon addition of the chemicals where the hydrophobic groups no longer are exposed to the solvent but instead are present inside the hydrophobic rods [90,115]. Fig. 16 provides a schematic presentation of the rod-like micelles. This transition to the rod-like conformation leads to more intermolecular bridging and thus enhanced solution properties. This is evident from the signicantly higher viscosity of the solutions compared to the viscosity of the individual components [90,115]. Both, rheopectic and thixotropic, behaviors have been observed for polymer solutions in the presence of surfactant [92].

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

1587

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

Using polymer 23 (R2 = 2, 3 or 4 series) Gouveia and Muller [306] observed a large increase in the solution viscosity upon the addition of CTAT. Adding salt (NaCl) to this solution leads to a further increase in the solution viscosity of the mixture. According to the authors the worm-like micelles increase in size by the addition of salt and at high salt concentration salting out effects arise. A peculiar behavior is observed for the interactions between uoro- and hydrocarbon copolymers 12 and surfactants. At low incorporation level of the hydrophobic groups the interactions with surfactants is selective to surfactants with the same chemical nature as the hydrophobic groups [212]. Different oil reservoirs possess different temperatures. The polymer solutions that are to be used must be able to cope with the different temperatures, i.e. no loss of viscosity with alteration in temperature. The viscosity of polymeric solutions employing the polymers 10 (terpolymer of AMNaAMPSAMPDAC) or 12 (Table 1) is minimally dependent on the temperature in the range 3060 C [204,209,210]. In addition polymer 12 displayed good retention of solution viscosity for a prolonged period (40 days) [209]. The viscosity of the polymer solution containing the zwitterionic (monomer) polyelectrolyte,

polymer 4, displays a unique behavior when the temperature is increased. The intrinsic viscosity of the polymer solution increased (from 6.5 to 8.5 dl/g) when the temperature is increased from 25 to 60 C [197]. This has been observed for polymer 14 in the temperature range of 2040 C. According to the authors [217] the increase can be attributed to the fact that hydrophobic associations are endothermic in the investigated temperature range, as hypothesized by McCormick et al. [103]. Using the latter terpolymers 14 LAlloret et al. [310] observed a signicant increase in the solution viscosity when the temperature is increased from room temperature to 80 C. In addition the authors demonstrated that the thermothickening behavior of the polymeric solution is more pronounced at lower salt concentration and lower shear rates. The latter effect is also observed with polymer 27 (6) [243] and is attributed to the increase in PPO concentration in the hydrophobic micro-domains caused by an increase in the mobility of the chains at relatively high temperature. At higher temperatures, the reduction in viscosity is attributed to the loss in connectivity of the network due to changes in the hydrophobic microdomains. According to the authors the size of the microdomains 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) 15581628

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 loss. The viscosity of the AMAMPDAC copolymers in the polymer series 1 decreased as expected with increasing temperature [193]. The same behavior is observed for a polymeric solution employing polymer 2 [195]. Unlike polymers 1 and 2, an intriguing property of polymer 4 is the increase of the intrinsic viscosity with temperature (in the range 2560 C). Another polymeric solution that displays peculiar behavior with increase in the temperature is the solution of terpolymer (9, AMNaAMPSAMPDAC), which was minimally dependent on the temperature [204]. The LCST concept, discussed in the introduction of this chapter, also applies to polymer 33 where the viscosity of the solution can increase with an increase in temperature. However, one important nding reported by the authors [235] is the presence of hydrophobic associations below the LCST. An increase of the solution temperature (towards the LCST) leads to disruption of these associations. However, only for long or high hydrophobicity groups do the associations arise below the LCST. This behavior was observed with the amphiphilic polymer 23 (NIPAM-CN AM). Four decades ago, it was proposed that the LCST should decrease with the polymer hydrophobicity [311]. However, the study by Ringsdorf et al. [235] demonstrated that the proposed trend is not followed when increasing the length of the hydrophobes in polymer 23 (NIPAM series). This is attributed to the formation of a micellar structure by the hydrophobes, which prevents them from contributing to the LCST suppression. On the basis of this hypothesis, the authors have proposed a model representing the thermally induced polymer conformational changes for both, 23 (NIPAM series) and 33, polymers (see Fig. 17). In addition, increasing the temperature from 25 to 55 C leads to the preferential formation of intermolecular hydrophobic associations and thus an increase in the viscosity. The pH of the polymeric solution also affects the solution viscosity. The effect on classic polyelectrolyte is easily understood given the charged nature of the polymers. Polyanions will have a low viscosity at low pH and high viscosity at high pH while polycations display the opposite behavior. Polymer 2 is a classic example of a polyanion where the solution viscosity decreases as the pH of the solution is reduced. The effect of pH is more pronounced on polyelectrolyte amphiphilics where the associations are a balance between hydrophobic and electrostatic interac-

tions. Zhou et al. [225] demonstrated, using polymer 18 (AA series), that increasing the pH from 4 to 5 and from 11 to 12, a signicant increase in the solution viscosity is observed while at a pH between 5 and 11 the solution viscosity is lower. A similar behavior was observed for polymer 17 (AA series) although the pH where the increase in solution viscosity is observed is shifted. The increase is observed between a pH of 56 and 1213 [221]. Although the effect of pH on polyelectrolyte amphiphilics is complex (vida supra), the investigation of the polymer network structures of solutions employing polymer 15 using NRET measurements led to the development of a model illustrating the effect of the solution pH on the conformation of the polymer chains and their associations [218] (see Fig. 18). Increasing the pH will lead in principle to the transition towards more intermolecular associations rather than intramolecular. The high degree of ionization (at high pH) will disrupt the intramolecular hydrophobic associations causing a rearrangement of the network with a preference for intermolecular associations [218]. Amphiphilics are affected by the solution pH mainly due to the carboxylic groups of the polymers. Polymer 26 is a classic example showing what the effect of the pH is on the behavior of an amphiphilic polymer. The solution viscosity passes through a maximum as the pH rises from 4 to 12. This behavior is attributed to two different effects: neutralization of the carboxylic groups leading to (1) intramolecular electrostatic repulsion and thus chain extension and (2) intermolecular electrostatic repulsion leading to disruption of the intermolecular associations. The rst effects dominate at low pH, the latter one at high pH [240]. However, in contrast to this classic polyelectrolyte behavior, increase in the pH results in the formation of intramolecular hydrophobic association for polymer 29. The carboxyl groups repel each other in classic polyelectrolytes to form expanded coils, however the presence of large hydrophobic groups on the polymer backbone impart the repulsion between the carboxylic groups. The polymer forms a much more compact structure, which is called a hypercoil [312] (Fig. 19). The hydrophobic groups associate to form the interior of the hypercoil, which is surrounded by carboxyl groups. The same behavior has been observed for polyacids bearing large hydrophobic groups [95,96]. McCormick et al. [219] noticed the same behavior for polymer 16. By increasing the ionization degree of the polymers containing 50 mol% 4-butylaniline (4-BA) the polymers undergo a transition from closed (intramolecular) associations to open (inter-

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

1589

Fig. 19. Schematic presentation of a hypercoil. Reproduced with permission from [312] 1986, ACS Publishing.

Fig. 20. Schematic presentation of transition to intramolecular association. 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 ionization degrees. Unlike polymer 16 and 29 no intermolecular hydrophobic associations could be detected for polymer 31. This is attributed to the high number of carboxylate units at high pH which will disturb the water structure to such an extent that the driving force responsible for hydrophobic associations is suppressed. At low pH, intramolecular hydrogen bonding stabilizes intramolecular hydrophobic associations thus excluding intermolecular association. The transition from closed (intramolecular) to open (intermolecular) associations has been conrmed using the model polymer 32. This transition, which is triggered by changes in the pH and salt concentration can be viewed as a molecular level event instead of a macroscopic phenomenon [178]. A schematic presentation of this behavior has been provided by Kramer et al. [179] (see Fig. 20). Although a transition arises for polymer 19 with changes in the ionic strength and/or pH, the restructuring of

the polymer is different. Two models have been proposed by the two studies [226,227] illustrating the effect of the ionic strength and pH on the structure of polymer 19 that is obtained (see Fig. 21). Indeed, for classic polyelectrolytes, increasing the ionic strength of the solution (i.e. increasing the salt concentration) will lead to a decrease in the solution viscosity due to screening effects. However for polymer 9, which bears both electronic charges, a peculiar behavior is observed where the intrinsic viscosity increases with increase in the ionic strength [205,206]. 3.2. Hydrophobically modied ethoxylated urethane (HEUR) The interest in HEUR focussed on fundamental research [313] and commercial applications in paint formulation [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) 15581628

Fig. 22. Chemical structure HM PEO.

formulation [314,315]. Their use as model compounds can signicantly increase the scientic understanding of associating polymers. Table 2 presents an overview of the published systems with the corresponding chemical structure. HEUR polymers are prepared by chain extension of poly(ethylene oxide) (PEO) oligomer with a narrow molecular weight distribution (MWD) using di-isocyanate and end-capping with a hydrophobic group [105]. HEUR polymers can be categorized into two main classes: step growth (S-G) HEUR and uni-HEUR. The differences between the two classes relates to the synthesis method and the corresponding MWD of the product [336]. The S-G HEUR involves a procedure where PEO (of a given Mw) is reacted with a large excess of a di-isocyanate agent to produce a precursor which contains isocyanate functional groups at both ends. This precursor is subsequently reacted with a hydrophobic group to end-cap both ends yielding a telechelic polymer. This procedure leads to a HEUR polymer displaying a broad MWD. The procedure for uni-HEUR involves the direct addition of a mono-isocyanate containing a hydrophobic group to the PEO. To this a hydrophobic group is added. The resulting polymer has a relatively narrow MWD, which is related to that of the parent PEO. The chemical structure of a hydrophobically modied PEO is given in Fig. 22. HEUR polymers are designed to be end-capped at both ends with a hydrophobic group. However it has been demonstrated that, in practice, a signicant fraction of these polymers is end-capped at only one side [337]. The polymers usually comprise a mixture of telechelic polymer and nonionic surfactant [125]. In most cases long chain alcohols [105,122,322,324,325] are used as hydrophobic groups although other types of hydrophobic groups have been investigated. These include uorocarbon containing hydrophobes [186,329331], amine containing hydrophobes [122], alkyl groups [124,128,170,316,327] and alkoxy groups [129,318,319]. Combinations of telechelic and internal hydrophobes [122] and comb-like structures using a long chain 1,2-diol [125] have been investigated. As mentioned earlier an isocyanate agent is used in the synthesis. Different isocyanate agents have been used such as isophorone di-isocyanate (IPDI) [122,125,129,186,318,319,329331,338], dicyclohexylmethane di-isocyanate [322,324,325], octadecyl (alkyl) isocyanate [317,323,339] and different polyisocyanates [317,339]. Differences in the synthetic procedures and the used hydrophobic groups clearly result, as discussed below, in

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 elucidated using uorescence studies. In principle rosette-like [124,126,318,319,328] aggregates are formed with diameters in the nanometer range resembling star polymers [340,341]. The formation of these structures starts above the critical micelle concentration (CMC) [124]. The rosettelike structure comprises an inner core rich in hydrophobic groups (ranging from 4 to 20 hydrophobic groups) and a corona composed of the hydrophilic PEO chains (approximately 510 looped chains) [130,318,319,342]. The number of hydrophobic groups present in the hydrophobic core depends on the type of group used [326]. The number of hydrophobic groups in the inner core and the structure of the corona are independent of polymer concentration [318,319], only the number of rosettes is affected by the polymer concentration. Although the exact dimensions vary with the polymer molecular structure, the polymers 35, 36, 43 and 44 all were demonstrated to form the aforementioned structures in an aqueous solution. Fig. 23, provided by Xu et al. [343], shows the micelle structures and the effect of polymer concentration on the conformation of the polymer (for polymer 36, AT 223). Several other publications [105,126,329,338] provided similar schemes illustrating the different concentration regimes. Aside from the independency of the structures on concentration, the type of end group is another parameter that does not affect the network structure. According to Fonnum et al. [328], only the relaxation time of the network depends on the end group as observed with polymer 45 (16). The rosettes can be bridged by polymeric chains, at higher concentration, leading to a network as displayed in Fig. 23. The higher the tendency towards bridging, the more stable the network will be. The bridging is affected by the polymer molecular structure. Xu et al. [343] have demonstrated that the comb polymer 36 has a much higher fraction of bridging chains and a lower fraction of loop chains when compared to the telechelic analogues. When investigating the structures of HEUR polymers, polymer 38(2) containing only one hydrophobic group, a peculiar behavior is observed. A smaller hydrophobicity group (C6 F13 ) leads to dimer formation whereas a larger hydrophobicity group (C8 F17 ) leads to the formation of micelles (Fig. 24) [189]. For HEUR polymers hydrophobic associations start at very low polymer concentration (as low as 10 ppm). Although the exact critical concentration for the onset of hydrophobic association varies, the solutions employing polymers 39, 43 and 44 have a critical concentration in the same range [125,327]. Most of the HEUR polymers in Table 2 contain urethane linkages. A study by Alami et al. [344] demonstrated that at low polymer concentration the presence of urethane linkages dramatically inuence the initial association and the clouding of the polymer in water. HEUR polymers containing urethane linkages 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 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.

[89,317]

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

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 structure. Either end capped (traditional HEUR polymers, 1) or the hydrophobic end groups are separated by oxyethylene groups around an internal hydrophobe, 24.

[122]

1591

1592

R = different amine and alcohol containing groups

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

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

[189,320]

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) 15581628

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:

[322325]

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

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

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

[188,326]

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) 15581628

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) with octadecyl groups (1) or with nonylphenol (2) groups:

[187] D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

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

[329]

Polymers based on ethylene oxide end capped either at both (1) or one side (2) with uorocarbon hydrophobes.

[186,330,331]

Table 2 (Continued) Polymer End group: F(CF2 )8 (CH2 )11 Structure References D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

Polymers based on ethylene oxide end capped either at both or one side with uorocarbon hydrophobes:

[182,332334]

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) 15581628

Fig. 23. Concentration regimes of a telechelic PEO. Reproduced with permission from [343] 1997, ACS Publishing.

urethane linkages. The polymer series 49 was synthesized with the idea that this would be better model polymers since no interference by urethane linkages is present [332]. Indeed, it was already demonstrated by another study [344] that signicant changes in associating behavior are observed when telechelic polymers with an ether versus a urethane bond connecting the hydrophobic groups with the polyethylene oxide chain are compared. The association behavior of the polymer in solution depends on the molecular structure of the polymer. When Liu et al. [321] investigated polymer 40 a signicantly different association mechanism was observed when comparing comb and telechelic polymers. Comb associative polymers display a more compact structure in comparison to the association of equivalent telechelic polymers. The authors [321] hypothesized, similar to the conclusion of Jimenez-Regalado et al. [245], that comb associative polymers favour the formation of intramolecular associ-

ations instead of intermolecular association which occurs for telechelic polymers. When using a combination of small spacers between the end groups and the PEO backbone, internal hydrophobe phase separation was noted (polymer 37, 24). According to the authors [122] this behavior can be attributed to the dominance of intramolecular hydrophobic associations. To resolve this, the use of SDS was suggested. The polymer series 38(1) are prepared by step growth polymerization and, as a result, a mixture of components is obtained [320]. In principle the strength of the network in solution depends on the type of hydrophobic group used and the length of the PEO chain. The strength of the network can be increased by increasing the hydrophobicity of the end groups. The hydrophobicity can be increased by using larger sized hydrophobes, using uorocarbon rather than hydrocarbons, or by increasing the size of the di-isocyanate group [124,320,329,330]. An example of such a polymer 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) 15581628

1599

of the end group leads to an enhanced network. It was demonstrated that the thickening efciency decreased in the order of octadecyl > pentadecyl > 9-heptadecenyl [124]. The same trend was observed for polymer 43 (2) where the thickening efciency decreased in the order of dodecyl > octyl > hexyl [188] and for polymer 45 (16) [328]. Polymers 46, 47 and 48 (1 and 2) represent polymers with increased network strength due to the presence of uorocarbons instead of hydrocarbons. 3.2.1. Rheological properties Several publications presented the rheological properties of HEUR polymers in aqueous solutions. In aqueous solution HEUR polymers form transient networks, which follow a Maxwellian behavior with a single characteristic relaxation time [89,124,318,328,338,343,345349]. This relaxation time was attributed to the residence time of a hydrophobic end (group) inside a hydrophobic microdomain. At low polymer concentration, no micelles are present and the viscosity of the solution equals the viscosity of the solute. Increasing the polymer concentration above the CMC leads to micelle formation without a signicant increase in viscosity. A further increase in the polymer concentration leads to more micelles. As mentioned before, these micelles can be interpreted as star-like polymers, which are known to repel each other in good solvents [89,345347,350]. However, telechelic micelles have the possibility of forming bridges between two separate hydrophobic cores, thus leading to cluster formation between them [341]. The viscosity of the polymer solution now changes due to an increase in the hydrodynamic volume of the micellar clusters formed. Raising the polymer concentration leads to networks (microgels) which are connected through bridges [351,352]. These microgels break upon dilution of the polymer solution. Two studies from the Annable group [89,350] proposed similar network formation where micelles (or a collection of micelles) are connected through superbridges and superloops. Superbridges connect several micelles in a linear fashion whereas superloops connect several micelles in a non-linear fashion, but the start and end of this superloop rest in the same micelle [89,350]. In addition, dangling ends are present in the microgels, which are chains with only one hydrophobic end associated with other hydrophobic entities (Fig. 25). By using uorescence spectroscopy, it has been demonstrated that a critical association concentration (CAC) exists for telechelic polymers end-capped with alkyl groups [89,128] (polymer 34). Above the CAC the viscosity of the polymer solution increases more sharply with increasing concentration when compared to polymers lacking the hydrophobic groups [128]. In HEUR solution the interchange of the bridges connecting the micellar clusters is fast at low shear rates, without breakage by shear [170] and thus no loss of solution viscosity. Nevertheless, HEUR polymers also display pseudoplastic and pseudodilatant behavior. Under steady shear, the HEUR polymeric solutions display a strong pseudoplastic behavior that is preceded by a weak shear thickening. The pseudoplastic behavior arises due to the

breakage of the bridges as the shear increases [328]. The microgels reduce in size and thus also the hydrodynamic volume leading to a lower viscosity of the solution. As mentioned before, prior to shear thinning a weak pseudodilatant behavior is observed. At moderate shear the microgels are deformed, which entails the stretching of the bridges connecting the microgels [105,318,319,328]. The bridges resist breakage by stretching (up to a certain point) and therefore induce a shear thickening [105,318,319]. A further increase in the shear rate leads to the aforementioned pseudoplastic behavior. Tam et al. [105] investigated the network structure of polymer 36 (AT22-3) in shear ows and proposed a more rened model than the one provided by Yekta et al. [338], depicting the different structures the network attains under shear (Fig. 26). As the shear is increased shorter superbridges are formed but the density of mechanically active, i.e. part of the polymer network, chains increases. This leads to a stronger network and thus a higher solution viscosity. A further increase of shear will lead to the shear thinning regions. According to the study [105] three different shear thinning regions can be distinguished for polymer 36, i.e. regions A, B and C. Regions 1 and 2 are characterized by a signicant loss in viscosity and the difference between the two is a reduced slope (viscosity versus shear stress plot) of the latter one. Region 3 is characterized by a large drop in the solution viscosity with increasing shear. The difference between regions 12 and 3 is the mechanism causing the loss in viscosity. It was demonstrated that in regions 12 the superbridge network junctions disrupt, but they rearrange quickly without a loss in the number of bridges. The reduction in the relaxation time leads to the reduction in viscosity for regions 1 and 2 given that the viscosity is dependent on two separate processes according to the = G0 equation ( = relaxation time and G0 = plateau modulus) for a Maxwell uid [105]. The loss in viscosity for region 3 is attributed to fragmentation of the network. 3.2.2. Effects of inherent parameters Telechelic polymers such as the polymer series 35 display in steady shear a Newtonian behavior up to relatively high shear rates [105]. This is in stark contrast with polymers bearing hydrophobic moieties distributed along their backbone, where shear thinning predominates. When incorporating the hydrophobic group along the polymer backbone, the thickening capability, i.e. solution viscosity as a function of the polymer concentration, is lower when compared to telechelic polymer [89]. The comb associative polymer 40 is an example of a polymer with less pronounced increase in solution viscosity when compared to a telechelic equivalent. However, it must be mentioned that although both polymers are quite similar, the molecular weight of the comb type is higher than for the telechelic one. In addition, it was demonstrated that the distribution of the hydrophobic groups plays a crucial role in the observed behavior. When each of the hydrophobic moieties consist of two hydrocarbon groups the thickening capability is signicantly enhanced. This is similar to the behavior observed for twin-tailed HMPAM polymer

1600

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

15 [321]. The hydrophobicity of the groups increases and therefore a stronger interaction gives rise to the observed enhanced thickening capability. In light of this, the comparison between a single comb and telechelic polymers of polymer 40 is not viable given the low hydrophobicity of the groups in the comb polymer. Internal hydrophobes affect the rheological properties of HEUR solution due to the preference for intramolecular hydrophobic associations. However, there are conicting reports on the effects of internal hydrophobes. According to Fonnum et al. [328] internal hydrophobes do not affect the rheology of polymeric solution employing polymer 45. On the other hand Lundberg et al. [122] concluded that using HEUR polymer 37 (with an internal hydrophobe in combination with two terminal hydrophobes) resulted in a more compact structure due to intramolecular hydrophobic associations rather than intermolecular (vida supra). This leads to a reduction of the hydrodynamic volume of the polymer chains, and thus a reduction in the solution viscosity. The hydrophobicity of the end groups affects the rheological properties. According to Calvet et al. [330] the increase in hydrophobicity of the groups leads to an increase in the lifetime that the hydrophobic end groups are present in the hydrophobic core. It has also been demonstrated that the monofunctionalized polymer 48 (2) displays a marked difference in rheology compared to the difunctionalized polymer 48 (1) [330]. The concentration at which the onset of signicant (static) viscosity increase is observed is much lower for the difunctionalized polymer. However, it cannot be concluded that this difference is only caused by the single end capping since the molecular weight of the monofunctionalized polymer is half that of the difunctionalized one. The rheological properties of a HEUR solution also depend on the molecular weight of the polymers. In principle, the molecular weight of the HEUR polymer has an optimal intermediate range where the highest thickening capability can be obtained [331]. The critical polymer

concentration for the onset of hydrophobic association is also affected by the molecular weight of the polymer. According to Kaczmarski and Glass [187] the increase in the number of EO units (increase in molecular weight) promotes the formation of intermolecular associations and thus increasing the network strength leading to an enhanced solution viscosity. At low molecular weights, a reduction in the solution viscosity is observed due to a reduction in the hydrodynamic volume of the polymer. At high molecular weights, also a reduction in solution viscosity is observed. This is attributed to the decrease in the concentration of the hydrophobic groups, which leads to less hydrophobic micro-domains being formed and thus a weaker network. Polymers 42, 43, 45 (16) and 47 are classic examples of polymers whose rheological properties depend on the molecular weight. The molecular weight is closely related to the PDI. The PDI also affects the rheological properties of the polymeric solution. A study [325] compared the thickening capability of the polymers and found signicant differences in the viscosity enhancement capability. For instance, when a broad MWD polymer is used twice as much polymer (concentration) is required to attain the same viscosity of a narrow MWD polymer. However, the storage and loss moduli were similar despite the concentration difference. Broad MWD polymers contain small polymer chains of low molecular weight and these polymers favor the formation of intramolecular hydrophobic associations. This leads to a weakened polymer network, which reduces the viscosity enhancement capability. 3.2.3. Effects of external parameters Interactions between surfactants and HEUR polymers in solution have been demonstrated by several authors [127,353,354]. Depending on the concentration of the surfactant, an increase in viscosity of the solution can be observed. Interactions between polymers 34, 36 (AT 22-3), 37 (1), 41, 49 and 50 and the surfactants SDS [124,164,182,335,353357], 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) 15581628

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 [353], polyoxyethylene glycol (C12 E8 , nonionic surfactant) [127] and polyoxyethylene stearyl ether (C18 [EO]20 ) [358] have been investigated. The effect of the surfactant has been summarized by Annable et al. [353] in terms of the plateau modulus: (1) at low polymer concentration the plateau modulus is signicantly raised using a low to moderate amount of surfactant, (2) using higher surfactant concentration leads to a monotonic decrease in the plateau modulus, (3) at high polymer concentration no peak in the modulus is observed and (4) the modulus peak shifts towards lower surfactant concentration as the polymer concentration is lowered. According to Alami et al. [127] the polymer concentration at which the onset of hydrophobic micro-domains arise can be inuenced by the addition of nonionic surfactants. The hydrophobic micro-domains form at much lower concentrations than the CAC or CMC when nonionic surfactants are present in the polymer solution. The study by Alami et al. [127] provided a schematic presentation of the interactions between the SDS surfactant and HEUR polymers in solution. Depending on the concentration of the surfactant and the polymer, different structures (associated polymer aggregates) are formed (Fig. 27). The study [127] also shows that it is difcult to distinguish a concentration region where only micelles are formed when a HEUR polymer with a high molecular weight is used. The formation of micelle clusters starts close to the CMC. At low (SDS) surfactant concentration (csurfactant < cmcsurfactant ) the interaction of the surfactant is mainly with the hydrophobic ends of the polymer chains, which leads to a stronger network and thus an enhanced solution viscosity [124,127,164,182,353356]. High surfactant concentration leads to a destabilized network as the hydrophobic end groups are solubilized, which results in a reduction in the viscosity. In addition an interaction between the SDS surfactant and oxyethylene

of the polymer chain has been found [164,182,354356]. In comparison, cationic surfactants only interact with the hydrophobic end groups [127,354]. Interactions between nonionic surfactants have been demonstrated where a large increase in solution viscosity is observed upon addition of the surfactant [124,127,353,358,359]. The study by Kim et al. [358] investigated the stability of the polymeric solution employing the HEUR polymer 37(1) with the polyoxyethylene stearyl ether (C18 [EO]20 ) surfactant in complex formulations. An important observation made, in the context of EOR-applications, was the stability of the micellar network in solution in an oil-water emulsion, despite the changes in ionic strength (different NaCl concentrations) and pH. In order to facilitate the investigation of molecular weight of HEUR polymer the hydrophobic associations have to be suppressed. Cyclodextrins, cyclic oligomers consisting of 6, 7 or 8 glucose units (corresponding to -, -, -CD) connected by -1,4-glucosidic linkages [360], are capable of encapsulating the hydrophobic groups thus inhibiting hydrophobic associations. The interaction between cyclodextrins and the HEUR polymers 37(1), 44 and 49 (2, F8 F8 ) support this. A schematic presentation illustrating the interactions at different stages is given in Fig. 28. Depending on the cyclodextrin concentration three different regions can be distinguished [314,361,362]. At low cyclodextrin concentration encapsulation of the end groups of the loops occurs making the network more susceptible to shear thinning. At moderate cyclodextrin concentration detachment of the bridging chains between micelle clusters occurs, which weakens the network and thus results in a lower viscosity. A further increase in the cyclodextrin concentration leads to the complete destruction of the network formation leaving only individual clusters or polymer chains in solution. The viscosity of such a polymeric solution equals the solute viscosity.

1602

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

Fig. 28. Interactions between cyclodextrins and a HEUR polymer. Reproduced with permission from [361] 2007, ACS Publishing.

3.3. Hydrophobically modied alkali swellable emulsion (HASE) HASE systems are emulsions at low pH where the polymers exist as macroscopic particles (non-transparent solution). As the pH is increased the particles will swell

and dissolution follows. These polymers have not been studied as extensively as the HMPAM or HEUR polymers [363]. However they represent an alternative to the nonionic HEUR polymers for applications as thickeners [363]. HASE polymers are terpolymers consisting of methacrylic 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) 15581628

1603

The HASE polymers are, as mentioned before, insoluble in low pH solutions and require higher pH (above pH 6) for solubilization. This, given the salinity of most reservoir water, is benecial for application in EOR. In most cases the ratio between methacrylic acid and ethyl acrylate (i.e. molar ratio x/y) equals 1 and the hydrophobic group is incorporated at a level of a few mol percentages (%). The HASE polymers can be seen as polyelectrolytes, which have been hydrophobically modied by the introduction of a few mol% of a hydrophobic group. Comb-like structure where the hydrophobic groups are placed on short PEO side chains randomly distributed along the backbone of the HASE polymer chains are favored [169,363]. The backbone of the HASE polymers possesses moderate high molecular weight and can reach 200,000250,000 Da [169]. The general chemical structure of a HASE polymer [106,169,366,367] is given in Fig. 29 while a more general overview of all studied systems can be found in Table 3. HASE polymers are synthesized using an emulsion copolymerization process [181,363,372]. The three constituents, methacrylic acid, ethyl acrylate and the hydrophobic macromonomer, are added to the reaction mixture along with an initiator. The hydrophobic macromonomers act as the surfactant for the polymerization reaction. These are produced by rst ethoxylating a hydrophobic group and subsequently reacting the product of the ethoxylation with an unsaturated isocyanate [363,364,367]. The reaction product of the emulsion polymerization is the HASE polymer. Different types of hydrophobic groups have been used which include alkyl groups [106,363,367,369,370], complex poly(alkylaryl) moieties [91,363] and hydrophobic groups containing pyrene for investigation of the association structures [365,373375]. The length of the PEO groups to which the hydrophobic groups are attached can be manipulated by controlling the number of CH2 CH2 O moieties [169,363]. In addition the length of the hydrophobic group can be altered by using longer or shorter hydrophobic agents [367]. The hydrophobicity of the hydrophobic groups can be manipulated [169]. The number of hydrophobes per chain can be controlled by the variation of the proportion of the macromonomer (Fig. 29) in the recipe during the polymerization stage. These controls offer unique possibilities for tailoring the chemical structure, and thus the rheological properties of the HASE polymer solutions [376]. The structure in solution is dependent on the molecular structure of the HASE polymer. Giving the molecular structure of the polymers 51, a comb-like structure is obtained with hydrophobic groups tethered through oxyethylene spacers to the polymer backbone [363]. The structure of HASE polymer in aqueous solution has been elucidated (see Fig. 30). By using superposition of oscillations on steady shear ow [366] of HASE solutions or in simple salt solution [369] a model was developed. Above a given polymer concentration, the unimers will form oligomers consisting of two or more polymer chains. When the polymer concentration increases, the size of the aggregates will increase. Using spin-echo (PGSE) NMR Nagashima et al. [370] concluded that the distribution of hydrophobes among the polymer chains is inhomoge-

neous. Two distinct regions can be identied, which are a hydrophobe enriched one and a hydrophobe depleted one. The hydrophobe enriched region will form more compact structures compared to the other region. By using uorescence studies, Prazeres et al. [374] investigated the association level in the polymer series 54 and found that in aqueous solution 58% of the hydrophobic pendants associate compared to only 5% in tetrahydrofuran (THF). The same authors demonstrated though, that the coil expansion caused by the aqueous basic conditions will lead to a large portion of the hydrophobic pendants (21 6%) being excluded from associated microdomains. This portion is much larger when compared to polymer solutions in THF. Varying the length of the hydrophobe groups and the oxyethylene spacers will induce changes in the strength of the polymer network. Increasing the hydrophobe length, i.e. the hydrophobicity, will lead to stronger associations and thus a stronger network. This behavior was demonstrated by Tirtaatmadja et al. [367] which is in line with ndings using polyacrylamides and poly(ethylene oxide). Increasing the length of the oxyethylene spacers increases the accessibility of the hydrophobes, which in turn leads to easier formation of hydrophobic domains [169] and therefore a stronger network. However, if the length exceeds a certain critical length intramolecular associations are favored causing a collapse of the network [169]. 3.3.1. Rheological properties The rheological properties of HASE polymers in aqueous solution are dependent on the solubilization conditions. When using alkali media, the backbone of the HASE polymer chain becomes hydrophilic and will dissolve in aqueous solutions. The hydrophobic groups of the HASE polymer will form associations, either intra- or interchain interactions, which will cause a signicant increase in the solution viscosity [369]. The dissolution of HASE polymers has been investigated by several authors [167,168,377]. The data obtained using potentiometric titration, isothermal titration calorimetry and dynamic light scattering studies, led to the elucidation of the dissolution mechanism of the HASE polymers, either in the presence or absence of salt (Fig. 31). The size (hydrodynamic radius, Rh ) of the latex particles increase in size as the pH rises (increase in base concentration) due to the electrostatic repulsions. For a HASE polymer latex particle in a salt solution the size increases but to a lesser extent due to shielding of the negative charges on the polymer backbone by the cations, i.e. Rh, no salt > Rh, salt The electrostatic repulsion forces between the chains also alter the conguration of the polymer aggregates. The thickening capability of the HASE polymers is a combination of hydrophobic association and the expansion of the polymer chains [181,370]. At low polymer concentration, the intramolecular hydrophobic interactions dominate while at higher polymer concentration the situation arises where predominantly intermolecular hydrophobic interactions are present [370]. The intermolecular hydrophobic interactions give rise to network formation thus enhancing the viscosity of the polymer solution [366,367,370]. 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) 15581628

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

[169]

R = H (1), C8 H17 (2), C12 H25 (3), C16 H33 (4) or C20 H41 (5) m = 33 (for 2) m = 35 (for 1, 35) XYZ (1) = 48.7651.240.00 (mol%) XYZ (2) = 49.0050.001.00 (mol%) XYZ (3) = 49.0550.050.90 (mol%) XYZ (4) = 49.0650.040.90 (mol%) XYZ (5) = 49.0650.040.90 (mol%)

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

[106,169,366,367]

Serie A, N = 19 and m = 0, 10, 15, 35 or 40 Serie B, m = 35 and N = 4, 8, 12, 16 or 20 XYZ (A&B) = 49.0650.010.90 (mol%)

Table 3 (Continued) Polymer Structure References D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

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

[363,366368]

m = 80 XYZ = 40.0040.0020.00 (wt.%)

1605

1606 D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

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

[91,371]

m = 53 XYZ (Refs. [169,369,370]) = 45.0055.00<1.00 (mol%) XYZ (HASE65) = 0.490.490.021 XYZ (HASE48) = 0.500.490.011 XYZ (HASE42) = 0.440.550.011 XYZ (HASE35) = 0.450.550.006 XYZ (HASE12) = 0.400.590.002 XYZ (HASE11) = 0.400.690.0035 XYZ (HASE0) = 0.400.590.002

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

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 undergoes when the polymer concentration is increased, from low to a high concentration, in an alkaline polymer solution (Fig. 32). A schematic presentation of the microstructure of the polymer chains at very low polymer concentration in an alkaline solution, as proposed by Dai et al. [369], is given in Fig. 32. At low polymer concentration a colloidal suspension is present, i.e. no solubilization of the polymer chains. The microstructure of the polymer chains at low polymer

concentration consist of a unimer with the hydrophobic domain in the center and the hydrophilic backbone extended in the solution. As the polymer concentration is increased a transition to an overlap regime occurs. This transition takes place above the cross-over concentration (Cg ) [378] and leads to a signicant rise in the viscosity of the HASE polymer solution due to the interactions between the hydrophobic groups. The structure of the network has been investigated and a schematic presentation has been 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) 15581628

There are a number of different interactions in the network of HASE polymers, which are the intraand intermolecular hydrophobic interactions, polymer chain entanglements and electrostatic interactions. Araujo et al. [180] demonstrated the presence of two different hydrophobic micro-domains using various pyrene derivatives as probes. The ethyl acrylate rich micro-domains (Fig. 33) are what Araujo et al. [180] envisaged. Most HASE polymer solutions exhibit a strong pseudoplastic behavior when sheared, which is preceded by a weak shear thickening region [106,378]. The latter arises due to the disruption of the hydrophobic intermolecular associations by shear [365]. Thixotropic behavior was also noticed by Knaebel et al. [378], similar to polymeric solutions containing HMPAM. Increasing the polymer concentration leads to an increase in the zero-shear viscosity. However, the shear rate at the onset of shear thinning shifts to lower values when the polymer concentration is increased [378]. 3.3.2. Effect of inherent parameters The effects of length of the PEO spacers, hydrophobicity of the hydrophobic group, length of the hydrophobic groups and the molecular weight of the HASE polymer on the rheological properties of HASE polymer solution have been elucidated. As discussed earlier, when the length of the PEO spacers is increased the accessibility of the hydrophobic groups for hydrophobic associations increases. This leads to a lower number of larger hydrophobic micro-domains. However if the length of the PEO spacers surpasses a critical length intramolecular associations will be the dominant type of hydrophobic interaction [169]. The strength of the polymer network reduces and a decrease in the solution viscosity follows. This was observed by Tam et al. [380] who demonstrated that PEO spacers with a moderate length displayed the highest steady shear and zero shear solution viscosity. The more pronounced shear thinning and viscoelastic behavior arose with the same polymer. Changing the type of hydrophobic groups, i.e. its hydrophobicity or length, also affects the polymer network and thus its rheological properties. Increasing the hydrophobicity or the length of the hydrophobic groups will increase the strength of the hydrophobic associations [169,367,376]. This translates

into a higher steady shear and zero shear solution viscosity but also to a more pronounced pseudoplastic behavior. Using the polymer series 51 the positive effect of the length or size of the hydrophobe on the solution viscosity was demonstrated [367,376]. The highest solution viscosity was obtained with the longest/largest hydrophobic group (C20 ). Using polymer 52 English et al. [371] demonstrated that by using more complex hydrophobic groups the network dynamics of the polymer is no longer represented by a single characteristic relaxation time but by rather the hindered reptation model put forward by Leibler et al. [291]. The deviation of the classical Maxwellian behavior arises by the coexistence of hydrophobic associations and chain entanglements [371]. Ng et al. [381] investigated the difference in solution viscosity in HASE polymers compared to the unmodied analogue. Fig. 34 schematically represents the differences hypothesized by the authors. The hydrophobic groups will interact with the ethyl acrylate blocks in the semi-dilute concentration regime. These in turn will enhance the intermolecular associations between different clusters. The higher solution viscosity of the hydrophobically modied polymers compared to the unmodied analogues is attributed to this enhancement. As mentioned before the hydrophobicity of the hydrophobic groups will affect the rheological properties of the HASE polymer. A classic example of this is the polymer series 54 [365]. The polymer with the pyrene functionality displayed a higher solution viscosity compared to the same polymer without pyrene. However, the number of intermolecular hydrophobic associations that arise is dependent on the pyrene content with more being formed at lower pyrene contents. The degree of ethoxylation affects the thickening capability of HASE polymers. The optimum ethoxylation degree lies between 5-40 mol% of the ethylenoxide on the hydrophobic monomer [380]. According to Wu and Shay [376], the balance between hydrophobic associations and electrostatic repulsion can be manipulated by the incorporation level of the methacrylic acid moiety [376]. The maximum zero shear rate viscosity is obtained with an incorporation rate of 40 wt%. A classic example of this behavior is observed for polymer 51. In addition it was 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) 15581628

1609

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

responds to the dominant elastic property of the polymer system at 10 mol% ethyleneoxide [380]. Another factor that affects the rheological properties is the molecular weight and this can be controlled by the amount of chain transfer agent (CTA) used during the emulsion polymerization. Increase in molecular weight results in higher zero shear and steady shear solution viscosity [376], unlike HEUR polymers where the exact opposite behavior is observed. In addition the authors observed that a reduction in molecular weight leads to a system where the Newtonian plateau is more extended. 3.3.3. Effect of external parameters The effect of pH and ionic strength on the rheology of the HASE polymer solutions has also been investigated. At relatively low pH, the HASE polymers are insoluble given the uncharged nature of the carboxylic acid groups [181]. Due the presence of methylacrylic acid in the backbone of the HASE polymer alkaline pH will ionize these groups and will lead to swelling of the polymer and the formation of a polyelectrolyte with solubility in water [169,181,367,368,370]. The swelling of the polymer is a result of the electrostatic repulsion between the polyelectrolyte chains [169,368,379]. As the pH is increased a smaller number of larger hydrophobic micro-domains are formed, which leads to an increase in the solution viscosity [181]. The expansion of the polymer chain is partly a result of the dissolution of the particle (present prior to the solubilization) [181]. In order to investigate the association mechanisms of HASE polymers the hydrophobic micro-domains of polymer 51 (5) were probed using

pyrene [181]. At a pH of 5.7 a large number of hydrophobic micro-domains are present. As the pH of the polymer solution is raised a smaller number of larger hydrophobic domains are formed. The authors suggest a transition from intramolecular to intermolecular hydrophobic associations. This leads to a signicant increase in the solution viscosity. Adding salt to the polymer solution affected its rheological properties. Increase in the salt concentration resulted in a decrease of the solution viscosity with the effect being more pronounced for groups with higher hydrophobicity. However, the polymer 51 (with C20 H41 as the hydrophobe) displayed a shear thickening behavior at a given salt concentration [368]. The rst effect, signicant reduction in the solution viscosity [379] with the addition of salt, is attributed to the shielding of the charges on the polymer backbone leading to the shrinkage of the polymer backbone. The intermolecular hydrophobic associations will be destroyed due to the shrinkage. The second effect, shear thickening with the addition of more salt, is attributed to the reformation of the network. The authors [379] have provided a model illustrating the network structures in salt solution under shear deformation (Fig. 35). Several different studies [91,106,375,382384] have been published on the interactions between HASE polymers and different surfactants. Polymers 51 (5) and 53 have been shown to interact with surfactants. Depending on the nature of the surfactant different behaviors can arise [91]. Using a partially water-soluble nonionic surfactant (nonylphenol ethoxylate) English et al. [91] demonstrated 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) 15581628

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. The use of water-soluble nonionic surfactant, at high sufcient surfactant concentration, resulted in the loss of shear induced structuring of the polymer network. It was further demonstrated that, when using water-soluble surfactants, segregation occurs where two distinct systems exists, one phase rich in polymer and one rich in surfactant. On the contrary, by using a partially soluble surfactant no evidence of complete segregation was found. A few studies [382384]demonstrated the increase in viscosity of the polymeric solution, employing polymer 51 (5), when using nonionic surfactant (C12 EO4 ). The increase in the viscosity was noticed up to a surfactant concentration of 0.1 M. The studies attributed the increase in solution viscosity to the formation of mixed micelles of surfactant and the hydrophobic groups of the HASE polymer chains. At surfactant concentrations above 0.01 M stronger interactions arise, thus a strengthening of the polymer network, where the formation of a stiff gel-like structure occurs. At this stage a bilayer is formed by the surfactant molecules, which are stabilized as large vesicles. A schematic presentation, as provided by Tirtaatmadja et al. [384], of the interactions between the nonionic surfactant, at different surfactant concentration, and the HASE polymer is given in Fig. 36. The temperature affects the interactions between the surfactant and HASE polymers. At low to moderate (nonionic) surfactant concentration the solution shear viscosity decreases with increasing temperature. Above a certain shear rate a substantial decrease in viscosity is observed, which can be related to the characteristic relaxation time associated with the dissociation of hydrophobes from hydrophobic domains. The rate of dissociation is much higher than the rate of association [385]. However this concept, i.e. the relation between characteristic relaxation time and exit rate of hydrophobic groups, as proposed by Aubry and Moan is controversial and needs verication [383]. Tam et al. [106] investigated the interactions between polymer 51 (5) and the nonionic surfactant (C12 EO23 ). The addition of the nonionic surfactant resulted in the strengthening of the polymer network by the increase of the number of intermolecular hydrophobic associations similar to when C12 EO4 was used as the nonionic surfactant. According to the study [106] the polymer-surfactant interactions are entropically driven. Interactions between ionic surfactants and HASE polymers have also been demonstrated [373,382]. The positive effect, i.e. increase in solution viscosity, of the surfactant

on the rheological property arises due to the same effect as with nonionic surfactant. Seng et al. [382] investigated the interactions between polymer 51 (5) and the SDS surfactant. Above a given critical surfactant concentration (CS*), closed to the CMC of the surfactant, the aforementioned positive effect is lost due to the disruption of the polymer network [382]. This leads to a reduction in the number of hydrophobic micro-domains. In addition free cations of the surfactant molecules screen the anionic charges on the HASE polymer backbone, which reduces the electrostatic repulsion between chains. A model representing the different molecular interactions between a ionic surfactant and polymer 51 (5) has been provided by Seng et al. [382] (Fig. 37). The interactions of SDS with polymer 51 (5) can be divided into three regions (1, 2 and 3 in Fig. 37). In regions 1 and 2 the effects of the polymersurfactant interactions on the rheology are limited. However, in region 3 weakening and destruction of the polymer network arise due to on one hand shielding of the charges (less electrostatic repulsion and thus decrease in hydrodynamic volume) and on the other hand the isolation of the hydrophobes in micelles thus suppressing intermolecular hydrophobic associations. Both of these effects lead to a lower solution viscosity. The same interaction behavior is observed with SDS and polymer 53 [373]. This in marked contrast to the behavior of nonionic surfactants discussed earlier. Nevertheless the same behavior, loss in solution viscosity at high surfactant concentration, has also been reported for a nonionic surfactant (polyoxyethylene) [106]. 3.4. Hydrophobically modied cellulose derivatives The current demand for environmental-friendly materials is relevant also for EOR applications. Hydrophobically modied hydroxyethyl cellulose (HM-HEC), hydrophobicallyhydrophilically modied hydroxyethyl cellulose (HHM-HEC), hydrophobically modied hydroxypropyl cellulose (HM-HPC) or hydrophobically modied ethyl hydroxyethyl cellulose (HM-EHEC) polymers can meet these demands. The parent cellulose polymer is a nonionic polysaccharide, which possesses anti-bacterial properties [386]. Parent cellulose polymers are hydrophobically or/and hydrophilically modied to produce the HM-HEC, HHM-HEC, HM-EHEC or HM-HPC polymers. A general structure of hydrophobically modied cellulosic polymer is presented in Fig. 38.

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

1611

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

A general overview of the reported systems is given in Table 4. A synthesis method, described by Miyajima et al. [406], for HM-HEC polymers involves a one-pot synthesis. A reactor is charged with the HEC polymer and isopropyl alcohol together with an aqueous solution (48%) of sodium hydroxide. This slurry is stirred at room temperature for 30 min under a nitrogen atmosphere after which the hydrophobic reagent is introduced. The reaction is then performed for several hours (e.g. 8 h) at 80 C. Methods for the synthesis of different HM-EHEC polymers have been developed and provided by Bostrm et al. [407] and Landoll [98]. These comprise several steps. First the cellulose is isolated from wood pulp using standard techniques, i.e. using a 21.5 wt% aqueous sodium hydroxide solution for 30 min. Then hydroxyethyl groups are introduced by reaction with ethylene oxide at 50 C for 75 min. The last step involves the incorporation of the hydrophobic group by reaction at 105 C for 120 min. The presence of hydrophobic groups on the polymer backbone of the polysaccharides will cause the formation of associations just like the HMPAM, HEUR and HASE analogues. The network structure is enhanced due to these groups and the hydrophobic associations arise at very low polymer concentration [118,399,405,407409]. Polymers 56, 57, 58, 61 and 63 are classic examples of this behavior with associations commencing at very low polymer concentrations. When good solvents are used, e.g. alcohols or dioxane, no associations arise [118]. Polymers 56, 57, 58 and 63 impart a stronger network when compared to their unmodied analogues. The type of hydrophobic

group also affects the strength of the network. Increasing the hydrophobicity will increase the strength. Polymer 57 and 60 (R1 = 2) are examples where the polymer with uorocarbon provide a stronger network than the hydrocarbon (un)modied analogue [391]. The distribution of the hydrophobic groups is also of crucial importance. Polymer 62 represents an example in which, due to inhomogeneous distribution of the hydrophobic group, two separate phases of polymers are observed [401]. One containing highly substituted chains and another with less substitution. The highly substituted chains will be preferentially incorporated into the network. Classical polyelectrolyte behavior can be obtained by the incorporation of charged moieties in the backbone of the polysaccharides. An example of this behavior is observed with polymer 60 (R1 = 3) where cationic moieties are incorporated into the backbone of the polysaccharide [397].

Fig. 38. General chemical structure of hydrophobically modied cellulosic polymer.

3.4.1. Rheological properties Although different types of hydrophobically modied polysaccharides have been synthesized their rheological properties in aqueous solution depend on the modication rather than on the type of polysaccharide used. The hydrophobic modications introduced on the different cellulose derivatives decrease the solubility of the polysaccharide. Several studies have demonstrated the pseudoplasticity of hydrophobically modied polysaccharides. Typical shear thinning is observed similar to the hydrophobically modied 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 addition shear thickening behavior is observed. The presence of shear thickening only arises within a certain concentration range of approximately 0.150.5 wt%. The zero-shear solution viscosity of a polymer solution is much more dependent on the concentration in the lower concentration regime. Increasing the polymer concentration also decreases the dependency of the zero-shear viscosity on concentration [410]. Examples of the shear thickening

1612

Table 4 Structure of different water-soluble polymers, HMHEC. Polymer HHM-HEC Structure References

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

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

[106]

HHM-HEC, R2 = nC18 H37 and R3 = 1 S-HEC, R2 = H and R3 = 1 R-HEC, R2 = nC18 H37 and R3 = H HM-HEC

Commercial product, produced by different suppliers.

[386]

Table 4 (Continued) Polymer HM-HEC Structure References

The hydrophobic groups are uorocarbon chains

[386389] D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

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) 15581628

R3 , R4 and R5 = alkyl, aryl, aralkyl, alkaryl, cycloalkyl, alkoxyalkyl and alkoxyaryl where a 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.

[392394]

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, ourocarbon groups R1 = 3, cationic groups R1 = 3 and 4, cationic and hydrophobic groups randomly distributed R1 = 5, cationic hydrophobic groups

Table 4 (Continued) Polymer HM-HPC Structure References

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

The polysaccharide is labeled with pyrene (1) or uorene (2) groups for uorescence studies

[395,396]

HM-EHEC

The hydrophobic groups are either long alkyl chains or a nonylphenol group

[397399]

1615

1616

HM-HEC-py-1

The hydrophobic groups are an alkyl chain and a pyrene group for uorescence studies

[118,400]

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

HM-CMC

The hydrophobic groups are amino-terminated poly(N-isopropylacrylamide)

[307,401404]

HM-Alginate

The hydrophobic groups are n-octylamines

[405]

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

1617

behavior are polymers 55 [386,387] and 59 [410,411] and an example of a polymer which does not display shear thickening due to a high concentration is polymer 58 [392394]. Thixotropic behavior is also observed for the hydrophobically modied polysaccharides. Similar to other hydrophobically modied polymers, such as HMPAM, hysteresis effects can be seen for polymeric solution employing polymer 55 (HHM-HEC, R-HEC) [387], 56 [412], 58 [392394], 59 (hydrophobic groups are hexadecyl moieties) and 60 (R1 = 2, R1 = 5) [397,399]. The viscosity of the polymeric solution at increasing shear rates is higher compared to the viscosity of the same solution when the shear rates decrease. According to Maestro et al. [413] hydrophobically modied polysaccharides display thixotropic behavior due to the diffusion controlled migration of hydrophobes to micelles. 3.4.2. Effect of inherent parameters The presence of hydrophobic groups leads to an increase in the solution viscosity of HM-polysaccharides. Similar to other hydrophobically modied polymers associations between the hydrophobic groups arise which lead to a strong polymer network. The behavior, signicant increase in the zero-shear viscosity, was noticed for solutions employing either polymer 58 [392394] or polymer 64 [405]. As mentioned before, the incorporation of charged moieties affects the network structure of HHMpolysaccharides. Therefore the rheological properties are also affected. In principle the thickening capability of the charged polysaccharide is higher compared to the uncharged analogue [386]. The HHM-HEC polymer 55 is an example of such a charged polysaccharide where sulfonic groups are incorporated into the backbone. 3.4.3. Effect of external parameters In principle the presence of salt does not affect the rheological properties of HM-polysaccharides solutions. However, it has been demonstrated that the presence of salt leads to a stronger hydrophobic network and thus an increase in the solution viscosity. Polymer 57 is an example of this behavior [391]. When charged moieties are incorporated into the backbone of the HM-polysaccharides, screening of the electrostatic repulsion arises in the presence of electrolyte. In classic polyelectrolytes, this screening leads to a collapse of the polymer coils and a signicant reduction in the solution viscosity. However, for the polysaccharides modied with hydrophobic groups (in addition to the charged groups), this behavior is not observed. Indeed, the presence of salt will weaken the electrostatic repulsion. However this leads to the enhancement of the hydrophobic associations and thus a strengthening of the polymer network. In addition to the presence of salt, the type of salt is also important for the observed behavior. According to Picton and Muller [412], the presence of salts that are able to induce a structuring of the water molecules, will lead to an enhancement of the viscosity due to the increase number of hydrophobic association. However using other types of salts, e.g. potassium thiocyanate, the enhancement

in solution viscosity is not as pronounced [412]. Solutions containing polymer 56 display this behavior. The ratio between the hydrophobic/hydrophilic substitution signicantly affects the solution properties at low polymer concentration [389]. The thickening capability is increased when this ratio is high. Polymer 55 is an example of a polymer whose thickening capability increases with a high hydrophobic/hydrophilic substitution [387,388]. Fig. 39 displays the effect of the presence of salt on the polymer network that has been envisaged by Kawakami et al. [387]. As the concentration of the salt increases, the electrostatic repulsion is reduced. This leads to a reinforcement of the hydrophobic associations and thus a stronger polymer network. Akiyama et al. [389] agree up to a point with the model presented in Fig. 39. However, an expansion is introduced in that the salt effect is also dependent on the hydrophobic/hydrophilic ratio. When the hydrophobic microdomains of the polymer network are saturated, i.e. when there are no free hydrophobes for association, the effect (reinforcement of the hydrophobic association) depicted in the model will not occur. The addition of salt will lead to a reduction of the electrostatic repulsion but the hydrophobic associations will not be reinforced given their strength prior to the addition of the salt. On the other hand if the hydrophobic microdomains are not saturated, i.e. when there are free hydrophobes, than the model presented in Fig. 39 can be held valid. The solution hysteresis (as a function of the shear rate) is affected by the presence of salt. Increasing the salt concentration will suppress the tendency of the solution to display hysteresis, and increasing the salt concentration beyond a certain limit no hysteresis behavior is observed. An example of such a solution is the solution with polymer 55 (R-HEC) [387]. The temperature affects the viscosity of hydrophobically modied polysaccharides. In most cases the viscosity is reduced as the temperature is increased. This behavior has been observed with polymer 55 and 62 (R = 1) [402]. In addition, the aggregation of hydrophobes is also affected by temperature. By using polymer 61, Winnik [414] demonstrated that upon increasing the temperature the hydrophobic associations are less stable, which leads to a weaker polymer network. The dependence of the viscosity on temperature was also observed with polymer 56 [415]. The shear thickening and shear thinning remained the same but the viscosity values changed as a function of temperature. Similar to HMPAM, HEUR and HASE polymers hydrophobically modied polysaccharides also display interactions with different types surfactants, i.e. anionic, cations or nonionic. The surfactants that have been demonstrated to interact with hydrophobically modied polysaccharides are listed in Table 5. The interaction between HM-HEC and a surfactant leads to different behaviors depending on the concentration of the surfactant and the type of surfactant. When using the anionic surfactant SDS, the surfactant molecule will interact with the hydrophobic groups forming mixed micelles. The formation of these mixed micelles will increase the number of hydrophobic microdomains and will

1618

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

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 polymer network [307,415,420,423]. Increasing the surfactant concentration further will lead to electrostatic repulsions between the mixed micelles and subsequently a drop in the solution viscosity of the ternary (waterpolymersurfactant) system [423]. The viscosity of the polymeric solution employing polymer 55, 57, 62 (R=1) or 65 increases and passes through a maximum as the concentration of the anionic SDS surfactant is increased [408,416,432]. The reduction in solution viscosity at elevated surfactant concentration can be counteracted by increasing the surfactant aggregation number in the mixed micelles. This can be achieved by either adding screening electrolytes or oppositely charged surfactants [433]. A polymersurfactant solution employing a surfactant with a lower CMC will reach a maximum solution viscosity at lower surfactant concentration compared to a surfactant with a higher CMC [430]. The zero-shear viscosity of a polymer solution passes through a maximum with the addition of a nonionic surfactant (Cm EOn ), similar to the behavior with an anionic surfactant [415]. Increasing the temperature of the polymersurfactant solution leads at rst to a reduction in the solution viscosity. However, beyond this region a considerable increase in the viscosity is observed until the cloud point temperature of the solution is reached. This increase in solution viscosity is attributed to the interchain association caused by micellar aggregation [415].

The surfactant binding mechanism to the hydrophobes of HM-EHEC depends on the surfactant concentration. It was demonstrated that at low surfactant concentration the hydrophobic microdomains act as nucleation sites onto which the surfactants preferentially absorb [426,428]. An increase in the surfactant concentration leads to a transition from non-cooperative binding to a more cooperative binding [417,426] similar to the unmodied analogue (EHEC) in polymersurfactant solution [426]. In addition, the data indicate that the number of hydrophobic microdomains is constant with respect to the surfactant concentration. According to Evertsson and Nilsson [417], the strongest interaction, in terms of the highest microviscosity, is achieved when the surfactant binds in non-cooperative way at very low adsorption degree of SDS. The polymer concentration also has an effect on the type of interaction that is observed between surfactants and HM-EHEC. With the addition of surfactant (SDS) to the polymeric solution employing polymer 62 (R = 1), Lauten and Nystrm [419] demonstrated that the association complexes between the polymer and surfactant will grow with time. These complexes will eventually sedimentate due to their size. This effect is observed at low SDS concentration at both high and low polymer concentration. At high SDS concentration the intermolecular associations are dissolved and no time dependence is observed for the system [419]. By using cloud point experiments, Thuresson et al. [425] demonstrated that the solubility of the

Table 5 The different surfactants shown to interact with hydrophobically modied polysaccharides. Surfactant Sodium dodecyl sulfate Sodium decyl sulfate Sodium tetradecyl sulfate Sodium dodecyl-di(ethyleneoxide)-sulfate Sodium alkanoate Sodium octyl benzene sulfonate Cetyltrimethylammonium bromide Tetradecyltrimethyl ammonium bromide Dodecylammonium chloride Dodecyltrimethylammonium chloride Dodecyltrimethylammonium bromide Hexadecyltrimethylammonium bromide Hexadecyltrimethylammonium acetate Alkyl betainate Octylphenol adduct Polyoxyethylene alkyl-ether Abbreviation SDS (NaC12 S) NaC10 S NaC14 S (NaC12 [EO]2 S) CH3 [CH2 ]n COONa, n = 4, 6, 8 and 10 SOBS CTAB TTAB C12 ACl C12 TACl C12 TABr C16 TABr C16 TAAc CH3 [CH2 ]n COOCH2 N[CH3 ]3 , n = 9, 11, 13 C8 H17 C6 H4 O[EO]10 H (Triton X-100) Cm Hn , m = 11, 12 and 13/n = 4 Type Anionic Anionic Anionic Anionic Anionic Anionic Cationic Cationic Cationic Cationic Cationic Cationic Cationic Cationic Nonionic Nonionic Source [117,307,408,416428] [117] [424] [424] [427] [427,429] [430] [420,430] [424] [424] [424] [424] [424] [431] [418] [117,415,420]

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628

1619

hydrophobically modied polymer is suppressed at low SDS concentrations compared to its unmodied analogue. HM-EHEC also interacts with different surfactant in a different manner. The addition of ionic surfactants leads in most cases to an enhancement of the hydrophobic associations at moderate surfactant concentration. Increasing the surfactant concentration further leads to the disruption of the hydrophobic micro-domains and thus of the polymer network. According to Kjonikson et al. [434] the polymersurfactant system transits from a heterogeneous system at low surfactant concentration to a homogeneous one at high surfactant concentration. The interactions of nonionic surfactants with HM-EHEC, i.e. polymer 62 (R = 1), are weaker when compared to ionic surfactants [427]. Nevertheless, for both ionic and nonionic surfactants the length of the surfactant chain is crucial for the solution behavior. If this length is increased the disruption of the network, i.e. loss of solution viscosity, is postponed due to an increase in the aggregation number of the mixed micelles [427]. One striking behavior is the inuence of a spacer (EOspacer) between the hydrophobic group and the polymer backbone on the solution viscosity [434] in the presence of surfactants. At high surfactant concentrations, the behavior of the polymer with or without the spacer is similar to its unmodied non-hydrophobic analogue. The viscosity of the polymeric solution employing the HM-EHEC with spacers of 4 EO units with no surfactant is much lower compared to the one without spacers. However, the authors [434] correctly mention the positive effect a spacer should have on the viscosity given the studies of Hwang and HogenEsch [241] and Tam et al. [380]. Therefore the authors [434] attribute the discrepancy to the difference in distribution of the hydrophobes along the polymer backbone. The presence of salt does not affect the rheological properties of a solution containing a HM-polysaccharide and a surfactant. At intermediate salt concentration no change in the solution viscosity is observed. However a further increase in the salt concentration leads to a solution viscosity which is higher than for a solution without the surfactant. 4. Conclusion This review clearly indicates that the successful design of new water-soluble polymers for a given application requires an integral multiscale and multidisciplinary approach. Proper denition of the required product properties is in this case crucial. Knowledge of polymer chemical architecture (and thus of the synthetic methods used) must be conceptually linked to the desired product application requirements. In this case viscosity measurements under different shear conditions are of paramount importance and should be ideally correlated with the nature (i.e. polymer architecture and overall chemical composition) of the corresponding water solution. The inuence of external parameters (e.g. pH and temperature) on the rheological behavior must be coupled with an in depth knowledge of the relationship between the chemical structure and architecture of the polymer and the rheological behavior. In this respect, an overall correlation cannot be dened only as a function of the

chemical/molecular structure. Rheological properties will be affected by a combination of external parameters and the chemical nature and molecular structure of the polymer. For instance, the rheological properties of an aqueous solution of an amphiphilic polyelectrolyte are similar to those of an unmodied analogue without amphiphilic moieties. However, in the presence of salt a markedly different behavior is observed. The solution viscosity of the unmodied polymer decreases with increasing salt concentration, whereas the solution viscosity of the aqueous solution containing the amphiphilic polyelectrolyte is not affected. Another good example is the thermal performance of amphiphilic polymers: the rheological properties of an aqueous solution containing the amphiphilic polymer or its unmodied analogue, i.e. without the NIPAM monomer, are quite similar. However, when exposing both solutions to higher temperatures signicant differences arise. The effect of temperature on the solution viscosity of an aqueous solution containing the amphiphilic polymer is limited, whereas the viscosity of the unmodied analogue changes signicantly. Although there are many different water-soluble polymers capable of enhancing the solution viscosity, it is important to understand their differences and analogies. Different polymers display in general differences in the agglomeration principles governing their water behavior. On a molecular level the basic principle is indeed quite general: the presence of relatively weak inter(macro)molecular interactions (e.g. hydrophobic association and hydrogen bonding) factually increases the molecular weight of the polymer coils. As a consequence the solution viscosity increases. However, a careful balance must be observed here since predominantly weak interactions (both in terms of strength and number thereof) do not result in observable rheological differences while excessively strong ones might compromise the solubility of the system by leading for example to gel formation.

Acknowledgement This work is part of the Research Programme of the Dutch Polymer Institute DPI, Eindhoven, the Netherlands. References
[1] Thomas S. Enhanced oil recoveryan overview. Oil Gas Sci Technol 2008;63:919. [2] Han D-K, Yang C-Z, Zhang Z-Q, Lou Z-H, Chang Y-I. Recent development of enhanced oil recovery in China. J Pet Sci Eng 1999;22:1818. [3] Li G, Zhai L, Xu G, Shen Q, Mao H. Current tertiary oil recovery in China. J Disper Sci Technol 2000;21:367408. [4] Lake LW. Enhanced oil recovery. Englewood Cliffs, NJ: Prentice-Hall Inc.; 1989. [5] Sorbie KS. Polymer-improved oil recovery. Boca Raton, FL: CRC Press; 1991. [6] Levitt DB, Pope GA. Selection and screening of polymers for enhanced-oil recovery. In: SPE Improved Oil Recovery Symp. 2008. p. 118, 113845. [7] Gaillard N, Giovannetti B, Favero C. Improved oil recovery using thermally and chemically protected compositions based on co- and ter-polymers containing acrylamide. In: SPE Improved Oil Recovery Symp. 2010. p. 111, 129756.

1620

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628 [35] Sukpisan J, Kanatharana J, Sirivat A, Wang Y. The specic viscosity of partially hydrolyzed polyacrylamide solutions: effects of degree of hydrolysis, molecular weight, solvent quality and temperature. J Polym Sci Part B Polym Phys 1998;36:74353. [36] Khune GD, Donaruma LG, Hatch MJ, Kilmer NH, Shepitka JS, Martin FD. Modied acrylamide polymers for enhanced oil recovery. J Appl Polym Sci 1985;30:87585. [37] McCormick CL, Neidlinger HH, Hester RD, Wildman GC. Synthetic random and graft copolymers for utilization in enhanced oil recoverysynthesis and rheology. In: Shah DO, editor. Surface phenomena in enhanced oil recovery. 1st ed. New York: Plenum; 1981. p. 74172. [38] Dautzenberg H. Polyelectrolyte complex formation in highly aggregating systems. 1. Effect of salt: polyelectrolyte complex formation in the presence of NaCl. Macromolecules 1997;30:78105. [39] Peng S, Wu C. Light scattering study of the formation and structure of partially hydrolyzed poly(acrylamide)/calcium(II) complexes. Macromolecules 1999;32:5859. [40] Ohmine I, Tanaka T. Salt effects on the phase transition of ionic gels. J Chem Phys 1982;77:57259. [41] Ben Jar P-Y, Wu YS. Effect of counter-ions on swelling and shrinkage of polyacrylamide-based ionic gels. Polymer 1997;38:255760. [42] Cook Jr RL, King HE, Peiffer DG. High-pressure viscosity of dilute polymer solutions in good solvents. Macromolecules 1992;25:292834. [43] Martin FD, Hatch MJ, Shepitka JS, Ward JS. Improved water-soluble polymers for enhanced recovery of oil. In: SPE Ann Tech Confr. 1983. p. 114, 11786. [44] Martin FD. Mechanical degradation of polyacrylamide solutions in core plugs from several carbonate reservoirs. SPE Formation Eval 1986;1:13950. [45] Lewandowska K. Comparative studies of rheological properties of polyacrylamide and partially hydrolyzed polyacrylamide solutions. J Appl Polym Sci 2007;103:223541. [46] Hu Y, Wang SQ, Jamieson AM. Rheological and rheooptical studies of shear-thickening polyacrylamide solutions. Macromolecules 1995;28:184753. [47] Seright RS, Seheult M, Talashek T. Injectivity characteristics of EOR polymers. SPE Reserv Eval Eng 2009;12:78392. [48] Chauveteau G. Molecular interpretation of several different properties of ow of coiled polymer solutions through porous media in oil recovery conditions. In: SPE Ann Tech Confr. 1981. p. 114, 10060. [49] Choplin L, Sabati J. Threshold-type shear-thickening in polymeric solutions. Rheol Acta 1986;25:5709. [50] Ferguson J, Walters K, Wolff C. Shear and extensional ow of polyacrylamide solutions. Rheol Acta 1990;29:5719. [51] Shepitka JS, Case CE, Donaruma LG, Hatch MJ, Kilmer NH, Khune GD, Martin FD, Ward JS, Wilson KV. Partially imidized, water-soluble polymeric amides. 1. Partially imidized polyacrylamide and polymethacrylamide. J Appl Polym Sci 1983;28:36117. [52] Durst F, Haas R, Interthal W. The nature of ows through porous media. J Non-Newtonian Fluid Mech 1987;22:16989. [53] James D, MacLaren DR. Laminar-ow of dilute polymer-solutions through porous media. J Fluid Mech 1975;70:73352. [54] Magueur A, Moan M, Chauveteau G. Effect of successive contractions and expansions on the apparent viscosity of dilute polymer solutions. Chem Eng Commun 1985;36:35166. [55] Silberberg A, Mijnlieff PF. Study of reversible gelation of partially neutralized poly(methacrylic acid) by viscoelastic measurements. J Polym Sci Part A2 Polym Phys 1970;8:1089. [56] Quadrat O. Negative thixotropy in polymer solutions. Adv Colloid Interface Sci 1985;24:4575. [57] Ohya S, Matsuo T. Time-dependent change of viscosity in polymethacrylic acid solution. J Colloid Interface Sci 1979;68:5935. [58] Bradna P, Quadrat O, Dupuis D. The inuence of salt concentration on negative thixotropy in solutions of partially hydrolyzed polyacrylamide. Colloid Polym Sci 1995;273:4215. [59] Quadrat O, Bradna P, Dupuis D, Wolff C. Negative thixotropy of solutions of partially hydrolyzed polyacrylamide. Part I. The inuence of shear rate on time changes of ow characteristics. Colloid Polym Sci 1992;270:10579. [60] Bradna P, Quadrat O, Dupuis D. Negative thixotropy of solutions of partially hydrolyzed polyacrylamide. Part II. The inuence of glycerol content and degree of ionization. Colloid Polym Sci 1995;273:6427. [61] Bradna P, Quadrat O, Titkova L, Dupuis D. The inuence of multivalent cations on negative thixotropy in aqueous glycerol solutions of partially hydrolyzed polyacrylamide. Acta Polym 1997;48:4469.

[8] Wu Y, Wang K-S, Hu Z, Bai B, Shuler P, Tang Y. A new method for fast screening of long-term thermal stability of water soluble polymers for reservoir conformance control. In: SPE Improved Oil Recovery Symp. 2009. p. 111, 124257. [9] Pancharoen M, Thiele MR, Kovscek AR. Inaccessible pore volume of associative polymer oods. In: SPE Improved Oil Recovery Symp. 2010. p. 115, 129910. [10] Buchgraber M, Clemens T, Castanier LM, Kovscek AR. The displacement of viscous oil by associative polymer solutions. In: SPE Ann Tech Confr. 2009. p. 119, 122400. [11] Zhang L-J, Yue X-A. Displacement of polymer solution on residual oil trapped in dead ends. J Cent South Univ Tech 2008;15:847. [12] Zhang L-J, Yue X-A, Guo F. Micro-mechanisms of residual oil mobilization by viscoelastic uids. Pet Sci 2008;5:5661. [13] Wang D, Xia H, Liu Z, Anda Q, Yang Q. Study of the mechanism of polymer solution with visco-elastic behavior increasing microscopic oil displacement efciency and the forming of steady oil thread ow channels. In: SPE Asia Pacic Oil Gas Confr. 2001. p. 19, 68723. [14] Yin H, Wang D, Zhong H. Study on ow behaviors of viscoelastic polymer solution in micropore with dead end. In: SPE Ann Tech Confr. 2006. p. 110, 101950. [15] Wang D, Cheng J, Yang Q, Gong W, Li Q, Chen F. Viscous-elastic polymer can increase microscale displacement efciency in cores. In: SPE Ann Tech Confr. 2000. p. 110, 63227. [16] Xia H, Ju Y, Kong F, Wu J. Effect of elastic behavior of HPAM solutions on displacement efciency under mixed wettability conditions. In: SPE Ann Tech Confr. 2004. p. 18, 90234. [17] Glass JE, editor. Hydrophilic polymers. Performance with environmental acceptance. Advances in chemistry series 248. Washington, DC: American Chemical Society; 1996. [18] Schulz DN, Glass JE, editors. Polymers as rheology modiers. ACS symposium series no. 462. Washington, DC: American Chemical Society; 1991. [19] Shalaby SW, McCormick CL, Butler GB, editors. Water soluble polymers. Synthesis, solution properties and applications. ACS symposium no. 467. Washington, DC: American Chemical Society; 1991. [20] Glass JE, editor. Polymers in aqueous media: performance through association. Advances in chemistry series 223. Washington, DC: American Chemical Society; 1989. [21] Morgan SE, McCormick CL. Water-soluble polymers in enhanced oil recovery. Prog Polym Sci 1990;15:10345. [22] Sabhapondit A, Borthakur A, Haque I. Characterization of acrylamide polymers for enhanced oil recovery. J Appl Polym Sci 2003;87:186978. [23] Sabhapondit A, Borthakur A, Haque I. Water soluble acrylamidomethyl propane sulfonate (AMPS) copolymer as an enhanced oil recovery chemical. Energy Fuels 2003;17:6838. [24] Song H, Zhang S-F, Ma X-C, Wang D-Z, Yang J-Z. Synthesis and application of starch-graft-poly(AM-co-AMPS) by using a complex initiation system of CS-APS. Carbohydr Polym 2007;69:18995. [25] Vega I, Snchez L, DAccorso N. Synthesis and characterization of copolymers with 1,3-oxazolic pendant groups. React Funct Polym 2008;68:23341. [26] Borthakur A, Rahman M, Sarmah A, Subrahmanyam B. Partially hydrolysed polyacrylamide for enhanced oil recovery. Res Ind 1995;40:904. [27] Stokes RJ, Evans DF. Fundamentals of interfacial engineering. New York: Wiley-VCH; 1997. [28] Shupe RD. Chemical stability of polyacrylamide polymers. J Pet Technol 1981;33:151329. [29] Fuoss RM, Strauss UP. Polyelectrolytes. 2. Poly-4-vinylpyridonium chloride and poly-4-vinyl-N-N-butylpyridonium bromide. J Polym Sci 1948;3:24663. [30] Fuoss RM, Strauss UP. Electrostatic interaction of polyelectrolytes and simple electrolytes. J Polym Sci 1948;3:6023. [31] Fuoss RM. Viscosity function for polyelectrolytes. J Polym Sci 1948;3:6034. [32] Ait-Kadi A, Carreau PJ, Chauveteau G. Rheological properties of partially hydrolyzed polyacrylamide solutions. J Rheol 1987;31:53761. [33] Dupuis D, Lewandowski FY, Steiert P, Wolff C. Shear thickening and time-dependent phenomena: the case of polyacrylamide solutions. J Non-Newtonian Fluid Mech 1994;54:1132. [34] Ellwanger RE, Jaeger DA, Barden RE. Use of the empirical hill equation for characterization of the effect of added cations on the viscosity of aqueous solutions of partially hydrolyzed polyacrylamide. Polym Bull 1980;3:36974.

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628 [62] Ghannam MT, Esmail MN. Rheological properties of aqueous polyacrylamide solutions. J Appl Polym Sci 1998;69:158797. [63] Xin X, Xu G, Gong H, Bai Y, Tan Y. Interaction between sodium oleate and partially hydrolyzed polyacrylamide: a rheological study. Colloids Surf A 2008;326:19. [64] Methemitis C, Morcellet M, Sabbadin J, Francois J. Interactions between partially hydrolyzed polyacrylamide and ionic surfactants. Eur Polym J 1986;22:61927. [65] Leela JK, Sharma G. Studies on xanthan production from Xanthomonas campestris. Bioprocess Eng 2000;23:6879. [66] Nasr S, Soudi MR, Haghighi M. Xanthan production by a native strain of X. campestris and evaluation of application in EOR. Pak J Biol Sci 2007;10:30103. [67] Garca-Ochoa F, Santos VE, Casas JA, Gmez E. Xanthan gum: production, recovery, and properties. Biotechnol Adv 2000;18:54979. [68] Becker A, Katzen F, Puhler A, Ielpi L. Xanthan gum biosynthesis and application: a biochemical/genetic perspective. Appl Environ Microbiol 1998;50:14552. [69] Morris ER, Rees DA, Young G, Walkinshaw MD, Darke A. Orderdisorder transition for a bacterial polysaccharide in solutionrole for polysaccharide conformation in recognition between Xanthomonas pathogen and its plant host. J Mol Biol 1977;110:116. [70] Dentini M, Crescenzi V, Blasi D. Conformational properties of xanthan derivatives in dilute aqueous-solution. Int J Biol Macromol 1984;6:938. [71] Norton IT, Goodall DM, Frangou SA, Morris ER, Rees DA. Mechanism and dynamics of conformational ordering in xanthan polysaccharide. J Mol Biol 1984;175:37194. [72] Holzwarth G. Conformation of extracellular polysaccharide of Xanthomonas-campestris. Biochemistry 1976;15:43339. [73] Chen CSH, Sheppard EW. Conformation and shear stability of xanthan gum in solution. Polym Eng Sci 1980;20:5126. [74] Yoshida T, Tanner RD, editors. Bioproducts and bioprocess. Berlin: Springer-Verlag; 1993. [75] Richardson RK, Rossmurphy SB. Nonlinear viscoelasticity of polysaccharide solutions. 2. Xanthan polysaccharide solutions. Int J Biol Macromol 1987;9:25764. [76] Kierulf C, Sutherland IW. Thermal-stability of xanthan preparations. Carbohydr Polym 1988;9:18594. [77] Lambert F, Rinaudo M. On the thermal-stability of xanthan gum. Polymer 1985;26:154953. [78] Seright RS, Henrici BJ. Xanthan stability at elevated temperatures. SPE Reserv Eng 1990;1:5260. [79] Wellington SL. Bio-polymer solution viscosity stabilization polymer degradation and antioxidant use. Soc Pet Eng J 1983;23:90112. [80] Ash SG, Clarke-Sturman AJ, Calvert R, Nisbet TM. Chemical stability of biopolymer solutions. In: SPE Ann Tech Confr. 1983. p. 18, 12085. [81] Sutherland IW. An enzyme system hydrolysing the polysaccharides of Xanthomonas species. J Appl Bacteriol 1982;53:38593. [82] Cadmus MC, Jackson LK, Burton KA, Plattner RD, Slodki ME. Biodegradation of xanthan gum by Bacillus sp. Appl Environ Microbiol 1982;44:511. [83] Bragg JR, Maruca SD, Gale WW, Gall LS, Wernau WC, Beck D, Goldman IM, Laskin AI, Naslund LA. Control of xanthan-degrading organisms in the loudon pilot: approach, methodology, and results. In: SPE Ann Tech Confr. 1983. p. 112, 11989. [84] Hou CT, Barnabe N, Greaney K. Biodegradation of xanthan by salttolerant aerobic microorganisms. J Ind Microbiol 1986;1:317. [85] Taugbol K, Ly TV, Austad T. Chemical ooding of oil-reservoirs. 3. Dissociative surfactantpolymer interaction with a positive effect on oil-recovery. Colloids Surf A 1995;103:8390. [86] Austad T, Taugbol K. Chemical ooding of oil-reservoirs. 2. Dissociative surfactantpolymer interaction with a negative effect on oil-recovery. Colloids Surf A 1995;103:7381. [87] Taylor KC, Nasr-El-Din HA. Water-soluble hydrophobically associating polymers for improved oil recovery: a literature review. J Pet Sci Eng 1998;19:26580. [88] Hill A, Candau F, Selb J. Properties of hydrophobically associating polyacrylamides: inuence of the method of synthesis. Macromolecules 1993;26:452132. [89] Annable T, Buscall R, Ettelaie R, Whittlestone D. The rheology of solutions of associating polymer: comparison of experimental behavior with transient network theory. J Rheol 1993;37:695726. [90] Panmai S, Prudhomme RK, Peiffer DG. Rheology of hydrophobically modied polymers with spherical and rod-like surfactant micelles. Colloids Surf A 1999;147:315.

1621

[91] English RJ, Laurer JH, Spontak RJ, Khan SA. Hydrophobically modied associative polymer solutions: rheology and microstructure in the presence of nonionic surfactants. Ind Eng Chem Res 2002;41:642535. [92] Biggs S, Selb J, Candau F. Effect of surfactant on the solution properties of hydrophobically modied polyacrylamide. Langmuir 1992;8:83847. [93] Sarkar N, Kershner LD. Rigid rod water-soluble polymers. J Appl Polym Sci 1996;62:393408. [94] Maia AMS, Borsali R, Balaban RC. Comparison between a polyacrylamide and a hydrophobically modied polyacrylamide ood in a sandstone core. Mater Sci Eng C 2009;29:5059. [95] Dubin PL, Strauss UP. Hydrophobic bonding in alternating copolymers of maleic acid and alkyl vinyl ethers. J Phys Chem 1970;74:28427. [96] Dubin PL, Strauss UP. Hydrophobic hypercoiling in copolymers of maleic acid and alkyl vinyl ethers. J Phys Chem 1967;71:27579. [97] Schwab FC, Sheppard EW, Chen CSH. Waterooding process employing cationic-nonionic copolymers. US Pat 4,110,232; 1978. [98] Landoll LM. Hydrophobically modied polymers. US Pat 4,529,523; 1984. [99] Bock J, Pace SJ, Schulz DN. Enhanced oil recovery with hydrophobically associating polymers containing N-vinyl pyrrolidone functionality. US Pat 4,709,759; 1987. [100] Bock J, Valint PL, Pace SJ. Enhanced oil recovery with hydrophobically associating polymers containing sulfonate functionality. US Pat 4,702,319; 1987. [101] Bock J, Siano DB, Pace SJ. Enhanced oil recovery with hydrophobically associating polymers. Can Pat 1300362; 1992. [102] Evani S. Enhanced oil recovery process using a hydrophobic associative composition containing a hydrophilic/hydrophobic polymer. US Pat 4,814,096; 1989. [103] McCormick CL, Nonaka T, Johnson CB. Water-soluble copolymers. 27. Synthesis and aqueous solution behavior of associaacrylamide/N-alkylacrylamide copolymers. Polymer tive 1988;29:7319. [104] Lara-Ceniceros AC, Rivera-Vallejo C, Jimenez-Regalado EJ. Synthesis, characterization and rheological properties of three different associative polymers obtained by micellar polymerization. Polym Bull 2007;58:42533. [105] Tam KC, Jenkins RD, Winnik MA, Bassett DR. A structural model of hydrophobically modied urethane-ethoxylate (HEUR) associative polymers in shear ows. Macromolecules 1998;31:414959. [106] Tam KC, Seng WP, Jenkins RD, Bassett DR. Rheological and microcalorimetric studies of a model alkali-soluble associative polymer (HASE) in nonionic surfactant solutions. J Polym Sci Part B Polym Phys 2000;38:201932. ois J. Hydrophobi[107] Feng Y, Billon L, Grassl B, Khoukh A, Franc cally associating polyacrylamides and their partially hydrolyzed derivatives prepared by post modication. 1. Synthesis and characterization. Polymer 2002;43:205564. [108] Argillier J-F, Audibert A, Lecourtier J, Moan M, Rousseau L. Solution and adsorption properties of hydrophobically associating watersoluble polyacrylamides. Colloids Surf A 1996;113:24757. [109] Volpert E, Selb J, Candau F. Inuence of the hydrophobe structure on composition, microstructure, and rheology in associating polyacrylamides prepared by micellar copolymerization. Macromolecules 1996;29:145263. [110] Ezzell SA, McCormick CL. Water-soluble copolymers. 39. Synthesis and solution properties of associative acrylamido copolymers with pyrenesulfonamide uorescence labels. Macromolecules 1992;25:18816. [111] Ezzell SA, Hoyle CE, Creed D, McCormick CL. Water-soluble copolymers. 40. Photophysical studies of the solution behavior of associative pyrenesulfonamide-labeled polyacrylamides. Macromolecules 1992;25:188795. [112] Klucker R, Candau F, Schosseler F. Transient behavior of associating copolymers in shear ow. Macromolecules 1995;28:641622. [113] Dowling KC, Thomas JK. A novel micellar synthesis and photophysical characterization of water-soluble acrylamidestyrene block copolymers. Macromolecules 1990;23:105964. [114] Schulz DN, Kaladas JJ, Maurer JJ, Bock J, Pace SJ, Schulz WW. Copolymers of acrylamide and surfactant macromonomers: synthesis and solution properties. Polymer 1987;28:21105. [115] Peiffer DG. Hydrophobically associating polymers and their interactions with rod-like micelles. Polymer 1990;31:235360. [116] Iliopoulos I, Wang TK, Audebert R. Viscometric evidence of interactions between hydrophobically modied poly(sodium acrylate) and sodium dodecyl sulfate. Langmuir 1991;7:6179.

1622

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628 mer solutions: thermothinning versus thermothickening. Macromolecules 2005;38:851221. Mortensen K, Brown W, Jrgensen E. Phase behavior poly(propylene oxide)poly(ethylene oxide)poly(propylene oxide) triblock copolymer melt and aqueous solutions. Macromolecules 1994;27:565466. Nishinari K, Hofmann KE, Moritaka H, Kohyama K, Nishinari N. Gelsol transition of methylcellulose. Macromol Chem Phys 1997;198:121726. Doelker E. Cellulose derivatives. Adv Polym Sci 1993;107:199265. Sarkar N. Thermal gelation properties of methyl and hydroxypropyl methylcellulose. J Appl Polym Sci 1979;24:107387. Yoshioka H, Mikami M, Mori Y, Tsuchida E. A synthetic hydrogel with thermoreversible gelation. 1. Preparation and rheological properties. J Macromol Sci Part A Pure Appl Chem 1994;31: 11320. Nystrm B, Walderhaug H, Hansen FK. Dynamic light scattering and rheological studies of thermoreversible gelation of a poly(ethylene oxide)poly(propylene oxide)poly(ethylene oxide) triblock copolymer in aqueous solution. Faraday Discuss 1995;101:33544. Wang G, Lindell K, Olofsson G. On the thermal gelling of ethyl(hydroxyethyl)cellulose and sodium dodecyl sulfate. Phase behavior and temperature scanning calorimetric response. Macromolecules 1997;30:10512. Loyen K, Iliopoulos I, Audebert R, Olsson U. Reversible thermal gelation in polymer/surfactant systems. Control of the gelation temperature. Langmuir 1995;11:10536. Sarrazin-Cartalas A, Iliopoulos I, Audebert R, Olsson U. Association and thermal gelation in mixtures of hydrophobically modied polyelectrolytes and nonionic surfactants. Langmuir 1994;10: 14216. Greenhill-Hooper MJ, OSullivan TP, Wheele PA. The aggregation behavior of octadecylphenylalkoxysulfonates. 1. Temperature dependence of the solution behavior. J Colloid Interface Sci 1988;124:7787. Bokias G, Hourdet D, Iliopoulos I, Staikos G, Audebert R. Hydrophobic interactions of poly(N-isopropylacrylamide) with hydrophobically modied poly(sodium acrylate) in aqueous solution. Macromolecules 1997;30:82937. Hourdet D, LAlloret F, Audebert R. Synthesis of thermoassociative copolymers. Polymer 1997;38:253547. Hourdet D, LAlloret F, Audebert R. Reversible thermothickening of aqueous polymer solutions. Polymer 1994;35:262430. de Vos S, Mller M, Visscher K, Mijnlieff PF. Synthesis and characterization of poly(acrylamide)-graft-poly(ethylene oxideco-propylene oxide). Polymer 1994;35:264450. Hutchinson BH, McCormick CL. Water-soluble copolymers. 15. Studies of random copolymers of acrylamide with N-substituted acrylamides by C-13 NMR. Polymer 1986;27:6236. Magny B, Lafuma F, Iliopoulos I. Determination of microstructure of hydrophobically modied water-soluble polymers by C-13 NMR. Polymer 1992;33:31514. Newman JK, McCormick CL. Water-soluble copolymers. 51. Copolymer compositions of high-molecular-weight functional acrylamido water-soluble polymers using direct-polarization magic-angle-spinning C-13 nuclear-magnetic-resonance. Polymer 1994;35:9358. McCormick CL, Chen GS, Hutchinson BH. Water-soluble copolymers. 5. Compositional determination of random copolymers of acrylamide with sulfonated co-monomers by infraredspectroscopy and C-13 nuclear magnetic-resonance. J Appl Polym Sci 1982;27:310320. Newman JK, McCormick CL. Water-soluble copolymers. 53. Na-23 NMR-studies of hydrophobically-modied polyacidscopolymers of 2-(1-naphthylacetamido)ethylacrylamide with acrylic-acid and methacrylic-acid. Macromolecules 1994;27:51238. Newman JK, McCormick CL. Water-soluble copolymers. 52. Na23 NMR-studies of ion-binding to anionic polyelectrolytespoly(sodium 2-acrylamido-2-methylpropanesulfonate), poly(sodium 3-acrylamido-3-methylbutanoate), poly(sodium acrylate), and poly(sodium galacturonate). Macromolecules 1994;27:511422. Furo I, Iliopoulos I, Stilbs P. Structure and dynamics of associative water-soluble polymer aggregates as seen by F-19 NMR spectroscopy. J Phys Chem B 2000;104:48594. Walderhaug H, Hansen FK, Abrahmsen S, Persson K, Stilbs P. Associative thickenersNMR self-diffusion and rheology studies of aqueous-solutions of hydrophobically-modied poly(oxyethylene) polymers. J Phys Chem 1993;97:833642.

[117] Senan C, Meadows J, Shone PT, Williams PA. Solution behavior of hydrophobically modied sodium polyacrylate. Langmuir 1994;10:24719. [118] Tanaka R, Meadows J, Williams PA, Phillips GO. Interaction of hydrophobically modied (hydroxyethyl)cellulose with various added surfactants. Macromolecules 1992;25:130410. [119] Winnik FM. Association of hydrophobic polymers in water: uorescence studies with labeled (hydroxypropyl)celluloses. Macromolecules 1989;22:73442. [120] Durand A, Hourdet D. Synthesis and thermoassociative properties in aqueous solution of graft copolymers containing poly(Nisopropylacrylamide) side chains. Polymer 1999;40:494151. [121] Maia AMS, Costa M, Borsali R, Garcia RB. Rheological behavior and scattering studies of acrylamide-based copolymer solutions. Macromol Symp 2005;229:21727. [122] Lundberg DJ, Brown RG, Glass JE, Eley RR. Synthesis, characterization, and solution rheology of model hydrophobically-modied, water-soluble ethoxylated urethanes. Langmuir 1994;10:302734. [123] Winnik MA, Yekta A. Associative polymers in aqueous solution. Curr Opin Colloid Interface Sci 1997;2:42436. [124] Huldn M. Hydrophobically modied urethane-ethoxylate (HEUR) associative thickeners. 1. Rheology of aqueous solutions and interactions with surfactants. Colloids Surf A 1994;82:26377. [125] Xu B, Li L, Zhang K, MacDonald PM, Winnik MA, Jenkins R, Basset D, Wolf D, Nuyken O. Synthesis and characterization of comb associative polymers based on poly(ethylene oxide). Langmuir 1997;13:6896902. [126] Rao B, Uemura Y, Dyke L, Mcdonald PM. Self-diffusion coefcients of hydrophobic ethoxylated urethane associating polymers using pulsed-gradient spin-echo nuclear magnetic resonance. Macromolecules 1995;28:5318. [127] Alami E, Almgren M, Brown W. Interaction of hydrophobically end-capped poly(ethylene oxide) with nonionic surfactants in aqueous solution. Fluorescence and light scattering studies. Macromolecules 1996;29:502635. [128] Alami E, Almgren M, Brown W. Aggregation of hydrophobically end-capped poly(ethylene oxide) in aqueous solutions. Fluorescence and light-scattering studies. Macromolecules 1996;29:222943. [129] Yekta A, Duhamel J, Brochard P, Adiwidjaja H, Winnik MA. A uorescent probe study of micelle-like cluster formation in aqueous solutions of hydrophobically modied poly(ethylene oxide). Macromolecules 1993;26:182936. [130] Maechling-Strasser C, Clouet F, Francois J. Hydrophobically end-capped polyethylene-oxide urethanes: 2. Modelling their association in water. Polymer 1992;33:10215. [131] Maechling-Strasser C, Francois J, Clouet F. Hydrophobically endcapped poly(ethylene oxide) urethanes. 1. Characterization and experimental study of their association in aqueous solution. Polymer 1992;33:62736. [132] Xie X, Hogen-Esch TE. Copolymers of N,N-dimethylacrylamide and 2-(N-ethylperuorooctanesulfonamido)ethyl acrylate in aqueous media and in bulk. Synthesis and properties. Macromolecules 1996;29:173445. [133] Abu-Sharkh BF, Yahaya GO, Ali SA, Kazi IW. Solution and interfacial behavior of hydrophobically modied water-soluble block copolymers of acrylamide and N-phenethylacrylamide. J Appl Polym Sci 2001;82:46776. [134] Volpert E, Selb J, Candau F. Associating behaviour of polyacrylamides hydrophobically modied with dehexylacrylamide. Polymer 1998;39:102533. [135] Candau F, Selb J. Hydrophobically-modied polyacrylamides prepared by micellar polymerization. Adv Colloid Interface Sci 1999;79:14972. [136] Yahya GO, Ali SA, Alnaafa MA, Hamad EZ. Preparation and viscosity behavior of hydrophobically-modied poly(vinyl alcohol) (PVA). J Appl Polym Sci 1995;57:34352. [137] Yahya GO, Hamad EZ. Solution behaviour of sodium maleate/1alkene copolymers. Polymer 1995;36:370510. [138] Kopperud HM, Hansen FH, Nystrm B. Effect of surfactant and temperature on the rheological properties of aqueous solutions of unmodied and hydrophobically modied polyacrylamide. Macromol Chem Phys 1998;199:238594. [139] Shaikh S, Asrof Ali SK, Hamad EZ, Abu-Sharkh BF. Synthesis and solution properties of poly(acrylamide-styrene) block copolymers with high hydrophobic content. Polym Eng Sci 1999;39: 19628. [140] Hourdet D, Gadgil J, Podhajecka K, Badiger MV, Brlet A, Wadgaonkar PP. Thermoreversible behavior of associating poly-

[141]

[142]

[143] [144] [145]

[146]

[147]

[148]

[149]

[150]

[151]

[152] [153] [154]

[155]

[156]

[157]

[158]

[159]

[160]

[161]

[162]

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628 [163] Nystrom B, Walderhaug H, Hansen FK. Dynamic crossover effects observed in solutions of a hydrophobically associating watersoluble polymer. J Phys Chem 1993;97:774352. [164] Abrahmsn-Alami S, Stilbs P. 1 H NMR self-diffusion and multield 2H spin relaxation study of model associative polymer and sodium dodecyl sulfate aggregation in aqueous solution. J Phys Chem 1994;98:635967. [165] Persson K, Abrahmsen S, Stilbs P, Hansen FK, Walderhaug H. The association of urethane-polyethyleneoxide (Heur) thickeners. As studied by NMR self-diffusion measurements. Colloid Polym Sci 1992;270:4659. [166] McCormick CL, Elliott DL. Water-soluble copolymers. 14. Potentiometric and turbidimetric studies of water-soluble copolymers of acrylamidecomparison of carboxylated and sulfonated copolymers. Macromolecules 1986;19:5427. [167] Wang C, Tam KC, Jenkins RD, Bassett DR. Potentiometric titration and dynamic light scattering of hydrophobically modied alkali soluble emulsion (HASE) polymer solutions. Phys Chem Chem Phys 2000;2:196772. [168] Wang C, Tam KC, Jenkins RD. Dissolution behavior of HASE polymers in the presence of salt: potentiometric titration, isothermal titration calorimetry, and light scattering studies. J Phys Chem B 2002;106:1195204. [169] Dai S, Tam KC, Jenkins RD, Bassett DR. Light scattering of dilute hydrophobically modied alkali-soluble emulsion solutions: effects of hydrophobicity and spacer length of macromonomer. Macromolecules 2000;33:70218. [170] Chassenieux C, Nicolai T, Durand D. Association of hydrophobically end-capped poly(ethylene oxide). Macromolecules 1997;30:49528. [171] Prochazka K, Martin TJ, Webber SE, Munk P. Onion-type micelles in aqueous media. Macromolecules 1996;29:652630. [172] Prochazka K, Martin TJ, Munk P, Webber SE. Polyelectrolyte poly(tert-butyl acrylate)-block-poly(2-vinylpyridine) micelles in aqueous media. Macromolecules 1996;29:651825. [173] Biggs S, Selb J, Candau F. Copolymers of acrylamide/Nalkylacrylamide in aqueous solution: the effects of hydrolysis on hydrophobic interactions. Polymer 1993;34:58091. [174] Biggs S, Hill A, Selb J, Candau F. Copolymerization of acrylamide and a hydrophobic monomer in an aqueous medium: effect of the surfactant on the copolymer microstructure. J Phys Chem 1992;96:150511. [175] Branham KD, Davis DL, Middleton JC, McCormick CL. Water-soluble polymers. 59. Investigation of the effects of polymer microstructure on the associative behaviour of amphiphilic terpolymers of acrylamide, acrylic acid and N-[(4-decyl)phenyl]acrylamide. Polymer 1994;35:442936. [176] Branham KD, Snowden HS, McCormick CL. Water-soluble copolymers. 64. Effects of pH and composition on associative properties of amphiphilic acrylamide/acryl acid terpolymers. Macromolecules 1996;29:25462. [177] Branham KD, Shafer GS, Hoyle CE, McCormick CL. Watersoluble copolymers. 61. Microstructural investigation of pyrenesulfonamide-labeled polyelectrolytes. Variation of label proximity utilizing micellar polymerization. Macromolecules 1995;28:617582. [178] Kramer MC, Steger JR, Hu Y, McCormick CL. Water-soluble copolymers. 65. Environmentally responsive associations probed by nonradiative energy transfer studies of naphthalene and pyrene-labeled poly(acrylamide-co-sodium 11-(acrylamido)undecanoate). Macromolecules 1996;29:19927. [179] Kramer MC, Welch CG, Steger JR, McCormick CL. Water-soluble copolymers. 63. Rheological and photophysical studies on the associative properties of pyrene-labeled poly[acrylamide-co-sodium 11-(acrylamido)undecanoate]. Macromolecules 1995;28:524854. [180] Araujo E, Rharbi Y, Huang X, Winnik MA, Bassett DR, Jenkins RD. Pyrene excimer kinetics in micelle like aggregates in a 0-HASE associating polymer. Langmuir 2000;16:866471. [181] Kumacheva E, Rharbi Y, Winnik MA, Guo L, Tam KC, Jenkins RD. Fluorescence studies of an alkaline swellable associative polymer in aqueous solution. Langmuir 1997;13:1826. [182] Ruer C, Collet A, Viguier M, Oberdisse J, Mora S. SDS interactions with hydrophobically end-capped poly(ethylene oxide) studied by 13 C NMR and SANS. Macromolecules 2009;42:522635. [183] Hu Y, Kramer MC, Boudreaux CJ, McCormick CL. Water-soluble copolymers. 62. Nonradiative energy-transfer studies of pHresponsive and salt-responsive associations in hydrophobicallymodied, hydrolyzed maleic anhydride-ethyl vinyl ether copolymers. Macromolecules 1995;28:71006.

1623

[184] Hu YX, Smith GL, Richardson MF, McCormick CL. Water soluble polymers. 74. pH responsive microdomains in labeled poly(sodium maleate-alt-ethyl n-octylamide-substituted vinyl ethers): synthesis, steady-state uorescence, and nonradiative energy transfer studies. Macromolecules 1997;30: 352637. [185] Smith GL, McCormick CL. Water-soluble polymers. 78. Viscosity and NRET uorescence studies of pH-responsive twin-tailed associative terpolymers based on acrylic acid and methacrylamide. Macromolecules 2001;34:91824. [186] Cathbras N, Collet A, Viguier M, Berret J-F. Synthesis and linear viscoelasticity of uorinated hydrophobically modied ethoxylated urethanes (F-HEUR). Macromolecules 1998;31:130511. [187] Kaczmarski JP, Glass JE. Synthesis and solution properties of hydrophobically-modied ethoxylated urethanes with variable oxyethylene spacer lengths. Macromolecules 1993;26: 514956. [188] May R, Kaczmarski JP, Glass JE. Inuence of molecular weight distributions on HEUR aqueous solution rheology. Macromolecules 1996;29:474553. [189] Zhang HS, Pan J, Hogen-Esch TE. Synthesis and characterization of one-ended peruorocarbon-functionalized derivatives of poly(ethylene glycol)s. Macromolecules 1998;31:281521. [190] McCormick CL, Hester RD, Morgan SE, Saeddine AM. Watersoluble copolymers. 31. Effects of molecular-parameters, solvation, and polymer associations on drag reduction performance. Macromolecules 1990;23:21329. [191] McCormick CL, Hester RD, Morgan SE, Saeddine AM. Watersoluble copolymers. 30. Effects of molecular-structure on drag reduction efciency. Macromolecules 1990;23:212431. [192] McCormick CL, Blackmon KP. Water-soluble copolymers. 21. Copolymers of acrylamide with 2-acrylamido-2methylpropanedimethylammonium chloride: synthesis and characterization. Polymer 1986;27:19715. [193] McCormick CL, Blackmon KP, Elliot DL. Water-soluble copolymers. 22. Copolymers of acrylamide with 2-acrylamido-2methylpropanedimethylammonium chloride: aqueous solution properties of a polycation. Polymer 1986;27:197680. [194] McCormick CL, Blackmon KP. Water-soluble copolymers. 17. Copolymers of acrylamide with sodium 3-methacrylamido-3methylbutanoate: synthesis and characterization. Macromolecules 1986;19:15125. [195] McCormick CL, Elliot DL, Blackmon KP. Water-soluble copolymers. 18. Copolymers of acrylamide with sodium 3-methacrylamido-3methylbutanoate: microstructural studies and solution properties. Macromolecules 1986;19:151622. [196] Kathmann EE, White LA, McCormick CL. Water-soluble polymers. 73. Electrolyte- and pH-responsive zwitterionic copolymers 4-[(2-acrylamido-2-methylpropyl)-dimethylammonio]of butanoate with 3-[(2-acrylamido-2-methylpropyl)dimethylammonio]propanesulfonate. Macromolecules 1997;30:5297304. [197] McCormick CL, Salazar LC. Water soluble copolymers. 46. Hydrophilic sulphobetaine copolymers of acrylamide and 3-(2-acrylamido-2-methylpropanedimethyl-ammonio)-1propanesulphonate. Polymer 1992;33:461724. [198] Kathmann EE, White LA, McCormick CL. Water soluble polymers. 69. pH and electrolyte responsive copolymers of acrylamide and the zwitterionic monomer 4-(2-acrylamido-2-methylpropyldimethylammonio) butanoate: synthesis and solution behaviour. Polymer 1997;38:8718. [199] Kathmann EE, White LA, McCormick CL. Water soluble polymers. 70. Effects of methylene versus propylene spacers in the pH and electrolyte responsiveness of zwitterionic copolymers incorporating carboxybetaine monomers. Polymer 1997;38:87986. [200] Kathmann EE, McCormick CL. Water-soluble polymers. 72. Synthesis and solution behavior of responsive copolymers of acrylamide and the zwitterionic monomer 6-(2-acrylamido-2methylpropyldimethylammonio) hexanoate. J Polym Sci Part A Polym Chem 1997;35:24353. [201] Kathmann EEL, Davis DD, McCormick CL. Water-soluble polymers. 60. Synthesis and solution behavior of terpolymers of acrylic acid, acrylamide and the zwitterionic monomer 3-[(2-acrylamido2-methylpropyl)dimethylammonio]-1-propanesulfonate. Macromolecules 1994;27:315661. [202] Kathmann EE, McCormick CL. Water-soluble polymers. 71. pH responsive behavior of terpolymers of sodium acrylate, acrylamide, and the zwitterionic monomer 4-(2-acrylamide-2methylpropanedimethylammonio)butanoate. J Polym Sci Part A Polym Chem 1997;35:23142.

1624

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628 isopropylacrylamide) on its association in water. Macromolecules 1998;31:252732. Zhang Y, Wu C, Fang Q, Zhang YX. A light scattering study of the aggregation behavior of uorocarbon modied polyacrylamides in water. Macromolecules 1996;29:24947. Zhang YX, Fang Q, Fu YQ, Da AH, Zhang Y, Wu C, Hogen-Esch TE. Synthesis and characterization of uorocarbon-modied poly(Nisopropylacrylamide). Polym Int 2000;49:76374. Zhou H, Song GQ, Zhang YX, Chen JY, Jiang M, Hogen-Esch TE, Dieing R, Ma L, Haeussling L. Hydrophobically modied polyelectrolytes. 4. Synthesis and solution properties of uorocarboncontaining poly(acrylic acid). Macromol Chem Phys 2001;202: 305764. Armentrout RS, McCormick CL. Water soluble polymers. 76. Electrolyte responsive cyclopolymers with sulfobetaine units exhibiting polyelectrolyte or polyampholyte behavior in aqueous media. Macromolecules 2000;33:41924. Armentrout RS, McCormick CL. Water-soluble polymers. 77. Amphoteric cyclopolymers with sulfobetaine units: phase behavior in aqueous media and solubilization of p-cresol in microdomains. Macromolecules 2000;33:294451. Chang YH, McCormick CL. Water-soluble copolymers. 57. Amphiphilic cyclocopolymers of diallylalkoxybenzylmethylammonium chloride and diallyl-dimethylammonium chloride. Polymer 1994;35:350312. McCormick CL, Middleton JC, Cummins DF. Water-soluble copolymers. 37. Synthesis and characterization of responsive hydrophobically modied polyelectrolytes. Macromolecules 1992;25:12016. McCormick CL, Middleton JC, Grady CE. Water soluble copolymers. 38. Synthesis and characterization of electrolyte responsive terpolymers of acrylamide, N-(4-butyl)phenylacrylamide, and sodium acrylate, sodium-2-acrylamido-2-methylpropanesulphonate or sodiu-acrylamido-3-methylbutanoate. Polymer 1992;33:418490. Ali SA, Umar Y, bu-Sharkh BF, Al-Muallem HA. Synthesis and comparative solution properties of single-, twin-, and triple-tailed associating ionic polymers based on diallylammonium salts. J Polym Sci Part A Polym Chem 2006;44:548094. Umar Y, Al-Muallem HA, bu-Sharkh BF, Ali SA. Synthesis and solution properties of hydrophobically associating ionic polymers made from diallylammonium salts/sulfur dioxide cyclocopolymerization. Polymer 2004;45:365161. Ali SA, Umar Y, Al-Muallem HA, bu-Sharkh BF. Synthesis and viscosity of hydrophobically modied polymers containing dendritic segments. J Appl Polym Sci 2008;109:178192. Umar Y, bu-Sharkh BF, Ali SA. The effects of charge densities on the associative properties of a pH-responsive hydrophobically modied sulfobetaine/sulfur dioxide terpolymer. Polymer 2005;46:1070917. Ringsdorf H, Venzmer J, Winnik FM. Fluorescence studies of hydrophobically modied poly(N-isopropylacrylamides). Macromolecules 1991;24:167886. Wang TK, Iliopoulos I, Audebert R. Aqueous-solution behavior of hydrophobically modied poly(acrylic acid). In: Shalaby SW, McCormick CL, Butler GB, editors. Water soluble polymers. ACS symposium series no. 467. Washington, DC: American Chemical Society; 1991. p. 21831. Xue W, Hamley IW, Castelletto V, Olmsted PD. Synthesis and characterization of hydrophobically modied polyacrylamides and some observations on rheological properties. Eur Polym J 2004;40:4756. Yahaya GO, Ahdab AA, Ali SA, Abu-Sharkh BF, Hamad EZ. Solution behavior of hydrophobically associating water-soluble block copolymers of acrylamide and N-benzylacrylamide. Polymer 2001;42:336372. Reddy GJ, Naidu SV, Reddy AVR. Synthesis and characterization of phenyl methacrylamide copolymers. Adv Polym Technol 2006;25:4150. Zhang HP, Xu K, Ai H, Chen DH, Xv LL, Chen MC. Synthesis, characterization and solution properties of hydrophobically modied polyelectrolyte poly(AA-co-TMSPMA). J Solution Chem 2008;37:113748. Hwang FS, Hogen-Esch TE. Effects of water-soluble spacers on the hydrophobic association of uorocarbon-modied poly(acrylamide). Macromolecules 1995;28:332835. Zhang YX, Da AH, Butler GB, Hogen-Esch TE. A uorine-containing hydrophobically associating polymer. 1. Synthesis and solution properties of copolymers of acrylamide and uorine-containing acrylates or methacrylates. J Polym Sci Part A Polym Chem 1992;30:138391.

[203] Grassl B, Francois J, Billon L. Associating behaviour of polyacrylamide modied with a new hydrophobic zwitterionic monomer. Polym Int 2001;50:11629. [204] McCormick CL, Johnson CB. Water-soluble polymers: 33. Ampholytic terpolymers of sodium 2-acrylamido-2-methylpropanesulphonate with 2-acrylamido-2-methylpropanedimethylammonium chloride and acrylamide: Synthesis and aqueous solution behaviour. Polymer 1990;31:11007. [205] Mumick PS, Welch PM, Salazar LC, McCormick CL. Water-soluble copolymers. 56. Structure and solvation effects of polyampholytes in drag reduction. Macromolecules 1994;27:32331. [206] Mumick PS. Structurally tailored water soluble polymers for the study of drag reduction. PhD Dissertation, Hattiesburg MS: The University of Southern Mississippi; 1993. [207] Salazar LC. Synthesis and solution behavior of electrolyteresponsive polyampholytes. PhD Dissertation, Hattiesburg MS: The University of Southern Mississippi; 1991. [208] McCormick CL, Johnson CB. Water-soluble polymers. XXXIV. Ampholytic terpolymers of sodium 3-acrylamido-3methylbutanoate with 2-acrylamido-2-methylpropanedimethylammonium chloride and acrylamide: synthesis and absorbancy behavior. J Macromol Sci Part A Pure Appl Chem 1990;27:53947. [209] McCormick CL, Johnson CB. Water-soluble copolymers. 29. Ampholytic copolymers of sodium 2-acrylamido-2methylpropanesulfonate with (2-acrylamido-2-methylpropyl)dimethylammonium chloride: solution properties. Macromolecules 1988;21:6949. [210] McCormick CL, Johnson CB. Water-soluble polymers. 28. Ampholytic copolymers of sodium 2-acrylamido-2-methylwith (2-acrylamido-2-methylpropyl)propanesulfonate dimethylammonium chloride: synthesis and characterization. Macromolecules 1988;21:68693. [211] McCormick CL, Salazar LC. Water-soluble copolymers. 43. Ampholytic copolymers of sodium 2-(acrylamido)-2-methylpropanesulfonate with[2-(acrylamido)-2-methylpropyl]trimethylammonium chloride. Macromolecules 1992;25:1896900. [212] Petit F, Iliopoulos I, Audebert R, Szonyi S. Associating polyelectrolytes with peruoroalkyl side chains: aggregation in aqueous solution, association with surfactants, and comparison with hydrogenated analogues. Langmuir 1997;13:422933. [213] Shedge AS, Lele AK, Wadgaonkar PP, Hourdet D, Pcrrin P, Chassenieux C, Badiger MV. Hydrophobically modied poly(acrylic acid) using 3-pentadecylcyclohexylamine: synthesis and rheology. Macromol Chem Phys 2005;206:46472. [214] Tomatsu I, Hashidzume A, Yusa S, Morishima Y. Unique associative properties of copolymers of sodium acrylate and oligo(ethylene oxide) alkyl ether methacrylates in water. Macromolecules 2005;38:783744. [215] Noda T, Hashidzume A, Morishima Y. Effects of spacer length on the side-chain micellization in random copolymers of sodium 2-(acrylamido)-2-methylpropanesulfonate and methacrylates substituted with ethylene oxide-based surfactant moieties. Macromolecules 2001;34:130817. [216] Noda T, Hashidzume A, Morishima Y. Rheological properties of transient networks formed from copolymers of sodium acrylate and methacrylates substituted with amphiphiles: comparison with sodium 2-(acrylamido)-2-methylpropanesulfonate copolymers. Langmuir 2001;17:598491. [217] Zhong C, Luo P, Ye Z, Chen H. Characterization and solution properties of a novel water-soluble terpolymer for enhanced oil recovery. Polym Bull 2009;62:7989. [218] Smith GL, McCormick CL. Water-soluble polymers. 80. Rheological and photophysical studies of pH-responsive terpolymers containing hydrophobic twin-tailed acrylamide monomers. Macromolecules 2001;34:557986. [219] McCormick CL, Hoyle CE, Clark MD. Water-soluble copolymers. 26. Fluorescence probe studies of hydrophobically modied maleic acidethyl vinyl ether copolymers. Polymer 1992;33:2437. [220] Chang Y, McCormick CL. Water-soluble copolymers. 49. Effect of the distribution of the hydrophobic cationic monomer dimethyldodecyl(2-acrylamidoethyl)ammonium bromide on the solution behavior of associating acrylamide copolymers. Macromolecules 1993;26:61216. [221] Zhuang DQ, Da JCAH, Zhang YX, Dieing R, Ma L, Haeussling L. Hydrophobically modied polyelectrolytes II: synthesis and characterization of poly(acrylic acid-co-alkyl acrylate). Polym Adv Technol 2001;12:61625. [222] Zhang Y, Mei L, Fang Q, Zhang YX, Jiang M, Wu C. Effect of incorporating a trace amount of uorocarbon into poly(N-

[223]

[224]

[225]

[226]

[227]

[228]

[229]

[230]

[231]

[232]

[233]

[234]

[235]

[236]

[237]

[238]

[239]

[240]

[241]

[242]

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628 [243] Bastiat G, Grassl B, Francois J. Study of sodium dodecyl sulfate/poly(propylene oxide) methacrylate mixed micelles for the synthesis of thermo-associative polymers by micellar polymerization. Polym Int 2002;51:95865. [244] Lara-Ceniceros AC, Rivera-Vallejo C, Jimenez-Regalado EJ. Synthesis and characterization of telechelic polymers obtained by micellar polymerization. Polym Bull 2007;59:499508. [245] Jimenez-Regalado EJ, Cadenas-Pliego G, Perez-Alvarez M, Hernandez-Valdez Y. Study of three different families of watersoluble copolymers: synthesis, characterization and viscoelastic behavior of semidilute solutions of polymers prepared by solution polymerization. Polymer 2004;45:19932000. [246] McCormick CL, Hoyle CE, Clark MD. Water-soluble copolymers. 36. Photphysical investigations of water-soluble copolymers of 2-(1-naphthylacetamide)ethylacrylamide. Macromolecules 1991;24:2397403. [247] McCormick CL, Hoyle CE, Clark MD. Water-soluble copolymers. 35. Photophysical and rheological studies of the copolymer of methacrylic acid with 2-(1-naphthylacetyl)ethyl acrylate. Macromolecules 1990;23:31249. [248] Emmons WD, Stevens TE. Acrylamide copolymer thickener for aqueous systems. US Pat 4,395,524; 1983. [249] Emmons WD, Stevens TE. Acrylamid-copolymers suitable as thickening agent for aqueous systems and aqueous compositions containing them. Eur Pat 63018; 1982. [250] Kulicke W-M, Kniewske R, Klein J. Preparation, characterization, solution properties and rheological behaviour of polyacrylamide. Prog Polym Sci 1982;8:373468. [251] Evani S. Water-dispersible hydrophobic thickening agent. Eur Pat 57875; 1982. [252] Evani S. Water-dispersible hydrophobic thickening agent. US Pat 4,432,881; 1984. [253] Bock J, Siano DB, Kowalik RM, Turner SP. Process for forming acrylamide-alkyl acrylamide copolymers. Eur Pat 115213; 1984. [254] Turner SP, Siano DB, Bock J. Acrylamidealkylacrylamide copolymers. US Pat 4,520,182; 1985. [255] Turner SP, Siano DB, Bock J. Microemulsion process for producing acrylamide-alkyl acrylamide copolymers. US Pat 4,521,580;1985. [256] Turner SP, Siano DB, Bock J. Micellar process for the production of acrylamide-alkyl acrylamide copolymers. US Pat 4,528,348;1985. [257] Vittadello ST, Biggs S. Shear history effects in associative thickener solutions. Macromolecules 1998;31:76917. [258] Regalado EJ, Selb J, Candau F. Viscoelastic behavior of semidilute solutions of multisticker polymer chains. Macromolecules 1999;32:85808. [259] Shashkina YA, Zaroslov YD, Smirnov VA, Philippova OE, Khokhlov AR, Pryakhina TA, Churochkina NA. Hydrophobic aggregation in aqueous solutions of hydrophobically modied polyacrylamide in the vicinity of overlap concentration. Polymer 2003;44:228993. [260] Ma J-T, Huang R-H, Zhao L, Zhang X. Solution properties of ionic hydrophobically associating polyacrylamide with an arylalkyl group. J Appl Polym Sci 2005;97:31621. [261] Bock J, Valint PL. Process for preparing hydrophobically associating terpolymers containing sulfonate functionality. US Pat 4,730,028; 1988. [262] Valint PL, Bock J. Hydrophobically associating terpolymers containing sulfonate functionality. US Pat 5,089,578; 1992. [263] Bock J, Siano DB, Turner SP. Hydrophobically associating terpolymers of acrylamide, salts of acrylic acid and alkyl acrylamide. US Pat 4,694,046; 1987. [264] Schulz DN, Berluche E, Maurer JJ, Bock J. Tetrapolymers of N-vinyl pyrrolidone/acrylamide/salt of acrylic acid/N-alkyl acrylamide. US Pat 4,663,408; 1987. [265] Hill A, Candau F, Selb J. Aqueous solution properties of hydrophobically associating copolymers. Prog Colloid Polym Sci 1992;84:615. [266] Camail M, Margaillan A, Martin I. Copolymers of N-alkyl- and Narylalkylacrylamides with acrylamide: inuence of hydrophobic structure on associative properties. Part 1. Viscometric behaviour in dilute solution and drag reduction performance. Polym Int 2009;58:14954. [267] Camail M, Margaillan A, Martin I. Copolymers of N-alkyl- and Narylalkylacrylamides with acrylamide: inuence of hydrophobic structure on associative properties. Part 2. Rheological behaviour in semi-dilute solution. Polym Int 2009;58:15562. [268] Zhao Y, Zhou J, Xu X, Liu W, Zhang J, Fan M, Wang J. Synthesis and characterization of a series of modied polyacrylamide. Colloid Polym Sci 2009;287:23741. [269] Seery TAP, Yassini M, Hogen-Esch TE, Amis EJ. Static and dynamic light scattering characterization of solutions of hydrophobically

1625

[270]

[271]

[272]

[273]

[274]

[275]

[276]

[277]

[278]

[279]

[280]

[281]

[282]

[283]

[284] [285]

[286]

[287]

[288]

[289]

[290]

[291]

associating uorocarbon-containing polymers. Macromolecules 1992;25:478491. Zhang YX, Da AH, Hogen-Esch TE. A uorocarbon-containing hydrophobically associating polymer. J Polym Sci Part C Polym Lett 1990;28:2138. Amis EJ, Hu N, Seery TAP, Hogen-Esch TE, Hwang F. Associating polymers containing uorocarbon hydrophobic units. In: Glass JE, editor. Hydrophilic polymersperformance with environmental acceptance. Advances in chemistry series 248. Washington, DC: American Chemical Society; 1996. p. 279302. Umar Y, Abu-Sharkh BF, Ali SA. The effects of zwitterionic and anionic charge densities in polymer chains on the viscosity behavior of a pH-responsive hydrophobically modied ionic polymer. J Appl Polym Sci 2005;98:140411. Fevola MJ, Bridges JK, Kellum MG, Hester RD, McCormick CL. pH-responsive polyzwitterions: a comparative study of acrylamide-based polyampholyte terpolymers and polybetaine copolymers. J Appl Polym Sci 2004;94:2439. Johnson KM, Poe GD, Lockhead RY, McCormick CL. The synthesis of hydrophobically modied water-soluble polyzwitterionic copolymers and responsiveness to surfactants in aqueous solution. J Macromol Sci Part A Pure Appl Chem 2004;41:587611. Candau F, Biggs S, Hill A, Selb J. Synthesis, structure and properties of hydrophobically associating polymers. Prog Org Coat 1994;24:119. Branham KD, Middleton JC, McCormick CL. Photophysical and rheological properties of naphthalene-labeled water-soluble copolymers polymerized in surfactant solution. Polym Prepr Am Chem Soc 1991;32(1):106. Candau F, Volpert E, Lacik I, Selb J. Free-radical polymerization in micellar media: effect of microenvironment. Macromol Symp 1996;111:8594. Wang G-J, Engberts JBFN. Synthesis of hydrophobically and electrostatically modied polyacrylamides and their catalytic effects on the unimolecular decarboxylation of 6-nitrobenzisoxazole-3carboxylate anion. Langmuir 1995;11:385661. Vaskova V, Renoux D, Bernard M, Selb J, Candau F. Mechanism of copolymerization of acrylamide with a polymerizable surfactant. Polym Adv Technol 1995;6:44151. Renoux D, Selb J, Candau F. Aqueous solution properties of hydrophobically associating copolymers. Prog Colloid Polym Sci 1994;97:2137. Charalambopoulou A, Bokias G, Staikos G. Template copolymerisation of N-isopropylacrylamide with a cationic monomer: inuence of the template on the solution properties of the product. Polymer 2002;43:263743. Rainaldi I, Cristallini C, Ciardelli G, Giusti P. Copolymerization of acrylic acid and 2-hydroxyethyl methacrylate onto poly(Nvinylpyrrolidone): template inuence on comonomer reactivity. Macromol Chem Phys 2000;201:242431. Frisch HL, Xu QH. Copolymerization of styrene and methacrylicacid in the presence of poly(2-vinylpyridine) as the template. Macromolecules 1992;25:51459. Jantas R, Polowinski S. Copolymerization of a multiallyl monomer with styrene. Acta Polym 1989;40:2258. Polowinski S, Janowska G. Thermal copolymerization of acrylonitrile with methacrylate units arranged in matrix. Eur Polym J 1975;11:1835. Efng JJ, McLennan IJ, Kwak JCT. Associative phase separation observed in a hydrophobically modied poly(acrylamide)/sodium dodecyl sulfate system. J Phys Chem 1994;98:2499502. Efng JJ, McLennan IJ, van Os NM, Kwak JCT. 1 H NMR investigations of the interactions between anionic surfactants and hydrophobically modied poly(acrylamide)s. J Phys Chem 1994;98:12397402. McCormick CL, Salazar LC. Water-soluble copolymers. 45. Ampholytic terpolymers of acrylamide with sodium 3acrylamido-3-methylbutanoate and 2-acrylamido-2-methylpropanetrimethylammonium chloride. J Appl Polym Sci 1993;48:111520. Peiffer DG, Lundberg RD. Synthesis and viscometric properties of low charge-density ampholytic ionomers. Polymer 1985;26:105868. Petit-Agnely F, Iliopoulos I, Zana R. Hydrophobically modied sodium polyacrylates in aqueous solutions: association mechanism and characterization of the aggregates by uorescence probing. Langmuir 2000;16:99217. Leibler L, Rubinstein M, Colby RH. Dynamics of reversible networks. Macromolecules 1991;24:47017.

1626

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628 differential scanning calorimetry (DSC), and turbidity measurements. J Colloid Interface Sci 1996;179:2033. Emmons WD, Valley H, Stevens TE. Polyurethane thickeners in latex compositions. US Pat 4,079,028; 1978. Xu B, Yekta A, Li L, Masoumi Z, Winnik MA. The functionality of associative polymer networks: the association behavior of hydrophobically modied urethane-ethoxylate (HEUR) associative polymers in aqueous solution. Colloids Surf A 1996;112:23950. Yekta A, Xu B, Duhamel J, Adiwidjaja H, Winnik MA. Fluorescence studies of associating polymers in water: determination of the chain end aggregation number and a model for the association process. Macromolecules 1995;28:95666. Kaczmarski JP, Glass JE. Synthesis and characterization of stepgrowth hydrophobically-modied ethoxylated urethane associative thickeners. Langmuir 1994;10:303542. Liu F, Frere Y, Francois J. Association properties of poly(ethylene oxide) modied by pendant aliphatic groups. Polymer 2001;42:296983. Barmar M, Barikani M, Kaffashi B. Synthesis of ethoxylated urethane and modication with cetyl alcohol as thickener. Iran Polym J 2001;10:3315. Barmar M. Study of the effect of PEG length in uni-HEUR thickener behavior. J Appl Polym Sci 2009;111:17514. Barmar M, Barikani M, Kaffashi B. Steady shear viscosity study of various HEUR models with different hydrophilic and hydrophobic sizes. Colloids Surf A 2005;253:7782. Barmar M, Barikani M, Kaffashi B. The effect of molecular weight on the behaviour of step-growth hydrophobically modied ethoxylated urethane (S-G HEUR) end-capped with dodecyl alcohol. Iran Polym J 2004;13:2416. Vorobyova O, Yekta A, Winnik MA, Lau W. Fluorescent probe studies of the association in an aqueous solution of a hydrophobically modied poly(ethylene oxide). Macromolecules 1998;31:89989007. Wang YC, Winnik MA. Onset of aggregation for water-soluble polymeric associative thickenersa uorescence study. Langmuir 1990;6:14379. Fonnum G, Bakke J, Hansen FK. Associative thickeners. Part 1. Synthesis, rheology and aggregation behavior. Colloid Polym Sci 1993;271:3809. Xu B, Li L, Yekta A, Masoumi Z, Kanagalingam S, Winnik MA, Zhang K, MacDonald PM, Menchen S. Synthesis, characterization, and rheological behavior of polyethylene glycols end-capped with uorocarbon hydrophobes. Langmuir 1997;13:244756. Calvet D, Collet A, Viguier M, Berret J-F, Srro Y. Peruoroalkyl end-capped poly(ethylene oxide). Synthesis, characterization, and rheological behavior in aqueous solution. Macromolecules 2003;36:44957. Srro Y, Aznar R, Porte G, Berret J-F, Calvet D, Collet A, Viguier M. Associating polymers: from Flowers to transient networks. Phys Rev Lett 1998;82:55847. Hartmann P, Collet A, Viguier M. Synthesis and characterization of model uoroacylated poly(ethylene oxide). J Fluorine Chem 1999;95:14551. Zhang HS, Hogen-Esch TE, Boschet F, Margaillan A. Complex formation of beta-cyclodextrin- and peruorocarbon-modied water-soluble polymers. Langmuir 1998;14:49727. Boschet F, Branger C, Margaillan A, Condamine E. Synthesis, characterisation and aqueous behaviour of a one-ended peruorocarbonmodied poly(ethylene glycol). Polymer 2002;43:532934. Hu YZ, Zhao CL, Winnik MA, Sundararajan PR. Fluorescence studies of the interaction of sodium dodecyl-sulfate with hydrophobically modied poly(ethylene oxide). Langmuir 1990;6:8803. Wetzel WH, Chen M, Glass JE. Associative thickeners. An overview with an emphasis on synthetic procedures. In: Glass JE, editor. Hydrophilic polymersperformance with environmental acceptance. Advances in chemistry series 248. Washington, DC: American Chemical Society; 1996. p. 16379. Yekta A, Nivaggioli T, Kanagalingam S, Xu B, Masoumi Z, Winnik MA. Urethane-coupled poly(ethylene glycol) polymers containing hydrophobic end groups: NMR characterization as a step toward determining aggregation numbers in aqueous solutions. In: Glass JE, editor. Hydrophilic polymersperformance with environmental acceptance. Advances in chemistry series 248. Washington, DC: American Chemical Society; 1996. p. 36376. Yekta A, Duhamel J, Adiwidjaja H, Brochard P, Winnik MA. Association structure of telechelic associative thickeners in water. Langmuir 1993;9:8813. Singer W, Tenneck NJ, Driscoll AE. Method of improving viscosity stability of aqueous compositions. US Pat 3,770,684; 1973.

[292] Candau F, Regalado EJ, Selb J. Scaling behavior of the zero shear viscosity of hydrophobically modied poly(acrylamide)s. Macromolecules 1998;31:55502. ois J. [293] Feng Y, Billon L, Grassl B, Bastiat G, Borisov O, Franc Hydrophobically associating polyacrylamides and their partially hydrolyzed derivatives prepared by post modication. 2. Properties of non-hydrolyzed polymers in pure water and brine. Polymer 2005;46:928395. [294] Feng Y, Grassl B, Billon L, Khoukh A, Francois J. Effects of NaCl on steady rheological behaviour in aqueous solutions of hydrophobically modied polyacrylamide and its partially hydrolyzed analogues prepared by post modication. Polym Int 2002;51:93947. [295] Witten Jr TA, Cohen MH. Cross-linking in shear-thickening ionomers. Macromolecules 1985;18:19158. [296] Ballard MJ, Buscall R, Waite FA. The theory of shear-thickening polymer solutions. Polymer 1988;29:128793. [297] Maerker JM, Sinton SW. Rheology resulting from shear-induced structure in associating polymer-solutions. J Rheol 1986;30:7799. [298] Berret JF, Calvet D, Collet A, Viguier M. Fluorocarbon associative polymers. Curr Opin Colloid Interface Sci 2003;8:296306. [299] Kujawa P, udibert-Hayet A, Selb J, Candau F. Rheological properties of multisticker associative polyelectrolytes in semidilute aqueous solutions. J Polym Sci Part B Polym Phys 2004;42:164055. [300] Magny B, Iliopoulos I, Zana R, Audebert R. Mixed micelles formed by cationic surfactants and anionic hydrophobically modied polyelectrolytes. Langmuir 1994;10:31807. [301] Iliopoulos I, Olsson U. Polyelectrolyte association to micelles and bilayers. J Phys Chem 1994;98:15005. [302] Chang Y, Lochhead RY, McCormick CL. Water-soluble copolymers. 50. Effect of surfactant addition on the solution properties of amphiphilic copolymers of acrylamide and dimethyldodecyl(2bromide. Macromolecules acrylamidoethyl)ammonium 1994;27:214550. [303] Winnik FM, Ringsdorf H, Venzmer J. Interactions of surfactants with hydrophobically modied poly(N-isopropylacrylamides). 1. Fluorescence probe studies. Langmuir 1991;7:90511. [304] Winnik FM, Ringsdorf H, Venzmer J. Interaction of surfactants with hydrophobically modied poly(N-isopropylacrylamides). 2. Fluorescence label studies. Langmuir 1991;7:9127. [305] Penott-Chang EK, Gouveia L, Fernandez IJ, Muller AJ, az-Barrios A, Saez A. Rheology of aqueous solutions of hydrophobically modied polyacrylamides and surfactants. Colloids Surf A 2007;295:99106. [306] Gouveia LM, Muller AJ. The effect of NaCl addition on the rheological behavior of cetyltrimethylammonium p-toluenesulfonate (CTAT) aqueous solutions and their mixtures with hydrophobically modied polyacrylamide aqueous solutions. Rheol Acta 2009;48:16375. [307] Nystrm B, Thuresson K, Lindman B. Rheological and dynamic light-scattering studies on aqueous solutions of a hydrophobically modied nonionic cellulose ether and its unmodied analogue. Langmuir 1995;11:19942002. [308] Goddard ED, Leung PS. Studies of gel formation, phase behavior and surface tension in mixtures of a hydrophobically modied cationic cellulose polymer and surfactant. Colloids Surf 1992;65:2119. [309] Goddard ED, Leung PS. Interaction of cationic surfactants with a hydrophobically modied cationic cellulose polymer. Langmuir 1992;8:1499500. [310] LAlloret F, Hourdet D, Audebert R. Aqueous solution behavior of new thermoassociative polymers. Colloid Polym Sci 1995;273:116373. [311] Taylor LD, Cerankowski LD. Preparation of lms exhibiting a balanced temperature-dependence to permeation by aqueoussolutionsstudy of lower consolute behavior. J Polym Sci Part A Polym Chem 1975;13:255170. [312] Guillet JE, Rendall WA. Studies of the antenna effect in polymermolecules. 8. Photophysics of water-soluble copolymers of 1-naphthylmethyl methacrylate and acrylic-acid. Macromolecules 1986;19:22430. [313] Jenkins RD. The fundamental thickening mechanism of associative polymers in latex systems, a rheological study. PhD Dissertation, Bethlehem PA: Lehigh University; 1990. [314] Liao D, Dai S, Tam KC. Rheological properties of hydrophobic ethoxylated urethane (HEUR) in the presence of methylated bcyclodextrin. Polymer 2004;45:833948. [315] Glass JE, Karunasena A. Associative thickeners: from nonsense to reality. Polym Mater Sci Eng 1989;61:14552. [316] AbrahmsenAlami S, Alami E, Francois J. The lyotropic cubic phase of model associative polymers: small-angle X-ray scattering (SAXS),

[317] [318]

[319]

[320]

[321]

[322]

[323] [324]

[325]

[326]

[327]

[328]

[329]

[330]

[331]

[332]

[333]

[334]

[335]

[336]

[337]

[338]

[339]

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628 [340] Daoud M, Cotton JP. Star shaped polymers, a model for the conformation and its concentration dependence. J Phys 1983;43:5318. [341] Witten TA, Pincus PA. Colloid stabilization by long grafted polymers. Macromolecules 1986;19:250913. [342] Semenov AN, Joanny J-F, Khokhlov AR. Associating polymers: equilibrium and linear viscoelasticity. Macromolecules 1995;28:106675. [343] Xu B, Yekta A, Winnik MA, Sadeghy-Dalivand K, James DF, Jenkins R, Basset D. Viscoelastic properties in water of comb associative polymers based on poly(ethylene oxide). Langmuir 1997;13:690311. [344] Alami E, AbrahmsenAlami S, Vasilescu M, Almgren M. A comparison between hydrophobically end-gapped poly(ethylene oxide) with ether and urethane bonds. J Colloid Interface Sci 1997;193:15262. [345] Tanaka F, Edwards SF. Viscoelastic properties of physically crosslinked networks. Part 1. Non-linear stationary viscoelasticity. J Non-Newtonian Fluid Mech 1992;43:24771. [346] Tanaka F, Edwards SF. Viscoelastic properties of physically crosslinked networks. Part 2. Dynamic mechanical moduli. J NonNewtonian Fluid Mech 1992;43:27388. [347] Tanaka F, Edwards SF. Viscoelastic properties of physically crosslinked networks. Part 3. Time-dependent phenomena. J NonNewtonian Fluid Mech 1992;43:289309. [348] Tanaka F, Edwards SF. Viscoelastic properties of physically crosslinked networks. Transient network theory. Macromolecules 1992;25:151623. [349] Groot RD, Agterof WGM. Dynamic viscoelastic modulus of associative polymer networks: off-lattice simulations, theory and comparison to experiments. Macromolecules 1995;28: 628495. [350] Annable T, Buscall R, Ettelaie R. Network formation and its consequences for the physical behaviour of associating polymers in solution. Colloids Surf A 1996;112:97116. [351] Milner ST, Witten TA. Bridging attraction by telechelic polymers. Macromolecules 1992;25:5495503. [352] Witten TA. Associating polymers and shear thickening. J Phys 1988;49:105563. [353] Annable T, Buscall R, Ettelaie R, Shepherd P, Whittlestone D. Inuence of surfactants on the rheology of associating polymers in solution. Langmuir 1994;10:106070. [354] Zhang K, Xu B, Winnik MA, MacDonald PM. Surfactant interactions with HEUR associating polymers. J Phys Chem 1996;100:983441. [355] Binana-Limbele W, Clouet F, Francois J. Hydrophobically endcapped poly(ethylene oxide) urethanes. Part 3. Effect of sodium dodecyl sulfate on their association in aqueous solution. Colloid Polym Sci 1993;271:74858. [356] Persson K, Wang G, Olofsson G. Self-diffusion, thermal effects and viscosity of a monodisperse associative polymer. Self-association and interaction with surfactants. J Chem Soc Faraday Trans 1994;90:355562. [357] Persson K, Bales BL. EPR study of an associative polymer in solution. Determination of aggregation number and interactions with surfactants. J Chem Soc Faraday Trans 1995;91:286370. [358] Kim D-H, Kom J-W, Oh S-G, Kim J, Han S-H, Chung DJ, Suh K-D. Effects of nonionic surfactant on the rheological property of associative polymers in complex formulations. Polymer 2007;48:381721. [359] Lundberg DJ, Ma Z, Alahapperuna K, Glass JE. Surfactant inuences on hydrophobically modied thickener rheology. In: Schulz DN, Glass JE, editors. Polymers as rheology modiers. ACS symposium series no. 462. Washington, DC: American Chemical Society; 1991. p. 23453. [360] Harada A. Design and construction of supramolecular architectures consisting of cyclodextrins and polymers. Metal complex catalysts supercritical uid polymerization supramolecular architecture. Adv Polym Sci 1997;133:14191. [361] Liao D, Dai S, Tam KC. Rheological properties of a telechelic associative polymer in the presence of a- and methylated b-cyclodextrins. J Phys Chem 2007;111:3718. [362] Liao D, Dai S, Tam KC. Interaction between uorocarbon endcapped poly(ethylene oxide) and cyclodextrins. Macromolecules 2007;40:293645. [363] English RJ, Raghavan SR, Jenkins RD, Khan SA. Associative polymers bearing n-alkyl hydrophobes: rheological evidence for micro-gel behavior. J Rheol 1999;43:117594. [364] Jenkins RD, DeLong LM, Bassett DR. Inuence of alkali-soluble associative emulsion polymer architecture on rheology. In: Glass JE, editor. Hydrophilic polymersperformance with environmental acceptance. Advances in chemistry series 248. Washington, DC: American Chemical Society; 1996. p. 42547.

1627

[365] Prazeres TJV, Duhamel J, Olesen K, Shay G. Correlations between the viscoelastic behavior of pyrene-labeled associative polymers and the associations of their uorescent hydrophobes. J Phys Chem B 2005;109:1740616. [366] Tirtaatmadja V, Tam KC, Jenkins RD. Superposition of oscillations on steady share ow as a technique for investigating the structure of associative polymers. Macromolecules 1997;30:142633. [367] Tirtaatmadja V, Tam KC, Jenkins RD. Rheological properties of model alkali-soluble associative (HASE) polymers: effect of varying hydrophobe chain length. Macromolecules 1997;30:327182. [368] Tam KC, Guo L, Jenkins RD, Bassett DR. Viscoelastic properties of hydrophobically modied alkali-soluble emulsion in salt solutions. Polymer 1999;40:636979. [369] Dai S, Tam KC, Jenkins RD. Microstructure of dilute hydrophobically modied alkali soluble emulsion in aqueous salt solution. Macromolecules 2000;33:40411. [370] Nagashima K, Strashko V, MacDonald PM, Jenkins RD, Bassett DR. Diffusion of model hydrophobic alkali-swellable emulsion associative thickeners. Macromolecules 2000;33:932939. [371] English RJ, Gulati HS, Jenkins RD, Khan SA. Solution rheology of a hydrophobically modied alkali-soluble associative polymer. J Rheol 1997;41:42744. [372] Jenkins RD, Bassett DR, Shay GD. Polymers containing macromonomers. US Pat 5,292,843; 1994. [373] Siu H, Duhamel J. Comparison of the association level of a pyrenelabeled associative polymer obtained from an analysis based on two different models. J Phys Chem B 2005;109:177080. [374] Prazeres TJV, Beingessner R, Duhamel J, Olesen K, Shay G, Bassett DR. Characterization of the association level of pyrene-labeled HASEs by uorescence. Macromolecules 2001;34:787684. [375] Siu H, Duhamel J. Associations between a pyrene-labeled hydrophobically modied alkali swellable emulsion copolymer and sodium dodecyl sulfate probed by uorescence, surface tension, and viscometry. Macromolecules 2006;39:114455. [376] Wu W, Shay GD. Tailoring HASE rheology through polymer design: effects of hydrophobe size, acid content, and molecular weight. J Coat Technol Res 2005;2:42333. [377] Siddiq M, Tam KC, Jenkins RD. Dissolution behaviour of model alkali-soluble emulsion polymers: effects of molecular weights and ionic strength. Colloid Polym Sci 1999;277:11728. [378] Knaebel A, Skouri R, Munch JP, Candau SJ. Structural and rheological properties of hydrophobically modied alkali-soluble emulsion solutions. J Polym Sci Part B Polym Phys 2002;40:198594. [379] Tan H, Tam KC, Jenkins RD. Network structure of a model HASE polymer in semidilute salt solutions. J Appl Polym Sci 2001;79: 148696. [380] Tam KC, Farmer ML, Jenkins RD, Bassett DR. Rheological properties of hydrophobically modied alkali-soluble polymerseffects of ethylene-oxide chain length. J Polym Sci Part B Polym Phys 1998;36:227590. [381] Ng WK, Tam KC, Jenkins RD. Rheological properties of methacrylic acid/ethyl acrylate co-polymer: comparison between an unmodied and hydrophobically modied system. Polymer 2001;42:24959. [382] Seng WP, Tam KC, Jenkins RD. Rheological properties of model alkali-soluble associative (HASE) polymer in ionic and non-ionic surfactant solutions. Colloids Surf A 1999;154:36582. [383] Tirtaatmadja V, Tam KC, Jenkins RD. Effects of temperature on the ow dynamics of a model HASE associative polymer in nonionic surfactant solutions. Langmuir 1999;15:753745. [384] Tirtaatmadja V, Tam KC, Jenkins RD. Effect of a nonionic surfactant on the ow dynamics of a model HASE associative polymer. AIChE J 1998;44:275665. [385] Aubry T, Moan M. Inuence of a nonionic surfactant on the rheology of a hydrophobically associating water soluble polymer. J Rheol 1996;40:4418. [386] Kjoniksen AL, Beheshti N, Kotlar HK, Zhu KZ, Nystrom B. Modied polysaccharides for use in enhanced oil recovery applications. Eur Polym J 2008;44:95967. [387] Kawakami K, Ihara T, Nishioka T, Kitsuki T, Suzuki Y. Salt tolerance of an aqueous solution of a novel amphiphilic polysaccharide derivative. Langmuir 2006;22:333743. [388] Ihara T, Nishioka T, Kamitani H, Kitsuki T. Solution properties of a novel polysaccharide derivative. Chem Lett 2004;33:10945. [389] Akiyama E, Yamamoto T, Yago Y, Hotta H, Ihara T, Kitsuki T. Thickening properties and emulsication mechanisms of new derivatives of polysaccharide in aqueous solution. 2. The effect of the substitution ratio of hydrophobic/hydrophilic moieties. J Colloid Interface Sci 2007;311:43846.

1628

D.A.Z. Wever et al. / Progress in Polymer Science 36 (2011) 15581628 [415] Zhao GQ, Khin CC, Chen SB, Chen BH. Nonionic surfactant and temperature effects on the viscosity of hydrophobically modied hydroxyethyl cellulose solutions. J Phys Chem B 2005;109:14198204. [416] Dualeh AJ, Steiner CA. Hydrophobic microphase formation in surfactant solutions containing an amphiphilic graft copolymer. Macromolecules 1990;23:2515. [417] Evertsson H, Nilsson S. Microstructures formed in aqueous solutions of a hydrophobically modied nonionic cellulose derivative and sodium dodecyl sulfate: a uorescence probe investigation. Carbohydr Polym 1999;40:2938. [418] Kaczmarski JP, Tarng MR, Ma ZY, Glass JE. Surfactant and salinity inuences on associative thickener aqueous solution rheology. Colloids Surf A 1999;147:3953. [419] Lauten RA, Nystrom B. Time dependent association phenomena in dilute aqueous mixtures of a hydrophobically modied cellulose derivative and an anionic surfactant. Colloids Surf, A 2003;219:4553. [420] Maestro A, Gonzalez C, Gutierrez JM. Interaction of surfactants with thickeners used in waterborne paints: a rheological study. J Colloid Interface Sci 2005;288:597605. [421] Nilsson S, Goldraich M, Lindman B, Talmon Y. Novel organized structures in mixtures of a hydrophobically modied polymer and two oppositely charged surfactants. Langmuir 2000;16: 682532. [422] Nilsson S, Thuresson K, Lindman B, Nystrom B. Associations in mixtures of hydrophobically modied polymer and surfactant in dilute and semidilute aqueous solutions. A rheology and PFG NMR selfdiffusion investigation. Macromolecules 2000;33:96419. [423] Patruyo LG, Muller AJ, Saez AE. Shear and extensional rheology of solutions of modied hydroxyethyl celluloses and sodium dodecyl sulfate. Polymer 2002;43:648193. [424] Piculell L, Egermayer M, Sjostrom J. Rheology of mixed solutions of an associating polymer with a surfactant. Why are different surfactants different? Langmuir 2003;19:36439. [425] Thuresson K, Nystrom B, Wang G, Lindman B. Effect of surfactant on structural and thermodynamic properties of aqueous-solutions of hydrophobically-modied ethyl(hydroxyethyl)cellulose. Langmuir 1995;11:37306. [426] Thuresson K, Soderman O, Hansson P, Wang G. Binding of SDS to ethyl(hydroxyethyl)cellulose. Effect of hydrophobic modication of the polymer. J Phys Chem 1996;100:490918. [427] Thuresson K, Lindman B, Nystrom B. Effect of hydrophobic modication of a nonionic cellulose derivative on the interaction with surfactants. Rheology. J Phys Chem B 1997;101:64509. [428] Thuresson K, Lindman B. Effect of hydrophobic modication of a nonionic cellulose derivative on the interaction with surfactants. Phase behavior and association. J Phys Chem B 1997;101: 64608. [429] Joabsson F, Rosen O, Thuresson K, Piculell L, Lindman B. Phase behavior of a Clouding nonionic polymer in water. Effects of hydrophobic modication and added surfactant on phase compositions. J Phys Chem B 1998;102:29549. [430] Panmai S, Prudhomme RK, Peiffer DG, Jockusch S, Turro NJ. Interactions between hydrophobically modied polymers and surfactants: a uorescence study. Langmuir 2002;18:38604. [431] Karlberg M, Stjerndahl M, Lundberg D, Piculell L. Mixed solutions of an associating polymer with a cleavable surfactant. Langmuir 2005;21:975663. [432] Bu HT, Kjoniksen AL, Knudsen KD, Nystrom B. Effects of surfactant and temperature on rheological and structural properties of semidilute aqueous solutions of unmodied and hydrophobically modied alginate. Langmuir 2005;21:1092330. [433] Nilsson S, Thuresson K, Hansson P, Lindman B. Mixed solutions of surfactant and hydrophobically modied polymer. Controlling viscosity with micellar size. J Phys Chem B 1998;102:7099105. [434] Kjoniksen AL, Nilsson S, Thuresson K, Lindman B, Nystrom B. Effect of surfactant on dynamic and viscoelastic properties of aqueous solutions of hydrophobically modied ethyl(hydroxyethyl)cellulose, with and without spacer. Macromolecules 2000;33:87786.

[390] Karlberg M, Thuresson K, Piculell L, Lindman B. Mixed solutions of hydrophobically modied graft and block copolymers. Colloids Surf A 2004;236:15964. [391] Hwang FS, Hogenesch TE. Fluorocarbon-modied water-soluble cellulose derivatives. Macromolecules 1993;26:315660. [392] Dai SS, Ye L, Huang RH. A study on the solution behavior of IPBChydrophobically-modied hydroxyethyl cellulose. J Appl Polym Sci 2006;100:282431. [393] Ye L, Li Q, Huang RH. Study on the rheological behavior of the hydrophobically modied hydroxyethyl cellulose with 1,2epoxyhexadecane. J Appl Polym Sci 2006;101:29539. [394] Li Q, Ye L, Cai Y, Huang RH. Study of rheological behavior of hydrophobically modied hydroxyethyl cellulose. J Appl Polym Sci 2006;100:334652. [395] Gonzalez JM, Muller AJ, Torres MF, Saez AE. The role of shear and elongation in the ow of solutions of semi-exible polymers through porous media. Rheol Acta 2005;44:396405. [396] Shaw KG, Leipold DP. New cellulosic polymers for rheology control of latex paints. J Coat Technol 1985;57:6372. [397] Kstner U, Hoffmann H, Donges R, Ehrler R. Interactions between modied hydroxyethyl cellulose (HEC) and surfactants. Colloids Surf A 1996;112:20925. [398] Stone FW, Rutherford JM Jr. Quaternary nitrogen-containing cellulose ethers. US Pat 3,472,840; 1969. [399] Kstner U, Hoffmann H, Donges R, Ehrler R. Hydrophobically and cationically modied hydroxyethyl cellulose and their interactions with surfactants. Colloids Surf A 1994;82:27997. [400] Winnik FM, Winnik MA, Tazuke S, Ober CK. Synthesis and characterization of pyrene-labeled (hydroxypropyl)cellulose and its uorescence in solution. Macromolecules 1987;20:3844. [401] Karlson L, Joabsson F, Thuresson K. Phase behavior and rheology in water and in model paint formulations thickened with HMEHEC: inuence of the chemical structure and the distribution of hydrophobic tails. Carbohydr Polym 2000;41:2535. [402] Badiger MV, Lutz A, Wolf BA. Interrelation between the thermodynamic and viscometric behaviour of aqueous solutions of hydrophobically modied ethyl hydroxyethyl cellulose. Polymer 2000;41:137784. [403] Um SU, Poptoshev E, Pugh RJ. Aqueous solutions of ethyl (hydroxyethyl) cellulose and hydrophobic modied ethyl (hydroxyethyl) cellulose polymer: dynamic surface tension measurements. J Colloid Interface Sci 1997;193:419. [404] Zou S, Zhang WK, Zhang X, Jiang BZ. Study on polymer micelles of hydrophobically modied ethyl hydroxyethyl cellulose using single-molecule force spectroscopy. Langmuir 2001;17:4799808. [405] Nishikawa K, Yekta A, Pham HH, Winnik MA, Sau AC. Fluorescence studies of hydrophobically modied hydroxyethylcellulose (HMHEC) and pyrene labeled HMHEC. Langmuir 1998;14:711929. [406] Miyajima T, Kitsuki T, Kita K, Kamitani H, Yamaki K. Polysaccharide derivative, and preparation process and use thereof. US Pat 5,891,450; 1999. [407] Bostrm P, Ingvarsson I, Sundberg K. Water soluble nonionic cellulose ethers and their use in paints. US Pat 5,140,099; 1992. [408] Tanaka R, Meadows J, Phillips GO, Williams PA. Viscometric and spectroscopic studies on the solution behavior of hydrophobically modied cellulosic polymers. Carbohydr Polym 1990;12:44359. [409] Landoll LM. Modied nonionic cellulose ethers. US Pat 4,228,277; 1980. [410] Zhao GQ, Chen SB. Nonlinear rheology of aqueous solutions of hydrophobically modied hydroxyethyl cellulose with nonionic surfactant. J Colloid Interface Sci 2007;316:85866. [411] Maestro A, Gonzalez C, Gutierrez JM. Thickening mechanism in associative polymers. Macromol Symp 2002;187:91927. [412] Picton L, Muller G. Rheological properties of modied cellulosic polymers in semi-dilute regime: effect of salinity and temperature. Prog Colloid Polym Sci 1996;102:2631. [413] Maestro A, Gonzalez C, Gutierrez JM. Shear thinning and thixotropy of HMHEC and HEC water solutions. J Rheol 2002;46:144557. [414] Winnik FM. Effect of temperature on aqueous-solutions of (hydroxypropyl)cellulose. Macromolecules pyrene-labeled 1987;20:274550.

Potrebbero piacerti anche