188
  Microwave Assisted Extraction of Xylan By Fathimathul Suhara Panthapulakkal A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Graduate Department of Chemical Engineering and Applied Chemistry University of Toronto © Copyright by Fathimathul Suhara Panthapulakkal (2014)  

By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

 

 

MicrowaveAssistedExtractionofXylan

By

FathimathulSuharaPanthapulakkal

Athesissubmittedinconformitywiththerequirements

forthedegreeofDoctorofPhilosophy

GraduateDepartmentofChemicalEngineeringandAppliedChemistry

UniversityofToronto

©CopyrightbyFathimathulSuharaPanthapulakkal(2014)

 

Page 2: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

ii 

 

AbstractMicrowave Assisted Extraction of Xylan Doctor of Philosophy, 2014 Fathimathul Suhara Panthapulakkal Graduate Department of Chemical Engineering and Applied Chemistry University of Toronto

Xylan is one of the major hemicelluloses present in plant cell wall matrix, where it is closely

associated with other cell wall components, cellulose and lignin. Xylan has enormous potential

as a renewable biopolymer and recently, research in the direction of isolation and utilization of

xylan is gaining lot of research attention. Extraction of xylan from the plant cell walls involves

the hydrolysis of xylan and its transfer from the plant cell wall matrix to the hydrolyzing media.

Current process of extraction involves prolonged heating of the biomass with the hydrolysis

media at high temperature and/or pressure that leads to molecular degradation of xylan and

limits its high potential polymeric applications. In this research, microwave assisted alkaline

extraction of polymeric xylan from birch wood is investigated as an alternative to the time

intensive conventional extraction. The hypothesis to be tested is that the microwave’s selective

heating ability leads to the generation of hot spots through its interaction with the alkali present

in the fibers and the resulting "explosion effect" loosen the recalcitrant fiber structure network

thereby facilitating the hydrolysis of xylan and its dissolution before undergoing significant

degradation. Effect of microwave extraction on the yield of xylan and wood solubilization,

physico-chemical properties of wood fibers and of isolated xylan were investigated in

comparison with conventional extraction. Low power input microwave (110 W) alkaline

extraction was found to be an efficient alternative to the conventional extraction. FTIR and

chemical composition of wood fibers after extraction demonstrated an increased removal of

xylan from the wood fibre using microwave extraction. SEM, X-ray microtomography, and X-

Page 3: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

 

 

iii 

 

ray crystallinity studies of wood fibers demonstrated a porous and loosened fibre structure after

microwave extraction confirming the hypothesis. Molecular weight of the isolated xylan using

microwave extraction was found to be higher compared to the xylan isolated using conventional

extraction indicating less molecular degradation. About 75% of xylan present in birch wood

could be extracted using a low power input microwave heating under optimized extraction

conditions of 8wt% NaOH solution, 1:8 (g:mL) solid to liquid ratio, and 25 minutes of

extraction time.

Page 4: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

 

 

iv 

 

ACKNOWLEDGEMENT

It is my utmost pleasure to thank the many people who made this thesis possible.

I would like to express my sincere gratitude to my advisor Prof. Mohini M Sain for his

continuous guidance, generous support and valuable advice throughout my research.

I am deeply grateful to Prof. Donald Kirk, Department of Chemical Engineering and Applied

Chemistry, University of Toronto, and Professor D. N. Roy, Faculty of Forestry, University of

Toronto for their valuable suggestions, and constructive support throughout this work.

I wish to express my warm and sincere thanks to my fellow students and other department staff

at the Faculty of Forestry and Department of Chemical Engineering and Applied Chemistry,

with a special mention to Mr. S. Law, Dr. S. Kunar, and Dr. O. Faruk for their help and

support during this research.

A special note to my husband Sreekumar and my kids Abhinav, Tejas and Rithik - Thank you

for your constant support and encouragement and with so much of happiness let me say that

this effort is dedicated to you all.

Finally, to my parents, your love and thoughts have been my strength and inspiration and would

always remain.

Thank you all.

Page 5: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

 

 

 

Table of Content  

 

Abstract ......................................................................................................................................................... ii 

Acknowledgements....................................................................................................................................iv 

List of Tables.............................................................................................................................................viii 

List of Figures..............................................................................................................................................x 

Chapter 1  Scientific Background ............................................................................................................ 1 

1.1   Introduction .......................................................................................................................................... 1 

1.2  Wood Hemicelluloses ........................................................................................................................... 4 

1.3.  Xylan   ................................................................................................................................................ 7 

1.3.1  Hardwoodxylan ................................................................................................................... 7 

1.4   Current and potential Applications of xylan ......................................................................................... 9 

1.5   Challenges in the extraction of hemicelluloses .................................................................................. 10 

1.5.1  Interactionofhemicelluloseswithcellulose ...................................................................... 10 

1.5.2  Interactionofhemicellulosewithlignin ............................................................................ 12 

1.6   Isolation options of xylan from wood ................................................................................................. 13 

1.6.1  Acidhydrolysis .................................................................................................................... 14 

1.6.2  Hydrothermaltreatments .................................................................................................. 15 

1.6.3  Alkalineextraction ............................................................................................................ 18 

1.7  Microwave Technology ....................................................................................................................... 22 

1.7.1  Microwaveassistedsolventextraction .............................................................................. 25 

1.7.2  Microwaveassistedextractionofplantmaterials ............................................................ 26 

Chapter 2  Research Hypothesis and Objectives .................................................................................. 30 

2.1   Hypothesis ........................................................................................................................................... 30 

2.2   Research Objectives ............................................................................................................................ 30 

Chapter 3   Materials and Methods ....................................................................................................... 31 

3.1  Materials ............................................................................................................................................. 31 

3.1.1  Birchwoodfibers ................................................................................................................ 31 

3.1.2  Chemicals ............................................................................................................................ 31 

3.2  Methods  .............................................................................................................................................. 32 

3.2.1  Isolationofxylanfrombirch .............................................................................................. 32 

3.2.2  CharacterizationofBirchwood ......................................................................................... 35 

3.2.3  Analysisofliquidextracts .................................................................................................. 42 

Page 6: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

 

 

vi 

 

3.2.4  Characterizationofxylan .................................................................................................. 43 

Chapter 4  Evaluation of microwave assisted alkaline extraction of birch xylan .................................. 46 

4.1.  Materials and Methods ....................................................................................................................... 46 

4.1.1  Birchwoodfibers ................................................................................................................ 46 

4.1.2  Extractionofxylanfrombirch ........................................................................................... 48 

4.2  Results and Discussion ........................................................................................................................ 49 

4.2.1  Effectofsodiumhydroxideconcentrationontheextractionofxylan .............................. 49 

4.2.2   Effectofirradiationtimeandmicrowavepowerontheextractionofxylan ................... 51 

4.2.3   MaterialBalance ................................................................................................................ 60 

4.2.4   Characterization of wood: Crystallinity study using X‐ray diffraction ................................. 62 

4.2.5   Characterization of the precipitated xylan .......................................................................... 63 

4.3  Conclusions ......................................................................................................................................... 68 

Chapter 5  Investigation of the mechanism of microwave assisted alkaline extraction of birch wood69 

5.1  Introduction ........................................................................................................................................ 69 

5.2  Methods  .............................................................................................................................................. 69 

5.3  Results and Discussion ........................................................................................................................ 72 

5.3.1  Comparisonofmicrowaveandconventionalalkalineextraction:Woodsolubilization 72 

5.3.2   Comparisonofmicrowaveandconventionalalkalineextraction:Yieldofxylan ............ 81 

5.3.3   Comparisonofmicrowaveandconventionalalkalineextraction:Physico‐chemicalstructuralanalysisofwood ................................................................................................................ 89 

5.3.4  XylanCharacterization .................................................................................................... 102 

5.4  Conclusions ....................................................................................................................................... 106 

Chapter 6  Investigation of structural changes of alkaline extracted wood using X‐ray 

microtomograhy: A comparison  of microwave versus conventional method of extraction ........... 107 

6.1  Introduction ...................................................................................................................................... 107 

6.2  Material  ............................................................................................................................................ 109 

6.3  X‐ray computed microtomography image processing ...................................................................... 110 

6.4  Results and Discussion ...................................................................................................................... 111 

6.4.1  Imageresolutionanalysis ................................................................................................ 113 

6.4.2  Representativevolumeofinterest ................................................................................... 115 

6.4.3  Comparisonofconventionalextractedandmicrowaveextractedwoodanatomy ...... 116 

6.5  Conclusions ....................................................................................................................................... 122 

Page 7: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

 

 

vii 

 

Chapter 7  Statistical optimization of microwave assisted alkaline extraction of xylan from birch 

wood using response surface methodology ............................................................................................. 123 

7.1  Introduction ...................................................................................................................................... 123 

7.2  Materials ........................................................................................................................................... 125 

7.3  Microwave assisted extraction ......................................................................................................... 125 

7.4  Experimental designing using CCD .................................................................................................... 126 

7.4.1   Statisticalanalysisandthemodelevaluation ................................................................. 128 

7.4.2   Optimizationoftheprocessingvariables ........................................................................ 130 

7.5  Results and Discussion ...................................................................................................................... 130 

7.5.1  Effectofextractionvariablesontemperatureofthewoodslurry ................................. 130 

7.5.2   Effectofextractionvariablesonwooddissolution ......................................................... 136 

7.5.3  Effectofextractionvariablesonyieldofxylan................................................................ 140 

7.5.4   Optimizationofmicrowaveassistedextractionofxylanandvalidationofthemodel . 143 

7.6  Conclusions ....................................................................................................................................... 146 

Chapter 8  Summary and Conclusions ................................................................................................ 147 

Chapter 9  Recommendations for Future Research ........................................................................... 149 

References ................................................................................................................................................. 151 

Page 8: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

 

 

viii 

 

Listoftables

Table 1.1 Cellulose, hemicellulose and lignin content of woody biomass compiled from

different authors

Table 1.2 Wood hemicellulose characteristics

Table 1.3 Available literature on microwave assisted extraction of polysaccharides from

various biomass

Table 4.1 Chemical composition of extractive-free birch wood fibers

Table 4.2 Experimental conditions used for microwave extraction

Table 4.3 Material balance analysis of wood fibers after microwave assisted extraction

Table 4.4 Chemical composition of xylan extracted using different experimental conditions

Table 4.5 Molecular mass distribution of xylan extracted using different experimental

conditions

Table 5.1 Experimental conditions used for microwave and conventional extraction

Table 5.2 Time-temperature combinations used in the microwave and conventional

extractions

Table 5.3 Temperature of the slurry after different duration of microwave extractions

Table 5.4 Mass balance analysis of wood fibers after xylan extraction

Table 5.5 Effect of alkaline extraction on the chemical composition of birch wood fibers

Table 5.6 Peak assignment of wood fibers before and after different extraction processes

Table 5.7 Crystallinity index of wood fibers before and after different extraction processes

Table 5.8 Chemical composition of xylan isolated under different extraction conditions

Table 5.9 Intrinsic viscosity, viscosity average molecular weight and number average

degree of polymerization of xylan obtained by microwave convetnional extraction

Table 6.1 Experimental conditions used for extraction

Table 6.2 Porosity of wood samples after different extraction methods (sample size > 2 mm)

Table 6.3 Calculated porosities of wood samples after sodium hydroxide extraction using

conventional heating and microwave irradiation process ( sample size < 2 mm)

Page 9: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

 

 

ix 

 

Table 6.4 Comparison of the calculated porosities of wood samples after 10 minutes of

sodium hydroxide extraction using conventional heating and microwave

irradiation process ( sample size < 2 mm)

Table 7.1 Independent variables studied in the CCD with their coded and uncoded levels

Table 7.2 DOE design matrix and the results

Table 7.3 Polynomial equations for the quadratic model and the regression coefficients

Table 7.4 Analysis of variance (ANOVA) for the RSM model

Table 7.5 Experimental and predicted values of wood dissolution, yield of xylan, and

temperature at the optimum extraction conditions used for the alkaline extraction

of xylan

Page 10: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

 

 

 

ListofFigures

Figure 1.1 Typical structures of hemicelluloses in wood

Figure 1.2 Structure of hardwood xylan

Figure 1.3 Polysaccharide network in lignocellulosic matrix

Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix

Figure 1.5 Representation of peeling and stopping reactions of polysaccharides

Figure 1.6 Schematic of a sinusoidal electric field to an ideal dielectric and the out-of-phase

displacement current which is induced

Figure 1.7 (a) Phase diagram for an ideal dielectric where the energy is transmitted without

loss; (b) phase diagram where there is a phase displacement and (c) the

relationship between ’ and ’’

Figure 3.1 Protocol for the xylan extraction and characterization

Figure 3.2 Experimental protocol for characterization of wood

Figure 3.3 HPLC calibration graphs for sugars

Figure 3.4 Calibration graphs for total sugar content determination

Figure 3.5 Calibration graph for molar mass determination

Figure 4.1. Effect of NaOH concentration on the wood dissolution and xylan yield (Power

level 110W)

Figure 4.2. Mechanism of alkaline hydrolysis and dissolution of hemicelluloses

Figure 4.3. Effect of irradiation time and microwave power on the solubilization of wood (a)

110W (b) 330W, and 550 W (c) 770W and 1100W

Figure 4.4. Temperature of the wood fiber slurry after different microwave irradiation time

Figure 4.5. Effect of irradiation time and microwave power on the yield of xylan (a) 110W

(b) 330W and 550 W (c) 770W and 1100W

Figure 4.6. Effect of microwave power and irradiation time on the dissolved sugar content of

the liquid phase after the precipitation of xylan

Figure 4.7. Effect of microwave power and irradiation time on dissolution of dissolved lignin

at 110 W, 330 W, 550 W, 770 W, and 1100 W

Page 11: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

 

 

xi 

 

Figure 4.8. X-ray crystallographs of wood fiber before and after different duration of

microwave extraction

Figure 4.9. FTIR spectra of xylans obtained at different extraction conditions

Figure 4.10. Typical SEC signal for xylan obtained by microwave assisted method of

extraction

Figure 5.1 Percentage of wood solubilized and yield of xylan after microwave and

conventional extraction

Figure 5.2 Comparison of the amount of wood solubilized during microwave and

conventional extraction

Figure 5. 3 SEM photomicrographs of wood fiber before and after extraction

Figure 5. 4 Effect of microwave and conventional extraction of xylan on the solubilization of

wood

Figure 5.5 Comparison of the yield of xylan (based on the total xylan) during microwave and

conventional extraction

Figure 5.6 Sugar content of the supernatant after precipitating the extracted xylan using two

different processes

Figure 5.7 Effect of microwave and conventional extraction of xylan on the yield of xylan

Figure 5.8 Effect of temperature on the yield of xylan

Figure 5.9 Effect of microwave and conventional extraction of xylan on the chemical

composition of wood

Figure 5.10 FTIR spectra of wood fibers before and after microwave and conventional

extraction

Figure 5.11 XRD crystallographs of birch wood fibers before and after microwave and

conventional extraction

Figure 5.12 SEM photomicrographs of birch wood fibers before and after microwave and

conventional extraction

Figure 5.13 FTIR spectra of xylan isolated using microwave and conventional extractions at

two different experimental conditions

Figure 5.14 ηsp/C vs. Concentration of xylan solutions in CED

Figure 6.1 Binary images of birch wood and the 3D image of the vessels extracted from the

binary images

Page 12: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

 

 

xii 

 

Figure 6.2 Typical X-ray microtomographic binary images of birch wood obtained at

different resolution

Figure 6.3 Average porosity of wood samples using different resolutions

Figure 6.4 Average porosity of wood samples using different volume sizes

Figure 6.5 Typical 3D images of wood chips after different extraction conditions (sample

size >2mm)

Figure 6.6 Typical 3D tomographic images of wood samples, samples after conventional and

microwave assisted extraction process (sample size < 2 mm)

Figure 6.7 Typical 3D images of wood chips after 10 minutes of extraction at 100oC, 120oC,

and 140oC

Figure 7.1 Actual temperature vs. predicted temperature

Figure 7.2 Response surface plots showing the interaction between the variables affecting the

temperature of wood slurry

Figure 7.3 Actual wood dissolution vs. predicted wood dissolution

Figure 7.4 Response surface plots showing the interaction between the variables affecting the

wood dissolution

Figure 7.5 Actual yield of xylan vs. predicted yield of xylan

Figure 7.6 Response surface plots showing the interaction between the variables affecting the

yield of xylan

Figure 7.7 ηsp/C vs. Concentration of xylan obtained at the optimal conditions of microwave

extraction

Page 13: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

 

Chapter1 ScientificBackground 

1.1 Introduction

During the last decade, an increased interest has been observed in the research of biopolymers

from renewable sources. The main driving forces behind the research activities in the area of

polysaccharides, including hemicelluloses are (i) increased awareness of the future shortage of

natural energy sources, (ii) replacement of petroleum-based products as a solution of worldwide

environmental problems, and (iii) demands for healthy food and alternative medicines. In the

past, research activities in the field of hemicellulose were designed mainly for utilizing plant

biomass by conversion into sugars, chemicals, fuel, and as sources of heat energy. Nevertheless,

these polysaccharides can also be used as biopolymers in their native or modified forms

(Ebringerova et al., 1994; Gabrielii et al., 2000; Sun et al., 1999; Jain et al., 2000, Lindblad et al.,

2005). Structural varieties and diversity of these biopolymers provides enormous application

potentials in the fields of food, medical, and industrial applications.

Forest and agricultural biomass are the inexpensive resources for hemicellulosic polysaccharides.

Canadian forests sector is the world’s second largest supplier of woody biomass, behind United

States and as such, an annual supply of more than 200 million m3 of biomass through

commercial operations is reported (FAO, 2003; NRCAN, 2003). Pulp and paper industry is the

major end user of the forest biomass and has been the major contributor to the North American

economy for many years. Recently, the pulp and paper sector has been experiencing a down turn

due to simultaneous impact factors such as immense worldwide competition from paper

Page 14: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

 

 

industries producing fibers from fast growing species, the high cost of the energy, and the

reduced demand for the news print and pulp (Helmerius et al., 2010). The pulp and paper

industry needs to identify additional economic strategies in the coming year to revitalize the

forest sector and to strengthen its competitiveness in the current global market. Authors

including Mabee at al. (2005), Towers et al. (2007), Thorpe (2005), Carvalheiro (2008), and Mao

et al. (2008 ), have explored the concept of integrated forest biorefinery as an opportunity for the

forest product industry to increase revenues and improve environmental sustainability. In this

concept, all the components of biomass can be fractionated to utilize in a most profitable manner

to make high value chemicals, fuels, materials, heat, and power in addition to the traditional core

products. (van Heiningen, 2006, Huang et al., 2008).

The woody biomass is a highly complex material with three major components such as cellulose,

hemicellulose and lignin in different proportions along with minor amounts of extractives and

ash. In general, about 70 % of all wood is polysaccharides (cellulose and hemicelluloses)

(Sjostrom, 1993) and the composition of these components vary species to species, age of wood

and ecological factors. Compositions of different wood species are summarized in the Table 1.1.

Current practice in the pulp and paper industry to fractionate the woody biomass into cellulose

fibers is a “destructive” technology, where other components such as lignin and hemicelluloses

undergo degradation during the pulping process. Majority of hemicelluloses (approximately

20% of the dry weight of wood) are degraded during the chemical pulping process to simple

sugars or isosacharinic acid along with lignin (Casebier and Hamilton, 1965; Kleppe, 1970;

Roberts and El-Karim, 1983; Gustavsson and Al-Dajani, 2000). The sugars and lignin extracted

Page 15: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

 

 

Table 1.1. Cellulose, hemicellulose and lignin content of woody biomass compiled from different authors (Sjostrom, 1993; Fengel and Wegener, 1984, and Garrote et al., 1999)

Raw material Species Cellulose Hemicellulose Lignin Hardwoods Red maple Sugar Maple Trembling aspen Common beech Silver birch Paper Birch Blue gum Softwoods Balsam fir Douglas fir Eastern hemlock Monterey pine Norway Spruce White spruce

Acer rubrum Acer saccharum Populus tremuloides Fagus sylvatica Betula pendula Betula papyrifera Eucalyptus globulus Abies balsamea Pseudotsuga menziesii Tsuga Canadensis Pinus radiate Picea abies Picea glauca

42 40.7 49.4 39.4 41.0 39.4 51.3 38.8 38.8 37.7 37.4 41.7 39.5

25.2 27.3 26.6 29.1 29.8 31.1 21.3 25.8 22.9 23.1 28.9 24.9 27.9

25.4 25.2 18.1 24.8 22 21.4 21.9 29.1 29.3 30.5 27.2 27.4 27.5

during kraft pulping are subsequently concentrated and incinerated to recover the heat energy

(Chakar and Ragauskas, 2004). Since the heating value of wood carbohydrates (about

13.6MJ/kg) is only about half that of lignin (heating value of lignin is 25MJ/kg), the extracted

hemicelluloses provide only ~25% of energy resources for a recovery furnace compared to the

combustion of lignin (van Heiningen, 2006; Tunc and Heiningen, 2009; Yoon and Heiningen,

2010). In general, hemicelluloses in the woody biomass is currently an underutilized renewable

resource that has potential utility for the production of bio-fuels, chemicals, and polymeric

materials (Liu et al., 2006; Amidon and Liu, 2006, Stipanovic et al., 2006; Peng et al., 2007).

Therefore, a more economical use of the hemicelluloses would be to extract them prior to

Page 16: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

 

 

pulping, and then convert them to higher value-added products. The advantages of pre-extraction

of hemicelluloses before pulping was reported by Ragauskas et al., (2006); Mao et al., (2008);

Al-Dajani and Tschirner, (2008); and Yoon and Heiningen, (2008). It was found that the pre-

extraction of hemicelluloses before pulping could substantially improve pulp mill operations by

reducing the cooking times, and improving the pulp production capacity for kraft pulp mills that

are recovery-furnace limited.

1.2 WoodHemicelluloses

The term hemicellulose was originally proposed to denote a substance somewhat similar in

character to cellulose and can easily extract, in comparison to cellulose, from plants by aqueous

alkaline solutions (Dwayer, 1923; Aspinal, 1959). These are the second most abundant

polysaccharides, after cellulose, and approximately 25-35% by weight (Aspinal, 1959) of wood

belongs to this group. However, unlike cellulose (linear homo polymer with a degree of

polymerization of 305-15300), hemicelluloses are branched heteropolysaccharides having β-

(1→4)-linked backbones with an equatorial configuration made up of pentoses such as xylose,

and arabinose, hexoses such as mannose, glucose, and galactose, and/ or sugar acids such as -

D-glucuronic, - D- 4-O-methylgalacturonic and - D-galacturonic acids (Timell, 1964).

Hemicelluloses are non-crystalline, with a low molecular weight range of 10 - 26 kDa and a

relatively low DP of 80 – 200 (Glaudemans and Timell, 1958; Goring and Timell, 1960;

Koshijima et al., 1965).

Xylans and glucomannans are the two major hemicelluloses found in woody biomass.

Glucuronoxylans (about 15-30w/wt%) are the main hemicelluloses of hardwoods, with small

amount of glucomannans (about 2-5w/w%), whereas acetylgalactoglucomannan, (AcGGM;

Page 17: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

 

 

approximately 20 w/w %) and arabinoglucuronoxylan, (approximately 5-10 w/w %) are the

major hemicelluloses of soft wood (Sjostrom, 1993). Typical structures and composition of

hemicelluloses of both hard wood and soft wood are shown in Figure (1.1) and Table 1.2

respectively.

Table 1.2. Wood hemicellulose characteristics (Timell and Syracuse, 1967) Wood type

Hemicellulose Percent of wood

Composition Parts Linkages Mol.wt. (Mn)

Mol.wt. (Mw)

Hard wood

O-acetyl-4-O-methylglucurono xylan Glucomannan

10-35 3-5

β-D-Xylp 4-O-Me-α-D-GlupA O-acetyl β-D-Manp β-D-Glup

10 1 7 1-2 1

1-4 1-2 1-4 1-4

200 >70

180-250 >120

Soft wood

Arabino-4-O-methylglucurono xylan Galacto glucomannan (water-soluble) Galacto glucomannan (alkali-soluble)

10-15 5-10 10-15

β-D-Xylp 4-O-Me-α-D-GlupA L-Araf β-D-Manp β-D-Glup α-D-Galp O-acetyl β-D-Manp β-D-Glup α-D-Galp O-acetyl

10 2 1.3 3 1 1 0.24 3 1 0.1 0.24

1-4 1-2 1-3 1-4 1-4 1-6 1-4 1-4 1-6

>120 >100 >100

>150 >150

Xylp- xylopyranose; Glup- glucopyranose; Manp- mannopyranose; Galp- galactopyranose; Arf- arabinofuranose; GlupA- glucopyranosyl

acid(glucuronic acid)

Page 18: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

 

 

Figure 1.1. Typical structures of hemicelluloses in wood: (a) 4-O-methy-D-Glucuronoxylan (b) D-galacto-D-mannan; (c) D-gluco-D-mannan; (d)(L-arabino)-4-O-methyl-D-glucurono-D-xylan

(a)

(b)

(c)

(d) 

Page 19: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

 

 

1.3. Xylan

Xylans are linear or branched polymers comprised of β-1,4 - linked D-xylopyranosyl residues as

the linear backbone with various substituents depending on the origin and the method of

extraction (Aspinall, 1959; Hirst, 1962; Timell, 1964). The substituents include acetyl,

arabinosyl, and glucuronosyl (Glucuronic acid and 4-O-methyl glucuronic acid) residues. It is the

principal component of the hemicelluloses in many plants, and it contributes about 20–35% to

the total dry matter of the solids (Whistler, 1950; Wilkie, 1979). 4-O-methylglucuronoxylans are

the major xylans present in the lignified cell walls of dicotelydons, whereas

glucuronoarabinoxylans with a low degree of substitution of glycosyl residues are the major

xylans in lignified cell walls of Poacease (Vogel, 2008). In some tissues of cereals and grasses

even up to 50% of the biomass accounted for the xylans (Ebringerova, 2005). These complex

polysaccharides are the second most abundant polysaccharides in nature. The xylans present in

agricultural biomass such as maize, rice, wheat, corn stover, and oats are more complex and

diverse in their structure. The xylan backbone can be heavily branched with acetyl, 4-O-methyl-

GlcpA, GlcpA, Xylp, Araf, and Galp groups (Ebringerova, 2005).

1.3.1Hardwoodxylan

Glucuronoxylans, (O-acetyl-4-O-methylglucuronoxylan), are the main hemicellulose of

hardwoods, such as aspen, birch, and beech accounting up to 15-30% of their dry mass (Timell,

1967; Alen, 2000) and consists of β -D-xylopyranosyl units (xylp) linked though β-1,4-

glycosidic linkages (Timell, 1967). Most of the glucuronoxylans have single 4-O-methyl-α-d-

glucopyranosyl uronic acid residues (MeGlcA) attached by -(1-2) linkages of the main chain

Xylp units (Figure 1.2). However, the glucuronic acid side chain may be present in both the 4-O-

Page 20: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

 

 

methylated and non-methylated forms (GlcA). The ratio of uronic acid residues to xylopyranosyl

units in hard woods (xyl:MeGlcA) varies from 4 :1 to 16:1 depending on the extraction

conditions used; on average, the ratio is about 10 : 1. (Timell, 1964; Koshijima et al., 1965).

Figure 1.2. Structure of hardwood xylan

An unusual methyl glucuronoxylan (MGX) was isolated from the wood of Eucalyptus globulus

(Shatalov et al., 1999; Evtuguin et al., 2003) and it contains α-D-galactose substitution at O-2 in

addition to terminal methyl glucuronic acid (MeGlcA) residues on the xylp residues. In the

native state, the xylan is supposed to be O-acetylated usually at C2 and C3 positions of xylp

residues (Timell, 1967; Teleman et al. 2000 ; 2002). Teleman et al. (2000) reported the degree of

substitution of native aspen xylan as 0.6-0.7. The percentage of acetyl groups of MGX isolated

from hardwoods varies between 3% and 17% of total xylan, corresponding an average of 3.5-7

acetyl groups per 10 xylose units (Alen, 2000). The 4-O-methylglucuranic side groups are more

resistant to acids than the xylp and acetyl groups. The acetyl groups are split during the alkaline

extraction conditions resulting in partial or full water-insolubility of the xylan preparations. But

the acetyl groups may be, at least in part, preserved by treating with hot water or steam. In

addition to these main structural units, glucuronoxylan may also contain minor amounts of

Page 21: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

 

 

galacturonic acid and L-rhamnose, which increases the resistance of the polymer to alkaline

agents. The average degree of polymerization of this polysaccharide is in the range of 100-200

(Goring and Timell, 1960; Timell, 1960; Koshijima et al., 1965)

1.4 CurrentandpotentialApplicationsofxylan

Some of the important applications of xylans were explored and include their use as cholesterol

depressant (Scheller and Ulvskov, 2010), tablet disintegrant (Juslin and Paronen, 1984), dietary

fiber (Barnett et al., 1989), and reagents for chiral separations (Okamoto et al., 1984). Promising

results of xylan were obtained in the field of papermaking, baking, and food additives and are

well documented in several review papers (Ebringerova and Heinz, 2000; Ebringerova, 2005;

Hansen and Placket, 2008; Sedlmeyer, 2011). Xylooligomers are currently used as sweeteners in

food additives as they have beneficial effects on nutrition and health care (Garrote et al., 1999,

Crittenden and Playne, 1996; Vazquez et al., 2000). Hemicelluloses are reported as potential

resources of pharmacologically active polysaccharides; glucuronic acid containing acidic xylans

have been reported to markedly inhibit the growth of sarcoma-180 and other tumors (Sun, 2008).

The unique advantages of these polysaccharides (biocompatibility, non-toxicity and

biodegradability) have explored in areas such as drug delivery, wound closures, surgical

implantation, encapsulation, and cellular therapy (Chen et al., 1995; Van de Velde and Keikens,

2002; Ebringerova, 2005; Coviello et al., 2007). Barrier properties of hemicellulosic

polysaccharides, their application potential in the field of films and coatings in the packaging

films and coatings for foodstuffs, are the topic of a recent review by Hansen and Placket (2008).

Xylans, obtained from wood or cereal straw, have been tested as gel-forming or thermoplastic

materials (Rajesh et al., 2001), as fillers for polypropylene (Amash and Zugenmaier, 1998), as a

component for paint formulation (Fang et al., 2001), and as a coating for cellulosic fibers

Page 22: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

10 

 

 

(Henricksson and Gatenholm, 2001). Micro- and nanoparticles were prepared by a coacervation

method from xylan isolated from corncobs (Garcia et al., 2001) with the aim of application in

drug delivery systems.

1.5 Challengesintheextractionofhemicelluloses

The pre-extraction of hemicelluloses from wood prior to kraft pulping has been extensively

studied and practiced since the early 1930s for the production of dissolving grades of pulps

(Lonnberg, 2005). However, the increased renewed interest in the isolation of hemicelluloses

from biomass to develop value added products is observed recently. This is due to the increased

awareness of the limited petroleum resources and to develop an environmentally sustainable

platform for fuel, chemicals and materials. The challenges involved in the extraction of

hemicellulose are the recalcitrance of biomass due to its close association with the other

components of cell walls such as cellulose and lignin through physical and chemical bonds as

described in the following section.

1.5.1 Interactionofhemicelluloseswithcellulose

Five concepts were reported to explain the network formation of hemicelluloses with cellulose

microfibrils and are as follows (Cosgrove, 2005) (see Figure 1.3):

Page 23: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

11 

 

 

Figure 1.3. Polysaccharide network in lignocellulosic matrix (Source: Cosgrove, Nature

Reviews, 2005)

(i) Hemi polysaccharides spontaneously bind to the surfaces of cellulose microfibrils and

tether adjacent microfibrils together (Hayashi, 1989; Fry, 1989)

(ii) Polysaccharides (xyloglucan) might become entrapped during formation of the

ordered microfibril (b in the figure) and the untrapped remainder of the

polysaccharides would be free to bind to other cellulose surfaces or to other matrix

polymers, thereby anchoring the microfibril firmly to its neighbors (Baba et al., 1994;

Hayashi et al., 1994; Brett and Waldron,1996).

(iii) Cellulose microfibrils might be simply coated with hemicelluloses (xyloglucan) that

adhere to other matrix polysaccharides, without direct linkage between microfibrils,

(Tabolt and Ray, 1992)

(iv) Hemicelluloses (blue strands) might be covalently attached to pectin polysaccharides

(red strands), forming a macromolecule that anchors the microfibrils by sticking of

polysaccharides to cellulose surfaces (Thompson and Fry, 2000)

Page 24: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

12 

 

 

(v) Polysaccharides (arabinoxylans) might bind cellulose and be cross linked by ferulic

acid esters (A-F-F-A) (e in the figure). This type of phenolic crosslink might also

crosslink other hemicelluloses and pectins, particularly in grass cell walls (Zykwinska

et al., 2005)

1.5.2 Interactionofhemicellulosewithlignin

Lignin in the plant cell wall is chemically bonded to carbohydrate materials present in the cell

wall in addition to the lignin networks and such bonds are referred to as L-C (lignin-

carbohydrate) bonds (Sjostrom, 1993). Figure 1.4 shows the four types of lignin carbohydrate

bonds which have been reported in the literature: benzyl ethers, benzyl esters, phenyl glycosides

and acetal bonds (Freudenberg and Grion, 1959; Freudenberg and Harkin, 1960; Kosikova et al.,

1972; Yuka et al., 1976; Eriksson and Lindgren, 1977; Kosikova et al., 1979; Eriksson et al.,

1980; Koshijima et al., 1984; Joseleau and Kesaraoui, 1986; Watanabe and Koshijima, 1988;

Lam et al., 2001). Ether type linkages between lignin and carbohydrates are more common and

stable while the ester linkages are easily cleaved by alkali (Sjostrom, 1993). Benzyl ester bonds

are possible through an uronic acid group on a hemicellulose and a hydroxyl group on lignin, and

benzyl ether and glycosidic bonds when the hemicellulose moiety is bonded through an oxygen

atom to an aliphatic or aromatic carbon atom in lignin (Panshin and de Zeeuw, 1980). Lignin is

difficult to isolate from wood and is usually modified or degraded during the extraction process.

Separation of hemicelluloses from the lignin network and cellulose microfibril network is

difficult to achieve. Therefore, extracting pure hemicelluloses in this complex scenario is quite

challenging.

 

Page 25: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

13 

 

 

 

Figure 1.4. Suggested lignin-carbohydrate bonds in lignocellulosic matrix

1.6 Isolationoptionsofxylanfromwood

Isolation of hemicelluloses from wood typically involves the hydrolysis of the covalent bonds

(ester and ether linkages) to liberate them from the lignocellulosic matrix followed by extracting

or dissolution into the extraction media (Sjostrom, 1993). A number of methods are used to

isolate hemicelluloses from plant biomass including extraction with alkali, steam or water, and

acid hydrolysis (Lindbald and Albertsson, 2005). The composition of extracted hemicelluloses

highly dependent on the method of isolation and the variation can be observed in the

Page 26: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

14 

 

 

deacetylation, degradation of the polysaccharide, and contamination with lignin. Among the

separation methods, several different extraction methods other than those conventionally used in

the pulp and paper industry have shown promising results for hemicelluloses recovery (Aspinall,

1959; Erins et al., 1976; Fengel, 1971; Koshijima et al., 1976; Mora et al., 1989; Timell, 1964;

Yoon and van Heiningen, 2010). Ward and Morak detailed an account of sequential extraction

methods using different types of extractants for fractionation of hemicelluloses from various

wood species (Ward and Morak, 77-88). A massive research undergoing in this direction is well

evidenced from the available publications (Koshijima et al., 1965; Casebier et al., 1969, 1973;

Fang et al., 2000; Garrote et al., 1999, 2004; Glasser et al., 2000; Puis and Saake, 2004; Yoon

and van Heiningen, 2010). In spite of these research attempts, only a few processes are reported

as under pilot scale projects (Glasser et al., 1995; Gabrieli et al., 2000; Puis and Saake, 2004).

However, an economic and effective commercial process to extract hemicelluloses from the

biomass is still needs to be exploited.

1.6.1 Acidhydrolysis

Acid hydrolysis processes can be categorized based on two approaches as concentrated acid/low

temperature and dilute acid/high temperature hydrolysis. Hemicelluloses are very sensitive to

depolymerization under acidic conditions at high temperatures (Carrasco et al., 1994). Sulfuric

acid is the most commonly used acid, whereas other mineral acids such as hydrochloric, nitric

and trifluroacetic acid have also been evaluated (Fengel and Wegener, 1984; Camacho et al.,

1996). Weak organic acids and phosphoric acids were studied in the dilute acid hydrolysis.

Typical sulfuric acid concentration and temperature ranges used for hemicellulose acid

hydrolysis are 0.5%-1.5% and 121-160OC respectively. (Girio et al., 2010). Depending on the

intensity of acid hydrolysis, sugar dehydration reactions also take place (Fengel and Wegener,

Page 27: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

15 

 

 

1984) leading to hydroxyfurfural (5-(hydroxymethyl)-2-furaldehyde or HMF) and furfural from

hexose and pentose mono sugars respectively. Under acidic conditions, glucose is subject to

isomerization reactions to produce a small amount of its isomers, fructose and mannose, and

undergoes reversible reactions which lead to formation of oligosaccharides, disaccharides and

anhydrosugars (Conner and Lorenz, 1986). Equipment corrosion, high acid recovery costs and

high capital investment are the hurdles for economic feasibility of the concentrated acid process

(Goldstein, 1983). Ramos (2003) studied the evaluation of acid hydrolysis process at a

commercial scale and found out that sugar degradation was the main constrain for the efficiency

of the process. However, there is a renewed interest in this process in the bio-ethanol industry

because of the moderate operation temperatures (Zhang et al., 2007).

1.6.2Hydrothermaltreatments

Liquid hot water or autohydrolysis and steam explosion are the two major hydrothermal

treatments for the isolation of hemicelluloses. In autohydrolysis, hot compressed water with high

pressure is used to hydrolyze the hemicellulose, whereas in steam explosion, breakdown of

polysaccharides is attained by the breakdown of structural components by steam, shear forces

due to the expansion of moisture and hydrolysis of glycosidic bonds (Chornet and Overend,

1988). Autohydrolysis is of interest because water is the only reagent making it an

environmentally friendly, inexpensive process compared to dilute mineral acid prehydrolysis

(Conner and Lorenz, 1986). Autohydrolysis is a hydronium ion self-catalyzed process with the

mechanism similar to that of dilute acid hydrolysis. The hydronium ions are generated in situ by

the autoionization of water in the autohydrolysis process, hydrolyses acetyl and uronic acid ester

substitutions that result in the formation of acetic acid and uronic acids which further hydrolyzes

polysaccharides to oligomers and monomers possible (Heitz et al., 1986; Nimela and Alen,

Page 28: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

16 

 

 

1999). The acetic acid released from the acetylated polysaccharides during autohydrolysis lowers

the pH of the extract to a range of 3 to 4 (Brasch and Free, 1965), allowing the removal of

hemicelluloses from the wood. The role of hot water extraction or hydrolysis of woody biomass

in the context of a biorefinery approach was reviewed recently by Liu (2010). The role of acetic

acid on the hemicellulose hydrolysis is reported as higher than that of the hydronium ion

generated during autohydrolysis of water (Al-Dajani and Tschirner, 2008). Autohydrolysis

extraction of Loblolly pine was reported to pre-extract hemicelluloses before kraft pulping (Yoon

and van Heiningen, 2008) and the study reveals that acidity of the reaction medium due to the

generation of acetic acid during hydrolysis can possibly lead to a serious degradation of

polysaccharides by acid hydrolysis. Though the role of uronic acid in the hydrolytic processes

has not been completely understood yet, it has been reported that these acids may also contribute

to the generation of hydronium ions (Conner, 1984). The operational temperature of

autohydrolysis varies between 150-230oC, and the reaction times vary from seconds to hours

depending on the reaction conditions such as temperature, and solid to liquid ratio (Garrot et al.,

1999a, b, and van Walsum et al., 1996; Tunc et al., 2008). Depending on the intensity of the

hydrolysis, sugar degradation also takes place resulting hydroxyl methyl furaldehyde (HMF) and

furfural compounds from hexoses and pentoses respectively (Fengel and Wegener, 1984). A

relatively high level of hemicelluloses recovery (55-84%), mainly a mixture of oligomers and

monomers has been obtained through autohydrolysis without affecting cellulose and lignin

significantly. (Conner and Lorenz, 1986; Nimela and Alen, 1999; Allen et al. 2001; Garrote and

Parajo, 2002). Hemicellulose extracts (liquor) after autohydrolysis of hardwoods contain mainly

xylooligosaccharides, low molar-mass cellulose, lignin, acetic acid, uronic acid and dehydration /

decomposition products such as furfural and HMF (Casebier et al., 1969, 1973; Lora and

Page 29: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

17 

 

 

Waymen, 1978; Garrote et al., 1999, Tunc et al., 2008). Also, at higher prehydrolysis

temperatures resinous deposits of lignin are formed (Leschinsky et al., 2007) and these deposits

hinder commercial implementation of the water pre-hydrolysis technology.

Steam explosion process, in general, results in (i) the hydrolysis of glycosidic bonds in the

hemicellulose and to a less extent to cellulose, and (ii) cleavage of hemicellulose-lignin bonds

leading to increased solubilization of hemicellulose and lignin degraded products. The process

involves heating woody biomass at high temperatures (preferably below 240oC) and pressure

followed by mechanical disruption of the material either by explosion (violent discharge in to a

collecting tank) (Schwald et al., 1989; Ramos et al., 1992, Saddler et al., 1982; Delong, 1983) or

without explosion (mild discharge after bringing the steam pressure down to atmospheric

pressure) (Brownell et al., 1986, 1987). Steam explosion carried out without any added catalyst

is an autohydrolytic process and most of steam treatments produce mainly oligosaccharides

along with lignin degradation products. A review of the chemistry involved in the steam

pretreatment was reported by Ramos (2003). Sugar recoveries of 45-69% are reported in the

absence of added catalysts (Heitz et al., 1991; Ballestors et al., 2000; Martin et al., 2008).

Impregnation of biomass with acid catalysts was also common to lower the treatment

temperatures and reaction time and thereby increasing hemicelluloses recovery (Boussaid et al.,

2001; Galbe and Zachi, 2007). Comparatively low molecular weight oligosaccharides results

from the severe degradation of wood and agro-residues during steam explosion treatment is a

bottleneck of the process (Glasser et al., 2000)

Page 30: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

18 

 

 

1.6.3 Alkalineextraction

Alkaline extraction at moderate temperatures is another common method for extracting

hemicelluloses from biomass (Dashek, 1997; Vuorinen and Alen, 1998; Ebringerova and Heinze,

2000; Glasser et al., 2000; Gabrieli et al., 2000). Hemicelluloses are partly alkali soluble and

highly accessible under hot alkali and acid conditions (Puis and Saake, 2004). Alkali hydrolyses

the ester linkages between plant polysaccharides and lignin, which releases polysaccharides and

increases the solubility of polysaccharides, without reducing their molar mass, under moderate

reaction conditions. The very common procedure to extract xylan is the alkaline extraction of

delignified wood. (Ebringerova and Heinze, 2000; Gustavasson et al., 2001). However,

delignification process modifies the chemistry of polysaccharides. (Sjostrom, 1993). Unlike hot

water or steam explosion process, alkaline treatment deacetylates the polysaccharide (van

Hazendonk et al., 1996).

In alkaline hydrolysis, several physical and chemical changes occur on hemicelluloses (Teleman

et al., 1995; Alen, 2000; Lai 2001; Sun and Cheng, 2002; Sun and Sun, 2002): alkali induces

swelling, leading to an increased surface area, a decrease in degree of polymerization, and

crystallinity occurs with a consequent separation of structural linkages between lignin and

carbohydrates. This results in the disruption of the lignin structure followed by saponification of

intermolecular ester bonds that cross link hemicelluloses and other components leading to an

increased porosity. In addition, removal of acetyl and uronic acid substitutions of hemicelluloses

also occurs.

Xylan is a hydrophilic polysaccharide shown to contribute to fiber wall swelling and flexibility

(Eriksson et al., 1991, Laine and Stenius, 1997). Swelling is the major factor in achieving a good

Page 31: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

19 

 

 

fractionation of hemicelluloses, where increasing concentration of alkali used. It was reported

that there is an optimum swelling concentration above which the swelling of the cellulose

inhibits the removal of hemicelluloses (Nelson and Schuerch, 1957). To improve the efficiency

of extraction of hemicelluloses using a given extractant, the structure must be as accessible as

possible to the extractant, and this can only be achieved by complete swelling. The importance of

swelling to remove hemicelluloses of different types of delignified wood in different extractants

was also emphasized by Hamilton and Thompson (1959), Dutton and Hunt (1958), and Morak

and Ward (1960;1961). The results of extraction depend on the type of alkali used and its

concentration and by developing a suitable extraction it is possible to attain a considerable

fractionation of xylan.

It was reported that hardwood xylans can be extracted in significant amounts using aqueous

alkaline solutions. The extractive power of lithium, sodium and potassium hydroxide solutions,

for hemicelluloses removal from Western Hemlock wood and its holocellulose was investigated

by Hamilton and Quimby (1957). They reported that sodium and lithium hydroxides had a

greater extractive ability than potassium hydroxide, for the isolation of glucomannans. However,

for xylose polymers containing arabinose and galactose, the extractive nature of all three alkali

salts was similar when used at concentrations equal to or greater than 2.8 N. Koshijima et al.

(1965) succeeded in removing 90% of xylan from trembling aspen wood using a 24% aqueous

potassium hydroxide solution. For the isolation of xylans from hardwood, in particular, two-step

procedures with a NaOH/H2O2 delignification step were shown to be more acceptable in practice

than the hazardous delignification with sodium chlorite. In a detailed study, Gabrielii et al.

(2000) described the isolation procedure and material properties of the methyl glucuronoxylan

(MGX) from aspen wood by an alkali extraction followed by a hydrogen peroxide treatment to

Page 32: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

20 

 

 

remove lignin, ultra filtration, and spray drying. The approach removed 55% of hemicelluloses

and the recovered xylan was reported to contain 65.2%xylan, 6.4% lignin, and 2.6% ash. A

comparison of the mild steam extraction and alkaline extraction of poplar wood and agricultural

residues showed that alkaline extraction has the advantage of providing polymeric xylan

compared to steam extraction (Glasser et al., 2000). The study also compares the efficiency of

various methods for the removal of associated lignin with the extraction and found that bleaching

with hydrogen peroxide followed by ultra filtration is a good procedure for isolation of xylan in

its polymeric form. A thermo-mechanical chemical fractionation of poplar hemicelluloses using

a twin screw extruder equipped with a filtration module in alkaline pulping was reported by

N’Diaye et al. (1996) and N’Diaye and Rigal (2000). Westbye et al. (2008) used a solid –liquid

extraction followed by a liquid-liquid extraction using pyridine/acetic acid mixture in an attempt

to produce xylan in a very pure form and was able to reduce the lignin content from 12% to 3%.

End-wise depolymerization (peeling) is one of the most important reactions of carbohydrate in

the alkaline reaction media that restricts the molecular mass of the extracted polysaccharides.

Hemicelluloses, like cellulose, possess a non-reducing end at one end of the chain and a reducing

end at the other. The peeling reaction removes one unit at a time from the reducing end of the

hemicelluloses chain until the chain end transforms to a stable oxidized carboxyl containing end

group, through a competing stopping reaction mechanism that hinders the peeling reaction

(Fengel and Wegener, 1984) (Figure 5). The rate of these reactions largely depends on the pH

and temperature of the alkaline medium. The activation energies for the peeling and stopping

reactions have been estimated to be 103 kJ/mol and 135 kJ/mol respectively (Haas et al., 1967).

Page 33: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

21 

 

 

Figure 1.5. Representation of peeling and stopping reactions of polysaccharides

The peeling reaction is nearly 65 times as fast as the stopping reaction (Franzon and Samuelsson,

1957; Lai and Sarkanen, 1967). Under alkaline conditions, deacetylated oligomeric xylan is more

stable due to the 4-O-methylglucuronic acid side chains and so alkaline extraction is suitable for

extraction of xylan from hardwood. However, solubilized polysaccharides can completely

degraded to sacharinic and hydroxyl acids under drastic conditions (Bhaskaran and von Koepen,

1970; Niemela and Alen 1999). Thus, it is possible to isolate xylan in its high molecular weight

before undergoing a complete peeling reaction by proper designing of the extraction method.

Selection of method of fractionation or pretreatments depends on the final requirement of the

product and the best pretreatment options are those which combine elements of both physical and

Page 34: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

22 

 

 

chemical methods (Ghosh and Singh, 1993; Saddler et al., 1993). In this regard, a fast and mild

alkaline extraction utilizing mild microwave energy would be an ideal choice for extracting

xylan in its polymeric form.

1.7 MicrowaveTechnology

Microwave radiation in the electromagnetic spectrum lies between infra red radiation and

radiofrequency waves. Frequencies of microwaves range from 300 MHz to 30 GHz

corresponding to the wavelength of 1cm-1m. The frequencies allotted for microwave heating

applications are 915 MHz and 2.45 GHz corresponding to a wavelength of 33.3 and 12.2 cm and

the latter is being used most often. Microwave chemistry has an edge over the conventional

methods for conducting chemical experiments and has the potential to achieve cleaner and

efficient reactions over conventional methods. Over the past three decades, this technology has

laid the foundation in the field of science with its applications extend from analytical

applications such as ashing, digestion, extraction, and fat analysis to synthetic organic and

polymer chemistry as synthesis of fine chemicals, organometallic intercalation, coordination, and

polymer curing (Bond et al., 1993; Caddick, 1995; Thostenson and Chou, 1999; Jones et al.,

2002; Hoz et al., 2005; Strauss and Varma, 2006; Zhang and Hayward, 2006). The enhanced

chemistry is attributed to the direct interaction between microwave energy and the materials of

interest which is explained in the following section.

The interaction between microwave and dielectric materials occurs through the dipoles or

induced dipoles present in the materials. The polarization of the dielectrics arises from the finite

displacement of or rotation of dipoles in an electric field. At the molecular level, this is due to the

distribution of the electron cloud within a molecule or the physical rotation of dipoles in the

Page 35: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

23 

 

 

presence of an alternating electric field and the latter being responsible for microwave radiation

heating.

Figure 1.6. Schematic of a sinusoidal electric field to an ideal dielectric (top) and the out-of-phase displacement current which is induced (bottom)

Microwaves consist of electric and magnetic field components and the electric field component

of microwaves interacts with the dipoles in the materials resulting in the rotation of the

molecules. In the microwave frequency range, the electric field component of microwave

radiation changes its direction 2.4 x 109 times per second (Metaxas and Meridith, 1983; Datta,

2001). The re-orientation of the dipoles and the resultant displacement of charge is equivalent to

an electric current, known as the Maxwell displacement current. For materials where the

molecules can keep pace with the field changes (ideal dielectrics), there is no lag between the

orientation of the molecules and the variations of the alternating voltage and the resultant

displacement current is 90° out-of-phase with the oscillating electric field as shown in Figure

Page 36: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

24 

 

 

1.6. There is no component of current in phase with the electric field. Rotation of polar

molecules lags behind the electric field oscillation and the resulting phase displacement

acquires a current component in-phase (I sin) with the electric field leading to resistive heating

mechanism in the medium and is described as dielectric loss (Figure1.7). If the phase angle

significantly differs from 90o, the materials can act both as a dielectric and conductor. The total

permittivity of the material - the parameter showing the materials ability to interact with the

radiation – is thus have a complex character and is expressed mathematically as

where ’ is the real part of the relative permittivity (the dielectric constant) and ’’ is the loss

factor which reflects the conductance of the material.

The phase diagrams representing ideal behavior and phase lag displacement and the relationship

between ’ and ’’ are shown in the Figure 1.7. In short, the interaction of microwaves with any

material depends on its dielectric properties: dielectric constant and dielectric loss factor. The

dielectric constant is a measure of the ability of a material to store electromagnetic energy and

the dielectric loss factor is a measure of the ability of a material to convert electromagnetic

energy into heat (Kumar et al., 2007). In the phase diagram, ''/' = tan, and is described as

energy dissipation factor (loss tangent) and is the parameter generally used to describe the

overall efficiency of a material to utilize energy from microwave radiation and convert to

thermal energy within the dielectric (Newham et al., 1991; Gabriel et al., 1998; Datta, 2001).

Page 37: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

25 

 

 

Figure 1.7. (a) Phase diagram for an ideal dielectric where the energy is transmitted without

loss; (b) phase diagram where there is a phase displacement and (c) the

relationship between ’ and ’’

1.7.1 Microwaveassistedsolventextraction

Microwave assisted extraction is a process where microwave energy is used to partition the

material of interest from the sample to the surrounding liquid. This process has been significantly

exploited in the pharmaceutical, fine chemicals, and environmental applications. Unlike

conventional thermal processing, where energy is transferred to the material through convection,

conduction, and radiation of heat from the surfaces of the material, microwave energy is

delivered directly to materials through molecular interaction with the electromagnetic field and

deposit energy within the material leading to heat generation throughout the material (which is

Page 38: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

26 

 

 

called volumetric heating). Moreover, when materials with different dielectric properties exposed

to microwave radiation microwaves will selectively couple with the higher lossy material. In

general, the higher the dielectric constant, the more efficient the molecules absorb energy and

hence are heated more efficiently. Molecules with low dielectric constant and loss factor cannot

absorb energy as they cannot couple efficiently with the microwave radiation. The special

heating mechanism and the fact that different chemical substances absorb microwaves at

different levels can make microwave assisted extraction an efficient and selective method of

extraction of target compounds.

1.7.2 Microwaveassistedextractionofplantmaterials

Extraction of plant compounds using microwave assisted process is highly dependent on the

solvents used. Selection of solvent depends on the solubility of the target compound and the

dielectric properties of the solvent. When a polar solvent with a relatively high dielectric

constant and loss factor is used, the solvent will be heated through dipole rotation by interaction

with microwave energy. The heated solvents will accelerate the process of desorption of the

matrix-solvent interface and the diffusion of the target compounds into the solvent (Hawthorne et

al., 1995). The special extraction mechanism of microwave assisted extraction can be better

interpreted when a non-polar solvent is used for extraction, where the material interacts with the

microwave radiation. Studies on the solvent extraction of leaves and seeds with highly absorbing

solvents and non-polar solvents have been demonstrated to selectively extract the compound of

interest. It was found that in the case of non-polar solvent extraction, a significant fraction of

microwave energy is absorbed by the sample mainly by the water in the glandular and vascular

systems and the volumetric heating effect enhanced the extraction process and reduced the time

of extraction (Pare and Belanger, 1994; Chen and Spiro, 1994; 1995; Chemat et al., 2005).

Page 39: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

27 

 

 

Ganzler et al. (1986) reported that for extraction of crude fat, and other components from oil

seeds, the duration of microwave assisted extraction was almost 100 times less than the

traditional methods. Pare (1995) also reported similar observation to extract plant components

from cedar, garlic, parsley and mint leaves. The reduction in the time of extraction compared to

conventional extraction is the unique feature in all these extraction studies.

The earliest known study in the lignocellulosic materials on microwave pretreatment examined

the effect of microwave radiation on rice straw and bagasse immersed in water and reported an

improvement in total reducing sugar production by a factor of 1.6 for rice straw and 3.2 for

bagasse in comparison to untreated biomass (Ooshima et al., 1984). Microwave pretreatment of

sugar cane bagasse and rice hulls soaked in water followed by lignin extraction was reported to

yield 77-84% of total available reducing sugars (Azuma et al., 1984). A similar study involving

microwave pretreatment of rice straw and sugarcane bagasse followed by lignin extraction

reported a yield of 43-55% of total available reducing sugars (Kitchaiya et al., 2003). Microwave

pretreatment of rice straw soaked in dilute alkali resulted in a glucose yield of 65% and total

carbohydrate conversion of 78% (Zhu et al., 2005). A recent study on microwave based alkali

pretreatment of switch grass investigated low power levels for extended pretreatment time and

reported 70-90% sugar yields (Hu and Wen, 2008). Application of microwave technology in the

extraction of hemicelluloses from various biomasses including flax shives (Jacobs et al., 2003;

Burnov and Mazza, 2010), wood (Teleman et al., 2000; Palm et al., 2003; Jacobs et al., 2002;

Lundquist et al., 2002), corn pericarp and fiber (Yoshida et al., 2010; Benko et al., 2007), and

barley husk (Roos et al., 2009) were reported. The details of the studied parameters such as yield

of xylan, molecular weight distribution, and lignin content are cited in the Table 1.3.

Page 40: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

28 

 

 

There is great fundamental and tremendous environmental and economic importance entailed to

achieve the extraction of hemicellulose, xylan, with minimum molecular alteration. But, the

studies reported so far either performed at a high temperature (>120oC) using water, acid, or base

catalyzed reaction for extraction or a pressurized solvent assisted extraction. The severe

conditions of extraction leads to degradation of the xylan and further, such processes require

costly experimental set up which has a negative impact on the cost efficiency. Also, these

studies reported different yields (lower and higher) and different characteristics of the xylan

extracted compared to conventional methods. Moreover, there is no available literature on

systematic analysis to differentiate the microwave assisted extraction from that of conventional

extractions. Hence, a fundamental study focused on the mechanism of extraction of xylan from

the plant materials using microwave energy will be justified for understanding and developing an

efficient physico-chemical extraction process for extracting xylan from the plant biomass. The

study would provide not only an innovative knowledge on the fundamentals of microwave

assisted extraction but also, an environmentally friendly method for the isolation of

hemicellulose (xylan) with minimum degradation, which is highly valuable to the bio-refinery

industry.

Page 41: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

29 

 

 

Table 1.3: Available literature on microwave assisted extraction of polysaccharides from various biomass Authors Raw material Solvent used Yield & Lignin

content Molecular weight

Burnov and Mazza (2010) Teleman et al., (2000) Palm and Zacchi (2003) Benko et al., (2007) Jacobs et al., (2002) Alexandra et al., 2009 Lundquist et al. 2002 Jacobs et al., (2003) Yoshida et al., (2010)

Flax Shives Aspen wood Spruce Corn fiber Spruce, Aspen Barley husk Spruce Flax shives Corn pericarp

Water; ethanol Water Water; ethanol Sulfuric acid, sodium hydroxide and water water Water, dil. Alkali, Dil.acid (NaOH, H2SO4) Water, NaOH, KOH and H2SO4 Acid and base catalysed hydrothermal microwave treatment Alkaline Hydrothermal treatment

xylooligosaccharides; Lignin 6.4% Predominant compound in the extract were Xylo-oligomers No data on lignin Yield: Water (20%), Ethanol (6%) No data on lignin No data 11-30% NaOH-4-11% H2SO4 -2-11 Water-20% No data Water- 50% Acid -35% Base – 9% Galactoglucomannan oligopolysaccharide yield- 70-60%

No data No data Mw-water 32251 Mw-alcohol 3965 Average degree of polymerization 2-35 24000-870 g/mol (DP-133) Mw 136000-172000 13000-18000 127000 DP- 6-20 Mw-13,000 Mw-13,000 Mw-52,000 Mol.Wt., 800-80000 Mw-low 1700-4500

Page 42: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

 

30 

 

Chapter2 ResearchHypothesisandObjectives

2.1 Hypothesis

Selective interaction of microwave energy with the more lossy component (alkali) in the wood

fiber generates hot-spots and the resulting "explosion effect" rupture the recalcitrant

lignocellulosic structure and increase porosity thereby enhancing the mass transfer of hydrolyzed

components to the solution and reduce the time of extraction.

2.2 ResearchObjectives

1. Investigate and establish an efficient physico-chemical process using microwave energy

for the extraction of polymeric xylan from lignocellulosics for developing value added

polymers.

2. Investigate the effect of interaction of microwave energy with the wood slurry on the

extraction of xylan in comparison with conventional heating process

3. Elucidate the mechanism involved in the microwave assisted alkaline extraction and

study the effect of microwave irradiation on the physico-chemical properties of xylan

4. Optimize microwave assisted extraction of xylan from birch wood.

Page 43: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

 

 

31 

 

Chapter3MaterialsandMethods 

  

3.1 Materials

3.1.1 Birchwoodfibers

Birch wood fibers were prepared using chips obtained from St. Mary’s Paper Company, Ontario,

Canada, and from birch wood log collected near the University of Toronto’s premises. The bark

of the logs was removed and was cut into smaller blocks. The chips and small wood blocks were

dried at 40oC for 5 days, and were separately ground to 2 mm size using a Wiley mill. The

ground wood fibers were sieved using a 0.42 mm sieves to remove the fines. Extractives were

removed from these fibers before the xylan extraction. Preparation of extractive-free wood fibers

involved the following procedure. The fibers were treated with 0.05 M hydrochloric acid (solid

to liquid ratio of 1:10 g/mL) at 70oC for 2 hours. After cooling the suspension, ammonium

hydroxide (14 M) was added to a pH of 9-10 and the fibers were allowed to swell, in the

suspension, overnight to remove pectins, starch and fat (Gabrielii et al., 2000). The suspension

was then filtered through 0.42 mm sized screen and washed thoroughly with water until the

washings were neutral. The extractive-free wood fibers were then dried at 40oC for 72-96 hours

and stored in air tight containers until used.

3.1.2 Chemicals

All the chemicals used in this study were of analytical or reagent grade and are listed below:

0.05 M hydrochloric acid, ammonium hydroxide, sodium hydroxide, sulfuric acid (72% and

96%), and dilute acetic acid were of reagent grade; HPLC grade de-ionized water for HPLC

Page 44: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

32 

 

 

 

analysis; 50 wt% HPLC grade sodium hydroxide solution for HPLC analysis; birch wood xylan

used as a reference for FTIR characterization, sugars used as standards for chromatographic

analysis (glucose, xylose, mannose, arabinose, and galactose), and dextran standards with

different molecular weights used for size exclusion chromatography were of analytical reagents

and were obtained from Sigma Aldrich.

3.2 Methods

3.2.1 Isolationofxylanfrombirch

The protocol used for the isolation of xylan from birch fibers is given in the Figure 3.1. Slurry of

the wood fibers in sodium hydroxide solution was subjected to conventional and microwave

assisted extraction. Xylan was precipitated from the extract and the wood residue was

characterized to investigate the structural changes occurred during the extraction process. The

extract after precipitation of xylan was used to quantify the molecular degradation of the xylan

during the extraction process.

Page 45: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

33 

 

 

 

Figure 3.1. Protocol for the xylan extraction and characterization

3.2.1.1 Extractionofxylanfrombirchusingmicrowave 

A general purpose microwave oven (Sanyac Model) operating at 2450MHz and with adjustable

microwave power input (110 to 1100W) was used for microwave assisted extraction. Wood

fiber slurry was prepared by adding required amount of sodium hydroxide (NaOH) solution

(concentration of NaOH solution: 1 wt% to 4 wt%) to the weighed extractive-free wood fibers

taken in an Erlenmeyer flask. The slurry was kept at room temperature for 5 minutes to

Conventional method of 

extraction

Filtrate 

Characterization 

Degraded sugar content 

Microwave assisted 

extraction 

Extractive free 

wood sample

Filtrate 

Xylan 

Characterization 

Sugar composition, Lignin 

content, Molar mass,  & 

FTIR, 

Precipitation of xylan

Residue Residue 

Characterization 

SEM & X‐ray tomography 

FTIR, X‐ray crystallography 

Page 46: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

34 

 

 

 

completely wet the wood fibers and placed in the middle of the oven over the rotating plate and

xylan was extracted by exposing the slurry to microwave radiation according to the experimental

conditions explained in the following chapters. At the end of heating, temperature of the mixture

was noted. Immediately after, the mixture was filtered through a previously weighed crucible

and the residue was washed with distilled water (50-100 mL depending on the conditions of

extraction) and combined the washings with the filtrate. The filtrate was then neutralized with

acetic acid to a pH of 4.6 to precipitate dissolved xylan. The precipitate was allowed to settle

overnight and separated from the liquid by centrifugation at room temperature. An aliquot of

liquid phase after the separation of precipitated xylan was collected to determine the dissolved

sugar content. The separated xylan was washed with water to remove the salt, 95% ethanol,

centrifuged, and freeze-dried or oven dried at 40oC. In some cases, the precipitated xylan was

re-dissolved in minimum amount of sodium hydroxide and the solution was poured into 3

volumes of ethanol to re-precipitate the xylan. The precipitated xylan was then washed with

water, centrifuged and freeze dried. The wood fibers after each extraction was washed until the

washings were neutral to pH paper and were dried at 40oC for 72-96 hours. All the extractions

were carried out at least three times to get an average result.

3.2.1.2 Extractionofxylanfrombirchusingconventionalheating Conventional extraction was performed at different temperatures according to the experimental

conditions described in the respective chapters using a water bath or oil bath depending on the

temperature used for the study. The wood fiber slurry was extracted isothermally using sodium

hydroxide solution at different temperatures. The protocol used for separation of wood residue

Page 47: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

35 

 

 

 

from the liquid phase and precipitation of xylan after extraction were the same as those used in

microwave extraction.

3.2.2 CharacterizationofBirchwood

Figure 3.2. Experimental protocol for characterization of wood

Extractive-free wood fibers were characterized for moisture content, chemical composition, and

ash content. A sample of the extractive-free wood fiber was ground to 0.25 mm size using a

Wiley mill and used for the determination of chemical composition. The experimental protocol is

demonstrated in Figure 3.2. For all the quantitative analysis, three samples were used and the

results expressed are the average of three measurements.

3.2.2.1 Holocellulose Holocellulose represents the entire polysaccharide, cellulose and hemicelluloses, portion of

wood. Determination of holocellulose followed the method described by Zobel and McElvee

(1966) by dissolving the lignin in an acidic medium with chlorine based solution. The procedure

used was as follows: About 0.3-0.5 g of air dried extractive-free wood was accurately weighed

Moisture content  

Holocellulose 

Extractive free Birch woodAsh content 

 ‐ cellulose  Lignin Sugar composition 

Page 48: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

36 

 

 

 

into a 125 ml Erlenmeyer flask and added 32 ml of distilled water, followed by 0.1 ml glacial

acetic acid and 0.3 g reagent grade sodium chlorite. The flask was capped with a loose fitting

inverted 10 ml Erlenmeyer flask and heated in a temperature controlled water bath at 70oC. The

contents were heated for 1 hour with occasional swirling to ensure a uniform mixing of the

reaction mixture. Without cooling, an additional 0.1 ml of glacial acetic acid and 0.3 g sodium

chlorite were added successively after one hour for two times. After three hours of heating at

70oC, the flasks were placed in an ice bath and cooled the reaction mixture below 10oC. The

contents of the flask were filtered through a previously weighed coarse fritted glass crucible

using a minimum quantity (25 ml) of ice distilled water to transfer all the holocellulose and

remove the color and odor of the chlorine dioxide. The contents were then washed with 100 ml

of hot distilled water under suction, washed with acetone without suction, dried by suction and

dried in an oven at 105°C for 24 hours. The crucible with the sample was allowed to cool in a

desiccator before weighing. Holocellulose was determined using the equation 3.1.

%

100

3.1

3.2.2.2 ‐Celluloseandhemicelluloses 

-Cellulose was determined from holocellulose after hydrolyzing the heterogeneous

polysaccharides for 2 hours using 17.5% alkali, as per the TAPPI procedure T203 om-93.

Holocellulose used for - Cellulose determinations were not oven dried, but air dried for

overnight in a conditioning chamber. -Cellulose in the sample was calculated using the

equation 3.2.

Page 49: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

37 

 

 

 

%

100

3.2

The difference between the holocellulose and -cellulose was accounted as hemicelluloses

content.

3.2.2.3 AshContent 

Ash content of the samples was carried out using a modified Tappi procedure T413 pm 95.

Representative samples were weighed accurately in previously weighed silica crucibles, and

heated at 600oC for 4 hours to remove the carbonaceous materials. The samples were then cooled

to room temperature in desiccator before weighing. Ash content was calculated based on the

oven dry weight of the wood sample using the equation 3.3.

%

100 

 

3.3.

3.2.2.4 SugarCompositionandLigninContent 

The constituent sugars and lignin content of the wood fibers were determined using a two-step

hydrolysis with 72 and 4% sulfuric acid (Sluiter et al., 2008). Samples were hydrolyzed using

72% sulfuric acid, followed by a second hydrolysis of the samples with 4% sulfuric acid for 1

hour at 121oC in an autoclave. The procedure involved is as follows: About 0.3 g of the wood

sample was measured in to a tared glass tube followed by 3 mL of 72% sulfuric acid and mixed

the reaction mixture thoroughly with a stirring rod. The glass tubes were incubated in a water

bath set at a temperature of 30oC with frequent stirring, without removing the sample from the

water bath. After 1 hour of incubation, the samples were completely transferred to pressure tubes

Page 50: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

38 

 

 

 

using 84 g of deionized water, using a balance, in order to dilute the acid concentration to 4%.

The samples were then autoclaved at 121oC for 1 hour. A set of sugar recovery standards with

4% sulfuric acid was also autoclaved to determine the loss of sugar during the hydrolysis.

The hydrolyzed samples were then filtered through a previously weighed medium fritted glass

crucible, and an aliquot of the filtrate was kept aside for the determination of sugar composition

and acid soluble lignin. The solid residue from the flask was completely transferred

quantitatively to the crucible using deionized water. The residue was then washed with hot

deionized water until the washings were free from acid. The crucibles with the acid insoluble

residue were dried at 105oC for 24 hours and were then cooled to room temperature before

weighing. Acid insoluble lignin was determined using the equation 3.4.

100 

                        3.4 

 

For determining acid soluble lignin, absorbance of the collected filtrate was measured using a

UV-visible spectrophotometer at a wavelength of 240 nm. An aliquot of the sample was diluted

using deionized water to get the absorbance reading in the range of 0.7-1. The amount of acid

soluble lignin was calculated using the equation 3.5.

∈100 

3.5

where, UVabs = average UV-visible absorbance for the sample at 240 nm;

Volume filtrate = volume of the filtrate (87 mL)

Page 51: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

39 

 

 

 

              3.6 

 

where, ε = absorptivity of biomass at specific wavelength ( at 240 nm, ε is 25 L/g cm).

For the monosaccharide composition of the hydrolyzed samples, an aliquot (30 mL) of the

filtrate sample was transferred to an Erlenmeyer flask, and neutralized using calcium carbonate

to a pH of 5-6. The sample was allowed to settle, and decanted the supernatant. The neutralized

samples were analyzed using a high performance anion exchange chromatography (HPAEC)

(Dionex system) equipped with a Carbopac PA10 column, a pulsed amperometric detector (ED-

40) and an auto sampler. Sugars were separated using 3 mM NaOH solution under isocratic

conditions for 25 minutes. Flow rate of eluent was 1 mL/minute and the detection was performed

by pulsed amperometry (ED-40, Dionex). After each isocratic elution, the column was cleaned

with 250 mM NaOH for 10 minutes and then equilibrated with the eluent for further 20 minutes.

The total time for the analysis of each sample was 57 minutes. Calibration was performed with

five sets of standard solutions containing L(+) arabinose, D(+) galactose, D(+) glucose, D(+)

xylose, and D(+) mannose in appropriate concentrations. Calibration plots obtained for different

sugars are shown in Figure 3.3. Two replicates of each sample prepared were analyzed and for

each sample two replicates of hydrolysis were performed. Chromeleon software was used for

the quantification of the monosaccharides using the calibration curves. Content of individual

polysaccharides was calculated by multiplying the content of corresponding mono sugar with the

correlation factors; 0.88 for xylose and arabinose, and 0.9 for glucose, mannose and galactose

(Sluiter et al., 2008). The results reported were the average of four analysis results.

Page 52: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

40 

 

 

 

Figure 3.3. HPLC calibration graphs for sugars  

3.2.2.5 ScanningElectronMicroscopy(SEM) 

The wood fibers before and after both extraction processes, were subjected to SEM (Hitachi S-

2500, Tokyo, Japan) to study the changes in the microstructure during extraction. The dried

fibers were mounted on metal stubs and were sputter coated with a thin layer of gold to avoid

electrostatic charge during examination. Images were taken using an accelerating voltage of

15kV and a working distance of 10 mm.

3.2.2.6 X‐raydiffraction 

XRD- measurements were conducted on a Bruker AXS D8 Discovery diffraction system (Bruker

AXS Inc., Madison, WI, USA) equipped with a high power point focus (1x1mm) Cu-kα target,

graphite monochromator (26.53o) for elimination of Cu-kβ lines, and a Hi-Star GADDS area

y = 425.37xR² = 0.9956

y = 505.37xR² = 0.996

y = 534.71xR² = 0.9959

y = 620.45xR² = 0.9964

y = 501.5xR² = 0.9979

0

4

8

12

16

20

0 0.005 0.01 0.015 0.02 0.025 0.03 0.035

Are

a o

f th

e p

ea

k

Concentration (mg/mL)

Arabinose Galactose Glucose Xylose Mannose

Page 53: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

41 

 

 

 

detector for 2-D images. The powdered samples having a size of 0.25 mm were pressed into the

rectangular sample holder with dimensions of 15 x 20 mm and with a thickness of 1mm. The

mass of the sample in the sample holder was approximately the same for all the experiments. The

X-ray diffractometer was operated at a voltage of 40 kV with a current density of 40 mA. The

scanning range was from 2 - 5o to 40o at a scan speed of 2o/s. The data was collected using a

continuous mode with angular intervals of 0.02o. The percentage of crystalline material in the

wood fibers was evaluated as a crystallinity index (CrI) using Segal's method (Segal et al.,1959)

as shown in the equation 3.7.

%

100 (3.7)

where I002 is the maximum intensity (in arbitrary units) of the 002 diffraction peak at 2 = 22.8o

and Iam is the intensity of diffraction in the same units at 2 =18o.

3.2.2.7 X‐raymicrotomography

X-ray microtomographic technique was used to quantify the porosity of the wood samples. A

Skyscan 1172 high resolution desktop X-ray micro CT system (Skyscan, Artselaar, Belgium:)

was used to obtain the microtomographical images of the samples. The system consists of a

microfocus sealed X-ray tube with high voltage power supply, an object stage with a precision

manipulator, a 2D X-ray CCD detector and a computer with tomographic reconstruction

software (NRecon from Skyscan). The X-ray CCD is based on a medium resolution of cooled

CCD sensor with a fiber optic coupling to an X-ray scintillator. The samples were scanned using

a medium resolution of 2014x1024 at a source voltage of 46 kV and a current of 180 µA for

thicker samples and 30 kV and 122 µA for thinner samples. The X-ray radiographs were

recorded over the interval of 0 to 360 degrees using a stepwise rotation of 0.4 degree. The X-ray

Page 54: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

42 

 

 

 

source and detector were remained at a fixed position. In order to improve the signal to noise

ratio, and to get a high quality radiograph images a 6 frame averaging, (each radiograph image

obtained is the average of 6 shadow projections) was used. After the image acquisition, the 2D

cross sectional images of the samples were reconstructed using Skyscan's reconstruction

software (NRECON_version 1.4.4). The 2D binary images of the samples obtained were

processed for 3D construction and porosity analysis using image analysis software Image Pro

(version 6) and the method is described in Chapter 6.

3.2.3 Analysisofliquidextracts

Lignin content of the liquid extract after extraction processes was determined by measuring the

UV absorbance at a wavelength of 240 nm (Sluiter et al., 2008) using a similar equation for the

soluble lignin as described in section 3.2.3.4.

The total sugar content of the liquid phase after precipitating the xylan was determined by

phenol-sulfuric acid method using glucose as the standard (Dubois et al., 1979; Fournier, 2001).

A series of 10 standard solutions with different concentrations were prepared using a glucose

standard solution (1 mg/mL). A definite volume of the standard sugar solutions (5 µL-50 µL)

were mixed with 500 µL of 4% phenol and 2.5 mL of concentrated sulfuric acid (96%) to

develop the color and the absorbance of the solutions was then measured using UV-visible

spectrophotometer at 490 nm to prepare the calibration curve. Similar procedure was used to

develop color for the samples of liquid phase after the precipitation of xylan. Volume of the

sample taken was adjusted to obtain the absorbance within the range of the absorbance of the

standard solutions. Concentration of the sugar (mg/mL) was determined from the absorbance of

the sample solution, using the calibration curve (Figure 3.4).

Page 55: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

43 

 

 

 

The percentage of total sugar based on oven dried wood was calculated using the equation 3.8.

% .

100 3.8

Figure 3.4. Calibration graphs for total sugar content determination

3.2.4 Characterizationofxylan

3.2.4.1 Sugarcompositionandlignincontent 

The sugar composition and lignin content of the extracted xylan was determined by a two-step

acid hydrolysis using the same procedure as described in the wood characterization.

3.2.4.2 Molecularweightdetermination 

The molecular weight of the extracted xylan was determined using an HPLC equipped with a

size exclusion PL Aquagel-OH mixed column (300 x7.5 mm) (Agilent technologies), auto

y = 0.0189xR² = 0.9731

0

0.2

0.4

0.6

0.8

1

1.2

0 10 20 30 40 50 60

Ab

so

rpti

on

at

49

0 n

m

Concentration of sugar (mg/mL)

Page 56: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

44 

 

 

 

sampler, and refractive index (RI) detector. The eluent used was 0.02 M NaCl in 0.005 M

sodium phosphate buffer (pH 7) at a flow rate of 1 mL/minute. Dextran (Sigma Aldrich) with

molecular mass of 5000, 25000, 50000, 150000, 250,000, and 670,000 Da were used as

standards. Aqueous solutions (1 mg/mL) of dextran standards were used to prepare the

calibration curve (Figure 3.5). Samples of xylan (1 mg/mL) were prepared in the eluent. The

standards were detected by RI and calibration curve was prepared using the elution volume

obtained from the standards. Molecular mass (Mw) of the samples was calculated as dextran

equivalents from the elution peaks obtained using the calibration curve. The relative percentage

of each fraction was reported as the relative ratio of the area of each fraction.

Figure 3.5. Calibration graph for molar mass determination

Page 57: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

45 

 

 

 

3.2.4.3 Viscositymeasurement 

 

Viscosity of the xylan solution was determined using cupriethylenediamine (CED) as the

solvent. The xylan obtained was purified further by re-dissolving in sodium hydroxide solution

followed by re-precipitation by adding 3 volume of ethanol. The precipitated xylan was then

washed with ethanol and freeze-dried. The purified xylan powder was dissolved in 0.5 M CED to

prepare solutions of various concentrations (0.1-0.5 g/dL). The relative viscosity of the xylan

solutions was determined using a Cannon-Fenske viscometer at 25oC. Specific viscosity (ηsp) and

reduced viscosity (ηsp/C) was calculated according to the following equations.

Relative viscosity (ηrel) = t/t0 (3.9)

Specific viscosity (ηsp) = (t-t0)/t0 (3.10)

Reduced viscosity (ηsp/C) = (t-t0)/t0C (3.11) where C is the concentration of the xylan solution (g/dL), t and t0 are the times required for the

solvent and solution respectively to flow down the capillary tube of the viscometer. Intrinsic

viscosity was obtained by extrapolating the reduced viscosity to zero concentration by the least

square method.

 

3.2.4.4 FTIRSpectra

Approximately 2 mg of the powdered sample was mixed with 200 mg of dry potassium bromide

(KBr) and pelletized the mixture. A Bruker Optics FTIR was used to record the absorbance

Page 58: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

46 

 

 

 

between 4000 and 400 cm-1 with a resolution of 4 cm-1. The FTIR backgrounds were measured

using KBr pellet. Each spectrum represents the average of 64 scans and each spectrum was

corrected for atmospheric compensation, followed by baseline correction, and normalization

using minimum-maximum method (available in Opus software, v. 5.0, Bruker Optics). The

spectra reported are the average spectra of 4 measurements.

Page 59: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

 

 

47 

 

Chapter4 Evaluationofmicrowaveassistedalkalineextractionofbirchxylan

 

Aim of this study was to investigate the efficacy of microwave energy for the extraction of

xylan, the major hemicelluloses present in birch wood, using mild alkali under atmospheric

conditions. The study focused on the effect of microwave power input and time of irradiation on

the solubilization of the wood fiber and yield of xylan to understand the effect of process

variables on the extraction of xylan and to determine the process parameters for obtaining

maximum yield. X-ray diffraction studies were performed to study changes in the crystallinity of

the fibers after extraction. Chemical composition, molecular weight and FTIR analysis of

isolated xylan under different experimental conditions were used to evaluate the effect of process

parameters on the structural characteristics of xylan.

4.1.MaterialsandMethods

4.1.1 Birchwoodfibers

Birch wood fibers obtained from St. Mary’s Paper, ON, Canada were used for extraction of

xylan in this study. Birch wood fibers believed to be a good source of xylan and typically

hardwood contains about 20-30% of xylan. The chemical composition of extractive-free birch

wood fibers is given in the table 4.1. All the results given are based on the average of three

Results of this chapter have been published in the Journal of Polymers and the Environment.

December 2013, Volume 21, Issue 4, pp 917-929

Page 60: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

48 

 

 

 

samples. Birch wood fibers contained about 73.4 % of holocellulose, 25% of lignin, and 0.3 %

of ash content. Carbohydrate composition of wood fibers includes 42% glucan, 22% xylan, and

1.8% of mannan, accounting a total of 65.8%. The difference between the holocellulose and the

total sugar content 7.6% is accounted for the acetyl and glucuronic acid residues.

Table 4.1. Chemical composition of extractive-free birch wood fibers

Component Amount (%) Holocellulose Alpha cellulose Hemicelluloses Lignin Soluble lignin Insoluble lignin Ash Carbohydrate composition

Glucan Xylan Mannan Arabinan Galactan

73.4 ± 3.4 40.1 ± 4.7 33.3 ± 5.8 5.1 ± 1.2 19.9 ± 0.7 0.3 ± 0.01 41.6 ± 1.1 22.3 ± 1.5 1.8 ± 0.8 Not detected Not detected

4.1.2 Extractionofxylanfrombirch

Table 4. 2. Experimental conditions used for microwave extraction

Microwave power level (W)

Time (s)

110 330 550 770 110

60, 120, 240, 360, 480, 600 30,60,90,120,180 30,60,90,120,180 10,20,30,40,60 10,20,30,40,60

Page 61: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

49 

 

 

 

Experimental conditions used for the extraction of xylan from birch wood using NaOH solution

are given in the table 4.2. The protocol used for the extraction, separation of the residue from the

liquid extract, and precipitation of xylan from the extract were explained in section 3.2.1.

The amount of wood dissolved (solubilization of wood) during extraction was expressed as the

ratio of the wood dissolved to the weight of the wood used for extraction as shown in the

equation 4.1.

%

. .

. 100

4.1 where OD stands for the term oven dry.

The yield of xylan was calculated as the ratio of the amount of xylan obtained to the total amount

of xylan present in dry wood as shown in equation 4.2.

%

100 4.2

4.2 ResultsandDiscussion

4.2.1 Effectofsodiumhydroxideconcentrationontheextractionofxylan

Extraction of hemicelluloses highly depends on the concentration of the alkali used, as it affects

the hydrolysis of the hemicelluloses-lignin linkages. In order to make the process less harsh and

more environmentally friendly, lower concentration of alkali is preferred. To find out the

minimum concentration at which a moderate extraction of xylan is possible, different

concentrations of alkali (1 wt%, 3 wt%, and 4 wt %) were used for the extraction at a microwave

power input of 110W from 2-10 minutes. In birch, since majority of the hemicelluloses

Page 62: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

50 

 

 

 

accounted for xylan (22%) with only a small amount of mannan (1.8%), our study focused on the

xylan.

Figure 4.1. Effect of NaOH concentration on the wood dissolution and xylan yield (Power

level 110W) The amount of wood solubilized and the yield of xylan obtained after different extraction

conditions are shown in the Figure 4.1. As expected, wood solubilization and the yield of xylan

0

5

10

15

20

25

0 2 4 6 8 10 12

Wo

od

lo

ss

, %

Time, min

1% 3% 4%

0

10

20

30

40

50

60

70

0 2 4 6 8 10 12

Xy

lan

Yie

ld,

%

Time, min

1% 3% 4%

Page 63: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

51 

 

 

 

increased with the increase of NaOH solution concentration. The amount of wood solubilized

and the yield obtained using 4 wt% NaOH solution was almost double as that of 1 wt% and 3

wt% solution and hence 4 wt% NaOH solution was used for the rest of the study.

4.2.2 Effectofirradiationtimeandmicrowavepowerontheextractionofxylan

In order to study the effect of microwave power input and irradiation time on the efficiency of

xylan extraction, wood fibers were extracted with 4 wt% sodium hydroxide solution (solid to

liquid ratio: 1: 10 g/mL) at different microwave power inputs (110 W – 1100 W) for different

periods of time (10 sec-10 minutes) (Table 4.2). The results were studied in terms of

solubilization of wood and the yield of precipitated xylan.

4.2.2.1 Solubilizationofwood 

Solubilization of wood components and the resultant wood loss are caused by the hydrolysis of

the esterified linkages of hemicelluloses and lignin and the transfer of these wood components

into the alkali. The mechanism of transport phenomena occurs during the alkaline reactive

extraction of hemicelluloses from wood consists of the following steps as shown in Figure 4.2:

(i) diffusion of NaOH solution (axially) from the bulk phase to the vessels or central canal in the

cells (ii) diffusion of NaOH from the voids (radially) to the reaction sites into the cell walls

(primary and secondary cell walls) (iii) hydrolysis of the hemicelluloses-lignin linkages,

liberation and solubilization of the hemicelluloses (iv) radial diffusion of hemicelluloses from the

walls to voids, and (v) axial diffusion of hemicelluloses from the central voids to the bulk phase.

Dissolution of lignin along with the hemicelluloses is also expected under alkaline extraction

conditions.

Page 64: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

52 

 

 

 

Figure 4.3 shows the amount of solubilized wood after each extraction for different microwave

power inputs and time of irradiation. It is obvious that dissolution of wood depends on the

extraction conditions such as irradiation time and microwave power as these are the two main

factors affecting the radiation intensity on the reaction mixture and the resulting temperature rise.

At all the power levels studied, wood solubilization increased with the time of irradiation. This

is attributed to the temperarure increase of the wood slurry as a result of longer irradiation time

(Figure 4.4). As the temperature increases, diffusion of NaOH solution as well as the hydrolysed

components (hemicelluloses and lignin) increases that respectively leads to the enahnced

Figure 4.2. Mechanism of alkaline hydrolysis and dissolution of hemicelluloses (Prat et al., 2002)

hydrolysis of hemicelluloses-lignin linkages and their transfer to the solution. For example, at a

power input of 110W, solubilization of wood increases from 9% at an irradiation time of 1

minute (temperature was 43oC) to 19.7% for 10 minutes of radiation (temperature was 96oC),

(Figure 4.3a) indicating a high dissolution at severe conditions of extraction. A similar trend in

Page 65: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

53 

 

 

 

wood dissolution with time was observed at all the power inputs studied.

Figure 4.3. Effect of irradiation time and microwave power on the solubilization of wood (a) 110W (b) 330W, and 550 W (c) 770W and 1100W

0

5

10

15

20

25

0 200 400 600 800

Wo

od

so

lub

iliza

tio

n (

% O

D b

as

is)

Time (s)

0

5

10

15

20

25

0 50 100 150 200 250

Wo

od

so

lub

iliza

tio

n (

% O

D b

as

is)

Time (s)

330W 550W

0

5

10

15

20

25

0 20 40 60 80 100

Wo

od

so

lub

iliza

tio

n (

% O

D b

as

is)

Time (s)

770 W 1100W

(a)

(b)

(c)

110 W 

Page 66: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

54 

 

 

 

The effect of power input on the solubilization of wood was also demonstrated in the Figure 4.3.

As the power input increased, the time required for approximately the same amout of wood

dissolution decreased significantly. Maximum wood solubilization obtained at the experimental

conditions studied was about 20%. Time required to obtain this solubilization was found to

decrease from 600 seconds to 180 sec to 60 seconds when the power level increased respectively

from 110 W to 550 W to 1100W. Similarly, the amount of wood solubilized after 60 seconds of

extraction at each power level studied are found to increase and are reported as 8.9%, 13.7%,

15.0%, 15.9%, and 18.6% respectively for 110 W, 330W, 550W, 770W and 1100 W. The

increasd solubilization for the same duration of irradiation or decreased time of irradiation for the

same amount of dissolution was attributed to the rapid increase in the temperature of the reaction

slurry with the increased power level as seen in the Figure 4.4.

Figure 4.4. Temperature of the wood fiber slurry after different microwave irradiation time

0

20

40

60

80

100

120

0 100 200 300 400 500 600 700

Te

mp

era

ture

(oC

)

Time (s)

330W 550W 770 W

1100W 110 W

Page 67: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

55 

 

 

 

4.2.2.2 Yieldofxylan 

It was demonstrated from the wood solubilization study that longer irradiation time at lower

power input as well as shorter irradiation time at higher power input leads to higher dissolution

of wood. A similar trend was expected in the yield of xylan, as more xylan dissolves with the

increase in the wood solubilization. Figure 4.5 shows the effect of irradiation time as well as

power input on the yield of xylan obtained during different extraction conditions. At all the

power levels studied, similar to wood solubilization, yield of the precipitated xylan increased

with time of irradiation (Figure 4.5). However, at higher power inputs, yield of xylan obtained

was found to be leveling out at longer duration of extraction. Moreover, the maximum yield

obtained at higher power input was lower than the lower power input extractions. During

microwave heating the temperature of the reaction media increases due to the direct interaction

of the microwaves with the media. Figure 4.4 shows that a rapid rise in temperature occured

wihin very short time at the higher power inputs compared to the lower power input. It is already

known that wood hemicelluloses are relatively easily hydrolyzed by alkali, and at the same time

undergoes degradation or peeling of the dissolved carbohydrates in the presence of alkali

(Kleppe, 1970). It is anticipated that the in-situ heat generation and the increase in the

temperature of the reaction slurry during microwave assisted extraction leads to these two

opposing phenomena. The rise in temperature increases the diffusion of NaOH and the

hydrolyzed xylan leads to an increase in the hydrolysis rate as well as mass transfer of the

hydrolyzed xylan to the solution. On the other hand, the increase in temperature may also leads

to the degradation of the already dissolved xylan. Yield of the xylan obtained depends on these

two opposing phenomena. At the lowest power level studied, (110 W), yield of xylan increased

Page 68: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

56 

 

 

 

Figure 4.5. Effect of irradiation time and microwave power on the yield of xylan (a) 110W

(b) 330W and 550 W (c) 770W and 1100W

0

10

20

30

40

50

60

70

0 200 400 600 800

Yie

ld o

f x

yla

n (

%)

Time (s)

0

10

20

30

40

50

60

70

0 50 100 150 200 250

Yie

ld o

f x

yla

n (

%)

Time (s)

330W 550W

0

10

20

30

40

50

60

0 20 40 60 80 100

Yie

ld o

f x

yla

n (

%)

Time (s)

770 W 1100W

(a)

(b)

(c)

110 W

Page 69: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

57 

 

 

 

steadily with time of irradiation, from 33% at the mildest condition (60 seconds) to 60% at the

severe conditions (600 seconds). The amount of wood solubilized at these conditions was 9%

and 20% respectively and the temperature of the slurry was 43oC and 96oC respectively. Though

there is degradation of the carbohydrates due to increase in temperature of the alkaline reaction

media (43oC - 96oC), the results indicates that increased hydrolysis rate and transfer of the

dissolved components to the solution is the major process at this low power extraction.

At higher power levels, even though the yield of xylan increased with time of radiation, the

maximum yield obtained was significantly lower than the xylan obtained at the lowest power

level (110W) studied. It was also observed that the yield of xylan obtained decreased with the

increase in the power input. The yield of xylan obtained after maximum duration of irradiation at

each power levels studied are 45, 46 wt% respectively for the power levels of 330 and 550 W

(for 180 seconds), and 38 and 35 wt% respectively for 770 and 1100 W (for 60 seconds). The

amount of wood solubilized at these extraction conditions were 17.4 %, 19.5 %, 16 %, and

18.6% respectively. A comparison of the yield obtained at the lowest power extraction studied

(110 W) indicated that for a wood dissolution of about 15% to 20%, the yield of xylan obtained

varied from 44 % to 60%. Further, comparison of the yield of xylan obtained for the similar

extraction duration (Figure 4.5 b, and 4.5.c) indicates that, the yield obtained was leveled out

faster at the higher power level extractions (550 W for 3 minutes and 1100 W for 1 minute)

compared to the respective lower power inputs (330 W for 3minutes and 770 W for 1 minute).

These results indicate that during extractions using high power input, degradation of the

carbohydrate (xylan) becomes significant compared to the similar duration at the low power

extraction. This can be attributed to the rapid increase in the temperature of the wood slurry as a

result of high intensity of microwave irradiation during the higher power input extraction. As the

Page 70: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

58 

 

 

 

power input increased, the temperature reaches rapidly to about 95-97oC ( 1-3 minutes) and this

might enhance the degradation of the dissolved xylan lowering the yield.

It is expected that the supernatant after the precipitation of xylan contains lower molecular

weight hemicelluloses and/or monomeric sugars and/or other degraded products. Hence, to study

the degradation of the carbohydrates, total sugar content of the supernatant after the precipitation

of xylan was determined using phenol- sulfuric acid method. Figure 4.6 shows the amount of

total sugar in the supernatant obtained after different duration of extractions using different

power inputs. Earlier studies reported that under mild conditions of alkali hydrolysis (below

100oC) dissolution of cellulose is negligible compared to hemicelluloses (Kleppe, 1970) and

hence the dissolved sugar is assumed to be derived from the hemicelluloses present in the wood.

In the figure, the sugar obtained was presented in terms of the total sugar content in the wood as

well as the sugar content in terms of the hemicellulose present in the wood. At all the power

levels studied, total sugar content in the supernatant increased with the time of irradiation (Figure

4.6). The total sugar content in the supernatant at the conditions of maximum wood

solubilization at each power level studied is about 2.9 to 3.3 % of the total sugar content, which

is about 7.6% to 9% of the total hemicelluloses. The results demonstrated that the amount of

total sugar in the supernatant after 600 seconds of irradiation at a microwave power input of 110

W is about the same as the sugar obtained after 180 seconds of irradiation at a power input of

550 W and 60s of irradiation at a power input of 1100 W. Similarly, for the same duration of

irradiation at each power input, the amount of degraded sugar increased with the power level

indicating more degradation of the carbohydrates at higher power input for the same duration.

The degradation rate of xylan increases with the increase of the power input, which is due to the

Page 71: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

59 

 

 

 

rapid rise of temperature and this leads to the lower yield of xylan at higher power microwave

extraction. These results demonstrated that lower microwave power input and longer duration of

irradiation will be the most suitable processing parameters for the efficient extraction of xylan.

Figure 4.6. Effect of microwave power and irradiation time on the dissolved sugar content of

the liquid phase after the precipitation of xylan (A and B represents sugar content based on the total sugar content; A1 and B1represents sugar content based on the hemicelluloses)

Page 72: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

60 

 

 

 

4.2.2.3 Dissolvedlignin

Dissolution of lignin along with the hemicelluloses is also expected under alkaline extraction

conditions. Percentage of dissolved lignin was calculated on the basis of the lignin originally

present in the wood. As expected, at all the power levels studied, amount of dissolved lignin

increased with the increase of irradiation time (Figure 4.7). At a power output of 110 W,

amount of soluble lignin showed an increase with the time (5% at 2 minute vs. 10% at 10

minutes) of irradiation indicating the dependency of hydrolysis on the time and temperature. At

all power levels, amount of dissolved lignin increased at a particular irradiation time and the

results are in accordance with the high dissolution of wood as explained in the previous section.

Figure 4.7. Effect of microwave power and irradiation time on dissolution of dissolved lignin at 110 W, 330 W, 550 W, 770 W, and 1100 W (Secondary axis are for 110 W as shown by the arrows)

0 100 200 300 400 500 600 700

0

2

4

6

8

10

12

0

2

4

6

8

10

0 50 100 150 200

Time, (s)

Dis

solv

ed li

gnin

con

ten

t, %

Dis

solv

ed l

ign

in c

onte

nt,

%

Time, (s)330W 550W 770W 1100W 110 W

Page 73: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

61 

 

 

 

4.2.3 MaterialBalance

A material balance analysis of the wood fiber extracted at different power inputs for the

maximum duration at each power input has also been performed to demonstrate the difference in

the yield of xylan obtained at different power levels and is given in table 4.3. It is noticed from

the table that as the power input or energy increased, the difference in the mass balance is also

increased and this may be due to the formation of other degradation products, such as sacharinic

acid from sugar degradation, formed as a result of the rapid generation of high temperature.

Higher amount of solubilized wood, lower yield of precipitated xylan and higher amount of

degraded products for shorter duration of microwave irradiation at higher power levels indicates

degradation of the carbohydrates.

Table 4.3. Material balance analysis of wood fibers after microwave assisted extraction Material 110W, 10min 330W, 3 min 550W, 3 min 770W, 1min 1100W, 1min Initial Wood fiber (g) Solid residue after extraction, (g) High mol.wt. xylan (g) Soluble lignin (g) Total soluble sugar after precipitating xylan (g) Total (g)

100 80.89 ± 0.66 13.31 ± 0.21 2.51 ± 0.06 2.07 ± 0.12 98.78 ± 0.71

100 82.70 ± 0.67 9.96 ± 0.85 1.95 ± 0.02 1.69 ± 0.11 96.3 ± 1.08

100 81.49 ± 1.35 10.72 ± 0.68 2.01 ± 0.01 1.78 ± 0.04 96 ± 1.5

100 84.07 ± 0.33 8.45 ± 0.17 1.76 ± 0.17 1.36 ± 0.05 95.64 ± 0.41

100 81.39 ± 1.33 9.73 ± 0.87 1.73 ± 0.28 1.77 ± 0.06 94.62 ± 1.61

Page 74: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

62 

 

 

 

 

4.2.4 Characterizationofwood:CrystallinitystudyusingX‐raydiffraction

X-ray diffraction (XRD) is one of the primary tools used in the determination of the crystallinity,

conformation and structure of cellulose microfibrils (Segal et al., 1959; Foner and Adan, 1983;

Cave, 1997; Borysiak and Doczekalska, 2005; Jiang et al., 2007). Crystallinity of fibers is one of

the structural characteristics related to the strength of the fibers and helps to determine the post

application of the extracted fibers. XRD analysis was performed to see if the microwave

extraction made any changes in the crystallinity of wood fibers. The XRD pattern of wood fibers

and wood fibers after different duration of microwave extraction using 110 W are shown in the

Figure 4.8. Two major diffraction peaks are observed in all the samples approximately at 15o

and 22o which corresponds to the 101 and 002 crystal planes, respectively. Variations are

Figure 4.8. X-ray crystallographs of wood fiber before and after different duration of

microwave extraction (Power level: 110 W)

observed in the XRD pattern of wood fibers in the peak maxima (l002) and the minima around 2θ

of 18o (IAM) after both extraction. This might be due to the changes in the crystallinity of the

cellulose in the fibers. Relative degree of crystallinity determined using the equation 3.7 for

wood, for different periods of microwave extraction (2, 6, 10 minutes) is 52.5±1.3%, 56.6±2.3%,

0

200

400

600

800

1000

1200

1400

0 10 20 30 40 50

Inte

ns

ity

(a.u

)

2 (degree)

wood PL1-2min PL1-6min PL1-10min

Page 75: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

63 

 

 

 

58.8±1.9% and 60.2±1.8% respectively. Percentage crystallinity increased after the extraction

and is because of the hydrolysis of the amorphous region of the wood, namely hemicelluloses

and part of the lignin. The results indicates that even up to 10 minutes of microwave extraction

affected mainly the amorphous matrix of the wood fibers and did not have a significant effect on

the crystallinity of the cellulose and hence the strength of the fibers. Hence, the fibers after

extraction would be good resource for production of cellulosic pulp fibers, as well as

reinforcements in composites.

4.2.5 Characterizationoftheprecipitatedxylan

 

4.2.5.1 Sugarcomposition

The composition of neutral sugars and the lignin content of the selected number of extracted

xylan at different conditions of microwave extraction are given in the table 4.4. It is obvious that

the isolated xylan contains about 68 - 88% of xylose and minor amounts of glucose (about 0.1%)

indicating the extracted polymer is xylan. It is hard to isolate xylan without bound lignin from

the woody biomass and the xylan extracted contains about 9-12 % of lignin.

4.2.5.2 FTIRspectroscopy 

FTIR spectra of the precipitated xylan after different periods of microwave irradiation were

taken for a comparison of the structural features of the xylans isolated under different conditions,

and are shown in the Figure 4.9. All the spectral profiles were similar indicating the isolated

xylans under different conditions of extraction are similar. Two strongest absorption bands were

Page 76: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

64 

 

 

 

Table 4.4. Chemical composition of xylan extracted using different experimental conditions

Extraction Conditions

Xylose (%) Glucose (%)

Klason Lignin (%)

Total lignin (%)

110 W 120 s 240 s 360 s 600 s

88.3 78.1 74.8 73.9

0.08 0.07 0.05 0.06

7.0 9.5 9.2 9.1

9.6 9.7 10.8 11.1

330 W 60 s 120s 180s

74.9 74.0 71.4

0.07 0.07 0.06

9.3 9.8 6.7

9.8 12.1 12.6

550 W 60 s 90 s

79.0 74.8

0.08 0.06

7.3 9.2

9.2 12.8

770 W 40 s 60 s

77.3 76.0

0.05 0.03

7.5 9.4

10.2 10.8

1100 W 40 s 60 s

78.4 68.5

0.04 0.04

8.1 9.4

10.5 12.3

 

observed at 3700-3000cm-1 and 1200-1100 cm-1 in all the spectra of the examined xylans, and

these absorption bands were due to C-O-H vibrations. The broad intense band with a maximum

around 3400 cm-1 correspond to O-H stretching vibrations; the complicated absorption band in

the region of 1200-1000 cm-1 with a principal maximum at 1043 cm-1 shows the C-O stretching

vibrations (Marchessault and Liang, 1962; Sun et al., 2004; Buslov et al., 2009; Burnov and

Mazza, 2010). The two absorption maxima observed at around 2930 and 2870 cm-1 correspond

to asymmetric and symmetric stretching vibrations of CH2 groups. A polymeric chain formed

from pure xylanopyranose units should not have noticeable peaks in the absorption region of

Page 77: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

65 

 

 

 

Figure 4.9. FTIR spectra of xylans obtained at different extraction conditions

1800 -1500 cm-1. Hence, the absorption peaks in this region was due to the vibrations of different

substituents in the main polymeric chain. All the spectra were similar in the region of 1300 cm-1

to 800cm-1. The peaks observed at around 1600-1620cm-1 and 1420cm-1 indicates the presence of

carboxylate ions [ νas(COO-) =~ 1600 cm-1; νs (COO-) = 1425 cm-1) (Marchessault and Liang

1962; Buslov et al., 2009 ) present in the xylan and may be due to the presence of uronic acid on

the side chain. A shoulder peak was observed at around 1595cm-1 in the spectra of xylans

Page 78: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

66 

 

 

 

obtained at power level 110W after 4minutes, and all the xylans obtained from higher power

levels indicating the presence of a small amount of lignin residues in the precipitated xylan. A

sharp peak observed in all spectra at 1160-1170 cm-1 corresponds to the C-O-C vibrations in the

xylan and the peaks at 1252 cm-1 relates to the OH in-plane bending. The sharp peak observed

at around 896 cm-1 was due to the characteristic ring frequency of the xylanopyranose units

(Marchessault and Liang, 1962).  

4.2.5.3 Molecularmass

The SEC chromatograms of the precipitated xylans were multimodal in nature (Figure 4.10). The

multi-model nature of xylan might be due to the presence of different components in the xylan,

including the aggregated xylan, non-aggregated xylan, and or lignin-carbohydrate complexes etc.

Figure 4.10. Typical SEC signal for xylan obtained by microwave assisted method of

extraction

Page 79: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

67 

 

 

 

Table 4.5. Molecular mass distribution of xylan extracted using different experimental conditions

Xylan sample ID

*Molar mass

D

Relative %

*Molar mass

D

Relative %

*Molar mass

D

Relative %

110 W 4 min 8 min 10 min

21 x 105 20 x 105 20 x 105

36.5 38.8 25.1

19 x 104 19 x 104

19 x 104

38.1 39.1 35.4

6 x 103 6 x 103 6 x 103

25.3 22.2 39.5

1 minute 330 W 550 W 770 W 1100 W

21x105 20x105 22x105 22x105

35.8 20.9 18.4 13.0

21 x104 19 x104 21 x104 22 x 104

39.4 22.8 24.8 12.4

6 x 103 6 x 103 6 x 103 6 x 103

24.8 56.3 56.8 74.6

* Molar mass in terms of dextran equivalents

Similar studies were reported earlier (Jacobs and Dahlman, 2001; Bikova and Treimanis, 2002).

Molecular mass (Mw) corresponding to the peaks in the chromatograms as dextran equivalents

with the respective relative percentage of each fraction is given in table 4.5. Different molecular

mass fractions observed are of the order of 105, 104, and 103 Da. The very high molecular mass

fractions may be due to the aggregation of the xylan fractions, and the lowest mass fraction is of

the non-aggregated xylan. Xylan extracted using microwave process consists of about 60-75% of

high molecular mass fraction (105-104) and 20-40% are low molecular mass fractions (103). As

the time of extraction increased from 4 minutes to 10 minutes, the high molecular mass fraction

decreased from 36 to 25%, while the low molecular mass fraction increased from 21 to 30%,

Page 80: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

68 

 

 

 

indicating a change in the molecular structure. A similar change was observed with the increase

of microwave power level indicating degradation of the polymer with the higher power input.

4.3 Conclusions

The results of the microwave assisted extraction demonstrated that a short duration of micro

wave irradiation of the wood slurry (10 min) leads to an extraction yield of 60% based on the

original xylan present in the wood. At all the higher power levels studied yields of extracted

xylan was found to increase with the time of extractions (between 10 s – 180 s). However, the

maximum yield obtained was decreased with the increase of power level and may due to the

degradation of the xylan occurred as a result of a rapid increase in the temperature within a short

duration of time. Lower power level and longer duration of irradiation (110 W, 10 min) was

found to be the most suitable extraction conditions studied for a better yield of xylan. Sugar

composition and FTIR spectra indicated that the extracted polymer structure mainly consisted of

a backbone of (1-4) β-D-xylopyransoyl residues. Comparison of molecular mass demonstrated

relatively higher amount of higher molecular fractions for microwave extracted xylan can be

obtained with low power input compared to the high microwave power.

Page 81: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

 

 

69 

 

Chapter5 Investigationofthemechanismofmicrowaveassistedalkalineextractionofbirchwood

  

 

5.1 Introduction

Microwave heating is known to be an efficient and environmentally friendly alternative to

conventional heating where the heating occurs through the direct interaction between the

electromagnetic energy and the material of interest (Kappe, 2008). Advantages of using

microwave energy as an alternative method to conventional heating include (i) acceleration of

reaction rates, due to the efficient internal heating produced by the direct coupling of microwave

energy with the molecules present in the reaction mixture against conduction or convection

heating, where energy is transferred based on the intra molecular interaction and (ii) more

specific and selective reactions, due to the preferential absorption of energy by the target

compounds with high dielectric loss and high polarity in the reaction mixtures (Newnham et al.,

1991; Venkatesh & Raghavan, 2004; Kappe, 2008).

From Chapter 4, it was found that a low power input microwave assisted alkaline extraction is an

efficient method of extraction of xylan from birch wood. In this chapter the research emphasize

on two objectives. The first objective was to compare the efficiency of microwave assisted

extraction in comparison with the conventional method of extraction and to investigate if there is

any "microwave effects" underlying during the microwave extraction compared to the

conventional method of extraction. The second objective was to investigate the mechanism

involved in the microwave assisted alkaline extraction by studying the physico-chemical changes

Page 82: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

70 

 

 

 

of the wood after extraction. Two different approaches were used to accomplish the research

objectives. In the first approach, characteristics of the low power input microwave assisted

extraction (110 W) were compared with a conventional isothermal extraction at 90oC. The

temperature selected for the conventional extraction was based on the maximum temperature

generated during the microwave extraction. The maximum temperature generated during

microwave extraction at which comparatively high yield of xylan obtained was in the range of 86

- 96oC (for 6-10 minutes of duration) (Chapter 3), and hence we performed the conventional

extraction at 90oC. Rate of wood dissolution and rate of formation of xylan were compared to

establish the efficiency of the microwave extraction over conventional method of extraction. In

the second approach, conventional extraction was performed for the same duration as that of

microwave extraction at different temperatures to differentiate the effect of temperature effects

from the microwave effects. The temperatures selected for conventional extraction was the final

temperature of the slurry obtained after the same duration of microwave assisted extraction.

Comparison of wood dissolution and yield of xylan was again performed to see if there is any

difference in the extraction process. In addition to this, physical and chemical changes in the

woody biomass after extraction was analyzed to establish the mechanism involved in the

processes. Characterization of xylan obtained was also performed to evaluate if there is any

differences in the structure as a result of the difference in the heating mechanisms.

5.2 Methods

In the first part of the study, 3 g of wood fiber (OD basis), supplied by St. Mary's Paper

Company, Ontario, Canada, was extracted using 30 mL of 4 wt% NaOH solution (solid to liquid

ratio of 1:10 g:mL) using microwave and conventional methods of extraction. Experimental

Page 83: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

71 

 

 

 

conditions used are given in the table 5.1. The procedures used for the extraction and separation

of xylan are described in Chapter 3.

Table 5.1. Experimental conditions Method of extraction Microwave (110 W) Conventional (90oC) Time, (min) 0.5, 1, 2, 4, 6, 8, 10, 12, 15, 18 5,10, 30, 90, 120, 150, 180, 240

Time temperature combination of the microwave assisted and conventional extraction used in the

second part of the study are given in table 5.2. The temperatures selected for each conventional

extraction performed was the final temperature of the slurry obtained after the same duration of

microwave assisted extraction. Since conventional heating has a long thermal lag time, NaOH

solution was preheated to the required temperature before being mixed with the wood fibers

(which is previously weighed and kept at 40oC). The protocol used for the separation of wood

fibers and precipitation of xylan was same as the procedure described in Chapter 3. In this study,

5 g of the wood fibers (OD basis), prepared from birch wood logs obtained from premises near to

the University of Toronto, were extracted using 50 mL of 4 wt% NaOH solution ((solid to liquid

ratio of 1:10 g:mL). A higher amount of material (volume of solution) was used in this study to

prevent evaporation of the solution due to longer duration of extraction.

Table 5.2. Time-temperature combinations used in the microwave and conventional extractions *Microwave Conventional (isothermal) Temperature , (oC) 37,60, 80, 95, 98,100 37, 60, 80, 95, 98, 100 Time, (min) 1, 5, 10, 20, 30, 40 1, 5, 10, 20, 30, 40 * final temperature of the slurry after extraction

 

Page 84: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

72 

 

 

 

5.3 ResultsandDiscussion

5.3.1 Comparisonofmicrowaveandconventionalalkalineextraction:Wood solubilization

During alkaline extraction of wood, the linkages between the hemicelluloses and lignin undergo

hydrolysis and the hydrolyzed components dissolve into the solution. The dissolved xylan was

separated from the solubilized part of wood by precipitation using acetic acid. The percentage of

solubilized wood and the yield of precipitated xylan were calculated using the equation 4.1 and

4.2 respectively for different duration of extraction under two different processes and are shown

in the Figure 5.1. It is clear from the figure that the time required for a definite amount of

Figure 5.1. Percentage of wood solubilized and yield of xylan after microwave and conventional extraction

0

10

20

30

40

50

60

70

0 50 100 150 200 250 300

Wo

od

so

lub

iliza

tio

n &

Yie

ld o

f x

yla

n (

%)

Time (Minutes)

Wood dissolution (M) Yield of xylan (M) Wood solubilization (C) Yield of xylan (C)

Page 85: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

73 

 

 

 

solubilization of wood during microwave assisted extraction is significantly lower compared to

the conventional extraction. This is attributed to the difference in the heating mechanism

involved in the two processes. Microwave heating is volumetric and rapid as the microwave

heats the target compound through the direct interaction of the objects with the applied

electromagnetic field, whereas, in conventional heating, heat transfers from the heating device to

the sample in a rather slow and homogeneous manner through conduction and/ or convention,

which is based on inter- and intra-molecular heat transfer. The rapid temperature rise in the

system can accelerate the hydrolysis reaction and hence the faster dissolution during microwave

assisted extraction.

For a detailed analysis of the effect of heating on the solubilization of wood during each

extraction process, the graphs of wood solubilization was re-plotted as in the Figure 5.2. The

amount of wood dissolved during both the process can be empirically fitted by two linear

relations as shown in the figure. It is clear that solubilization of wood can be considered as two

stages under the experimental conditions studied; where in the first stage a fast dissolution occurs

(about 2 minutes for microwave and about 10 minutes for conventional extraction) and is

followed by second stage where in a steady increase in the dissolution occurs. The faster initial

stage dissolution is believed to be due to the dissolution of the surface debris present on the

fibers during the fiber preparation. Removal of the debris present on the fiber is clear from the

scanning electron microphotographs (Figure 5.3) of the fibers before and after extraction. The

fiber surface becomes smooth after both extraction processes. The slower second stage

dissolution is believed to be the result of various mass transfer processes during alkaline

hydrolysis of the wood fibers. The major mass transfer processes during alkaline extraction

Page 86: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

74 

 

 

 

include hydrolysis of hemicellulose-lignin linkages and their dissolution, diffusion of NaOH to

the fibers to replace the alkali consumed by the hydrolysis and diffusion of the hydrolyzed

components (hemicelluloses and lignin) to the fibre surface and dissolution to the mass solution.

Figure 5.2. Comparison of the amount of wood solubilized during microwave and

conventional extraction. (a) Microwave extraction; (b) Conventional extraction

y = 10.4xR² = 0.8256

y = 1.2074x + 9.1817R² = 0.982

0

5

10

15

20

25

30

35

0 5 10 15 20 25 30

Wo

od

so

lub

iliza

tio

n (

%)

Time (Minutes)

y = 1.56xR² = 0.8342

y = 0.0625x + 14.603R² = 0.9656

0

5

10

15

20

25

30

35

0 50 100 150 200 250 300

wo

od

so

lub

iliza

tio

n (

%)

Time (Minutes)

(b) 

(a) 

Page 87: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

75 

 

 

 

The results pointed out that wood solubilization during both stages of dissolution using

microwave assisted extraction are faster than the conventional extraction. The slopes of the two

linear portions of the graph provide the rate of dissolution of wood during hydrolysis. Rate of the

dissolution in the first stage , (Y/t)Microwave = 10.4 >> (Y/t)Conventional = 1.56, where Y is the

percentage of wood solubilized and t is the time in minutes. Comparison of the slopes at this

stage indicates that wood solubilization during microwave extraction is about 6.6 times faster

than the isothermal conventional extraction. Similarly, slope of the second stage dissolution,

(Y/t)Microwave = 1.21 >> (Y/t)Conventional = 0.06, indicates that rate of the second stage

dissolution of wood during microwave assisted extraction is 19 times faster than the

conventional extraction.

Difference in the rate of wood solubilization can be explained in terms of the difference in the

heating mechanism. During microwave extraction, temperature of the system increases rapidly

because of the direct interaction of the microwave with the NaOH solution compared to the

conventional extraction. The rapid increase in temperature increases the initial dissolution

compared to the conventional extraction. Though there is rise in temperature of the system

during microwave extraction that enhances the hydrolysis reaction and the diffusion, this might

not be sufficient to enhance the diffusion of the hydrolyzed components to the solution during

the early stages. This explains the comparatively smaller difference (6.6 times in the early stage

compared to 19 times in the later stage) in the observed rates of wood dissolution during the

early stage of the two different processes. As the duration of extraction increased, temperature of

the slurry increases and enhancement of all the three mass transfer processes mentioned above

occurs and leads to an increased dissolution. Temperature measured at the end of each

Page 88: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

76 

 

 

 

Table 5.3. Temperature of the slurry after different duration of microwave extractions

Time (Minutes) Temperature (oC)

0.5 1 2 4 6 8 10 12 15 18

35.6 ± 0.9 43 ± 0.5 56.5 ± 0.4 65.7 ± 0.8 86.3 ± 0.1 89.7 ± 0.1 96 ± 0.6 97.0 ± 1.0 98.2 ± 0.3 99.5 ± 0.2

microwave extraction performed is given in table 5.3. The slurry temperature varied from 56oC

to 100oC during the second stage of wood dissolution in the microwave extraction. The

conventional extraction performed is an isothermal extraction at a temperature of 90oC, which is

closer to the temperature generated after 8 minutes of microwave extraction. Considerably

higher rate of dissolution observed at this stage during microwave extraction can be attributed to

the rapid increase in temperature of the fiber slurry which increase the hydrolysis and the

dissolution of the wood components. The maximum amount of wood dissolved during extraction

is about 29-32% under the experimental conditions studied using conventional and microwave

extraction. The time required to obtain the maximum dissolution is 240 minutes in conventional

extraction whereas it took only 18 minutes in microwave extraction indicating significant

reduction (about 13 times) in the extraction duration using microwave heating. It is anticipated

that about 10 degree rise in temperature double or triple the rate of the reaction, and by changing

the temperature from 90oC to 100oC, the rate of dissolution can be doubled or tripled. However,

the rate of dissolution of wood in the second stage of microwave extraction is significantly

Page 89: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

77 

 

 

 

higher (19 times) than conventional extraction. In most of the microwave heating applications

spectacular accelerations was observed, due to the efficient internal heating produced by the

direct coupling of microwave energy with the molecules present in the reaction mixture

(Kingston and Haswell, 1997; Kappe, 2008). Lewis et al. (1992) observed enhanced reaction

kinetics during microwave assisted chemical reactions and the authors attributed this

phenomenon to the non-uniform energy distribution at the molecular level during the chemical

reactions and are reported as the major microwave effects produced. It was assumed that,

microwave heating can induce “hot spots”, in the biomass due to the selective interaction of the

microwaves with the inhomogeneous lignocellulosic materials, and disrupt the structure of the

materials. For example, Hu and Wen, (2008) found that microwave treatment enhances surface

disruption and breaking of lignin structures in alkaline treated switch grass, whereas

conventional heating the fibers remain intact. They also reported that microwave treated switch

grass undergoes easy enzymatic saccharification to produce sugars compared to the conventional

treated one. These studies leads to the speculation that the significant enhancement observed in

the rate of wood dissolution or reduction in the time of extraction could be due to the combined

effect of thermal and the "microwave effects” caused during the microwave extraction. It is

hypothesized that the selective heating feature of microwaves with the lossier component in the

reaction system leads to direct interaction of the microwaves with the alkali present in the fibers,

and the temperature rise inside the fibers produces an “explosion” effect in the wood fibers. This

leads to the rupture of the fiber structure and facilitates the hydrolysis and diffusion of the

hydrolyzed components to the solution. Morphological study of the fibers before and after

extraction (discussed below) supported this hypothesis.

Page 90: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

78 

 

 

 

Figure 5. 3. SEM photomicrographs of wood fiber before and after extraction. (a) before extraction (b) after conventional extraction (90 minutes at 90oC) under two different magnifications (c) after microwave extraction (10 minutes at a power level of 110 W) under two different magnifications

SEM photomicrographs of the wood fibers before and after both extraction processes are shown

in Figure 5.3. Figure 5.3a represents the fiber before extraction. Fiber surface becomes smooth

after both extractions indicating the dissolution of debris present on the fibers. After

conventional extraction, no significant change is noticed in the overall fiber structure (Figure

5.3b1 and b2), whereas the fibers after microwave extraction becomes more porous (Figure 5.3c1

(a)

(b1)  (b2)

(c1)  (c2)

Page 91: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

79 

 

 

 

and c2.). The more porous structure leads to enhanced hydrolysis, faster diffusion and dissolution

of the hydrolyzed components. An attempt to quantify the micro structural changes has also been

performed and will be discussed in Chapter 6.

Though there is difference observed in the rate of wood solubilization during conventional and

microwave assisted extraction it is essential to have similar experimental conditions for both

processes to accurately differentiate the two processes. So, another set of conventional

experiments were performed at different temperatures for the same duration as that of microwave

extraction to separate the temperature effects from the microwave effects. In this part of the

study different microwave extractions were performed for different extraction duration (time of

extraction vary from 1 minute to 40 minutes) as given in the table 5.2. Temperature of the wood

fiber slurry was noted after each microwave extraction. Conventional experiments were

performed for the same duration as that of microwave extraction, in an isothermal extraction at a

particular temperature, which is the final temperature of the slurry after the microwave extraction

for that particular time of extraction (Experimental conditions are given in table 5.2). Further, to

decrease the temperature gradient occurred during the conventional heating of the wood slurry,

and for faster attainment of the temperature, the weighed wood fibers were kept at 40oC and

NaOH solution was preheated to the required temperature before mixing with the wood fiber.

The amount of wood solubilization, calculated using the equation 4.1, during microwave assisted

and conventional alkaline extraction of wood fibers at different duration of extraction is shown in

Figure 5.4. The results indicated the wood dissolution behavior is similar in both cases. Again,

for shorter duration the difference in the wood dissolution is not separable from each other and

this confirms that early stage wood dissolution was mainly due to the dissolution of the surface

Page 92: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

80 

 

 

 

 Figure 5. 4. Effect of microwave and conventional extraction of xylan on the solubilization of

wood

debris on the fiber. As the duration of extraction increased, temperature of the slurry increased

and the rate of wood dissolution and amount of solubilized wood increased in both extractions.

Higher wood dissolution was expected during conventional extraction, as the temperature for the

extraction was kept constant at that particular temperature, whereas in microwave assisted

extraction, the temperature of the slurry increased fast and reached to the final temperature. The

rate of wood solubilization in microwave extraction [(Y/t)Microwave = 0.45] is greater than the

rate of wood dissolution in conventional extraction [(Y/t)Conventional = 0.39] . The rate of wood

dissolution and the amount of solubilized wood in microwave extraction would have been the

same or lower than the conventional extraction if the wood dissolution is only because of the

increase in the temperature. However, it was found that the rate of dissolution and the amount of

wood dissolved for particular duration is higher in the microwave extraction. This confirms the

hypothesis that the synergic effect observed during microwave extraction is not only due to the

y = 0.4508x + 16.37R² = 0.9911

y = 0.3888x + 16.09R² = 0.9608

0

5

10

15

20

25

30

35

40

0 10 20 30 40 50

Wo

od

dis

so

luti

on

(%

)

Time (minutes)

Microwave Conventional

Page 93: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

81 

 

 

 

temperature effect on the hydrolysis and wood solubilization, but also due to the structural

changes caused by the "explosion effect" that favors hydrolysis and mass transfer of the

hydrolyzed components to the solution. The hypothesis will be verified later by analyzing the

physical structural changes of the wood fibers after the two different extraction processes.

5.3.2 Comparisonofmicrowaveandconventionalalkalineextraction:Yieldofxylan

The yield of precipitated xylan calculated using the equation 4.2 for different duration of

extraction under two different experimental conditions (table 5.1) is shown in the Figure 5.1. It is

clear from the figure that under these experimental conditions, time required to extract a definite

amount of xylan during microwave assisted extraction is significantly lower compared to the

conventional extraction. In order to understand how conventional and microwave extraction

affects the yield of xylan, the graphs were re-plotted with time as done in the case of wood

dissolution and is shown in Figure 5.5. The figure pointed out that the yield obtained during

both extractions followed a similar path; increased very fast initially (stage 1), then a slow and

steady increase (stage 2), which is followed by a decrease in yield (stage 3). Yield of xylan

during alkaline extraction depend on the following simultaneous and multiple reactions and or

processes involved in the system; (i) hydrolysis of easily hydrolysable fraction of xylan and their

dissolution to the solution (ii) hydrolysis of xylan from the relatively tougher zone of the fibers

(iii) degradation (peeling) of the xylan to low molecular weight fractions and sugars, and (iv)

further degradation of the sugars to other products such as sacharinic acid. The yield of xylan

obtained during both the processes under the experimental conditions can be explained with

these different reactions involved and are discussed below.

Page 94: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

82 

 

 

 

Figure 5.5. Comparison of the yield of xylan (based on the total xylan) obtained during microwave and conventional extraction. (a) Microwave extraction; (b) Conventional extraction

y = 33xR² = 0.9555

y = 3.0537x + 29.956R² = 0.9834

y = ‐0.5949x + 64.917R² = 0.899

0

10

20

30

40

50

60

70

0 5 10 15 20 25 30

Yield of xylan (%

)

Time (Minutes)

y = 3.312xR² = 0.9071

y = 0.3045x + 30.637R² = 0.9027

y = ‐0.1045x + 66.303R² = 0.951

0

10

20

30

40

50

60

70

0 50 100 150 200 250 300

Yie

ld o

f x

yla

n (

%)

Time (Minutes)

(a)

(b) 

Page 95: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

83 

 

 

 

It is clear from the Figure 5.5 that a portion of xylan can be extracted easily within a short

duration of extraction (within 1-2 minutes for microwave extraction and within 10 minutes

during conventional extraction). This represents the yield obtained from the susceptible or easily

hydrolysable fraction of xylan present in the birch wood. This easily accessible fraction of the

xylan obtained during the initial stages in both processes of extraction is about the same and is

found to be around 32%. After this period of extraction, yield of xylan obtained increased with

the duration of extraction up to about 10 minutes in microwave heating (Figure 5.5. a) and

decreased thereafter. It is expected that after the removal of easily accessible xylan, hydrolysis

and dissolution of the relatively inaccessible xylan and the degradation of the already dissolved

xylan occur simultaneously and the yield obtained is possibly due to the result of these two

processes. The observed increase in the yield of xylan up to 10 minutes of microwave extraction

indicates that at this stage, hydrolysis and dissolution of the xylan is higher compared to

degradation of the dissolved xylan. Temperature of the wood slurry increased from 56oC to about

96oC during this period. A similar observation was found during 10 to 90 minutes of

conventional extraction. Lower yield of xylan after 10 minutes of microwave extraction

demonstrates that during this period, degradation of xylan is more compared to the hydrolysis

and dissolution. This could be due to the longer duration of exposure of the xylan at high

temperature (> 95oC). Such a decrease was observed in the conventional extraction at about 120

minutes of extraction (Figure 5.5.b).

Similar to the wood solubilization, a comparison of the slope of the different stages of extraction

was analyzed to investigate the effect of heating on the extraction of xylan. It was found that

rate of extraction of xylan in microwave assisted process is significantly higher than the

Page 96: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

84 

 

 

 

conventional extraction. The rate of xylan extraction in the first and second stages in the

microwave and conventional extraction is as follows:

(Y/t)Microwave, stage 1 = 33 >> (Y/t) Conventional, stage 1 = 3.3 and

(Y/t)Microwave, stage 2 = 3.05 >> (Y/t)Conventional, stage 2 = 0.31 ,

where Y is the percentage yield of xylan calculated based on the amount of xylan present in the

raw wood, and t is the time of extraction in minutes. At both stages the rate of extraction of xylan

is almost 10 times higher than the conventional extraction. The increased rate of the xylan

extraction in the microwave assisted process can be explained using the selective heating ability

of the microwave energy and the resulted exploded structure as in the wood dissolution.

However, the sudden increase in temperature may also leads to the degradation of the already

dissolved xylan and hence the yield of high molar fraction of xylan. The degradation becomes

more prominent for longer duration, as the temperature of the slurry is higher and leading to a

lower yield of xylan. It is found that maximum yield is obtained at about 10 minutes and

thereafter, the yield is decreased during microwave assisted extraction. Similarly in conventional

extraction, maximum yield is obtained at about 90 minutes and thereafter the yield of xylan is

decreased.

The degradation of the dissolved xylan was confirmed by quantifying the total sugar content of

the supernatant after precipitating the xylan (Figure 5.6). The Figure shows that as the duration

of the extraction increased the sugar content left in the supernatant also increased and is

attributed to the degradation of the high molecular weight xylan. However, the difference in the

dissolved sugar content obtained in both processes was found to be more or less similar. As a

Page 97: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

85 

 

 

 

result of longer duration extraction at 90oC, there may be degradation of the carbohydrates that

not only leads to sugar but also to other products such as sacharinic acid, which are not

quantified in the process, and this might be the reason for a low xylan yield.

Figure 5.6. Sugar content of the supernatant after precipitating the extracted xylan using two different processes

Mass balance analysis of the wood fibers after the two extraction processes, where the maximum

yield of xylan obtained has also been performed (table 4.3) to demonstrate the degradation of

xylan to other products. The wood residue left after both extractions was found to be more or

less similar, indicating a similar amount of wood dissolution. However, the mass balance

difference between the original fiber and the total amount after extraction is higher in the

conventional extraction (4.2 g) compared to microwave extraction (1.22 g). The increased

0

1

2

3

4

Su

ga

r Co

nte

nt,

%

Time, minutes

90C110W

10 30 60 90 120110W: 2 4 6 8 10

Page 98: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

86 

 

 

 

difference observed in the mass balance of the fibers after conventional extraction could be due

to the degradation of the xylan to other products as explained above.

Table 5.4. Mass balance analysis of wood fibers after xylan extraction

Material Microwave (10 minutes)

Conventional (90 minutes)

Original wood fiber (g) Wood residue left after extraction(g) Xylan Precipitated (g) Lignin dissolved (g) Total sugar left in the supernatant (g) Total (g)

100 80.89 ± 0.66 13.31 ± 0.21 2.51 ± 0.06 2.07 ± 0.12 98.78 ± 0.71

100 79.9 ± 0.5 12.1 ± 0.8 2.0 ± 0.2 1.8 ± 0.1 95.8 ± 1.0

Figure 5.5 also shows that the rate of degradation was higher in the microwave extraction

((Y/t)Microwave, stage 3 = 0.595 ) compared to conventional extraction ((Y/t)Conventional, stage 2 =

0.104) . This could be attributed to the higher temperature (above 95oC) generated during

microwave extraction after 10 minutes. Xylan may undergo faster degradation at this high

temperature compared to the xylan obtained using the conventional extraction performed at

90oC. Degradation of xylan at high temperature was further confirmed by determining the degree

of polymerization of the xylan and will be discussed in the section 5.3.4.3.

For further differentiating the two process of extraction, the yield obtained from the two different

extraction processes for different duration of extraction at different temperatures (experimental

conditions are given in table 5.2) were compared. The yield of xylan calculated using the

equation 4.2 for the microwave and conventional extraction with the respective duration of

Page 99: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

87 

 

 

 

extraction is shown in Figure 5.7. The yield obtained was higher for the microwave assisted

process and the difference is more prominent for longer duration. The higher yield of xylan

indicates that microwave heating enhances the faster dissolution of xylan through the more

loosened fiber structure. This again due to the difference in the heating mechanism involved in

microwave and conventional extraction, as explained in 5.3.1.

Figure 5.7. Effect of microwave and conventional extraction of xylan on the yield of xylan

The yield obtained in both extraction processes followed an increase in the yield for about 20-25

minutes of extraction and the yield is decreased thereafter. The temperature of the slurry was

above 95oC after 20 minutes of extraction and the decrease in the yield after this again indicates

the enhanced degradation of the dissolved xylan at high temperature. However, the decrease in

yield is more significant in conventional extraction compared to microwave extraction. This

could be possibly due to the increased degradation of xylan caused by the prolonged exposure to

R² = 0.9756

R² = 0.9947

0

10

20

30

40

50

60

70

1 5 10 20 30 40

Yie

ld o

f x

yla

n (

%)

Time (minutes)

Microwave

Conventional

Page 100: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

88 

 

 

 

higher temperature (>95oC), as the extraction is held at that particular temperature. Unlike

conventional extraction, during microwave extraction the temperature is reaching rapidly to that

temperature favoring the dissolution of xylan rather than degradation. These results indicate that

microwave heating leads to a higher yield of xylan compared to conventional heating.

Figure 5.8. Effect of temperature on the yield of xylan

Figure 5.8 shows the dependence on the yield of xylan with the temperature, where the yield

obtained from two different processes was plotted against temperature. In both processes the

yield obtained could empirically fitted to a logarithmic trend: yield increase is lower with the

increase in temperature and is due to the degradation of the dissolved xylan. As the duration of

extraction increased, the temperature of the media also increased and leads to two opposing

phenomena; (i) enhanced hydrolysis and dissolution and (ii) degradation of the dissolved xylan

as explained in the previous section. As explained above, increased difference in the yield of

xylan obtained at longer duration could be possibly due to the increased degradation of xylan

R² = 0.9873

R² = 0.9896

0

10

20

30

40

50

60

70

0 20 40 60 80 100 120

Yie

ld o

f xy

lan

(%

)

Temperature (oC)

microwave conventional

Page 101: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

89 

 

 

 

caused by the prolonged exposure to higher temperature, as the extraction is held at that

particular temperature.

5.3.3 Comparisonofmicrowaveandconventionalalkalineextraction:Physico‐chemicalstructuralanalysisofwood

The results in the previous section on wood solubilization and the yield of xylan obtained have

shown the difference in the microwave and conventional extraction of xylan from wood. The

other goal of this study was to establish the mechanism involved in the microwave assisted

process. In the following section, the results of physico-chemical structural investigation of the

wood fibers before and after extraction will be discussed to distinguish the mechanism involved

in the microwave process compared to the conventional process of extraction of xylan. The

results discussed are based on the experiments listed in table 5.2, where the extractions are

performed under more or less similar conditions.

5.3.3.1 Chemicalcompositionofbirchwoodfibersbeforeandafterextraction 

The wood fibers used for this study contained about 42.3±1% glucan, 29.4±0.7% xylan, 1±0.3%

mannan, 21.3±0.4% Klason lignin, 1.4±0.1% acid soluble lignin, and 0.5±0.04% ash content.

The remaining portion (4.1 ± 1.3%) is accounted for acetyl and 4-O-methyl glucuronic acid

residues. Since the major chemical constituents of birch wood fibers were glucan, xylan, and

Klason lignin, the changes in these constituents were used in this study for the comparison of

microwave and conventional alkaline extraction.

The amount of glucan, xylan, and lignin components present in the wood fibers before and after

conventional and microwave assisted alkaline extraction is given in table 5.5. All the

compositions reported in table 5.5 are on the basis of the original 100 g of the oven-dry wood

Page 102: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

90 

 

 

 

fibers used for the alkaline extraction. As can be seen from the table, xylan content of the wood

fibers was decreased from 29.4 g xylan / 100 g wood fibers to 15.7 - 12. 1 g xylan/100 g wood

fibers after 5-30 minute duration of microwave assisted extraction. On the other hand, during the

same duration of 5-30 minutes of conventional extraction the amount of xylan decreased to 15.1-

13.5 g xylan/100 g wood fibers. An increase in the xylan removal from the wood fibers using

microwave assisted extraction was observed compared to the conventional extraction.

Table 5.5. Effect of alkaline extraction on the chemical composition of birch wood fibers No. Method of

extraction Temperature(oC)

Time (Minutes)

Glucan (%)

Xylan (%)

Lignin (%)

1 2 3 4 5 6 7 8 9

- I I I I II II II II

- 60 80 98 98 60 80 98 98

- 5 10 20 30 5 10 20 30

42.3±1.0 54.3±1.2 58.5±0.3 60.9±1.8 60.3±1.9 52.7±1.8 53.9±0.3 57.6±1.6 58.6±0.1

29.4±0.7 15.7±0.2 14.3±0.2 13.2±0.6 12.1±0.1 16.3±0.7 15.1±0.2 14.5±0.1 13.5±0.2

21.3±0.4 19.0±0.7 18.6±0.3 17.8±0.4 17.4±0.1 19.1±0.8 18.8±0.2 17.9±0.6 17.6±0.7

I – Microwave extraction, II- conventional extraction

The percentage removal of xylan on the basis of the amount of xylan present in the original

fibers are shown in the Figure 5.9.A, which also confirmed that the microwave heating resulted

in faster xylan removal compared to the conventional extraction. Similar results were reported

earlier in the pre-treatment of corn stover using water and acid treatment (Shi et al., 2011),

alkaline pre-treatment of switch grass (Hu and Wen, 2008), and FeCl3 treatment of rice straw (Lu

and Zhou, 2011).

Page 103: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

91 

 

 

 

Figure 5.9. Effect of microwave and conventional extraction of xylan on the chemical

composition of wood (A) xylan removal (B) lignin removal (c) Glucan left in the fiber

0

10

20

30

40

50

60

70

5 /60 10 /80 20 /98 30 /98

% r

em

ov

al o

f x

yla

n f

rom

wo

od

(b

as

ed

on

ori

gin

al

xy

lan

co

nte

nt)

Time (minutes)/ Temperature (oC)

microwave convetnional

0

5

10

15

20

25

5 /60 10 /80 20 /98 30 /98

% r

em

ov

al

of

lign

in w

oo

d, (

ba

se

d

on

ori

gia

nl

lign

in c

on

ten

t)

Time (minutes)/ Temperature (oC)

microwave convetnional

(A)

(B)

0

10

20

30

40

50

60

5 /60 10 /80 20 /98 30 /98

% in

cre

as

e o

f G

luc

an

in t

he

wo

od

Time (minutes)/ Temperature (oC)

microwave convetnional (C)

Page 104: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

92 

 

 

 

The lignin left in the alkaline extracted wood decreased from 21.3 g per 100 g of oven-dry wood

fiber to 19 - 17.4 g lignin per 100 g oven-dry wood fiber over the duration of 5-30 minutes of

microwave assisted extraction, and a similar decrease in lignin content was observed in the

conventional extraction for the same duration of extraction. The percentage removal of lignin

on the basis of the original lignin content in the wood fibers (Figure 5.9B) showed that about 11

to 18 % of lignin removed during the period of 5-30 minutes of microwave assisted extraction

and no significant difference was observed during the conventional extraction.

Unlike xylan and lignin reduction after alkaline extraction, glucan content increased as a result of

both microwave and conventional extraction. This is because of the removal of xylan and lignin

from the lignocellulose matrix which makes the fiber richer in cellulose. The glucan left in the

extracted fiber using microwave extraction increased from 42.3 g glucan per 100 g of oven-dry

wood fibers to 54.3-60 g of glucan per 100g of oven-dry wood fibers. The increase during

conventional extraction from the similar experimental conditions varied from 52.7-58.6 g. The

increase in the glucan content on basis of the original glucan present in wood fibers with the

time/temperature conditions are shown in Figure 5.9C, confirming that the increase in glucan in

the microwave extracted fibers are higher and is due to the higher removal of xylan and lignin

from the wood fibers during microwave assisted extraction. Comparatively higher amount of

xylan, and lignin removal (lower xylan and lignin, and higher glucan) is expected in the fibers

after conventional extraction, as the conventional performed isothermally at the final temperature

of the microwave assisted extraction. However, the lower removal of xylan and lignin from the

fibers after conventional process compared to microwave assisted extraction again indicates the

role of “microwave explosion effects" in enhancing the hydrolysis and dissolution of xylan.

Page 105: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

93 

 

 

 

 

5.3.3.2 FT‐IRspectraofwoodfibers 

Fourier transform infrared spectroscopy (FTIR) is a very powerful tool for obtaining rapid

information about the chemistry of wood, wood constituents and chemical changes taking place

during weathering, decay and chemical treatments, thermal treatments or natural aging (Faix et

al., 1993; Korner et al., 1992; Rodriguez et al., 1998; Popescue et al., 2007, 2009,2010). FTIR is

used in this study to see the chemical structural changes happened during the alkaline extraction

and to evaluate if there is any difference in the structure of wood after microwave extraction

compared to conventional extraction. To simplify the complexity of the spectra, they were

separated into two regions, namely the OH and CH stretching vibrations in the 3900 -

2700cm-1 and the finger print region in 1900-800 cm-1 (shown in Figure 5.10). The major peaks

assigned for wood taken from different references (Korner et al., 1992; Faix et al., 1993; 1998;

Pandey, 1999; Moore and Owen, 2001; Kubo and Kadla, 2005; Popescue et al., 2007, 2010) and

that obtained for wood after both extraction processes are given in the table 5.6. The broad band

at 3400 cm-1 represents the hydrogen bonded O-H stretching due to the presence of aliphatic

hydroxyl groups of cellulose, hemicelluloses and lignin. The peaks at 2936 cm-1, 2898 cm-1, and

2847 cm-1 represents C-H symmetric stretching in methyl and methylene groups, asymmetric

methoxyl stretching and symmetric CH2 stretching respectively. The enhancement in the

intensity of these peaks is more in the microwave assisted extracted fibers compared to the

conventionally extracted fibers. There are many well defined peaks in the finger print region of

1900-400 cm-1. The bands at 1743 cm-1 assigned to C=O stretching of acetyl and carbonyl

groups of hemicelluloses, and the disappearance of this peak suggests the susceptibility of the

removal of xylan during alkaline hydrolysis. The peak at 1645 cm-1 represents the absorbed OH

Page 106: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

94 

 

 

 

Table 5.6. Peak assignment of wood fibers before and after different extraction processes

Band assignment Wave number, cm-1

Wood before extraction

Wood after conventional extraction

Wood after microwave extraction

OH and CH stretching region Multiple formation of intra-molecular hydrogen bonds between phenolic groups and that of alcoholic groups

3421

3352

3354-3310

C-H stretching in methyl and methylene groups 2936

2936

2937

Asymmetric methoxyl C-H stretching 2898 2899 2899

Symmetric CH2 stretching 2847 2844 2844

Finger print region Unconjugated C=O in xylans

1737 (1743)

-

-

Absorbed O-H and conjugated C=O; Xyloglucan C=O vibration of carboxylic acids

1645 Decreased intensity; peak observed at 1658

Decreased intensity; peak observed at 1658

C=C of aromatic skeletal (lignin)

1596, 1506

1596, 1506

1595, 1506

C-H deformation in lignin and carbohydrates

1464,1426

1464,1426

1464,1426

C-H deformation in cellulose and hemicelluloses

1372

Decreased intensity

Decreased intensity

C-H vibration in cellulose and C1-O vibration in syringyl derivatives

1329

1329

1331

Guaiacyl ring breathing, C-O stretch in lignin and for C-O linkage in guaiacyl aromatic methoxyl groups

1269

Decreased intensity

Decreased intensity

Syringyl ring and C-O stretch in lignin and xylan 1242

1232 1235

C-O-C vibration in cellulose and hemicelluloses

1162

1162

1162

Aromatic skeletal and C-O stretch

1115

1121

1123

C-O stretch mainly from C3 and C5 in cellulose

1055

1055

1056

C-O and C-C stretching ring in celluloses and hemicelluloses

1034

1034

1035

Pyran ring stretching 897 (898) 898 897

Page 107: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

95 

 

 

 

2600280030003200340036003800

wood

M-5

M-10

M-20

M-30

(A1)

26002800300032003400360038004000

wood

C-5

C-10

C-30

C-20

Ab

sorb

an

ce U

nits

Wave number, cm-1

(B1)

Page 108: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

96 

 

 

 

Figure 5.10. FTIR spectra of wood fibers before and after microwave and conventional extraction (A1 and A2 : microwave assisted extraction and B1and B2 conventional extraction). M5-M30 and C5-C30 represents 5 - 30 minutes of extraction using microwave and conventional heating respectively.

80010001200140016001800

wood

M-5

M-10

M-20

M-30

Ab

sorb

ance

Un

its

Wave number, cm-1

(A2)

80010001200140016001800

wood C-5

C-10

C-30

C-20

Absorbance Units

Wave number, cm‐1

(B2)

Page 109: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

97 

 

 

 

and conjugated C= O (Ph-C=O) absorption and the intensity of this peak is decreased with the

time of extraction. After 5 minutes of microwave assisted extraction the intensity of this peak

decreased considerably, whereas in conventional extracted fibers the peak intensity decreased

only after 10 minutes of extraction, indicating faster removal of ester linkages between

hemicelluloses and lignin. The observed peak at 1658cm-1 in the fiber after 5 -10 minutes of

extraction is due to the absorbed OH peak in the fibers and may be due to the increase in the OH

concentration as a result of the hemicellulose/lignin hydrolysis. The peaks at 1596 and 1506 cm-1

represents the C=C stretching of lignin skeletal vibration, and C-H deformation in lignin

respectively. Intensity of the peak at 1329 cm-1 represents C-H vibrations in cellulose and C1-O

vibrations in syringyl derivatives of the lignin moieties, whereas the peak at 1269 cm-1 represents

guaiacyl ring breathing. The decrease in the intensity of the peak at 1329 cm-1 indicates the

susceptibility of syringyl groups for the alkaline hydrolysis. The intensity of the peak at 1260 cm-

1 (1269 - 1242 cm-1) decreased significantly in the extracted fibers indicating the ether bonds

(aromatic-C-O-C-aliphatic) also undergoes changes in the alkaline extraction. The intensity

decrease is higher in microwave assisted extraction compared to conventional extraction. This

could be due to the increased breakage of the ether linkages (a higher energy reaction compared

to ester bond hydrolysis) between lignin and hemicelluloses caused by the focused temperature

rise in the fiber matrix. The peaks at 1055 and 1034 cm-1 represents the C-O and C-C stretching

in cellulose and hemicelluloses respectively. The sharp band at 898 cm-1 is characteristic of β-

(1,4)-glycosidic linkages between the sugar units in cellulose and hemicelluloses (characteristic

peak of pyran ring stretching). However, it has to be noted that many of the peaks in this region

is the overlap of frequencies of lignin and carbohydrates.

Page 110: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

98 

 

 

 

5.3.3.3 Crystallinityofwoodfibers

Figure 5.11. XRD crystallographs of birch wood fibers before and after microwave and

conventional extraction

0

400

800

1200

0 10 20 30 40 50

Inte

nsi

ty, A

rbit

rary

un

its

Angle, 2, degree

Birch wood 

0

400

800

1200

0 10 20 30 40 50

Inte

ns

ity,

Arb

itra

ry u

nit

s

Angle, 2, degree

5 minutes 10 minutes 20 minutes 30 minutes

0

400

800

1200

0 10 20 30 40 50

Inte

ns

ity,

Arb

itra

ry u

nit

s

Angle, 2, degree

5 minutes 10 minutes 20 minutes 30 minutes

Conventional

Microwave

Page 111: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

99 

 

 

 

Crystallinity of wood fibers is important as it determines the downstream application of the

fibers after xylan extraction including enzymatic hydrolysis, pulping to produce cellulose fibers,

and for using reinforcements in bio-composites. The removal of lignocellulosic components such

as hemicellulose (xylan) and lignin during the extraction process can change the crystal structure

of cellulose by altering inter- and intra- hydrogen bonds present in the lignocellulosic matrix. X-

ray diffraction patterns of the wood fibers before and after microwave and conventional alkaline

extraction of wood fibers are shown in Figure 5.11 and the crystallinity index, which is a

measure of crystallinity, calculated using the equation 3.7 is given in table 5.7. The crystallinity

index of original wood fibers is 41.8% and this value increased after both extraction processes,

indicating the removal of amorphous portion of the lignocellulosic matrix. Crystallinity index

increased to about 61-62% after 10 minutes of extraction and thereafter decreased in both

extraction processes. The reduction in the crystallinity at high temperature and for longer

duration may be due to the changes in the crystal structure of the cellulose during extraction. The

Table 5.7. Crystallinity index of wood fibers before and after different extraction processes Sample ID Time

(Minutes) Intensity of I002 Intensity at

2=18o

Crystallinity index (%)

Birch wood fibers Wood fibers after microwave extraction Wood fibers after conventional extraction

- 5 10 20 30 5 10 20 30

870 1260 1136 1296 1246 1149 1024 1129 1183

506 576 433 586 586 475 396 454 502

41.8 54.3 61.9 54.8 54.7 58.7 61.3 59.8 57.6

Page 112: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

100 

 

 

 

reduction in crystallinity of the fibers after microwave assisted extraction is higher compared to

the conventional extraction, confirming the structural explosion that affects the crystal structure

of cellulose. The structural explosion was further confirmed by the SEM analysis of the fibers in

the following section. The high crystallinity of these extracted fibers makes them suitable for the

production of cellulose fibers/ cellulose nano fibers and also as the reinforcements for the

plastics, as the strength of the fibers increases with the increase in crystallinity.

5.3.3.4 Scanningelectronmicroscopy(SEM) 

SEM images of the wood fibers before and after microwave assisted and conventional alkaline

extraction are given in the Figure 5.12. The fiber morphology has changed after both extraction

processes. However the changes observed are different from each other. After conventional

extraction, for all the time and temperature combinations, the fiber structure becomes smooth

and intact and as the time of extraction increased the fiber become striated and became thinner as

a result of the amorphous hemicellulose and lignin removal from the matrix. Unlike conventional

extraction at all duration of extraction, the fiber structure is fibrillated and a more loosened or

porous structure was observed in fibers after microwave extraction, and as the time increased

more fibrillar structure can be seen with thinner fiber surfaces. This porous structure of the wood

after the microwave assisted extraction explained the higher dissolution of the wood components

and the decreased crystallinity for longer duration of extraction. Further, this confirms the

hypothesis that the explosion effect caused by the interaction of microwave and alkali in the

fibre increase the porosity and loosens the recalcitrant fibre structure thus enhancing the

hydrolysis and the mass transfer of the dissolved components to the solution.

Page 113: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

101 

 

 

 

Figure 5.12. SEM photomicrographs of birch wood fibers before and after microwave and

conventional extraction (A) wood fiber before extraction ; M -microwave extraction and C- conventional extraction

(A) 

Page 114: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

102 

 

 

 

5.3.4 XylanCharacterization

5.3.4.1 Chemical composition of xylan

Chemical composition of xylan isolated using microwave and conventional extraction for

different duration are given in table 5.8. Xylan contained about 78-93% xylose, and 0.3-1.1%

glucose, and 2-3.8% lignin, indicating the polymer extracted was a polymer with xylopyranose

backbone.

Table 5.8. Chemical composition of xylan isolated under different extraction conditions Sample ID Glucose (%) Xylose (%) Lignin (%) XM- 5minutes XM- 10 minutes XM- 20 minutes XM- 30 minutes XC- 5minutes XC- 10 minutes XC - 20 minutes XC- 30 minutes

0.3±0.04 0.5±0.07 1.1±0.2 1.0±0.1 0.5±0.07 0.7±0.0 0.9±0.1 1.1±0.1

78.9±2.7 81.2±1.9 92.1±3.4 93.4±1.2 74.5±1.4 75.5±1.5 87.6±3.7 87.2±1.7

2.2±0.3 2.1±0.1 2.7±0.5 3.6±0.3 2.1±0.6 2.7±0.8 3.8±0.5 3.7±0.3

XM- xylan obtained after microwave extraction; XC- xylan obtained after conventional extraction

5.3.4.2 FTIR Spectroscopy

The FTIR spectra of the isolated xylan (Figure 5.13) also show characteristic peaks of xylan, as

explained in Chapter 4. There is no significant difference was observed in the FTIR, suggesting

the isolation processes did not affect the structure of the polymer isolated.

Page 115: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

103 

 

 

 

Figure 5.13. FTIR spectra of xylan isolated using microwave and conventional extractions at

two different experimental conditions

5.3.4.3 Viscosity and degree of polymerization

Viscosity measurements of xylan dissolved in 0.5 M cupriethylenediamine solution were used to

find out the intrinsic viscosity of the solutions. Plot of viscosity data to determine the intrinsic

viscosity is shown in the Figure 5.14. The intrinsic viscosity, [η], has to reported to be related to

molecuar weight of a polymer by the Mark- Houwink equation [η] = K Mva, where Mv is the

viscosity average molecular weight and K and a are constants. The values of 'K' and 'a' depends

on the polymer-solution interaction. The values of 'K' and 'a' for xylan in CED are 8.5 x10-6 and

1.15 respectively (USDA higher education challenge program, course module 6). Number

500100015002000250030003500

Wavenumber cm-1

0.0

0.5

1.0

1.5

2.0

Abs

orba

nce

Uni

ts

500100015002000250030003500

Wavenumber cm-1

0.0

0.5

1.0

1.5

2.0

Abs

orba

nce

Uni

ts

XM‐5

XC‐5

XM‐30

XC‐30

Page 116: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

104 

 

 

 

average degree of polymerization is more informative rather than the viscosity average molecular

weight. Hence, conversion of intrinsic viscosity value into number average degree of

polymerization was carried out according to the folowing equation [η]CED = DPn x 4.7 x10-3

(Lebel et al., 1963), where DPn is the number average degree of polymerization. The intrinsic

viscosity, viscosity average molecular weight, and number average degree of polymerization of

xylan obatined at different experimental conditions are given in the table 5.9. It is clear from the

table that intrinsic viscosity decreased (22.5%) during 5 - 30 minuts of conventional

extraction,whereas viscosity values remain relatively constant during microwave assisted

extraction until about 20 minutes of extraction and then decreased rapidly (15.6 %). Further, in

both extractions, the decrease in the degree of polymerization is significant after 20 minutes of

extraction. This again confirms that higher temperture above 95oC enhanced the degrdation of

xylan in alkali. However, the percentage decrease in the intrinsic visocisty is lower (15.6%)

compared to the xylan obtained by convetnional method implying that molecular degradation is

higher during convetnional extraction. Calculated viscosity average molecular weight of the

xylan is in the range of 18000-19000 and the corresponding degree of polymerization is about

150. Such a degree of polymerization of birch xylan was reported earlier by Lebel et al. (1963).

Similar to viscosity, degree of polymerization is decreased (from 149-115) during convetnional

extraction, but remains almost constant during microwave extraction until 20 minutes of

extraction and then decreased to 126. The degradation of the xylan during convetnional

extraction may be due to the prolonged exposure of the dissolved xylan to high temperature.

These findings support the degradation of xylan and hence the lower yield of high molecular

weight xylan during convetnional extraction due to prolonged extraction duration.

Page 117: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

105 

 

 

 

Figure 5.14. ηsp/C vs. Concentration of xylan solutions in CED (a) Xylan obtained using microwave extraction (b) Xylan obtained using conventional extraction (C and M represents conventional and microwave extraction respectively; 5,10,20 and 30 represents the duration of time in each process)

0.4

0.5

0.6

0.7

0.8

0.9

1

1.1

1.2

0 0.2 0.4 0.6 0.8

ηs

p/C

(d

L/g

)

Concentration of xylan solution (g/dL)

M‐5 M‐10 M‐20 M‐30

0.4

0.5

0.6

0.7

0.8

0.9

1

0 0.2 0.4 0.6 0.8

(ηs

p/C

) (d

L/g

)

Concentration of xylan solution (g/dL)

C-5 C-10 C-20 C-30

Page 118: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

106 

 

 

 

Table 5.9 Intrinsic viscosity, viscosity average molecular weight and number average degree of polymerization of xylan obtained by microwave convetnional extraction Sample ID Temperature

(oC) Intrinsic viscosity (dL/g)

Viscosity average molecular weight

Number average degree of polymerization

XM- 5minutes XM- 10 minutes XM- 20 minutes XM- 30 minutes XC- 5minutes XC- 10 minutes XC - 20 minutes XC- 30 minutes

60 80 95 98 60 80 95 98

0.7019 0.7164 0.7148 0.5923 0.7036 0.6719 0.5858 0.5446

18859 19198 19160 16271 18899 18156 16116 15125

149 152 152 126 149 142 124 115

XM- xylan obtained after microwave extraction; XC- xylan obtained after conventional extraction

5.4 Conclusions

Alkaline extraction of xylan from birch wood fibers using microwave and conventional heating

were studied with two different experimental set-ups to demonstrate the efficiency of the

microwave process and the mechanism involved in the process. The rate of wood dissolution was

found to be higher compared to the conventional extraction. Comparatively higher wood

solubilization, and xylan removal during microwave assisted extraction for similar duration of

extraction indicated that the mechanism involved is different from that of conventional

extraction. Comparison of the physico-chemical structural changes observed for the fibers

subjected to microwave and conventional extraction demonstrated that temperature dependence

on the rate of hydrolysis is not the only factor responsible for the faster extraction using

microwave extraction but the structural effect, due to the interaction effect of microwave

radiation and the alkali within the fibers, also contributes to the hydrolysis and dissolution.

Chemical structure of the xylan did not exhibit dependence on the extraction process. Intrinsic

Page 119: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

107 

 

 

 

viscosity analysis showed that microwave assisted xylan has higher intrinsic viscosity and the

calculated degree of polymerization was about 150. Disintegration of xylan at high temperature

was observed from the molecular weight and degree of polymerization of the isolated xylan and

it is believed that this is one of the factors that leads to the lower yield of xylan obtained during

conventional extraction.

 

 

 

 

 

 

Page 120: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

 

 

108 

 

Chapter6 InvestigationofstructuralchangesofalkalineextractedwoodusingX‐raymicrotomograhy:Acomparisonofmicrowaveversusconventionalmethodofextraction

6.1 Introduction

X-ray computed microtomography (micro CT) has been used for qualitative as well as

quantitative analysis in a number of research applications including soil science, geology,

hydrology, and material science (Thovert et al., 1993; Perret eta l., 1999; Wildenschild et al.,

2002; Muler et al., 2002; Salvo et al., 2003). In this technique, a series of radiographic

projections at different viewing angles of the samples are reconstructed mathematically to

visualize and analyse the architecture of the materials with a micrometer resolution. The

visualization of the internal structure of the materials depends on the travel of the X-rays through

the structure which in turn depends on the mass density and absorption coefficient of the

materials of interest. Literatures are available on the use of X-ray CT for the analysis of the pore

structures of fruits and vegetables such as internal quality changes of peaches and cucumber

during storage, microstructure of cereal products, and pore space analysis of apples (Harker and

Hallet, 1992; Barcelon et al., 1999; Mendoza et al., 2007). The advantage of this process over the

conventional microtomy is that the analysis can be done without any special sample preparation

or chemical fixation which usually a time and effort intensive process in the conventional

microtomy analysis. Moreover, X-ray micro CT enables the study of 3- dimensional internal

Results of this chapter have been published in Journal of Wood Chemistry and Technology.

Volume 33, Issue 2, June 2013, pages 92-102

Page 121: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

109 

 

 

 

microstructure of the materials of interest. The method has been recently used for the

quantification of the plant microstructure (Fromm et al., 2001; Stuppy et al., 2003; Steppe et al.,

2004; Mayo et al., 2009). It has been reported that the micro CT method can be used effectively

to quantify the microstructure of the wood; particularly the xylem anatomy and porosity

provided a proper image analysis method is used for the processing of the images (Steppe et al.,

2004). In this study, we tried to make an attempt to quantify the structural changes of the wood

during conventional and microwave assisted alkaline extraction using X-ray computed

microtomography to understand the difference in the mechanism of microwave and conventional

method of extraction. Further, we used this technique as an indirect means for measuring the

temperature generated inside the wood fibers during microwave extraction.

6.2 Material

The birch wood chips used in this study were obtained from St. Mary's Paper, Ontario, Canada.

The chips were oven dried at 40oC for 72 hours. Rectangular samples with approximate

thickness of 1-2 mm and 3-5 mm were cut from the dried wood chips. The samples were soaked

in a 4 wt% solution of sodium hydroxide (with a sample to liquid ratio of 1:10) for 2 hours to

completely wet the chips. The samples with the solution were then subjected to microwave

irradiation at a power level of 110 W for different times.

Another set of samples were extracted using the conventional method of heating using an oil

bath for different duration at different temperatures. For temperatures higher than 100oC,

extractions were performed using small laboratory digestion cylinders. Extraction time was

determined after the temperature of the solution reaches the set temperature. Extraction

conditions used in this study are given in the table 6.1. After extraction, the samples were

Page 122: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

110 

 

 

 

washed thoroughly and air dried overnight. Three different samples from each set were used for

the X-ray CT imaging to accommodate the variability of the wood samples.

Table 6.1. Experimental conditions used for extraction

Extraction conditions

Conventional Microwave assisted

Temperature Time (minutes)

Power output Time (minutes)

70oC 90oC 100oC 120oC 140oC

30 60 120 120 10 10 10

110W

1 6 10

6.3 X‐raycomputedmicrotomographyimageprocessing

The procedure used for the 3D construction and quantitative analysis of porosity is demonstrated

in Figure 6.1. The scanned images were reconstructed to grey scale image using the NRecon

software from Skyscan. The grey scale images were converted to black and white binary images

to specifically select the voids from the matrix, and were done by the segmentation technique.

The simplest method for segmentation is thresholding. In the figure (Figure 6.1) the bright areas

represent the xylem vessels whereas the dark areas represent the remaining wood matrix.

The threshold value should define the boundary between the objects in the area of interest and

should be selected in between the intensity of both the void and the solid phases, as the selection

of threshold highly influences the analysis. A similar procedure used by Steppe et al. (2004) was

Page 123: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

111 

 

 

 

Figure 6.1. Binary images of birch wood and the 3D image of the vessels extracted from the binary images (the 3D structure consists of a finite number of voxels with a length of voxel side of 7.4 um)

used to reduce artifacts from the image. A global thresholding followed by a morpho-

mathematical filter with a rounding structural element (available in Image Pro software) was

used to extract the vessel elements. The 2D images were then converted to 3D images. After

processing, the images were labeled and performed a quantitative measurement of each

individual in the volume of interest (VOI) to get the volume of the voids. Porosity of the wood

samples was calculated using the equation 6.1.

Volume of vessel elements

Porosity (%) = -------------------------------- x100 (6.1) Volume of interest selected 6.4 ResultsandDiscussion

The interaction of microwaves with any material depends on its dielectric constant and dielectric

loss factor. The overall efficiency of a material’s ability to utilize or absorb microwave energy

Page 124: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

112 

 

 

 

from the radiation is obtained from the loss tangent parameter, which is a ratio of dielectric loss

factor to the dielectric constant (Nelson and Datta, 2001). Theoretically, the rate at which

electromagnetic energy is converted to thermal energy in a load which is being subjected to

microwave frequency "f" is given by the equation 6.2.

dT/dt = k’’fE2

rms/Cp (6.2) where k is a constant, Erms is the electric field intensity and and Cp are the density and specific

heat capacity of the material being irradiated (Metaxas and Meridith, 1983). The electric field

intensity depends on ''. The volumetric heating produced by microwave energy leads to an in-

situ increase in temperature of the load within short duration compared to conventional heating

method. The recorded temperature of the solution after 1, 6, and 10 minutes of irradiation was

52.3oC, 82.3oC, and 96.7oC respectively. For lignocellulosic biomass the values of loss tangent

is low compared to other solvents such as water, and ethanol, and the value is highly dependent

on the moisture and temperature. The loss tangent values of birch wood was reported to be 0.15-

0.19 at different values of moisture content (Koubaa et al., 2008) and that of sodium hydroxide

solution is around ~1-5 from ambient temperature to 100oC (Keshwani, 2010). Since the tan

value is higher for NaOH solution, it will be the most lossy of the components present in the

system and hence is expected that sodium hydroxide solution absorbs most of the energy in the

extraction mixture of wood fibers and sodium hydroxide solution. The energy absorbed by the

sodium hydroxide solution inside the fiber produces instantaneous heat and pressure and this

may lead to the "explosion" of the wood structure and increase in the void volume of the fiber

structure. The porous fiber structure after microwave assisted extraction was shown in the SEM

photomicrographs of the fibers after extraction (Figures 5.3 c and 5.12), demonstrating the

Page 125: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

113 

 

 

 

difference in the micro structure of fibers before and after extraction. In this study, an attempt is

given to quantify the void fraction or porosity of the wood fiber structure using computerized X-

ray microtomography to understand the physical mechanism involved in the process.

6.4.1 Imageresolutionanalysis

Accuracy of the extracted parameters from the processed image highly depends on the resolution

of the images. Low resolution images can be improved by many methods including image

enhancing filters, sharpening filters and deconvolution techniques etc. However, this affects the

accuracy of the results extracted and hence a less processed image with an appropriate resolution

should always preferred to get the finest details of the structure. In order to determine the

optimum resolution required for the samples using the shortest scanning time, the samples were

scanned at different image resolution (3.4, 7.4, 11.4, 15.4, and 17.4 m/pixel) of the wood

samples and the quantitative measurements were performed as described previously. Image

resolution depends on the contrast features of the image and to get a high contrast image, a good

spatial resolution is needed. Figure 6.2 shows the representative cross-sectional

microtomography images of birch wood samples at different resolutions with a length side of 1

mm and the average porosity (calculated as the ratio of the volume of voids to that of the volume

of interest) with their standard deviation computed from three samples are shown in the Figure

6.3. It is clear from the figure that low resolution results into the scattering of images and hence

the visualization and analysis of the microstructure becomes more difficult. Similar results were

reported earlier in a study where the authors used X-ray microtomography to study the three-

dimensional microstructure of apple tissues (Mendoza et al., 2007). The quality of the images is

significantly affected by the resolution when magnifications are lower than 11.4 m/pixel was

Page 126: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

114 

 

 

 

used for scanning of the samples. The image at the lowest resolution is blurred and contrast is

very poor.

Figure 6.2. Typical X-raymicrotomographic binary images of birch wood obtained at

different resolution

The dependence of sensitivity to the resolution changes is evident from the calculated porosity.

The porosity was decreased with the decrease in the resolution. Statistical analysis of results

shows the resolution has a significant effect on the porosity measurement (ANOVA, P-value

<0.005). When comparing the porosity values of each resolution with that of the highest

resolution (3.4 µm/pixel), a significant statistical difference was observed from 11.4 µm/ pixels.

Based on these results the resolution for better contrast image was selected as 7.4 µm/pixel and

was used for further analysis of the wood chips after extraction.

7.4 µm 3.4 µm  11.4 µm 

15.4 µm  17.4 µm 

Page 127: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

115 

 

 

 

Figure 6.3. Average porosity of wood samples using different resolutions

6.4.2 Representativevolumeofinterest

Since the complexity of the 3D image analysis and computation depends on the size of the

analyzed images, it is desirable to have a representative sub volume, (volume of interest, RVI)

that can be treated as a mathematical point of the continuum scale or representing the

macroscopic properties of the porous medium. The RVI is defined as the range of volumes over

which a statistical average can be performed. To find out the representative volume of interest,

the effect on the porosity of six volume sizes extracted from the same stacks of images were

analyzed in all the three samples of wood. Then the average porosity and standard deviations

were calculated for each sub volume. A minimum of 9 sub volumes (3 from each sample) were

used for the averaging. Figure 6.4 shows the average porosity with standard deviations calculated

from different volume size of the images scanned. Statistical analysis shows that there is

significant difference in the RVI's between the smallest ones and the largest volume and the RVI

Page 128: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

116 

 

 

 

difference between 0.41 to 3.24 mm3 are not statistically significant (P>0.05). To estimate the

variability in the measurements, standard error for each sub-volume was calculated using the

relation SE= STD/n1/2, where SE is the estimated standard error, STD is the standard deviation

and n the respective number of sub-volumes in each sample set. The values of standard error

calculated are: 0.66%, 0.48%, 0.47% and 0.52% respectively for 0.41 mm3, 0.79 mm3, 1.67

mm3, and 3.24 mm3. Since the standard error calculated was very close for 0.79 and 1.67 mm3,

we have selected 0.79 mm3 (125 pixels/side) as the volume of interest for comparing the porosity

of the wood samples.

Figure 6.4. Average porosity of wood samples using different volume sizes

6.4.3 Comparisonofconventionalextractedandmicrowaveextractedwood anatomy

Typical 3D images used for quantitative determination of porosity of different samples are

shown in the Figure 6.5. The resolution of the images and the volume of interest used for

porosity determination are 7.1µm/pixel and 0.79 mm3 (125 pixels/side) respectively. Average

Page 129: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

117 

 

 

 

porosity calculated from different samples having a thickness greater than 2 mm are given in the

table 6.2. Porosity of wood increased after both methods of extraction and is due to the

dissolution of the wood components during alkaline extraction. Porosity of wood increased

about 10% after 2 hours of extraction at 70oC and remains more or less constant with the

duration and temperature of extraction. This indicates that during conventional method of

alkaline extraction, time and temperature did not have a significant effect on the microstructure

of the vessel elements.

Figure 6.5. Typical 3D images of wood chips after different extraction conditions

(sample size >2 mm)

Contrary to this, porosity of the wood after microwave extraction increased 15% after 10 minutes

of microwave irradiation, and also increased with the time of irradiation. This suggests that more

rupture of the fiber structure is occurred in the fiber during microwave irradiation that affects the

vessel elements. It is noted from the table 6.2 that standard deviation of the porosity values of

the wood samples after microwave irradiation is higher compared to the conventionally extracted

samples. In addition, statistical analysis of the porosities calculated for the wood samples after

both type of extraction processes, showed that the differences between average porosities are not

70oC, 2 h 90oC, 2 h 110W, 10 min

Page 130: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

118 

 

 

 

significant (ANOVA P value > 0.05). This may be due to the variations in the pore structure

occurred as a result of the non-uniform structural opening of the fiber structure during

microwave irradiation which in turn due to the difference in the heating occurred inside the

wood. However, SEM analysis of the wood fibers of size 2 mm has exhibited a considerable

difference in the pore structure of the fibers after the two methods of extraction (Figure 5.3 b and

c and 5.12). Hence the effect of sample thickness on the microwave extraction was studied

further.

Table 6.2. Porosity of wood samples after different extraction methods (sample size > 2 mm)

  Porosity (%)

Conventional Microwave assisted

Porosity of wood (%): 18.53 ± 1.52

70oC 30 minutes 60 minutes 120 minutes 90oC 120 minutes

20.04 ± 0.86 20.04 ± 1.11 20.21 ± 1.31 20.39 ± 1.57

1 minute 6 minutes 10 minutes

20.07 ± 2.34 21.39 ± 2.10 22.31 ± 2.99

To determine the correlation between the thickness of the samples and the porosity of the fiber

structure during alkaline extraction, samples of thickness less than 2 mm were used for the

extraction and computation of porosity. Typical 3-dimensional tomographic images are shown in

the Figure 6.6 and the calculated porosities are given in the table 6.3. It is clear from the figure

Page 131: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

119 

 

 

 

Figure 6.6. Typical 3D tomographic images of wood samples, samples after conventional and

microwave assisted extraction process (sample size < 2 mm) (a) wood (b) wood after 2 hours of extraction @90oC (c) wood after 10 minutes of microwave assisted extraction

and table that porosities are different for samples after conventional and microwave extraction.

Porosity is increased with temperature in the case of conventional extraction; however, there is

no significant difference observed between thicker and thinner samples (table 6.2 & 6.3). On the

other hand, after microwave extraction thinner samples exhibit a higher porosity. As the sample

(a) 

(b)  (c)

Page 132: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

120 

 

 

 

Table 6.3. Calculated porosities of wood samples after sodium hydroxide extraction using conventional heating and microwave irradiation process (sample size < 2 mm)

Sample Porosity (%) Wood Wood after 2 hours of extraction @70oC Wood after 2 hours of extraction @90oC Wood after 10 minutes of microwave assisted extraction

18.2 ± 0.4 20.1 ± 0.9 22.9 ± 1.2 25.9 ± 0.8

Table 6.4. Comparison of the calculated porosities of wood samples after 10 minutes of sodium hydroxide extraction using conventional heating and microwave irradiation process ( sample size < 2 mm)

Sample Porosity (%)

Wood Wood after extraction @100oC Wood after extraction @120oC Wood after extraction @140oC After Microwave extraction (10minutes)

18.2 ± 0.4 18.6 ± 1.1 24.1 ± 3.1 24.8 ± 3.8 25.9 ± 0.8

thickness decreases, the amount of energy per unit mass increases and may leads to more rupture

of the fiber structure resulting an increased porosity. Further, smaller variation in the porosity

(standard deviation) of the thinner samples indicates the irradiation produce a uniform changes

in the porous structure. This may be because of the uniformity in the interaction of microwave

radiation with the solvent present in the thinner samples. Statistical analysis of the average

porosities computed showed that the difference in porosity is significant (P value << 0.05)

compared to the conventionally extracted samples. The results indicate that focused heating

during microwave extraction and/or direct interaction of the microwave energy with the solution

Page 133: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

121 

 

 

 

inside the wood fiber leads to a more porous structure compared to the conventional heating

methods where the heating occurs through a temperature gradient.

Figure 6.7. Typical 3D images of wood chips after 10 minutes of extraction at 100oC, 120oC, and 140oC

In conventional extraction process, energy is transferred due to thermal gradients, but during

microwave extraction, heating is occurred by the direct interaction of electromagnetic energy

with the material. The temperature generated inside the fibers during the extraction may not be

the same as that of the reported temperature of the solution after the extraction process. In order

Page 134: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

122 

 

 

 

to determine the temperature generated inside the fibers during microwave extraction, anatomy

of wood chips were studied after performing conventional extraction at three different

temperatures; 100oC, 120oC and 140oC for 10 minutes and compared that with the samples

extracted for 10 minutes of microwave extraction. The typical tomographic pictures are shown in

the Figure 6.7 and the respective average porosities calculated are given in the table 6.4. Porosity

of the samples did not changed after 10 minutes of extraction at 100oC and was increased

thereafter for 120oC and 140oC respectively. Comparison of the structural anatomy demonstrated

that the average porosity of the samples extracted for 10 minutes of irradiation is similar to that

of the ones extracted at 140oC. However, statistical analysis of the porosities of the samples

extracted at 120oC and 140oC were not significantly different. This led to the conclusion that the

structural changes produced from 10 minutes of extraction is similar to that of conventional

extraction at 120oC indicating the temperature generated inside the fibers during 10 minutes of

microwave extraction was about 120oC. Such an increase in temperature within a short duration

might leads to the rupture of the fiber structure confirming the enhanced dissolution kinetics and

mass transfer.

 

6.5 Conclusions

X-ray microtomography was used to study the structural changes of wood during conventional

and microwave alkaline extraction of wood. The Following conclusions were drawn from this

study:

1. X-ray microtomography was found to be an effective method for 3-dimensional

characteristics of wood samples.

Page 135: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

123 

 

 

 

2. Porosity of wood after conventional extraction was not changed significantly with the

duration, and temperature of extraction; however, in thinner samples porosity increased at

higher temperature.

3. Porosity of wood after microwave extraction was found to be varied with the duration of

irradiation and thickness of the samples and was higher than the conventional extraction.

Increased porosity revealed the microstructure rupture as a result of the selective heating

effect due to the direct interaction of microwave energy with the solution inside the wood

fiber structure.

4. Porosity of wood samples extracted at different temperatures for the same duration (10

minutes) demonstrated that the structural changes associated with microwave extraction is

more or less similar to the conventionally extracted samples at 120oC.

Page 136: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

 

 

124 

 

Chapter7 Statisticaloptimizationofmicrowaveassistedalkalineextractionofxylanfrombirchwoodusingresponsesurfacemethodology

 

7.1 Introduction

The results obtained so far indicated that microwave assisted extraction can be used as an

efficient alternative to the conventional extraction. However, the process has to be optimized

further with respect to sample size, alkali loading, time of irradiation, and concentration of alkali,

as these are the parameters affecting the efficiency of the microwave extraction. In this study,

our objective was the optimization of the microwave assisted alkaline extraction using statistical

design of experiments.

Studies on the effect of one-factor at a time on the microwave assisted extraction can hardly

provide the relationship between all the experimental input parameters (factors) and the output

responses, as the results are valid only for the fixed experimental conditions and the prediction of

other conditions are vague. Design of experiments (DOE) using response surface methodology

(RSM), where several factors can be varied simultaneously, and each factor may be evaluated

independently, can be a better alternative to study and establish the relationship between all the

factors studied on the output responses (Montgomery, 2001). Advantage of this method is the

minimum number of experiments required for predicting the relationship between variables and

Results of this chapter have been published in the Journal of Material Science and Chemical

engineering. Vol.1 No.6, November 2013, 38-50.

Page 137: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

125 

 

 

 

responses, and hence the optimization process, rather than studying all possible combinations of

the experiment. Basically, RSM is a compilation of statistical techniques for designing

experiments, building models, and estimating the effect of factors on the responses and searching

for the optimum conditions of factors for a particular experiment (User guide, Design expert

software). Further, this technique can quantify the relationships among one or more measured

responses and the critical input factors. DOE has been widely used in the optimization

applications in the bio-refinery industry (Yoshida et al., 2010; Lu and Zhou, 2011; Gosh et al.,

2011; Ender and Perendeci, 2012; Brudecki et al., 2013). In this study, we used DOE to

investigate the effect of four variables such as time of extraction, sample size, solid to liquid

ratio, and concentration of alkali on the microwave-assisted extraction of xylan from birch wood.

The responses studied were solubilization of wood (wood dissolution), temperature, and yield of

xylan. Finally, the extraction process was optimized to maximize the yield of xylan.

7.2 Materials

The birch wood fibers used in this study were prepared from birch wood logs obtained from

premises near to University of Toronto. Preparation of extractive free wood was explained under

section 3.1.1. The dried extractive-free wood consists of 42.95±1.05% glucan, 29.35±0.69%

xylan, 0.96±0.27% mannan, 21.26±0.41% acid insoluble lignin, 1.43±0.13% soluble lignin,

and 0.52±0.04% ash content. Sodium hydroxide solutions of different concentrations used for the

extraction were of reagent grade.

7.3 Microwaveassistedextraction

Microwave extraction procedure used was explained in Chapter 3. The wood fiber slurry with

different combinations of alkali was subjected to microwave extraction using a power input of

Page 138: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

126 

 

 

 

110 W, as it was found to the best power input for getting the xylan without much degradation.

Immediately after extraction, temperature of the reaction media was noted. The wood residue

was separated from the liquid phase by filtration and the residue was washed to remove the

alkali. Xylan was precipitated from the liquid phase by neutralizing the solution to a pH of 4.6

and allowed to settle overnight. The precipitate was separated by centrifugation, washed with

95% ethanol, and then freeze-dried. Wood dissolution and yield of xylan were expressed as per

the equations 7.1 and 7.2.

%

. .

. 100

7.1

% .

. 100 7.2

 

 

7.4 ExperimentaldesigningusingCCD

CCD is the standard RSM to optimize the response, estimate the second order polynomial

relationship between independent variables and the dependent variables, and the interaction

between the independent variables with the dependent variables (Montegomery, 2001). In order

to optimize the microwave assisted extraction, a four-factor, five-level CCD with replicates at

the center point was used. The variables studied were extraction time (A), NaOH concentration

(B), solid to liquid ratio (C), and sample size (D). The variables ranges were selected based on

the preliminary experiments. The number of experiments required for 4 variables CCD was

calculated using the equation 7.3.

2 2 (7.3)

Page 139: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

127 

 

 

 

where n is the number of independent variable, first term (2n) represents the number of factorial

points, second term (2n) represents the number of axial points (star points), and the third term

(nc) represents the number of replicates at the centre point. The whole CCD design of this study

consists of a total of 30 experiments including 16 factorial points, 8 axial points (=2) with 6

replicates at the centre point. The replicates of the centre point were used to determine the

reproducibility and reliability of the experiments and the error occurred during the experiment.

The variables were coded for statistical estimation as per equation 7.4.

∆ (7.4)

where, xi is the independent coded variable, Xi is the real value of the independent variable, Xo

is the real value of the independent variable at the centre point and Xi is the step value change.

Table 7.1. Independent variables studied in the CCD with their coded and uncoded levels

Coded variable level

Variable 1 (A) Extraction time

(min)

Variable 2 (B) NaOH

concentration (%)

Variable 3 (C) Solid to liquid ratio (g:mL)

Variable 4 (D) Sample size (g)

- (-2) -1 0

+1

+ (+2)

2.5 10

17.5 25

32.5

2 4 6 8 10

2 8 14 20 26

2.5 5

7.5 10

12.5

The range of variables and their levels studied are given in the table 7.1. The variables coded as

-1 and +1 represents the low and high levels of the variables studied, zero represents the centre

point of the design and -, and + represents the axial or star points of the design. In the study

Page 140: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

128 

 

 

 

the value of was fixed at 2 to make the design a rotatable one. The complete CCD design was

generated using the Design Expert 8.0.7.1(trial version, Stat-Ease Inc., Minneapolis, MN, USA)

software. The complete design matrix obtained with the coded and uncoded independent

variables and the corresponding responses obtained are given in the table 7.2.

As in any statistical designing of experiment, the experiments performed were randomized to

minimize the unpredictable variations in the observed responses due to uncontrolled extraneous

factors. The responses studied include dissolution of wood, yield of xylan, and temperature of

the slurry. A second degree polynomial quadratic equation was used to develop the empirical

model to establish the correlation between the variables and each response Y, as shown in the

following equation 7.5.

,

(7.5)

where , , , and are the constant tem, regression coefficients of the individual linear

effects, quadratic effects, and interaction effects between the variables respectively. These

polynomial equations were used to create the surface or contour plots to visualize the

relationship between the process variables and the responses studied.

7.4.1 Statisticalanalysisandthemodelevaluation

Statistical analysis of the experimental design was carried out using the Design Expert software.

Multiple linear regression analysis of the experimental data was used to evaluate the statistical

Page 141: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

129 

 

 

 

Table 7.2. DOE design matrix and the results

Run Independent variables studied Responses A

Time (min) B NaOH concentration (%)

C Solid to liquid ratio (g/mL)

D Sample size (g)

Wood dissolution (%)

Yield of xylan (%)

Temperature (oC)

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

10.0 (-1) 17.5 (0) 10.0 (-1) 17.5 (0) 25.0 (+1) 25.0 (+1) 17.5 (0) 10.0 (+1) 17.5 (0) 17.5 (0) 10.0 (-1) 25.0 (+1) 2.5 (2) 17.5(0) 25.0 (+1) 10.0 (-1) 17.5 (0) 32.5 (+2) 25.0 (+1) 17.5 (0) 10.0 (-1) 25.0 (+1) 17.5 (0) 25.0 (+1) 25.0 (+1) 10.0 (-1) 17.5 (0) 17.5 (0) 10.0 (-1) 17.5 (0)

8.0 (+1) 6.0 (0) 4.0 (-1) 6.0 (0) 4.0 (-1) 4.0 (-1) 6.0 (0) 4.0 (-1) 6.0 (0) 6.0 (0) 4.0 (-1) 4.0 (-1) 6.0 (0) 2.0 (-2) 8.0 (+1) 8.0 (+1) 6.0 (0) 6.0 (0) 8.0 (+1) 6.0 (0) 8.0 (+1) 4.0 (-1) 6.0 (0) 8.0 (+1) 8.0 (+1) 4.0 (-1) 6.0 (0) 10.0(+2) 8.0 (+1) 6.0 (0)

20.0 (+1) 14.0 (0) 8.0 (-1) 14.0 (0) 20.0 (+1) 8.0 (-1) 14.0 (0) 20.0 (+1) 14.0 (0) 14.0 (0) 20.0 (+1) 20.0 (+1) 14.0 (0) 14.0 (0) 20.0 (+1) 8.0 (-1) 26.0 (+2) 14.0 (0) 8.0 (-1) 14.0 (0) 20.0 (+1) 8.0 (-1) 14.0 (0) 20.0 (+1) 8.0 (-1) 8.0 (-1) 14.0 (0) 14.0 (0) 8.0 (-1) 2.0 (-2)

10.0 (+1) 7.5 (0) 10.0 (+1) 7.5 (0) 5.0 (-1) 5.0 (-1) 7.5 (0) 5.0 (-1) 2.5 (-2) 7.5 (0) 10.0 (+1) 10.0 (+1) 7.5 (0) 7.5 (0) 10.0 (+1) 5.0 (-1) 7.5 (0) 7.5 (0) 5.0 (-1) 7.5 (0) 5.0 (-1) 10.0 (+1) 12.5 (+2) 5.0 (-1) 10.0 (+1) 5.0 (-1) 7.5 (0) 7.5 (0) 10.0 (+1) 7.5 (0)

23.73 26.77 19.83 26.99 26.59 26.91 26.85 21.52 30.24 26.81 20.15 24.39 21.96 16.87 29.60 27.04 25.03 33.69 34.62 27.38 26.15 26.40 26.64 31.92 33.55 20.44 26.18 26.81 29.88 26.64

15.18 18.88 12.95 17.96 13.16 9.81 17.97 10.41 13.10 17.96 8.40 13.38 9.58 4.33 22.49 16.11 14.82 16.79 14.95 20.56 16.62 15.13 19.93 20.03 21.88 7.43 18.58 19.82 21.82 14.88

50.0 87.5 78.0 88.0 96.0 98.0 88.0 72.5 96.5 87.5 56.0 87.0 51.0 93.5 83.5 85.0 70.5 98.5 98.5 88.0 70.5 99.5 74.0 96.5 97.0 87.0 86.5 87.0 74.0 95.0

significance of the model developed. Analysis of variance (ANOVA) of the model for each

response was also performed to study the significance of the regression coefficient of

determinations for each effect (linear, quadratic, and interaction) in the model, and to assess

response model fit to the data (the lack of fit parameters), and hence to evaluate the significance

of the models for further rationalization. The competence of the model was also evaluated using

adjusted R2 and predicted R2, rather than R2 values, as the value of R2 increase with the addition

of variables despite the significance of the added variables.

Page 142: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

130 

 

 

 

7.4.2 Optimizationoftheprocessingvariables

The microwave extraction process was optimized using the numerical optimization approach

available in the software with the objective of maximizing the yield of xylan. Four additional

experiments were performed using the optimal conditions selected and the results (responses)

were compared with the predicted responses to verify the validity of the surface response model

developed.

7.5 ResultsandDiscussion

The results of the optimization experimentation of the four variables (extraction time,

concentration of alkali, solid to liquid ratio, and sample size) on the responses (wood dissolution,

yield of xylan, and temperature generated) are given in the table 7.2. The surface response

Table 7.3. Polynomial equations for the quadratic model and the regression coefficients Quadratic model equations (A= extraction time, B= NaOH concentration, C= solid to liquid ratio, and D= sample size)

R2 Adjusted R2

Predicted R2

Standard deviation

Adequate precision

87.58 11.58 1.33 6.42 5.00 0.56 2.81 2.19 0.19 0.81 2.44

3.2 0.68 1.20 0.82

% 26.94 3.02 3.02 0.59 0.77 0.27 0.44 0.052 0.57 0.12 0.33

0.19 0.92 0.30 0.35

% 18.65 1.51 3.72 0.22 1.51 0.17 0.94 0.45 0.030 0.29 1.52

1.23 1.51 0.81 0.40

0.9987 0.9962 0.9715

0.9974 0.9927 0.9450

0.9937 0.9804 0.8722

0.70 0.36 1.08

100.489 69.354 22.918

quadratic models used to study the relationship between the independent variables and to that of

the responses are shown in the table 7.3. Model regression coefficients were also reported in the

table for checking the adequacy of the model. The positive values in the model show the

Page 143: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

131 

 

 

 

synergistic effect of the variables on the response, whereas the negative values show the

antagonistic effect of the respective effect of the variables. The correlation coefficient values,

standard deviation, and adequate precision (a term that measures signal to noise ratio) of the

model indicate the quality of the model.

7.5.1 Effectofextractionvariablesontemperatureofthewoodslurry

During microwave assisted extraction, the temperature of the wood slurry increases as a result of

direct interaction with the electromagnetic radiation and the reaction media. In this case most of

the radiation absorption is caused by sodium hydroxide solution, as the microwave radiation

couple with the components of high loss value in a system where the components have different

dielectric properties. Sodium hydroxide solution has a high loss factor (~1-5) (Keshwani, 2009)

compared to wood (0.15-0.19) (Koubaa, 2008), and hence heating of the wood slurry occurs

mainly through the interaction of microwave and the solution. The effect of independent

variables (time of irradiation, concentration of sodium hydroxide, solid to liquid ratio and sample

size) on the temperature of the reaction media was studied using the experimental design, and the

results are given in the table 7.2. The temperature reported here is the temperature of the slurry

after particular time of irradiation. The lowest temperature reported was for the run #1 (50oC),

whereas the highest temperature was for the run #22 (99.5oC). The lowest temperature obtained

when all the variables studied except time were on the higher levels. On the other hand, the

highest temperature was obtained when the variables time and sodium hydroxide concentration

were on the higher levels indicating time of irradiation increases temperature of the slurry.

The surface response quadratic model using the CCD to establish the relationship between the

variables and temperature of the system is given in the table 7.3. The regression coefficients, R2,

Page 144: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

132 

 

 

 

adjusted R2 and predicted R2 for the model are 0.9987, 0.9974, and 0.9937 respectively. The high

values of R2, close to unity, indicate that the final temperature of the system can be predicted

with the model suggested. Further, close agreement within these values also indicates the

accuracy of the predictive model.

Figure 7.1. Actual temperature vs. predicted temperature

Figure 7.1 shows the predicted temperature versus the actual temperature, and the strong

correlation between the predicted and actual temperature indicates the significance of the model

within the experimental window. Standard deviation of the model was 0.70 and the small

standard deviation indicates the reproducibility of the model. Further, the adequate precision of

the model reported was very high (100.489); better model will have a value greater than 4, which

again indicates the suitability of the model for navigating the design space.

32

40.00

50.00

60.00

70.00

80.00

90.00

100.00

50.00 60.00 70.00 80.00 90.00 100.00

Actual temperature (oC)

Pre

dict

ed t

empe

ratu

re (

oC

) (c)

Page 145: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

133 

 

 

 

The ANOVA for the regression model for temperature is given in the table 7.4. The coefficients

with p-values less than 0.05 are considered as significant effect. The p value (<0.0001) and high

Fischer variance value (F-value) (281.50) for the quadratic model implied this model was

significant. From the regression coefficients of the variables, it was found that the linear terms

for all the variables, and interaction term between all the variables, except the interaction

between sodium hydroxide concentration and solid to liquid ratio, and all the quadratic terms in

the model were significant. The effect of interaction terms between variables on the response,

extent of interaction, and the nature of interaction can be obtained from the surface contour plots

of the models. The response surfaces of the significant interaction effects are shown in Figure

7.2. In all the figures, in order to describe the interactive effects of the independent variables on

temperature, the other two variables were kept constant at their mid levels. Figure 7.2.A shows

that, at 4% NaOH concentration the temperature was increased from 70oC to 98oC when the time

of irradiation increased from 10 minutes to 25 minutes, whereas at 8% NaOH concentration, the

temperature was increased from 7oC to 96oC for the same duration of extraction. Effect of

interaction between time and solid to liquid ratio on temperature shows a minimum temperature

for 10 minutes of extraction at a solid to liquid ratio of 1:20, and showed a maximum

temperature for 25 minutes of extraction at a solid to liquid ratio of 1:8 (Figure 7.2.B). Sample

size and time also had a similar effect on the temperature, and the maximum temperature was

observed at longer duration of extraction and for a small sample size (Figure 7.2.C). In Figure

7.2.E, maximum temperature was observed for a larger sample size and a lower solid to liquid

ratio (10 g, 1:8 g/mL). These results indicate that irradiation time had a positive impact on the

temperature, whereas sample size and the solid to liquid ratio had a negative impact under the

experimental conditions studied. Figures 7.2.A and 7.2.D shows the effect of interaction terms

Page 146: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

134 

 

 

 

Table 7.4. Analysis of variance (ANOVA) for the RSM model Source Degree

of freedom

Sum of squares

Mean square

F-value p-value

Dissolution of wood (%)

Model A B C D AB AC AD BC BD CD A2 B2 C2 D2 Residual Lack of fit Pure error

14 1 1 1 1 1 1 1 1 1 1 1 1 1 1 15 10 5

505.53 218.41 219.49 8.33 14.35 1.12 3.10 0.044 5.13 0.25 1.73 1.02 23.22 2.55 3.29 1.92 1.66 0.27

36.11 218.41 219.49 8.33 14.35 1.12 3.10 0.044 5.13 0.25 1.73 1.02 23.22 2.55 3.29 0.13 0.17 0.053

281.50 17.02.64 1711.11 64.94 111.89 8.76 24.15 0.34 39.99 1.91 13.48 7.92 180.98 19.89 25.64 3.11

<0.0001 <0.0001 <0.0001 <0.0001 <0.0001 0.0097 0.0002 0.5664 <0.0001 0.1872 0.0023 0.0131 <0.0001 0.0005 0.0001 0.114 (not significant)

Yield of xylan (%)

Model A B C D AB AC AD BC BD CD A2 B2 C2 D2 Residual Lack of fit Pure error

14 1 1 1 1 1 1 1 1 1 1 1 1 1 1 15 10 5

596.11 54.98 332.97 0.012 55.08 0.45 14.05 3.20 0.014 1.33 36.74 41.41 62.22 18.10 4.29 17.46 12.34 5.12

42.58 54.98 332.97 0.012 55.08 0.45 14.05 3.20 0.014 1.33 36.74 41.41 62.22 18.10 4.29 1.16 1.23 1.02

36.59 47.24 286.10 0.010 47.33 0.38 12.07 2.75 0.012 1.14 31.57 35.58 53.47 15.56 3.69 1.20

<0.0001 <0.0001 <0.0001 0.9209 <0.0001 0.5448 0.0034 0.1181 0.9140 0.3027 <0.0001 <0.0001 <0.0001 0.0013 0.0740 0.4435 (not significant)

Temperature (oC)

Model A B C D AB AC AD BC BD CD A2 B2 C2 D2 Residual Lack of fit Pure error

14 1 1 1 1 1 1 1 1 1 1 1 1 1 1 15 10 5

5508.88 3220.17 42.67 988.17 600 5.06 126.16 76.56 0.56 10.56 95.06 280.50 12.57 39.36 18.57 7.29 5.58 1.71

393.49 3220.17 42.67 988.17 600 5.06 126.16 76.56 0.56 10.56 95.06 280.50 12.57 39.36 18.57 7.29 5.58 1.71

809.47 6624.34 87.77 2032.80 1234.29 10.41 260.36 157.50 1.16 21.73 195.56 577.03 25.87 80.97 38.21 1.630

<0.0001 <0.0001 <0.0001 <0.0001 <0.0001 0.0056 <0.0001 <0.0001 0.2991 0.0003 <0.0001 <0.0001 0.0001 <0.0001 <0.0001 0.3063 (not significant)

Page 147: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

135 

 

 

 

Figure 7.2. Response surface plots showing the interaction between the variables affecting the temperature of wood slurry (A) interaction between time and NaOH solution concentration (B) interaction between time and solid to liquid ratio (C) interaction between time and sample size (D) interaction between sample size and NaOH concentration and (E) interaction between sample size and solid to liquid ratio

between NaOH solution concentration and time (at a constant level of sample size and solid to

liquid ratio: 7.5 g, and 1:14 g/mL), and that of NaOH solution concentration and sample size (at

a constant level of time and solid to liquid ratio; 17.5 minutes and 1:14 g/mL) on temperature.

The NaOH concentration under these experimental conditions did not affect the temperature of

the slurry. Increase in temperature leads to increased hydrolysis and leads to increased wood

dissolution, and so a high yield of xylan is expected at these conditions.

4.00

5.00

6.00

7.00

8.00

10.00

13.00

16.00

19.00

22.00

25.00

60

70

80

90

100

Tem

pera

ture

A: Time B: NaOH Concentration 8.00

11.00

14.00

17.00

20.00

10.00

13.00

16.00

19.00

22.00

25.00

60

70

80

90

100

Tem

pera

ture

A: Time C: Solid to liquid ratio

5.00

6.00

7.00

8.00

9.00

10.00

10.00

13.00

16.00

19.00

22.00

25.00

60

70

80

90

100

Tem

per

ature

A: Time D: Sample size

5.00

6.00

7.00

8.00

9.00

10.00

4.00

5.00

6.00

7.00

8.00

60

70

80

90

100

Tem

pera

ture

B: NaOH Concentration D: Sample size 5.00

6.00

7.00

8.00

9.00

10.00

8.00

11.00

14.00

17.00

20.00

60

70

80

90

100

Tem

pera

ture

C: Solid to liquid ratio

D: Sample size

A B C

DE

Page 148: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

136 

 

 

 

7.5.2 Effectofextractionvariablesonwooddissolution

Wood dissolution during alkaline extraction occurs due to the hydrolysis of the lignin-

carbohydrate linkages. The amount of wood dissolved during the experimentation (Table 7.2)

varied from about 17 wt% (run #14) to 35% (run#19). The highest amount of wood dissolution

during the microwave assisted extraction was observed at high levels of extraction time (+1, 25

minutes) and concentration of sodium hydroxide (+1, 8 wt %), and low levels of sample size (-1,

5 g) and solid to liquid ratio (-1, 1:8 g/mL). On the other hand, the lowest amount of wood

dissolution occurred at a low level of sodium hydroxide (-2, 2 wt %) and mid levels of time (0,

14 minutes), solid to liquid ratio (0, 1:14 g/mL) and sample size (0, 7.5 g). It is clear that higher

concentration of alkali and longer duration of extraction increased the wood dissolution. Longer

duration of irradiation increases the temperature of the system (respective temperatures for the

low and high dissolution of wood were 93.5oC to 98.5oC) as described in the previous section

and this enhances the wood dissolution.

The simultaneous effect of the independent variables on the dissolution of wood is established by

the surface response quadratic model using the CCD (Table 7.3). The relatively high values of R2

(0.9962) and adjusted R2 (0.9927) of the response surface model indicate that the model

considers only the significant terms and so a good agreement between the experimental and the

predicted responses. The predicted R2 (0.9804) was also found to be closer to the adjusted R2

values indicating the adequacy of the model. Relatively smaller standard deviations of the

model (0.36) used for the prediction of wood dissolution indicated the better reproducibility of

the results. Further, adequate precision for the wood dissolution model was reported to be 69.30,

which is well above the required value of 4. The higher values of R2, smaller standard deviation

Page 149: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

137 

 

 

 

and high value of adequate precision of the model demonstrated the suitability of the model for

the prediction of the real relationship among the variables studied to the wood dissolution. Figure

7.3 demonstrates the strong correlation between the predicted and observed wood dissolution

during microwave assisted extraction of xylan.

Figure 7.3. Actual wood dissolution vs. predicted wood dissolution 

The ANOVA for the response surface quadratic models for wood dissolution are given in the

table 7.4. The p-value for the quadratic model for the wood dissolution (<0.0001) and high F-

value (281.50) implied that this model is significant. From the regression coefficients of the

variables, it is clear that the linear terms of all the variables studied, interaction terms between

variables except the two interaction terms (time and sample size; and NaOH solution

concentration and sample size) and the quadratic terms were significant. Significance of the

2

15.00

20.00

25.00

30.00

35.00

15.00 20.00 25.00 30.00 35.00

Actual wood dissolution (%)

Pre

dict

ed w

ood

diss

olu

tion

(%) (a)

Page 150: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

138 

 

 

 

interaction terms between time and NaOH solution concentration (F value, 8.76, p-value 0.0097),

and that of solid to liquid ratio and sample size (F-value 13.48, p-value 0.0023) are less

significant than the interaction terms between time and solid to liquid ratio (F-value 24.15, p-

value 0.0002), and NaOH concentration and solid to liquid ratio (F-value 39.99, p-value

<0.0001).

The response surfaces of the significant interaction effects on the wood dissolution are shown in

Figure 7.4. The percentage of wood dissolution affected by different NaOH concentration and

extraction time was shown in Figure 7.4A, where solid to liquid ratio and sample size were kept

constant at 1:14 (g/mL) and 7.5 g respectively. Extraction time and NaOH concentration time

exhibited a positive impact on the wood dissolution and the wood dissolution increased with the

increase of time of extraction and sodium hydroxide concentration. Maximum wood dissolution

was obtained at an irradiation time of 25 minutes with a 8wt% NaOH solution. Figure 7.4B

depicts the interaction of time and solid to liquid ratio on the wood dissolution at a constant level

of NaOH solution (6 wt %) and sample size (7.5 g). The figure indicated that maximum wood

dissolution occurred for 25 minutes of extraction at a solid to liquid ratio of 1:8 g/mL. Effect of

interaction terms of NaOH solution concentration and solid to liquid ratio on the wood

dissolution is shown in Figure 7.4C. The extractions were performed using 7.5 g of wood sample

for 17.5 minutes. The wood dissolution increased as the level of NaOH concentration increased

Page 151: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

139 

 

 

 

Figure 7.4. Response surface plots showing the interaction between the variables affecting the wood dissolution (A) interaction between time and NaOH solution concentration (B) interaction between time and solid to liquid ratio (C) interaction between solid to liquid ratio and NaOH concentration (D) interaction between solid to liquid ratio and sample size

and a maximum of about 30 wt% was observed when the NaOH concentration was at 8 wt%, and

the solid to liquid ratio was 1:8 g/mL. Increase in the sample size and solid to liquid ratio has a

negative impact on the wood dissolution (Figure 7.4D), when the extraction was performed for

17.5 minutes using 6 wt% NaOH. Hence, 25 minutes of microwave irradiation of the slurry

using 8 wt% of NaOH solution at a solid to liquid ratio of 1:8 (g/mL) is expected to provide a

maximum amount of wood dissolution.

A B

C

D

5.00

6.00

7.00

8.00

9.00

10.00

8.00

11.00

14.00

17.00

20.00

25

26

27

28

29

Woo

d di

ssol

utio

n

C: Solid to liquid ratio

D: Sample size 8.00

11.00

14.00

17.00

20.00

4.00

5.00

6.00

7.00

8.00

22

24

26

28

30

32

Woo

d di

ssol

utio

n

B: NaOH C

C: Solid to liquid ratio

8.00

11.00

14.00

17.00

20.00

10.00

13.00

16.00

19.00

22.00

25.00

22

24

26

28

30

32

Woo

d di

ssol

utio

n

A: Time

C: Solid to liquid ratio 4.00

5.00

6.00

7.00

8.00

10.00

13.00

16.00

19.00

22.00

25.00

20

22

24

26

28

30

32

34

Woo

d di

ssol

utio

n

A: Time

B: NaOH Concentration

Page 152: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

140 

 

 

 

7.5.3 Effectofextractionvariablesonyieldofxylan

The yield of xylan (based on oven dry wood) obtained in the CCD design are given in the table

7.2. The lowest and highest yields of xylan obtained in the design space were 4.33% and 22.49%

respectively (run #14 and #15). Lowest yield of xylan observed for the experiment corresponds

to the lowest amount of wood dissolution where the levels of independent variables were:

sodium hydroxide (-2, 2 wt %), time (0, 14 minutes), solid to liquid ratio (0, 1:14 g/mL) and

sample size (0, 7.5 g). Contrary to the maximum wood dissolution (run# 19), the highest yield of

xylan (run #15) was obtained when all the independent variables are at the higher levels;

extraction time (+1,25 minutes), concentration of sodium hydroxide (+1, 8 wt%), solid to liquid

ratio (+1, 1:20 g/mL), and sample size (+1, 10 g). The yield of xylan obtained for run #19, where

maximum wood dissolution was observed, was 14.9%. The variation in yield can be attributed to

the rise in temperature of the wood slurry. The temperature of the slurry after run #19 was

98.5oC, whereas that of run #15 was 83.5oC. High wood dissolution is expected at higher

temperature, as the hydrolysis rate and dissolution of the hydrolyzed components (hemicelluloses

and lignin) increase with the temperature. However, the high temperature generated lead to the

degradation of the hemicelluloses resulting in a low yield of xylan, which is in agreement with

the earlier results discussed in Chapter 5.

The effect of independent variables on the yield of xylan was studied using the surface response

quadratic model given in the table 7.3. The model was found to be significant from the values of

regression coefficients (R2 = 0.9715 and adjusted R2 = 0.9450). The predicted R2 (0.8722) was

closer to the adjusted R2 and would give a good fit to the statistical model used. The observed

small standard deviation of the model (1.08) and the high adequate precision value (22.91) also

indicates the reproducibility and applicability of the quadratic model for the prediction of yield

Page 153: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

141 

 

 

 

of xylan. Figure 7.5 shows the predicted and actual yield of xylan obtained during the

experimental design studied and the strong correlation observed (R2=0.9856) indicates the

accuracy of the model.

Figure 7.5. Actual yield of xylan vs. predicted yield of xylan

Analysis of variance (ANOVA) of the regression model is given in the table 7.4 and was used to

evaluate the statistical significance of the model. The F-value for the quadratic model for the

yield of xylan is 36.59 with a p-value <0.0001 indicating the model is significant. The effect of

linear terms of extraction time, NaOH solution concentration, and sample size, interaction terms

between time and solid to liquid concentration and between solid to liquid ratio and sample size,

2

0.00

5.00

10.00

15.00

20.00

25.00

0.00 5.00 10.00 15.00 20.00 25.00

Actual yield of xylan (%)

Pre

dict

ed y

ield

of x

ylan

(%) (b)

Page 154: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

142 

 

 

 

and the quadratic terms of time, NaOH concentration, and solid to liquid ratio are found be

significant on the yield of xylan. The Lack of Fit F-value of 1.20 (p-value 0.4435) imply the

lack of fit is not significant relative to the pure error and hence the model can be used to predict

the yield of xylan in the present study.

Figure 7.6. Response surface plots showing the interaction between the variables affecting the

yield of xylan (A) interaction between time and solid to liquid ratio (B) interaction

between solid to liquid ratio and sample size

The response surface contour plots of the significant interaction terms are shown in Figure 7.6.

Figure 7.6A depicts effect of the interaction between time and solid to liquid ratio on the yield of

xylan (NaOH concentration and sample size was kept at 6 wt% and 7.5 g respectively), whereas

7.6B shows the effect of interaction between solid to liquid ratio and sample size (NaOH

concentration and time of extraction was kept at a constant level of 6 wt% and 17.5 minutes

respectively) on the yield of xylan. Larger sample size and smaller solid to liquid ratio resulted in

a higher yield of xylan (Figure 7.6B), whereas longer extraction time and larger solid to liquid

8.00

11.00

14.00

17.00

20.00

10.00

13.00

16.00

19.00

22.00

25.00

0

5

10

15

20

25

Yie

ld

A: Time C: Solid to liquid ratio 5.00

6.00

7.00

8.00

9.00

10.00

8.00

11.00

14.00

17.00

20.00

0

5

10

15

20

25

Yie

ld

C: Solid to liquid ratio D: Sample size

A B

Page 155: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

143 

 

 

 

ratio resulted comparatively higher yield. It is clear from all these results that larger sample size,

higher concentration of NaOH, longer irradiation time, and smaller solid to liquid ratio would

lead to a higher yield of xylan.

7.5.4 Optimizationofmicrowaveassistedextractionofxylanandvalidationofthe model

The objective of the study was to find the microwave assisted extraction conditions that result in

the highest yield of xylan from birch wood. The optimal extraction conditions for obtaining the

highest yield of xylan were extracted by the Design Expert software using a numerical

optimization. Several potential solutions were provided by the software, and based on the

experimental feasibility, the following optimum operating settings were selected for obtaining a

higher yield of xylan: time (A) = 25 minutes, NaOH concentration (B) = 8 wt%, solid to liquid

ratio (C) = 1:8 g/mL, and sample size (D) = 10 g. Four replicates of the extraction were

performed at this optimal extraction conditions. The results of the extraction with the respective

predicted values are given in the table 7.5. Based on the numerical optimization condition

selected, the maximum yield predicted was 22.66% (OD wood basis) with a standard deviation

of 1.08. The predicted wood dissolution, and temperature at this point were 33.53% (standard

deviation 0.36) and 96.46oC (standard deviation 0.69) respectively. The yield of xylan, wood

dissolution, and temperature of the slurry obtained were 21.27±2.65%, 33.78±0.21%, and

97.67±0.47 respectively. The experimental values obtained were found to be in good agreement

with the values calculated from the models suggesting that the quadratic surface models were

adequate for the optimization of microwave assisted extraction under study. Wood used for this

study contained 29.4% of xylan. Conversion of the yield of xylan obtained based on the oven dry

Page 156: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

144 

 

 

 

Table 7.5. Experimental and predicted values of wood dissolution, yield of xylan, and temperature at the optimum extraction conditions used for the alkaline extraction of xylan

Optimized extraction conditions

Wood dissolution (%) Yield of xylan (%) Temperature (oC) Experimental Predicted Experimental Predicted Experimental Predicted

Time, (min) NaOH concentration (wt%) Solid to liquid ratio (g:mL) Sample size (g)

25 8 8 10

33.78 ± 0.21

33.53 ± 0.36

21.27 ±2.6

22.66 ±1.08

97.63 ±0.5

96.46 ±0.69

basis of wood to that of the original xylan present in wood indicate that about 72.5 % of the

xylan present in birch wood can be extracted by 25 minutes of microwave assisted extraction .

Though there is no available literature for direct comparison, Sun et al. (1995) reported that a

low temperature alkaline extraction of wheat provided 76.4% of xylan, when wheat straw was

treated with 1.5% NaOH at solid loading of 2.5% dry matter for 144 hours at room temperature.

It is clear that great reduction in the extraction time can be obtained using microwave assisted

extraction for a similar yield of xylan. In a recent study (Obermeier et al., 2012) using

microwave assisted alkaline extraction, a similar type of xylan yield (73.6%) was reported in the

supernatant obtained from wheat straw extraction. The respective reaction conditions reported

were 5% NaOH solution at a temperature of 140oC with 10 minutes preheating time and 10

minutes of residence time for 3.5 g of wheat straw with a solid to liquid ratio of 1:10 using an

input power of 265 W. A total of 14 g of samples were used for extraction in a single batch. The

energy input for the extraction was calculated as the multiple of the power input into the reaction

chamber and the sum of the preheating and residence time in seconds. The energy calculated for

this extraction was 318 kJ and the energy input per g of wheat straw calculated as 22.7 kJ/g. In

our study, the energy input at the optimized extraction conditions for getting maximum yield of

Page 157: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

145 

 

 

 

xylan was calculated as 66 kJ and the energy input per g of wood fibers was 6.6 kJ/g. Though it

is not appropriate to compare two different raw materials, it is clear that optimization of the

reaction conditions are necessary to obtain maximum yield of xylan with lower input of energy.

Figure 7.7. ηsp/C vs. Concentration of xylan obtained at the optimal conditions of microwave extraction

The high temperature generated at the optimized conditions of extraction might degrade the

degree of polymerization of the polymer. As expected, viscosity measurement of the polymer in

CED solution showed a low intrinsic viscosity of 0.605 (Figure 7.7) and the corresponding

molecular weight and degree of polymerization calculated was 16564 and 128 respectively.

Comparison of the results discussed in section 5.3.4.3, the xylan obtained at the optimal

conditions is similar to the xylan obtained after 30 minutes of irradiation using 4 wt% NaOH

solution for 30 minutes.

y = 0.7311x + 0.6046R² = 0.9764

0.4

0.5

0.6

0.7

0.8

0.9

1

1.1

0 0.2 0.4 0.6 0.8

ηsp

/C (

dL

/g)

Concentration of xylan solution (g/dL)

Page 158: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

146 

 

 

 

7.6 Conclusions

Central composite design and surface response methodology were employed to optimize the low-

temperature microwave assisted alkaline extraction of xylan from birch wood. The effect of four

extraction parameters (time, alkali concentration, solid to liquid ratio and sample size) on the

temperature of the wood slurry, wood dissolution, and yield of xylan were studied. Three

quadratic polynomial models were developed to correlate the extraction variables with the three

responses. Analysis of variance (ANOVA) of the models developed indicated the statistical

significance of the models for the prediction of each response studied. Temperature of the slurry,

and wood dissolution were significantly influenced by all the four extraction parameters,

whereas the yield of extraction was not affected by the linear effect of solid to liquid ratio.

Process optimization was conducted to maximize the yield of xylan and the experimental values

obtained for the respective yield of xylan, wood dissolution and temperatures were found to

agree satisfactorily with the quadratic model used for xylan extraction. Optimum conditions used

for the maximum extraction of xylan from birch wood under microwave irradiation were: 10 g of

wood fibers, 8 wt% of NaOH solution, 1:10 solid to liquid ratio (g: mL) and 25 minutes of

irradiation time.22 wt% of xylan (OD) was obtained from about 34% of the dissolved wood at

the optimal extraction conditions.

Page 159: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

 

 

147 

 

Chapter8 SummaryandConclusions 

The following conclusions were drawn from the various results discussed in this thesis.

a) Direct interaction of microwaves with the wood slurry reduce the time required for the

extraction of xylan from wood fibers compared to conventional extraction

b) Ten minutes of microwave assisted extraction of wood using a power input of 110 W offered

about 60% of high molecular weight xylan polymers with degree of polymerization of 150.

c) Xylan obtained from microwave assisted extraction provided a higher number average degree

of polymerization compared to conventional process under similar conditions (150 vs. 125 for 20

minutes of extraction)

c) Interaction of microwaves with the alkali in the fibers resulted in a rapid increase in the

temperature and produced an "explosion effect” that loosened the recalcitrant lignocellulosic

structure and increased the porosity of the fiber thus enhancing the mass transfer of the

hydrolyzed components to the solution

d) During microwave assisted extraction of xylan, temperature as well as the fibre structure

rupture leads to the enhanced hydrolysis and dissolution of wood.

e) SEM analysis of the fibers after extraction demonstrated the "explosion effect" induced during

the microwave extraction and this was supported by the internal structure exploration by X-ray

tomography. Porosity of the wood fibers after microwave extraction was higher than the

conventional extraction.

Page 160: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

148 

 

 

 

f) Wood dissolution during the extraction process consisted of two stages, whereas the yield

obtained was assumed to be the consequence of three stages including the dissolution of easily

accessible xylan, followed by the steady increase in the yield where hydrolysis and dissolution of

xylan is higher, which is then followed by the decrease in yield, where degradation of the

dissolved polymer is higher than the dissolution of the hydrolyzed polymer.

g) Temperature of the extraction was found to be related with the yield of xylan as well as the

degree of polymerization, the increase in temperature and time of extraction decreased the yield

of xylan as a result of the molecular breakdown and hence the degree of polymerization.

h) Degradation of xylan in alkali was found to be significant at temperatures above 95oC.

i) Decrease in the crystallinity of the fibers after microwave extraction also supported the

hypothesis of the "explosion effect" and the structural changes produced by the interaction of

microwave with the alkali in the fiber slurry.

j) X-ray microtomography can be used as technique to determine the internal temperature

generated inside the fibers, and demonstrated that a 10 minute of microwave extraction using a

power input of 110 W makes structural changes equivalent to the temperature of 120oC.

k) Optimization of the microwave extraction process demonstrated that 25 minutes of microwave

extraction using 10g of the fibers at a solid to liquid ratio of 1:10 with 8 wt% NaOH solution can

yield up to 75% of the xylan present in the fibers.

Page 161: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

 

 

149 

 

Chapter9 RecommendationsforFutureResearch

Extraction of xylan from hard wood species birch has demonstrated that microwave assisted

extraction can considerably reduce the time of the conventional extraction. This research

established a synergetic microwave effect produced by the hot spot or explosion effect of the

fibers accelerated the extraction of the xylan polymers from the woody biomass. The study also

established that a low power input microwave can be more efficient for the extraction of the

polymeric xylan compared to high energy input and this method can be used as an effective

alternate to the conventional extraction of xylan polymers. In this section, recommendations for

future research in terms of the extension of the current research and the future necessary research

needed for the implementation and application of the microwave assisted extraction.

The dominant factors governing the extraction of the hemicelluloses from the lignocellulosic

matrix by microwave assisted extraction are the solubility of the hemicelluloses in the solvent,

the strength of hemicellulose/matrix interaction and the mass transfer kinetics of the

hemicelluloses from the matrix to the solution phase.

Interaction of hemicelluloses with other structural components such as cellulose and lignin

highly depends on the origin or type of the lignocellulosics under study. Extension of the

process of microwave extraction to other lignocelluloses such as other hard wood species, soft

wood species, as well as agro-residues such as wheat straw, bagasse, or corn stover would be

needed to investigate the effect of structural differences of the biomaterials on the microwave

Page 162: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

150 

 

 

 

extraction of the hemicelluloses and also to establish the effectiveness of the method. Effect of

pre-treatments of the lignocellulosic materials such as delignification and size reduction are also

to be considered respectively for further elucidation of the structural effects on the microwave

extraction, and to provide additional knowledge on the optimum size of the materials used for an

efficient extraction of polymers using microwave heating.

An important component of the engineering aspect of the process development is the

understanding of the process kinetics. Extraction process kinetics becomes difficult when there is

mass transfer along with chemical reactions. Understanding the mass transfer mechanisms and

kinetics is important for the proper designing of the equipment as well as optimizing the process

conditions during the scale-up of the process. Hence, investigation and establishment of the

extraction kinetics of the microwave assisted extraction has to be explored. This study can also

focus on the development of models to predict the mass transfer kinetics as well as the validation

of the developed models.

For the successful demonstration and implementation of the microwave extraction technology,

energy efficiency and economic viability (cost analysis) of the process has to be studied

compared to the conventional method of extraction processes.

Application potential of the lignocellulosic residue after hemicelluloses extraction and

modification of the polymeric hemicelluloses are the two other areas of research that would be

highly beneficial for the biorefinery industry.

Page 163: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

 

 

151 

 

References

Al-Dajani, W. F., Tschirner, U.W., Jenson, T. (2008). Pre extraction of hemicelluloses and

subsequent kraft pulping. Part II, Acid and autohydrolysis. Tappi Journal, 3-8

Alén, R., 2000. Structure and chemical composition of wood. In: Stenius, P. (Ed.), Forest

Products Chemistry. Chapter 1, pp. 11–57

Allen, S. G., Schulman, D., Lichwa, J., et al. (2001). A Comparison of Aqueous and Dilute-

Acid Single-Temperature Pretreatment of Yellow Poplar Sawdust. Industrial and Engineering

Chemistry Research. 40(10): 2352-2361

Amash, A.; Zugenmaier, P. (1998). Study on cellulose and xylan filled polypropylene

composites. Polymer Bulletin. 40: 251–258

Amidon, T.E., Liu, S.J. (2009). Water-based woody biorefinery. Biotechnology Advances.

27: 542–50.

Aspinall, G.O. (1959). Structural chemistry of the hemicelluloses. Advances in Carbohydrate

Chemistry. 14: 429-468

Azuma J., Tanaka F., Koshijima, T. (1984) Enhancement of enzymatic suceptibility of

lignocellulosic wastes by microwave irradaiation. Journal of Frermentation Technology. 62:

377-84

Baba, K., Sone, Y., Misaki, A., Hayashi, T. (1994). Localization of Xyloglucan in the

Macromolecular Complex Composed of Xyloglucan and Cellulose in Pea Stems. Plant Cell

Physiology. 35: 439–444.

Page 164: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

152 

 

 

 

Ballestors, I., Oliva, J.M., Cabanas, A., Manzanares, P., Ballestors, M. (2000). Effect of chip

size on steam explosion treatment of softwood. Applied Biochemistry and Biotechnology.

84: 97-110

Barnett, E. R., Dikeman, R., Pantaleon, D. P. et al. (1989). European Patent EP 0,301,440 A1

Barcelon, E.G., Tojo, S., Watanabe, K. (1999). X-ray computed tomography for internal

quality evaluation of peaches. Journal of Agricultural Engineering Research. 73:323-330.

Benko, Z., Andersson, A., Szengyel, Z., Gasper, M., Reczey K., et al. (2007). Heat extraction

of corn fiber hemicelluloses. Applied Biochemistry and Biotechnology. 137-140: 253-265

Bhaskaran, T.A., von Koeppen, A. (1970). The degradation of wood carbohydrates during

sulphate pulping. Holzforschung. 24: 14-19

Bikova, T., Treimanis, A. (2002). Solubility and molecular weight of hemicelluloses from

Almus incana and Almus glutinosa, Effect of tree age. Plant Physiol. Biochem. 40:347

Biosucceed, A USDA higher education challenge program, Course module 6; Fundamentals

of polymer chemistry, www.nscdu.edu/biosucceed/courses

Bond, G., Moyers, R.B., Whan, D.A. (1993). Recent applications of microwave heating in

catalysis. Catalysis Today. 17: 427-437

Borysiak, S., Doczekalska, B. (2005). X-ray diffraction study of the pine wood treated

NaOH. Fibers and Textiles in Eastern Europe 13(5): 87

Boussaid, A., Robinson, J., Cai, Y., Gregg, D.J., Saddler, J.N. (2001). Sugar recovery and

fermentability of hemicellulose hydrolysates from steam-exploded softwoods containing bark.

Biotechnology Progress. 17(5): 887-892

Brasch, D. J., Free, K. W. (1965). Prehydrolysis-kraft pulping of Pinus radiata grown in New

Zealand. Tappi Journal. 48(4): 245-248

Page 165: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

153 

 

 

 

Brett, C.T., Wladron, K.W. (1996). Physiology and Biochemistry of Plant Cell Walls (2nd

Ed.). Chapman and Hall, London.

Brownell, H.H., Yu, E.K.C., Saddler J.N. (1986). Steam-explosion pretreatment of wood:

Effect of chip size, acid, moisture content and pressure drop. Biotechnology and

Bioengineering. 28:792-801

Brownell, H.H., Saddler J.N. (1987). Steam pretreatment of lignocellulosic material for

enhanced enzymatic hydrolysis. Biotechnology and Bioengineering. 29:228-235

Brudecki, G., Cybulska, I., Rosentrater, K. (2013). Optimization of clean fractionation

process applied to switch grass to produce pulp for enzymatic hydrolysis. Bioresource

Technology. 131:101–112.

Burnov, A.U., Mazza, G. (2010). Extraction and characterization of hemicelluloses from flax

shives by different methods. Carbohydrate Polymers. 79: 17-25

Caddick, S. (1995). Microwave assisted organic reactions. Tetrahedron. 51: 10403-32

Camacho, F., Gonzallez-Tello, P., Jurado,E., Robles, A. (1996). Micro-crystalline cellulose

with concentrated sulfuric acid. Journal of Chemical Technology and Biotechnology. 67: 350-

356

Carrasco, J.E., Saiz, M.C., Navarro, A., Soriano, P. Saef, F., Martinez, M. (1994). Effect of

dilute acid and steam explosion pretreatments on lignocellulosic materials. Applied

Biochemistry and Biotechnology. 45/46: 23-34.

Carvalheiro, F., Duarte, L. C., Girio, F. M. (2008). Hemicelluloses biorefineries: A review on

biomass pretreatments. Scientific and Industrial Research Journal. 67: 849-864.

Casebier, R.L., Hamilton, J. K. (1965). Alkaline degradation of glucomannans and

galactoglucomannans. Tappi Journal. 48(11): 664-669 .

Page 166: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

154 

 

 

 

Casebier, R.L., Hamilton, J.K., Hergert, H.L. (1969). Chemistry and mechanism of water

prehydrolysis on southern pine wood. Tappi Journal. 52(12): 2369-2377

Casebier, R.L., Hamilton, J.K., Hergert, H.L. (1973). The chemistry and mechanism of water

prehydrolysis on black gumwood. Tappi Journal. 56(3): 135-139

Cave, I.D. (1997). Theory of X-ray measurement of microfibril angle. Wood Science and

Technology. 31: 143

Chakar, F. S., Ragauskas, A.J. (2004). Review of current and future softwood kraft lignin

process chemistry. Industrial Crops and Products. 20(2): 131-141

Chemat, S., Ait-Amar, H., Lagha, A., Esveld, D.C. (2005). Microwave assisted extraction

kinetics of terpenes from caraway seeds. Chemical Engineering Proceedings. 44: 1320.

Chen, S.S., Spiro, M. (1994). Study of microwave extraction of essential oil constituents from

plant materials. Journal of Microwave Power. 29: 231-241

Chen, S.S., Spiro, M. (1995). Kinetics of microwave extraction of rosemary leaves in hexane,

ethanol and a hexane + ethanol mixture. Flavour and Fragrance Journal. 10: 101-112

Chen, J., Jo, S., Park, K. (1995). Polysaccharide hydrogels for protein drug delivery.

Carbohydrate Polymers. 28: 69-76

Chornet, E., Overend, R. P. (1988). Phenomenological Kinetics and. Reaction Engineering

Aspects of Steam/Aqueous Treatments. In: Proceedings of the international workshop on

steam explosion techniques: Fundamentals and industrial applications, pp 21-58

Conner, A.H. (1984). Kinetic modeling of hardwood prehydrolysis. 1. Xylan removal by

water prehydrolysis. Wood fiber Science. 16 (92): 268-277

Page 167: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

155 

 

 

 

Conner, A. H., Lorenz, L. F. (1986). Kinetic modeling of hard wood prehydrolysis.III.

Water and dilute acetic acid prehydrolysis of southern red oak. Wood fiber Science. 18(2):

248-263

Cosgrove, D.J. (2005). Growth of the plant cell wall. Nature Reviews, Molecular Cell

Biology. 6: 851-861

Coviello, T., Matricardi, P., Marianecci, C., Alhaique, F. (2007). Polysaccharide hydrogels for

modified release formulations. Journal of Controlled Release. 119: 5-24

Crittenden, R. G., Playne, M. J. (1996). Production, properties and applications of food-grade

oligosaccharides. Trends in Food Science and Technology. 7: 353–361.

Dashek, W.V. (1997). In: Methods in plant biochemistry and molecular biology. Dashek

W.V. (Ed). CRC Press, New York, pp.29-47

Datta, A.K. (2001). Fundamentals of heat and moisture transport for microwavable food

product and process development. In: Datta AK, Anantheswaran RC (eds.) Handbook of

microwave technology for food applications, Marcel Dekker INC

Delong, E.A. (1983). Canadian Patent, Number 1,141,376

Dubois, M., Gilles, K., Hamilton, J.K, et al. (1979). Colorimetric method for the

determination of sugars and related substances. Analytical Chemistry. 28:350-356

Dutton, G.G.S., Hunt, K. (1958). The constitution of hemicelluloses. Part II. Structure of the

mannan portion. Journal of American Chemical Society. 80: 5697-5701

O’Dwyer, M.H. (1923). The Hemicelluloses. III. The Hemicellulose of American White Oak.

Biochemistry Journal. 17 (4-5): 501-509

Ebringerová, A., Hromadkova, Z., Kacurakova, M., Antal, M. (1994). Quarternized xylans-

synthesis and structural characterization. Carbohydrate Polymers. 24: 301-308

Page 168: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

156 

 

 

 

Ebringerova. A., Heinze, T. (2000). Xylan and xylan derivatives – biopolymers with valuable

properties, 1. Naturally occurring xylans structures, isolation procedures and properties- A

review. Macromolecular Rapid Communications. 21: 542-556

Ebringerová, A. (2005). Structural diversity and application potential of hemicelluloses.

Macromolecular Symposia. 232 (1): 1-12

Ender Us, N., Perendeci, A. (2012). Improvement of methane production from greenhouse

residues: Optimization of thermal and H2SO4 pretreatment process by experimental design.

Chemical Engineering Journal. 181– 182: 120– 13.

Eriksson, I., Haglind, I., Lidbrandt, O., Salmén, L. (1991). Fiber swelling favoured by lignin

softening. Wood Science Technology. 25:135–144.

Eriksson, O., Lindgren, B.O. (1977). About the linkage between lignin and hemicelluloses in

wood. Svensk Papperstidning 80(2): 59-63

Eriksson, O., Goring, D.A.I, Lindgren, B.O. (1980). Structural studies on the chemical bonds

between lignin and carbohydrates in spruce wood. Wood Science and Technology. 14: 267-

279

Erins, P., Cinite, V., Jacobsons M., Gravitis, J. (1976). Wood as a multi component cross

linked polymer system. Applied Polymer Symposium. 28: 1117-1138

Evtuguin, D.V., Tomas, J.L., Silva, A.M.S., Net, C.P. (2003). Characterization of an

acetylated heteroxylan from Eucalyptus. Carbohydrate Research. 338: 597-604

Faix, O., Bremer, J., Schmidt, O., Stevanovic, T. (1993). Monitoring of chemical changes in

white-rot degraded beech wood by pyrolysis-gas chromatography and Fourier transform

infrared spectroscopy, Journal of Analytical and Applied Pyrolysis. 21: 147–162.

Page 169: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

157 

 

 

 

Fang, J.M., Sun, R.C., Tomkinson, J.( 2000). Isolation and characterization of hemicelluloses

and cellulose from rye straw by alkaline peroxide extraction. Cellulose. 7: 87-107

FAO. (2003). State of the World’s Forests 2003, FAO, Rome, Italy, pp. 143–145.

Franzon, O., Samuelsson, O. (1957). Degradation of cellulose by alkali cooking. Svensk

Papperstidning. 60(23): 872-877.

Fromm, H., Sautter, I., Matthies, D., Kremer, J., Schumacher, P. et al. (2001). Xylem water

content and wood density in spruce and oak trees detected by high resolution computed

tomography. Plant Physiology.127: 416-425.

Fengel, D. (1971). Ideas on the structural organization of the cell wall components. Journal

of Polymer Science C. 36: 383-392

Fengel, D., Wegener, G. (1984). Wood: Chemistry, Ultrastructure, Reactions. Walter de

Gruyter and Co., Berlin.

Foner, H.A., Adan, N. (1983). The Characterization of papers by X-ray diffraction. Journal of

Forensic Science Society. 23(4): 313-321.

Fournier, E. (2001). Colorimetric quantifiaction of carbohydrates in Current protocols in

Food. Anal Chem E1.1.1-E1.1.8., John Wiley and Sons, Inc.

Freudenberg, K., Grion, G. (1959). Beitrag zum Bildungsmechanis- mus des Lignins und der

Lignin-Kohlenhydratbindung. Chemische Berichte. 92: 1355-1363

Freudenberg, K., Harkin, J.M. (1960). Modelle für die Bindung des Lignins an die

Kohlenhydrate. Chemische Berichte. 93: 2814-2819

Fry, S. C. (1989). Cellulases, hemicellulases and auxin-stimulated growth: A possible

relationship. Physiology Plant. 75: 532–536.

Page 170: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

158 

 

 

 

Gabriel, C., Gabriel, S., Grant, E.H., et al. (1998). Dielectric parameters relevant to

microwave dielectric heating. Chemical Society Review. 27: 213-224.

Gabrielii, I., Gatenholm, P., Glasser, W.G., Jain, R.K., Kenne, L. (2000) Separation,

characterization and hydrogel formation of hemicelluloses from aspen wood. Carbohydrate

Polymers. 43: 367-374

Galbe, M., Zacchi, G. (2007). Pretreatment of lignocellulosic materials for efficient

bioethanol production. Advances in Biochemistry Engineering and Biotechnology. 108: 41-

65

Ganzler, K., Salgo, A., Valko, K. (1896). Microwave extraction. A novel sample preparation

method for chromatography. Journal of Chromatography. 371:299 -306

Garcia, R.B., Nagashima, T., Praxedes, A.K.C., Raffin, et al. (2001). Preparation of micro and

nanoparticles from corn cobs xylan. Polymer Bulletin. 46:371-379

Garrote, G., Dominguez, H., Parajo, J.C. (1999a). Mild autohydrolysis: an environmentally

friendly technology for xylooligosacharide production from wood. Journal of Chemical

Technology and Biotechnology.74(11): 1101-1109

Garrote, G., Dominguez, H., Parajo, J.C. (1999b). Hydrothermal processing of lignocellulosic

materials. Holz als Roh-und Werkstoff. 57: 191-202

Garrote, G., Parajo, J.C. (2002). Non-isothermal autohydrolysis of Eucalyptus wood. Wood

Science and Technology. 36(2): 111-123

Garrote, G., Dominguez, H., Parajo, J.C. (2004). Production of substituted oligosaccharides

by hydrolytic processing of barley husks. Industrial and Engineering Chemistry Research.

43(7): 1608-1614

Page 171: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

159 

 

 

 

Ghosh P., Singh, A. (1993). Physico-chemical and biological treatments for

enzymatic/microbial conversion of lignocellulosic biomass. Advances in Applied

Microbiology. 39: 295

Girio, F.M., Fonseca, C., Carvalheiro, F. et al. (2010). Hemicelluloses for fuel ethanol: A

review. Bioresource Technology. 101: 4775-4800

Glasser, W.G., Jain, R.K., Sjostedt, M.A. (1995). US Patent # 5,430,142

Glasser, W. G., Kaar, W.E., Jain, R.K. (2000). Isolation options for non-cellulosic

heteropolysaccharides. Cellulose. 299-317

Goh, C. S., Tan, H. T., Lee, K. T., Brosse, N. (2011) Evaluation and optimization of

organosolv pretreatment using combined severity factors and response surface methodology.

Biomass and Bioenergy. 9: 4025–4033

Goldstein, I.S. (1983). Acid processes for cellulose hydrolysis and their mechanisms. Wood

and Agricultural residues. Academic Press Inc., New York. pp. 315-328

Goring, D.A.I., Timell, T. E. (1960). Molecular properties of six 4-O-methylglucurono

xylans. Journal of Physical Chemistry. 64: 1426-1430

Gustavsson, C.A.S., Al-Dajani, W.W.( 2000). The influence of cooking conditions on the

degradation of hexenuronic acid, xylan, glucomannan, and cellulose during kraft pulping of

softwood. Nordic Pulp and Paper Research Journal. 15(2): 160-67 .

Gustavsson, M., Bengtsson, M., Gatenholm, P., et al. (2001). In: Biorelated Polymers:

Sustainable Polymer Science and Technology, Eds. Chiellinin, E., Gil, H., Braunegg G., et

al., Kluwer Academic/Plenum Publishers, New York. pp. 41-52

Harker, F.R., Hallet, I.C. (1992). Physiological changes associated with development of

mealiness of apple fruit during cool storage. Journal of Horticultural Science. 27:1291-1294.

Page 172: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

160 

 

 

 

Haas, D. W., Hrutfiord, B.F., Sarkanen, K.V. (1967). Kinetic study on the alkaline

degradation of cotton hydrocellulose. Journal of Applied Polymer Science. 11: 587-600

Hamilton, J.H., Quimby, G.R. (1957). The extractive power of lithium, sodium, and

potassium hydroxide for hemicelluloses associated with wood cellulose and holocellulose

from western hemlock. Tappi Journal. 40(9): 781-786

Hamilton, J.K., Thompson, S. N. (1959). A comparison of the carbohydrates of hardwoods

and softwoods. Tappi Journal. 42: 752-760

Hansen, M.L., Placket, D. (2008). Sustainable films and coatings from hemicelluloses: A

review. Biomacromolecules. 9(6): 1493-1505.

Hayashi, T. (1989). Xyloglucans in the primary cell wall. Ann. Rev. Plant Physiol: Plant

Molecular Biology. 40: 139–168

Hawthorne, S.B., Galy, A.B., Schnitt, V. O., Miller, D. (1995). Effect of SFE flow rate on

extraction rates: classifying sample extraction behavior. Journal of Analytical Chemistry. 67:

2723-2732

Heitz, M., Carrasco, F., Rubio, M. et al. (1986). Generalized correlations for the aqueous

liquefaction of lignocellulosics. Canadian Journal of Chemical Engineering. 64(4): 647-650.

Heitz, M., Capek-Menard, E., Koeberle, P.G. et al. (1991). Fractionation of Populus

tremuloides at the pilot plant scale: Optimization of steam pretreatment conditions using the

STAKE II technology. Bioresource Technology. 35(1):23-32

Helmerius, J., Von Walter, V. J., Rova, U., Berglund, K. A., Hodge, D. B. (2010). Impact of

hemicelluloses pre-extraction for bioconversion on birch kraft pulp properties. Bioresource

Technology. 101: 5996-6005.

Page 173: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

161 

 

 

 

Henriksson, A., Gatenholm, P. (2001). Controlled assembly of glucuronoxylans onto cellulose

fibers. Holzforschung. 55: 494–502.

Hirst, E. L. (1962). The chemical structure of the hemicelluloses, Proceedings of the Wood

Chemistry Symposium, Montreal, Canada, London, Butterworth53-66

Hosseini, S.A., Shah, N. (2009). Multiscale modeling of hydrothermal biomass pretreatment

for chip size optimization. Bioresource Technology. 100: 2621–2628.

Hoz, A.D., Díaz-Ortiz, A., Moreno, A. (2005). Microwaves in organic synthesis. Thermal and

non-thermal microwave effects. Chemical Society Review. 34: 164–178

Hu, Z. H., Wen, Z. Y. (2008). Enhancing enzymatic digestibility of switch grass by

microwave assisted alkali pre treatment. Biochemical Engineering Journal. 38(3): 369-378.

Huang, H.J., Ramaswamy, S., Tschirner, U.W., Ramrao, B.V. (2008). A review of separation

technologies in current and future biorefineries. Separation and Purification Technology. 62:

1-21

Jacobs, A., Dahlman, O. (2001). Characterization of the molar masses of hemicelluloses from

wood and pulps employing size exclusion chromatography and matrix-assisted laser

desorption ionization time-of-flight mass spectrometry. Biomacromolecules. (2): 894–905

Jacobs, A., Lundqvist, J., Stalbrand, H., Tjerneld, F., Dahlman, O. (2002). Characterization of

water-soluble hemicelluloses from spruce and aspen employing SEC/MALDI mass

spectroscopy. Carbohydrate Research. 337:711-7

Jacobs, A., Palm, M., Zacchi, G., Dahlman, O. (2003). Isolation and characterization of water

soluble hemicelluloses from flax shives. Carbohydrate Research. 338: 1869-1876.

Jain, R.K., Sjostedt, M., Glasser, W.G. (2000). Thermoplastic xylan derivatives with

propylene oxide. Cellulose. 7: 319-336

Page 174: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

162 

 

 

 

Jiang, Z,H., Yang, Z., So, C.L., Hse C.Y. (2007). Rapid prediction of wood crystallinity in

Pinus Elliotti plantation wood by near- infrared spectroscopy. Journal of Wood Science. 53:

449-453.

Jones, D.A., Lelyved, T.P., Mavrofidis, S.D., Kingman, S.W., Miles, N.J. (2002).

Microwave applications in environmental engineering - a review. Resources, Conservation

and Recycling. 34: 75-90.

Joseleau, J.P., Kesaraoui, R. (1986). Glycosidic bonds between lignin and carbohydrates.

Holzforschung. 40: 163-168

Juslin, M., Paronen, P. J. (1984). Xylan – a possible filler and disintegrant for tablets.

Pharmacy and Pharmacology. 36: 256-257

Kappe, C.O. (2008). Microwave dielectric heating in synthetic organic chemistry. Chemical

Society Reviews. 37(6):1127-1139

Keshwani, D.R. (2009). Microwave pretreatment of Switch grass for bio-ethanol production.

Ph.D. Thesis, North Carolina State University, North Carolina, United States.

Kingston, H.M., Haswell, S.J. (1997). (Eds.) Microwave-Enhanced Chemistry –

Fundamentals, Sample Preparation, and Applications. American Chemical Society,

Washington

Kitchiya, P., Intankul, P., Krairish, M. (2003). Enhancement of enzymatic hydrolyiss of

lignocellulosic wastes by microwave pre treatmet under atmospheric pressure. Journal of

Wood Chemistry and Technology. 23: 217-25

Kleppe, P.J. (1970) Kraft pulping- Review. Tappi Journal. 53(1): 35-47

Korner, I., Faix,O., Wienhaus,O. (1992). Attempts to determine the degradation of pine

wood due to brown rot with the aid of FTIR spectroscopy. Holz Roh Werkstoff.50: 363–367.

Page 175: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

163 

 

 

 

Koshijima, T., Timell, T., Zinbo, M. (1965). The number-average molecular weight of native

hardwood xylans. Journal of Polymer Science C: Polymer Symposia.11:265–79.

Koshijima, T., Yaku, F., Tanaka, R. (1976). Fractionation of Bjorkman LCC from pinus

densiflora. Applied Polymer Symposia. 28: 1025-1025

Koshijima, T., Watanabe, T., Azuma, J. (1984). Existence of benzylated carbohydrate moiety

in lignin-carbohydrate complex from pine wood. Chemistry Letters. 1737-1740

Kosikova, B., Joniak, D., Skamla, J. (1972). Lignin carbohydrate bonds in beech wood.

Cellulose Chemistry and Technology. 6: 579-588

Kosikova, B., Joniak, D., Kosikova, L. (1979). On the properties of benzyl ether bonds in the

lignin-saccharidic complex isolated from spruce. Holzforschung. 33: 11-14

Koubaa, A., Perre, P., Hutcheon, R.M., Lessard, J. (2008). Complex dielectric properties of

sapwood of aspen, white birch, yellow birch and sugar maple. Drying Technology. 26: 568-

578

Kubo, S., Kadla, J.F. (2005). Hydrogen bonding in lignin. A Fourier transform infrared model

compound study. Biomacromolecules. 6: 2815-2821.

Kumar, P., Coronel, P., Simunovic, J., Truong, V.D., Sandeep, K.P. (2007). Measurement of

dielectric properties of pump able food materials under static and continuous flow conditions.

Journal of Food Science. 72: E177-E183

Lai, Y.-Z. (2001) Chemical degradation. In: Wood and Cellulose chemistry. D. N. -S. Hon

and N. Shiraishi (Eds.) Marcel Dekker, New York, 443-512

Lai, Y.-Z., Sarkanen, K.V. (1967). Kinetics of alkaline hydrolysis of glycosidic bonds in

cotton cellulose. Cellulose Chemistry and Technology. 1: 517-527.

Page 176: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

164 

 

 

 

Laine, J., Stenius, P. (1997). Effect of charge on the fiber and paper properties of bleached

industrial kraft pulps. Paperi Puu. 79: 257 -266

Lam, T.B.T., Kadoya, K., Iiyama, K. (2001). Bonding of hydroxycinnamic acids to lignin:

ferulic and p-coumaric acids are predominantly linked at the benzyl position of lignin, not the

β-position, in grass cell walls. Phytochemistry. 57: 987–992

LeBel, R. G., Goring, D. A. I., Timell, T. E. (1963). Solution properties of Birch xylan I.

Measurement of molecular weight. Journal of Polymer Science. C. 2: 9-28. 

Leschinsky, M., Patt, R., Sixta, H. (2007). Water prehydrolysis of E. Globulus with the main

emphasis on the formation of insoluble components. Pulp & Paper Conference, Helinski,7-14

Lewis, D. A., Summers, J. D., Ward, T.C., McGrath, J. E. (1992). Accelerated imidization

reactions using microwave radiation. Journal of Polymer Science Part B: Polymer Chemistry.

30: 1647-1653

Lindblad, M.S., Albertsson A.C.(2005). In Polysaccharides: Structural diversity and

functional versatility; Dumitriu, S. Ed.; Marcel Dekker: New York. pp. 491-508

Lü, J., Zhou, P. (2011). Optimization of microwave-assisted FeCl3 pretreatment conditions of

rice straw and utilization of Trichoderma viride and Bacillus pumilus for production of

reducing sugars, Bioresource Technology. 102: 6966–6971.

Liu, S. (2010). Woody biomass : Niche position as a source of sustainable renewable

chemicals and energy and kinetics of hot water extraction/hydrolysis. Advances in

Biotechnology. 28: 563-582

Liu, S.J., Amidon, T.E., Francis, R.C., Ramarao B.V., Lai, Y-Z., Scott, G.M. (2006). From

forest biomass to chemicals and energy. Industrial Biotechnology. 2: 113–20.

Page 177: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

165 

 

 

 

Lonnberg, B. (2005). In: In Polysaccharides (2nd Edition). Dumitriu S (ed.). Marcel Dekker

Inc., New York. pp.1035-1054

Lora, J.H., Wayman, M. (1978). Delignification of hardwoods by autohydrolysis and

extraction. Tappi Journal. 61(6): 47-50

Lundqvist, J., Teleman, A., Junel, L., Zacchi, G., Dalhman, O., Tjerneld, F. et al. (2002).

Isolation and characterization of galactoglucomannan from spruce (Picea abies).

Carbohydrate Polymers. 48:29-39.

Mabee, W.E., Gregg, D.J., Saddler, J.N. (2005). Assessing the emerging biorefinery sector in

Canada. Applied Biochemistry and Biotechnology. 121: 765–778.

Mao, H., Genco, J.M., Yoon, S.H., van Heiningen, A., Pendse, H. (2008). Technical

economic evaluation of a hardwood biorefinery using the “near neutral” hemicelluloses pre-

extraction process. Journal of Biobased Materials and Bioenergy. 2:177–85.

Marchessault, R.H., Liang, C.Y. (1962). The infrared spectra of crystalline polysaccharides

VIII. Xylans. Journal of Polymer Science. 59: 357-378

Martin, C., Marcet, M., Thomsen, A.B. (2008). Comparison between wet oxidation and steam

explosion as pretreatment methods for enzymatic hydrolysis of sugarcane bagasse.

Bioresources. 3(3): 670-683

Mayo, S., Evans, R., Chen, F., Lagerstrom, R. (2009). X-ray phase-contrast micro-

tomography and image analysis of wood microstructure. Journal of Physics: Conference

Series. 186(1): 012105.

Mendoza, F., Verboven, P.,Mebatsion, H.K., Kerckhofs, G., Wevers, M. B., Nicolai, B.

(2007). Three dimensional pore space quantification of apple using X-ray computed

microtomography. Planta. 226: 559-570.

Page 178: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

166 

 

 

 

Metaxas, A.C., Meridith, R.J. (1983). Industrial microwave heating. Peter Peregrinus.

London. Chapter 1-3.

Montgomery, D. C. (2001) Design and analysis of experiments, Fifth ed. Wiley, New York

Moore, A.K., Owen, N.L. (2001). Infrared spectroscopic studies of solid wood, Applied

Spectroscopy Reviews. 36: 65–86.

Mora F., Pla, F., Gandini A. (1989). The interactions between wood components and

formaldehyde-based resins. I. Monofunctional resin model compounds. Angewandte

Macromolecular Chemistry. 173:137-152

Morak, A.J., Ward Jr., K. (1960). Tappi Journal. 43: 413

Morak, A.J., Ward Jr., K. (1961). Fractional extraction and characterization of hemicelluloses.

Tappi Journal. 44: 12.

Muler, B., Beckmann, F., Huser, M., Maspero, F., Szekely, G. et al. (2002). Non-destructive

three dimensional evaluation of a polymer sponge by microtomography using synchrotron

radiation. Biomolecular Engineering. 19: 73-78.

Nelson, S.O., Datta, A.K. (2001). Dielectric properties of food materials and electric field

interactions. In: A.K. Datta and R.C. Anantheswaran (Editors), Handbook of Microwave

Technology for Food Applications. Marcel Dekker, Inc., New York

N’Diaye, S., Rigal, L. (2000). Factors influencing the alkaline extraction of poplar

hemicelluloses in a twin-screw reactor: correlation with specific mechanical energy and

residence time distribution of the liquid phase. Bioresource Technology. 75: 13-18

N’Diaye, S., Rigal, L., Larocque, P., Vidal, P.F. (1996). Extraction of hemicelluloses from

poplar, populus tremuloides, using an extruder-type twin-screw reactor: a feasibility study.

Bioresearch Technology. 57, 61-67

Page 179: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

167 

 

 

 

Nelson, R., Schuerch, C. (1957). The extraction of pentosans from woody tissues. Tappi

Journal. 40: 419-426.

Newnham, RE., Jang, SJ., Xu, M., Jone, F. (1991). Fundamental interaction mechanisms

between microwave and matter. In: Clarke D.E., Gac, FD., Sutton, WH (eds.). Ceramic

Transactions; Microwaves: Theory and applications in materials processing. Vol.21,

American Ceramic Society, Westerville, Ohio

Niemela, K., Alen R. (1999). In: Analytical methods in wood chemistry, pulping and paper

making, Sjorstrom E., Alen R. Eds. Springer-Verlag, Bermin, pp.193-232

NRCan. (2003). The State of Canada’s Forests 2002–2003, Natural Resources Canada,

Ottawa, ON

Obermeir, I. J., Seiber, V., Faulstich, M. and Scheider, D. (2012). Solubilization of

hemicellulose and lignin from wheat straw through microwave-assisted alkali treatment.

Industrial Crops and Products. 39: 98-203.

Okamoto, Y., Kawashima, M., Hatada K. (1984). Useful chiral packing materials for high

performance liquid chromatographic resolution of enantiomers: phenylcarbamates of

polysaccharides coated on silica gel. Journal of American Chemical Society. 106: 5357-5359

Ooshima, H., Aso, K., Harano, Y. (1984). Microwave treatemnt of cellulosic materials for

their enzymatic hydrolysis. Biotechology Letters. 289-94

Palm, M., Zacchi, G. (2003). Extraction of Hemicellulosic Oligosaccharides from spruce

using microwave oven or steam treatment. Biomacromolecules, 4, 617-623

Palm, M., Zacchi G., (2004). Separation of hemicellulosic oligomers from steam-treated

spruce wood using gel filtration. Separation and Purification Technology. 36: 191-201

Page 180: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

168 

 

 

 

Pandey, K.K. (1999). A study of chemical structure of soft and hardwood polymers by FTIR

spectroscopy, Journal of Applied Polymer Science. 71: 1969–1975.

Panshin, A.J., de Zeeuw, C. (1980). Textbook of Wood Technology. 4th edition. McGraw-

Hill Book Co., New York

Pare, J.R.J., Belanger, J.M.R. (1994). Microwave-Assisted Process: A New Tool for the

Analytical Laboratory. Trends in Analytical Chemistry. 13(4): 176

Pare, J. R. (1995). U.S. Patent # 5 458 897

Peng, Y., Wu, S. (2010). The structural and thermal characteristic of wheat straw

hemicellulose. Journal of Analytical and Applied Pyrolysis. 88: 134-139

Peng, F., Ren, L.L., Xu, F., Sun, R.C. (2011). Chemicals from hemicelluloses- A review. In:

Sustainable production of fuels, chemicals, and fibers from forest biomass. ACS Symposium

Series, ACS, Washington DC, pp 219-259.

Perret, J., Prasher, S.O., Kantzas, A., Langford, C. (1999), Three dimensional quantification

of macrospore networks in undisturbed soil cores. Soil Science Society of America Journal.

63: 1530-1543

Popescu, C.M., Popescu, M.C., Singurel, G., Vasile, C., Argyropoulos, D.S., Willfor, S.

(2007). Spectral characterization of eucalyptus wood. Applied Spectroscopy. 61:1168–1177.

Popescu, C.M., Singurel, G., Popescu, M.C., Vasile, C., Argyropoulos, D.S., Willfor, S.

(2009). Vibrational spectroscopy and X-ray diffraction methods to establish the differences

between hardwood and softwood. Carbohydrate Polymers. 77:851

Popescu, C.M., Popescu, M.C., Vasile, C. ( 2010). Structural changes in biodegraded wood.

Carbohydrate Polymers. 79: 362, 372.

Page 181: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

169 

 

 

 

Prat, L., Guiraud, P., Rigal, L., Gourdon, C. (2002). A one dimensional model for the

prediction of extraction yields in a two phases modified twin-screw extruder. Chemical

Engineering and Processing. 41: 743-751.

Puis, J., Saake, B. (2004). Industrial Isolated Hemicelluloses. Chapter 2, p. 24,

Hemicelluloses: Science and Technology, P. Gatenholm and M. Tenkanen, ed., ACS Symp.

Series 864, ACS, Washington, DC

Ragauskas, A.J., Nagy, M., Kim, D.H., et al. (2006). From wood to fuels, integrating

biofuels and pulp production. Industrial Biotechnology. 2(1): 55-65,

Ramos L.P., Breuil, C., Saddler J.N. (1992). Comparison of steam pretreatment of

eucalyptus, aspen, and spruce wood chips and their enzymatic hydrolysis. Applied

Biochemistry and Biotechnology. 34/35:37-48

Ramos, L.P. (2003). Chemistry involved in the steam treatment of lignocellulosic materials.

The Quim Nova. 26(6): 863-871

Roberts, J.C., EI-Karim, S.A. (1983). The behaviour of surface adsorbed xylans during the

beating of a bleached kraft pine pulp. Cellulose Chemistry and Technology. 17: 379-386

Rodrigues, J., Faix, O., Pereira, H. (1998). Determination of lignin content of Eucalyptus

globules wood using FTIR spectroscopy. Holzforschung. 52: 46–50.

Roos, A.A., Persson, T., Krawczyk, H., Zacchi, G., Stålbrand, H. (2009). Extraction of water-

soluble hemicelluloses from barley husks. Bioresource Technology. 100: 763-769

Saddler, J.N., Brownell, H.H., Clermont, L.P., Levitin N. (1982). Enzymatic hydrolysis of

cellulose and various pretreated wood fractions. Biotechnology and Bioengineering. 24:

1389-1402

Page 182: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

170 

 

 

 

Saddler, J.N., Ramos L.P., Breuil, C. (1993). In: Bioconversion of forest and agricultural plant

wastes; Saddler, J.N. ed.; C. A. B. International: London. pp.73-92

Salvo, L., Clotenes, P., Maire, E., Zabler, S., Blandin, J.J. et al. (2003). X-ray

microtomography an attractive characterization technique in material science. Nuclear

Instruments Methods Physics Research. B. 200: 273-286.

Scheller, H.V., Ulvskov, P. (2010). Hemicelluloses. Annual Review, Plant Biololgy. 61:263-

89

Schwald,W., Breuil C., Brownell, H.H., Chan M., Saddler, J.N. (1989). Assessment of

pretreatment conditions to obtain fast complete hydrolysis on high substrate concentrations.

Applied Biochemistry and Biotechnology. 20/21: 29

Sedlmeyer, F. B. (2011). Xylan as by-product of biorefineries: characteristics and potential

use for food applications. Food Hydrocolloids. 25: 1891–1898.

Segal, L., Creely, J.J., Martin Jr, A.E., Conrad, C.M. (1959). An empirical method for

estimating the degree of crystallinity of native cellulose using the X-ray diffractometer.

Textile Research Journal. 29:786-794

Sjostrom, E. (1993). Wood Chemistry Fundamentals and Applications. Second ed. Academy

Press Inc., San Diego

Shatalov, A. A., Evutuguin, D.V., Neto C. P. (1999) . 2-O- -D-Galactopyranosyl-4-O-

methyl- -D-glucurono-D-xylan from Eucalyptus globulus Labill. Carbohydrate Research.

320: 93-99

Shi, J., Pu, Y., Yang, B., Ragauskas, A., Wyman, C.E. (2011). Comparison of microwaves to

fluidized sand baths for heating tubular reactors for hydrothermal and dilute acid batch

pretreatment of corn stover. Bioresource Technol. 102: 5952–5961.

Page 183: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

171 

 

 

 

Sluiter, A., Hames, B., Ruiz, R., Scarlata, C., Sluiter, J., Templeton, D., Crocker, D. (2008).

Determination of structural carbohydrates and lignin in biomass, Laboratory Analytical

Procedure. National Renewable Energy laboratory technical report, NREL/TP-510-42618

State-ease, version 8.0.7.1. User's guide, Design Expert Software, Inc., Minneapolis, USA

Steppe, K., Cnudde, V., Girard, C., Lemeur, R., Cnudde, J.P. et al. (2004), Use of X-ray

computed microtomography for non-invasive determination for wood anatomical

characteristics. Journal of Structural Biology. 148: 11-21.

Stipanovic, A., Amidon, T.E., Scott, G.M., Barber, V., Blowers, M.K. (2006). Hemicellulose

from biodelignified Wood: A Feedstock for Renewable Materials and Chemicals. Feedstocks

for the Future, ACS Symposium Series 921, Joseph J. Bozell Ed., Oxford University Press.

Strauss, C.R., Varma, R.S. (2006). Microwaves in green and sustainable chemistry. Topics in

Current Chemistry. 266: 199–232.

Stuppy, W.H., Maisano, J.A., Colbert, M.W., Rudall, P.J., Rowe, T.B. (2003). Three

dimensional analysis of plant structure using high resolution X-ray computed tomography.

Trends in Plant Science. 8: 2-6.

Sun, J.X., Sun, R.C., Sun, X.F., Su, Y.Q. (2004). Fractional and physico-chemical

characterization of hemicelluloses from ultrasonic irradiated sugarcane bagasse. Carbohydrate

Research. 339: 291-300

Sun, R.C., Lawther, J.M., Banks, W.B. (1995). Influence of alkali pre-treatment on wheat

straw cell wall components. Industrial Crops and Products. 4: 127–145

Sun, R.C., Fanga, J.M., Motta, L., Boltona, J. (1999). Extraction and Characterization of

Hemicelluloses and Cellulose from Oil Palm Trunk and Empty Fruit Bunch Fibers. Journal of

Wood Chemistry and Technology. 19 (182): 167-185

Page 184: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

172 

 

 

 

Sun, R.C., Sun, X.F. (2002). Fractional and structural characterization of hemicelluloses

isolated by alkali and alkaline peroxide from barley straw. Carbohydrate Polymers. 49: 415–

423.

Sun Y., Cheng, J. (2002). Hydrolysis of lignocellulosic materials for ethanol production: a

review. Bioresource Technology. 83: 1-11

Sun, R.C. (2008). Detoxification of biomass for bioethanol. Bioresources. 4(2): 452-455.

Talbott, L. D., Ray, P. M. (1992). Molecular size and separability features of pea cell wall

polysaccharides. Plant Physiology. 92: 357–368

Teleman, A., Harjunpa, V., Tenkanen, M., et al. (1995). Characterization of 4-deoxy-β-L-

threo-hex-4-enopyranosyluronic acid attached to xylan in pine kraft pulp and pulping liquor

by 1H and 13C NMR spectroscopy. Carbohydrate Research. 272: 55-71

Teleman, A., Lundqvist, J., Tjerneld, F., Stalbrand, H., Dahlman, O. (2000). Characterization

of acetylated 4-O-methylglucuronoxylan isolated from aspen employing 1H and 13C NMR

spectroscopy. Carbohydrate Research. 329: 807-815.

Teleman, A., Tenkanen, M., Jacobs, A., Dahlman, O. (2002). Characterization of O-acetyl-(4-

O-methylglucurono)xylan isolated from birch and beech. Carbohydrate Research. 337: 373-7

Thompson, J.E., Fry, S.C. (2000). Evidence for covalent linkage between xyloglucan and

acidic pectins in suspension-cultured rose cells. Planta. 211: 275-286

Thostenson, E.T., Chou, T.W. (1999). Microwave processing: Fundamentals and

applications. Composites Part A. 30: 1055-1071

Thovert, J., Salles, J., Adler, P. (1993). Computerized characterization of the geometry of real

porous media: their discretization, analysis and interpretation. Journal of Microscopy.170:

65-79.

Page 185: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

173 

 

 

 

Timell, T. E., Syracuse, N.Y. (1967) Recent progress in the chemistry of wood

hemicelluloses. Wood Science and Technology. 1:45-40

Timell, T. E. (1964). Wood hemicelluloses. Part 1. Advances in Carbohydrate Chemistry. 19:

247-302

Thorpe, B. (2005). Biorefinery offers industry leaders business model for major change. Pulp

and Paper Canada. 79(11): 35-39.

Towers, M., Browne, T., Kerekes, R., Paris, J., Tran, H. (2007). Biorefinery opportunities for

the Canadian pulp and paper industry. Pulp Paper Canada. 108 (6): 26–29.

Tunc, M.S., Adriaan, R.P., van Heiningen, R.P. (2008). Hemicellulose extraction of mixed

southern hardwood with water at 150 °C: Effect of time. Industrial Engineering and

Chemistry Research. 47: 7031-7037

Tunc, M.S., van Heiningen, A. (2009). Hydrothermal dissolution of mixed southern

hardwoods: Effect of P factor. Nordic Pulp and Paper Science Journal. 24(1): 42-47

Van de Velde, K., Kiekens, P. (2002). Biopolymers: overview of several properties and

consequences on their applications. Polymer Testing. 21: 433-42

van Hazendonk, J.M., Reinerink, J.M., de Waard, P., van Dam, J. E. G. (1996). Structural

analysis of acetylated hemicellulose polysaccharides from fiber flax (Linum usitatissimum

L.). Carbohydrate Research. 291: 141-154

van Heiningen, A. (2006). Converting a kraft pulp mill into an integrated forest biorefinery.

Pulp and Paper Magazine Canada. 10(6): 38-46.

van Wulsum, G.P., Allen, S.G., Spencer, M. J., et al. (1996). Conversion of lignocellulosics

pretreated with liquid hot water to ethanol. Applied Biochemistry and Biotechnology. 57-58:

157-170

Page 186: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

174 

 

 

 

Vazquez, M. J., Alonso, J. L., Dominguez, H., Parajo, J. C. (2000). Xylooligosaccharides:

manufacture and applications. Trends in Food Science and Technology. 11: 387–393.

Venkatesh, M.S., Raghavan, G.S.V. (2004). An Overview of Microwave Processing and

Dielectric Properties of Agri-food Materials. Biosystems Engineering. 88 (1): 2004

Vogel, J. (2008). Current Opinions in Plant Biology. 11:30-307

Vuorinen, T., Alén, R. (1998). Carbohydrates. In: Sjöström, E., Alén, R. (Eds.), Analytical

Methods in Wood Chemistry, Pulping, and Papermaking. Springer- Verlag, New York. pp.

37–71.

Ward Jr, K., Muray, M.L. (1959). Tappi Journal. 42, 17

Ward, K., Morak A. J. Fractional extraction and properties of hemicelluloses.

iupac.org/publications/pac/5/1/0077/pdf/ 77-88

Watanabe, T., Koshijima, T. (1988). Evidence for an ester linkage between lignin and

carbohydrate complexes by DDQ-oxidation. Agricultural and Biological Chemistry. 52(11):

2953-2955

Westbye P., Kohnke, T., Gatenholm, P. (2008). Fractionation and characterization of xylan

rich extracts from birch. Holzforschung. 62: 31-37

Whistler, R. L. (1950). Xylan. Advances in Carbohydrate Chemistry. 5: 269-290

Wildenschild, D., Hopsman, J.W., Vaz, C.M.P., Rivers, M.L., Rikard, D. et al. (2002) Using

X-ray computed tomography in hydrology: systems, resolutions, and limitations. Journal of

Hydrology. 267: 285-297.

Wilkie, K.C.B. (1979). The hemicelluloses of grasses and cereals. Advances in Carbohydrate

Chemistry and Biochemistry. 36: 215-264.

Page 187: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

175 

 

 

 

Yoon S.H., van Heiningen, A. (2008). Kraft pulping and paper making properties of hot water

pre-extracted Loblolly Pine in an integrated forest product biorefinery. Tappi Journal.

(July)22- 27

Yoon S-H., van Heiningen A. (2010). Green liquor extraction of hemicelluloses from

southern pine in an integrated forest biorefinery. Journal of Industrial and Engineering

Chemistry. 16: 74-80

Yoshida T., Tsubaki, S., Teramoto, Y., Azuma J. (2010). optimization of microwave assisted

extraction of carbohydrates from industrial wastes of corn starch production using response

surface methodology. Bioresource Technology. 101: 7820–7826

Yuka, F., Yamada, Y., Koshijima, T. (1976). Lignin-carbohydrate complex. Part IV. Lignin as

side chain of the carbohydrate in Bjorkman LCC. Holzforschung 30: 148-156

Zhang, X., Hayward, D.O. (2006) Applications of microwave dielectric heating in

environment related heterogeneous gas phase catalytic systems. Inorganica Chimia Acta. 359:

3421-3433

Zhang, Y.H.P., Ding, S.Y., Mielenz, J.R., Cui J.B. et al. (2007). Fractionating recalcitrant

lignocellulose at modest reaction conditions. Biotechnology and Bioengineering. 97: 214-

223

Zhu, S.D., Wu, Y.X., Yu, Z.N., Liao J.T., Zhang, Y. (2005). Pretreatment by

microwave/alkali of rice straw and its enzymatic hydrolysis. Process Biochemistry. 40 (9):

3082-3086

Zobel, B., McElvee, R. (1966). Variation of cellulose in loblolly pine. Tappi Journal. 49 (9):

383

Page 188: By Fathimathul Suhara Panthapulakkal · Figure 1.3 Polysaccharide network in lignocellulosic matrix Figure 1.4 Suggested lignin-carbohydrate bonds in lignocellulosic matrix Figure

176 

 

 

 

Zykwinska, A. W., Ralet, M. C., Garnier, C. D., Thibault, J. F. (2005). Evidence for in vitro

binding of pectin side chains to cellulose. Plant Physiology. 139: 397–407