Sulfonic Acid Group Determination in Lignosulfonates by Headspace

Apr 16, 2018 - Available techniques to address sulfonic acid groups in LS can be separated into two groups: indirect methods measuring total sulfur co...
1 downloads 4 Views 580KB Size
Subscriber access provided by UNIV OF DURHAM

Sulfonic Acid Group Determination in Lignosulfonates by Headspace Gas Chromatography Philipp Korntner, Andreas Schedl, Ivan Sumerskii, Thomas Zweckmair, Arnulf Kai Mahler, Thomas Rosenau, and Antje Potthast ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.8b00011 • Publication Date (Web): 16 Apr 2018 Downloaded from http://pubs.acs.org on April 16, 2018

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

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

Page 1 of 24 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

1

Sulfonic Acid Group Determination in

2

Lignosulfonates by Headspace Gas

3

Chromatography

4

Philipp Korntner1, Andreas Schedl1, Ivan Sumerskii1, Thomas Zweckmair1, Arnulf Kai

5

Mahler2, Thomas Rosenau1, Antje Potthast1*

6 7

1

University of Natural Resources and Life Sciences, Department of Chemistry, Division of

8 9

Chemistry of Renewable Resources, Konrad-Lorenz-Str. 24, A-3430 Tulln, Austria 2

SAPPI Europe, Gratkorn Mill, SAPPI Paper Holding, Brucker Str. 21, A-8101 Gratkorn,

10 11

Austria Corresponding author: [email protected]

12 13

Keywords

14

Lignin, technical lignins, lignosulfonates, functional groups, sulfonate, quantification,

15

titration, elemental analysis, gas chromatography

16

Abstract

17

A method for the determination of the sulfonic acid content in lignosulfonates via sulfur

18

dioxide quantification upon treatment with H3PO4 by headspace gas chromatography is 1 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 24

19

presented and compared to two already established alternatives, conductometric titration and

20

elemental analysis. Several lignosulfonates, purified from various industrial sources, were

21

examined by all three methods. Limitations and possible interference by other functional

22

groups, such as carboxylic acids, are discussed and suitable solutions are presented. Results of

23

the novel approach are comparable to those of the established techniques in terms of accuracy

24

and precision. The LOD is 0.88 µmol SO3H and the corresponding LOQ is 3.78 µmol SO3H.

25

At the same time, this method outperforms conductometric titration in terms of a higher

26

sample throughput and a much smaller sample amount needed for analysis. Treatment of the

27

sample with highly concentrated phosphoric acid and simple heating is straightforward

28

enough to render the method a valuable tool in lignosulfonate analysis of industrial problem

29

sets, such as screenings or optimization of process parameters, without compromising

30

analytical accuracy.

31

Introduction

32 33

Lignin, as a unique, renewable resource for aromatics, has moved into the spotlight of

34

industry and bioeconomy. Technical lignin, having a largely altered structure compared to the

35

native parent polymer, is produced in vast amounts as a byproduct in the pulp and paper

36

industry. Therefore, it possesses immense potential to be exploited as a feedstock for

37

chemicals; while currently mostly being used to generate heat because of its high calorific

38

value to ensure independence autarky of pulping mills.1,2,3

39

The general term lignin, especially technical lignin, describes a family of polymers sharing

40

similar structural features. So-called native lignin is mostly the product of a combinatorial-

41

like radical coupling reaction of the three main phenylpropanoid units, the monolignols (p-

42

coumaryl, coniferyl, and sinapyl alcohol). Unlike proteins or other biopolymers, this results in

43

a purely chemically controlled synthesis with an astronomical number of possible lignin 2 ACS Paragon Plus Environment

Page 3 of 24 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

44

isomers. Additionally, the ratio of the starting monolignols is highly dependent on the type of

45

biomass. In fact, lignin shows a high propensity for the incorporation of other structurally

46

related monomers of phenylpropanoid type, sometimes called “metabolic plasticity” of

47

lignin.4,5,6 The structure of the native lignins is significantly altered upon pulping processes,

48

most of which aim at breaking down and/or dissolving the lignin. In terms of molecular

49

structure, these processes leave behind a profoundly modified lignin with high polydispersity

50

and a variety of functional groups, such as the hydroxyl, carboxyl, keto, thiol, and sulfonic

51

acid groups.

52

Functional groups containing sulfur are introduced by both of the two most commercially

53

relevant pulping processes, the kraft and the sulfite processes. The sulfite process introduces

54

sulfonic acid groups, mostly in the benzylic positions to the lignin, forming lignosulfonates

55

(LS); this is one of the driving factors for the dissolution of the lignin in the pulping

56

process.7,8 Today, the analysis of native or unaltered lignins as well as kraft lignins is rather

57

well established, including structural information ranging from wet-chemical analysis5,9,10,11

58

to advanced 2D and 3D nuclear magnetic resonance (NMR) protocols12,13,14,15 or various

59

techniques for hydroxyl group determination. Examples are 1H NMR16 and

60

derivatization,17

61

of these methods can be used without modification for (LS)21,22,23 in the same way as they are

62

used for most native and technical kraft lignins; others are limited by the low solubility of LS

63

in nonpolar organic solvents.

13

31

P NMR after

C NMR18 or gas chromatography/mass spectrometry (GC/MS).19,20 Many

64

Sulfur contents of around 5–6% in LS are higher than those in Kraft lignins, and correspond

65

to about 0.4–0.7 sulfonate groups per phenylpropanoid (C9) unit.24,25,26 Although sulfur may

66

not be desirable for several applications,27 such as work with catalysts that might be poisoned,

67

the sulfonic acid groups in LS provide unique physicochemical properties of a polyelectrolyte.

68

At present, applications of LS include oil well drilling additives, several types of dispersants,

69

emulsion stabilizers, and plasticizers in concrete,28 which all fall into the lower-value 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

Page 4 of 24

70

segment. Most of the applications exploit the surface activity of the polymer, which in turn is

71

connected to the molecular weight and the content of functional groups, sulfonic acid groups

72

in particular. Traditionally, LS was used in the form it was extracted from the spent pulping

73

liquor. However, for replacing higher-performance petro-based chemicals, the tailoring of its

74

physicochemical properties, and thereby the degree of sulfonation, becomes important.3,29,30 It

75

is obvious that appropriate analytical techniques to quantify those active functional groups are

76

needed, especially when considering the rising number of potential LS applications.

77

Available techniques to address sulfonic acid groups in LS can be separated into two

78

groups: indirect methods measuring total sulfur contents, assuming that all sulfur is present in

79

the form of sulfonic acid groups, and those measuring sulfonic acid groups directly.31

80

Examples of indirect determination are X-ray fluorescence spectroscopy,32 combustion/ion

81

chromatography,33,34 and elemental analysis. With regard to direct sulfonic acid

82

determination, well-known examples are the quantitation of retained benzidinium ions by

83

UV-spectroscopy,35 or the direct conductometric titration of the acid form by a strong base.36

84

While most methods have certain advantages for different problem sets, many of them share

85

the disadvantage of tedious sample preparation or the necessity of relatively high sample

86

amounts, limiting their value to industrial sample sets.

87

This study presents a comparison of two established methods for sulfonic acid

88

determination, namely elemental analysis41 and conductometric titration,31,9 with a novel

89

headspace GC/MS method, by means of various LS samples. The two conventional methods

90

provided reference data against which the new method for the determination of sulfonic acid

91

was evaluated. The method aims at higher throughput using smaller sample volumes and

92

easier sample handling.

93 94

Material and Methods

4 ACS Paragon Plus Environment

Page 5 of 24 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

95

Chemicals. All LSs were provided by project partners in the FLIPPR (Future Lignin and

96

Pulp Processing Research) project. All solutions of hydrochloric acid (HCl) and sodium

97

hydroxide (NaOH) for titration were prepared from 0.1M FIXANAL® (Sigma-Aldrich)

98

standard solutions. Solutions of lithium hydroxide (LiOH) were prepared from LiOH

99

monohydrate (p.a.). Ortho-phosphoric acid was of p.a. quality, and 3,4-dihydroxybenzoic acid

100

(protocatechuic acid, DHBA) had a purity of >97% (purum). All other solvents and chemicals

101

were of p.a. quality or higher. All chemicals were purchased from Sigma-Aldrich.

102

Preparation of lignin samples. The LS was prepared by Amberlite® XAD7 resin

103

purification according to Sumerskii et al.25 in order to remove all processing chemicals and

104

low molar mass carbohydrates. In short, a 5 ml polypropylene syringe was filled with 0.5 g of

105

Dowex 50WX8 (H-form) and 1 g of XAD-7 resin. Glass wool was plugged in the tip to

106

prevent the resins from leaking. The final concentration of LS solution to be purified was

107

chosen to be 100–150 mg LS/g resin. The syringe filled with the LS solution was closed and

108

placed on a shaker for 5–6 hours. After shaking to ensure complete adsorption of the LS to the

109

resin, the supernatant was released from the syringe. The LS was desorbed and re-dissolved

110

by repeated washing of the resin with ethanol to a volume of about 15 ml (ensuring complete

111

desorption of the LS). As a last step, the ethanol was evaporated on a rotary evaporator after

112

diluting the solution with water by a factor of approximately two. The volumes will vary to a

113

certain extend for different LS samples. The resulting aqueous solution of the LS was freeze-

114

dried and carefully homogenized with a glass rod.

115

Alternatively, the LS samples were also purified by ultrafiltration of black liquor through a

116

1 kDa membrane and ion exchange into the acid form by Dowex 50WX8 resin. Subsequently,

117

the resulting aqueous solutions were freeze-dried and carefully homogenized with a glass rod.

118

Solid-state nuclear magnetic resonance (NMR) spectroscopy. Solid-state nuclear

119

magnetic resonance (NMR) experiments were performed on a Bruker Avance III HD 400

120

spectrometer (resonance frequency of 1H of 400.13 MHz, and

13

C of 100.61 MHz, 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

121

respectively), equipped with a 4 mm dual broadband cross-polarization magic-angle spinning

122

(CP-MAS) probe.

123

(TOSS) sequence at ambient temperature with a spinning rate of 5 kHz and a CP contact time

124

of 2 ms. Quantitative 13C NMR spectra were recorded according to the multiple CP approach

125

described by Johnson and Schmidt-Rohr38 at a rotational frequency of 12 kHz. A total of ten

126

CP blocks were implemented, each with a contact time of 1 ms and a ramped CP contact 1H

127

(70–100%) followed by a 1 s repolarization delay. For both experiments, a recycle delay of 2

128

s, SPINAL-64 1H decoupling, and an acquisition time of 49 ms were used, with the spectral

129

width set to 250 ppm. Chemical shifts were referenced externally against the carbonyl signal

130

of glycine with δ = 176.03 ppm. The integration limits for carboxylic acids were δ = 160 ppm

131

to δ = 210 ppm. Apodization with an exponential function prior to Fourier transformation

132

(FT) was applied.

13

C NMR spectra were acquired with the total sideband suppression

133

Quantitative 13C NMR (solution state). LS (200 mg) was dissolved in 600 µl of a 0.01 M

134

chromium(III) acetylacetonate (Cr(acac)3) solution in DMSO-d6. The spectra were acquired

135

on a Bruker Avance II 400 instrument (resonance frequency 400.13 MHz for 1H and 100.61

136

MHz for

137

acquisition of 64 k data points and 30.000 scans. The spectral width was set to 240 ppm. An

138

acquisition time of 1.4 s and a relaxation delay of 2 s were used. Prior to Fourier

139

transformation (FT), apodization with an exponential window function (lb = 70 Hz) was done.

140

The integration limits for carboxylic acids were δ = 160 ppm to δ = 210 ppm18. Standard

141

Bruker NMR experiments were used for acquisition. Data processing was performed with

142

Topspin 2.1 (Bruker, Rheinstetten, DE).

13

C), equipped with a 5 mm broadband probe head (BBFO) with a z-gradient,

143

Conductometric titration. Conductometric titration was performed on a Metrohm Titrando

144

device (Metrohm, Herisau, CH), equipped with an 856 conductivity module and an electrode

145

with a 5-ring conductivity cell combined with a PT1000 thermo sensor. 40–80 mg of sample

146

(in protonated form) were dissolved in 40 ml of HQ water and titrated in 0.05 ml steps against 6 ACS Paragon Plus Environment

Page 6 of 24

Page 7 of 24

147

0.1 M LiOH solution (titer determined by 0.1 M HCl standard solution) and subsequently

148

back-titrated with 0.1 M HCl solution. The titration was stopped after 6 ml of base or acid

149

were added to the LS solution. Metrohm tiamo™ software was used for data acquisition.

150

The sulfonic acid groups were determined by straight-line fitting of the linear regions of the

151

titration curves and determination of the points where the lines intersect. The first intersection

152

point of the base titration provided the strong acid content, which was equal to the sulfonic

153

acid content (since no other strong acids were present in the sample). The second intersection

154

point accounted for the weak acids, mainly phenolic hydroxyl groups and, to a lesser extent,

155

carboxylic acids31,9. The curve of the back titration was evaluated in the same way, providing

156

the data in the reversed order for confirmation.9 Fig. 1 provides a graphic illustration of the

157

data evaluation procedure.

Conductivity (mS/cm (20°C))

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

1.6

1.2

0.8 c

0 0.4

b

2

a

4

6

Volume HCL 0.1M, ml a

0.0 0

b

c

2

4

6

Volume LiOH 0.1M, ml

158 159

Figure 1: Conductometric titration of a LS sample with 0.1 M LiOH and the back titration

160

with 0.1 M HCl. Strong acids (sulfonic acids) were calculated from volume a, weak acids

161

(such as carboxylic acids and phenolic hydroxyls) from volume b, and the excess amount of

162

LiOH is shown as volume c.

163

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

Page 8 of 24

164

Elemental analysis. All samples (3 mg per analysis) were thoroughly dried prior to

165

elemental analysis and stored under an inert atmosphere. Elemental analysis was performed as

166

C/H/N/S (oxygen was determined indirectly) analyses on an EA 1108 CHNS-O (Carlo Erba

167

Instruments, CE Elantech, Inc.) elemental analyzer.41

168

Preparation of lignin samples for headspace GC/MS. For the determination of sulfonic

169

acid groups in LS, 5–15 mg of LS was weighed into 10 ml headspace vials. As an internal

170

standard, approximately 5 mg DHBA was added to the sample. Then 3 ml of 85% ortho-

171

phosphoric acid was added. The vials were tightly sealed with crimp caps, shaken, and heated

172

to 110 °C for three hours. In particular the heating of the reaction mixture should be done with

173

suitable safety protection during operation since boiling of the concentrated phosphoric acid

174

(b.p=158°C) can lead to exploding vials and severe spills of hot, concentrated acid.

175

Calibration curve. Aliquots from 5–50 mg of a selected LS were prepared as described

176

above, again with the addition of around 5 mg of DHBA as an internal standard. The selected

177

LS was well characterized beforehand, in terms of sulfonic acid content, by conductometric

178

titration and elemental analysis, ash content, residual carbohydrates, and methoxyl groups. A

179

minimum of 14 calibration points have been measured evenly distributed over the calibration

180

range.

181

Determination of the analytical figures of merit. Estimation of cLOD (Limit of detection)

182

was done according to DIN 32645:2008-1140 by measurement of blank samples (n=8) within

183

a confidence level of 99%. A 3-point calibration with a minimum concentration of 5*LOD

184

was used for calculation. The Limit of Quantification, cLOQ was calculated according to

185

cLOQ=3.3*cLOD

186 187 188 189

Relative standard deviation (RSD) was determined by independent analysis of sample material (the values of n are provided) ). Accuracy was evaluated by comparison to reference methodologies, such as elemental analysis or conductometric titration. 8 ACS Paragon Plus Environment

Page 9 of 24 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

190

Table 1: Lignin standard used for calibration. The number of repeated determinations and the

191

RSDs (%) of the corresponding analysis are shown in brackets if applicable.

Pulping process

Magnefite

Residual non-cellulosic polysaccharide, wt% < 1 (n=2) Ash content, wt% (n=2)