Next Article in Journal
Strategies to Fabricate Polypeptide-Based Structures via Ring-Opening Polymerization of N-Carboxyanhydrides
Previous Article in Journal
Photosensitive Layer-by-Layer Assemblies Containing Azobenzene Groups: Synthesis and Biomedical Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Experimental Evidence of Large Amplitude pH Mediated Autonomous Chemomechanical Oscillation

1
Department of Chemistry, Capital Normal University, Beijing 100048, China
2
School of Materials Science and Engineering, Beijing Institute of Technology, Beijing 100081, China
3
Xi’an Satellite Control Center, Xi’an 710043, China
*
Authors to whom correspondence should be addressed.
Co-first author.
Polymers 2017, 9(11), 554; https://doi.org/10.3390/polym9110554
Submission received: 29 September 2017 / Revised: 23 October 2017 / Accepted: 25 October 2017 / Published: 25 October 2017

Abstract

:
Large amplitude autonomous chemomechanical oscillations were observed in a coupled system consisting of a porous pH-responsive hydrogel and a bromate-sulfite-manganese (II) pH oscillatory reaction. The porous structure effectively improves the chemomechanical response speed, and the negative feedback species of the bulk oscillation Mn2+ takes part in the coupling by forming complex and physical crosslinks with the responsive group in the gel. It strengthens the porous gel by forming additional networks, which may contribute to sustaining the long-lasting chemomechanical oscillation. Additionally, the interaction between Mn2+ and the hydrogel alters the period of the oscillatory reaction due to its binding competition with H+, the positive feedback species.

Graphical Abstract

1. Introduction

Stimuli-responsive hydrogels have recently been coupled with oscillatory chemical reactions to produce autonomous chemomechanical materials, which have wide potential applications in many areas [1,2,3,4,5,6,7] such as drug release [1], homeostatic materials [4], and biomimetic actuators [5,6,7]. Autonomous chemomechanical systems can generally be classified into two categories. In the first type, the mechanical oscillations of the gel are merely slaved to the chemical oscillation kinetics. The gel is immersed in oscillatory reaction components [2,8,9,10,11] or the oscillations can take place inside the gel when introducing appropriate reactants into the polymer network [5,6,12,13,14] or when the reactant species act as an active cross-linker of the gel [15]. In the second type, instabilities are generated by the interactions between the gel and non-oscillatory reaction, which can be further classified into two scenarios: “long-range activation” and “synergetic chemomechanical oscillation”. In the case of long-range activation [16,17], differential diffusion contributes to the spatiotemporal oscillations. In synergetic chemomechanical [4,18,19,20,21,22,23,24,25] oscillations realized in spatial bistable systems [18,19,20,23,24] or enzymatic reactions realized in reactive porous membranes [21,22], gel size change exerts feedback on the reaction so that an oscillation can be induced in non-oscillatory bulk reactions.
Although pH-responsive gels are easy to obtain, and there are many pH oscillators available [26], the realization of pH-mediated autonomous chemomechanical systems is not so expected. Yoshida et al. had reported a self-oscillating gel by coupling p(N-Isopropylacrylamide-co-acrylic acid) with a pH oscillatory reaction [8]. pH responsive hydrogel polymethacrylic acid was also realized to oscillate in a bromate-sulfite-ferrocyanide pH oscillation reaction [9]. Additionally, the gels employed in the reported systems are often small. One of the constraints is the dependence of the rate of volume change of the gel on the mass transport of the solvent [2], whereby the rate of volume change is inversely proportional to the square of the linear dimension of the hydrogel [27]. Accordingly, large pH-responsive gels do not have enough time to change size on the time scale of the pH oscillation period, which is often tens of minutes [10,11]. Incorporating pores into hydrogels [10,11,28,29,30] to facilitate the migration of water through enlarging the surface area is often employed to improve the hydrogel response speed. Additionally, the oscillatory reaction constitutes an environment with high ion strength. Obtaining large amplitude responsive size changes is a demanding requirement for pH sensitive hydrogel, which is a polyelectrolyte in nature. Moreover, reversible volume changes may induce high strains on the gel network, and such effects are more significant if the gel is large. Chemical cross-linking in pH-responsive gels inevitably introduces spatial heterogeneity in the cross-link density, leading to local anisotropic expansion and contraction. Such an expansion can instigate network fracture, thus lowering the mechanical strength of the gel [2,31,32]. In the present study, a porous pH-sensitive hydrogel with a special porogen alginate is synthesized to accelerate the diffusion inside the gel. The bromate-sulfite-manganese (II) (BSM) pH oscillatory reaction [33], which provides large amplitude periodical pH changes, was employed as the driving force. In this reaction, Mn2+ (or MnO4) acts as the negative feedback species that consumes H+, and the reaction of BrO3−-SO32 in an acidic environment produces H+ in an autocatalytic manner. The hydrogel interacts with the oscillation not only through H+, but also through the negative feedback species Mn2+. The special interaction between Mn2+ and the hydrogel strengthens the gel by forming complex and physical crosslinks and weakens the salt effect. Although the chemomechanical instability of this system still originates from the original oscillatory kinetics, the oscillation ingredient also contributes to strengthen the gel and bring about the large amplitude chemomechanical oscillation.

2. Materials and Methods

2.1. Hydrogel Synthesis

Hydrogel poly(acrylic acid-co-acrylamide) (p(AA-co-AM)) was synthesized by free-radical solution polymerization. The monomer acrylamide (Amethyst Chemicals, Beijing, China), acrylic acid (Alfa Aesar, Ward Hill, MA, USA), (monomer ratio = 1:1 (mass ratio), concentration of monomers in pregel solution: 10 wt %), cross-linker N,N′-methylenebis(acrylamide) (0.4 wt % of total monomer, Sinopharm Chemical Reagent, Beijing, China), initiator ammonium persulfate (2 wt % of total monomer, Sinopharm Chemical Reagent, Beijing, China), and porogen sodium alginate (SA; viscosity ≥ 0.02 Pa∙s, 29.3 wt % of total monomer, Sinopharm Chemical Reagent, Beijing, China) were dissolved in 0.6 M NaCl solution in a glass beaker. The salt solution was used to instigate a coil-to-globule transition of the SA chains so that a porous structure is obtained after the SA chains are washed out. Then, the glass beaker was sealed and transferred to a thermostatic water bath at 65 °C for 4 h. Finally, the mixture was cooled to room temperature. Hydrogels were immersed in deionized water for seven days, with regular water replacement to eliminate impurities. Before testing, all the fully-swollen samples were immersed in buffer solutions of pH 6.8 for 12 h and stained with methylene blue. Samples prepared in this way are referred to as the porous hydrogels. For comparison, non-porous hydrogels were also synthesized by similar processes without adding SA and NaCl. The morphology of the porous and non-porous hydrogels was observed by scanning electron microscopy (SEM; SU8010, Hitachi, Tokyo, Japan). The element component of gel soaked in oscillatory reaction is tested by energy dispersive spectrometer (EDS; SU8010, Hitachi, Tokyo, Japan). The gel component was characterized by Fourier transform infrared (FT-IR; Bruker Tensor 27, Karlsruhe, Germany) spectroscopy. The bulk gel synthesized in the beaker was cut into cubes with a razor blade.

2.2. pH Oscillation Setting

Oscillation experiments were performed in a water-jacketed glass continuous stirred tank reactor (CSTR, 24 mL). The reaction temperature was maintained to 45 ± 0.l °C by circulating water between the water jacket and the thermostatic water bath (PolyScience 9006, Niles, IL, USA). Four stock solutions of BSM reactant, containing SO32−, H+, Mn2+, and BrO3, respectively, were separately introduced into the reactor by a peristaltic pump (Waston Marlow-205S, Cornwall, UK). The volume of the mixture in the reactor was maintained by continuous aspiration of the solution from the top of the reactor. The solution mixture in the CSTR was stirred magnetically by a Teflon-coated magnetic stirrer bar. Sufficient time was maintained at each flow rate to obtain stable oscillatory states. The pH was monitored using a combined glass electrode connected to a data acquisition system (CONSORT D230, Turnhout, Belgium).

2.3. Hydrogel pH-Responsiveness Test

The gel responsiveness to the pH oscillator was tested using both direct and indirect methods. In the direct method, the hydrogel was coupled with the BSM oscillation by suspending the sample on a small plastic plate that was immersed in the CSTR solution. The volume changes of the gel were monitored by a video camera, and the corresponding volume (length) change was determined from the images. In the indirect method, the sample was alternately immersed in buffer solutions of pH 1.7 and 6.8 for 8 min, with its weight being recorded every 2 min. The ionic strength of the buffer solutions was adjusted to that of the BSM system [33], which is calculated with the input concentration.

3. Results and Discussion

3.1. Coupled Chemomechanical Oscillations

As observed in the snapshot (Figure 1a) and the video (Supplementary materials), the porous p(AA-co-AM) hydrogel showed sustained reversible volume (length) changes with the pH changes of the BSM oscillation [33]. The BSM system provides a large amplitude pH change, whose period is long and its duration in the high pH parts are comparative to that in the low pH parts. The Mn2+ plays a clear and independent role of negative feedback in the mechanism. Here, the oscillation parameter setting was chosen under the principle that the oscillation is stable and the period is long. The measured length change was maintained at more than 10% for over 1 h, which is significant for the macroscopic chemomechanical energy transition. The contribution of porous acceleration is tested by control experiments of nonporous gels. Naked eye or charge-coupled devices (CCD) cannot recognize the volume change of the non-porous gel. Thus, the difference between the responses of the porous and non-porous gels was compared with an indirect test, whereby the weight of the gels was measured after the gels were soaked in buffer solutions. The periodic response of the porous hydrogel was considerably more obvious than that of the non-porous hydrogel (Figure 1b). In another quantitative test, the porous gel deswelled to 40% of its original weight in 8 min, whereas the non-porous gel took 32 min to achieve the same result. Therefore, the non-porous gel could not respond within the time scale of the oscillatory period to produce visible volume changes. The accelerated response rate in the porous gel was attributed to the porous structure formed upon contact of SA with NaCl solution. SA is a natural polyelectrolyte that undergoes a coil-to-globule transition when in contact with the NaCl solution to generate a porous gel structure (Figure 1c). The SEM images of the porous and non-porous gels in Figure 1d, e clearly demonstrate the pore-forming effect. Owing to the chain transfer reactions, part of the SA could graft to the copolymer [34,35]. FT-IR spectroscopy results (Figure 1f) of the porous gels have some very weak absorptions at 880 cm−1 and 1050 cm−1, which suggests the presence of residual SA. SA is proposed to graft to the hydrogel through the macroradicals obtained by abstracting hydrogen from the hydroxyl groups in SA by the sulfate anion radical (from the initiator [36,37]).
However, reversible, large amplitude chemomechanical oscillation is not realized by simply increasing the gel volume change rate. Control experiments with porous gels obtained by other methods, such as a gas blowing (or foaming) techniques or using silica particles as a porogen, showed no evidence of autonomous gel volume oscillation. Moreover, some samples do not swell in the indirect test at all. This is because the fragile nature of the gel network is accentuated since there are more walls inside the gel. Note that the ionic strength (~0.5 M) of the current chemomechanical system is considerably higher than that for other pH-responsive gels [38,39,40]. The good salt resistance displayed by the porous p(AA-co-AM) hydrogel here may be related to the special interaction of Mn2+ with the gel, which will be further discussed in later sections.

3.2. Mn2+ Coupled to the Hydrogel

Instead of being brittle and weak, the porous p(AA-co-AM) hydrogel appears to be even stronger when soaked in the oscillatory solution. It is speculated to be an effect of the Mn2+ involvement, which will interact with carboxyl groups, as well as the porogen SA. To identify the presence of Mn2+ in the hydrogel, energy dispersive spectrometry tests were performed for porous gels soaked in MnSO4 solution and BSM solution after they were washed with distilled water. EDS results in Figure 2 confirm the presence of Mn2+ in the gel, and the local Mn2+ content in the gel soaked in BSM solution is much higher than that in the MnSO4 solution. This might be because BSM is an open system with continuous input.
The responsive performance of the porous (with SA) and non-porous (SA-free) gels were compared by alternatively immersing the gel in the buffer solutions of pH 1.7 and 6.8 with the addition of MnSO4 and NaCl, respectively. The ionic strength of the buffer solution was adjusted to be the same. The concentration of MnSO4 and NaCl in the solution are the same as that in the BSM system. Following the duration of low and high pH part in the BSM oscillation, the gels are first immersed in the pH 1.7 solution for 10 min, and then cleaned to put into the pH 6.8 solution for 25 min. This cycle is then repeated. It is obvious from the snapshots in Figure 3a that the nonporous gel shows only limited volume change due to the limited response speed, while the porous gels exhibit a definite reversible size change. Comparing the last snapshot in each group, the color of the gel fades rapidly, except for the porous gel in the Mn2+ solution. This indicates that there are some additional networks formed inside the gel with the aid of Mn2+ ions in solution, which helps to hold the dye molecules in the gel.
Previous investigations had reported the formation of complex of Mn2+ with polyelectrolytes, like ployacrylic acid [41,42]. Additionally, SA is known to possess a strong gelation capability in the presence of divalent cations [43,44]. The regular intramolecular configuration of SA facilitates the formation of an “egg carton” shaped physical cross-linked conformation with divalent cations (Figure 3b). Therefore, Mn2+ will lead to additional gel-oscillation interactions, especially with the presence of SA in the gel. These may finally form additional networks inside the hydrogel. Further experiments with SA and divalent cations show the binding of Mn2+ on SA in solution is quite limited compared with Ca2+ or Cu2+. In the hydrogel, this ability may be enhanced since SA molecules have low mobility when grafted on the polymer chain. This moderate strength of binding provides an appropriate physical crosslink network that strengthens the hydrogel network without destroying its responsivity. The color difference between the non-porous gel in Mn2+ and Na+ solutions suggests that the complex formation between Mn2+ and the gel can also form some networks inside the gel to prevent the dye molecules from leaving. However, probably due to the irregular distribution of AA on the gel chain, the strength of this latter network appears to be weaker compared to that with SA. From SEM images of the porous and nonporous gels soaked in the MnSO4 and NaCl solution (Figure 3c–e) we can see that there are still three-dimensional network structures in the porous gel sample after it was soaked in MnSO4 solution. For the nonporous gel or the porous gels soaked in NaCl solution, network structures are not clear. This suggests that there are complexation- and crosslinking-induced network structures in the presence of Mn2+. These interactions between Mn2+ and the gel suggests that Mn2+ in the solution influences the gel mainly by forming complex and physical cross-links, thus reducing electrostatic screening effects. Considering that Mn2+ contributes significantly to the overall ionic strength in the solution, this should be the main factor that improves the salt resistance of the gels.
FT-IR spectra of the gel samples soaked in MnSO4 solution provides additional evidence that Mn2+ should be involved in the gel structure. Figure 4 shows that there are acrylic acid-associated structures in the gel (peaks around 1670 cm−1). It should be the Mn2+ formed complex and physical crosslinks that contribute to the formation of the associated structures, since the peak is strengthened significantly with the increase of MnSO4 concentration. In the MnSO4 solution with the same concentration, the porous gel has a stronger associated structure absorption than the non-porous gel, which shows from another angle that SA in the porous gel enhanced the binding of Mn2+.

3.3. Mn2+ Binding Alternated Oscillation

The special coupling mechanism also changes the bulk oscillation character. Figure 5a shows two independent oscillation sequences when porous gels of the same size were put into oscillation with identical parameter settings. The oscillation period is randomly reduced or increased by the addition of the gel. This is also associated with the binding of Mn2+. When pH-dependent processes are driven by pH oscillations, they usually coupled through H+ and have a buffer-like effect to the oscillation, which will reduce the amplitude as well as the period of the oscillation [45,46]. This is because H+ consumption in pH-dependent processes speeds up the H+ production in the reversible ionization process in the oscillation. For the BSM system, its amplitude is robust [47], but its period will be reduced. However, coupling through Mn2+ has quite the opposite effect. Mn2+ consumes H+ in BSM, and its binding in the gel will retard the negative feedback of BSM and prolong its period. The experimental results of the BSM system (Figure 5b) also verifies that the BSM period increases with the decreasing of Mn2+ concentration. Note that the system described in Figure 5b has no hydrogel in it. Therefore, the irregular period change in Figure 5a should be the result of competition binding of the two coupling media, i.e., H+ and Mn2+. Since the monomer distribution inside the gel is not homogenous, the oscillation sequence cannot be quantitatively reproducible.
Increasing the gel size or enhancing the negatively-charged gel ingredient content (introducing some 2-acrylamide-2-methylpropanesulfonic acid (AMPS) in the copolymer) may stop the oscillation at high pHs, with a large amount of white precipitate found around the gel. The precipitation should be the insoluble Mn(OH)2 formed in the high pH environment when local Mn2+ concentration is high. When BSM is conducted at higher MnSO4 concentration (above 13 mM), a small amount of white precipitate could be found at the bottom of reactor, indicating the possibility of forming the insoluble Mn(OH)2 in BSM reactions. In the current system, most precipitates are formed around the gel, suggesting Mn2+ ions are gathered near the gel so that its local concentration is high enough to form a precipitate. The formation of such a large amount of precipitate makes the Mn2+ in the solution less than necessary for the oscillation, i.e., it cuts off the feedback loop. Therefore, the oscillation stops at a high pH when the precipitate condition is satisfied.

4. Conclusions

Large amplitude hydrogel volume oscillation was observed in a chemomechanical system constituted of a BSM pH oscillatory reaction and porous p(AA-co-AM) gel treated with alginate as the pore-forming agent. The negative feedback species Mn2+ is involved in the chemomechanical coupling by forming complex and physical crosslinks within the gel. These constitute additional three-dimensional networks in the hydrogel, which considerably reinforced the gel network structure. Despite that the nonlinear stability still originating from the oscillatory reaction, Mn2+ binding also helps to maintain the chemomechanical oscillation by strengthening the gel. As oscillatory reactions are currently used as a non-equilibrium driving force to produce autonomous behaviors in many areas, the negative feedback species involved coupling mechanism provide new sight for the related chemomechanical transitions.

Supplementary Materials

Supplementary materials are available online at www.mdpi.com/2073-4360/9/11/554/s1.

Acknowledgments

We are grateful to Stephen Scott for helpful discussions and valuable suggestions, we also thank John Pojman for helpful discussions. This work was supported by the Beijing Natural Science Foundation Program (2152009), and Beijing Youth talent cultivation plans (CIT&TCD201404161).

Author Contributions

Xin Yang, Yanhui Ding, and Yi Zhou performed the experiments and prepared figures; Lin Ji and Jianquan Wang designed the whole work and arranged the investigation; Lin Ji wrote the main manuscript; and Xin Liang contributed to the revision and additional experimental processes.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Siegel, R.A. Autonomous Rhythmic Drug Delivery Systems Based on Chemical and Biochemomechanical Oscillators; Borckmans, P., Kepper, P., Khokhlov, A.R., Métens, S., Eds.; Springer: Dordrecht, The Netherlands, 2009; pp. 175–201. [Google Scholar]
  2. Howse, J.R.; Topham, P.; Crook, C.J.; Gleeson, A.J.; Bras, W.; Jones, R.A.; Ryan, A.J. Reciprocating power generation in a chemically driven synthetic muscle. Nano Lett. 2006, 6, 73–77. [Google Scholar] [CrossRef] [PubMed]
  3. Dong, L.; Jiang, H. Autonomous microfluidics with stimuli-responsive hydrogels. Soft Matter 2007, 3, 1223–1230. [Google Scholar] [CrossRef]
  4. He, X.; Aizenberg, M.; Kuksenok, O.; Zarzar, L.D.; Shastri, A.; Balazs, A.C.; Aizenberg, J. Synthetic homeostatic materials with chemo-mechano-chemical self-regulation. Nature 2012, 487, 214–218. [Google Scholar] [CrossRef] [PubMed]
  5. Hara, Y.; Yoshida, R. Control of oscillating behavior for the self-oscillating polymer with pH-control site. Langmuir 2005, 21, 9773–9776. [Google Scholar] [CrossRef] [PubMed]
  6. Yoshida, R. Design of self-oscillating gels and application to biomimetic actuators. Sensors 2010, 10, 1810–1822. [Google Scholar] [CrossRef] [PubMed]
  7. Yoshida, R.; Ueki, T. Evolution of self-oscillating polymer gels as autonomous polymer systems. NPG Asia Mater. 2014, 6, e107. [Google Scholar] [CrossRef]
  8. Yoshida, R.; Ichijo, H.; Hakuta, T.; Yamaguchi, T. Self-oscillating swelling and deswelling of polymer gels. Macromol. Rapid Commun. 1995, 16, 305–310. [Google Scholar] [CrossRef]
  9. Ryan, A.J.; Crook, C.J.; Howse, J.R.; Topham, P.; Jones, R.A.L.; Geoghegan, M.; Parnell, A.J.; Ruiz-Perez, L.; Martin, S.J.; Cadby, A.; et al. Responsive brushes and gels as components of soft nanotechnology. Faraday Discuss. 2005, 128, 55–74. [Google Scholar] [CrossRef] [PubMed]
  10. Bilici, C.; Karayel, S.; Demir, T.T.; Okay, O. Self-oscillating ph-responsive cryogels as possible candidates of soft materials for generating mechanical energy. J. Appl. Polym. Sci. 2010, 118, 2981–2988. [Google Scholar] [CrossRef]
  11. Wang, L.P.; Ren, J.; Zhang, X.Y.; Yang, X.C.; Yang, W. Synthesis and Characterization of pH-sensitive and Self-oscillating IPN Hydrogel in a pH Oscillator. Polym. Korea 2015, 39, 359–364. [Google Scholar] [CrossRef]
  12. Sasaki, S.; Koga, S.; Yoshida, R.; Yamaguchi, T. Mechanical oscillation coupled with the belousov-zhabotinsky reaction in gel. Langmuir 2003, 19, 5595–5600. [Google Scholar] [CrossRef]
  13. Sakai, T.; Yoshida, R. Self-oscillating nanogel particles. Langmuir 2004, 20, 1036–1038. [Google Scholar] [CrossRef] [PubMed]
  14. Zhang, Y.; Zhou, N.; Li, N.; Sun, M.; Kim, D.; Fraden, S.; Epstein, I.R.; Xu, B. Giant volume change of active gels under continuous flow. J. Am. Chem. Soc. 2014, 136, 7341–7347. [Google Scholar] [CrossRef] [PubMed]
  15. Zhang, Y.; Zhou, N.; Akella, S.; Kuang, Y.; Kim, D.; Schwartz, A.; Bezpalko, M.; Foxman, B.M.; Fraden, S.; Epstein, I.R.; et al. Active cross-linkers that lead to active gels. Angew. Chem. Int. Ed. Engl. 2013, 52, 11494–11498. [Google Scholar] [CrossRef] [PubMed]
  16. Boissonade, J.; de Kepper, P.; Gauffre, F.; Szalai, I. Spatial bistability: A source of complex dynamics. From spatiotemporal reaction-diffusion patterns to chemomechanical structures. Chaos 2006, 16, 037110. [Google Scholar] [PubMed]
  17. Fuentes, M.; Kuperman, M.N.; Boissonade, J.; Dulos, E.; Gauffre, F.; de Kepper, P. Dynamical effects induced by long range activation in a nonequilibrium reaction-diffusion system. Phys. Rev. E Stat. Nonlinear Soft Matter Phys. 2002, 66, 056205. [Google Scholar] [CrossRef] [PubMed]
  18. Boissonade, J. Simple chemomechanical process for self-generation of rhythms and forms. Phys. Rev. Lett. 2003, 90, 188302. [Google Scholar] [CrossRef] [PubMed]
  19. Borckmans, P. Chemomechanical Instabilities in Responsive Materials; Springer: Dordrecht, The Netherlands, 2009. [Google Scholar]
  20. Boissonade, J. Self-oscillations in chemoresponsive gels: A theoretical approach. Chaos 2005, 15, 23703. [Google Scholar] [CrossRef] [PubMed]
  21. Dhanarajan, A.P.; Urban, J.; Siegel, A.R. Nonlinear Dynamics in Polymeric Systems; Tran-Cong-Miyata, J.P.A.Q., Ed.; American Chemical Society: Washington, DC, USA, 2003; pp. i–v. [Google Scholar]
  22. Dhanarajan, A.P.; Misra, G.P.; Siegel, R.A. Autonomous Chemomechanical Oscillations in a Hydrogel/Enzyme System Driven by Glucose. J. Phys. Chem. A 2002, 106, 8835–8838. [Google Scholar] [CrossRef]
  23. Horváth, J.; Szalai, I.; Boissonade, J.; Kepper, P.D. Oscillatory dynamics induced in a responsive gel by a non-oscillatory chemical reaction: Experimental evidence. Soft Matter 2011, 7, 8462. [Google Scholar] [CrossRef]
  24. Horvath, J. Sustained large-amplitude chemomechanical oscillations induced by the Landolt clock reaction. J. Phys. Chem. B 2014, 118, 8891–8900. [Google Scholar] [CrossRef] [PubMed]
  25. Horvath, J. Peristaltic waves in a responsive gel sustained by a halogen-free non-oscillatory chemical reaction. Polymer 2015, 79, 243–254. [Google Scholar] [CrossRef]
  26. Orbán, M.; Kurin-Csörgei, K.; Epstein, I.R. pH-Regulated Chemical Oscillators. Acc. Chem. Res. 2015, 48, 593–601. [Google Scholar] [CrossRef] [PubMed]
  27. Tanaka, T.; Fillmore, D.J. Kinetics of swelling of gels. J. Chem. Phys. 1979, 70, 1214–1218. [Google Scholar] [CrossRef]
  28. Zhang, X.Z.; Yang, Y.Y.; Chung, T.S.; Ma, K.X. Preparation and characterization of fast response macroporous poly(N-isopropylacrylamide) hydrogels. Langmuir 2001, 17, 6094–6099. [Google Scholar] [CrossRef]
  29. Chaterji, S.; Kwon, I.K.; Park, K. Smart polymeric gels: Redefining the limits of biomedical devices. Prog. Polym. Sci. 2007, 32, 1083–1122. [Google Scholar] [CrossRef] [PubMed]
  30. Zhang, Y.; Zhou, R.; Shi, J.F.; Zhou, N.; Epstein, I.R.; Xu, B. Post-self-assembly cross-linking to integrate molecular nanofibers with copolymers in oscillatory hydrogels. J. Phys. Chem. B 2013, 117, 6566–6573. [Google Scholar] [CrossRef] [PubMed]
  31. Shibayama, M.; Tanaka, T. Volume phase-transition and related phenomena of polymer gels. Adv. Polym. Sci. 1993, 109, 1–62. [Google Scholar]
  32. Shinohara, S.-I.; Seki, T.; Sakai, T.; Yoshida, R.; Takeoka, Y. Chemical and optical control of peristaltic actuator based on self-oscillating porous gel. Chem. Commun. 2008, 4735–4737. [Google Scholar] [CrossRef] [PubMed]
  33. Okazaki, N.; Rabai, G.; Hanazaki, I. Discovery of novel bromate-sulfite pH oscillators with Mn2+ or MnO4− as a negative-feedback species. J. Phys. Chem. A 1999, 103, 10915–10920. [Google Scholar] [CrossRef]
  34. Yashin, V.V.; Balazs, A.C. Modeling polymer gels exhibiting self-oscillations due to the Belousov-Zhabotinsky reaction. Macromolecules 2006, 39, 2024–2026. [Google Scholar] [CrossRef]
  35. Kim, J.H.; Lee, S.B.; Kim, S.J.; Lee, Y.M. Rapid temperature/pH response of porous alginate-g-poly(N-isopropylacrylamide) hydrogels. Polymer 2002, 43, 7549–7558. [Google Scholar] [CrossRef]
  36. Pouijavadi, A.; Ghasemzadeh, H.; Hosseinzadeh, H. Preparation and swelling behaviour of a novel anti-salt superabsorbent hydrogel based on κ-carrageenan and sodium alginate grafted with polyacrylamide. E-Polymers 2004, 4. [Google Scholar] [CrossRef]
  37. Pourjavadi, A.; Samadi, M.; Ghasemzadeh, H. Temperature sensitive superabsorbent hydrogels from poly(N-t-butyl acrylamide-co-acrylamide) grafted on sodium alginate. Macromol. Symp. 2008, 274, 177–183. [Google Scholar] [CrossRef]
  38. Lee, W.-F.; Hsu, C.-H. Thermoreversible hydrogels: 3. Synthesis and swelling behavior of the (N-isopropylacrylamide-co-trimethylacrylamidopropyl ammonium iodide) copolymeric hydrogels. Polymer 1998, 39, 5393–5403. [Google Scholar] [CrossRef]
  39. Lee, W.F.; Yuan, W.Y. Thermoreversible hydrogels XI: Effect of salt on the swelling properties of the (N-isopropylacrylamide-co-sodium 2-acrylamido-2-methylpropyl sulfonate) copolymeric hydrogels. J. Appl. Polym. Sci. 2001, 79, 1675–1684. [Google Scholar] [CrossRef]
  40. Marandi, G.B.; Sharifnia, N.; Hosseinzadeh, H. Synthesis of an alginate-poly(sodium acrylate-co-acrylamide) superabsorbent hydrogel with low salt sensitivity and high pH sensitivity. J. Appl. Polym. Sci. 2006, 101, 2927–2937. [Google Scholar] [CrossRef]
  41. Harry, P.G.; Lionel, B.L.; Ernest, M.L. Metal-Polyelectrolyte Complexes. IV. Complexes of Polyacrylic Acid with Magnesium, Calcium, Manganese, Cobalt and Zinc. J. Phys. Chem. 1955, 55, 990–991. [Google Scholar]
  42. Koda, S.; Nomura, H.; Nagasawa, M. Raman-Spectroscopic Studies on the Interaction between Divalent Counterion and Polyion. Biophys. Chem. 1983, 18, 361–367. [Google Scholar] [CrossRef]
  43. De Gennes, P.G. Scaling Concepts in Polymer Physics; Cornell University Press: Ithaca, NY, USA, 1979. [Google Scholar]
  44. Tokarev, I.; Gopishetty, V.; Zhou, J.; Pita, M.; Motornov, M.; Katz, E.; Minko, S. Stimuli-Responsive Hydrogel Membranes Coupled with Biocatalytic Processes. ACS Appl. Mater. Interfaces 2009, 1, 532–536. [Google Scholar] [CrossRef] [PubMed]
  45. Misra, G.P.; Siegel, R.A. Ionizable drugs and pH oscillators: Buffering effects. J. Pharm. Sci. 2002, 91, 2003–2015. [Google Scholar] [CrossRef] [PubMed]
  46. Kurin-Csörgei, K.; Epstein, I.R.; Orbán, M. Systematic design of chemical oscillators using complexation and precipitation equilibria. Nature 2005, 433, 139–142. [Google Scholar] [CrossRef] [PubMed]
  47. Horvath, K.K.-C.V.; Epstein, I.R.; Orban, M. Oscillations in the Concentration of Fluoride Ions Induced by a pH Oscillator. J. Phys. Chem. A 2008, 112, 4271–4276. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) Variations in the length of the porous gel and pH of the BSM system as a function of time. Snapshots were taken at the time points noted in the time sequence graph. The BSM oscillation parameters are [NaBrO3]0 = 90 mM, [Na2SO3]0 = 118 mM, [MnSO4]0 = 9 mM, [HClO4]0 = 16.5 mM, T = 45 °C, k0 = 2.945 × 10−3·s−1. (b) Comparison of the responsive swelling of the porous and non-porous gels determined in the offline study; the ionic strength of the buffer solution was adjusted to 0.5 M. The unshaded areas of the graph represent instances when the pH is 1.7, whereas the shaded areas of the graph represent instances when the pH is 6.8. (c) Schematic of the formation of the porous (gel) structure upon addition of SA to NaCl solution. SEM images of the (d) porous and (e) non-porous gels. FT-IR spectrum of porous, nonporous hydrogel (f).
Figure 1. (a) Variations in the length of the porous gel and pH of the BSM system as a function of time. Snapshots were taken at the time points noted in the time sequence graph. The BSM oscillation parameters are [NaBrO3]0 = 90 mM, [Na2SO3]0 = 118 mM, [MnSO4]0 = 9 mM, [HClO4]0 = 16.5 mM, T = 45 °C, k0 = 2.945 × 10−3·s−1. (b) Comparison of the responsive swelling of the porous and non-porous gels determined in the offline study; the ionic strength of the buffer solution was adjusted to 0.5 M. The unshaded areas of the graph represent instances when the pH is 1.7, whereas the shaded areas of the graph represent instances when the pH is 6.8. (c) Schematic of the formation of the porous (gel) structure upon addition of SA to NaCl solution. SEM images of the (d) porous and (e) non-porous gels. FT-IR spectrum of porous, nonporous hydrogel (f).
Polymers 09 00554 g001
Figure 2. Energy dispersive spectrometry of the porous p(AA-co-AM) hydrogel taken from the 9 mM MnSO4 and BSM (with the same parameter as Figure 1a) solution.
Figure 2. Energy dispersive spectrometry of the porous p(AA-co-AM) hydrogel taken from the 9 mM MnSO4 and BSM (with the same parameter as Figure 1a) solution.
Polymers 09 00554 g002
Figure 3. (a) Comparison of the swelling/deswelling behaviors of the porous and non-porous gels immersed in 9 mM MnSO4 and 326 mM NaCl of buffer solution with different pH. T = room temperature, initial pH: 6.8 (b) Schematic of the physical cross-linking between the porous gel and divalent metal ions. An additional network is formed upon interaction of Mn2+ with SA through an “egg carton” conformation and with carboxyl groups on AA through a complex conformation. SEM images of the (c) porous gel in MnSO4 solution, (d) non-porous gels soaked in MnSO4 solution, and (e) porous gel in NaCl solution.
Figure 3. (a) Comparison of the swelling/deswelling behaviors of the porous and non-porous gels immersed in 9 mM MnSO4 and 326 mM NaCl of buffer solution with different pH. T = room temperature, initial pH: 6.8 (b) Schematic of the physical cross-linking between the porous gel and divalent metal ions. An additional network is formed upon interaction of Mn2+ with SA through an “egg carton” conformation and with carboxyl groups on AA through a complex conformation. SEM images of the (c) porous gel in MnSO4 solution, (d) non-porous gels soaked in MnSO4 solution, and (e) porous gel in NaCl solution.
Polymers 09 00554 g003aPolymers 09 00554 g003b
Figure 4. FT-IR spectrum of porous (nonporous) hydrogels soaked in the MnSO4 solution with different concentrations as noted in the legend.
Figure 4. FT-IR spectrum of porous (nonporous) hydrogels soaked in the MnSO4 solution with different concentrations as noted in the legend.
Polymers 09 00554 g004
Figure 5. (a) pH oscillation sequences of the porous hydrogel-BSM coupled system. The two sequences are realized with identical parameter settings, i.e., both the BSM setting and gel component size are the same. The arrows indicate the time at which a new porous gel sample was added. The sizes of the three gels added in sequence are 10 × 7 × 3, 10 × 5 × 3, and 10 × 3 × 3 mm3, respectively. (b) Dependence of the BSM oscillation period on the concentration of MnSO4, and other experimental parameters, are the same as Figure 1a.
Figure 5. (a) pH oscillation sequences of the porous hydrogel-BSM coupled system. The two sequences are realized with identical parameter settings, i.e., both the BSM setting and gel component size are the same. The arrows indicate the time at which a new porous gel sample was added. The sizes of the three gels added in sequence are 10 × 7 × 3, 10 × 5 × 3, and 10 × 3 × 3 mm3, respectively. (b) Dependence of the BSM oscillation period on the concentration of MnSO4, and other experimental parameters, are the same as Figure 1a.
Polymers 09 00554 g005

Share and Cite

MDPI and ACS Style

Yang, X.; Zhou, Y.; Ji, L.; Ding, Y.; Wang, J.; Liang, X. Experimental Evidence of Large Amplitude pH Mediated Autonomous Chemomechanical Oscillation. Polymers 2017, 9, 554. https://doi.org/10.3390/polym9110554

AMA Style

Yang X, Zhou Y, Ji L, Ding Y, Wang J, Liang X. Experimental Evidence of Large Amplitude pH Mediated Autonomous Chemomechanical Oscillation. Polymers. 2017; 9(11):554. https://doi.org/10.3390/polym9110554

Chicago/Turabian Style

Yang, Xin, Yi Zhou, Lin Ji, Yanhui Ding, Jianquan Wang, and Xin Liang. 2017. "Experimental Evidence of Large Amplitude pH Mediated Autonomous Chemomechanical Oscillation" Polymers 9, no. 11: 554. https://doi.org/10.3390/polym9110554

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop