Enhanced Water Retention Maintains Energy Dissipation in

May 10, 2018 - ... in dry weather when the indoor R.H. was also measured to be ∼16%. ... Derivative thermogravimetric analysis (DTG) was performed u...
2 downloads 0 Views 1MB Size
Subscriber access provided by Universitaetsbibliothek | Johann Christian Senckenberg

Article

Enhanced water retention maintains energy dissipation in dehydrated metal-coordinate polymer networks: Another role for Fe-catechol crosslinks? Sungjin Kim, Amy M Peterson, and Niels Holten-Andersen Chem. Mater., Just Accepted Manuscript • DOI: 10.1021/acs.chemmater.7b05246 • Publication Date (Web): 10 May 2018 Downloaded from http://pubs.acs.org on May 10, 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 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 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.

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 26 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

Chemistry of Materials

Submitted to

1. Introduction Understanding the process by which biological materials are transformed from wet, compliant and soft into dry, stiff and hard materials is not only intriguing from a fundamental scientific perspective, but also important from the viewpoint of bioinspired sustainable materials engineering. While bones and shells are well-studied materials resulting from one such process (i.e., biomineralization1–7), wood,4,8–11 the outer layer of horns,12 fingernails,13 marine worm jaws,14–17 insect cuticles,18–21 and squid beaks18,22–24 are examples of materials solidified via a synergistic process of macromolecular crosslinking and dehydration, in which little to no mineralization occurs.19,24,25 The resulting mechanical properties of some of these unmineralized materials are comparable to biomineralized tissues such as bone or dentin, but with lower density.17,18 Nature also displays examples of loadbearing materials that experience repeated dehydrationrehydration cycles throughout their functional lifetime. For example, the byssal threads of intertidal mussels experience dehydrating-rehydrating dynamic conditions daily due to tidal cycles.26–28 Yet, to our knowledge, only a few studies29,30 have explored the effect of dehydration on mussel thread mechanical properties. In general, the mechanical robustness of the mussel holdfast is caused by the high energy dissipation of its threads31 and by the strong adhesion of the plaque that firmly tethers each thread to the substrate,32,33 both material properties which are partially attributed to the presence of Fe-catechol coordinate crosslinks.34–38 Inspired by their apparent role in the robust mussel thread mechanics, Fe-catechol coordinate crosslink dynamics have been explored as an energy-dissipative crosslinking mechanism in the engineering of tough hydrogels.39,40 However, the contribution of metal-coordinate crosslink dynamics to the mechanics of mussel threads, or mussel-inspired polymer hydrogels, upon dehydration, has not yet been 2 ACS Paragon Plus Environment

Chemistry of Materials 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 2 of 26

Submitted to

explored.29,30,41 Using 4-arm-PEG-catechol (4cPEG) hydrogels, herein we investigate if Fecatechol crosslink dynamics continue to impact network mechanics upon gel dehydration.42,43 We present evidence to suggest that, in addition to their well-known dynamic mechanical contributions to polymer networks swollen in water, Fe-catechol coordinate crosslinks also significantly increase energy dissipation of dehydrated polymer networks. To explain these observations, we present additional evidence to support the hypothesis that the Fe-catechol crosslinks remain dynamic upon bulk network dehydration due to local water binding.

2. Experimental Section 2.1. Materials. 10 kDa 4-arm-PEG-NHS (4aPEG) was purchased from JenKem Technology USA, Inc. Iron(III) chloride hexahydrate (FeCl3∙6H2O) was purchased from Fluka. All other chemicals were purchased from Sigma-Aldrich unless indicated otherwise. 2.2. Synthesis of 4cPEG. 4-arm-PEG-catechol (4cPEG) was synthesized with some modifications to a procedure reported elsewhere.44 In short, to substitute –NHS to –catechols 2 g of 4aPEG was mixed with dopamine hydrochloride (dopa-HCl) (1.5× molar relative to NHS) in a round bottom flask, dissolved by adding 10 mL of anhydrous N,Ndimethylformamide (DMF) whereafter the reaction was started by adding triethylamine (TEA) (2.5× molar equivalent relative to –NHS). The reaction was protected with N2 gas and allowed to proceed for 12 hours at 55 ºC using a silicone oil bath. The reaction was stopped via cooling to room temperature. Then, 20 mL of chloroform and 20 mL of water was added so that salts in the aqueous phase and polymer product in the organic phase can be separated. After extracting the organic phase from the mixture, Na2SO4 was added to this organic solution to get rid of remnant water. The organic solvent was then vaporized by rotary evaporation. The product was then purified via precipitating in diethylether, followed by re-dissolving in 5 mL 3 ACS Paragon Plus Environment

Page 3 of 26 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

Chemistry of Materials

Submitted to

of dichloromethane (DCM). After repeating this cycle three or more times, the precipitate was dried under vacuum. Finally, the precipitate was re-dissolved in Milli-Q (DI) water, frozen under – 80 ℃, then lyophilized to obtain the final product. 2.3. Formation of Fe-4cPEG hydrogels. Fe-4cPEG hydrogels were formed by mixing aqueous solutions of 4cPEG (200 mg/mL) and FeCl3∙6H2O (400 mM) in a [catechol]:[Fe3+] ratio of 1:1. Dark green hydrogels instantly formed via Fe3+-oxidation-induced crosslinking of catechols, and after adjusting the gel pH to 3 by adding NaOH (200 mM) we denoted these gels as “Fe (low pH)”. To generate “Fe (high pH)” gels the pH was instead increased to 12 by adding NaOH (400 mM) quickly after mixing 4cPEG and Fe3+, which immediately produced dark red gels upon further mixing. The final concentrations of NaOH in the hydrogels were 142 mM for Fe (low pH) and 571 mM for Fe (high pH). The final concentrations of 4cPEG and Fe3+ were 142.86 mg/mL and 57.1 mM, respectively, for all types of Fe-4cPEG hydrogels. 2.4. Formation of NaIO4–induced covalently crosslinked 4cPEG hydrogels. To form the covalent-only crosslinked 4cPEG network, 4cPEG solution (200 mg/mL) was mixed with an equal volume of the oxidant NaIO4 (40mM) in a [catechol]:[IO4-] ratio of 2:1. The mixture immediately developed an orange color, and at least 6 hours of curing time was allowed for covalent network crosslinking before further characterization. The resulting gels were denoted as “NaIO4” gels. 2.5. UV-Vis Spectroscopy. The UV-spectroscopic measurement was performed on hydrogels using a DS-11 Spectrophotometer/Fluorometer (DeNovix) in microvolume measurement mode. Approximately 1 μL of hydrogel specimen was placed onto the lower stage surface; then the top arm was lowered for the top stage surface to contact the sample, followed by the measurement. DI water, loaded in the same way, was used to set the baseline of spectra. The UV-Vis absorbance spectra were obtained between 250 nm to 650 nm. 4 ACS Paragon Plus Environment

Chemistry of Materials 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 4 of 26

Submitted to

2.6. Tensile Test. To obtain Young’s moduli (𝐸 ) of dehydrated gels, tensile tests were performed using a tensile tester Z2.5 (Zwick/Roell) with a 20 N load cell. The initially wet specimens were dehydrated for 48 hours in dog bone or bar-shaped acrylic molds under 16 % R.H. at 20 ºC in a desiccator unless specified otherwise. The measurement was performed under constant strain rate of 10 mm/min (i.e., ~ 0.11 s-1 for the 1.5 mm measured sample length) and data were collected in the range of 0.1 % maximum strain. Young’s moduli were calculated at 0.06 % strain in the linear region. The modulus was calculated by the following equation: 𝐸=

𝜎 𝜀

(1)

where 𝐸 is the Young’s modulus, 𝜎 is the stress, and 𝜀 is the strain. Experiments were performed in triplicate. We also note that tensile tests and other characterization under ambient conditions were performed in dry weather when the indoor R.H. was also measured to be ~ 16 %. 2.7. Dynamic Mechanical Analysis (DMA). To assess compressive storage and loss moduli (𝐾′ and 𝐾′′, respectively) and loss factor (tan 𝛿 = 𝐾′′/𝐾 ′ ) of the specimens, a DMA 242 E Artemis (NETZSCH-Gerätebau) was used. An individual hydrogel sample was pre-dehydrated for at least 1 hour within the DMA furnace (25 ºC, 0 % R.H., air atmosphere) before measurement to prevent damage of the initially soft sample by the load applied by the compression geometry. The isothermal analysis was followed for 9 or more hours at 1 Hz frequency (max amplitude: 240 µm, proportional factor: 1.1) at the same humidity and temperature in the furnace (i.e., 25 ºC, 0 % R.H., air atmosphere). Data were collected for 1 hour after a total of 9 hours of dehydration after the moduli had reached plateau values. Experiments were performed in triplicate. 2.8. Dehydration Test. To compare the dehydration profiles of Fe-catechol gels (i.e., 4cPEG, Fe (low pH), Fe (high pH)), samples (17.5 μL) were dehydrated in a desiccator with constant 5 ACS Paragon Plus Environment

Page 5 of 26 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

Chemistry of Materials

Submitted to

R.H. of 16 % at 20 ºC and weighed using a balance AG 204 (Mettler Toledo) at different time points during up to 50 hours of dehydration. The weight was measured instantly after taking the sample out of the desiccator to minimize possible re-absorption of moisture. The mass loss, i.e., the ratio between the mass of lost water and the mass of the initial hydrogel (𝛥𝑚/𝑚𝑖 , wt %) was calculated using the following equation: Mass Loss (wt %) =

𝛥𝑚 𝑚𝑖 − 𝑚 × 100 = × 100 𝑚𝑖 𝑚𝑖

(2)

where 𝛥𝑚 is the mass of lost water, 𝑚𝑖 is the initial mass of the hydrogel, and 𝑚 is the mass of the gel at the selected time point. Note that the full dehydration profile for covalent hydrogels produced by NaIO4-oxidation is omitted, since their initial composition and gelation process is different from all other samples, which prevented meaningful comparison. Experiments were performed in triplicate. 2.9. Thermogravimetric Analysis (TGA). A TGA Q500 (TA Instruments) was used to obtain the dehydration profiles upon temperature increase of samples prepared as described above. Samples were heated from 23 ºC to 300 ºC at a rate of 5 ºC/min in the furnace (0 % R.H.) with an N2 atmosphere unless indicated otherwise. The mass loss was calculated by equation (2). Derivative thermogravimetric analysis (DTG) was performed using the TA Analysis 2000 Software (TA Instruments) to examine the rate of mass loss (%) as a function of temperature (dm/dT, %/ºC) and time (dm/dt, %/min). 2.10. Dissolution Test. Dissolution tests were performed by immersing gels in over 50× volume of DI water for over 40 hours. After carefully removing the excess aqueous solution, the mass fraction (Ф, %) of the permanent (Ф𝑝 ) and transient (Ф𝑡 ) fractions of the gels were determined by the following equations: Ф𝑝 =

𝑚𝑑 × 100 𝑚0

(3)

6 ACS Paragon Plus Environment

Chemistry of Materials 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 26

Submitted to

Ф𝑡 =

𝑚0 − 𝑚𝑑 × 100 𝑚0

(4)

where 𝑚0 is the dehydrated mass of a sample before dissolution and 𝑚𝑑 is the dehydrated mass after any dissolution. The dehydrated mass (𝑚𝑑 ) was obtained by dehydrating the wet, swollen gel using the standard dehydration (20 ºC, 16 % R.H. for 48 hours). After dissolution tests, the samples were subjected to additional TGA testing with the same experimental conditions as described before. Experiments were performed in triplicate. 2.11. Raman Spectroscopy. Raman spectra were obtained using a Raman Spectrometer – LabRAM Raman confocal microscope (HORIBA Jobin Yvon). A 785 nm near-IR laser excitation was used in combination with a 10x lens. Samples were loaded on a glass substrate, which was placed on a Märzhäuser stage (Märzhäuser Wetzlar), followed by positioning under the microscope lens. The laser power was 30 mW. The grating was 600 g/mm, and the filter was adjusted to 10 %. The integration time was 0.5 s (or longer, up to 20 s, to increase the signal intensity) with an accumulation of 60 times. The samples were dehydrated for 24 hours at 16 % R.H. and 20 ºC before the measurement. Measurements and data collection were performed using LabSpec 6 (Horiba Scientific) software.

3. Results 3.1. Crosslinking of PEG-catechol Polymer Networks. We follow a PEG-catechol hydrogel assembly method inspired by the proposed mussel thread processing pathway (Figure S1),45–48 with an initial low pH mixing step at a 1:1 molar ratio between Fe3+ and catechol and subsequent Fe-catechol coordinate crosslink formation induced by a pH jump.48 For comparison, we also assembled PEG-catechol hydrogels with only covalent crosslinks, using NaIO4 as the oxidizer and without pH jump.42 The two different 7 ACS Paragon Plus Environment

Page 7 of 26 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

Chemistry of Materials

Submitted to

hydrogels are labeled “Fe (high pH)”, for the mussel-inspired hybrid network produced from the Fe-mediated crosslinking resulting in both covalent and coordination crosslinks, and “NaIO4”, for the only covalently crosslinked network formed by NaIO4-induced oxidative crosslinking (Figure 1a). Please see Experimental Section for details of the hydrogel processing methods.

Figure 1 (a) The polymer crosslinks formed via mussel-inspired Fe3+-mediated network crosslinking45– 48 - Fe (high pH) (left) and via catechol oxidation using NaIO4 (right). Note that the illustration is conceptual and not meant to represent the detailed chemical structure of the network. For more details, please see Fig. S1 (b) Young’s moduli (E, left axis, solid columns) and loss factors (tan δ, right axis, dotted columns) of the Fe (high pH) (red) and NaIO4 (orange) hydrogels after dehydration obtained using tensile testing and compressive DMA, respectively.

8 ACS Paragon Plus Environment

Chemistry of Materials 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 8 of 26

Submitted to

Figure 2 UV-Vis absorbance spectra of hydrogel samples before dehydration; Fe (high pH) (red), Fe (low pH) (green), NaIO4 (orange) and a pure solution of 4cPEG (navy). The insets are photographs of the corresponding hydrogels.

The two distinct hydrogel crosslinking processes were confirmed by UV spectroscopy (Figure 2 and S1). An UV absorbance maximum ~ 480 nm is evidence of the Fe-catechol charge transfer complex,43,49 which is clearly seen in the Fe (high pH) hydrogel, but is absent in the NaIO4 hydrogel. This observation is in agreement with a high pH-induced tris-catechol-Fe coordination complex. The catechol absorbance peak ~ 280 nm,49,50 observed in the Fe (high pH) gel, and a pure 4cPEG control, shifts to a slightly shorter wavelength ~ 266 nm in the NaIO4 hydrogel spectrum, which is in agreement with the oxidative covalent coupling of catechol.49 The additional shoulder at ~ 395 nm in the NaIO4 spectrum likely corresponds to quinone groups,50 another indicator of oxidation-induced covalent crosslinking of catechols. 9 ACS Paragon Plus Environment

Page 9 of 26 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

Chemistry of Materials

Submitted to

The peak at ~ 320 nm in a hydrogel control sample with a 1:1 Fe:catechol ratio but no pH jump (labeled Fe (low pH)), also supports oxidation-induced covalent crosslinking of catechols via DOPA-indole derivatives or α,β-dehydro-DOPA as reported in melanin formation or insect cuticle sclerotization, respectively.42 Although it should be noted that the absorbance spectra of catechol compounds in the UV range have not been fully characterized, these results support covalent catechol-based crosslinking in both Fe (high pH) and NaIO4 gels. 3.2. Mechanical Properties of the Dehydrated Crosslinked Networks To compare the mechanical properties of the Fe (high pH) and NaIO4 gels after dehydration, we utilized tensile testing and dynamic mechanical analysis (DMA) in compression (Figure 1b, S2). Tensile tests were performed on samples after 48 hours of dehydration at 20 ºC, 16 % R.H. (See Experimental Section for details on dehydration conditions). Fe (high pH) networks display a Young’s modulus of ~ 1.8 GPa, approximately twice that of NaIO4. In DMA, the Fe (high pH) similarly show higher compressive storage modulus as well as higher loss modulus (Figure S3a) and loss factor (Figure 1b, S3b). In fact, with a loss factor above 0.7, the dehydrated Fe (high pH) gels display good mechanical damping properties.51 These results demonstrate that the dehydrated Fe (high pH) network is not only stiffer under both tension and compression, but also significantly more energy-dissipative than a dehydrated covalentonly network (NaIO4). 3.3. Water Retention by Fe-Catechol Coordination To better understand the enhanced energy-dissipative properties of the dehydrated Fe (high pH) network, we investigated its dehydration profile in comparison to the Fe (low pH) hydrogel, as well as pure 4cPEG (Figure 3, S4, S5). We used Fe (low pH) in lieu of NaIO4 as a comparison to the dehydration kinetics of a covalent network, to keep the gelation time and the material composition similar during the dehydration. The Fe (low pH) and pure 4cPEG display a mass 10 ACS Paragon Plus Environment

Chemistry of Materials 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 10 of 26

Submitted to

loss equivalent to the original amount of water present in the hydrated gels within 24 hours of dehydration, whereas Fe (high pH) gels show slower and lower mass loss, which suggests increased water retention. We note that the NaIO4 gel also reached its maximum possible water mass loss within 4 hours (Figure S4a). TGA further confirms the increased water retention of Fe (high pH) upon heating to 300 ºC (Figure 3b). To investigate the role of Fe-catechol coordination in this water-retention effect, we performed a control experiment on a sample of non-functionalized 4-arm-PEG (4aPEG), i.e. without catechol modification, prepared identically to Fe (high pH) (Figure S4). This specimen (4aPEG Fe (high pH)) displayed faster and greater dehydration than the Fe (high pH) gel under both ambient conditions and in a heating scan (Figure S4a and S4b, respectively), ultimately reaching its maximum possible water mass loss. DTG analyses (Figure S5) further confirmed that the Fe (high pH) gel shows the slowest dehydration rate upon heating. These results provide further evidence of the enhanced affinity to water of the Fe (high pH) network. They also indicate that the interaction between water and Fe3+ coordinated with catechols is stronger than the possible water-retaining effect of salts in the Fe-containing networks.

11 ACS Paragon Plus Environment

Page 11 of 26 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

Chemistry of Materials

Submitted to

Figure 3 (a) The dehydration profiles of Fe (high pH) (red), Fe (low pH) (green) hydrogels and a 4cPEG aqueous solution (blue) with equivalent polymer concentration at 20 ℃, 16 % R.H.. The mass loss (wt %) is estimated as the ratio between the lost mass and the initial mass of the hydrogel (equation (2) in the Experimental Section). The horizontal dashed lines indicate the maximum possible water mass loss from each sample calculated from the composition of water and nonwater components of the initial hydrogel. The inset shows the dehydration profile from 10 hr to 50 hr. (b) Mass loss measured by TGA. (c) Raman spectra of the given samples after dehydration. Please see Table 1 for peak assignments.

12 ACS Paragon Plus Environment

Chemistry of Materials 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 12 of 26

Submitted to

Approx. Raman Shift (cm-1) 315

Assignment

Present in

O-Fe-O stretching from Fe-H2O interaction

Fe (high pH)

charge transfer interaction from bidentate chelate

Fe (high pH), Fe (low pH)*

590, 635

interaction between the oxygens of catechols and Fe3+

Fe (high pH), Fe (low pH)*

1250 - 1480

phenyl ring vibrations

Fe (high pH), Fe (low pH), 4cPEG

530

Note

* broad, undistinguished peak

Table 1 Assignment of the observed Raman peaks in Figure 3c. The vibration band from the O-Fe-O deformation at ~ 315 cm-1 attributed to the Fe-H2O interaction52–54 is only present in the Fe (high pH) sample.

To explore the possible molecular mechanism underlying the increased water retention of Fe (high pH) networks, we performed Raman spectroscopic analyses on dehydrated gels of Fe (high pH), Fe (low pH), and dried out 4cPEG (Figure 3c). The measured vibration bands for catechols agree well with previous studies43,50 (Table 1): (i) phenyl ring vibrations from ~ 1250 to ~ 1480 cm-1, (ii) distinct coordination peaks at ~ 590 and ~ 635 cm-1 from the interactions between the oxygen of catechols and Fe3+, as well as a peak at ~ 530 cm-1 from the charge transfer interaction from bidentate chelates clearly seen in Fe (high pH), whereas (iii) a broad peak from ~ 500 to ~ 650 cm-1 of Fe (low pH) indicates weak coordination between catechol and Fe3+. In addition, a vibration band ~ 315 cm-1 is observed uniquely in the Fe (high pH) network, which correlates with previous peak assignments of Fe-H2O interactions around 315332 cm-1 originating from O-Fe-O stretching.52–54 In another catechol-Fe system, Schmitt et al.54 suggested that the peak at this Raman shift is caused by water molecules filling positions in the first shell of the catechol-Fe coordination sphere. To further support the hypothesis that this peak corresponds to unique Fe-H2O interactions in the catechol-Fe coordination sphere, 13 ACS Paragon Plus Environment

Page 13 of 26 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

Chemistry of Materials

Submitted to

we performed two additional experiments. First, the Fe (high pH) sample was prepared in D2O instead of H2O, which caused a slight downshift of the peak as expected if this peak correlates to water-binding (Figure S6).55 Second, the Fe (high pH) sample was lyophilized, which resulted in an additional ~ 1 % water removal (Figure S7a) in correlation with a decrease in the relative peak intensity at ~ 315 cm-1 (Figure S7b). Finally, to confirm that the proposed enhanced water-binding capability of Fe-catechol coordination complexes is thermodynamic rather than kinetic in nature, we verified that the water retention of Fe (high pH) is indeed independent of temperature increase rates (Figure S8).

Figure 4 (a) Fe (high pH) (left) and NaIO4 (right) immersed in DI water for 40+ hours (upper row) and after the removal of water (lower row). (b) Comparison between the covalently crosslinked permanent fraction Ф𝑝 (solid) and the Fe-catechol coordinate crosslinked transient fraction Ф𝑡 (patterned) of the two networks calculated by equations (3) and (4) in the Experimental Section, respectively. (c) TGA of Fe (high pH) hydrogels before (solid red) or after (dashed red) dissolution of the Ф𝑡 and of a NaIO4 hydrogel (dotted orange). 14 ACS Paragon Plus Environment

Chemistry of Materials 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 14 of 26

Submitted to

Since both the increased mechanical damping and water retention correlate with the presence of Fe-catechol coordinate crosslinks in the dehydrated network, we tested whether removing the Fe-catechol transient network fraction in the Fe (high pH) gel would decrease the water binding capacity upon dehydration of the remaining covalently crosslinked network. By using mass lost by dissolution as a measure of the transient fraction Ф𝑡 of the network, and the remaining undissolved mass as a measure of the permanent covalent fraction Ф𝑝 , we found Ф𝑡 in Fe (high pH) gels to be ~ 67 %, whereas in NaIO4 gels Ф𝑡 ~ 0 % (i.e., Ф𝑝 ~ 100 %) (Figure 4a, 4b). After removal of the Fe-catechol transient network fraction, the water binding ability of the remaining covalent fraction of the Fe (high pH) network decreases and becomes similar to that of a NaIO4 network with only covalent bonds (Figure 4c). This observation further supports that the water retained in the dehydrated Fe (high pH) network is preferentially bound to the transient Fe-catechol crosslinks.

4. Discussion In this work, we report that a hybrid covalent and Fe-catechol coordinate crosslinked PEG network, initially established as a hydrogel via a mussel-inspired two-step crosslinking process, after dehydration displays considerably enhanced stiffness and energy dissipation when compared to a dehydrated NaIO4–induced covalently crosslinked network. A previous study by Barrett et al. likewise reported increased stiffness and energy dissipation of a similar hybrid network compared to a covalent-only network.48 However, that study was performed with hydrogels, whereas we have demonstrated for the first time that these mechanical differences persist even after dehydration. The network shrinkage during dehydration will induce significant physical interactions 15 ACS Paragon Plus Environment

Page 15 of 26 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

Chemistry of Materials

Submitted to

between the transient and permanent network fractions of the hybrid gels, even if, as suggested by the data in Figure 4b, the two fractions are covalently unconnected (see Figure S1a for further details). Since Fe-catechol bonds are stronger than other types of noncovalent interactions,56 the combined coordinate and covalent crosslinks will increase the overall effective crosslink density and thereby increase the elastic modulus of this dehydrated hybrid network compared to the covalent-only network. Observations similar to this synthetic hybrid polymer network demonstration have been documented in the cuticles of the mussel thread and the nereis jaw; in both materials, metal-coordinate crosslinking has been reported to significantly increase stiffness compared to the chelated metal-free covalently crosslinked material.17,54,57 Importantly, in addition to the increased stiffness compared to covalent-only networks, the synthetic hybrid networks also retain an enhanced energy dissipation upon dehydration. We speculate that the proposed water retention helps to preserve a more dynamic energydissipative state of the metal coordinate bonds in the hybrid networks after dehydration. Interestingly, Fe-catechol coordinate bonds were also recently demonstrated to maintain their kinetically labile state and dramatically enhance energy dissipation in a mussel-inspired dehydrated elastomer.58,59 We are currently further investigating the relationship between bond dynamics and the proposed water retention capability of metal-coordinate crosslinks. Using the Fe (high pH) data in Figure 3a, we can crudely estimate the number of water molecules retained in our dehydrated hybrid networks to be ~ 50-fold higher than the equimolar number of Fe3+ or catechols (see SI for details on calculations). While attempted explanations of how this water is bound within the hybrid network remain speculative, our Raman data (Figure 3c) and dehydration control tests (Figure S4 and S5) suggest that catechol-Fe coordination complexes are the main source of the water binding.54 Hence, although the details 16 ACS Paragon Plus Environment

Chemistry of Materials 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 16 of 26

Submitted to

of these proposed water-Fe-catechol coordinate complex interactions are not yet understood, our estimates would suggest multiple layers of hydration shells per coordination complex, which is in agreement with the numbers reported for biomacromolecules in general.60,61 It is possible that part of the population of Fe-catechol complexes are of the bis-type form, wherein water molecules are bound extra tight in the first coordination shell by serving as ligands. It is generally known that the potential energy of water molecules can be significantly lower near the polar or dipolar region of a macromolecule due to their non-random average orientation compared to that in bulk water.62 In addition, if the water molecules interact favorably with the macromolecule surface (e.g., via H-bonds, dipole-dipole interactions, etc.), the water close to the surface becomes very difficult to remove.62,63 Likewise, we infer that the water molecules near a Fe-catechol complex can be strongly bound, possibly via the coordination we propose, effectively resisting external perturbations such as heating (Figure 3b, S4b) or freezing (Figure S7). We also note that in a recent study of another catechol-functionalized polymer material, Ca-catechol ionic complexes were suggested to aggregate upon dehydration to form concentrated metal ion-clusters with a high water uptake.64 The water retention of Fe-catechol coordination complexes presented in our study could plausibly be enhanced by a similar cluster formation induced upon gel dehydration, a mechanism which will be further investigated in future studies. Furthermore, we note that a Raman peak at ~ 315 cm-1 possibly associated with water-Fe-catechol coordinate complex interactions as we propose here, was also observed in the dehydrated Fe-catechol-crosslinked elastomer mentioned above,58 as well as in Fe-treated mussel thread cuticle.37 Hence, the proposed water binding by Fe-catechol coordination complexes may be affecting crosslink bond dynamics, and thereby possibly dehydrated material mechanics, of both native mussel threads and their bio-inspired derivatives.

17 ACS Paragon Plus Environment

Page 17 of 26 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

Chemistry of Materials

Submitted to

5. Conclusion Using the pH-controlled two-step macromolecular crosslinking mechanism inspired by the mussel holdfast, we have presented evidence to suggest that Fe-catechol coordinate crosslinking can increase the energy dissipation in largely dehydrated polymer networks by more than three-fold through enhanced water-retention. The possible interplay between transient crosslink bond dynamics and microscopic water binding of Fe-catechol, and perhaps other metal-ligand coordinate complexes, may allow energy dissipative mechanical properties to be retained during dehydration-processed viscoelastic-to-solid transitions of existing biological and future bioinspired macromolecular materials.

Acknowledgments Financial support of this work by the Office of Naval Research (ONR) under the Young Investigators Program Grant ONR.N00014-15-1-2763 is gratefully acknowledged. This work made use of the MRSEC Shared Experimental Facilities at MIT, supported by the National Science Foundation under Award No. DMR-1419807. The first author thankfully acknowledges financial support by the 12th Samsung Scholarship Program and helps by Hyunwoo Yuk from Prof. Xuanhe Zhao’s Group (Department of Mechanical Engineering, MIT) for tensile tests and Dr. Eugene N. Cho from Prof. Jeffrey C. Grossman’s Group (Department of Materials Science and Engineering, MIT) for TGA. He also thanks Xuejian Lyu from Prof. Amy M. Peterson’s Group (Department of Chemical Engineering, WPI) for his assistance in using DMA. Finally, we thank Prof. Herbert Waite (The Department of Molecular, Cellular, and Developmental Biology, UCSB) and Dr. Emmanouela Filippidi from Prof. Megan T. Valentine’s Group (Department of Mechanical Engineering, UCSB) for constructive discussion. Supporting Information Available for calculation of remnant water molecules and supporting figures (Figure S1-S8)

18 ACS Paragon Plus Environment

Chemistry of Materials 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 18 of 26

Submitted to

References (1)

Evans, J. S. “Tuning in” to Mollusk Shell Nacre- and Prismatic-Associated Protein Terminal Sequences. Implications for Biomineralization and the Construction of High Performance Inorganic - Organic Composites. Chem. Rev. 2008, 108, 4455–4462.

(2)

Westbroek, P.; Marin, F. A Marriage of Bone and Nacre. Nature 1998, 392, 861–862.

(3)

Lakes, R. Materials with Structure Hierarchy. Nature 1993, 361, 511–515.

(4)

Fratzl, P.; Weinkamer, R. Nature’s Hierarchical Materials. Prog. Mater. Sci. 2007, 52, 1263–1334.

(5)

Meyers, M. A.; Chen, P.-Y.; Lin, A. Y.-M.; Seki, Y. Biological Materials: Structure and Mechanical Properties. Prog. Mater. Sci. 2008, 53, 1–206.

(6)

Ortiz, C.; Boyce, M. C. Materials Science. Bioinspired Structural Materials. Science 2008, 319, 1053–1054.

(7)

Kim, S.; Park, C. B. Bio-Inspired Synthesis of Minerals for Energy, Environment, and Medicinal Applications. Adv. Funct. Mater. 2013, 23, 10–25.

(8)

Renaud, M.; Rueff, M.; Rocaboy, A. C. Mechanical Behaviour of Saturated Wood under Compression Part 2: Behaviour of Wood at Low Rates of Strain Some Effects of Compression on Wood Structure. Wood Sci. Technol. 1996, 30, 237–243.

(9)

Ji, B.; Gao, H. Mechanical Principles of Biological Nanocomposites. Annu. Rev. Mater. Res. 2010, 40, 77–100.

(10)

Olsson, R. T.; Azizi Samir, M. A. S.; Salazar-Alvarez, G.; Belova, L.; Ström, V.; Berglund, L. A.; Ikkala, O.; Nogués, J.; Gedde, U. W. Making Flexible Magnetic Aerogels and Stiff Magnetic Nanopaper Using Cellulose Nanofibrils as Templates. Nat. Nanotechnol. 2010, 5, 584–588.

(11)

Alméras, T.; Gronvold, A.; van der Lee, A.; Clair, B.; Montero, C. Contribution of 19 ACS Paragon Plus Environment

Page 19 of 26 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

Chemistry of Materials

Submitted to

Cellulose to the Moisture-Dependent Elastic Behaviour of Wood. Compos. Sci. Technol. 2017, 138, 151–160. (12)

Kitchener, A.; Vincent, J. F. V. Composite Theory and the Effect of Water on the Stiffness of Horn Keratin. J. Mater. Sci. 1987, 22, 1385–1389.

(13)

Farran, L.; Ennos, A. R.; Starkie, M.; Eichhorn, S. J. Tensile and Shear Properties of Fingernails as a Function of a Changing Humidity Environment. J. Biomech. 2009, 42, 1230–1235.

(14)

Broomell, C. C.; Mattoni, M. A.; Zok, F. W.; Waite, J. H. Critical Role of Zinc in Hardening of Nereis Jaws. J. Exp. Biol. 2006, 209, 3219–3225.

(15)

Lichtenegger, H. C.; Schöberl, T.; Ruokolainen, J. T.; Cross, J. O.; Heald, S. M.; Birkedal, H.; Waite, J. H.; Stucky, G. D. Zinc and Mechanical Prowess in the Jaws of Nereis, a Marine Worm. Proc. Natl. Acad. Sci. U. S. A. 2003, 100, 9144–9149.

(16)

Lichtenegger, H. C.; Schöberl, T.; Bartl, M. H.; Waite, H.; Stucky, G. D. High Abrasion Resistance with Sparse Mineralization: Copper Biomineral in Worm Jaws. Science 2002, 298, 389–392.

(17)

Degtyar, E.; Harrington, M. J.; Politi, Y.; Fratzl, P. The Mechanical Role of Metal Ions in Biogenic Protein-Based Materials. Angew. Chem. Int. Ed. Engl. 2014, 53, 12026– 12044.

(18)

Broomell, C. C.; Khan, R. K.; Moses, D. N.; Miserez, A.; Pontin, M. G.; Stucky, G. D.; Zok, F. W.; Waite, J. H. Mineral Minimization in Nature’s Alternative Teeth. J. R. Soc. Interface 2007, 4, 19–31.

(19)

Schaefer, J.; Kramer, K. J.; Garbow, J. R.; Jacob, G. S.; Stejskal, E. O.; Hopkins, T. L.; Speirs, R. D. Aromatic Cross-Links in Insect Cuticle: Detection by Solid-State 13C and 15N NMR. Science 1987, 235, 1200–1204.

(20)

Andersen, S. O.; Hojrup, P.; Roepstorff, P. Insect Cuticular Proteins. Insect Biochem. 20 ACS Paragon Plus Environment

Chemistry of Materials 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 20 of 26

Submitted to

Mol. Biol. 1995, 25, 153–176. (21)

Kerwin, J. L.; Turecek, F.; Xu, R.; Kramer, K. J.; Hopkins, T. L.; Gatlin, C. L.; Yates, J. R. Mass Spectrometric Analysis of Catechol-Histidine Adducts from Insect Cuticle. Anal. Biochem. 1999, 268, 229–237.

(22)

Miserez, A.; Schneberk, T.; Sun, C.; Zok, F. W.; Waite, J. H. The Transition from Stiff to Compliant Materials in Squid Beaks. Science 2008, 319, 1816–1819.

(23)

Miserez, A.; Li, Y.; Waite, J. H.; Zok, F. Jumbo Squid Beaks: Inspiration for Design of Robust Organic Composites. Acta Biomater. 2007, 3, 139–149.

(24)

Miserez, A.; Rubin, D.; Waite, J. H. Cross-Linking Chemistry of Squid Beak. J. Biol. Chem. 2010, 285, 38115–38124.

(25)

Andersen, S. O. Insect Cuticular Sclerotization: A Review. Insect Biochem. Mol. Biol. 2010, 40, 166–178.

(26)

Carrington, E.; Waite, J. H.; Sarà, G.; Sebens, K. P. Mussels as a Model System for Integrative Ecomechanics. Ann. Rev. Mar. Sci. 2015, 7, 443–469.

(27)

Carrington, E. Seasonal Variation in the Attachment Strength of Blue Mussels: Causes and Consequences. Limnol. Oceanogr. 2002, 47, 1723–1733.

(28)

Carrington, E. The Ecomechanics of Mussel Attachment: From Molecules to Ecosystems. Integr. Comp. Biol. 2002, 42, 846–852.

(29)

Aldred, N.; Wills, T.; Williams, D. N.; Clare, A. S. Tensile and Dynamic Mechanical Analysis of the Distal Portion of Mussel (Mytilus Edulis) Byssal Threads. J. R. Soc. Interface 2007, 4, 1159–1167.

(30)

Troncoso, O. P.; Torres, F. G.; Grande, C. J. Characterization of the Mechanical Properties of Tough Biopolymer Fibres from the Mussel Byssus of Aulacomya Ater. Acta Biomater. 2008, 4, 1114–1117.

(31)

Carrington, E.; Gosline, J. M. Mechanical Design of Mussel Byssus: Load Cycle and 21 ACS Paragon Plus Environment

Page 21 of 26 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

Chemistry of Materials

Submitted to

Strain Rate Dependence. American Malacological Bulletin. 2004, pp 135–142. (32)

Lee, B. P.; Messersmith, P. B.; Israelachvili, J. N.; Waite, J. H. Mussel-Inspired Adhesives and Coatings. Annu. Rev. Mater. Res. 2011, 41, 99–132.

(33)

Stewart, R. J.; Ransom, T. C.; Hlady, V. Natural Underwater Adhesives. J. Polym. Sci. Part B Polym. Phys. 2011, 49, 757–771.

(34)

Waite, J. H.; Qin, X. X.; Coyne, K. J. The Peculiar Collagens of Mussel Byssus. Matrix Biol. 1998, 17, 93–106.

(35)

Holten-Andersen, N.; Fantner, G. E.; Hohlbauch, S.; Waite, J. H.; Zok, F. W. Protective Coatings on Extensible Biofibres. Nat. Mater. 2007, 6, 669–672.

(36)

Sun, C.; Waite, J. H. Mapping Chemical Gradients within and along a Fibrous Structural Tissue, Mussel Byssal Threads. J. Biol. Chem. 2005, 280, 39332–39336.

(37)

Harrington, M. J.; Masic, A.; Holten-Andersen, N.; Waite, J. H.; Fratzl, P. Iron-Clad Fibers: A Metal-Based Biological Strategy for Hard Flexible Coatings. Science 2010, 328, 216–220.

(38)

Hwang, D. S.; Zeng, H.; Masic, A.; Harrington, M. J.; Israelachvili, J. N.; Waite, J. H. Protein- and Metal-Dependent Interactions of a Prominent Protein in Mussel Adhesive Plaques. J. Biol. Chem. 2010, 285, 25850–25858.

(39)

Zhao, X. Multi-Scale Multi-Mechanism Design of Tough Hydrogels: Building Dissipation into Stretchy Networks. Soft Matter 2014, 10, 672–687.

(40)

Krogsgaard, M.; Nue, V.; Birkedal, H. Mussel-Inspired Materials: Self-Healing through Coordination Chemistry. Chem. - A Eur. J. 2016, 22, 844–857.

(41)

Gosline, J.; Lillie, M.; Carrington, E.; Guerette, P.; Ortlepp, C.; Savage, K. Elastic Proteins: Biological Roles and Mechanical Properties. Philos. Trans. R. Soc. Lond. B. Biol. Sci. 2002, 357, 121–132.

(42)

Lee, B. P.; Dalsin, J. L.; Messersmith, P. B. Synthesis and Gelation of DOPA22 ACS Paragon Plus Environment

Chemistry of Materials 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 22 of 26

Submitted to

Modified Poly(Ethylene Glycol) Hydrogels. Biomacromolecules 2002, 3, 1038–1047. (43)

Holten-Andersen, N.; Harrington, M. J.; Birkedal, H.; Lee, B. P.; Messersmith, P. B.; Lee, K. Y. C.; Waite, J. H. PH-Induced Metal-Ligand Cross-Links Inspired by Mussel Yield Self-Healing Polymer Networks with near-Covalent Elastic Moduli. Proc. Natl. Acad. Sci. U. S. A. 2011, 108, 2651–2655.

(44)

Grindy, S. C.; Learsch, R.; Mozhdehi, D.; Cheng, J.; Barrett, D. G.; Guan, Z.; Messersmith, P. B.; Holten-Andersen, N. Control of Hierarchical Polymer Mechanics with Bioinspired Metal-Coordination Dynamics. Nat. Mater. 2015, 14, 1210–1216.

(45)

Priemel, T.; Degtyar, E.; Dean, M. N.; Harrington, M. J.; Ghomi, M. Rapid SelfAssembly of Complex Biomolecular Architectures during Mussel Byssus Biofabrication. Nat. Commun. 2017, 8, 14539.

(46)

Yu, J.; Wei, W.; Danner, E.; Ashley, R. K.; Israelachvili, J. N.; Waite, J. H. Mussel Protein Adhesion Depends on Interprotein Thiol-Mediated Redox Modulation. Nat. Chem. Biol. 2011, 7, 588–590.

(47)

Fullenkamp, D. E.; Barrett, D. G.; Miller, D. R.; Kurutz, J. W.; Messersmith, P. B. PHDependent Cross-Linking of Catechols through Oxidation via Fe(3+) and Potential Implications for Mussel Adhesion. RSC Adv. 2014, 4, 25127–25134.

(48)

Barrett, D. G.; Fullenkamp, D. E.; He, L. H.; Holten-Andersen, N.; Lee, K. Y. C.; Messersmith, P. B. PH-Based Regulation of Hydrogel Mechanical Properties through Mussel-Inspired Chemistry and Processing. Adv. Funct. Mater. 2013, 23, 1111–1119.

(49)

Menyo, M. S.; Hawker, C. J.; Waite, J. H. Versatile Tuning of Supramolecular Hydrogels through Metal Complexation of Oxidation-Resistant Catechol-Inspired Ligands. Soft Matter 2013, 9, 10314–10323.

(50)

Li, Q.; Barrett, D. G.; Messersmith, P. B.; Holten-Andersen, N. Controlling Hydrogel Mechanics via Bio-Inspired Polymer–Nanoparticle Bond Dynamics. ACS Nano 2016, 23 ACS Paragon Plus Environment

Page 23 of 26 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

Chemistry of Materials

Submitted to

10, 1317–1324. (51)

Mok, M. M.; Kim, J.; Torkelson, J. M. Gradient Copolymers with Broad Glass Transition Temperature Regions: Design of Purely Interphase Compositions for Damping Applications. J. Polym. Sci. Part B Polym. Phys. 2008, 46, 48–58.

(52)

Jarzcki, A. A.; Anbar, A. D.; Spiro, T. G.; Jarzecki, A.; Anbar, A. D.; Spiro, T. G. DFT Analysis of Fe(H 2 O) 6 3+ and Fe(H 2 O) 6 2+ Structure and Vibrations; Implications for Isotope Fractionation. J. Phys. Chem. A 2004, 108, 2726–2732.

(53)

Best, S. P.; Beattie, J. K.; Armstrong, R. S. Vibrational Spectroscopic Studies of Trivalent Hexa-Aqua-Cations: Single-Crystal Raman Spectra between 275 and 1 200 Cm –1 of the Caesium Alums of Titanium, Vanadium, Chromium, Iron, Gallium, and Indium. J. Chem. Soc., Dalt. Trans. 1984, No. 12, 2611–2624.

(54)

Schmitt, C. N. Z.; Winter, A.; Bertinetti, L.; Masic, A.; Strauch, P.; Harrington, M. J. Mechanical Homeostasis of a DOPA-Enriched Biological Coating from Mussels in Response to Metal Variation. J. R. Soc. Interface 2015, 12, 20150466.

(55)

Brooker, M. H.; Hancock, G.; Rice, B. C.; Shapter, J. Raman Frequency and Intensity Studies of Liquid H2O, H218O and D2O. J. Raman Spectrosc. 1989, 20, 683–694.

(56)

Lee, H.; Scherer, N. F.; Messersmith, P. B. Single-Molecule Mechanics of Mussel Adhesion. Proc. Natl. Acad. Sci. U. S. A. 2006, 103, 12999–13003.

(57)

Broomell, C. C.; Chase, S. F.; Laue, T.; Waite, J. H. Cutting Edge Structural Protein from the Jaws of Nereis Virens. Biomacromolecules 2008, 9, 1669–1677.

(58)

Filippidi, E.; Cristiani, T. R.; Eisenbach, C. D.; Waite, J. H.; Israelachvili, J. N.; Ahn, B. K.; Valentine, M. T. Toughening Elastomers Using Mussel-Inspired Iron-Catechol Complexes. Science 2017, 358, 502–505.

(59)

Winey, K. I. Designing Tougher Elastomers with Ionomers. Science 2017, 358, 449– 450. 24 ACS Paragon Plus Environment

Chemistry of Materials 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 24 of 26

Submitted to

(60)

Laage, D.; Elsaesser, T.; Hynes, J. T. Water Dynamics in the Hydration Shells of Biomolecules. Chem. Rev. 2017, 117, 10694–10725.

(61)

Badger, J. Multiple Hydration Layers in Cubic Insulin Crystals. Biophys. J. 1993, 65, 1656–1659.

(62)

Wolfe, J.; Bryant, G.; Koster, K. L. WHAT IS “UNFREEZABLE WATER”, HOW UNFREEZABLE IS IT AND HOW MUCH IS THERE? CryoLetters 2002, 23, 157– 166.

(63)

Park, S.; Venditti, R. A.; Jameel, H.; Pawlak, J. J. Hard to Remove Water in Cellulose Fibers Characterized by High Resolution Thermogravimetric Analysis - Methods Development. Cellulose 2006, 13, 23–30.

(64)

Kim, C.; Ejima, H.; Yoshie, N. Non-Swellable Self-Healing Polymer with Long-Term Stability under Seawater. RSC Adv. 2017, 7, 19288–19295.

25 ACS Paragon Plus Environment

Page 25 of 26 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

Chemistry of Materials

Submitted to

TOC Image

26 ACS Paragon Plus Environment