wheat straw-based hydrogels as soil conditioners?

From agricultural by-products to value-added. 2 materials: wheat straw-based hydrogels as soil. 3 conditioners? 4. Katja Heise†τ*, Maximilian Kirst...
33 downloads 0 Views 3MB Size
Subscriber access provided by OCCIDENTAL COLL

Article

From agricultural by-products to value-added materials: wheat straw-based hydrogels as soil conditioners? Katja Heise, Maximilian Kirsten, Yvonne Schneider, Doris Jaros, Harald Keller, Harald Rohm, Karsten Kalbitz, and Steffen Fischer ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.9b00378 • Publication Date (Web): 28 Mar 2019 Downloaded from http://pubs.acs.org on April 3, 2019

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 45 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

254x190mm (96 x 96 DPI)

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 2 of 45

1

From agricultural by-products to value-added

2

materials: wheat straw-based hydrogels as soil

3

conditioners?

4

Katja Heise†*, Maximilian Kirsten‡, Yvonne Schneider#, Doris Jaros#, Harald Keller, Harald

5

Rohm#, Karsten Kalbitz‡, Steffen Fischer†

6

† Institute of Plant and Wood Chemistry, Technische Universität Dresden, Pienner

7

Strasse 19, 01737 Tharandt, Germany

8

‡ Institute of Soil Sciences and Site Ecology, Technische Universität Dresden, Pienner

9

Strasse 19, 01737 Tharandt, Germany

10

# Chair of Food Engineering, Technische Universität Dresden, Bergstrasse 120, 01069

11

Dresden, Germany

12

* Corresponding Author: Katja Heise, E-mail: [email protected]

13

 BASF SE, Carl-Bosch-Strasse 38, 67056 Ludwigshafen, Germany

14

KEYWORDS Biomass valorization, Moisture sorption, Soil amendment, Water retention

ACS Paragon Plus Environment

1

Page 3 of 45 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

15

ACS Sustainable Chemistry & Engineering

ABSTRACT

16

Herein, we present a simple synthetic approach to fabricate wheat straw-based

17

hydrogels, starting from the unfractionated and carboxymethylated lignocellulosic

18

matrix. Citric acid was used as a cheap and non-toxic crosslinker. The applied hydrogel

19

characterizations can be essentially distinguished into investigations on the synthetic

20

pathway and model-scale application-related tests. For the first part, three sample-

21

specific values were introduced: gel yield (%), swelling ratio (gwater/g) and gel stiffness (G´,

22

Pa). Optimized reaction conditions led to mechanically stable gels with a moderate

23

swelling ratio (up to 50 gwater/g). Moreover, dynamic vapor sorption analysis revealed

24

that these gels re-swell after complete drying. Finally, one selected hydrogel was

25

incorporated into two different model soil substrates, assessing its impact on the soils´

26

water retention. Our experiments showed that already low incorporation rates

27

(0.2 wt %) increased the water content of a sandy soil by 70 % (at pF 2.53). Overall, these

28

results are promising and may lead to new soil amendments based on a sustainable

29

source and a simple synthesis.

ACS Paragon Plus Environment

2

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

30

Page 4 of 45

INTRODUCTION

31

Inspired by Nature´s zero-waste principle, the valorization of lignocellulosic biomass

32

wastes has an enormous economic and ecological potential, as they are produced in

33

excess, globally, and their recycling allows for circular production strategies in

34

agriculture and forestry.1–3

35

To develop individual value-chains for cellulose, hemicelluloses and lignin, recent

36

biorefinery concepts prevalently target biomass fractionation using elaborate

37

pretreatments entailing large expenses for chemicals and energy.4 Alternatively, starting

38

from unfractionated lignocelluloses in simplified operations may considerably reduce

39

processing efforts and increase the economic value of biomass residues for selected

40

applications.2

41

In this regard, two product groups have been in the focus of applied research

42

recently:

lignocellulose-based

bioadsorbents

for

waste

43

lignocellulosic fiber-reinforcements for composite materials.5,6 A third and very

44

promising idea leads back to agriculture and considers applying lignocellulose-based

45

hydrogels to soils as miniature water reservoirs and nutrient carriers.7 These three

46

concepts have one essential principle in common: they use chemical and/or physical

47

treatments to optimize inherent polymer characteristics of the biomass matrix toward

48

the intended application.7–10

ACS Paragon Plus Environment

water

treatment

and

3

Page 5 of 45 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

49

Linear and crosslinked polymers have widely been investigated as soil amendments

50

to prevent soil erosion or for improving the soils´ water holding capacity.11,12 In many

51

cases, superabsorbent polymers (SAPs) are thereby based on crosslinked polyacrylate or

52

polyacrylamide, which were found to absorb enormous amounts of water (10-10,000

53

g/g) and improve the availability of water and fertilizers in soils.13 However, despite the

54

successful use of SAPs in horticultural industry, their high costs and hampered

55

biodegradability have so far restricted an application on agricultural scale.13,14

56

Lignocelluloses offer an ideal polymeric framework for the fabrication of hydrogels

57

for soil usage owing to their inherent hydrophilic nature, biodegradability and

58

modifiability. Moreover, compared to purely cellulosic substrates, residual lignin

59

within the gel matrix might slow down biodegradation preserving the gels´ structural

60

integrity over an adequate time period (e.g. one growth season).15,16 Nevertheless,

61

synthetic concepts for lignocellulosic hydrogels are still scarce. In current strategies,

62

lignocelluloses (e.g. cotton stalks, rice straw, sugarcane bagasse) basically serve as

63

biopolymer backbone for the synthesis of acrylate-based graft copolymers.7,17–20 The

64

obtained semisynthetic SAPs exhibited excellent water absorbencies (up to 350 g/g) and,

65

therefore, improved soils´ water retention. Nevertheless, the use of acrylate-derived

66

products may be questioned, as leaching and accumulation of potentially eco-toxic

67

monomers in natural systems should be mandatorily avoided.21,22 Furthermore, a

68

complete degradation of soil amendments should be guaranteed.

ACS Paragon Plus Environment

4

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 6 of 45

69

Polycarboxylic acids may be a very promising alternative in this regard, as they were

70

found to effectively crosslink cellulose without the need for initiators or organic

71

solvents.23–25 Among them, citric acid has gained increasing interest for preparing

72

cellulose-based films or superabsorbent hydrogels, taking into account its low price,

73

natural origin and non-toxicity.26,27 Moreover, implementations are simple, as the

74

esterification of cellulose follows an easy heat-induced mechanism: citric acid

75

dehydrates over two reactive anhydride stages and thereby crosslinks adjacent cellulose

76

chains.28

77

Inspired by this simple approach, we used citric acid to crosslink unfractionated and

78

carboxymethylated wheat straw, which represents a class of highly available biomass

79

residues with an estimated annual production of above 150 million tons in the

80

European Union (FAO, 2017).29 The obtained gels were analyzed extensively focusing

81

on the synthetic pathway and their potential application as soil conditioners.

82 83

MATERIALS AND METHODS

84

Materials. Chopped wheat straw (WS; Agrargenossenschaft Rossau e.G. Germany)

85

was ball milled (MM 400, Retsch® GmbH, Haan, Germany): 5 min, f = 20 rpm, median

86

particle size 43 m. All chemicals were applied in analytical state. Deionized water was

87

utilized for all experiments. For water retention tests, a sandy (from subsoil, soil depth

88

> 40 cm, Duebener Heide, Germany) and a silty model soil substrate (from limestone

ACS Paragon Plus Environment

5

Page 7 of 45 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

89

quarry, soil depth = 300 cm, Ostrau, Germany) were collected, air-dried, sieved to < 2

90

mm and stored at room temperature (RT). Soil characteristics are summarized in the SI

91

(Table S1).

92

Hydrogel preparation. Carboxymethyl (CM) intermediates were synthesized and

93

purified as described previously.30 Milled WS was modified with sodium

94

monochloroacetate (MCA) in isopropanol/NaOH (aq) as follows: (i) alkalization –

95

1 hour, RT; (ii) carboxymethylation after MCA addition – 55 °C, 3 hours. Per mol

96

anhydrous glucose units of cellulose (AGUcell, 48.9 % in wheat straw) 1.5, 3.0 and 6.0

97

mol MCA were used for WS1, WS2, and WS3, resulting in carboxymethyl group

98

contents of 1.75, 2.56 and 3.79 mmol/g, respectively. The product work-up was carried

99

out by: decanting remaining medium, product dispersion in 100 mL water,

100

neutralization with acetic acid, followed by disperser-assisted precipitation in 500 mL

101

96-% ethanol and, finally, 5 repeated dispersion-precipitation steps in water/ethanol.

102

Carboxymethyl group contents were determined by ICP-OES (Na determination,

103

device: SPECTRO CIROSCCD, SPECTRO Analytical Instruments, Kleve, Germany)

104

after dialysis against deionized water (membrane: SpectraPor® 3, MWCO: 3.5 kDa),

105

lyophilization

106

HNO3/HF/HClO4.30,31

and

microwave-assisted

decomposition

of

sample

aliquots

in

107

After carboxymethylation, CM-intermediates were crosslinked with citric acid

108

monohydrate. 2, 4, 8 or 16 wt % citric acid (w/w dry CM-intermediate) were dissolved

ACS Paragon Plus Environment

6

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 8 of 45

109

in 45 mL deionized water and 5 g of dry CM-intermediate were suspended in the

110

solution. The mixture was kept for 24 hours at RT to homogeneously distribute the

111

crosslinker. Subsequently, the mixture was spread on a Petri dish (thin layer) and pre-

112

dried at 60 °C for 24 hours, followed by crosslinking at temperatures of 120, 140 or

113

160 °C for durations of 15, 30, 60 or 300 minutes. Afterwards, the product was gently

114

crushed, sieved to a defined particle size (150 to 800 µm) and suspended in deionized

115

water for 24 hours at RT. Finally, by-products and water-soluble fractions were

116

removed by repeated water-washing and the purified gel particles were oven-dried at

117

45 °C.

118

Hydrogel characterization. Analytical characterizations of straw-based hydrogels

119

were distinguished into investigations on the synthetic pathway, gel-moisture-

120

interactions and model-scale water retention tests in soil substrates.

121

ATR-IR spectroscopy. Attenuated total reflection (ATR) FT IR spectroscopy was

122

performed with a Bruker Tensor 27 (Bruker Optik GmbH, Ettlingen, Germany), having

123

a RT-DLaTGS detector, KBr beam splitter and Platinum ATR (A225) diamond cell. ATR

124

spectra were measured as follows: range = 4000 – 400 cm-1, resolution = 4 cm-1, 400

125

scans. Average spectra were formed (OPUS Ver. 6.5, Bruker) out of three

126

determinations.

127

Gel yield. The gravimetric gel yield (in %) after esterification with citric acid was

128

determined in duplicate by suspending 0.1 g of dry and sieved (150-800 m) sample

ACS Paragon Plus Environment

7

Page 9 of 45 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

129

particles (before removing water-soluble fractions) in 10 mL deionized water for

130

24 hours at RT. Water-soluble proportions were then removed by washing the swollen

131

particles over a pre-dried (at 105 °C) and weighed glass fiber filter (WhatmanTM, GF6)

132

with deionized water, followed by drying the loaded filter (105 °C) to mass constancy.

133

The gel yield % is defined as: ((mtotal – mF) * 100 %)/ m0; with mtotal as total mass of

134

recovered gel particles on the filter, mF as mass of the filter and m0 the initial dry mass

135

(determined in duplicate by drying the samples to mass constancy at 105 °C).

136

Swelling ratio. 0.1 g dry and sieved (150-800 m) gel particles were weighed into a

137

heat-sealable teabag and were swollen in 700 mL deionized water for 24 hours at RT.

138

The teabag was then removed from the water, unabsorbed water drained off for 10

139

minutes, followed by weighing the swollen hydrogel within the bag. The swelling ratio

140

(in gwater g-1, duplicate measurements) is defined as: (mtotal - mTB - WTB)/ mGEL; where mtotal

141

is the total mass of swollen gel within the bag, mTB and WTB are mass and water uptake

142

of the teabag, respectively, and mGEL the dry-mass related sample mass.

143

Gel rheology. Rheological measurements were carried out in duplicate with sieved

144

(150-800 m) and fully water-swollen (24 h/ RT) gel particles using an AR-G2 rheometer

145

(TA Instruments, New Castle, USA) with parallel plates (d = 25.0 mm) at 23 ± 1 °C.

146

Dynamic strain sweep tests were conducted at  = 1.0 rad/s, in a strain range of 0.001 to

147

1.0 (10 points per decade) to determine the linear viscoelastic (LVE) region. Dynamic

148

frequency sweep tests ( = 0.1 to 100.0 rad/s, 10 points per decade) were performed at a

ACS Paragon Plus Environment

8

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 10 of 45

149

strain of 0.002. The gel stiffness (G´ Pa) was taken from frequency sweep experiments

150

at  = 1.0 rad/s.32

151

Dynamic vapor sorption analysis (DVS). Gel-moisture-interactions were investigated for

152

the following selected gel samples: WS1-CA4%, WS3-CA4% and WS3-CA8%, based on

153

WS1 (1.75 mmol –CH2COONa g-1) or WS3 (3.79 mmol –CH2COONa g-1), respectively;

154

esterification: 140 °C, 30 min, 4 or 8 wt % citric acid (as indicated). The measurements

155

were carried out in duplicate using a Q5000 SA dynamic vapor sorption analyzer (TA

156

Instruments, New Castle, USA). Approx. 5 mg sieved (150-800 m) and preconditioned

157

(10 d over P2O5) xerogel particles were loaded into a quartz crucible. Relative humidity

158

(RH) in the sample chamber (25 °C) was adjusted by mixing continuous streams of dry

159

and moisture-saturated nitrogen with continuous flow. Initial mass of the dry sample

160

was determined at RH = 0 %. In the first adsorption cycle (Ad1), RH was increased to 90

161

% in 10 % increments. At each RH step, equilibrium mass was taken when the relative

162

mass change was < 0.01 % for 5 min. The second cycle comprised of stepwise desorption

163

to RH = 0 % (De1), the third cycle another adsorption to RH = 90 % (Ad2).

164

Sorption isotherms were obtained by plotting equilibrium moisture contents of the

165

samples (Meq, mgwater g-1sample) against relative humidity (expressed as water activity aw

166

-). Sorption hysteresis was calculated as absolute difference between desorption and

167

first adsorption Meq. Deviations between the second (Ad2) and first adsorption Meq were

168

calculated to examine structural alterations within the gel matrices. Moreover, for

ACS Paragon Plus Environment

9

Page 11 of 45 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

169

determining monolayer capacity (M0 mgwater g-1sample) of the samples, sorption data

170

were fitted to the Guggenheim-Anderson-de Boer (GAB) model: cG, GAB monolayer

171

sorption energy; kG, GAB multilayer sorption energy; aW, water activity = RH/100:33,34 Meq =

M0 × cG × kG × aw (1 ― kG × aw) (1 + (cG ― 1)kG × aw)

(1)

172

Water-accessible surface area AG m2 g-1 was obtained by: N0, Avogadro constant

173

(6.022 x 1023 molecules mol-1); am, area captured by a single water molecule at 25 °C (1.06

174

x 10-19 m²); Msorb, molar mass of water (18 g mol-1):35 AG =

M0 × N0 × am Msorb

(2)

175 176

Water retention in model soil substrates. In model-scale, the impact of one selected

177

hydrogel (WS3-CA4% – based on WS3 (3.79 mmol –CH2COONa g-1), esterification:

178

4 wt % citric acid, 30 min, 140 °C) on the soils´ water retention was investigated. Two

179

defined model soil substrates were used, representing a sandy and a silty soil

180

(characteristics in SI, Table S1). Dry and sieved (150 - 800 m) gel granules were

181

incorporated homogeneously into the soil substrates at rates of 0.2 or 0.7 wt % (w/w

182

soil). Approx. 10.0 g of each soil/gel mixture were placed in aluminum cylinders

183

(2.8 x 3.0 cm, diameter/height) and saturated with deionized water for 24 hours at RT.

184

Subsequently, the cylinders were placed on a ceramic plate and the pressure was

185

reduced stepwise using a 100-centimeter water column. At each step (0, 10, 30, 40, 50,

ACS Paragon Plus Environment

10

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 45

186

60, 80 and 100 cm, corresponding to a pF value of 0.00, 1.00, 1.48, 1.60, 1.69, 1.78, 1.90

187

and 2.00) the samples were equilibrated for 24 hours, followed by recording their

188

weight. The pF value is an expression of soil water tension (common logarithm of cm

189

water column) describing the soil matrix potential (defined as energy required to move

190

water against adsorptive and capillary forces). Afterwards, the samples were exposed to

191

pF = 2.53 (water column: 338.8 cm) in a pressure chamber, again allowing an

192

equilibration for 24 hours. Finally, pF curves were obtained from average values out of

193

three replications and were compared to results of pristine soil substrates.

194

ACS Paragon Plus Environment

11

Page 13 of 45 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

195

ACS Sustainable Chemistry & Engineering

RESULTS AND DISCUSSION

196

Hydrogel synthesis and gel properties. Figure 1 represents the synthetic pathway

197

comprising carboxymethylation and citric acid crosslinking, and the resulting ATR-IR

198

spectra

199

carboxymethylation introduced a strong signal at 1591 cm-1 (–COO- asymmetric

200

stretching vibrations) into the biomass spectrum. The subsequent crosslinking reaction

201

gave two new peaks of weak nature, attributable to newly formed ester bonds ((C=O):

202

1728 cm-1, (C-O): 1230 cm-1). Besides, a significant alteration of the lignocellulosic

203

matrix – particularly in the course of the carboxymethylation – became evident through

204

a considerable decline of hemicellulose-derived acetyl signals ((C=O): 1733 cm-1; (C-

205

O): 1228 cm-1). Our previous studies on the carboxymethylation of lignocelluloses

206

already showed that the strongly alcoholic-alkaline medium (pH > 12) caused a partial

207

extraction of hemicelluloses and lignin throughout the reaction.30 For instance,

208

mercerization studies in isopropanol-NaOH mixtures showed that a typical

209

carboxymethylation conditions (55 °C, 3 hours) would reduce lignin and hemicellulose

210

contents to 15.1 % or 9.4 % from originally 22.8 % or 27.3 %, respectively. Nevertheless,

211

following the synthetic path, all lignocellulosic constituents may be involved in the

212

modification reactions as they bear accessible OH-groups.

of

the

modified

biomass

(band

ACS Paragon Plus Environment

assignments36).

Accordingly,

12

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 14 of 45

213 214

Figure 1. (A) Schematic representation of the synthetic path, and (B) ATR-IR spectra of

215

(top to bottom) unmodified wheat straw, carboxymethylated straw (WS2) and (c) citric

216

acid-crosslinked WS2 (16 wt % citric acid, 140 °C, 30 min).

217

Synthesis parameters (covering: –CH2COONa content of CM-intermediates, reaction

218

temperature and duration, and crosslinker concentration) of citric acid esterifications

219

were evaluated by introducing three sample-specific values: gel yield (%), swelling ratio

220

(gwater g-1) and gel stiffness (G´, Pa). Exact data for each parameter set are summarized in

221

the SI (Table S2).

ACS Paragon Plus Environment

13

Page 15 of 45 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

222

Strain sweep experiments showed that the strain resistance of the gel network

223

decreased with an increasing content of carboxymethyl groups in the biomass (WS3 >

224

WS2 > WS1) (see SI, Figure S1). This implies a more pronounced network extension

225

through electrical repulsion of charged carboxymethyl groups – a typical observation

226

for polyelectrolyte hydrogels.37 Besides, reaction temperatures above 120 °C and citric

227

acid concentrations of > 2 wt % significantly enhanced the strain resistance of the gels; a

228

further increase in temperature and crosslinker amount, as well as long reaction

229

durations (300 min) rather affected the overall gel stiffness G´ in the same strain range

230

than the strain resistance of the network (SI, Figure S2). The frequency responses

231

revealed a dominant elastic behavior over the entire frequency range, indicating that the

232

polymer networks were sufficiently crosslinked (SI, Figure S3).

233 234

Figure 2. Crosslinking of the carboxymethyl intermediates WS1, WS2 and WS3 –

235

development of gel yield (◼), swelling ratio (○) and gel stiffness (◆, G´) as a function

236

of the carboxymethyl group content.

ACS Paragon Plus Environment

14

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 16 of 45

237

Besides the impact on gel stiffness, carboxymethylation essentially determined yield

238

and swelling capacity of the hydrogels. Gel yield and stiffness decreased with

239

increasing carboxymethyl content, whereas the swelling capacity was significantly

240

enhanced (Figure 2). Following these trends, WS1-CA4% gave a high yield after

241

crosslinking (78 %) and rigid hydrogel particles (G´ = 1558 Pa), with a poor swelling

242

capacity (8 gwater g-1). Swelling increased from WS2-CA4% (38 gwater g-1) to WS3-CA4% (50

243

gwater g-1) with the higher content of ionic groups and corresponding to the increased

244

water-solubility of the carboxymethyl intermediates (WS2, WS3). However, the

245

crosslinking reaction required free and accessible hydroxyl groups. Therefore, yield and

246

gel stiffness decreased remarkably for the gel based on WS3 (45 %, 399 Pa).

247

Nevertheless, the product still appeared as mechanically stable and, thus, as promising

248

candidate for soil tests.

249

The impact of the reaction setting on gel yield, stiffness and swelling was evaluated

250

by esterifying WS2 applying various sets of reaction temperatures, durations and

251

crosslinker quantities (Figure 3). The actual amount of citric acid (Figure 3(A)) was

252

thereby naturally the most decisive factor: products crosslinked with 2 wt % citric acid

253

were of a very soft texture, whereas high crosslinker concentrations (8 or 16 wt %)

254

resulted in brittle gels and poor swelling. As depicted in Figure 3(B) crosslinking was

255

also reinforced through prolonged reaction durations. Considering both swelling and

256

structural stability of the hydrogels, 30 minutes were found to be the ideal curing time.

ACS Paragon Plus Environment

15

Page 17 of 45 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

257

A shorter duration (here: 15 min), however, insufficiently crosslinked the biopolymer

258

chains, resulting in a high proportion of water-soluble fraction. Elevating the reaction

259

temperature from 120 to 160 °C (Figure 3(C)) increased the product yield clearly (48 to

260

69 %), probably owing to an accelerated thermal degradation of citric acid to its reactive

261

anhydride at temperatures above 153 °C.38 The stiffness of these gels, however,

262

remained within almost the same range.

ACS Paragon Plus Environment

16

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 18 of 45

263 264

Figure 3. Esterification of the carboxymethyl intermediate WS2 (2.56 mmol –

265

CH2COONa g-1): gel yield (◼), swelling ratio in deionized water (○) and gel stiffness

ACS Paragon Plus Environment

17

Page 19 of 45 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

266

(◆, G´) as functions of (A) citric acid concentration, (B) reaction duration and (C)

267

temperature.

268

Overall, the ideal reaction setting was found to be: 140 °C, 30 minutes, 4 wt % citric

269

acid. Compared to citric acid gels based on pure carboxymethyl cellulose,27 swelling of

270

the straw gels was considerably lower indicating to remaining lignin-carbohydrate

271

interconnections. Also, gels obtained from lignocelluloses by graft-copolymerization, as

272

in studies of El-Saied et al. (e.g. 2000 and 2016)17,18, had a much higher water-uptake

273

(350 g/g), however, likely originating from the synthetic vinyl-based polymer part.

274

Moreover, it is questionable whether too high swelling is beneficial for the integrity of

275

the soils´ unique pore structure.

276

Gel-water-interactions. Dynamic vapor sorption (DVS) analysis delivers a valuable

277

insight into gel-water-interactions (e.g. accessibility of hydrophilic sites, swelling

278

behavior) by studying the response of an inititally dry sample (xerogel) to humidity

279

changes. In our study, DVS gave answers on two central questions: (i) How did the

280

synthetic path affect the interaction between lignocellulosic matrix and water? and (ii) Do the

281

gels retain their structural integrity upon the contact with water and in a sequenz of sorption-

282

desorption cycles? The later point thus gives vital information for a potential real

283

application, considering complete drying throughout the period of usage. To evaluate

284

both aspects, samples with different carboxymethyl contents and/or different network

285

densities were analyzed.

ACS Paragon Plus Environment

18

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 20 of 45

286

Figure 4 shows adsorption and desorption isotherms of the gels. Throughout Ad1 and

287

Ad2, all samples clearly displayed an increasing Meq with the rising aw. The sigmoid

288

curve shape – having a significant upswing at high aw – corresponds to type II isotherms

289

(Brunauer´s

290

physisorption.33,40 The water uptake at aw ≤ 0.2 is due to monolayer adsorption41, which

291

was fairly low for all samples. This indicates a poor accessibility of hydrophilic sites and

292

weak adsorbent-adsorbate interactions at low vapor pressures.40 Intermediate

293

humidities (0.3 ≤ aw ≤ 0.7) provoked a more convex isotherm shape. In this stage

294

multilayer sorption proceeded after the monolayer was saturated with water41, and the

295

deviation between the samples´ sorption isotherms started to grow remarkably.

classification39)

indicating

unrestricted

monolayer-multilayer

296 297

Figure 4. Sorption isotherms of the first and second adsorption (Ad1/Ad2) and the

298

intermediate desorption (De1) of gels based on WS1 (1.75 mmol –CH2COONa g-1) or

ACS Paragon Plus Environment

19

Page 21 of 45 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

299

WS3 (3.79 mmol –CH2COONa g-1). (crosslinking conditions: 140 °C, 30 min, 4 or 8 wt %

300

citric acid)

301

At high water activities (aw 0.8-0.9), Meq of the straw gels increased rapidly, probably

302

following moisture penetration into the sample particles, which led to swelling and

303

plasticization of the rigid polymer network and a liberation of further hydrophilic

304

sites.41,42 The curve upswing in this stage can be clearly linked to the degree of

305

carboxymethylation (WS3 > WS1), with respect to ascent and final Meq. Furthermore,

306

comparing isotherms of WS3-CA4% and WS3-CA8% reveals that the higher crosslinking

307

density of the latter gel has not reduced water sorption until aw 0.9. On the contrary, the

308

moisture uptake of WS3-CA8% was slightly enhanced, though equlibrium swelling was

309

much lower (WS3-CA4%: 50 g/g, WS3-CA8%:12 g/g). This finding can be likely attributed

310

to the DVS methodology, in which a thermodynamic equlibrium of the samples´

311

swelling process was not accomplished.

312

Deviations between adsorption and desorption processes are characterized by

313

hysteresis loops, implying (irreversable) structural alterations within the sorbents´

314

domains following their interaction with water.40 Sorption hysteresis occurred

315

throughout the entire humidity range for each sample. The extent and exact position of

316

hysteresis maxima within the humidity range, however, obviously depended on

317

individual sample characteristics, coinciding, in particular, with the crosslinking

318

density. Gels crosslinked with 4 wt % citric acid showed a maximum of hysteresis at

ACS Paragon Plus Environment

20

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 22 of 45

319

intermediate aw. WS3-CA8% evidently retained high amounts of moisture particularly at

320

low water activities (aw 0.1-0.4), pointing on an entrapment of water within the gel

321

network. This may be related to the affinity of water to fill pores, which increases with

322

the crosslinking density.42 However, it is further conceivable that citric acid

323

esterification increased the interaction with water molecules due to the introduction of

324

further OH-groups.

325

Deviations between Ad1 and Ad2, including shifted starting points and a reduced

326

moisture-absorbency at aw = 0.9, were most likely induced by structural alterations

327

during the first experiment cycle. Particularly, in the case WS3-CA8%, Ad2 started from

328

a significantly higher moisture level, which again indicates water entrapment.

329

Moreover, the fact that Meq,

330

humidity range implies that water penetrated more easily in Ad2. This effect may be

331

linked to the formation of new sorption sites with the irreversable cleavage of weaker

332

bonds (e.g. hydrogen bonds) during network extension in Ad1.42,43 High contents of

333

carboxymethyl groups would increase this effect as they promote swelling.

334

Furthermore, for WS1-CA4%, the final moisture load (at aw = 0.9) was slightly reduced,

335

which can be attributed to gel shrinkage during desorption entailing an irreversable

336

formation of hydrogen bonds (hornification).44,45 The icreasing carboxylate content from

337

WS1 to WS3, however, obviously mitigated this effect.

Ad2

exceeded Meq,

Ad1

for each sample in the intermediate

ACS Paragon Plus Environment

21

Page 23 of 45 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

338

To interpret gel-moisture interactions, sorption isotherms were fitted to the GAB

339

equation (SI, Table S3) followed by calculation of monolayer capacity and water-

340

accessible surface area (Figure 5). As indicated by the higher M0 and AG for WS3-based

341

gels, carboxymethylation increased the accessibility of sorption sites. The subsequent

342

crosslinking step enhanced the interaction with water, likely owing to the additional

343

introduction of OH-groups. As discussed previously, higher monolayer capacities

344

during De1 imply a hysteresis of sorption domains (i.a. entrapment of moisture).

345

346 347

Figure 5. Monolayer capacity M0 (blue bars/values) and water-accessible specific surface

348

area AG (◻, black values) obtained by GAB model fitting of sorption isotherms of first

349

and second adsorption (Ad1/Ad2) and intermediate desorption (De1). (Gels based on

350

WS1 (1.75 mmol –CH2COONa g-1) or WS3 (3.79 mmol –CH2COONa g-1), crosslinking

351

reaction: 140 °C, 30 min, 4 or 8 wt % citric acid)

ACS Paragon Plus Environment

22

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 24 of 45

352

Although it was shown before that WS3-CA8% retained water at low vapor pressures

353

during De1, the GAB model fit gave significantly decreased M0 and AG for the

354

desorption cycle. Both parameters further decreased in Ad2. The same phenomenon can

355

be observed for WS1-CA4%. In both cases, sorption sites obviously became unavailable

356

for water binding, which might be explained by an irreversable formation of hydrogen

357

bonds during De1. In contrast, WS3-CA4% gave a steady increase in monolayer moisture

358

levels and water-accessible surface area.

359

Overall, the development of M0 and AG underlines the impact of each synthesis step

360

on water-matrix-interactions. Sorption processes probably also correlate with the

361

disintegration of the lignocellulosic matrix and decrystallization of the cellulose fraction

362

accompanying carboxymethylation. The presence of charged carboxylate groups promoted

363

network extension, which accelerated moisture penetration and clearly mitigated the

364

irreversable aggregation of biopolymer chains during drying. This ensures a re-

365

swellability of the gel network, for example, in a real soil application. Particularly, WS3-

366

CA4% showed a promising structural integrity upon the contact with moisture

367

suggesting its further evaluation in a model-scale soil application.

368

Water retention in soil substrates. Using two different model soil substrates, water

369

retention tests were conducted to give a more application-related assessment of the gel

370

performance. Based on promising previous results, the straw gel WS3-CA4% was

371

exemplarily selected. Furthermore, the model soils chosen – classified as a sandy and a

ACS Paragon Plus Environment

23

Page 25 of 45 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

372

silty substrate – provided us information about the practical gel application (covering

373

particle sizes, exchangeable cations, etc. – SI, Table S1). However, in Nature, mixtures of

374

both soil substrates may rather be found. Increasing the storing capacity of plant

375

available water in sandy soils, for instance, may enable agricultural cultivation on

376

formerly unusable areas and could be, therefore, an important contribution to securing

377

food supply. Silty soils, on the other hand, as valuable agricultural grounds, may

378

benefit from an improved water storing capacity, particularly, in view of climate change

379

scenarios.

380

Figure 6 demonstrates how the amount of water retained in the sandy or silty soil and

381

in soil/gel mixtures developed as function of the applied matrix potential (pF). This

382

relationship represents important soil-moisture characteristics, e.g. the plant water

383

availability depending on water content. For both model soils, the incorporation WS3-

384

CA4% significantly increased the water retention throughout the pF range, with a direct

385

correlation to gel incorporation rate and, additionally, soil characteristics. The starting

386

point of the assay (pF = 0) enables the examination of unaffected soil-gel-interactions,

387

without external impairments. At that point (as well as throughout the pF-range), the

388

silty substrate inherently retained more water, owing to its smaller grain sizes and

389

consequently higher volume of medium sized pores compared to the sand (SI, Table

390

S1). However, swelling of the gel was obviously less restricted in the sandy soil,

391

resulting in a more remarkable improvement of the sands´ water retention. The

ACS Paragon Plus Environment

24

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 26 of 45

392

deviation between the water content retained within the different soil/gel mixtures

393

became more evident with the increasing matrix potential and incorporation time. From

394

pF ≥ 1.60 (experiment duration of ≥ 96 hours) the water retention of silt/gel mixtures

395

dropped down, which is especially significant for the higher gel incorporation rate of

396

0.7 wt %. In contrast, curve progressions of gel/sand mixtures at both rates remained

397

almost linear throughout the pF-range. Within the matrix potential range, the region

398

between pF 1.80 and 2.53 is of particularly interest for agronomic applications as a

399

control of irrigation.46 The amount of water in this pF region is thereby defined as field

400

capacity and represents the maximum of plant available water.46 The incorporation of

401

WS3-CA4% thus significantly improved the water retention of the sand at both rates

402

within this specific range (e.g. at pF 2.53: 0.2 or 0.7 wt % gel increased water retention by

403

70 or 300 %, respectively). However, the effect on the silty substrate at pF 2.53 in our

404

lab-scale survey was much smaller, with an increase of circa 12 or 20 %, by

405

incorporating 0.2 or 0.7 wt % dry gel granules, respectively.

406

ACS Paragon Plus Environment

25

Page 27 of 45 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

407 408

Figure 6. Water retention in soil substrates – water retained in a (A) sandy or (B) silty

409

soil or gel/soil mixtures as function of matrix potential (0 ≤ pF ≤ 2.53) and gel

410

incorporation rate (w/w soil). (hydrogel: WS3-CA4%; blue error bars: standard deviation,

411

n = 3)

412

This significant deviation between the two soil substrates thereby most probably

413

corresponds to both the different soil characteristics and the ionic character of the

414

lignocellulosic hydrogel. Cations – particularly of a divalent nature as Ca2+ – released

ACS Paragon Plus Environment

26

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 28 of 45

415

from the soil substrate may have strongly impaired the swelling of the gel particles over

416

time by physically crosslinking the carboxylated polymer chains. This aspect can be

417

demonstrated by a very simple experiment, swelling WS3-CA4% in aqueous saline

418

solutions (SI, Figure S5). Accordingly, the addition of cations to the surrounding

419

medium strongly limited the water absorbency of the hydrogel under equilibrium

420

swelling conditions. Another reason for diminished swelling of WS3-CA4% in the silty

421

soil might have been its fine texture and, therefore, strengthened capillary forces within

422

soil pores. With the increasing matrix potential, the induced capillarity might thus have

423

entailed a drainage of the gel particles and an alignment of the water potential in gel

424

and surrounding soil substrate. Also, the soil pH (sand: 4.07, silt: 6.98) might have

425

affected the swelling properties of the gel. Especially, a soil pH below the pKA of

426

carboxylic groups (4.6)47 would cause, over time, a collapse of the gel network due to

427

the induced ion exchange. Nevertheless, throughout our short-time lab-scale

428

experiments, WS3-CA4% had a higher swelling capacity in the more acidic sandy

429

substrate.

430

According to our experiments, the incorporation of WS3-CA4% showed promising

431

potential for the sand within the examined pF range. However, to thoroughly illuminate

432

the effect of the gel on the water accessibility in soils, it would be essential to extent

433

water retention tests to pF = 4.2. Beyond a matrix potential of 4.2 (permanent wilting

434

point) water would be unavailable for plants.46,48

ACS Paragon Plus Environment

27

Page 29 of 45 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

435

Finally, a crucial point in further studies should be to translate the model-scale into

436

real applications, involving optimized synthetic upscaled procedures and field

437

experiments that offer realistic conditions. Also, the high loss of biomass due to

438

extraction/dissolution throughout the synthetic path (> 60 % for, e.g., WS3-CA4%)

439

demands for further process optimization and/or strategies to recover and valorize lost

440

material fraction. For the actual soil application, the structural integrity of the gel over

441

an adequate time period (e.g. one growth season) and its final biodegradation should be

442

central points of assessment. For the upscale to an economic volume, citric acid

443

crosslinking, as a very simple water-based, heat-induced reaction would allow for

444

continuous processing (e.g. on belt furnaces).

445 446

CONCLUSIONS

447

Herein we presented a simple synthetic approach to fabricate wheat straw-based

448

hydrogels starting from the unfractionated and carboxymethylated lignocellulosic

449

matrix. Citric acid was used as a cheap and non-toxic low-molecular-weight crosslinker.

450

Optimized reaction conditions led to mechanically stable hydrogels with moderate

451

swelling capacities of up to 50 gwater/g. Gel swelling and rheology were thereby

452

essentially determined by both the content of ionic carboxylate groups and the

453

crosslinking density. Moreover, DVS analysis revealed that the straw-based gels re-

454

swell after complete drying, implying a promising durability in terms of a potential soil

ACS Paragon Plus Environment

28

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 45

455

application. Finally, incorporating one selected hydrogel into different model soil

456

substrates showed that already low gel rates (here: 0.2 wt %) may effectively increase

457

the water holding capacity of, especially, sandy soils.

ACS Paragon Plus Environment

29

Page 31 of 45 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

458

ACS Sustainable Chemistry & Engineering

ASSOCIATED CONTENT

459

Supporting Information. A supporting information file (PDF) is available free of

460

charge containing: soil substrates´ characteristics; additional results for gel rheology,

461

crosslinking conditions vs. gel properties, DVS analysis, and swelling experiments.

462 463

AUTHOR INFORMATION

464

Corresponding Author

465

* [email protected]

466

Present Addresses

467

 Department

468

00076 Aalto, Espoo, Finland

469

Author Contributions

470

The manuscript was written through contributions of all authors. All authors have

471

given approval to the final version of the manuscript.

472

Funding Sources

473

The authors gratefully acknowledge the financial support from BASF SE (Germany) and

474

from the Graduate Academy, Technische Universität Dresden (Excellence Initiative of

475

the German federation and the federal states).

of Bioproducts and Biosystems, Aalto University, P.O. Box 16300, FIN-

ACS Paragon Plus Environment

30

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

476

Notes

477

The authors declare no competing financial interests.

478

ACKNOWLEDGMENT

Page 32 of 45

479

The authors would like to thank Dr. Arndt Weiske (Institute of Plant and Wood

480

Chemistry, TU Dresden), Dr. Thomas Klinger and Gisela Ciesielske for ICP-OES

481

measurements and the assistance with water retention tests.

482

ABBREVIATIONS

483

Ad1(2), first (second) adsorption; CA, citric acid; De1, intermediate desorption; DI

484

water, deionized water; GAB, Guggenheim-Anderson-de Boer; ICP-OES, inductively

485

coupled plasma optical emission spectroscopy; IPA, isopropyl alcohol; RT, room

486

temperature; WS, wheat straw

487

REFERENCES

488

(1)

489 490

Tuck, C. O.; Pérez, E.; Horváth, I. T.; Sheldon, R. A.; Poliakoff, M. Valorization of Biomass: Deriving More Value from Waste. Science 2012, 337 (6095), 695–699.

(2)

Laufenberg, G.; Kunz, B.; Nystroem, M. Transformation of Vegetable Waste into Value

491

Added Products: (A) the Upgrading Concept; (B) Practical Implementations. Bioresour.

492

Technol. 2003, 87 (2), 167–198.

493

(3)

Daioglou, V.; Stehfest, E.; Wicke, B.; Faaij, A.; van Vuuren, D. P. Projections of the

494

Availability and Cost of Residues from Agriculture and Forestry. GCB Bioenergy 2016, 8

495

(2), 456–470.

ACS Paragon Plus Environment

31

Page 33 of 45 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

496

ACS Sustainable Chemistry & Engineering

(4)

Mosier, N.; Wyman, C.; Dale, B.; Elander, R.; Lee, Y. Y.; Holtzapple, M.; Ladisch, M.

497

Features of Promising Technologies for Pretreatment of Lignocellulosic Biomass.

498

Bioresour. Technol. 2005, 96 (6), 673–686.

499

(5)

Abdolali, A.; Guo, W. S.; Ngo, H. H.; Chen, S. S.; Nguyen, N. C.; Tung, K. L. Typical

500

Lignocellulosic Wastes and by-Products for Biosorption Process in Water and

501

Wastewater Treatment: A Critical Review. Bioresour. Technol. 2014, 160, 57–66.

502

(6)

503 504

Mohanty, A. K.; Misra, M.; Hinrichsen, G.; others. Biofibers, Biodegradable Polymers and Biocomposites: An Overview. Macromol. Mater. Eng. 2000, 276 (1), 1–24.

(7)

El-Saied, H.; Waley, A. I.; Basta, A. H.; El-Hadi, O. High Water Absorbents from

505

Lignocelluloses. II. Novel Soil Conditioners for Sandy Soil from Lignocellulosic Wastes.

506

Polym.-Plast. Technol. Eng. 2004, 43 (3), 779–795.

507

(8)

Wan Ngah, W. S.; Hanafiah, M. A. K. M. Removal of Heavy Metal Ions from

508

Wastewater by Chemically Modified Plant Wastes as Adsorbents: A Review. Bioresour.

509

Technol. 2008, 99 (10), 3935–3948.

510

(9)

Cao, W.; Dang, Z.; Zhou, X.-Q.; Yi, X.-Y.; Wu, P.-X.; Zhu, N.-W.; Lu, G.-N. Removal

511

of Sulphate from Aqueous Solution Using Modified Rice Straw: Preparation,

512

Characterization and Adsorption Performance. Carbohydr. Polym. 2011, 85 (3), 571–

513

577.

514

(10) Majeed, K.; Jawaid, M.; Hassan, A.; Abu Bakar, A.; Abdul Khalil, H. P. S.; Salema, A.

515

A.; Inuwa, I. Potential Materials for Food Packaging from Nanoclay/Natural Fibres Filled

516

Hybrid Composites. Mater. Des. 2013, 46, 391–410.

ACS Paragon Plus Environment

32

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 34 of 45

517

(11) Sojka, R. E.; Bjorneberg, D. L.; Entry, J. A.; Lentz, R. D.; Orts, W. J. Polyacrylamide in

518

Agriculture and Environmental Land Management. In Advances in Agronomy; Sparks, D.

519

L., Ed.; Academic Press, 2007; Vol. 92, pp 75–162.

520

(12) Hüttermann, A.; Orikiriza, L. J. B.; Agaba, H. Application of Superabsorbent Polymers

521

for Improving the Ecological Chemistry of Degraded or Polluted Lands. CLEAN – Soil

522

Air Water 2009, 37 (7), 517–526.

523

(13) Zohuriaan-Mehr, M. J.; Omidian, H.; Doroudiani, S.; Kabiri, K. Advances in Non-

524

Hygienic Applications of Superabsorbent Hydrogel Materials. J. Mater. Sci. 2010, 45

525

(21), 5711–5735.

526

(14) Wilske, B.; Bai, M.; Lindenstruth, B.; Bach, M.; Rezaie, Z.; Frede, H.-G.; Breuer, L.

527

Biodegradability of a Polyacrylate Superabsorbent in Agricultural Soil. Environ. Sci.

528

Pollut. Res. 2013, 21 (16), 9453–9460.

529 530

(15) Tuomela, M.; Vikman, M.; Hatakka, A.; Itävaara, M. Biodegradation of Lignin in a Compost Environment: A Review. Bioresour. Technol. 2000, 72 (2), 169–183.

531

(16) Martínez, Á. T.; Speranza, M.; Ruiz-Dueñas, F. J.; Ferreira, P.; Camarero, S.; Guillén, F.;

532

Martínez, M. J.; Gutiérrez Suárez, A.; Río Andrade, J. C. del. Biodegradation of

533

Lignocellulosics: Microbial, Chemical, and Enzymatic Aspects of the Fungal Attack of

534

Lignin. 2005.

535

(17) El-Saied, H.; Waley, A. I.; Basta, A. H. High Water Absorbents from Lignocelluloses. I.

536

Effect of Reaction Variables on the Water Absorbency of Polymerized Lignocelluloses.

537

Polym.-Plast. Technol. Eng. 2000, 39 (5), 905–926.

ACS Paragon Plus Environment

33

Page 35 of 45 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

538

(18) El-Saied, H.; El-Hady, O. A.; Basta, A. H.; El-Dewiny, C. Y.; Abo-Sedera, S. A. Bio-

539

Chemical Properties of Sandy Calcareous Soil Treated with Rice Straw-Based Hydrogels.

540

J. Saudi Soc. Agric. Sci. 2016, 15 (2), 188–194.

541

(19) Ibrahim, M. M.; Abd-Eladl, M.; Abou-Baker, N. H. Lignocellulosic Biomass for the

542

Preparation of Cellulose-Based Hydrogel and Its Use for Optimizing Water Resources in

543

Agriculture. J. Appl. Polym. Sci. 2015, 132 (42), n/a-n/a.

544 545

(20) Ren, J.; Kong, W.; Sun, R. Preparation of Sugarcane Bagasse/Poly(Acrylic Acid-CoAcrylamide) Hydrogels and Their Application. BioResources 2014, 9 (2), 3290–3303.

546

(21) Johnson, K. A.; Gorzinski, S. J.; Bodner, K. M.; Campbell, R. A.; Wolf, C. H.; Friedman,

547

M. A.; Mast, R. W. Chronic Toxicity and Oncogenicity Study on Acrylamide

548

Incorporated in the Drinking Water of Fischer 344 Rats. Toxicol. Appl. Pharmacol. 1986,

549

85 (2), 154–168.

550

(22) Greim, H.; Ahlers, J.; Bias, R.; Broecker, B.; Hollander, H.; Gelbke, H.-P.; Jacobi, S.;

551

Klimisch, H.-J.; Mangelsdorf, I.; Mayr, W.; et al. Assessment of Structurally Related

552

Chemicals: Toxicity and Ecotoxicity of Acrylic Acid and Acrylic Acid Alkyl Esters

553

(Acrylates), Methacrylic Acid and Methacrylic Acid Alkyl Esters (Methacrylates).

554

Chemosphere 1995, 31 (2), 2637–2659.

555

(23) Rowland, S. P.; Welch, C. M.; Brannan, M. A. F.; Gallagher, D. M. Introduction of Ester

556

Cross Links into Cotton Cellulose by a Rapid Curing Process. Text. Res. J. 1967, 37 (11),

557

933–941.

ACS Paragon Plus Environment

34

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 36 of 45

558

(24) Welch, C. M. Tetracarboxylic Acids as Formaldehyde-Free Durable Press Finishing

559

Agents Part I: Catalyst, Additive, and Durability Studies1. Text. Res. J. 1988, 58 (8),

560

480–486.

561

(25) Trask-Morrell, B. J.; Andrews, B. A. K. Thermoanalytical Ranking of Catalysts for Use

562

with Polycarboxylic Acids as Durable Press Reactants1. Text. Res. J. 1992, 62 (3), 144–

563

150.

564

(26) Coma, V.; Sebti, I.; Pardon, P.; Pichavant, F. H.; Deschamps, A. Film Properties from

565

Crosslinking of Cellulosic Derivatives with a Polyfunctional Carboxylic Acid.

566

Carbohydr. Polym. 2003, 51 (3), 265–271.

567

(27) Demitri, C.; Del Sole, R.; Scalera, F.; Sannino, A.; Vasapollo, G.; Maffezzoli, A.;

568

Ambrosio, L.; Nicolais, L. Novel Superabsorbent Cellulose-Based Hydrogels Crosslinked

569

with Citric Acid. J. Appl. Polym. Sci. 2008, 110 (4), 2453–2460.

570 571 572 573

(28) Yang, C. Q. FT-IR Spectroscopy Study of the Ester Crosslinking Mechanism of Cotton Cellulose. Text. Res. J. 1991, 61 (8), 433–440. (29) Food

and

Agriculture

Organization

of

the

United

Nations.

FAOSTAT

http://www.fao.org/faostat/en/#data/QC.

574

(30) Heise, K.; Rossberg, C.; Strätz, J.; Bäurich, C.; Brendler, E.; Keller, H.; Fischer, S.

575

Impact of Pre-Treatments on Properties of Lignocelluloses and Their Accessibility for a

576

Subsequent Carboxymethylation. Carbohydr. Polym. 2017, 161, 82–89.

ACS Paragon Plus Environment

35

Page 37 of 45 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

577

ACS Sustainable Chemistry & Engineering

(31) Gutachterausschuss

Forstliche

Analytik.

A:3.3.5

Gesamtaufschluss

578

HNO3/HClO4/HF

579

Bundesministerium für Ernährung und Landwirtschaft, Ed.; Berlin, 2014.

580 581

Mit

Mikrowelle.

In

Handbuch

Forstliche

Mit

Analytik;

(32) Rohm, H.; Ullrich, F.; Schmidt, C.; Löbner, J.; Jaros, D. Gelation of Cross-Linked Casein under Small and Large Shear Strain. J. Texture Stud. 2014, 45 (2), 130–137.

582

(33) Bratasz, Ł.; Kozłowska, A.; Kozłowski, R. Analysis of Water Adsorption by Wood Using

583

the Guggenheim-Anderson-de Boer Equation. Eur. J. Wood Wood Prod. 2011, 70 (4),

584

445–451.

585 586

(34) Timmermann, E. O. Multilayer Sorption Parameters: BET or GAB Values? Colloids Surf. Physicochem. Eng. Asp. 2003, 220 (1–3), 235–260.

587

(35) Passauer, L.; Struch, M.; Schuldt, S.; Appelt, J.; Schneider, Y.; Jaros, D.; Rohm, H.

588

Dynamic Moisture Sorption Characteristics of Xerogels from Water-Swellable

589

Oligo(oxyethylene) Lignin Derivatives. ACS Appl. Mater. Interfaces 2012, 4 (11), 5852–

590

5862.

591 592 593 594 595 596

(36) Socrates, G. Infrared and Raman Characteristic Group Frequencies: Tables and Charts, 3rd ed.; John Wiley & Sons Ltd: Chichester, England, 2004. (37) Omidian, H.; Park, K. Swelling Agents and Devices in Oral Drug Delivery. J. Drug Deliv. Sci. Technol. 2008, 18 (2), 83–93. (38) Barbooti, M. M.; Al-Sammerrai, D. A. Thermal Decomposition of Citric Acid. Thermochim. Acta 1986, 98, 119–126.

ACS Paragon Plus Environment

36

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

597 598

Page 38 of 45

(39) Brunauer, S.; Deming, L. S.; Deming, W. E.; Teller, E. On a Theory of the van Der Waals Adsorption of Gases. J. Am. Chem. Soc. 1940, 62 (7), 1723–1732.

599

(40) Mutungi, C.; Schuldt, S.; Onyango, C.; Schneider, Y.; Jaros, D.; Rohm, H. Dynamic

600

Moisture Sorption Characteristics of Enzyme-Resistant Recrystallized Cassava Starch.

601

Biomacromolecules 2011, 12 (3), 660–671.

602

(41) Li, S.; Tang, J.; Chinachoti, P. Thermodynamics of Starch-Water Systems: An Analysis

603

from Solution-Gel Model on Water Sorption Isotherms. J. Polym. Sci. Part B Polym.

604

Phys. 1996, 34 (15), 2579–2589.

605

(42) Lu, Y.; Pignatello, J. J. Sorption of Apolar Aromatic Compounds to Soil Humic Acid

606

Particles Affected by Aluminum(III) Ion Cross-Linking. J. Environ. Qual. 2004, 33 (4),

607

1314.

608 609 610 611 612 613

(43) Chirkova, J.; Andersons, B.; Andersone, I. Study of the Structure of Wood-Related Biopolymers by Sorption Methods. BioResources 2009, 4 (3), 1044–1057. (44) Weise, U. Hornification : Mechanisms and Terminology. Pap. Ja Puu 1998, 80 (2), 110– 115. (45) Hill, C. A. S.; Norton, A.; Newman, G. The Water Vapor Sorption Behavior of Natural Fibers. J. Appl. Polym. Sci. 2009, 112 (3), 1524–1537.

614

(46) Klute, A.; Cassel, D. K.; Nielsen, D. R. Field Capacity and Available Water Capacity. In

615

SSSA Book Series; Methods of Soil Analysis: Part 1 - Physical and Mineralogical

616

Methods; Soil Science Society of America, American Society of Agronomy, 1986; Vol.

617

5.1, pp 901–926.

ACS Paragon Plus Environment

37

Page 39 of 45 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

618

(47) Yang, S.; Fu, S.; Liu, H.; Zhou, Y.; Li, X. Hydrogel Beads Based on Carboxymethyl

619

Cellulose for Removal Heavy Metal Ions. J. Appl. Polym. Sci. 2011, 119 (2), 1204–1210.

620

(48) Fonteno, W. C.; Bilderback, T. E. Impact of Hydrogel on Physical Properties of

621

Coarse-Structured Horticultural Substrates. J. Am. Soc. Hortic. Sci. 1993, 118 (2),

622

217–222.

623 624

For Table of Contents Use Only

625 626

SYNOPSIS

627

Herein, lignocellulosic biomass wastes – a sustainable polymer source – are used to

628

fabricate hydrogels as miniature water reservoirs for agricultural applications.

ACS Paragon Plus Environment

38

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. (A) Schematic representation of the synthetic path, and (B) ATR-IR spectra of (top to bottom) unmodified wheat straw, carboxymethylated straw (WS2) and (c) citric acid-crosslinked WS2 (16 wt % citric acid, 140 °C, 30 min). 272x208mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 40 of 45

Page 41 of 45 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 2. Crosslinking of the carboxymethyl intermediates WS1, WS2 and WS3 – development of gel yield (◼), swelling ratio (○) and gel stiffness (◆, G´) as a function of the carboxymethyl group content. 160x94mm (300 x 300 DPI)

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 3. Esterification of the carboxymethyl intermediate WS2 (2.56 mmol –CH2COONa g-1): gel yield (◼), swelling ratio in deionized water (○) and gel stiffness (◆, G´) as functions of (A) citric acid concentration, (B) reaction duration and (C) temperature. 160x279mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 42 of 45

Page 43 of 45 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 4. Sorption isotherms of the first and second adsorption (Ad1/Ad2) and the intermediate desorption (De1) of gels based on WS1 (1.75 mmol –CH2COONa g-1) or WS3 (3.79 mmol –CH2COONa g-1). (crosslinking conditions: 140 °C, 30 min, 4 or 8 wt % citric acid) 175x80mm (300 x 300 DPI)

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 5. Monolayer capacity M0 (blue bars/values) and water-accessible specific surface area AG (◻, black values) obtained by GAB model fitting of sorption isotherms of first and second adsorption (Ad1/Ad2) and intermediate desorption (De1). (Gels based on WS1 (1.75 mmol –CH2COONa g-1) or WS3 (3.79 mmol – CH2COONa g-1), crosslinking reaction: 140 °C, 30 min, 4 or 8 wt % citric acid) 119x70mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 44 of 45

Page 45 of 45 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 6. Water retention in soil substrates – water retained in a (A) sandy or (B) silty soil or gel/soil mixtures as function of matrix potential (0 ≤ pF ≤ 2.53) and gel incorporation rate (w/w soil). (hydrogel: WS3-CA4%; blue error bars: standard deviation, n = 3) 144x209mm (300 x 300 DPI)

ACS Paragon Plus Environment