Temperature and pH Responsive Hydrogels Using Methacrylated

Dec 20, 2017 - Taking a Closer Look at Alternatives to Colloidal Cesium Lead Halide Perovskite Nanocrystals. Colloidal semiconductor nanocrystals (NCs...
0 downloads 8 Views 2MB Size
Subscriber access provided by READING UNIV

Article

Temperature and pH responsive hydrogels using methacrylated lignosulfonate crosslinker: Synthesis, characterization and properties Can Jin, Wenjia Song, Tuan Liu, Junna Xin, William C. Hiscox, Jinwen Zhang, Guifeng Liu, and Zhenwu Kong ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.7b03158 • Publication Date (Web): 20 Dec 2017 Downloaded from http://pubs.acs.org on December 24, 2017

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.

ACS Sustainable Chemistry & Engineering 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 30 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

ACS Sustainable Chemistry & Engineering

Temperature and pH responsive hydrogels using methacrylated lignosulfonate crosslinker: Synthesis, characterization and properties

Can Jin,1,2 Wenjia Song,2 Tuan Liu,2 Junna Xin,2,* William C. Hiscox,3 Jinwen Zhang,2,* Guifeng Liu,1 Zhenwu Kong1,*

1. Institute of Chemical Industry of Forest Products, Chinese Academy of Forestry, National Engineering Laboratory for Biomass Chemical Utilization, Nanjing, 210042, China

2. School of Mechanical and Materials Engineering, Composite Materials and Engineering Center, Washington State University, P.O. Box 641806, Pullman, Washington 99164, United States

3. Nuclear Magnetic Resonance Center, Washington State University, P.O. Box 4630, Pullman, Washington 99164, United States

*Corresponding

authors.

E-mail:

[email protected]

(J.

Xin),

[email protected] (Z. Kong)

1

ACS Paragon Plus Environment

[email protected]

(J.

Zhang),

ACS Sustainable Chemistry & Engineering 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 30

Abstract: In this work, biobased hydrogels with temperature and pH responsive properties were prepared by copolymerizing N-isopropylacrylamide (NIPAM), itaconic acid (IA) and methacrylated lignosulfonate (MLS), where the multifunctional MLS served as a novel macro-crosslinker. The network structures of the lignosulfonate-NIPAM-IA hydrogels (LNIH) were characterized and confirmed by elemental analysis, FT-IR and

13

C NMR. The equilibrium

swelling capacity of the LNIH hydrogel decreased from 31.6 to 19.1 g/g with MLS content increasing from 3.7 to 14.3%, suggesting a strong dependence of water absorption of the gel on MLS content. LNIH hydrogels showed temperature-sensitive behaviors with volume phase transition temperature (VPTT) around the body temperature, which was also influenced by MLS content. Moreover, all LNIH hydrogels exhibited pH sensitivity in the range of pH 3.0 to 9.1. Rheological study indicated that mechanical strength of the gel also increased with MLS content. The results from this study suggest that lignosulfonate derivative MLS is a potential feedstock serving both water-absorbing moiety and crosslinker for preparation of biobased smart hydrogels.

Keywords: Lignosulfonate, hydrogel, temperature-sensitive, pH-sensitive, crosslinker.

2

ACS Paragon Plus Environment

Page 3 of 30 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

ACS Sustainable Chemistry & Engineering

Introduction Smart hydrogels, which are sensitive to environmental stimuli including temperature,1 pH,2 photo3 and certain chemical triggers,4 have received considerable attentions in fields of biomedicine, tissue engineering, agriculture and material science.5-7 In particular, temperatureand pH-sensitive hydrogels are mostly studied because both parameters are important environmental factors in typical biological, physiological and chemical systems. On the other hand, multiple environmental stimuli may occur at the same time in many cases. Therefore, numerous smart hydrogels have been prepared by combining temperature-sensitive component (e.g., N-isopropylacrylamide, pluronics, and 2-oxazoline) and pH-sensitive component (e.g., alginate, acrylic acid and orthoester amide) to control their dual stimuli-responsive behaviors.8-10 Very recently, there has been a growing interest in the preparation of smart hydrogels from natural polymers including cellulose11, chitosan12, and starch13. Natural polymers endow smart hydrogels many distinct advantages such as biocompatibility and biodegradability. Besides, natural polymers may also offer excellent mechanical strength to the hydrogel networks.14 Lignosulfonate, accounting for 90% of commercial lignin, is an abundant (~1.8 million tons per year) and inexpensive byproduct generated from sulfite pulping industry.15 Lignosulfonate is a branched and aromatic polymer with hydrophilic, chemically reactive and bioactive features, which enables it to be utilized as dispersant,16 flocculant,17 ion-exchange resin18 and antioxidant19. In recent years, lignosulfonate has been chosen as a renewable candidate for fabricating “green” hydrogels. Sun et al. synthesized lignosulfonate-modified graphene hydrogel for Pb(II) adsorption through one-step method.20 Wang et al. designed lignosulfonate-grafted poly(acrylic 3

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

acid-co-poly(vinyl pyrrolidone)) hydrogel for drug delivery.21 Fu et al. reported the application in dye removal using lignosulfonate-grafted-acrylic acid hydrogels as a new adsorbent.22 Moreover, our group recently developed a new approach to prepare lignosulfonate amine-PEG hydrogel without any crosslinker.23 Interestingly, our study indicated that the utilization of lignosulfonate derivative as macro-crosslinker could eliminate use of a traditional crosslinker (e.g., N,N’-methylenebisacrylamide) which was made from petrochemical feedstock with high cost and serious environmental impacts.24 Despite lignosulfonate shows promising prospects in “green” hydrogels by taking its advantages of biocompatibility, biodegradability and abundance, lignosulfonate-based hydrogel with stimuli-responsive behavior has been rarely reported. In this work, a series of stimuli-responsive hydrogels which contained biocompatible N-isopropylacrylamide (NIPAM) and itaconic acid (IA) were chemically crosslinked by a lignosulfonate derivative, where NIPAM and IA were imported as temperature- and pH-sensitive component, respectively. Methacrylate groups were introduced into the structure of lignin, and the resulted methacrylated lignosulfonate (MLS) was used as a macro-crosslinker. Effects of MLS content in the reaction on the composition and properties of lignosulfonate-NIPAM-IA hydrogels (LNIH) were investigated. Temperature- and pH-sensitive swelling behaviors of MLS crosslinked hydrogels were systematically studied. Moreover, the mechanical properties of designed hydrogels were improved due to the incorporation of lignosulfonate backbone. The findings from this study may set up a framework for preparation of smart hydrogels from lignin feedstock.

Experimental Section 4

ACS Paragon Plus Environment

Page 4 of 30

Page 5 of 30 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

ACS Sustainable Chemistry & Engineering

Materials. Sodium lignosulfonate (SLS, 2.97 mmol -OH/g, Mw = 62200 g/mol from GPC analysis) was purchased from Borregaard LignoTech USA Inc. and used as received. N-isopropylacrylamide (NIPAM, 98%), itaconic acid (IA, 99%) and sodium bisulfite (SBS) were purchased from Acros-Organics. Methacrylic anhydride (MAA, 94%), potassium persulfate (KPS, 99%), triethylamine (TEA, 99%) and 2-chloro-4,4,5,5-tetramethyl-1,2,3-dioxaphospholane (TMDP, 95%) were purchased from Sigma-Aldrich. All reagents were used as received unless otherwise specified.

Synthesis of methacrylated lignosulfonate (MLS). SLS (3.5 g, 10.4 mmol -OH groups), MAA (8.5 g, 52 mmol) and TEA (0.1 g) were dissolved in 60 mL deionized water. The solution was heated to 70 °C and remained under stirring for 18 h. After cooled down to room temperature, the reaction mixture was poured into ethanol and washed three times with excess ethanol to obtain a brown precipitate. The precipitate was freeze dried to obtain methacrylated lignosulfonate (Mw = 71400 g/mol).

Synthesis of lignosulfonate-NIPAM-IA hydrogels (LNIH). To endow the hydrogels with temperature- and pH-responsive properties, the content of NIPAM and IA monomer as temperature- and pH-sensitive components were maintained at above 80 wt% and 4 wt% in copolymerization reaction, respectively. Therefore, NIPAM (1000 mg), IA (50 mg), a certain amount of MLS (up to 200 mg) and KPS (2 wt% on the basis of total weight of reactants) were weighed and dissolved in deionized water (5 mL/1 g reactant) in a glass vial under an argon 5

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

atmosphere. After stirred at 100 rpm for 10 min, SBS (50% equiv. of KPS) was added into the aqueous solution to start the polymerization. The reaction was carried out at 70 °C for 4 h and then the solution was placed and cooled down to room temperature to obtain a LNIH hydrogel. Subsequently, to remove non-crosslinked copolymer and/or residual monomers, the LNIH hydrogel was cut into small pieces and washed thoroughly with excess deionized water, which was changed twice every day for at least 5 days. Finally, the washed LNIH hydrogels were separated, cut into pieces and dried in oven to dry gel state with constant weight.

Characterization. Gel Content. The gel content (G%) is calculated according to equation 1:

(1)

G% =

where G is the gel content of LNIH, Wa the weight of dry gel (washed), and Wb the weight of unwashed hydrogel. Three replicates were performed to determine the average gel content of each sample.

Swelling Behaviors of Hydrogels. The dry gel (washed) with known weight was placed into a teabag and then immersed into deionized water for sufficient swelling at 25 °C. Next, the hydrogel sample was taken out from aqueous solution and weighted after the removal of free water on surface with filter paper. The equilibrium swelling capacity (SC) of hydrogel is calculated using equation 2: 6

ACS Paragon Plus Environment

Page 6 of 30

Page 7 of 30 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

ACS Sustainable Chemistry & Engineering

(2)

SC (g/g) =

where Ws and W0 represent the weights of the swollen hydrogel and dry gel (washed), respectively. Duplicates were performed to determine the average SC value of each sample.

The pH sensitivity of LNIH hydrogels (washed) was examined in buffer solution at 25 °C. The pH values of buffer solutions were adjusted by HCl and NaOH and ranged from 3.0 to 9.1, and the ionic strength of buffer solution was 50 mmol/L by adding appropriate amount of NaCl. After the hydrogel samples were sufficiently immersed in buffer solutions, and the equilibrium swelling capacities of hydrogels were obtained.

FT-IR Spectroscopy.

FT-IR spectra of dry gel samples (washed) were recorded on a Nicolet

iS50 spectrometer (Thermo-Fisher Corporation, USA) using KBr disc method, and collected ranging from 4000 to 500 cm-1 for 64 scans at a 4 cm-1 resolution.

Nuclear Magnetic Resonance (NMR).

1

H NMR,

13

C NMR and

31

P NMR spectra were

recorded on a Varian 400 MHz NMR spectrometer. Particularly, the hydroxyl values of SLS and MLS were analyzed by 31P NMR spectra using TMDP as a phosphitylating reagent according to our previous work.25 Solid-state cross-polarization magic angle spinning (CPMAS)

13

C NMR

spectra for dry gel (washed) and MLS were acquired at a frequency of 100.63 MHz on a Bruker Avance DRX-400 NMR spectrometer fitted with a Chemagnetics triple resonance 5 mm MAS 7

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

probe, with a spectral width of 5000 Hz, relaxation delay of 4 s, contact time of 0.5 s and spinning rate of 8000 Hz. The signal averaged over a total of 20,000 scans and then Fourier transformed after applying 50 Hz line broadening.

Elemental Analysis. Carbon, hydrogen, nitrogen and sulphur contents of dry LNIH samples (washed) were determined on a LECO CHN analyzer (LECO TruSpec, St. Joseph, MI).

Scanning Electron Microscopy (SEM) Measurement. The swollen hydrogel was frozen in liquid nitrogen and then freeze-dried. The freeze-dried hydrogel was surface-coated with Au and examined on a Quanta 200F instrument (FEI, US).

Differential Scanning Calorimetry (DSC) Analysis. The volume phase transition temperature (VPTT) of hydrogel sample (washed) was determined on a DSC 1 instrument (Mettler-Toledo, Switzerland). Sample was immersed into deionized water of 10 times its weight and allowed to reach an equilibrium swelling. Then the swollen sample was cut into pieces (6 - 10 mg), placed into aluminum pan and sealed. The sample was scanned from 12 to 80 °C at a heating rate 5 °C /min under a continuous N2 flow (40 mL/min).

Rheological Measurement. The hydrogel sample for rheology test was prepared in a cylinder-shaped mold with a diameter 25 mm. The resulting LNIH hydrogel was sliced into discs with a thickness of ~2 mm for tests. Rheological properties were measured on a Discovery HR-2 rheometer (TA Instruments, USA) with a parallel plate geometry (diameter 25 mm), and the 8

ACS Paragon Plus Environment

Page 8 of 30

Page 9 of 30 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

ACS Sustainable Chemistry & Engineering

sample was scanned from 0.1 to 160 rad/s at a 1% strain and 25 °C. Shear modulus was tested from stress-strain curve in the strain range of 0 - 10% at 25 °C.

Results and Discussion Preparation of Methacrylated Lignosulfonate (MLS) Sodium lignosulfonate (SLS) was modified through methacrylation reaction using TEA as a catalyst in aqueous media to give MLS (Scheme 1). In comparison with the unmodified SLS, 1H NMR spectrum of MLS clearly confirmed the methacrylation by the appearance of characteristic peaks that corresponded to protons of double bond =CH2 (δ 5.6 - 6.7 ppm) and methyl groups -CH3 (δ 2.1 ppm) from the methacrylate moiety (Figure 1), respectively. Quantitative

31

P NMR

spectrum were also performed to determine the reaction degree through the change of -OH values of SLS and MLS (Table S1). As shown in Figure 2, the content of methacrylate groups in MLS was determined to be 1.61 mmol/g, which was calculated by comparing the total hydroxy content of SLS before (2.97 mmol/g) and after (1.36 mmol/g) the methacrylation reaction.25

Scheme 1. Schematic synthesis route of methacrylated lignosulfonate (MLS).

9

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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. 1H NMR spectra of SLS (top) and MLS (bottom) in D2O.

Figure 2. 31P NMR spectra of SLS (top) and MLS (bottom).

Preparation and Characterization of Hydrogels 10

ACS Paragon Plus Environment

Page 10 of 30

Page 11 of 30 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

ACS Sustainable Chemistry & Engineering

LNIH hydrogels were synthesized through radical polymerization of NIPAM and IA in the presences of MLS as crosslinker and KPS as initiator at 70 °C (Scheme 2). The control experiment for the synthesis of LNIH hydrogel without use of MLS failed to receive a gel product, revealing that MLS played a crosslinker role in the formation of hydrogel network. Determinations of carbon (C%), hydrogen (H%), nitrogen (N%) and sulphur (S%) contents by elemental analysis further confirmed the formation of the LNIH hydrogels (Table 1). It was noted that the MLS content (MLS%) in the washed hydrogel product increased with the concentration of MLS crosslinker in the reaction, ranging from 3.7 to 14.3 wt%. These results are roughly in good agreement with the theoretical values of MLS% by assuming complete conversion of MLS, suggesting the polymerization condition in this study was sufficient. Hereafter, the hydrogels are named as LNIH-3.7%, LNIH-5.7%, LNIH-6.9%, LNIH-8.6% and LNIH-14.3% in which the numbers correspond to the actual MLS contents in individual hydrogels, respectively.

Scheme 2. Synthesis route of LNIH hydrogel from the reaction of NIPAM, IA and MLS.

11

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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 30

Table 1. The elemental compositions of MLS and LNIHs. MLS% MLS% (calculated (measured)b )c

MLS (mg) amount in copolymerization

C%

H%

N%

S%

MLS

-

48.98

5.68

0.18

4.05

-

-

LNIH-3.7%

30

58.86

9.08

10.37

0.15

3.7

2.8

LNIH-5.7%

50

58.26

9.06

10.28

0.23

5.7

4.5

LNIH-6.9%

75

58.89

8.91

10.24

0.28

6.9

6.7

LNIH-8.6%

100

58.98

8.93

10.16

0.35

8.6

8.7

LNIH-14.3%

200

58.81

8.70

9.61

0.58

14.3

16.0

a

Sample

a

Elemental composition

The hydrogels were prepared using NIPAM (1000 mg), IA (50 mg), various amounts of MLS

and KPS (2 wt% of total weight of reactants) in deionized water (5 mL/1 g reactants) under an argon atmosphere. The hydrogels were then washed with deionized water; b MLS content (wt%) in hydrogel was calculated from the measured S content by referring the S content of pure MLS (4.05 wt%); c Calculated MLS content (wt%) in the gel by assuming complete conversion for each comonomer.

The evidence for the successful preparation of crosslinked hydrogels could also be found from the FT-IR analysis (Figure 3). Several characteristic peaks of MLS, including hydroxyl groups (-OH) stretching at 3470 cm-1, benzene ring vibration at 1506 cm-1 and C-H bending from benzene ring at 1017 cm-1, were found in the spectrum of LNIH-8.6% sample, respectively. Meanwhile, the peak around 1545 cm-1 corresponding to the amide groups (-NHCO, amide II 12

ACS Paragon Plus Environment

Page 13 of 30 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

ACS Sustainable Chemistry & Engineering

band) from NIPAM was also noted in the spectrum of LNIH-8.6%. In addition, the LNIH-8.6% sample exhibited an obvious shoulder peak around 1724 cm-1 attributed to the carboxyl groups (-COOH) from the IA moiety.

Figure 3. FT-IR spectra of IA, NIPAM, MLS and LNIH-8.6%.

Figure 4 shows the 13C NMR spectra of IA (in solution), NIPAM (in solution), MLS (in solid) and LNIH-8.6% (in solid). The peaks at 126 - 134 ppm ascribed to the double bond =CH2 from IA, NIPAM and MLS, respectively, nearly disappeared in the spectrum of LNIH-8.6%, suggesting a better completion of the copolymerization. Since NIPAM was a major comonomer in the hydrogel, the characteristic signals at 21.26 ppm (-CH3) and 41.66 ppm (-CH) from the NIPAM moiety could be seen in the LNIH-8.6% spectrum, respectively. Furthermore, the signal 13

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

at 173.63 ppm related to the carbons (C=O) from NIPAM, MLS and IA could also be easily recognized in the spectrum of LNIH-8.6%. Both the FT-IR and

13

C NMR results confirmed the

successful copolymerization of MLS, NIPAM and IA.

Figure 4. 13C NMR spectra of IA (in solution), NIPAM (in solution), MLS (in solid) and LNIH-8.6% (in solid).

The morphologies of the freeze-dried hydrogel samples were investigated by SEM (Figure 5). The obtained hydrogel samples exhibited porous and mesh-like morphological structures after the removal of water from the hydrogel networks under the freeze-drying condition. This observed porous structure of freeze-dried hydrogel was due to the phase separation of the gel during rapid cooling (freezing) and subsequent removal of the solvent by sublimation which left voids in place where the solvent previously occupied. In addition, it was noted that the pore size 14

ACS Paragon Plus Environment

Page 14 of 30

Page 15 of 30 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

ACS Sustainable Chemistry & Engineering

of hydrogel became larger with MLS content. A similar phenomenon was also previously noted in the study of another lignin-based hydrogel material.23 This result was likely due to the increased crosslink density that caused faster phase separation during freezing, resulting in formation of more coarse structure.

(a) LNIH-3.7%

(b) LNIH-5.7%

(d) LNIH-8.6%

(c) LNIH-6.9%

(e) LNIH-14.3%

Figure 5. SEM micrographs of LNIH hydrogels (a) LNIH-3.7%, (b) LNIH-5.7%, (c) LNIH-6.9%, (d) LNIH-8.6% and (e) LNIH-14.3% crosslinked with different MLS contents.

Swelling Properties. In Figure 6, the gel content increased from 62.4 to 80.8% with MLS content increasing from 3.7 to 14.3% in the reaction. Meanwhile, the swelling capacity decreased from 32.1 to 20.2 g/g with 15

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

increasing MLS content. Based on these results, it can be assumed that the decrease in swelling capacity was associated with the increase in crosslink density of the gel.

Figure 6. Swelling capacity (g/g) and gel content (G%) of LNIH hydrogels.

The swelling rate of the hydrogel was also investigated as a function of time at 25 °C (Figure 7a). It was noted that the amount of absorbed water increased rapidly during the initial swelling for each hydrogel and then increased slowly until reached equilibrium. The swelling kinetics of LNIHs in deionized water were explored using a pseudo-second-order model based on equation 3.26

(3) 16

ACS Paragon Plus Environment

Page 16 of 30

Page 17 of 30 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

ACS Sustainable Chemistry & Engineering

where SCt is the swelling capacity of swollen hydrogel at swelling time t, K the swelling rate constant, and SCeq the swelling capacity at equilibrium time. As displayed in Figure 7b, all plots of t/SCt versus t exhibited an ideal straight line with high correlation coefficients, which demonstrated that the swelling behaviors of hydrogels could be well described by the pseudo-second-order model.27

Figure 7. (a) Swelling rate and (b) pseudo-second-order kinetics of hydrogels in deionized water. All the values of correlation coefficient (R2) > 0.998.

Table 2. Kinetic parameters of different LNIHs for absorption of deionized water. Sample

SCeq (g/g)a

SCexp (g/g)b

K [g/(g·s)]

LNIH-3.7%

32.6

31.6

3.6×10-5

LNIH-5.7%

28.9

28.4

6.0×10-5

LNIH-6.9%

25.4

24.5

3.7×10-5

17

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

a

Page 18 of 30

LNIH-8.6%

22.7

22.5

11.5×10-5

LNIH-14.3%

19.4

19.1

13.1×10-5

Calculated equilibrium swelling capacity; b Experimental equilibrium swelling capacity.

Based on the fitting curves, the kinetic parameters including swelling rate constant (K) and the theoretical equilibrium swelling capacity (SCeq) in swelling process were obtained and are listed in Table 2. The values of calculated equilibrium swelling capacity were in good agreement with the experimental data for all hydrogels. By analyzing all the values of SCeq, SCexp and K for each LNIH, it is noted that the most crosslinked LNIH-14.3% possessed the lowest calculated and experimental values of swelling capacity (SCeq and SCexp) but the highest swelling rate constant (K). This result is likely the consequence of two completing mechanisms in the hydrogel. As the sulfonate-containing MLS is a strong water-absorbing moiety in the resulting hydrogel, thus increasing MLS content would result in increase of swelling capacity; on the other hand, the crosslink density increases with MLS content, which would result in drastic decrease in swelling capacity. According to the assumption of pseudo-second-order model, the swelling rate of the crosslinked LNIH hydrogel was controlled by relaxation process of polymer chains.28

Temperature Responsive Behavior. The NIPAM-type hydrogel exhibits volume-phase transition performance due to the conformation variation of NIPAM chains on network while the temperature is set below or above the volume phase transition temperature (VPTT).29, poly-NIPAM hydrogels,31,

32

30

In Figure 8a, like other typical

LNIH-3.7%, LNIH-5.7% and LNIH-6.9% remained swollen in 18

ACS Paragon Plus Environment

Page 19 of 30 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

ACS Sustainable Chemistry & Engineering

deionized water at lower temperatures but experienced a sharp deswelling process with increase in temperature. However, the temperature responsive behavior was not clearly noted for both the LNIH-8.6% and LNIH-14.3% samples. The more rigid architecture of hydrogel was formed with higher MLS content, which overwhelmed the volume-phase transition performance of NIPAM chains in network above VPTT.

In Figure 8b, DSC analysis indicated the VPTTs of LNIH-3.7%, LNIH-5.7% and LNIH-6.9% were around 34.5, 34.5 and 35.5 °C, respectively, which were close to the human physiological temperature (37 °C). Increasing crosslinker content in polymeric network resulted in slightly higher VPTT values for LNIHs than that of a typical poly-NIPAM hydrogel (~32 °C ),33 revealing that the higher temperature was needed to drive the disruption of hydrogen bonds strengthened by the increased MLS content. Furthermore, LNIH-8.6% and LNIH-14.3% showed no obvious VPTT peak in DSC measurement, which were consistent with the results of temperature responsive swelling test.

19

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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 8. (a) Temperature-sensitive behaviors of LNIH hydrogels and (b) DSC curves of LNIH hydrogels in the temperature range from 12 to 80 °C.

pH Responsive Behavior. The pH responsive behaviors of LNIH hydrogels were studied within the range of pH 3.0 to 9.1 in buffers below and above VPTT (Figure 9, S2 and S3). At 25 ℃ which was below VPTT, all hydrogels exhibited lower swelling capacities in 20 mM buffers with a 50 mM ionic strength than in pure water, owing to the higher ionic strength of external solution.2 The swelling capacities of all hydrogels increased with pH increasing up to 7.4 and then decreased with further increase of the pH value (Figure 9a). In view of the chemical structure of the hydrogel, introduction of itaconic acid (pKa1 3.85 and pKa2 5.44) mainly contributed to the pH sensitivity of hydrogels. Under strong acidic condition, the carboxyl groups in the polymeric network existed in acid form (-COOH) and could form intermolecular hydrogen bonds. This internal hydrogen bonding resulted in the unfavorable swelling behavior and lower swelling capacity for hydrogels. As the pH value increased to 7.4, the carboxyl groups gradually transformed into the ionized carbonate form (-COO-) which led to stronger hydrophilicity and higher electrostatic repulsion of the network, triggering the increased water absorption capacity.34 However, the repulsion of the negative -COO- groups from itaconic acid and -SO3- groups from MLS would be shielded by more Na+ ions in the basic condition (pH > 7.4) for screening effect, leading to the shrinking of hydrogels, thus their swelling capacities decreased subsequently. At 42 ℃, all LNIH hydrogels exhibited relatively lower swelling capacities compared with that at 25 ℃ due to their temperature-responsive shrinking behaviors above VPTT (Figure 9b). Meanwhile, it should be 20

ACS Paragon Plus Environment

Page 20 of 30

Page 21 of 30 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

ACS Sustainable Chemistry & Engineering

noted that LNIH hydrogels still maintained pH responsive behaviors at 42 ℃ in the experimental pH range which were consistent with the observations at 25 °C,35 as has been discussed above. These results suggest that the swelling behavior of LNIH hydrogels can be manipulated by varying pH of the solution.

Figure 9. Swelling capacities of LNIH hydrogels at different pH values under (a) 25 °C and (b) 42 °C.

Rheological Properties. Rheological properties of LNIH hydrogels with different MLS contents were measured at 25 °C. Because of the hydrogel network structure, the storage modulus (G′) of all LNIH hydrogels were higher than the corresponding loss modulus (G″) over the whole selected angular frequency range (Figure 10a and 10b). Besides, G′ showed a monotonous increase with MLS content in the gel. This result was most likely attributed to the rigid structure of lignin and increased crosslink density,23, 36 leading to the increase of stiffness of the network structure. G″ also increased with 21

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

MLS content in the gel, which was a consequence of higher crosslink density of the gel leading to more heat dissipation for chain segment movement. The positive effect of MLS content on the mechanical properties of LNIH hydrogels could also be observed in their stress-strain curves (Figure 10c), where LNIH-14.3% presented much higher stress values than the other hydrogels over the entire examined strain range. Together with the observation of G′ and G″ values, these results indicate that increasing the content of the multifunctional MLS crosslinker can dramatically improve the mechanical properties of the hydrogels.

22

ACS Paragon Plus Environment

Page 22 of 30

Page 23 of 30 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

ACS Sustainable Chemistry & Engineering

Figure 10. Curves of (a) Storage modulus (G′, 1% strain), (b) loss modulus (G″, 1% strain) and 23

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

(c) stress-strain for LNIH hydrogels with various crosslinker contents. All the hydrogel test specimens contained 83.3 wt% water.

Conclusions We demonstrated that a methacrylated lignosulfonate derivative (MLS) was utilized as a novel crosslinker in the preparation of lignosulfonate-NIPAM-IA copolymer hydrogels (LNIH) with dual temperature and pH responsive properties. Use of methacrylated lignosulfonate as crosslinker enabled efficient crosslinking of the copolymer hydrogels with acceptable gel contents. Study of swelling kinetics revealed that the pseudo-second-order model was suitable for describing the water absorption of LNIH hydrogels. LNIH-3.7%, LNIH-5.7% and LNIH-6.9% hydrogels showed temperature-sensitive behaviors around 35 °C, which were very close to the physiological temperature (37 °C). Additionally, LNIH hydrogels also exhibited pH-sensitive due to the carboxyl groups from IA moieties. Moreover, shear strength and rheological properties of LNIH hydrogels increased with MLS content in the polymeric network. The lignosulfonate-based hydrogels with stimuli-responsive behaviors may have great potential for controlled release of some pesticides or drugs in various conditions.

Supporting Information The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acssuschemeng.xxxxx. Additional supporting figures (Figure S1-S3) for LNIH hydrogels under different pH and temperature conditions, and 31P NMR results (Table S1) for samples (PDF). 24

ACS Paragon Plus Environment

Page 24 of 30

Page 25 of 30 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

ACS Sustainable Chemistry & Engineering

Acknowledgements This work was supported by the National Natural Science Foundation of China (No. 31400516) and Fundamental Research Funds for the Central Non-profit Research Institution of Chinese Academy of Forestry (No. CAFYBB2014QB043).

References 1. Ni, M.; Zhang, N.; Xia, W.; Wu, X.; Yao, C.; Liu, X.; Hu, X. Y.; Lin, C.; Wang, L., Dramatically Promoted Swelling of a Hydrogel by Pillar[6]arene-Ferrocene Complexation with Multistimuli Responsiveness. J. Am. Chem. Soc. 2016, 138 (20), 6643-6649. 2. Chang, C.; He, M.; Zhou, J.; Zhang, L., Swelling Behaviors of pH- and Salt-Responsive Cellulose-Based Hydrogels. Macromolecules 2011, 44 (6), 1642-1648. 3. Tamate, R.; Ueki, T.; Kitazawa, Y.; Kuzunuki, M.; Watanabe, M.; Akimoto, A. M.; Yoshida, R.,

Photo-Dimerization

Induced

Dynamic

Viscoelastic

Changes

in

ABA

Triblock

Copolymer-Based Hydrogels for 3D Cell Culture. Chem. Mater. 2016, 28 (17), 6401-6408. 4. Liao, X. J.; Chen, G. S.; Jiang, M., Hydrogels locked by molecular recognition aiming at responsiveness and functionality. Polym. Chem. 2013, 4 (6), 1733-1745. 5. Pandey, M.; Mohamad, N.; Amin, M. C., Bacterial cellulose/acrylamide pH-sensitive smart hydrogel: development, characterization, and toxicity studies in ICR mice model. Mol. Pharmaceutics 2014, 11 (10), 3596-3608. 6. Culver, H. R.; Clegg, J. R.; Peppas, N. A., Analyte-Responsive Hydrogels: Intelligent Materials for Biosensing and Drug Delivery. Acc. Chem. Res. 2017, 50 (2), 170-178. 7. Kahn, J. S.; Hu, Y.; Willner, I., Stimuli-Responsive DNA-Based Hydrogels: From Basic Principles to Applications. Acc. Chem. Res. 2017, 50 (4), 680-690. 8. Hoogenboom, R.; Schlaad, H., Thermoresponsive poly(2-oxazoline)s, polypeptoids, and polypeptides. Polym. Chem. 2017, 8 (1), 24-40. 9. Zhang, J.; Chu, L.-Y.; Li, Y.-K.; Lee, Y. M., Dual thermo- and pH-sensitive 25

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

poly(N-isopropylacrylamide-co-acrylic acid) hydrogels with rapid response behaviors. Polymer 2007, 48 (6), 1718-1728. 10. Kocak, G.; Tuncer, C.; Bütün, V., pH-Responsive polymers. Polym. Chem. 2017, 8 (1), 144-176. 11. Mohd Amin, M. C. I.; Ahmad, N.; Halib, N.; Ahmad, I., Synthesis and characterization of thermo- and pH-responsive bacterial cellulose/acrylic acid hydrogels for drug delivery. Carbohydr. Polym. 2012, 88 (2), 465-473. 12. El-Sherbiny, I. M., Enhanced pH-responsive carrier system based on alginate and chemically modified carboxymethyl chitosan for oral delivery of protein drugs: Preparation and in-vitro assessment. Carbohydr. Polym. 2010, 80 (4), 1125-1136. 13. Tan, Y.; Xu, K.; Wang, P.; Li, W.; Sun, S.; Dong, L., High mechanical strength and rapid response rate of poly(N-isopropyl acrylamide) hydrogel crosslinked by starch-based nanospheres. Soft Matter 2010, 6 (7), 1467-1471. 14. Song, W.; Xin, J.; Zhang, J., One-pot synthesis of soy protein (SP)-poly(acrylic acid) (PAA) superabsorbent hydrogels via facile preparation of SP macromonomer. Ind. Crop. Prod. 2017, 100, 117-125. 15. Aro, T.; Fatehi, P., Production and Application of Lignosulfonates and Sulfonated Lignin. ChemSusChem 2017, 10 (9), 1861-1877. 16. Yang, D.; Qiu, X.; Pang, Y.; Zhou, M., Physicochemical Properties of Calcium Lignosulfonate with Different Molecular Weights as Dispersant in Aqueous Suspension. J. Dispersion Sci. Technol. 2010, 29 (9), 1296-1303. 17. He, K.; Lou, T.; Wang, X.; Zhao, W., Preparation of lignosulfonate-acrylamide-chitosan ternary graft copolymer and its flocculation performance. Int. J. Biol. Macromol. 2015, 81, 1053-1058. 18. Liang, F.-B.; Song, Y.-L.; Huang, C.-P.; Li, Y.-X.; Chen, B.-H., Synthesis of Novel Lignin-Based Ion-Exchange Resin and Its Utilization in Heavy Metals Removal. Ind. Eng. Chem. Res. 2013, 52 (3), 1267-1274. 19. Shogren, R. L.; Biswas, A., Preparation of starch-sodium lignosulfonate graft copolymers via laccase catalysis and characterization of antioxidant activity. Carbohydr. Polym. 2013, 91 (2), 26

ACS Paragon Plus Environment

Page 26 of 30

Page 27 of 30 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

ACS Sustainable Chemistry & Engineering

581-585. 20. Li, F.; Wang, X.; Yuan, T.; Sun, R., A lignosulfonate-modified graphene hydrogel with ultrahigh adsorption capacity for Pb(II) removal. J. Mater. Chem. A 2016, 4 (30), 11888-11896. 21. Wang, X.; Zhou, Z.; Guo, X.; He, Q.; Hao, C.; Ge, C., Ultrasonic-assisted synthesis of sodium lignosulfonate-grafted poly(acrylic acid-co-poly(vinyl pyrrolidone)) hydrogel for drug delivery. RSC Adv. 2016, 6 (42), 35550-35558. 22. Yu, C.; Wang, F.; Zhang, C.; Fu, S.; Lucia, L. A., The synthesis and absorption dynamics of a lignin-based hydrogel for remediation of cationic dye-contaminated effluent. React. Funct. Polym. 2016, 106, 137-142. 23. Teng, X.; Xu, H.; Song, W.; Shi, J.; Xin, J.; Hiscox, W. C.; Zhang, J., Preparation and Properties of Hydrogels Based on PEGylated Lignosulfonate Amine. ACS Omega 2017, 2 (1), 251-259. 24. Zhang, J.; Li, A.; Wang, A., Study on superabsorbent composite. VI. Preparation, characterization and swelling behaviors of starch phosphate-graft-acrylamide/attapulgite superabsorbent composite. Carbohydr. Polym. 2006, 65 (2), 150-158. 25. Jin, C.; Zhang, X.; Xin, J.; Liu, G.; Wu, G.; Kong, Z.; Zhang, J., Clickable Synthesis of 1,2,4-Triazole Modified Lignin-Based Adsorbent for the Selective Removal of Cd(II). ACS Sustainable Chem. Eng. 2017, 5 (5), 4086-4093. 26. Wang, W.; Wang, A., Nanocomposite of carboxymethyl cellulose and attapulgite as a novel pH-sensitive superabsorbent: Synthesis, characterization and properties. Carbohydr. Polym. 2010, 82 (1), 83-91. 27. Spagnol, C.; Rodrigues, F. H. A.; Pereira, A. G. B.; Fajardo, A. R.; Rubira, A. F.; Muniz, E. C.,

Superabsorbent

hydrogel

composite

made

of

cellulose

nanofibrils

and

chitosan-graft-poly(acrylic acid). Carbohydr. Polym. 2012, 87 (3), 2038-2045. 28. Schott, H., Swelling kinetics of polymers. J. Macromol. Sci. Part B 1992, 31 (1), 1-9. 29. Pelton, R., Poly(N-isopropylacrylamide) (PNIPAM) is never hydrophobic. J. Colloid Interface Sci. 2010, 348 (2), 673-674. 30. Su, W.; Yang, M.; Zhao, K.; Ngai, T., Influence of Charged Groups on the Structure of Microgel and Volume Phase Transition by Dielectric Analysis. Macromolecules 2016, 49 (20), 27

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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

7997-8008. 31. Zhang, X.; Wang, Y.; Zhao, J.; Xiao, M.; Zhang, W.; Lu, C., Mechanically Strong and Thermally Responsive Cellulose Nanofibers/Poly(N-isopropylacrylamide) Composite Aerogels. ACS Sustainable Chem. Eng. 2016, 4 (8), 4321-4327. 32. Vagias, A.; Košovan, P.; Koynov, K.; Holm, C.; Butt, H.-J.; Fytas, G., Dynamics in Stimuli-Responsive Poly(N-isopropylacrylamide) Hydrogel Layers As Revealed by Fluorescence Correlation Spectroscopy. Macromolecules 2014, 47 (15), 5303-5312. 33. Ono, Y.; Shikata, T., Contrary hydration behavior of N-isopropylacrylamide to its polymer, P(NIPAm), with a lower critical solution temperature. J. Phys. Chem. B. 2007, 111 (7), 1511-1513. 34. Huang, Y.; Zeng, M.; Ren, J.; Wang, J.; Fan, L.; Xu, Q., Preparation and swelling properties of graphene oxide/poly(acrylic acid-co-acrylamide) super-absorbent hydrogel nanocomposites. Colloids Surfaces A 2012, 401, 97-106. 35. Zhang, Q.; Zha, L.; Ma, J.; Liang, B., A novel route to prepare pH- and temperature-sensitive nanogels via a semibatch process. J. Colloid Interface Sci. 2009, 330 (2), 330-336. 36. Le Goff, K. J.; Gaillard, C.; Helbert, W.; Garnier, C.; Aubry, T., Rheological study of reinforcement of agarose hydrogels by cellulose nanowhiskers. Carbohydr. Polym. 2015, 116, 117-123.

28

ACS Paragon Plus Environment

Page 28 of 30

Page 29 of 30 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

ACS Sustainable Chemistry & Engineering

For Table of Contents Use Only

Biobased hydrogels with temperature- and pH- responsive properties have been fabricated by copolymerizing N-isopropylacrylamide (NIPAM) and itaconic acid (IA) using methacrylated lignosulfonate (MLS) as a macro-crosslinker.

29

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 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 30 of 30

Graphic Abstract

Biobased hydrogels with temperature- and pH- responsive properties have been fabricated by copolymerizing N-isopropylacrylamide (NIPAM) and itaconic acid (IA) using methacrylated lignosulfonate (MLS) as a macro-crosslinker.

ACS Paragon Plus Environment