Micelles from Doubly Thermoresponsive ... - ACS Publications

May 4, 2017 - Core−shell micelles in aqueous solution have been intensely explored, e.g. ... mass and concentration.14−17 However, if the hydropho...
8 downloads 0 Views 2MB Size
Article pubs.acs.org/Macromolecules

“Schizophrenic” Micelles from Doubly Thermoresponsive Polysulfobetaine‑b‑poly(N‑isopropylmethacrylamide) Diblock Copolymers Natalya S. Vishnevetskaya,† Viet Hildebrand,‡ Bart-Jan Niebuur,† Isabelle Grillo,§ Sergey K. Filippov,∥ André Laschewsky,*,‡,⊥ Peter Müller-Buschbaum,† and Christine M. Papadakis*,† †

Fachgebiet Physik weicher Materie/Lehrstuhl für Funktionelle Materialien, Physik-Department, Technische Universität München, James-Franck-Straße 1, 85748 Garching, Germany ‡ Institut für Chemie, Universität Potsdam, Karl-Liebknecht-Straße 24-25, 14476 Potsdam-Golm, Germany § Large Scale Structures Group, Institut Laue-Langevin, 6, rue Jules Horowitz, 38042 Grenoble, France ∥ Institute of Macromolecular Chemistry, Academy of Sciences of the Czech Republic, Heyrovsky Sq. 2, 16206 Prague 6, Czech Republic ⊥ Fraunhofer Institut für Angewandte Polymerforschung, Geiselbergstrasse 69, 14476 Potsdam-Golm, Germany S Supporting Information *

ABSTRACT: The 2-fold thermoresponsive diblock copolymer PSPP498-b-PNIPMAM144, which consists of a zwitterionic polysulfobetaine (PSPP) block and a nonionic poly(Nisopropylmethacrylamide) (PNIPMAM) block, is prepared by consecutive RAFT polymerizations. It combines the upper and lower critical solution temperature (UCST and LCST) behaviors, respectively, of the constitutive homopolymers in aqueous solution. We investigate the temperature-dependent phase behavior and the self-assembled structures of the block copolymer in D2O by turbidimetry and by small-angle neutron scattering (SANS) in salt-free solution and in the presence of small amounts of NaCl and NaBr. For comparison, solutions of PNIPMAM homopolymer in D2O are studied as well. Turbidimetry indicates thermally induced “schizophrenic” aggregation behavior for PSPP498-b-PNIPMAM144. SANS reveals that conventional star-like core−shell micellar structures are formed above the LCST transition, whereas below the UCST-transition, structure formation is much less pronounced. This is attributed to the different types of interactions, namely hydrophobic and ionic ones, dominating in the different regimes. Despite the increased polarity contrast between the zwitterionic and the nonionic blocks, and the much wider separation of the UCST- and LCST-based cloud points, CPUCST and CPLCST, the structural features of the new PSPP498-b-PNIPMAM144 resemble the ones found previously for the also 2-fold thermoresponsive analogue PSPP432b-PNIPAM200, for which both phase transition temperatures nearly coincide. Remarkably, the addition of small amounts of NaBr or NaCl to the solution of PSPP498-b-PNIPMAM144 causes a significant increase of CPUCST, as well as minor but notable changes in the self−assembled structures, but no gross alterations of the phase behavior.



INTRODUCTION Core−shell micelles in aqueous solution have been intensely explored, e.g. for forming hydrogels or for encapsulation, transport and delivery of substances that are not water-soluble. Traditionally, they have been prepared by self-assembly of diblock copolymers featuring a hydrophilic and a hydrophobic block.1−3 Such amphiphilic diblock copolymers typically form (star-like) micelles with the hydrophobic block forming the (usually considered as water-free) core and the hydrophilic block forming the shell, which is swollen by water. Micelle size and shape can be tuned by the absolute as well as the relative block lengths, while micellar dynamics are controlled by the hydrophobicity and the length of the hydrophobic block as well as by its glass transition temperature.4−6 Additional function© XXXX American Chemical Society

ality can be obtained by rendering one of the blocks responsive.7−11 If, for instance, the hydrophilic block is made thermoresponsive, the amphiphilic character and the ability for self-assembly can be turned “on” and “off” by changing the temperature. Upon a thermal stimulus, the shell block may become water-insoluble and collapses, thus enabling, e.g. control of phase separation or viscosity. A typical example for such “switchable” micelle-forming polymers are polystyrene-bpoly(N-isopropylacrylamide) (PS-b-PNIPAM) diblock copolymers,9,12,13 which exploit the thermoresponsive behavior of Received: February 16, 2017 Revised: April 21, 2017

A

DOI: 10.1021/acs.macromol.7b00356 Macromolecules XXXX, XXX, XXX−XXX

Article

Macromolecules PNIPAM. The particular focus on PNIPAM-containing polymers originates from its lower critical solution temperature (LCST) behavior of type II with a sharp coil-to-globule collapse transition at ∼32 °C, which depends only weakly on molar mass and concentration.14−17 However, if the hydrophobic PS block is not sufficiently short to enable direct water solubility,9 the preparation of micellar solutions requires the intermediate dissolution in a nonselective, water-miscible organic solvent, subsequent transfer into water, and final removal of the solvent. Apart from being a painstaking procedure, it is not always evident whether the organic solvent has been entirely removed. This is problematic for a number of applications, especially in the biomedical domain. The problem can be circumvented if the hydrophobic block is made thermoresponsive, enabling molecular dissolution in the double-hydrophilic state, and selfassembly in the amphiphilic state.7−11 This also allows, e.g., for induced gelling or for triggering the release of hydrophobic solubilizates. Up to now, most studies on such thermoresponsive block copolymers have dealt with LCST-based systems.7−11 Often, however, upper critical solution temperature (UCST) based responses would be particularly useful, e.g. in the case of the controlled release of drugs. Yet, only few polymers with UCSTbehavior in aqueous solution are known, 18−20 mostly polyzwitterions bearing sulfobetaine moieties,21−23 and little is known about the micellar self-assembly of the derived block copolymers. In addition to the frequently found UCSTbehavior in aqueous media, poly(sulfobetaine)s are also of high interest for responsive systems in the biomedical field by virtue of their high biocompatibility.21,24 Most interestingly, UCST- and LCST-based thermoresponsiveness can be also combined in block copolymers if composed of appropriate blocks. This gives rise to so-called “schizophrenic”25,26 selfassembly, enabling one to invert the roles of the respective blocks for forming the micellar core and shell (Figure 1).27−34

Figure 2. Chemical structure of the diblock copolymer under investigation (a) PSPPm-b-PNIPMAMn (m = 498, n = 144), (b) the corresponding homopolymers PSPPm′ (m′ = m) and (c) PNIPMAMn′ (n′ = 195).

passes through a maximum, beyond which it continuously decreases. These effects increase in the order SO42− < Cl− < Br−, in agreement with the Hofmeister series.42 The maximum value of CPUCST was reported for PSPP solutions in 0.004 M NaBr in D2O.37 In our previous study, solutions of PSPP432-b-PNIPAM200 with a molar mass of about 150 000 g mol−1 were investigated at concentrations of 10−50 g L−1 in D2O by means of turbidimetry, small-angle neutron scattering (SANS) and dynamic light scattering in dependence on temperature and electrolyte content.35 While above CPLCST, core−shell micelles were observed with a PNIPAM core and a PSPP shell, no analogous inverted core−shell structure could be detected for the aggregates formed below CPUCST. Apparently, the interaction between the PSPP and PNIPAM blocks was that strong that unstructured nanoparticles were formed, possibly surrounded by a thin well-hydrated PNIPAM layer. These nanoparticles formed large aggregates. In agreement, such a structure inversion, or “schizophrenic” aggregation, had previously been reported for dilute solutions of PSPP-bPNIPAM diblock copolymers with lower molar masses.27,28 These experiments revealed qualitatively the solubilization of a hydrophobic dye by the micellar aggregates formed by the collapse of the PNIPAM block at high temperatures, but not by the aggregates formed by the collapse of the PSPP block at low temperatures. Moreover, between CPUCST and CPLCST, where a homogeneous solution of PSPP432-b-PNIPAM200 was expected, concentration fluctuations were evident, as commonly observed in polyelectrolytes, which was attributed to the presence of the SPP block.35 This phenomenon was assumed to originate from the rather small difference between CPLCST and CPUCST of about 10 °C in D2O.

Figure 1. Schematic “schizophrenic” micellar self-assembly of a doubly thermoresponsive block copolymer showing CPUCST < CPLCST.

In this context, we have previously investigated in detail the self-assembly of a diblock copolymer of PNIPAM with a polysulfobetaine block, namely poly(N,N-dimethyl-N-(3(methacrylamido)propyl)ammoniopropane sulfonate) (PSPP, Figure 2b).35 The latter features UCST behavior with a clearing point (CPUCST) lower than the cloud point (CPLCST) of PNIPAM around 32 °C. The polyzwitterion’s CPUCST is markedly affected by the molar mass, by a H-D isotope effect (increasing in D2O by about 10 °C compared to H2O), and in particular by the type and concentration of added low molar mass electrolytes. 22,36−39 As reported for some other polysulfobetaines,40,41 the salt effects on PSPP are complex:36,37,39 CPUCST increases at low salt concentrations and B

DOI: 10.1021/acs.macromol.7b00356 Macromolecules XXXX, XXX, XXX−XXX

Article

Macromolecules

the polymer was dried under high vacuum. The purified PNIPMAM was obtained as yellow solid (yield 5.1 g, 50%). 1 H NMR (300 MHz, CD2Cl2, 298 K): δ (ppm) = 0.8−2.2 (broad 11H, −CH3 and −CH2−), 3.9−4.1 (1H, −CON−CHN+(CH3)2, 3.3−3.6 (−CH2−N+−CH2−, CON− CH2−), 3.9−4.1 (−CON−CH 4 to a concentration gradient near the aggregate surface.55,56 The Ornstein−Zernike (OZ) term describes the concentration fluctuations in semidilute polymer solutions,57 and gives the scaling factor, IOZ, and the correlation length, ξOZ. The SANS curves of the diblock copolymer solution in regimes I and II were analyzed using the model function:

(1)

where MCRU is the molar mass of the constitutional repeat unit, MCTA is the molar mass of the RAFT agent, cM is the molar concentration of the monomer, and cCTA is the molar concentration of the RAFT agent. Size exclusion chromatography (SEC) was performed with an apparatus SEC3010 (WGE - Dr. Bures, Dallgow-Döberitz, Germany) equipped with a refractive index detector, using PL-HFIPgel columns (Agilent Technologies), hexafluoroisopranol (HFIP) containing 50 mM of sodium trifluoroacetate as eluent, and poly(methyl methacrylate) standards (PSS Polymer Standard Service, Mainz, Germany) for calibration. Turbidimetry. Cloud points were determined by turbidimetry using a Cary 50 UV−vis spectrometer (Varian Inc., Palo Alto, CA) equipped with a single cell Peltier thermostat for the temperature control. Quartz glass cells (Hellma Analytics) with a light path of 10 mm were used. The copolymer solutions were cooled from regime II (35 °C) to regime I below CPUCST, or were heated from regime II to regime III above CPLCST. All transmittance measurements were performed at the wavelength of 500 nm during heating/cooling runs in steps of 0.5 °C with a thermal equilibration time of 5 min at each temperature. The data were normalized to the absorption of the solvent-filled cell, whose transmittance was set to 100%. The cloud points of the homopolymers were determined at a wavelength of 800 nm during cooling/heating runs in steps of 0.1 °C with an equilibration time of 12 s at each temperature. The solution was heated above the phase transition temperature and stirred prior measurement. The cloud points CPUCST and CPLCST were taken as the temperature where the normalized transmittance of the solution in the cooling and heating runs was reduced to 95% of its maximum value. Small-Angle Neutron Scattering (SANS). SANS experiments were performed at the instrument D11 at the Institut Laue-Langevin (ILL) in Grenoble, France. The incident neutrons had a wavelength λ = 6.0 Å with a spread of 9%. The 3He gas detector had an area 96 cm × 96 cm and a pixel size of 7.5 mm ×7.5 mm. A q-range from 0.02 to 5.2 nm−1 was covered. q is the momentum transfer, q = 4π × sin(θ/2)/λ with θ being the scattering angle. Samples were mounted in quartz glass cells (Hellma Analytics) with a neutron path of 1 mm. At the end of each run, the sample transmission was measured. Boron carbide was used for measurement of the dark current, and H2O for the detector sensitivity and calibration of the intensity. The scattered intensity curves were azimuthally averaged and corrected for background scattering from the solvent-filled cell and parasitic scattering with the software package LAMP.50 The homopolymer PNIPMAM195 was measured in pure D2O while heating from 22 to 50 °C in steps of 7−10 °C. The diblock copolymer PSPP498-b-PNIPMAM144 was measured in pure D2O and in 0.004 M NaBr in D2O while heating from 15 to 50 °C in steps of 7−10 °C. Measurements were performed using a copper sample holder and an inner flow circuit, which was connected to a thermostat. After each temperature change, a thermal equilibration time of 15 min was

I(q) = Iagg(q) + Isolv(q) + Ibkg

(3)

The first term, Iagg(q), describes the low-q scattering due to large aggregates, Isolv(q) is the so-called solvation term, which describes the high-q scattering, and Ibkg is the incoherent background.35 In many cases, Iagg(q) was approximated by a modified Porod term. In some cases, the Guinier approximation could be used for Iagg(q), enabling the determination of the radius of gyration of the aggregates, Rg, by

⎛ − q 2R 2 ⎞ g ⎟ Iagg(q) = IG exp⎜⎜ ⎟ ⎝ 3 ⎠

(4)

where IG is a scaling factor. Isolv(q) originates from the concentration fluctuations in semidilute polymer solutions and is governed by the solvation (hydration) of the polymer chains.58 It reveals the parameter C characterizing the solvation intensity, the solvation Porod exponent, m, which is expected to be 5/3 for a polymer coil in a good solvent in the case of noncharged polymers, while m = 2 corresponds to a Gaussian coil in a theta solvent;59 a correlation length, ξsolv; and the peak position, q0, which corresponds to an average distance d0 = 2π/q0 between the D

DOI: 10.1021/acs.macromol.7b00356 Macromolecules XXXX, XXX, XXX−XXX

Article

Macromolecules Table 1. Molecular Characteristics of the Polymers Studied Mn [g mol−1] polymer

theoretically expected

PNIPMAM195 PSPP498e PSPP498-b-PNIPMAM144

24 800 146 200 164 500

a

PDI

by end group analysisb

via 1H NMRc

by SECd

by SECd

29 000 185 000 210 000

− − 160 000

32 000 104 000 122 000

1.2 1.4 1.4

a

Calculated from monomer conversion and molar ratios of monomer and CTA; bBased on the UV/vis absorbance band at 442 nm of the R group; Calculated from the relative signal intensities of the 1H NMR signals of the PSPP and PNIPMAM blocks, assuming that the molar mass of the PSPP block is the same in the macroRAFT and in the block copolymer; dBy SEC in hexafluoro-2-propanol/50 mM CF3COONa, apparent value according to calibration with poly(methyl methacrylate) standards; eEmployed as macro CTA in the synthesis of the block copolymer. c

charged domains, d0 is finite for solutions of charged polymers and zero for neutral polymer solutions. The incoherent background was fixed at 0.06 cm−1. The scattering length density (SLD) of D2O for neutrons, ρD2O = 6.3 × 10−4 nm−2, was taken from the literature.60 The SLD value of the 0.004 M NaBr solution in D2O was calculated as 6.2 × 10−4 nm−2. The SLD values of the components of our systems were calculated using the mass densities, assuming 1.1 g cm−3 for PNIPMAM (estimated from the known density of PNIPAM, i.e., 1.1 g/cm3) and 1.0 g cm−3 for PSPP (estimated from the constituent elements and assuming a density of 1.0 g/cm3, typical for organic polymers). The following values were obtained: ρPNIPMAM = 6.8 × 10−5 nm−2 and ρPSPP ≈ 7.9 × 10−5 nm−2. According to the similarity of the SLD values for PNIPMAM and PSPP and the penetration of D2O into the micellar shells, the SLD values of the sphere ρsphere were kept in the range (6.8−7.9) × 10−4 nm−2. In the case of core−shell structures, a concentration gradient from the surface to the central part of the micelles may be present. Together with the above-described reasons, the SLD values were kept in the order ρcore < ρshell < ρD2O, namely (0.7−1.4) × 10−4 nm−2 for ρcore, (1.4−5.0) × 10−4 nm−2 for ρshell (obtained manually during fitting) and 6.3 × 10−4 nm−2 for ρD2O. The SANS curves were modeled using the SANS Data Reduction and Analysis software provided by the NIST Center for Neutron Research within the IGOR Pro software environment.61

D 2 O during cooling and heating runs (not shown): CPUCST(PSPP498) = 31.3 ± 0.5 °C and CPLCST(PNIPMAM195) = 38.0 ± 0.5 °C. The measured value of CPLCST(PNIPMAM195) is significantly lower than the ones around 44 °C mostly reported in the literature for PNIPMAM with molar masses in the range of 40 000−400 000 g mol−1.14,43−45 In analogy to similar findings for PNIPAM labeled with the same fluorophor,62 we attribute the low value of CPLCST(PNIPMAM195) to the hydrophobic fluorescence tag which was incorporated in order to facilitate the a priori difficult molar mass analysis of amphiphilic and zwitterionic block copolymers.37,39 A representative light transmission curve of PSPP498-bPNIPMAM144 in D2O at 10 g L−1 is shown in Figure 3a. Figure 3b shows the concentration dependence of the cloud points. Three regimes are distinguished in the transmission curve (Figure 3a) indicated as I, II, and III: In regimes I and III, the solution is turbid. In regime III, around 45% of light is still transmitted, while in regime I, the transmission of light is blocked. The transmission decreases sharply below CPUCST, but only gradually above CPLCST. This indicates the presence of



RESULTS Polymer Design. The doubly thermoresponsive diblock copolymer was synthesized by consecutive RAFT polymerizations of SPP and NIPMAM, following established procedures.35 The characteristics of the block copolymer as well as of the homopolymers used as references are compiled in Table 1. The results obtained by the various analytical techniques coincide well and also agree reasonably well with the theoretically expected values. Still, the numbers obtained from SEC have to be taken with some care. While the calibration of the SEC elugrams with poly(methyl methacrylate) standards seems to provide rather correct numbers for the apparent molar mass of PNIPMAM, the numbers for the apparent molar masses of PSPP seemingly underestimate the true values. Beside of the difficulties to find an appropriate calibration standard, the discrepancy between Mntheo and Mnapp might also result from a weak tendency of the zwitterionic polymers for interacting with the column material, considering a slight asymmetry in the elugrams toward long elution times. Importantly, the elugrams of the homopolymers as well as of the block copolymer show only monomodal distributions without signs of a shoulder or of a second polymer population. Accordingly, all data indicate a well-controlled polymer synthesis. Phase Behavior. Using turbidimetry, the cloud points of 50 g L −1 solutions of the homopolymers PSPP 498 and PNIPMAM195 (Figure 2b,c) were determined in salt-free

Figure 3. (a) Light transmission of a 10 g L−1 solution of PSPP498-bPNIPMAM144 in salt-free D2O (■), in 0.006 M NaCl (pink ▲), and in 0.004 M NaBr (green ●). The lines show CPUCST in salt-free solution (red − −), 0.006 M NaCl (pink -·-), and 0.004 M NaBr in D2O (green ···), as well as CPLCST in all solutions (blue -·-). I, II, and III indicate the three regimes identified. (b) Concentration dependence of the cloud points of PSPP498-b-PNIPMAM144 in salt-free D2O (closed symbols) and in 0.006 M NaCl (open symbols): (red ●,○) CPUCST, (blue ■, □) CPLCST. For comparison, (▲) CPUCST of PSPP432, (▼) CPLCST of PNIPMAM195 homopolymers (both in salt-free D2O, 50 g L−1), and (▽) cloud point for PNIPMAM from the literature are given as well.14,43−45 The lines are guides to the eye. E

DOI: 10.1021/acs.macromol.7b00356 Macromolecules XXXX, XXX, XXX−XXX

Article

Macromolecules

the LCST transitions is larger in the PSPP498-b-PNIPMAM144 solution. In both diblock copolymers in D2O, the turbidity is higher in regime I than in regime III, and whereas the transmittance decreases abruptly below CPUCST, it decreases gradually above CPLCST. The CPUCST values increase upon addition of salt up to a concentration of ca. 0.006 M with CPLCST being unchanged. The maximum transmittance in regime II of a 10 g L−1 PSPP498-b-PNIPMAM144 solution in salt-free D2O (ca. 60%) is higher than the one of the PSPP432-bPNIPAM200 solution (ca. 45%). Also, CPUCST values of PSPP498-b-PNIPMAM144 increase slightly more in dilute NaBr and NaCl solution than for PSPP432-b-PNIPAM200,35 but still much less than for PSPP homopolymers (increase by about 10 °C).37 The three regimes are also reflected in temperature-resolved 1 H NMR spectra of the block copolymer (10 g L−1 in D2O) (Figure 4). In regime II, at 35−45 °C, the proton signals of

different self-assembled structures in regimes I and III. In regime II, where the solvent conditions are good for both blocks, and where the polymer chains are expected to form a homogeneous transparent solution, the transmission is significantly lower than 100%, pointing to concentration fluctuations. The comparison of the cloud points reveals that CPUCST of the block copolymer is 29.5 ± 0.5 °C, thus about 2 °C lower than for the homopolymer PSPP498, which is a marginal difference. CPLCST of the block copolymer is 49.5 ± 0.5 °C, i.e., about 12 °C higher than for homopolymer PNIPMAM195. One possible reason for the marked increase of CPLCST of PNIPMAM is the different position of the dye end group: in the block copolymer, the label is not attached to the PNIPMAM, but to the PSPP block (Figure 2, parts a and c). Thus, the increased CPLCST in PSPP498-b-PNIPMAM144 may be due to the separation of the PNIPMAM block from the hydrophobic label. Still, when comparing the literature values of CPLCST of PNIPMAM (44 °C),14,43−45 with the one found for the diblock copolymer (49.5 °C), a difference remains, amounting to about 5 °C. Thus, CPLCST of PNIPMAM is altered by the coupling to PSPP; i.e., it depends on the polymer architecture and/or the environment, whereas CPUCST of PSPP is hardly affected by the attached PNIPMAM block. Figure 3b illustrates the concentration dependent CPUCST and CPLCST values of PSPP498-b-PNIPMAM144 in salt-free D2O and in 0.006 M NaCl D2O. In all cases, CPLCST stays constant at 49.5 ± 0.5 °C. In contrast, in salt-free solution, CPUCST increases from 21.0 ± 0.5 °C at 2 g L−1 to 29.5 ± 0.5 °C at 50 g L−1. In 0.006 M NaCl in D2O, CPUCST is slightly higher than in salt-free D2O, with the difference decreasing from 3.0 ± 0.5 °C at 2 g L−1 to 0.8 ± 0.5 °C at 50 g L−1. To summarize, CPUCST of the PSPP block is sensitive to the polymer concentration and even to low concentrations of added low molar mass salts, but changes only little by the attachment to the PNIPMAM block. In contrast, CPLCST of PNIPMAM is hardly concentration- and not salt-dependent, but markedly affected by neighboring groups, i.e., the presence of the PSPP block or of the fluorescence tag. The light transmission curves of 10 g L−1 solutions of PSPP498-b-PNIPMAM144 in salt-free D2O, in 0.006 M NaCl, and in 0.004 M NaBr (Figure 3a) show that the CPUCST value increases from 25.5 ± 0.5 °C in salt-free D2O to 27.5 ± 0.5 °C in 0.006 M NaCl and to 28.8 ± 0.5 °C in 0.004 M NaBr. Thus, NaBr has a stronger effect than NaCl, as expected from our previous studies on PSPP homopolymers in aqueous solutions, and in agreement with the Hofmeister series.37 For 50 g L−1 solutions of PSPP498-b-PNIPMAM144 in 0.004 M NaBr in D2O, CPUCST = 31.3 ± 0.5 °C and CPLCST = 49.5 ± 0.5 °C were found (not shown). While the three regimes are present in both PSPP498-bPNIPMAM144 and PSPP432-b-PNIPAM200 in D2O, the CPUCST and CPLCST values are ca. 9 and 16 °C higher respectively, in 50 g L−1 solutions of PSPP498-b-PNIPMAM144, and the width of regime II is thus considerably larger. The reasons are that (i) CPLCST of homopolymer PNIPMAM195 is ca. 10 °C higher than the one of PNIPAM200, and (ii) the higher molar mass of the PSPP498 block causes an increase of CPUCST,37 but enhances the increase of CPLCST of the nonionic block. These observations differ markedly from the ones for PSPP432-b-PNIPAM200, for which CPUCST of PSPP is reduced by 8 °C when the PNIPAM block is attached, while CPLCST of PNIPAM is unaffected by the PSPP block. As intended, the regime between the UCST and

Figure 4. Temperature dependent 1H NMR spectra of PSPP498-bPNIPMAM144 in D2O. Temperature decreases from 70 °C (top curve) to 10 °C (bottom curve) in steps of 5 °C. Left: full spectra. Right: section between 4.2 and 3.6 ppm, with signal intensities magnified by a factor of 12.5. The weak highly resolved signals originate from incomplete deuteration of the internal reference 3-trimethylsilylpropanesulfonic acid-d6 sodium salt.

both the PSPP and PNIPMAM blocks are well visible, relatively narrow, and the values of their integrals agree with the copolymer composition. In regime I, however, the signals characteristic for the PSPP block (e.g., at ∼2.0, ∼ 2.25, and 2.8−3.6 ppm)37 broaden increasingly and loose much of their intensity with decreasing temperature, while the signal characteristic for the PNIPMAM block at ca. 3.9 ppm remains virtually unchanged. Also, the intensity of the signal centered at ∼0.95 ppm, which is characteristic for the methyl group on the polymer backbone is markedly reduced, as the PSPP block is considerably longer than the PNIPMAM block. In comparison, the intensity of the signal centered at ∼1.15 ppm, which has contributions from the methylene protons of the backbone, but is in large parts due to the methyl groups of the isopropyl moiety of PNIPMAM, becomes much less attenuated when lowering the temperature. The scenario of regime I is reversed in regime III, when heating above 45 °C. Whereas the signals characteristic for the PSPP block remain nearly unchanged, the signal at 3.9 ppm of the PNIPMAM block continuously decreases with heating. Further, the intensity of the mixed signal group at 1.15 ppm is increasingly reduced in comparison F

DOI: 10.1021/acs.macromol.7b00356 Macromolecules XXXX, XXX, XXX−XXX

Article

Macromolecules

The SANS curves were fitted using the model described in eq 2. At 22 and 29 °C, i.e., below CPLCST, in addition to the expected Ornstein−Zernike (OZ) term, describing concentration fluctuations, giving rise to the decay at 0.1−2.5 nm−1, the scattering from larger polydisperse, homogeneous spherical particles (Psphere(q)) had to be taken into account as well as scattering due to large aggregates, described by the Porod term. The contributions from spherical particles vanish at 39 and 49 °C, i.e., above the CPLCST, and only the sum of the OZ and Porod terms are used in the analysis. For all temperatures, no hard-sphere structure factor is needed (set to one). At all temperatures, the achieved fits agree well with the data (best fit parameters are compiled in Table S1 in the Supporting Information). ξOZ increases from ca. 2.4 nm at 22 °C to ca. 4.2 nm at 39 °C (Figure 5b), and IOZ increases as well, reflecting critical behavior prior to the phase separation, as in PNIPAM solutions.63 At 49 °C, ξOZ is unchanged from 39 °C, but IOZ decreases abruptly, indicating that the contributions from single chain scattering are reduced. The scaling factor I0 of the decay at 0.01−2.5 nm−1 (describing spherical particles) slightly increases between 22 and 29 °C, indicating an increase of the volume fraction of the spherical particles close to the phase transition. At 22 and 29 °C, these particles have an average sphere radius ravg = 25 ± 4 nm with a polydispersity p = 0.21 ± 0.03. Thus, the PNIPMAM195 solution is not homogeneous on length scales larger than a few nm, but reveals particles at larger length scales, which collapse at CPLCST. At 22 and 29 °C, i.e. below CPLCST, their SLD value indicates penetration of D2O into these particles. The forward scattering at low q values is described by the Porod law with the exponent α growing from 3.0 ± 0.1 at 22 and 29 °C to 4.0 ± 0.2 at 39 and 49 °C, which confirms the presence of very large, compact aggregates with rough surfaces below the CPLCST, and with smooth surfaces above. The Porod scaling factor IP slightly decreases between 29 and 39 °C, reflecting the decrease of the specific surface of these large aggregates above the CPLCST, i.e., their increase in size. To summarize, the PNIPMAM195 homopolymers at 50 g L−1 in D2O form a solution with a correlation length ξOZ = 2.4−4.2 nm, and, in addition, polydisperse spherical particles with radii of around 25 nm below CPLCST, which form large aggregates with rough surfaces. Above CPLCST, the spherical particles collapse and form very large aggregates with smooth surfaces. Thus, the behavior of the PNIPMAM195 solution differs from the one of PNIPAM.63−65 In a solution of PNIPAM with a similar molar mass (Mw = 25 000 g mol−1) in D2O at a concentration of 130 g L−1, SANS revealed dissolved individual chains below the cloud point, whereas above, scattering from large aggregates dominates. In contrast, PNIPMAM forms larger particles and even large aggregates already below the cloud point.63 We assume two main reasons: (i) The additional methyl group on the backbone may favor hydrophobic intermolecular interactions already below the cloud point. (ii) The hydrophobic dye-labeled end group, which lowers the CPLCST of the homopolymer, may promote aggregation already below CPLCST. However, this effect is not expected to play a significant role in the diblock copolymer, where the label is separated from the PNIPMAM block by the long zwitterionic one. These effects may enforce the difference in the phase behavior and the structural changes in PSPP498-b-PNIPMAM144

to the mixed signal centered at 0.95 ppm, in agreement with the relative contributions of the two blocks to these signals (see above). The broadening and attenuation of the NMR signals is explained by aggregation and desolvatation of the individual PSPP and PNIPMAM blocks below CPUCST or above CPLCST, respectively. This behavior corroborates our view of temperature-controlled schizophrenic micellization of polymer PSPP498-b-PNIPMAM144, with the PSPP blocks forming the micellar core in regime I, and the PNIPMAM blocks in regime III. Nevertheless, we note that these changes do not take place abruptly, but rather gradually. Even at 10 °C, i.e., well below CPUCST, the signals of the PSPP block are still well visible, implying an appreciable residual hydration of the collapsed PSSP chains. The same remark holds true for the PNIPMAM signal at 3.9 ppm, which even at 70 °C, i.e., well above CPLCST, has not completely vanished. Structural Investigations Using SANS. The observed differences in the phase behavior of PSPP498-b-PNIPMAM144 and PSPP432-b-PNIPAM200 in salt-free D2O may be assigned to the different compositions and to the different nonionic thermoresponsive block. It is known that, in aqueous solution below the LCST, the conformation of PNIPMAM is stiffer and the chains are more expanded than in PNIPAM.45 The presence of a methyl group attached to the α-carbon of the repeat unit on the backbone restricts the free rotation due to the change of the hydration water, and thus, the hydrophobic moieties cannot associate in the most favorable way, giving rise to a counterintuitively higher LCST.47 Self-Assembled Structures in a PNIPMAM Solution. The PNIPMAM195 homopolymer is investigated with temperatureresolved SANS in salt-free D2O at a polymer concentration of 50 g L−1 (Figure 5). The high concentration together with the use of D2O ensures maximum scattering intensity.

Figure 5. (a) SANS curves from a 50 g L−1 PNIPMAM195 solution in D2O (symbols), where only every second point is shown for clarity, together with the fitting curves (). (b) Resulting temperature dependence of the correlation length, ξOZ. (blue -·-) CPLCST from turbidimetry.

The SANS curves at 22 and 29 °C feature two smooth intensity decays, which level off at q values above ca. 2.5 nm−1. At 39 °C, the decay in intensity at low q values is more pronounced, while it dominates at 49 °C. These curves differ strongly from the ones obtained from PNIPAM homopolymers.63 G

DOI: 10.1021/acs.macromol.7b00356 Macromolecules XXXX, XXX, XXX−XXX

Article

Macromolecules

of the PSPP498-b-PNIPMAM144 solution. The shallow maximum observed at a rather high q-value is typical for polyelectrolyte systems in salt-free solution,66 where the position of the correlation peak at q0 is related to an average distance d0 = 2π/q0 between charged domains.67 This solvation term (eq 3) is used to fit the SANS curves in regime II, in addition to the Guinier term (eq 4) describing the weak forward scattering at 49 °C (Table S2, Figure S1c,d). The correlation length from the solvation term, ξsolv, decreases from 8.5 ± 0.6 nm at 29 °C to 6.2 ± 0.4 nm at 49 °C. This decrease may be due to the different properties of the two blocks: At 29 °C (i.e., just above CPUCST), polymer−polymer intermolecular interactions still dominate over the interactions with water molecules for the PSPP block, whereas for the PNIPMAM block, both types of interactions are similar in strength. PNIPMAM may form a hydrophilic shell around PSPP, which is at the origin of the relatively high ξsolv value. In contrast, at 39 and 49 °C, both blocks are presumably in theta solvent conditions, and ξsolv reflects the molecular conformation of the entire polymer. At 29−49 °C, the solvation Porod exponent, m, is 2.0 ± 0.1, indicating that the solvent is a theta solvent for the PNIPMAM shell at 29 °C and for the entire polymer at 39−49 °C. The scaling factor of the solvation term, C, decreases from 11.1 ± 0.9 cm−1 at 29 °C to 6.1 ± 0.5 cm−1 at 49 °C, i.e. close to CPUCST, the value is maximum. The average distance between the charged domains, d0, is 75 ± 9 nm at 29−49 °C, which is quite large. Most probably, this is due to the expanded state of both blocks in regime II. This results from (i) the ionic attraction in the PSPP block which is weak above CPUCST, and (ii) the stiff chain conformation of PNIPMAM below CPLCST. At 49 °C, additional aggregates with an average radius of gyration Rg = 32 ± 3 nm are detected. Rg is slightly smaller than half the distance between the charged domains, d0/2 ≈ 38 nm, and may reflect the scattering from these domains. In regime I, the scattering intensity at q values below 0.2 nm−1 is higher than in regime II (Figure 6a), which may be due to higher contrast caused by aggregation. The shape of the curves is similar to those in regime II, but the maximum, which may be caused by ionic interactions in the PSPP block, is shifted to lower q values than in regime II. Moreover, the scattering curves differ strongly at low q values, where large aggregates cause an increased forward scattering. The curves in regime I can be fitted using the same model as in regime II (eq 3), but now the Porod term is needed to describe the strong forward scattering at low q values (Table S2, Figure S1b). ξsolv increases from 12 ± 1 nm at 22 °C to 38 ± 3 nm at 15 °C. Analogous to the correlation length in regime II at 29 °C, it describes mainly the correlation in the hydrophilic PNIPMAM shell surrounding the collapsed PSPP in the core. Thus, ξsolv increases with the strengthening of the attractive polymer− polymer interaction in the PSPP block. m is 2.0 ± 0.1, indicating that the PNIPMAM blocks, which keep the system in the solvated state, are in theta conditions. C increases strongly from C = 17 ± 2 cm−1 at 22 °C to 137 ± 14 cm−1 at 15 °C, indicating enhanced phase separation below CPUCST.58 d0 increases from 78 ± 8 nm at 22 °C to 118 ± 11 nm at 15 °C, due to the increasing polymer−polymer interactions in the PSPP block, causing formation of larger inhomogeneities at larger distance from each other. The size of the very large aggregates cannot be resolved, but the upturn at q values below 0.08 nm−1 can be approximated by the Porod law. The Porod exponent is α = 4.0 ± 0.4, which indicates the presence of compact aggregates with smooth surfaces.

and PSPP432-b-PNIPAM200 diblock copolymer solutions upon temperature variation. Self-Assembled Structures in the Salt-Free Diblock Copolymer Solution. A detailed structural characterization of a salt-free solution of PSPP498-b-PNIPMAM144 (50 g L−1 in D2O) was carried out using temperature-resolved SANS (Figure 6).

Figure 6. SANS curves from a 50 g L−1 salt-free solution of PSPP498-bPNIPMAM144 in D2O (symbols), where only every third point is shown for clarity, together with the fitting curves () obtained using eq 2 in regimes I and III and eq 3 in regime II, see details below (a). For better visibility, the curves are shifted in intensity by factors of 50 with respect to each other (b). Regimes I, II, and III are indicated by the blue, green and red color, respectively.

The SANS curves of the salt-free solution of PSPP498-bPNIPMAM144 in D2O during heating reflect the three regimes distinguished by turbidimetry: In regime I (blue), the SANS curves feature a smooth decay starting at the lowest q values with a shallow maximum, which moves from 0.06 nm−1 at 15 °C to 0.1 nm−1 at 22 °C. The curves become flat above ca. 3 nm−1. In regime II (green), the curves look similar, except that the decay at low q values is not present any longer. In regime III (red), the curve is flat up to ca. 0.06 nm−1, then decays steeply with a shallow second maximum at ca. 0.15 nm−1, before leveling off at ca. 3 nm−1. The changes in the curves occur at the CPUCST and at CPLCST values from turbidimetry (Figure 3b). To analyze the structures, the SANS data were fitted using the models described in the Experimental Section (Figure 6, for details see Figure S1 in the Supporting Information). The fits, which are good in all regimes, reveal substantial structural changes when CPUCST and CPLCST are crossed. The results of the fits using eq 3, applied in regimes I and II, as well as the ones using eq 2, applied in regimes I (as an alternative) and III, are summarized in Tables S2 and S3. As for the PSPP432-b-PNIPAM200 solutions,35 single chain scattering models do not describe the SANS curves in regime II H

DOI: 10.1021/acs.macromol.7b00356 Macromolecules XXXX, XXX, XXX−XXX

Article

Macromolecules

particles do not seem to contain D2O. The forward scattering is again described by the Porod law, with an exponent α = 4.1 ± 0.3, which confirms the presence of compact aggregates with smooth surfaces. Altogether, in regime I, small polydisperse spheres are present, which become larger upon cooling. The surface of the spheres is covered by the expanded, hydrated PNIPMAM blocks. Nevertheless, the spheres are weakly correlated and form large aggregates. The curves in regime III differ strongly from the ones in regimes I and II (Figure 6b): The characteristic correlation peak of the polyelectrolyte systems is not present, and the scattering up to 0.1 nm−1 points to self-assembly of the diblock copolymers on length scales smaller than hundred nm. The model described in eq 2 comprising the form factor for polydisperse core−shell spheres along with the OZ structure factor was used. The hard-sphere structure factor and the Porod term are not needed (Table S3, Figure S1e). The average core radius is rcore = 13 ± 1 nm with a polydispersity p = 0.44 ± 0.04, and the shell thickness is t = 15 ± 2 nm, resulting in the micellar radius rmic = 28 ± 3 nm. The SLD value of the core is 6.8 × 10−4 nm−2, whereas the SLD of the shell of (1.5 ± 0.2) × 10−4 nm−2 (obtained manually during fitting) indicates an amount of ca. 12% of D2O in the hydrophilic shell of the micelles. The correlation length ξOZ is 8.1 ± 0.8 nm, which describes the correlations in the PSPP shell around the collapsed PNIPMAM core. No forward scattering is observed in the SANS curves in regime III, which is in agreement with the results from turbidimetry, namely, that the solution is not completely turbid. The resulting structures are compiled in Figure 8. The difference in the micellar radii in regimes I (6−15 nm) and III

According to the previous observation for PSPP432-bPNIPAM20035 and to the scenario of the system depicted in Figure 1, the formation of spherical micelles is expected at temperatures below CPUCST. Therefore, the curves in regime I were fitted using the model described in eq 2 as well, describing polydisperse spheres correlated by a hard-sphere structure factor, the concentration fluctuations in the shell by an Ornstein−Zernike term and very large aggregates by the Porod term (Table S3, Figure S1a). The fits are equally good and give additional structural information. The average radius of the polydisperse spherical particles increases from ravg = 6.1 ± 0.5 at 22 °C to ravg = 15 ± 2 nm at 15 °C (Figure 7), and the

Figure 7. Results from fitting using eq 2 to the SANS curves in Figure 6 in regimes I and III. Temperature dependence of the micellar radius, ravg (■), the micellar radius of the core−shell structure, rmic (▲), and the core radius, rcore (▼). The hard-sphere radius, RHS (−) (a), of the correlation length, ξOZ (b), of the volume fraction, η (c), and of the SLD values of the polymer spheres (★) and the shell (▶) and the core (◀) of the core−shell polymeric micelles (d). In part b, a logarithmic axis is used. In some cases, the symbol size is larger than the error bar. (red − −) and (blue -·-): CPUCST and CPLCST values from turbidimetry. Regimes I, II, and III are indicated on top of the graph.

Figure 8. Schematic representation of the micelles/polymer coils in the three regimes as indicated. Red: PSPP block. Blue: PNIPMAM block.

polydispersity is rather high (p = 0.47 ± 0.09), pointing to aggregation. The hard-sphere radius (or half the interparticle distance), RHS, increases as well, namely from 22 ± 2 nm at 22 °C to 43 ± 3 nm at 15 °C. RHS is larger than ravg, i.e. the particles are spaced, and follows the same trend. The RHS value is smaller than d0/2, which is around 40 and 60 nm at 22 and 15 °C, respectively, but the trend is the same. The hard-sphere volume fraction of correlated micelles, η, is about 0.08, the correlation between the spherical particles is thus very weak. The OZ term reveals that ξOZ increases from 7.7 ± 0.3 nm at 22 °C to 12.3 ± 0.9 nm at 15 °C. These values are comparable with ravg, i.e., the presumed PNIPMAM shells are very loosely packed. ξOZ and ξsolv cannot be compared directly but their values follow the same trend at low temperatures. The large values may be due to the stiff chain conformation of the PNIPMAM blocks. The SLD value of the spherical particles was found at (6.8−7.9) × 10−5 nm−2, which corresponds to the SLD values of pure PSPP and PNIPMAM; i.e., the spherical

(28 nm) may be assigned to the block properties and the block copolymer architecture: The PSPP block, which is waterinsoluble in regime I, is about 3 times longer than the PNIPMAM block, which is water-insoluble in regime III. In regime II, both blocks are soluble, and the solubility of the PSPP block is driven by the ionic interactions between water and the charged polymer chains, resulting in concentration fluctuations. In regime I, small spherical domains (6−15 nm), presumably consisting of both, PSPP and PNIPMAM, are immersed in a PNIPMAM-rich matrix, are correlated and form very large aggregates. The expected core−shell structure of the micelles in regime I is not observed, probably because the PSPP and PNIPMAM blocks interact with each other. In contrast, in regime III, the micellar radius is much larger (28 nm), and a core−shell structure is evident. Neither a correlation of the micelles nor aggregation on larger length scales are observed, I

DOI: 10.1021/acs.macromol.7b00356 Macromolecules XXXX, XXX, XXX−XXX

Article

Macromolecules

reason may be the higher block length of PSPP, which is below its CPUCST in this regime, and is at the origin of structure formation. Another reason may be the tendency of PNIPMAM to expand, as evidenced by our measurements on PNIPMAM192, even though CPLCST is further away than it is in PSPP432-b-PNIPAM200. This effect may also promote the formation of very large aggregates in PSPP498-b-PNIPMAM144. In regime II, the scattering intensity of the PSPP498-bPNIPMAM144 solution is also higher, the polyelectrolyte peak is shifted to lower q values and the length scales are larger: The average distance between the charged domains, d0, is 1.5 times larger, which is presumably due to the expanded conformations of PNIPMAM block and the comparably smaller impact of the polyelectrolyte properties of the PSPP block. The tendency to aggregation of the PNIPMAM block close to CPLCST is corroborated by the appearance of inhomogeneities on a length scale of around 30 nm just below CPLCST in PSPP498-bPNIPMAM144. The largest differences between the two systems are observed in regime III. In this regime, the overall scattering intensity in PSPP498-b-PNIPMAM144 is lower, especially at low q values, and the core radius is significantly smaller (13 nm vs 73 nm). The smaller size of the PNIPMAM block compared to the PNIPAM block forming the micellar cores alone cannot explain this huge difference: Using ideal chain statistics, the core radii should only differ by a factor (200/144)1/2 = 1.18. The core radius of 13 nm is rather close to the expected value of the endto-end distance of PSPP, which may be roughly estimated using the end-to-end distance of poly(methyl methacrylate) of a molar mass of 27 900 g mol−1 (value from NMR, Table 1),68 namely 10 nm. Since the core radius of 73 nm found previously for PSPP432-b-PNIPAM200, is significantly larger, we suspect that, in PSPP432-b-PNIPAM200, clusters of micelles are formed, whereas they are separated from each other in PSPP498-bPNIPMAM144. This may be due to both reasons, the longer PSPP block and the stronger tendency for hydrophobic aggregation of PNIPMAM. The micellar shell formed by PSPP that has a thickness of 15 nm in PSPP 498 -bPNIPMAM 144 , is smaller than the one of PSPP 432 -bPNIPAM200 (19 nm), and contains about 5 times less D2O. Thus, the PSPP shell is more compact, which may point to its interaction with PNIPMAM. Electrolyte Effect. Turbidimetry reveals changes of the CPUCST values of the diblock copolymer in D2O upon addition of low molar mass electrolytes (Figure 3a). To gain information about the structural changes caused by electrolyte addition, temperature-dependent SANS measurements of a 50 g L−1 solution of PSPP498-b-PNIPMAM144 were performed in 0.004 M NaBr in D2O (Figure 9). In regime I (blue), the SANS curves of PSPP498-bPNIPMAM144 in 0.004 M NaBr in D2O feature significantly increased forward scattering at q values below ca. 0.09 nm−1 as well as a shift of the shallow maximum to lower q values (Figure 9b). In regime II (green), the forward scattering below 0.04 nm−1 at 29 and 39 °C, and below 0.06 nm−1 at 49 °C is slightly increased (Figure 9c,d). Above these q values, the curves stay virtually unchanged compared to the ones in salt-free D2O. The shape of the curves in regime III (red) remain unchanged over the q range above 0.09 nm−1, but at lower q values, the intensity is 1.5 times higher than in salt-free solution (Figure 9e). The same fitting models are used, and the resulting parameters are compiled in Tables S4 and S5 (using eqs 3 and 2, respectively). The results of the latter are shown in Figure 10.

which may be due to the thick hydrophilic PSPP shell, keeping the micelles in the solvated state. The different characters of the microphase-separated domains in regimes I and III are qualitatively corroborated by the temperature-dependent fluorescence characteristics of the solvatochromic end-group label. In the case of homopolymer PNIPMAM195 in dilute solution, the emission maximum of the fluorophore shifts from 546 nm below CPLCST to 536 nm above CPLCST, which goes along with a strong increase of the fluorescence intensity. This indicates a less polar environment of the fluorophore above the coil-to globule phase transition, quite similar to the behavior of analogously labeled PNIPAM.62 Accordingly, the hydrophobic dye partitions preferentially into the PNIPMAM-rich microphase. In contrast, in the case of PSPP498, the emission maximum of the fluorophore stays at 544 nm, independently of the temperature and of the coil-to globule phase transition, while the fluorescence intensity is always weak. Accordingly, the collapsed zwitterionic block does not accommodate the polar, but hydrophobic dye. The fluorescence behavior of the block copolymer PSPP498-b-PNIPMAM144, in which the solvatochromic label is attached to the zwitterionic block, corresponds to the one of the PSPP homopolymer. In regimes I and II, the emission maximum is always found at about 543 nm, supporting the view that the microphase separated domains of PSPP formed in regime I are much less hydrophobic than the domains formed by PNIPMAM in regime III. This would account for the reduced disposition of the block copolymer to produce well-defined core−shell aggregate structures in regime I. The self-assembly behavior is confirmed in temperatureresolved dynamic light scattering (DLS) experiments in backscattering geometry on a 10 g L−1 diblock copolymer solution in D2O (see the Supporting Information). DLS reveals very large aggregates (hydrodynamic radius Rh ≅ 700 nm) in regime I, unimers with Rh ≅ 8 nm in regime II (these are unconnected due to the lower concentration than in SANS), and aggregates having Rh ≅ 65 nm in regime III. The stability of a 150 g L−1 diblock copolymer solution in D2O is investigated in heating/cooling cycles using small-angle X-ray scattering (SAXS) (for experimental details and results see the Supporting Information). The same equilibration times as in the SANS experiments were applied. No difference between heating and cooling is found. Comparison of the Self-Assembled Structures in Salt-Free Solutions of PSPP498-b-PNIPMAM144 and PSPP432-b-PNIPAM200. When comparing the SANS results from PSPP498-bPNIPMAM144 with the previously reported ones for PSPP432-bPNIPAM200 in salt-free D2O,35 three factors have to be considered: (i) The overall molar mass of PSPP 498-bPNIPMAM144 is slightly higher. (ii) The block lengths and, accordingly, the composition of the diblock copolymers are different: In PSPP498-b-PNIPMAM144, the PSPP block is 1.2 times longer, whereas the PNIPMAM block is 1.4 times shorter; i.e., the fraction of PSPP is higher. (iii) PNIPMAM homopolymers form aggregates already below the cloud point, which is not the case in PNIPAM solutions. SANS curves from both diblock copolymers in regimes I−III are compiled in Figure S2. In regime I, the overall scattering intensity is higher for the PSPP498-b-PNIPMAM144 solution, which is mainly due to the significantly higher forward scattering due to very large aggregates and to the larger mesoscopic length scales, e.g. the radii of the spherical particles are 1.3−3.4 times larger. One J

DOI: 10.1021/acs.macromol.7b00356 Macromolecules XXXX, XXX, XXX−XXX

Article

Macromolecules

is unaffected by the presence of NaBr, as expected; i.e., heating leads to a weakening the attractive interactions between the PSPP segments. The solvation Porod exponent, m, is around 2.0, indicating θ solvent quality for the PNIPMAM shell. The scaling factor C decreases from 13.4 ± 1.2 cm−1 at 29 °C to 7.1 ± 0.7 cm−1 at 49 °C. These values are slightly higher than the ones in salt-free solution; thus, interpolymer interactions are strengthened due to the screening effect of the salt, in agreement with the upward shift of CPUCST. The average distance between the charged domains, d0, is 83 ± 11 nm at 29 °C and at 39 °C, and 103 ± 10 nm at 49 °C, i.e., on average 8 nm larger than in the salt-free solution with the difference being most pronounced close to CPLCST. Thus, in regime II, the salt screening effect causes larger distances between the inhomogeneities. In contrast to the salt-free polymer solution, the Guinier approximation has to be applied to describe the upturn at low q values at all temperatures in regime II. Aggregates with Rg decreasing from about 91 nm at 29 °C to 60 nm at 39 °C and further to 36 nm at 49 °C are formed. The Guinier scaling factor decreases from 29 to 39 °C from about 6 to 2 cm−1, indicating a decrease of the fraction of the aggregates above the UCST-type transition. It increases to 10 cm−1 at 49 °C, which is 3 times higher than the one in salt-free solution. Screening by salt thus breaks up the larger aggregates/more homogeneous large-scale structure observed in salt-free solution, and the scattering of the aggregates moves into the accessible q-range. In regime I, fitting eq 3 reveals a similar trend for all parameters as in salt-free solution, however, with the following differences (Table S4, Figure 10): The correlation length ξsolv increases by a factor of about 1.4. Thus, the screening of ionic interactions causes the formation of spherical particles with a less compact (presumed) PNIPMAM shell. The solvation Porod exponent m is slightly higher (by 0.05 and by 0.21 at 22 and 15 °C, respectively), indicating that the solvent quality of the shell improves at lower temperatures. The scaling factor C is 2 and 4 times higher at 22 and 15 °C, respectively, due to the incipient phase separation, which already starts in regime II. Moreover, d0 is on average about 20 nm larger. The Porod exponent and amplitude are unchanged. Comparison of the fitting parameters in regime I using eq 2 with and without NaBr reveals the following changes: The sphere radius ravg is only 1.3 nm larger at 22 °C, but 11 nm larger at 15 °C with a lower polydispersity at 15 °C: p = 0.49 ± 0.06 at 22 °C and p = 0.34 ± 0.02 at 15 °C (Table S5, Figure 10). RHS is on average 8 nm larger, indicating an increase in the distance between the (larger) spheres. ξOZ differs only at 15 °C, where it is around 4 nm smaller than in salt-free conditions. Most probably, the salt screening effect promotes the formation of more compact spherical particles at low temperatures and, at 15 °C, the correlation length may also include contributions from the inner part of the sphere. The Porod law reveals α = 4.1 ± 0.3 with nearly the same amplitude as in salt-free solution, indicating a similar rate of aggregation. The SLD values are unchanged as well. In regime III, the core−shell micelles are still uncorrelated and, within the uncertainties, unchanged from the salt-free solution, but have a slightly larger core: rcore = 15 ± 2 nm. The SLD value of the core is 6.8 × 10−5 nm−2, whereas the SLD of the shell is (9.2 ± 0.8) × 10−5 nm−2, indicating around 5% of D2O in the shell, which is much less than in salt-free conditions. The correlation length ξOZ in regime III, describing the correlation in the hydrophilic PSPP-rich shell, is 7.0 ± 0.6 nm, thus slightly lower than in salt-free conditions. ξOZ may be

Figure 9. SANS curves from a 50 g L−1 solution of PSPP498-bPNIPMAM144 in salt-free D2O (open symbols, from Figure 6) and in 0.004 M NaBr in D2O (closed symbols), where only every third point is shown for clarity, together with the fitting curves (−) obtained using eq 2 in regimes I and III and eq 3 in regime II. The curves are shifted in intensity as in Figure 6. (b−e) Zooms of the low q region: curves in regime I at 15 °C (b), in regime II at 22 °C (c) and at 49 °C (d), and in regime III at 59 °C (e), where only every second point is shown for clarity.

Figure 10. Results from model fitting to the SANS curves in Figures 6 and 9 from PSPP498-b-PNIPMAM144 in salt-free D2O (black symbols), and in 0.004 M NaBr in D2O (red symbols), respectively. Same designations as in Figure 7. CPUCST in salt-free D2O (red − −) and in 0.004 M NaBr in D2O (red ···) and CPLCST value (blue -·-), all determined by turbidimetry.

In regime II, the correlation length ξsolv (describing the correlation in the hydrophilic PNIPMAM shell) decreases from 8.8 ± 0.9 nm at 29 °C to 5.8 ± 0.5 nm at 49 °C (Table S4) and K

DOI: 10.1021/acs.macromol.7b00356 Macromolecules XXXX, XXX, XXX−XXX

Article

Macromolecules

effectively though in opposite ways, and to added low molar mass electrolytes, which is particular to the zwitterionic block. Turbidimetry shows for both block copolymers the same general pattern, revealing three regimes of association and selfassembly. At low and high temperatures, namely below the UCST-type and above the LCST-type phase transitions of the blocks, the solutions are turbid indicating the formation of aggregates, while at intermediate temperatures, they are translucent. Nevertheless, the results show also distinct differences between the block copolymers, demonstrating the importance of the precise molecular structure for the detailed responsive self-assembly. The choice of PNIPMAM as nonionic thermoresponsive block increases not only the width of the intermediate regime II by increasing CPLCST. Moreover, the apparently small molecular change by adding one methyl group to the repeat unit modulates markedly the interaction of the nonionic block with adjacent zwitterionic and hydrophobic segments in the block copolymer, quite differently from the use of PNIPAM. The reasons for the findings seem complex, suggesting a certain extent of hydrophobic interaction between PNIPMAM chains already below the phase transition, as well as interactions with the zwitterionic block. Temperature-resolved SANS measurements in D2O give structural information in the three regimes. PSPP498-bPNIPMAM144 exhibits behavior similar to polyelectrolytes in solution, due to the ionic interactions between the PSPP chains. Below CPUCST, the turbid solution is characterized by very large aggregates with a smooth surface. These are formed by correlated, small and homogeneous spherical particles. In the intermediate temperature range, between the UCST- and LCST-type transitions, the solution is optically not completely clear, but remains hazy. This is attributed to concentration fluctuations due to the high number of ionic groups bound to the PSPP block. Above the LCST-type transition of the PNIPMAM block, the partially turbid solution contains spherical core−shell micelles with a polydisperse core, which are not correlated. Dynamic light scattering in backscattering geometry confirmed the overall behavior. In addition, no differences are observed in SAXS measurements taken during heating and cooling runs. The difference in block lengths, the PSPP block being three times as large as the PNIPMAM block, and their individual properties, such as dominating ionic interpolymer interactions versus hydrogen bonding with water, yield the marked differences of the aggregation behavior below CPUCST and above CPLCST, respectively. Accordingly, the use of the LCST transition for the controlled core−shell formation and release of hydrophobic active substances is more favorable than the use of the UCST transition. This conclusion is supported by the temperature-dependent fluorescence characteristics of hydrophobic dye labels attached to the polymer ends. While the dye partitions preferentially into the microphase-separated domains of the nonionic block above its LCST-type coil-to-collapse transition, it does not incorporate into the microphase-separated domains of the zwitterionic block provided below the UCST-type transition. As the CPUCST of PSPP is sensitive to added electrolytes, salt effects on the aggregation behavior of the diblock copolymer solution in 0.004 M NaBr in D2O were investigated as well. Turbidimetry reveals a salt-induced upward shift of CPUCST by about 2 °C, while CPLCST remains virtually unchanged. Although the salt concentration (0.004 M NaBr) is too low to alter the general behavior of the polymer, slight structural changes are found in all regimes: enhanced aggregation and

decreased because of the screening effect. Again, no forward scattering is observed. Therefore, in the regime where micelles with a PNIPMAM-rich core and a PSPP-rich shell are formed, the addition of a small amount of salt causes the formation of the micelles with a slightly larger core and a slightly thinner shell than in salt-free solution. Comparison of the Effect of Salt in PSPP498-b-PNIPMAM144 and PSPP432-b-PNIPAM200. While the structural changes found in 0.004 M NaBr solutions of PSPP498-b-PNIPMAM144 and the reported earlier PSPP432-b-PNIPAM200 follow the same trend, we nevertheless observe some specific differences. In regime I of the PSPP498-b-PNIPMAM144 solution, the enhanced interpolymer interactions caused by the salt screening effect lead to a slight increase of the dimensions of the small spherical particles, and to an enhanced aggregation. These changes are more pronounced in the solution of PSPP498-bPNIPMAM144 than of PSPP432-b-PNIPAM200. Analogously, these enhanced attractive interactions cause the formation of a small fraction of large aggregates in regime II of the PSPP498-b-PNIPMAM144 solution, which was neither found in the salt-free solution of the diblock copolymer nor in regime II of the PSPP432-b-PNIPAM200 solution. The fact that, at temperatures slightly above CPUCST, larger aggregates appear in both systems upon salt addition, and that the solvation scaling factor increases, is in accordance with the shift of CPUCST to higher temperatures. Nevertheless, in regime II of both diblock copolymer solutions, the screening of the charges does not make the polyelectrolyte peak disappear. The concentration of NaBr, 0.004 M, is too low to alter the general behavior, and the salt only enhances the attractive interpolymer interactions, and hence, the distance between the charged domains increases. The changes of the core−shell micelles formed in regime III upon addition of electrolyte are only minor in PSPP498-bPNIPMAM144, where the micelles are uncorrelated. In contrast, in PSPP432-b-PNIPAM200, the micelles presumably form clusters and become more compact. Accordingly, the decrease of solvent content in the micellar shell in PSPP498-bPNIPMAM144 solution is around twice as large as in PSPP432b-PNIPAM200. This may be due to the fact that the PSPP shell is not only sensitive to the presence of salt, but is also prone to interact with the PNIPMAM block. The latter is characterized by a stronger tendency for aggregation due to the additional methyl group that increases its hydrophobicity. Nevertheless, analogous to PSPP432-b-PNIPAM200, low concentrations of added low molar mass salts of PSPP cause only minor structural changes in all regimes for PSPP498-b-PNIPMAM144 solutions in D2O upon the addition of NaBr at a concentration of 0.004 M.



CONCLUSION The influences of the architecture and of the chemical nature of the individual blocks on the solution phase behavior are investigated for a doubly thermoresponsive diblock copolymer consisting of a zwitterionic PSPP block and a nonionic PNIPMAM block, which exhibit UCST and LCST behavior, respectively. We compare the “schizophrenic” responsive aggregation behavior of PSPP498-b-PNIPMAM144 in D2O with the analogously designed copolymer PSPP432-b-PNIPAM200.35 For both systems, CPUCST is lower than CPLCST, so that the zwitterionic and the nonionic blocks are hydrophilic at intermediate temperatures, and the diblock copolymers are water-soluble in the entire temperature range studied (10−65 °C). Their aggregation behavior can be controlled by two stimuli, namely temperature to which both blocks respond L

DOI: 10.1021/acs.macromol.7b00356 Macromolecules XXXX, XXX, XXX−XXX

Macromolecules



increased radii of the small spheres in regime I, an increased distance between the charged domains in regime II, and enlarged micellar core dimensions in regime III. We explain this by screening of the ionic interactions between the charged groups attached to the polymers. When comparing the behaviors of PSPP498-b-PNIPMAM144 and of PSPP432-b-PNIPAM200, reported previously, we observe a similar overall behavior, yet, the cloud points are altered in the copolymers from the ones in the respective homopolymers in a different way: In PSPP432-b-PNIPAM200, the CPUCST value markedly decreases as compared to the PSPP430 homopolymer while CPLCST is unchanged. In contrast, for PSPP498-bPNIPMAM144, the CPLCST value markedly increases compared to the PNIPMAM195 homopolymer while CPUCST remains unchanged. This qualitative difference may be assigned to the steric hindrance in PNIPMAM due to the additional methyl group, resulting in a higher sensitivity to the environment, including the PSPP block and the dye-labeled end group attached. Apart from the difference in the cloud points, the aggregation behavior seems to be generic. Nevertheless, SANS reveals structural differences: Below CPUCST, the sphere radius is higher in PSPP498-b-PNIPMAM144 than in PSPP432-bPNIPAM200, whereas smaller aggregates are observed above CPLCST. This may be partially due to the different ratio of the lengths of the zwitterionic to the nonionic blocks. A more important reason may be the chemical difference between the PNIPMAM and PNIPAM nonionic blocks which may favor the formation of associated structures of PNIPMAM below its CPLCST, and stronger hydrophobic aggregation of the polymer chains above. An additional reason may be interactions of the PNIPMAM block with the PSPP block, both disposing of the same polymer backbone, which do not occur for PNIPAM Thus, the UCST- and LCST-type transition temperatures can be controlled by the selection of the nonionic thermoresponsive block as well as by the lengths of both blocks. Moreover, the UCST-type transition can be altered without affecting the LCST-type transition, by addition of low molar mass electrolytes. The particular tunable thermoresponsive properties of these block copolymers make them promising candidates for a variety of applications, e.g., in the biomedical field and in the development of switchable surfaces and thermo-optical devices.



Article

AUTHOR INFORMATION

Corresponding Authors

*E-mail [email protected]. Telephone +49 89 289 12 447. Fax +49 89 289 12 473 (C.M.P.). *E-mail [email protected]. Telephone +49 331 997 5225. Fax +49 331 997 5036 (A.L.). ORCID

Sergey K. Filippov: 0000-0002-4253-5076 Peter Müller-Buschbaum: 0000-0002-9566-6088 Christine M. Papadakis: 0000-0002-7098-3458 Author Contributions

The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript. Funding

We thank Deutsche Forschungsgemeinschaft (DFG) for financial support (PA 771/14−1, MU 1487/17−1, LA 611/ 11−1). Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We thank A. Lieske and M. Walter (Fraunhofer IAP) for support with the SEC measurements and M. Heydenreich, A. Krtitschka, and D. Schanzenbach (Universität Potsdam) for support with the temperature dependent 1H NMR studies. Institut Laue-Langevin is acknowledged for beam time allocation and for providing excellent equipment.



REFERENCES

(1) Harada, A.; Kataoka, K. Supramolecular Assemblies of Block Copolymers in Aqueous Media as Nanocontainers Relevant to Biological Applications. Prog. Polym. Sci. 2006, 31, 949−982. (2) Mai, Y.; Eisenberg, A. Self-Assembly of Block Copolymers. Chem. Soc. Rev. 2012, 41, 5969−5985. (3) Raffa, P.; Wever, D. A. Z.; Picchioni, F.; Broekhuis, A. A. Polymeric Surfactants: Synthesis, Properties, and Links to Applications. Chem. Rev. 2015, 115, 8504−8563. (4) Nicolai, T.; Colombani, O.; Chassenieux, C. Dynamic Polymeric Micelles Versus Frozen Nanoparticles Formed by Block Copolymers. Soft Matter 2010, 6, 3111−3118. (5) Schaeffel, D.; Kreyes, A.; Zhao, Y.; Landfester, K.; Butt, H.-J.; Crespy, D.; Koynov, K. Molecular Exchange Kinetics of Diblock Copolymer Micelles Monitored by Fluorescence Correlation Spectroscopy. ACS Macro Lett. 2014, 3, 428−432. (6) Wright, D. B.; Patterson, J. P.; Pitto-Barry, A.; Cotanda, P.; Chassenieux, C.; Colombani, O.; O’Reilly, R. K. Tuning the Aggregation Behavior of pH-Responsive Micelles by Copolymerization. Polym. Chem. 2015, 6, 2761−2768. (7) Dimitrov, I.; Trzebicka, B.; Müller, A. H. E.; Dworak, A.; Tsvetanov, C. B. Thermosensitive Water-Soluble Copolymers with Doubly Responsive Reversibly Interacting Entities. Prog. Polym. Sci. 2007, 32, 1275−1343. (8) Laschewsky, A. Smart Ingredients for Aqueous Formulations: Facts and Fancy. SÖ FW-Journal 2011, 137, 34−40. (9) Laschewsky, A.; Müller-Buschbaum, P.; Papadakis, C. M. Thermoresponsive Amphiphilic Di- and Triblock Copolymers Based on Poly(N-isopropylacrylamide) and Poly(methoxy diethylene glycol acrylate): Aggregation and Hydrogel Formation in Bulk Solution and in Thin Films. Prog. Colloid Polym. Sci. 2013, 140, 15−34. (10) Kelley, E. G.; Albert, J. N. L.; Sullivan, M. O.; Epps, T. H., III Stimuli-Responsive Copolymer Solution and Surface Assemblies for Biomedical Applications. Chem. Soc. Rev. 2013, 42, 7057−7071.

ASSOCIATED CONTENT

* Supporting Information S

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.macromol.7b00356. Fits to the SANS curves from a 50 g L−1 PSPP498-bPNIPMAM144 solution in salt-free D2O and best fit parameters for the SANS data of a 50 g L −1 PNIPMAM195 and PSPP498-b-PNIPMAM144 solution in D2O and in 0.004 M NaBr in D2O at different temperatures, SANS curves from PSPP498-b-PNIPMAM144 and PSPP432-b-PNIPAM200 solutions in saltfree D2O, dynamic light scattering, description of the experimental details, and SAXS results from PSPP498-bPNIPMAM144 in D2O obtained in heating/cooling cycles (PDF) M

DOI: 10.1021/acs.macromol.7b00356 Macromolecules XXXX, XXX, XXX−XXX

Article

Macromolecules (11) Strandman, S.; Zhu, X. X. Thermo-Responsive Block Copolymers with Multiple Phase Transition Temperatures in Aqueous Solutions. Prog. Polym. Sci. 2015, 42, 154−176. (12) Cammas, S.; Suzuki, K.; Sone, C.; Sakurai, Y.; Kataoka, K.; Okano, T. Thermo-Responsive Polymer Nanoparticles with a CoreShell Micelle Structure as Site-Specific Drug Carriers. J. Controlled Release 1997, 48, 157−164. (13) Nuopponen, M.; Ojala, J.; Tenhu, H. Aggregation Behaviour of Well Defined Amphiphilic Diblock Copolymers with Poly(Nisopropylacrylamide) and Hydrophobic Blocks. Polymer 2004, 45, 3643−2650. (14) Fujishige, S.; Kubota, K.; Ando, I. Phase Transition of Aqueous Solutions of Poly(N-isopropylacrylamide) and Poly (N-isopropylmethacrylamide). J. Phys. Chem. 1989, 93, 3311−3313. (15) Schild, H. G. Poly(N-isopropylacrylamide): Experiment, Theory and Application. Prog. Polym. Sci. 1992, 17, 163−249. (16) Aseyev, V.; Tenhu, H.; Winnik, F. M. Non-Ionic Thermoresponsive Polymers in Water. Adv. Polym. Sci. 2010, 242, 29−89. (17) Halperin, A.; Kröger, M.; Winnik, F. M. Poly(N-isopropylacrylamide) Phase Diagrams: Fifty Years of Research. Angew. Chem., Int. Ed. 2015, 54, 15342−15367. (18) Seuring, J.; Agarwal, S. Polymers with Upper Critical Solution Temperature in Aqueous Solution. Macromol. Rapid Commun. 2012, 33, 1898−1920. (19) Zhang, Q.; Hoogenboom, R. Polymers with Upper Critical Solution Temperature Behavior in Alcohol/Water Solvent Mixtures. Prog. Polym. Sci. 2015, 48, 122−142. (20) Niskanen, J.; Tenhu, H. How to Manipulate the Upper Critical Solution Temperature (UCST)? Polym. Chem. 2017, 8, 220−232. (21) Monroy Soto, V. M.; Galin, J. C. Poly(sulphopropylbetaines): 2. Dilute Solution Properties. Polymer 1984, 25, 254−262. (22) Laschewsky, A. Structures and Synthesis of Zwitterionic Polymers. Polymers 2014, 6, 1544−1601. (23) Zhu, Y.; Noy, J.-M.; Lowe, A. B.; Roth, P. J. The Synthesis and Aqueous Solution Properties of Sulfobutylbetaine (Co)Polymers: Comparison of Synthetic Routes and Tuneable Upper Critical Solution Temperatures. Polym. Chem. 2015, 6, 5705−5718. (24) Grainger, D. W. All Charged Up About Implanted Biomaterials. Nat. Biotechnol. 2013, 31, 507−509. (25) Bütün, V.; Liu, S.; Weaver, J. V. M.; Bories-Azeau, X.; Cai, Y.; Armes, S. P. A brief review of ’schizophrenic’ block copolymers. React. Funct. Polym. 2006, 66, 157−165. (26) Kotsuchibashi, Y.; Ebara, M.; Aoyagi, T.; Narain, R. Recent Advances in Dual Temperature Responsive Block Copolymers and Their Potential as Biomedical Applications. Polymers 2016, 8, 380. (27) Arotçaréna, M.; Heise, B.; Ishaya, S.; Laschewsky, A. Switching the Inside and the Outside of Aggregates of Water-Soluble Block Copolymers with Double Thermoresponsivity. J. Am. Chem. Soc. 2002, 124, 3787−3793. (28) Virtanen, J.; Arotçaréna, M.; Heise, B.; Ishaya, S.; Laschewsky, A.; Tenhu, H. Dissolution and Aggregation of a Poly(NIPA-blocksulfobetaine) Copolymer in Water and Saline Aqueous Solutions. Langmuir 2002, 18, 5360−5365. (29) Liu, S.; Billingham, N. C.; Armes, S. P. A Schizophrenic WaterSoluble Diblock Copolymer. Angew. Chem., Int. Ed. 2001, 40, 2328− 2331. (30) Maeda, Y.; Mochiduki, H.; Ikeda, I. Hydration Changes during Thermosensitive Association of a Block Copolymer Consisting of LCST and UCST Blocks. Macromol. Rapid Commun. 2004, 25, 1330− 1334. (31) Wang, D.; Wu, T.; Wan, X.; Wang, X.; Liu, S. Purely SaltResponsive Micelle Formation and Inversion Based on a Novel Schizophrenic Sulfobetaine Block Copolymer: Structure and Kinetics of Micellization. Langmuir 2007, 23, 11866−11874. (32) Shih, Y.-J.; Chang, Y.; Deratani, A.; Quemener, D. “Schizophrenic” Hemocompatible Copolymers via Switchable Thermoresponsive Transition of Nonionic/Zwitterionic Block SelfAssembly in Human Blood. Biomacromolecules 2012, 13, 2849−2858.

(33) Zhang, Q.; Hong, J.-D.; Hoogenboom, R. A Triple Thermoresponsive Schizophrenic Diblock Copolymer. Polym. Chem. 2013, 4, 4322−4325. (34) Cummings, C.; Murata, H.; Koepsel, R.; Russell, A. J. Dramatically Increased pH and Temperature Stability of Chymotrypsin Using Dual Block Polymer-Based Protein Engineering. Biomacromolecules 2014, 15, 763−771. (35) Vishnevetskaya, N. S.; Hildebrand, V.; Niebuur, B.-J.; Grillo, I.; Filippov, S. K.; Laschewsky, A.; Müller-Buschbaum, P.; Papadakis, C. M. Aggregation Behavior of Doubly Thermoresponsive Polysulfobetaine-b-poly(N-isopropylacrylamide) Diblock Copolymers. Macromolecules 2016, 49, 6655−6668. (36) Mary, P.; Bendejacq, D. D.; Labeau, M.-P.; Dupuis, P. Reconciling Low- and High-Salt Solution Behavior of Sulfobetaine Polyzwitterions. J. Phys. Chem. B 2007, 111, 7767−7777. (37) Hildebrand, V.; Laschewsky, A.; Zehm, D. On the Hydrophi licity of P ol yzwit terion P oly ( N , N- d i m e t h y l - N- ( 3 (methacrylamido)propyl)ammoniopropane sulfonate) in Water, Deuterated Water, and Aqueous Salt Solutions. J. Biomater. Sci., Polym. Ed. 2014, 25, 1602−1618. (38) Ranka, M.; Brown, P.; Hatton, T. A. Responsive Stabilization of Nanoparticles for Extreme Salinity and High-Temperature Reservoir Applications. ACS Appl. Mater. Interfaces 2015, 7, 19651−19658. (39) Hildebrand, V.; Laschewsky, A.; Wischerhoff, E. Modulating the Solubility of Zwitterionic Poly((3-methacrylamidopropyl)ammonioalkane Sulfonate)s in Water and Aqueous Salt Solutions via the Spacer Group Separating the Cationic and the Anionic Moieties. Polym. Chem. 2016, 7, 731−740. (40) Köberle, P.; Laschewsky, A.; Lomax, T. D. Interactions of a Zwitterionic Polysoap and its Cationic Analog with Inorganic Salts. Makromol. Chem., Rapid Commun. 1991, 12, 427−433. (41) Hildebrand, V.; Laschewsky, A.; Päch, M.; Müller-Buschbaum, P.; Papadakis, C. M. Effect of the Zwitterion Structure on the Thermoresponsive Behaviour of Poly(Sulfobetaine Methacrylate)s. Polym. Chem. 2017, 8, 310−322. (42) Lo Nostro, P.; Ninham, B. W. Hofmeister Phenomena: An Update on Ion Specificity in Biology. Chem. Rev. 2012, 112, 2286− 2322. (43) Chytrý, V.; Netopilík, M.; Bohdanecký, M.; Ulbrich, K. Phase Transition Parameters of Potential Thermosensitive Drug Release Systems Based on Polymers of N-alkylmethacrylamides. J. Biomater. Sci., Polym. Ed. 1997, 8, 817−824. (44) Duracher, D.; Elaissari, A.; Pichot, C. Characterization of CrossLinked Poly(N-Isopropylmethacrylamide) Microgel Latexes. Colloid Polym. Sci. 1999, 277, 905−913. (45) Djokpe, E.; Vogt, W. N-Isopropylacrylamide and NIsopropylmethacrylamide: Cloud Points of Mixtures and Copolymers. Macromol. Chem. Phys. 2001, 202, 750−757. (46) Tiktopulo, E. I.; Uversky, V. N.; Lushchik, V. B.; Klenin, S. I.; Bychkova, V. E.; Ptitsyn, O. B. “Domain” Coil-Globule Transition in Homopolymers. Macromolecules 1995, 28, 7519−7524. (47) Tang, Y.; Ding, Y.; Zhang, G. Role of Methyl in The Phase Transition of Poly(N-Isopropylmethacrylamide). J. Phys. Chem. B 2008, 112, 8447−8451. (48) Dybal, J.; Trchová, M.; Schmidt, P. The Role of Water in Structural Changes of Poly(N-isopropylacrylamide) and Poly(Nisopropylmethacrylamide) Studied by FTIR, Raman Spectroscopy and Quantum Chemical Calculations. Vib. Spectrosc. 2009, 51, 44−51. (49) Spěvácě k, J.; Dybal, J. Temperature-Induced Phase Separation and Hydration in Aqueous Polymer Solutions Studied by NMR and IR Spectroscopy: Comparison of Poly(N-vinylcaprolactam) and Acrylamide-Based Polymers. Macromol. Symp. 2014, 336, 39−46. (50) LAMP, the Large Array Manipulation Program, http://www.ill. eu/data_treat/lamp/the-lamp-book/. (51) Kotlarchyk, M.; Chen, S.-H. Analysis of Small Angle Neutron Scattering Spectra from Polydisperse Interacting Colloids. J. Chem. Phys. 1983, 79, 2461−2469. (52) Guinier, A.; Fournet, G. Small-Angle Scattering of X-Rays; John Wiley and Sons Inc.: New York, 1955; p 268. N

DOI: 10.1021/acs.macromol.7b00356 Macromolecules XXXX, XXX, XXX−XXX

Article

Macromolecules (53) Percus, J. K.; Yevick, G. Analysis of Classical Statistical Mechanics by Means of Collective Coordinates. Phys. Rev. 1958, 110, 1−13. (54) Porod, G. Die Röntgenkleinwinkelstreuung von dichtgepackten kolloiden Systemen. Kolloid Z. 1951, 124, 83−114. (55) Koberstein, J. T.; Morra, B.; Stein, R. S. The Determination of Diffuse-Boundary Thicknesses of Polymers by Small-Angle X-ray Scattering. J. Appl. Crystallogr. 1980, 13, 34−45. (56) Hammouda, B. New Guinier-Porod Model. J. Appl. Crystallogr. 2010, 43, 716−719. (57) Shibayama, M.; Tanaka, T.; Han, C. C. Small Angle Neutron Scattering Study on Poly(N-isopropyl acrylamide) Gels Near their Volume-Phase Transition Temperature. J. Chem. Phys. 1992, 97, 6829−6841. (58) Horkay, F.; Hammouda, B. Small-angle Neutron Scattering from Typical Synthetic and Biopolymer Solutions. Colloid Polym. Sci. 2008, 286, 611−620. (59) Feuz, L.; Strunz, P.; Geue, T.; Textor, M.; Borisov, O. Conformation of Poly(L-lysine)-graft-poly(ethylene glycol) Molecular Brushes in Aqueous Solution Studied by Small-Angle Neutron Scattering. Eur. Phys. J. E: Soft Matter Biol. Phys. 2007, 23, 237−245. (60) Sears, V. F. Neutron Scattering Lengths and Cross Sections. Neutron News 1992, 3, 26−37. (61) Kline, S. R. Reduction and Analysis of SANS and USANS Data Using IGOR Pro. J. Appl. Crystallogr. 2006, 39, 895. (62) Inal, S.; Kölsch, J. D.; Chiappisi, L.; Janietz, D.; Gradzielski, M.; Laschewsky, A.; Neher, D. Structure-Related Differences in the Temperature Regulated Fluorescence Response of LCST Type Polymers. J. Mater. Chem. C 2013, 1, 6603−6612. (63) Meier-Koll, A.; Pipich, V.; Busch, P.; Papadakis, C. M.; MüllerBuschbaum, P. Phase Separation in Semidilute Aqueous Poly(NIsopropylacrylamide) Solutions. Langmuir 2012, 28, 8791−8798. (64) Wu, C.; Li, W.; Zhu, X. X. Viscoelastic Effect on the Formation of Mesoglobular Phase in Dilute Solutions. Macromolecules 2004, 37, 4989−4992. (65) Cheng, H.; Shen, L.; Wu, C. LLS and FTIR Studies on the Hysteresis in Association and Dissociation of Poly(N-isopropylacrylamide) Chains in Water. Macromolecules 2006, 39, 2325−2329. (66) Bernstein-Levi, O.; Ochbaum, G.; Bitton, R. The Effect of Covalently Linked RGD Peptide on the Conformation of Polysaccharides in Aqueous Solution. Colloids Surf. B 2016, 137, 214−220. (67) Roe, R. J. Methods of X-ray and Neutron Scattering in Polymer Science; Oxford University Press: New York, 2000. (68) Fetters, L. J.; Lohse, D. J.; Colby, R. H. Chain Dimensions and Entanglement Spacings. In Physical Properties of Polymers Handbook. 2nd ed., Mark, J. E., Ed.; Springer: 2007; Chapter 25, pp 445−454.

O

DOI: 10.1021/acs.macromol.7b00356 Macromolecules XXXX, XXX, XXX−XXX