Salt-sensitivity of the thermoresponsive behavior of PNIPAAm-Brushes

Jan 22, 2018 - We report investigations on the salt-sensitivity of the thermoresponsive behavior of PNIPAAm-brushes applying the quartz-crystal microb...
1 downloads 0 Views 13MB Size
Subscriber access provided by READING UNIV

Article

Salt-sensitivity of the thermoresponsive behavior of PNIPAAm-Brushes Meike König, Keith Brian Rodenhausen, Sebastian Rauch, Eva Bittrich, KlausJochen Eichhorn, Mathias M. Schubert, Manfred Stamm, and Petra Uhlmann Langmuir, Just Accepted Manuscript • DOI: 10.1021/acs.langmuir.7b03919 • Publication Date (Web): 22 Jan 2018 Downloaded from http://pubs.acs.org on January 25, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Langmuir is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Salt-sensitivity of the thermoresponsive behavior of PNIPAAm-Brushes Meike Koenig,†,‡ Keith Brian Rodenhausen,¶,§ Sebastian Rauch,† Eva Bittrich,† Klaus-Jochen Eichhorn,† Mathias Schubert,k,⊥ Manfred Stamm,†,# and Petra Uhlmann∗,†,@ †Leibniz-Institut f¨ ur Polymerforschung Dresden e.V., Hohe Straße 6, 01069 Dresden, Germany ‡Present address: Karlsruhe Institute of Technology, Institute of Functional Interfaces, Hermann-von-Helmholtz-Platz 1, 76344 Eggenstein-Leopoldshafen, Germany ¶Department of Chemical and Biomolecular Engineering, University of Nebraska-Lincoln, 207 Othmer Hall, Lincoln, NE 68588, USA §Present address: Biolin Scientific, Inc., Paramus, New Jersey 07652, USA kDepartment of Electrical and Computer Engineering and Center for Nanohybrid Functional Materials, 209N Scott Engineering Center University of Nebraska-Lincoln, Lincoln, NE 68588, USA ⊥Department of Physics, Chemistry, and Biology, IFM, Link¨oping University, SE-581 83 Link¨oping, Sweden #Technische Universit¨at Dresden, Faculty of Science, Department of Chemistry, Chair of Physical Chemistry of Polymeric Materials, Bergstrae 66, 01069 Dresden,Germany @Department of Chemistry, Hamilton Hall, University of Nebraska-Lincoln, 639 North 12th Street, Lincoln, NE 68588, USA E-mail: [email protected]

1

ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Abstract We report investigations on the salt-sensitivity of the thermoresponsive behavior of PNIPAAm-brushes applying the quartz-crystal microbalance coupled with spectroscopic ellipsometry technique. This approach enables a detailed study of the optical and mechanical behavior of the polymer coatings. Additional conclusions can be drawn from the difference between both techniques due to a difference in the contrast mechanism of both methods. A linear shift of the phase transition temperature to lower temperatures with the addition of sodium chloride was found, similar to the behavior of free polymer chains in solution. The thermal hysteresis was found to be decreased by the addition of sodium chloride to the solution, hinting to the interaction of the ions with the amide groups of the polymer, whereby the formation of hydrogen bonds is hindered. The results of this study are of relevance to the application of PNIPAAm-brushes in biological fluids and demonstrate the additional potential of the ion-sensitivity besides the better known thermo-sensitivity.

2

ACS Paragon Plus Environment

Page 2 of 32

Page 3 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Introduction PNIPAAm probably is the most prominent example for polymers exhibiting LCST behavior and has been studied extensively. 1–4 One reason for this is that its phase transition temperature Tc in water lies in the vicinity of the human body temperature, rendering it interesting for biological applications. 5–7 Because biological fluids contain high amounts of salt, the effect of additional salt on the phase transition, as it was found for PNIPAAm chains in solution, 1,8,9 is of importance, as well. The addition of ions to an aqueous solution was reported to influence the position of the phase transition. 1,10 This effect is widely known as the “Hofmeister effect”, describing the ability of anions and cations to influence many crucial properties of a solution. 11–14 Initially it was discussed to stem from a change in the ordering of the water structure, but in recent reports, the Hofmeister effect was argued to partly arise from direct interaction between the ions and the solute. 15–18 In the case of PNIPAAm, Zhang et al. discussed three mechanisms contributing to the shift of Tc : (a) dehydration of the amide groups by kosmotropic ions, due to favorable hydration of the small ions, (b) a change in surface tension affecting the hydration of the hydrophobic parts of the polymer and (c) direct interaction of the anions with the amide groups. 16 While the first two effects lead to salting-out of the polymer (decrease in Tc ), the last effect leads to salting-in of the polymer (increase in Tc ). The extent to which each of these effects contributes to the overall behavior of an ion varies. Cl− is located in the middle of the Hofmeister series with weak chaotropic properties. Although weak binding to the amide group was measured, overall a salting-out effect is found due to the increase in surface tension. 16,19 Polymer brushes are a special type of polymeric coating, that have received increasing interest in recent years. 20–23 These systems consist of polymer chains, tethered by one end to a planar or curved substrate, in close proximity to each other such that the chains are forced to stretch away from the surface in a “brush conformation”. 24 Responsive polymer brushes are capable of reacting to external stimuli, generally by reversible swelling-deswelling 3

ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

behavior. PNIPAAm brushes have been reported to exhibit a broader transition than in solution, this has been discussed to stem from the increased interaction between the closely tethered polymer chains. 25,26 Few reports have dealt with the influence of additional salt on the thermoresponsive behavior of grafted PNIPAAm chains. 27–34 Jhon et al. investigated the influence of NaCl on grafting-from PNIPAAm brushes with unknown grafting density by Quarz Crystal Microbalance with Dissipation monitoring (QCM-D). They found a nonlinear dependence of the shift of Tc to lower temperatures, with a lower rate of decline at increased concentrations of NaCl. This is unlike the behavior found for PNIPAAm in solution, where the addition of NaCl lowers the transition temperature in a linear manner. 27 Ishida et al. investigated the response of PNIPAAm brushes with very low grafting density (∼0.008 chains nm−2 ) to the addition of Na2 SO4 using insitu-AFM and QCM-D. They found a very unusual peaklike change in frequency and dissipation with increasing salt content, which they attributed to the drastic structural changes occurring in the polymer layer. 28 Naini et al. measured the influence of different sodium halides on the switching kinetics and thermodynamics of PNIPAAM brushes using a laser temperature-jump technique 30 The halides were found to effect the Tc according to their order in the Hofmeister series. The effect of kosmotropic and chaotropic anions was compared by the group of Wanless using neutron reflectivity, atomic force microscopy and QCM-D 32 or QCM-D, ellipsometry and contact angle measurements. 33 Also here, a behavior in accordance with the Hofmeister series was found, with kosmotropic acetate ions shifting the Tc to lower temperatures and the chaotropic thiocyanate ion shifting the Tc to higher temperatures. A linear dependence on the concentration of the ions was detected, similar to the behavior of free PNIPAAm chains in solution. However, neither of these reports uses viscoelastic modeling to further elucidate the detected changes in QCMD data. Moreover, the discrepancies found by Humphreys et al. comparing the results of QCM-D and ellipsometry measurements could not be fully explained by the authors. 33 We used a combinatorial setup of QCM-D and Spectroscopic Ellipsometry (SE) to in-

4

ACS Paragon Plus Environment

Page 4 of 32

Page 5 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

vestigate the phase transition, and its shift with NaCl content, of grafting-to PNIPAAm brushes. Using viscoelastic and optical modeling, this setup allows a comprehensive study of the thickness, composition, optical and mechanical properties of the polymer layer. 35–37 Here, the combined setup ensures the direct comparability of the two techniques for further evaluation.

Results and Discussion

Figure 1: Exemplary data of QCM and SE measurements of the salt and temperature dependent swelling of PNIPAAm brushes; (a) ∆F of the seventh overtone, (b) ∆D of the seventh overtone, (c) Ψ at λ=501 nm, (d) ∆ at λ=501 nm

The effect of the concentration of NaCl was investigated by measuring the temperaturedependent swelling at four different salt concentrations (0 M, 0.1 M, 0.5 M and 1 M) for a PNIPAAm brush with a dry thickness of 12 ± 0.5 nm and a grafting density of 0.16 nm−1 . 5

ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 32

The measurements were recorded in quasi-static state, allowing the sample to equilibrate for 30 min at each temperature step, before the temperature was raised or cooled further. Since also a bare QCM-D sensor shows temperature-dependent frequency and dissipation shifts, the response of a bare sensor was subtracted from the data in order to only evaluate the effect of the polymer coating (see figure S1 in the Supporting Information). The effect of the salt concentration on the response of the bare sensor was found to be negligible. Figure 1 shows exemplary data from QCM-D and SE measurements of one heating and cooling cycle; similar curves can be plotted for each of the overtones and wavelengths measured. The transition temperature was determined from those curves as the maximum of the first derivative, by averaging the slopes of two adjacent points (Table 1). Table 1: Transition temperatures Tc determined from raw data of QCM and SE measurements of the salt and temperature dependent swelling of PNIPAAm brushes c(NaCl)[mol/l] 0 0.1 0.5 1

Tc,Heating (Tc,Cooling )[◦ C] ∆F ∆D Ψ 33.8 (33.8) 33.8 (32.8) 33.8 (31.8) 31.8 (31.8) 31.8 (31.8) 31.8 (30.9) 26.8 (26.8) 25.9 (25.9) 25.9 (24.9) 20.9 (19.9) 19.9 (19.9) 19.9 (19.5)

6

ACS Paragon Plus Environment

33.8 31.8 25.9 19.9

∆ (32.8) (30.9) (25.9) (19.9)

Page 7 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

The addition of salt to the solution has a very prominent effect: increasing the concentration of NaCl gradually shifts the phase transition to lower temperatures. The plot of Tc as calculated from the heating curves is shown in the Supporting Information (figure S2). As found for PNIPAAm in solution, the transition is linearly shifted upon addition of NaCl. At c(NaCl)=1 M the transition is shifted by ∼13 degrees to 20 ◦ C. This change is slightly larger than the change reported for PNIPAAm in solution, where Tc was found to shift from 31 ◦ C to about 22 ◦ C at c(NaCl)=1 M. 8 The shape of the transition curve was found to remain unchanged upon addition of NaCl and no broadening or sharpening was observed. This indicates that the addition of NaCl does not alter the mechanism of the deswelling of the chains. PNIPAAm is known from literature for exhibition of a hysteresis between the heating and the cooling cycle due to the formation of inter- and intramolecular hydrogen bonds in the collapsed state that have to be broken for reswelling. 38–40 For grafted polymer chains, the hysteresis was found to depend on the grafting density. 4 Comparing the results of the two techniques, a distinction can be seen in the hysteresis behavior: While QCM data show almost no difference in between heating and cooling cycle, the transition temperature derived from SE data varies by about one degree. This can be attributed to the different sensitivities of SE and QCM-D to the limit of the outer boundary of the brush against the bulk ambient. 41,42 There is very low optical contrast between the ambient bulk and the top of the swollen brush, where dilute polymer chains dangle out; SE is more biased to the denser region of the brush layer near the substrate. QCM-D is more sensitive to the dilute regime of the brush, where rehydration of the PNIPAAm chains and breaking of intermolecular hydrogen bonds occur more easily. To further investigate the hysteresis behavior, the difference between heating and cooling cycle at each temperature is plotted in Figure 2. The biggest difference between heating and cooling is found at Tc , as expected. Of greater interest is the observation that in almost all cases the hysteresis decreases with increasing salt content. A similar effect has been reported

7

ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 2: Difference between heating and cooling cycle in exemplary data of QCM and SE measurements of the salt and temperature dependent swelling of PNIPAAm brushes; (a) ∆F of the seventh overtone, (b) ∆D of the seventh overtone, (c) Ψ at λ=501 nm, (d) ∆ at λ=501 nm so far only for the strong chaotropic anion SCN− by Shechter et al. 43 They proposed that the presence of bound ions diminishes the segment-segment interaction leading to less interand intrachain hydrogen bonds, which would have to be broken for reswelling. Even though in our case the direct interaction of NaCl with the amide group is much weaker, 19 the effect is obviously large enough to be detected by QCM-D and SE. For kosmotropic anions a decreasing effect on the hysteresis was found as well, but here a relative stabilization of the collapsed coiled stage was hypothesized. 44 Since in our case the shape of the transition curve does not change upon the addition of NaCl, this effect can be excluded.

8

ACS Paragon Plus Environment

Page 8 of 32

Page 9 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Figure 3: Effect of salt and temperature on the amount of aqueous solution in PNIPAAm brushes; amount of aqueous solution detected by SE (a) and QCM-D (b), (c) difference in amount of aqueous solution detected by QCM-D and SE

A Voigt-Voinova approach was used to evaluate changes in areal mass and viscoelasticity of QCM-D data. Negative dissipation shift values are not treatable by the Voigt-Voinova model. In order to avoid these, ∆D was offset such that the minimum ∆D value obtained at the highest respective temperature was set as the zero reference point. At this reference point, the brush is already solvated to some extent, but the amount of aqueous solution acoustically coupled and adsorbed to the polymer brush in this state is unknown and the derived mass change only contains the amount of aqueous solution added to the polymer layer upon further swelling. SE data was modeled using a multilayer box model, applying a two component effective medium approach (EMA) to fit for the optical properties of the brush layer. This rather simple box model has turned out to work very well for thin polymer films as

9

ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

investigated here, since more complex models led to correlations in the modeled parameters. 45 Within the box layer a homogenous distribution of the polymer-segment density within the box is assumed. From this model, the thickness dbrush and the volume fraction of aqueous solution in the film fH2 O can be derived. The mass density of aqueous solution within the swollen polymer film was then calculated using ΓH2 O,SE = fH2 O · ρH2 O · dbrush . A density of 1 g cm−3 for the aqueous solution was assumed. Figure 3 displays the best-fit model data of the salt dependent measurements for the amount of aqueous solution ΓSE and ΓQCM . The critical temperature can be clearly detected as a strong decrease in the amount of aqueous solution in the polymer layer. The shift seen in the raw data upon addition of NaCl is mirrored in the amount of aqueous solution, as well. As in the raw data, the actual shape of the temperature dependent swelling is not changed by the presence of the ions. The difference in the amount of aqueous solution between the two techniques ∆ΓQCM −SE can be attributed to a difference in contrast, thereby differing in the definition of the outer boundary of the thin film, comprised of vibrationally coupled ambient molecules and dangling polymer chains. 41,42 Thus, the magnitude of the discrepancy between the two techniques can provide insight into the sharpness of the filmambient interface. Low values of ∆ΓQCM −SE imply that the contrast between the edge of the brush and the bulk ambient is more well defined, while higher values imply a decrease in optical contrast. As described in a previous report, 37 two different swelling regimes can be clearly defined: starting from low temperatures ∆ΓQCM −SE stays almost constant or slighlty increases, indicating that although the amount of aqueous solution decreases, the outer boundary of the polymer layer does not change. With a further rise of temperature, the amount of aqueous solution detected by both techniques quickly converges. This can be explained by the transition from a parabolic to a step-wise profile of the chain segment density as the polymer chains in the brush layer collapse, resulting in a more defined interface between polymer layer and ambient, which can analogously be detected by both techniques. Interestingly, ∆ΓQCM −SE of the swollen state decreases slightly with increasing NaCl con-

10

ACS Paragon Plus Environment

Page 10 of 32

Page 11 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

centration. This could be caused by a more defined interface between the brush and the ambient, due to the interaction with the ions.

Figure 4: Effect of salt and temperature on the viscosity of PNIPAAm brushes Best match model data for the layer viscosity, shown in Figure 4, exhibit a shift to higher values at high salt concentration. Here the polymer chains are surrounded by a more viscous ambient, which leads to a higher dissipation of vibrational energy of the swollen brush layer. For pure water and c(NaCl)=0.1 M, a local minimum is detected before the most collapsed stage with a substantial hysteresis between heating and cooling. The reason for this is unclear since this effect is not mirrored in the raw data. Further studies are needed in order to reveal the origin of this feature. To investigate the influence of the grafting density on the salt dependent swelling behavior, PNIPAAm brushes with a grafting density of 0.09 nm−1 were measured in a similar manner (see figures S4 and S5 in the Supporting Information). As described in a previous publication, 37 a lower grafting density results in a lower swelling ratio and a slight shift of Tc to lower temperatures. The thermal hysteresis was found to be diminished as well. However, the effect of additional sodium chloride was found to be similar as for the polymer brush with a higher grafting density.

11

ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Conclusion Summarizing, in this report the effect of salt-concentration on the thermoresponsive behavior of PNIPAAm brushes was investigated. A large linear shift of Tc to lower temperatures was observed upon addition of NaCl. This shift was in a similar range as found for PNIPAAm in solution. The comparison of the two applied techniques indicated that SE detects a larger thermal hysteresis than QCM-D, probably because SE has an increased sensitivity for the dense region of the brush, where the effect of interchain hydrogen bonds would be more prominent. The measured hysteresis effect decreased with increasing NaCl content, which indicates the interaction of ions with the amide groups and disruption of hydrogen bonding. The insights reported here, provide further knowledge for the development and application of thermoresponsive coatings utilizing PNIPAAm brushes and demonstrate the interesting potential for applications of the additional ion-sensitivity. The results also show that salt effects have to be considered when PNIPAAm brushes are supposed to be used for the temperature sensitive control of biointeractions.

Experimental Brush Preparation Polymer brushes were grafted on silica coated quartz crystals with a ca. 50 nm thick silicon oxide layer (QSX 303, QSense, Fr¨olunda, Sweden). Samples were cleaned with ethanol abs. (VWR, Germany) and dried with N2 gas. Then, samples were activated by oxygen plasma for 1 min at 100W (440-G Plasma System, Technics Plasma GmbH, Germany) A macromolecular anchoring layer was spin-coated onto the surface from a 0.02 wt.-% solution of poly-(glycidyl methacrylate) (PGMA, Mn =17 500 g mol−1 , Mw /Mn =1.12, Polymer Source, Inc., Canada) in chloroform (Sigma-Aldrich, Germany) and annealed in vacuum for 20 min at 100 ◦ C to chemically bind the polymer to the activated SiO2 surface. Carboxyterminated PNIPAAm (Mn =48 900 g mol−1 , Mw /Mn =1.3) was synthesized by controlled radical polymerization (atom transfer radical polymerization, ATRP) as described previously 46 12

ACS Paragon Plus Environment

Page 12 of 32

Page 13 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

and grafted onto the surface from a 0.8 wt.-% solution in chloroform, by annealing in vacuum at 150 ◦ C, overnight. Non-covalently bound polymer was extracted in ethanol. The process was followed by ellipsometry in dry state after each modification step. The grafting density of the resulting polymer brush was calculated from the dry thickness, 24 using literature values for the bulk density of PNIPAAm. 47

Combinatorial QCM-D/SE The combinatorial QCM-D/SE setup consists of an E1 QCM-D and an ellipsometry compatible module (Q-Sense, Fr¨olunda, Sweden, mounted onto the sample stage of an alpha-SE spectroscopic ellipsometer with a fixed angle of incidence (AOI) of 65◦ and a spectral range of 370-900 nm. Accidentally, measurements were performed with an instrument-setting of 70◦ AOI. Subsequent data analysis was corrected by appropriate numerical models for the liquid cell internal AOI of 68.8◦ by considering small, calibrated refractive polarization variations across the entrance and exit windows. The ellipsometrycompatible module is an air-tight liquid flow chamber, into which the sample is inserted for insitu analysis; the module has windows for the probing light beam of an ellipsometry measurement. Flow of liquid medium is facilitated by means of chemical resistant PTFE tubing (VICI AG International, Switzerland), which is connected to a peristaltic flow pump (Ismatec IPC high precision multichannel dispenser, IDEX Health & Science GmbH, WertheimR Ultrapure was used, Mondfeld, Germany). Millipore water purified with Purelab Plus

with and without addition of NaCl. The ellipsometry-compatible module temperature was controlled by the QCM-D software, and the system temperature was maintained, step-wise, at 30 min intervals. Temperature dependent experiments were done at stagnant flow conditions.

Ellipsometry modeling Modeling of the optical properties (refractive index) and the thickness (d) of the polymer films from the ellipsometry-measured primary data ∆ (relative phase shift) and tan(ψ) (relative amplitude ratio) was performed using a multilayer-boxmodel, consisting of a silica coated quartz sensor, PGMA as an anchoring layer, and PNI13

ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

PAAM as a brush layer, respectively. Measurements of the blank silica coated quartz sensors were parameterized by basis-spline (B-spline) functions and considered “virtual substrates” for further modeling. 48 The modification with polymer layers in dry state was modeled with fixed refractive indeces described by Cauchy dispersions (PGMA: A = 1.525 and B = 0 µm2 , PNIPAAm: A = 1.442 and B = 0.0056 µm2 , determined by measurements of a thick polymer layer) in an optical box model. For insitu measurements, a two-parameter Bruggemann-EMA model was used to determine the thickness of the polymer brush dbrush , the combined refractive index nef f and the volume fractions of the polymers fP olymer and the ambient solution fH2 O , using fixed values of refractive index as for the dry measurements. For the optical properties of the ambient the refractive index, n(λ) was determined with a digital multiple wavelength refractometer (DSR-lambda, Schmidt+Haensch GmbH& Co.) at eight different wavelengths from 435.8 nm to 706.5 nm.

QCM-D modeling Shifts in frequency and dissipation of the odd overtones (j= 3,5,7,9,11) were modeled using a Voigt-Voinova approach 49 for one homogeneous viscoelastic layer with a fixed density of 1 g cm−3 . Measurements on PNIPAAm brushes were referenced to the measurement with the smallest dissipation value. The temperature change in frequency and dissipation of a bare QCM-D sensor was subtracted from the QCM-D data, to eliminate the temperature-sensitivity of the bare quartz substrate. For modeling the software QTools (Q-Sense, Fr¨olunda, Sweden) was used.

Acknowledgement Financial support was granted by the German Science Foundation (DFG) within the DFGNSF cooperation project (DFG Proj. Nr. STA 324/49-1 and EI 317/6-1) in the frame of the “Materials World Network” and in the frame of the priority program SPP 1369 “Polymer-Solid Contacts: Interfaces and Interphases” (DFG Proj. Nr. STA 324/37-1). This work was supported in part by the National Science Foundation (NSF) through the Center 14

ACS Paragon Plus Environment

Page 14 of 32

Page 15 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

for Nanohybrid Functional Materials (EPS-1004094).

Supporting Information Available The following files are available free of charge. Supporting Information containing: Figure S1: Exemplary data of the temperature response of a bare QCM-D Sensor; Figure S2: Plot of the shift of the calculated Tc ; Figure S3: Modeled Thickness Data; Figure S4+S5: Swelling behavior at lower grafting density

References (1) Schild, H. G.; Tirrell, D. A. Microcalorimetric detection of lower critical solution temperatures in aqueous polymer solutions. J. Phys. Chem. 1990, 94, 4352–4356. (2) Balamurugan, S.; Mendez, S.; Balamurugan, S. S.; O’Brien, M. J.; L´opez, G. P. Thermal Response of Poly(N-isopropylacrylamide) brushes probed by surface plasmon resonance. Langmuir 2003, 19, 2545–2549. (3) Kooij, E.; Sui, X.; Hempenius, M. A.; Zandvliet, H. J. W.; Vancso, G. J. Probing the Thermal Collapse of Poly(N-isopropylacrylamide) Grafts by Quantitative in Situ Ellipsometry. J. Phys. Chem. B 2012, 116, 9261–9268. (4) Bittrich, E.; Burkert, S.; M¨ uller, M.; Eichhorn, K.-J.; Stamm, M.; Uhlmann, P. Temperature-Sensitive Swelling of Poly(N-isopropylacrylamide)Brushes with Low Molecular Weight and Grafting Density. Langmuir 2012, 28, 3439–3448. (5) Ebara, M.; Yamato, M.; Aoyagi, T.; Kikuchi, A.; Sakai, K.; Okano, T. TemperatureResponsive Cell Culture Surfaces Enable “On-Off” Affinity Control between Cell Integrins and RGDS Ligands. Biomacromolecules 2004, 5, 505–510. (6) Ionov, L.; Synytska, A.; Diez, S. Temperature-Induced Size-Control of Bioactive Surface Patterns. Adv. Funct. Mater. 2008, 18, 1501–1508. (7) Burkert, S.; Bittrich, E.; Kuntzsch, M.; M¨ uller, M.; Eichhorn, K.-J.; Bellmann, C.; Uhlmann, P.; Stamm, M. Protein Resistance of PNIPAAm Brushes: Application to Switchable Protein Adsorption. Langmuir 2010, 26, 1786–1795. (8) Zhang, Y.; Furyk, S.; Sagle, L. B.; Cho, Y.; Bergbreiter, D. E.; Cremer, P. S. Effects of Hofmeister Anions on the LCST of PNIPAM as a Function of Molecular Weight. J. Phys. Chem. C 2007, 111, 8916–8924.

15

ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(9) Du, H.; Wickramasinghe, R.; Qian, X. Effects of Salt on the Lower Critical Solution Temperature of Poly (N-Isopropylacrylamide). J. Phys. Chem. B 2010, 114, 16594– 16604. (10) Ataman, M. Properties of aqueous salt solutions of poly(ethylene oxide). Cloud points, θ temperatures. Colloid Polym. Sci. 1987, 265, 19–25. (11) Hofmeister, F. Zur Lehre von der Wirkung der Salze. 1888, 24, 247–260. (12) Von Hippel, P. H.; Schleich, T. Ion effects on the solution structure of biological macromolecules. Accounts Chem. Res. 1969, 2, 257–265. (13) Gurau, M. C.; Lim, S.-M.; Castellana, E. T.; Albertorio, F.; Kataoka, S.; Cremer, P. S. On the Mechanism of the Hofmeister Effect. J. Am. Chem. Soc. 2004, 126, 10522– 10523. (14) Kunz, W.; Lo Nostro, P.; Ninham, B. W. The present state of affairs with Hofmeister effects. Curr. Opin. Colloid In. 2004, 9, 1–18. (15) Florin, E.; Kjellander, R.; Eriksson, J. C. Salt effects on the cloud point of the poly(ethylene oxide)+ water system. J. Chem. Soc. Farad. T. 1 1984, 80, 2889–2910. (16) Zhang, Y.; Furyk, S.; Bergbreiter, D. E.; Cremer, P. S. Specific Ion Effects on the Water Solubility of Macromolecules: PNIPAM and the Hofmeister Series. J. Am. Chem. Soc. 2005, 127, 14505–14510. (17) Gibb, C. L. D.; Gibb, B. C. Anion Binding to Hydrophobic Concavity Is Central to the Salting-in Effects of Hofmeister Chaotropes. J. Am. Chem. Soc. 2011, 133, 7344–7347. (18) Thormann, E. On understanding of the Hofmeister effect: how addition of salt alters the stability of temperature responsive polymers in aqueous solutions. RSC Adv. 2012, 2, 8297–8305. (19) Von Hippel, P. H.; Peticolas, V.; Schack, L.; Karlson, L. Model studies on the effects of neutral salts on the conformational stability of biological macromolecules. I. Ion binding to polyacrylamide and polystyrene columns. Biochemistry 1973, 12, 1256–1264. (20) Cohen Stuart, M. A.; Huck, W. T.; Genzer, J.; M¨ uller, M.; Christopher, O.; Stamm, M.; Sukhorukov, G.; Szleifer, I.; Tsukruk, V. V.; Urban, M.; Winnik, F.; Zauscher, S.; Luzinov, I.; Minko, S. Emerging applications of stimuli-responsive polymer materials. Nat. Mater. 2010, 9, 101–113. (21) Azzaroni, O. Polymer brushes here, there, and everywhere: Recent advances in their practical applications and emerging opportunities in multiple research fields. J. Polym. Sci. Pol. Chem. 2012, 50, 3225–3258. (22) Ayres, N. Polymer brushes: Applications in biomaterials and nanotechnology. Polym. Chem. 2010, 1, 769–777.

16

ACS Paragon Plus Environment

Page 16 of 32

Page 17 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

(23) Chen, T.; Ferris, R.; Zhang, H.; Ducker, R.; Zauscher, S. Stimulus-responsive polymer brushes on surfaces: Transduction mechanisms and applications. Prog. Polym. Sci. 2010, 35, 94–112. (24) Brittain, W. J.; Minko, S. A Structural Definition of Polymer Brushes. J. Polym. Sci. Polym. Chem. 2007, 45, 3505–3512. (25) Zhulina, E. B.; Borisov, O. V.; Pryamitsyn, V. A.; Birshtein, T. M. Coil-Globule Type Transitions in Polymers. 1. Collapse of Layers of Grafted Polymer Chains. Macromolecules 1991, 24, 140–149. (26) Szleifer, I.; Carignano, M. A. Advances in Chemical Physics; John Wiley & Sons, Inc., 2007; pp 165–260. (27) Jhon, Y. K.; Bhat, R. R.; Jeong, C.; Rojas, O.; Szleifer, I.; Genzer, J. Salt induced depression of LCST in a surface-grafted neutral thermoresponsive polymer. Macromol. Rapid. Commun. 2006, 27, 697–701. (28) Ishida, N.; Biggs, S. Salt-Induced Structural Behavior for Poly(N-isopropylacryamide) Grafted onto Solid Surface Observed Directly by AFM and QCM-D. Macromolecules 2007, 40, 9045–9052. (29) Alem, H.; Jonas, A. M.; Demoustier-Champagne, S. Poly(N-isopropylacrylamide) grafted into nanopores: Thermo-responsive behaviour in the presence of different salts. Polym. Degrad. Stabil. 2010, 95, 327–331. (30) Naini, C. A.; Thomas, M.; Franzka, S.; Frost, S.; Ulbricht, M.; Hartmann, N. Hofmeister Effect of Sodium Halides on the Switching Energetics of Thermoresponsive Polymer Brushes. Macromol. Rapid Commun. 2013, 34, 417–422. (31) Zhao, X.-J.; Gao, Z.-F. Role of hydrogen bonding in solubility of poly(Nisopropylacrylamide) brushes in sodium halide solutions. Chinese Phys. B 2016, 25, 074703–1–074703–9. (32) Murdoch, T. J.; Humphreys, B. A.; Willott, J. D.; Gregory, K. P.; Prescott, S. W.; Nelson, A.; Wanless, E. J.; Webber, G. B. Specific Anion Effects on the Internal Structure of a Poly(N-isopropylacrylamide) Brush. Macromolecules 2016, 49, 6050–6060. (33) Humphreys, B. A.; Willott, J. D.; Murdoch, T. J.; Webber, G. B.; Wanless, E. J. Specific ion modulated thermoresponse of poly(N-isopropylacrylamide) brushes. Phys. Chem. Chem. Phys. 2016, 18, 6037–6046. (34) Christau, S.; Moeller, T.; Genzer, J.; Koehler, R.; von Klitzing, R. Salt-Induced Aggregation of Negatively Charged Gold Nanoparticles Confined in a Polymer Brush Matrix. Macromolecules 2017, 50, 7333–7343. (35) Bittrich, E.; Rodenhausen, K.; Eichhorn, K.-J.; Hofmann, T.; Schubert, M.; Stamm, M.; Uhlmann, P. Protein adsorption on and swelling of polyelectrolyte brushes: A simultaneous ellipsometry-quartz crystal microbalance study. Biointerphases 2010, 5, 159–167. 17

ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(36) Rodenhausen, K.; Schubert, M. Virtual separation approach to study porous ultra-thin films by combined spectroscopic ellipsometry and quartz crystal microbalance methods. Thin Solid Films 2011, 519, 2772–2776. (37) Adam, S.; Koenig, M.; Rodenhausen, K.; Eichhorn, K.-J.; Oertel, U.; Schubert, M.; Stamm, M.; Uhlmann, P. Quartz crystal microbalance with coupled Spectroscopic Ellipsometry-study of temperature-responsive polymer brush systems. Appl. Surf. Sci. 2017, (38) Wang, X.; Qiu, X.; Wu, C. Comparison of the Coil-to-Globule and the Globule-to-Coil Transitions of a Single Poly(N-isopropylacrylamide) Homopolymer Chain in Water. Macromolecules 1998, 31, 2972–2976. (39) 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. (40) Lu, Y.; Zhou, K.; Ding, Y.; Zhang, G.; Wu, C. Origin of hysteresis observed in association and dissociation of polymer chains in water. Phys. Chem. Chem. Phys. 2010, 12, 3188–3194. (41) Domack, A.; Prucker, O.; R¨ uhe, J.; Johannsmann, D. Swelling of a polymer brush probed with a quartz crystal resonator. Phys. Rev. E 1997, 56, 680–689. (42) St˚ algren, J.; Eriksson, J.; Boschkova, K. A Comparative Study of Surfactant Adsorption on Model Surfaces Using the Quartz Crystal Microbalance and the Ellipsometer. J. Colloid Interf. Sci. 2002, 253, 190–195. (43) Shechter, I.; Ramon, O.; Portnaya, I.; Paz, Y.; Livney, Y. D. Microcalorimetric Study of the Effects of a Chaotropic Salt, KSCN, on the Lower Critical Solution Temperature (LCST) of Aqueous Poly(N-isopropylacrylamide) (PNIPA) Solutions. Macromolecules 2010, 43, 480–487. (44) Paz, Y.; Kesselman, E.; Fahoum, L.; Portnaya, I.; Ramon, O. The interaction between poly(N-isopropylacrylamide) and salts in aqueous media: The salting-out phenomenon as studied by attenuated total reflection/fourier transform infrared spectroscopy. J. Polym. Sci. Pol. Phys. 2004, 42, 33–46. (45) Rauch, S.; Uhlmann, P.; Eichhorn, K.-J. In situ spectroscopic ellipsometry of pHresponsive polymer brushes on gold substrates. Anal. Bioanal. Chem. 2013, 405, 9061– 9069. (46) Rauch, S.; Eichhorn, K.-J.; Oertel, U.; Stamm, M.; Kuckling, D.; Uhlmann, P. Temperature responsive polymer brushes with clicked rhodamine B: synthesis, characterization and swelling dynamics studied by spectroscopic ellipsometry. Soft Matter 2012, 8, 10260–10270.

18

ACS Paragon Plus Environment

Page 18 of 32

Page 19 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

(47) Bae, Y. H.; Okano, T.; Kim, S. W. Temperature dependence of swelling of crosslinked poly(N,N-alkyl substituted acrylamides) in water. Journal of Polymer Science Part B: Polymer Physics 1990, 28, 923–936. (48) Johs, B.; Hale, J. S. Dielectric function representation by B-splines. Phys. Status Solidi A 2008, 205, 715–719. (49) Voinova, M.; Rodahl, M.; Jonson, M.; Kasemo, B. Viscoelastic Acoustic Response of Layered Polymer Films at Fluid-Solid Interfaces: Continuum Mechanics Approach. Physica Scripta 1999, 59, 391–396.

19

ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 1: Exemplary data of QCM and SE measurements of the salt and temperature dependent swelling of PNIPAAm brushes; (a) ∆F of the seventh overtone, (b) ∆D of the seventh overtone, (c) Ψ at λ=501 nm, (d) ∆ at λ=501nm 289x203mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 20 of 32

Page 21 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Figure 1: Exemplary data of QCM and SE measurements of the salt and temperature dependent swelling of PNIPAAm brushes; (a) ∆F of the seventh overtone, (b) ∆D of the seventh overtone, (c) Ψ at λ=501 nm, (d) ∆ at λ=501nm 289x203mm (300 x 300 DPI)

ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 1: Exemplary data of QCM and SE measurements of the salt and temperature dependent swelling of PNIPAAm brushes; (a) ∆F of the seventh overtone, (b) ∆D of the seventh overtone, (c) Ψ at λ=501 nm, (d) ∆ at λ=501nm 289x203mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 22 of 32

Page 23 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Figure 1: Exemplary data of QCM and SE measurements of the salt and temperature dependent swelling of PNIPAAm brushes; (a) ∆F of the seventh overtone, (b) ∆D of the seventh overtone, (c) Ψ at λ=501 nm, (d) ∆ at λ=501nm 289x203mm (300 x 300 DPI)

ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 2: Difference between heating and cooling cycle in exemplary data of QCM and SE measurements of the salt and temperature dependent swelling of PNIPAAm brushes; (a) ∆F of the seventh overtone, (b) ∆D of the seventh overtone, (c) Ψ at λ=501 nm, (d) ∆ at λ=501nm 290x203mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 24 of 32

Page 25 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Figure 2: Difference between heating and cooling cycle in exemplary data of QCM and SE measurements of the salt and temperature dependent swelling of PNIPAAm brushes; (a) ∆F of the seventh overtone, (b) ∆D of the seventh overtone, (c) Ψ at λ=501 nm, (d) ∆ at λ=501nm 290x203mm (300 x 300 DPI)

ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 2: Difference between heating and cooling cycle in exemplary data of QCM and SE measurements of the salt and temperature dependent swelling of PNIPAAm brushes; (a) ∆F of the seventh overtone, (b) ∆D of the seventh overtone, (c) Ψ at λ=501 nm, (d) ∆ at λ=501nm 290x203mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 26 of 32

Page 27 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Figure 2: Difference between heating and cooling cycle in exemplary data of QCM and SE measurements of the salt and temperature dependent swelling of PNIPAAm brushes; (a) ∆F of the seventh overtone, (b) ∆D of the seventh overtone, (c) Ψ at λ=501 nm, (d) ∆ at λ=501nm 290x203mm (300 x 300 DPI)

ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 3: Effect of salt and temperature on the amount of aqueous solution in PNIPAAm brushes; amount of aqueous solution detected by SE (a) and QCM-D (b), (c) difference in amount of aqueous solution detected by QCM-D and SE 290x203mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 28 of 32

Page 29 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Figure 3: Effect of salt and temperature on the amount of aqueous solution in PNIPAAm brushes; amount of aqueous solution detected by SE (a) and QCM-D (b), (c) difference in amount of aqueous solution detected by QCM-D and SE 287x201mm (300 x 300 DPI)

ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 3: Effect of salt and temperature on the amount of aqueous solution in PNIPAAm brushes; amount of aqueous solution detected by SE (a) and QCM-D (b), (c) difference in amount of aqueous solution detected by QCM-D and SE 287x201mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 30 of 32

Page 31 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Figure 4: Effect of salt and temperature on the viscosity of PNIPAAm brushes 290x203mm (300 x 300 DPI)

ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Table of Contents Graphic 84x49mm (150 x 150 DPI)

ACS Paragon Plus Environment

Page 32 of 32