transforming lignocelluloses to sugars and liquid fuels by li shuai a dissertation submitted in

223
Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy (Biological Systems Engineering) at the UNIVERSITY OF WISCONSIN-MADISON 2012 Date of final oral examination: 06/28/2012 The dissertation is approved by the following members of the Final Oral Committee: Xuejun Pan, Associate professor, Biological Systems Engineering John Ralph, Professor, Biological Systems Engineering and Biochemistry Sundaram GunasekaranProfessor, Biological Systems Engineering Troy Runge, Assistant Professor, Biological Systems Engineering John Grabber, Research Agronomist, US Dairy Forage Research Center

Upload: others

Post on 11-Sep-2021

13 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

Transforming Lignocelluloses to Sugars and Liquid Fuels

By

LI SHUAI

A dissertation submitted in partial fulfillment of

the requirements for the degree of

Doctor of Philosophy

(Biological Systems Engineering)

at the

UNIVERSITY OF WISCONSIN-MADISON

2012

Date of final oral examination: 06/28/2012

The dissertation is approved by the following members of the Final Oral Committee:

Xuejun Pan, Associate professor, Biological Systems Engineering

John Ralph, Professor, Biological Systems Engineering and Biochemistry

Sundaram Gunasekaran,Professor, Biological Systems Engineering

Troy Runge, Assistant Professor, Biological Systems Engineering

John Grabber, Research Agronomist, US Dairy Forage Research Center

Page 2: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

ii

Abstract

Extensive research has been done on the development of biofuel from low-cost and

abundant lignocelluloses. Unfortunately, cost-effectively producing sugars and sugar

derivatives still remains a barrier to developing a biorefining industry. In order to

overcome this barrier, a few innovative processes were developed for converting

lignocelluloses into sugars and liquid fuels, and are presented in this thesis.

First, a sulfite pretreatment (SPORL–Sulfite Pretreatment to Overcome

Recalcitrance of Lignocelluloses) developed by our group was compared with diluted acid

pretreatment (DA) to investigate the efficacy of this new pretreatment method on

enzymatic saccharification of spruce. Results show that addition of sulfite along with

sulfuric acid could remove more lignin, retain more carbohydrates in the substrate and

reduced formation of inhibitors in spent liquor than dilute acid pretreatment due to the

reaction of sulfite with lignin and the buffer effect of sulfite. Cellulose in SPORL-

pretreated spruce was completely digested by enzymes as compared to the cellulose

conversion of 60% in DA-pretreated spruce. Additionally, the formation of hydrophilic

sulfonic groups on lignin surface was believed to decrease non-productive adsorption of

enzymes on lignin, facilitating the enzymatic hydrolysis of cellulose.

Second, considering the acid corrosion and expense of enzymes involved in

pretreatment and enzymatic saccharification, preliminary work was conducted to

synthesize reusable cellulase-mimetic solid acid with both cellulose binding and

hydrolyzing domains for cellulose hydrolysis. The binding domain (-NH2, -OH or -Cl) on

the synthesized solid acids facilitated the association of substrates onto the catalyst surface,

which increased collision chance of substrate with acid sites (-SO3H) and therefore

Page 3: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

iii

accelerated the cellulose hydrolysis rate. Cellulose hydrolysis reactions catalyzed by the

synthesized solid acids showed much lower apparent activation energies than the ones

catalyzed by traditional liquid acids and general solid acids without binding domains.

Third, to avoid energy-intensive pretreatment and expensive cellulases involved in

traditional enzymatic saccharification of lignocelluloses, a one-step process to produce

concentrated sugar solution from lignocelluloses was developed. This process directly

converted cellulose and hemicellulose in lignocelluloses into sugars at moderate

temperature (100-160 °C) without pretreatment and enzymatic hydrolysis. Concentrated

LiCl, LiBr and CaBr2 solutions were found to have good cellulose dissolution abilities and

to be able to dissolve/hydrolyze cellulose from lignocelluloses at moderate temperatures.

Addition of small amount of acid into the concentrated salt solution accelerated the

hydrolysis of cellulose and hemicellulose. The batch-feeding of biomass allowed a high

final sugar concentration. After the saccharification, insoluble lignin was separated from

sugars and salt solution by filtration or centrifugation. Sugars and salt were separated

through a combination of organic solvent extraction of the salt and ion-exchange

chromatography. Solvent extraction separated approximately 95% of the salt from the

sugars, and the residual salt was removed by ion-exchange resins. In order to obtain a

purified sugar syrup of high concentration, air instead of water was used to push the sugar

stream through the ion-exchange columns. Ultimately, a sugar solution with a

concentration higher than 50% was recovered.

Fourth, a novel one-pot process of directly converting lignocelluloses into

hydrocarbon precursors without pretreatment and enzymatic saccharification was

developed. The reaction was conducted in a LiBr/acetone reaction system with small

Page 4: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

iv

amount of acid and water. Because of the deficiency of water in the LiBr/acetone system,

unsolvated Li+ and Br

- were able to disrupt the hydrogen bonding in cellulose crystals,

facilitating the hydrolysis of cellulose and hemicelluloses to monosugars. The Br- also

catalyzed the dehydration of the sugars into HMF (or furfural), which immediately reacted

with acetone to form furan-based hydrocarbon precursors with 5-21 carbons in high yield

and with high selectivity. Use of acetone as solvent prevented the self-condensation of

HMF (or furfural) into byproduct humins, thereby improving the selectivity of the sugars

to the precursors. Meanwhile, lignin was extensively depolymerized and dissolved in

acetone during the process. Because of very low molecular weight, the lignin could be

hydrodeoxygenated into hydrocarbon fuels (or fuel additives) without further

depolymerization, separately or jointly with the furan-based hydrocarbon precursors

derived from cellulose and hemicellulose.

Page 5: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

v

Acknowledgements

The work shown here was only possible under the support, guidance and

collaboration of many important people in the past five years. Firstly, I would like to thank

my advisor, Professor Xuejun Pan for his meticulous guidance on my research and his

cares and support for my life in the States. I feel very fortunate and grateful to work with

such an amiable, thoughtful and inspirational advisor.

I would like to thank our group members for their help, collaboration and

friendships. A special thank to Qiang who I worked with over the last five years. We are

the first two persons working in this group, and experienced and shared a lot together. I

would like to thank Dongsheng Zhang, Daeun Kim, Syrym Abylgaziyev, Sasikumar

Elumalai, Lis Nimani, and Chaoqun Mei for their collaboration and discussions.

In our department, I would like to thank Professor John Ralph for letting me access

NMR and GC-MS facilities and Dr. Fachuang Lu for NMR analysis. A thank-you to

Professor Sundaram Gunasekaran for letting me access his lab for the use of FT-IR and

particle size analyzer. A special thanks to Debby Sumwalt for her help and encouragement

in my life here.

My family deserve a huge thank-you. My mother and brother gave me strong

encouragement and support in my life. There is no way I would be where I am today

without them. Particularly, I want to thank my brother, Ke Shuai, who has taught me a lot

in all aspects of my life and who I always look up to as my role model. A special thank to

my fiancée, Ying Li, for her love and support every day. You are a beautiful girl and I am

blessed to have you in my life.

Page 6: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

vi

Table of Contents

Abstract

Acknowledgements

List of Figures

List of Tables

Chapter 1: Introduction to Biofuels .............................................................................................. 1

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

1.2 Chemistry of Biomass ................................................................................................. 4

1.3 Technical issues in biomass conversion ................................................................... 21

1.4 Project description ................................................................................................... 30

Chapter 2: Comparative Study of SPORL and Dilute Acid Pretreatments of Spruce for

Enzymatic Saccharification .................................................................................. 32

2.1 Introduction .............................................................................................................. 32

2.2 Experimental ............................................................................................................ 38

2.2.1 Materials ..................................................................................................... 38

2.2.2 Pretreatments .............................................................................................. 38

2.2.3 Enzymatic Hydrolysis .................................................................................. 39

2.2.4 Analytical Methods ...................................................................................... 40

2.2.5 Whole Cell-Wall NMR of substrates ........................................................... 41

2.2.6 Degree of Polymerization of Cellulose ....................................................... 42

2.2.7 Fermentability of SPORL and DA Pretreatment Liquors ........................... 42

2.3 Results and Discussion ............................................................................................. 43

2.3.1 Changes in Cell-Wall Components after pretreatments .............................. 43

2.3.2 Mass balance of sugars after pretreatments ............................................... 52

2.3.3 Enzymatic digestibility of pretreated spruce ............................................... 54

2.3.4 Fermentability of spent pretreatment liquors .............................................. 58

2.4 Conclusion and recommendations ........................................................................... 61

Page 7: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

vii

Chapter 3: Synthesis of Cellulase Mimetic Solid Acid for Cellulose Hydrolysis ..................... 63

3.1 Introduction .............................................................................................................. 63

3.2 Experimental ............................................................................................................ 69

3.2.1 Chemicals and materials ............................................................................. 69

3.2.2 Synthesis of cellulase mimetic solid acid .................................................... 69

3.2.3 Hydrolysis of biomass with solid acid ......................................................... 71

3.2.4 Adsorption of glucose and cellobiose on CP-SO3H .................................... 72

3.2.5 Determination of glucose ............................................................................ 72

3.2.6 FT-IR spectra of prepared resins ................................................................ 72

3.3 Results and discussion ............................................................................................. 73

3.3.1 Screening of binding groups ....................................................................... 73

3.3.2 Mechanism study ......................................................................................... 78

3.3.3 Hydrolysis of cellulose with CP-SO3H ........................................................ 82

3.4 Conclusion and recommendations ........................................................................... 90

Chapter 4: Saccharification of Lignocellulose in Concentrated Salt Solution ........................ 91

4.1 Introduction .............................................................................................................. 91

4.2 Experimental ............................................................................................................ 97

4.2.1 Materials and Chemicals ............................................................................ 97

4.2.2 Liquefaction of lignocellulose in concentrated LiBr solution ..................... 98

4.2.3 Hydrolysis of lignocellulose in acidic concentrated LiBr solution ............. 98

4.2.4 Extraction of LiBr by organic solvents........................................................ 99

4.2.5 Removal of salt by ion-exchange chromatography ................................... 101

4.2.6 Quantification of sugars and sugar derivatives ........................................ 102

4.2.7 Determination of LiBr amount .................................................................. 103

4.3 Result and discussion ............................................................................................. 103

4.3.1 Description of whole process .................................................................... 103

4.3.2 Liquefaction of lignocellulose in concentrated LiBr solution ................... 105

Page 8: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

viii

4.3.3 Dissolution mechanism of cellulose in concentrated LiBr solution .......... 109

4.3.4 Hydrolysis of lignocellulose in acidic concentrated LiBr solution ........... 113

4.3.5 Hydrolysis of lignocellulose through batch feeding .................................. 117

4.3.6 Saccharification of lignocellulose in concentrated solution of different salts

................................................................................................................... 119

4.3.7 Hydrolysis of lignocellulose in concentrated LiBr solution with different

acids ......................................................................................................... 124

4.3.8 Separation of LiBr and sugars by different methods ................................. 125

4.3.9 Removal of residual LiBr from sugar stream ............................................ 138

4.4 Conclusion and recommendations ......................................................................... 147

Chapter 5: Conversion of Lignocellulose into Hydrocarbons ................................................. 150

5.1 Introduction ............................................................................................................ 150

5.2 Experimental .......................................................................................................... 156

5.2.1 Chemicals and materials ........................................................................... 156

5.2.2 Production of hydrocarbon precursors from biomass .............................. 156

5.2.3 Determination of residual LiBr ................................................................. 157

5.2.4 Determination of sugars and sugar derivatives ........................................ 157

5.2.5 Qualitative analysis of hydrocarbon precursors using GC-MS ................ 157

5.2.6 Quantitative analysis of hydrocarbon precursors using ESI-MS .............. 158

5.2.7 Estimation of lignin molecular weight ...................................................... 160

5.2.8 Characterization of LiBr/acetone solvent systems .................................... 160

5.3 Results and Discussion ........................................................................................... 160

5.3.1 Description of the HDA process ................................................................ 160

5.3.2 LiBr/water and LiBr/acetone systems ....................................................... 163

5.3.3 One-step conversion of biomass into hydrocarbon precursors ................. 166

5.3.4 Effect of temperature and time on conversion of spruce powder .............. 171

5.3.5 Effect of different salts on conversion of spruce powder .......................... 172

5.3.6 Effect of different acids on conversion of spruce powder ......................... 173

Page 9: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

ix

5.3.7 Identification of products .......................................................................... 174

5.3.8 Quantification of conversion by ESI-MS ................................................... 181

5.3.9 Carbon number distributions of hydrocarbon precursors ........................ 185

5.3.10 Decomposition of lignin during HDA process .......................................... 187

5.3.11 Reaction mechanism of glucose to HMF ................................................... 188

5.3.12 Recycling/recovery of solvents and LiBr ................................................... 192

5.3.13 Hydrodeoxygenation of hydrocarbon precursor and lignin into

hydrocarbons ............................................................................................. 193

5.4 Conclusion and recommendations ......................................................................... 195

Reference ................................................................................................................................ 196

Page 10: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

x

List of Figures

Figure 1.1 A structure model for plant cell wall 6

Figure 1.2 Chemical structures of (a) cellulose and (b) starch 9

Figure 1.3 Hemicelluloses characterized by a β-(1→4)-linked backbone with an equatorial

configuration at C1 and C4 12

Figure 1.4 Schematic illustration of the types of hemicelluloses found in plant cell walls 14

Figure 1.5 Chemical structure of lignin monomers 15

Figure 1.6 Formation of resonance-stabilized phenoxyl radical catalyzed by enzyme 16

Figure 1.7 Structures of main linkages in lignin 18

Figure 1.8 A proposed model structure of lignin 19

Figure 1.9 Flowchart of biorefining 21

Figure 2.1 Flowchart of experiment and analysis 37

Figure 2.2 HSQC NMR spectra of untreated and pretreated spruce cell walls 51

Figure 2.3 Mass balance of saccharides during the DA and SPORL pretreatments 54

Figure 2.4 Comparison of time-dependent enzymatic hydrolysability of

SPORL and DA pretreated spruce at different levels of enzyme loading 55

Figure 2.5 Inhibitors formation from cellulose and hemicellulose during pretreatment 58

Figure 2.6 Fermentability of SPORL and DA pretreatment spent liquors by in

vitro ruminal fermentation assay 60

Figure 3.1 A proposed model of cellulase-mimetic solid acid and cellulose interactions 68

Figure 3.2 Synthesis of CP-SO3H 70

Page 11: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

xi

Figure 3.3 A model for cellulose hydrolysis on solid acid (CP-SO3H) surface 71

Figure 3.4 FT-IR spectra of (a) CP resin and (b) CP-SO3H resin 73

Figure 3.5 FT-IR spectra of solid acids with different binding domains 75

Figure 3.6 Cellobiose hydrolysis catalyzed by four types of solid acids as a function of time

76

Figure 3.7 Time course of adsorption curve of glucose and cellobiose onto resins in

aqueous solution 80

Figure 3.8 Comparison of cellobiose hydrolysis catalyzed by (a) CP-SO3H and (b) PS-

SO3H 81

Figure 3.9 Hydrolysis of cellobiose catalyzed by CP-SO3H and sulfuric acid 83

Figure 3.10 Cellobiose hydrolysis catalyzed by recycled CP-SO3H resin 85

Figure 3.11 Arrhenius plot for cellulose hydrolysis catalyzed by CP-SO3H 87

Figure 4.1 Process flow chart of biomass saccharification in concentrated salt solution 104

Figure 4.2 Hydrolysis way of biomass in concentrated salt solution 105

Figure 4.3 Hydrolysis of spruce powder in LiBr solution (no acid) as a function

of LiBr concentration at different temperatures 107

Figure 4.4 Models for cellulose dissolution in salt solutions of varied concentrations 111

Figure 4.5 The picture of the separated hydrolysate 119

Figure 4.6 The proposed structures of LiCl hydrates 123

Figure 4.7 Mechanism of extraction of glucose with boronic acid 126

Page 12: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

xii

Figure 4.8 Schematic diagram for ion-exclusion chromatography for separating

sugar and salt 128

Figure 4.9 Separation of LiBr and sugar solution using ion-exclusion chromatography 129

Figure 4.10 Separation of residual LiBr from sugar solution using ion-exclusion

chromatography 130

Figure 4.11 Flowchart for separating LiBr and sugars by solvent extraction 132

Figure 4.12 The picture of extraction of LiBr from hydrolysate with butanol-hexane 135

Figure 4.13 Formed sugar syrup after butanol extraction of hydrolysate 137

Figure 4.14 Sugars precipitated from solvent 142

Figure 4.15 The pictures of columns packed with cation and anion exchange resins 145

Figure 4.16 The picture of purified concentrated sugar solution 146

Figure 5.1 Reaction pathway of biomass to hydrcarbon precursors in

HDA process 160

Figure 5.2 Process flow chart of converting lignocelluloses into hydrocarbons 162

Figure 5.3 Interaction of LiBr with different solvents 164

Figure 5.4 FT-IR spectra of LiBr/acetone solution 165

Figure 5.5 The Hydrocarbon precursors from HDA process in acetone 171

Figure 5.6 Proposed self-condensation mechanism of acetone 175

Figure 5.7 Proposed condensation mechanism of furfural with acetone 175

Figure 5.8 GC-MS spectra of hydrocarbon precursors in CH2Cl2 178

Page 13: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

xiii

Figure 5.9 ESI-MS spectra of hydrocarbon precursors in CH2Cl2 180

Figure 5.10 Quantificaiton of the hydrocarbon precursors from lignocelluloses using 2D-

NMR method 183

Figure 5.11 Mechanism of quantifying hydrcarbon precursors by oxidation method 184

Figure 5.12 Carbon number distributions of hydrocarbon precursors from

different feedstocks 186

Figure 5.13 Gel permeation chromatograph of HDA lignin from spruce 188

Figure 5.14 Dehydration of glucose to HMF derivative at different LiBr concentrations 189

Figure 5.15 Dehydration of fructose to HMF derivative at different LiBr concentrations 189

Figure 5.16 The proposed mechanism of glucose to HMF in acetone/LiBr system 191

Figure 5.17 Images of products before and after hydrodeoxygenation 194

Page 14: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

xiv

List of Tables

Table 1.1 Composition of common lignocelluloses on dry basis 7

Table 1.2 Content of main linkages in lignin 18

Table 2.1 Chemical analyses of original spruce, pretreated spruce substrates

and spent pretreatment liquors 46

Table 2.2 Concentrations of major fermentation inhibitors in pretreatment

spent liquors 48

Table 2.3 Viscosity of cellulose solutions from untreated- and pretreated-spruces 49

Table 3.1 The apparent activation energies of cellobiose hydrolysis catalyzed

by four types of solid acids 77

Table 3.2 Hydrolysis of starch and Avicel cellulose catalyzed by CP-SO3H

and sulfuric acid 83

Table 3.3 Apparent activation energies of cellulose hydrolysis catalyzed by different

catalysts 89

Table 4.1 Composition analyses of different feedstocks 100

Table 4.2 LiBr-hydrolysis of various types of feedstock 106

Table 4.3 Monosaccharides in the hydrolysates from LiBr hydrolysis of different

feedstocks without acid 108

Table 4.4 Monosaccharides in hydrolysate from LiBr hydrolysis of different

feedstocks after autoclaving at 120 °C for 1h 109

Table 4.5. Hydrolysis of spruce powder under various conditions 116

Table 4.6 Hydrolysis of spruce powder in batch-feed mode 118

Page 15: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

xv

Table 4.7 Effect of different salts on hydrlysis of biomass 121

Table 4.8 Effect of different acids on hydrlysis of biomass 125

Table 4.9 Separation of glucose and LiBr by extraction with a mixture of

butanol and hexane 133

Table 4.10 Separation efficiency of LiBr and glucose by solvent extraction 133

Table 4.11 Effects of LiBr and water on the precipitation of glucose by ethanol 141

Table 5.1 Conversion of lignocelluloses or sugars into hydrocarbon precursors in

acetone/LiBr system 168

Table 5.2 Effect of temperature on conversion of spruce powder 172

Table 5.3 Effect of time on conversion of spruce powder 172

Table 5.4 Effect of different halogen salts on conversion of spruce powder 173

Table 5.5 Effect of different acids on conversion of spruce powder 174

Table 5.6 Proposed structure of products and corresponding molecular weight 176

Table 5.7 Quantitation of hydrocarbon precursors derived from carbohydrates 184

Table 5.8 Proton affinity of different functional groups in hydrocarbon precursors 186

Page 16: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

1

Chapter 1: Introduction to Biofuels

1.1 Introduction

A biofuel is a fuel that is derived from biological materials, and can include a

series of different fuels forms, such as solid carbon, liquid fuels, and gases. Biofuels are

typically deemed to be sustainable because the carbon in biofuel is in a closed carbon

cycle where carbon dioxide generated from combustion of biofuel could be fixed and

converted into biomass again through photosynthesis.

There has been a growing interest in liquid biofuels worldwide over the past few

years, mostly due to the rising price of fossil fuels, concerns over the rising CO2 and

other greenhouse gas emissions, climate change, and the expected depletion of world oil

reserves. It is believed that producing transportation fuels from renewable biomass

resources can reduce the dependence on traditional fossil fuel, relieve the energy crisis,

create new job opportunities, improve local economies, and reduce greenhouse gas

emissions. Recent legislation, government investment, and technology advances have

greatly promoted biofuels production and development. For example, the United States

produced 9 billion gallons corn ethanol and 500 million gallons biodiesel in 2009.

Page 17: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

2

Technologies for new generation cellulosic ethanol are under development and being

scaled-up in demonstration projects. While current research effort and investment

primarily focus on ethanol (from both starch and cellulose), biobutanol, and biodiesel,

the next generation liquid fuels, biohydrocarbons (gasoline, diesel and jet fuel from

biomass) also show promise because of their advantages over last generation liquid

biofuels (Demirbas, 2005; Demirbas, 2009; Limayem and Ricke, 2012).

Bioethanol including corn ethanol and cellulosic ethanol, is produced by the

fermentation of sugars. Commercialized bioethanol is currently produced from corn

kernels in the United States and sugarcane in Brazil. Corn kernels as grain have limited

yield and the massive consumption of corn kernels to produce bioethanol might

endanger the food security. Therefore, researchers now are focusing on obtaining sugars

from cheap and abundant lignocelluloses, such as wood, agricultural residues, and

dedicated bioenergy crops. A typical process for cellulosic ethanol production consists of

three steps: feedstock pretreatment to enhance the accessibility of cellulose to cellulases,

enzymatic saccharification of the cellulose to glucose, and fermentation of glucose to

ethanol. One of the barriers to the commercialization of cellulosic ethanol is the lack of

economical and effective technologies for the feedstock pretreatment. The significance

Page 18: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

3

and importance of the pretreatment step cannot be overemphasized, as the effectiveness

of pretreatment affects the upstream selection of biomass, the yield of fermentable sugars,

and the chemical and morphological characteristics of the pretreated substrate that in

turn govern downstream hydrolysis. An ideal pretreatment method should be both

economical (in terms of capital and operating costs) and effective for a variety of

lignocelluloses. Specifically, it should require minimal feedstock preparation and

preprocessing (such as size reduction) prior to pretreatment, maximally recover all

lignocellulosic components in usable forms with minimal formation of fermentation

inhibitors, and produce a readily digestible cellulosic substrate that can be easily

hydrolyzed with a low loading of enzymes. In the last several decades, research and

development efforts have made significant progress in pretreatment technologies for

lignocellulosic feedstocks (Chandra et al., 2007; Hendriks and Zeeman, 2009; Lynd et al.,

2002; Mosier et al., 2005; Wyman et al., 2005). Many pretreatment technologies, such as

lime, dilute acid, hot water, ammonia, steam explosion, SPORL (Sulfite Pretreatment to

Overcome Recalcitrance of Lignocelluloses), ionic liquid and organosolv pretreatments

(Chandra et al., 2007; Gupta and Sehgal, 1979; Kim et al., 2009; Monavari et al., 2009;

Nguyen et al., 2000; Pan et al., 2005; Sendich et al., 2008; Sierra et al., 2009; Wingren et

Page 19: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

4

al., 2008), have achieved varying levels of success.

Compared to bioethanol, next-generation liquid biofuels-hydrocarbons (or bio-

hydrocarbons) have advantages of high energy density and immiscibility with water. In

addition, limited water use for processing, short production cycle, and the elimination of

energy-intensive distillation will possibly lead to a low production cost of

biohydrocarbons. The biohydrocarbons can be produced through different processes

using a variety of technologies, including pyrolysis followed by upgrading of bio-oil,

gasification followed by Fischer-Tropsch synthesis, liquid-phase reforming of sugars or

sugar derivatives, decomposition and hydrodeoxygenation of lignin, and oligomerization

of alkenes derived from biomass. These methods typically involve multiple steps where

biomass is firstly converted into purer and simpler chemical states, such as sugars,

syngas, HMF, or levulinic acid, and then these chemicals are subsequently processed

into fuels catalytically.

1.2 Chemistry of Biomass

Lignocellulose is a scientific term for plant roots, stems and leaves, and is the

most abundant source of organic material on earth. Lignocelluloses include, for example,

Page 20: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

5

agricultural residues such as corn stover and wheat stalks, energy crops such as poplar

trees and switchgrass, and municipal solid waste.

The cell walls of lignocelluloses are composed of three major components:

cellulose, hemicelluloses, and lignin. The distributions of three components in cell walls

are schematically depicted in Figure 1.1. A cell wall is generally divided into several

layers, including the primary wall (P), the thin outer layer of the secondary wall (S1), the

substantial middle layer (S2), and the very thin inner layer or tertiary wall (S3).

Cellulose chains and other cell wall constituents are aggregated into bundles

called microfibrils within each layer of the secondary wall (S). The microfibrillar groups

are in helixes alternately crossed in the S1 layer; are oriented in bands nearly parallel to

the cell axis in the S2 layer; and in the S3 layer, is nearly perpendicular to that in the S2

layer. The primary wall (P) has an irregular helical arrangement around the cell axis. The

fibers are surrounded by the heavily lignified middle lamellae (M) which is shared by

adjacent fibers. These fibers are made more waterproof by lignin or waxy compounds,

which offer chemical and disease resistance of resulting plants. Hemicelluloses provide

an intimate interlacing and even bonding between the lignin and cellulose (Fan et al.,

1982).

Page 21: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

6

Figure 1.1 A structure model for plant cell wall (Fan et al., 1982).

The composition of lignocellulose depends on its source. Softwood, hardwood,

and herbaceous plants have different cellulose, hemicelluloses and lignin contents.

Page 22: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

7

Generally, cellulose and hemicelluloses account for 35-45% and 15-25% of the dry

matter, individually. Softwood has the highest lignin content of 25-35%, followed by

content of 18~25% in hardwood and content of 10-15% in herbaceous plant. Other than

the different lignin contents, another significant difference between woody biomass and

herbaceous biomass is that herbaceous biomass contains 5-10% ash, whereas woody

biomass hardly contains any ash. The composition of some common lignocelluloses is

summarized in Table 1.1.

Table 1.1 Composition of common lignocelluloses on dry basis (Sun and Cheng, 2002)

Lignocelluloses Cellulose (%) Hemicelluloses (%) Lignin (%)

Hardwoods stems 40–55 24–40 18–25

Softwood stems 45–50 25–35 25–35

Nut shells 25–30 25–30 30–40

Corn cobs 45 35 15

Grasses 25–40 35–50 10–30

Paper 85–99 0 0–15

Wheat straw 30 50 15

Newspaper 40–55 25–40 18–30

Waste papers 60–70 10–20 5–10

Swine waste 6.0 28 NA

Solid cattle manure 1.6–4.7 1.4–3.3 2.7–5.7

Coastal Bermuda grass 25 35.7 6.4

Switchgrass 45 31.4 12.0

Lignocellulose has a complicated structure, which makes it hard to utilize cost-

effectively. More processing steps and harsher conditions are required to convert

Page 23: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

8

lignocelluloses into usable sugars or sugar derivatives than those required for the

conversion of starch in corn kernels. The structures and chemical properties of each

component will be discussed below in detail.

(1) Cellulose

Cellulose with the formula of (C6H10O5)n is a strong and unbranched polymer of

glucose that is found in plant cell walls, as shown in Figure 1.2. Cellulose is produced by

terrestrial plant, from single-celled algae in the oceans to trees on the land. Cellulose is

the most abundant natural polymer on earth. A study completed by the USDA and the

U.S. Department of Energy indicated that at least 1 billion tons of cellulose in the form

of wheat straw, corn stover, other forages and residues, and wood wastes could be

sustainably collected and processed each year in U.S.. This resource represents an

equivalent of 67 billion gallons of ethanol and could replace 30% of the gasoline

consumption in the U.S. (Mosier, 2007).

Page 24: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

9

(a)

(b)

Figure 1.2 Chemical structures of (a) cellulose and (b) starch.

Unlike the α(1→4) and α(1→6) glycosidic linkages in the amorphous starch

polymer, glucose units in cellulose are linked together by β(1→4) glycosidic linkages.

The nature of β(1→4) glycosidic bond allows the cellulose chain to be straight. The

regular arrangement of these straight chains together with the abundant hydroxyl groups

favors the formation of orderly hydrogen bonding among cellulose chains, and allows

cellulose chains to form fibers through several orders of organization. In this way, most

of cellulose in plants exists in the form of a tight crystalline structure. The rigid crystal

structure of cellulose imparts exceptional strength to cellulose and the plant.

Page 25: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

10

However, the crystal structure becomes one of the barriers to the utilization of

lignocelluloses. Due to the strong intra- and inter-molecular hydrogen bonding in

cellulose crystals, cellulose is insoluble in water and also in dilute acid solution at low

temperature. The dissolution of cellulose depends on the disruption of hydrogen bonding

in crystalline cellulose. Cellulose can dissolve in concentrated acid at low temperature

where strong protons can penetrate into cellulose crystallites to break hydrogen bonding

and partially hydrolyze cellulose. In alkaline solutions, cellulose can swell and dissolve

when the polymerization degree of cellulose is lower than 200 (Krassig and Schurz,

2002). The previously mentioned cellulose dissolution is generally accompanied by

significant hydrolysis and degradation of the cellulose. In industry and laboratory,

cellulose solvents such as DMF/lithium halides complex, Cadoxen, and

cupriethylenediamine hydroxide, which work under mild condition, have been widely

investigated (Dawsey and Mccormick, 1990; Turbak et al., 1977). These solvents can

dissolve cellulose without substantially depolymerizing cellulose so that dissolved

material can retain its strength and be regenerated for fiber spinning. Salt solution such

as zinc chloride can also dissolve limited amounts of cellulose at certain temperature and

concentration, but may still lead to cellulose degradation because of the strong Lewis

Page 26: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

11

acidity of zinc ion. In summary, when extracting sugars from lignocelluloses, severe

conditions such as high temperature and concentrated acid are needed to disrupt

cellulose crystallinity to expose single cellulose chain to chemicals or enzymes for

hydrolysis.

(2) Hemicelluloses

Hemicelluloses are made up of highly branched polymers of glucose, arabinose,

galactose, mannose, xylose and uronic acids. The backbones of hemicelluloses include

xyloglucans, xylans, mannans, and glucomannans. These types of hemicelluloses are

present in the cell walls of all terrestrial plants. The detailed structure and content of the

hemicelluloses varies widely between different species and cell types. Hemicelluloses

are a heterogeneous group of polysaccharides. The term was coined at a time when the

structures were not well understood and biosynthesis was completely unknown.

Currently, most working with biomass uses the term hemicelluloses as a convenient

denotement for a group of cell wall polysaccharides that are characterized by having β-(1

→4)-linked backbones of glucose, mannose, or xylose. These polysaccharides all have

Page 27: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

12

an equatorial configuration at C1 and C4 and hence the backbones have significant

structural similarity, as shown in Figures 1.3 and 1.4.

Figure 1.3 Hemicelluloses characterized by a β-(1→4)-linked backbone with an equatorial

configuration at C1 and C4 (Scheller and Ulvskov, 2010).

Page 28: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

13

Xyloglucan [β-D-Glcp-(1→4)]n backbone substituted with side chains as seen in pea and Arabidopsis.

The arrow indicates the typical β-glucanase cleavage site.

Mixed linkage β-glucan [β-D-Glcp-(1→4)]n-β-D-Glcp-(1→3)-[β-D-Glcp-(1→4)]m, where n and m are 3 or 4;

typical of Poales.

Glucuronoarabinoxylan, GAX, typical of commelinid monocots.

Glucuronoxylan, typical dicot structure.

Galactomannan, typical of Fabaceae seeds.

Page 29: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

14

Galactoglucomannan, typical of conifer wood.

: :D-Glucose Glcp :D-Galactose Galp :D-Mannose Manp

:L-Arabinose Araf :D-Xylose Xylp :L-Fucose Fucp

:L-Rhamnose Rhap :D -Glucuronic acid GlcAp

Figure 1.4 Schematic illustration of the types of hemicelluloses found in plant cell walls (Scheller

and Ulvskov, 2010).

Note: The structure of the hemicelluloses varies greatly in different plant species and tissue types. ―Fer‖

represents acylation with ferulate acid (3-methoxy-4-hydroxycinnamate acid), which is characteristic of

xylans in commelinid monocots.

The most common types of polysaccharides that belong to the hemicelluloses are

xylan and mannan. As some minor sugars such as arabinose and galactose, as well as

acetyl groups, branch from the backbone chain, the resulting hemicelluloses are irregular

and therefore amorphous. Grasses generally contain a large amount of glucourono-

arabinoxylan and no mannan. Mannan is generally found in woody biomass, and

softwood contains higher amount of mannan than hardwood. The amorphous structure of

hemicelluloses allows them to be easily accessed by chemical reagent or enzymes.

Page 30: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

15

Hemicelluloses can be extracted by hot water or alkali solutions, and can also be easily

hydrolyzed into monosaccharides at low temperature by acid or enzymes.

(3) Lignin

Lignin occupies the interstitial space of plant cell walls, filling in the left space to

bond cellulose and hemicelluloses together, which results in the strength of the

lignocellulosic matrix, and thus of the entire plant. Lignin has very complicated

structure and is an amorphous polymer mainly composed of three basic monomer units

(monolignols), specifically p-coumaryl alcohol, coniferyl alcohol, and sinapyl alcohol,

as shown in Figure 1.5.

Figure 1.5 Chemical structure of lignin monomers.

Page 31: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

16

Figure 1.6 Formation of resonance-stabilized phenoxyl radicals catalyzed by enzyme.

The polymer forms through radical polymerization of the monolignols. An

example is shown in Figure 1.6 that a phenoxyl radical formed from coniferyl alcohol

and its resonance forms. It can be seen that stabilization of the radical occurs by

coupling to another radical at the position of the unpaired electrons. Coupling of two

radicals forms a dimer with linkages of β-O-4, 5-5, β-5, β-1, 4-O-5 or dibenzodioxocin,

as shown in Figure 1.7. Repeated radical couplings lead to formation of polymeric lignin,

and a model structure of lignin is shown in Figure 1.8. The typical contents of

dominating linkages in lignin are listed in Table 1.2. It can be seen that although the

content of these linkages varies with the types of wood, more than 2/3 of the linkages in

lignin are ether linkages. Hardwood lignin contains about 1.5 times more β-O-4 linkages

than softwood lignin because it contains more syringyl units which cannot couple at the

Page 32: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

17

5-positions. In lignin chemistry, the functional groups that have significant affects on

the reactivity of lignin are the methoxyl, phenolic and aliphatic hydroxyls, and β-O-4

linkages. It has been found that lignin from softwood is derived from more than 90%

coniferyl alcohol with the remaining being mainly p-coumaryl alcohol units. Contrary to

softwoods, lignin from hardwoods is derived from varying ratios of sinapyl and

coniferyl alcohol monomers. Grasses generally contain more p-coumaryl units than

hardwoods and softwoods. In chemical pulping, lignin needs to be removed to separate

fibers. When extracting sugars from lignocelluloses using enzymes, lignin should be

partially removed by cleaving ether linkages to increase the accessibility of the cellulose

to cellulases. After the extraction of sugars, lignin can be generally burned to produce

heat for the conversion process or develop value-added products. Alternatively, through

thermochemical methods, lignin along with cellulose and hemicelluloses can be

simultaneously converted into intermediates (such as syngas and biooil), which could be

catalytically processed into hydrocarbons for fuels.

Page 33: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

18

Table 1.2 Content of the main linkages in lignin (Achyuthan et al., 2010; Pandey and Kim, 2011;

Ralph, 2005; Zakzeski et al., 2010)

Linkage type Softwood (spruce) Hardwood (birch)

β-O-4-Aryl ether 46 60

Dibenzodioxocin 25-30 5-10

β-5-Phenylcoumaran 9-12 6

β-β-(Resinol) 2-6 3-12

4-O-5-Diaryl ether <4 <6.5

β-1-(1,2-Diarylpropane) 1-2 1-2

α-O-4-Aryl ether A few A few

Figure 1.7 Structures of main linkages in lignin (Ralph and Landucci, 2010).

Page 34: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

19

Figure 1.8 A proposed model structure of lignin.

(4) Ash and other components

Other than the three major components listed above, other organic materials

founded in plants include terpenes, resins, and phenols, which can be extracted by water

and organic solvents such as ethanol, benzene, and acetone, thereby named extractives.

Related to terpenes are terpene alcohols and ketones. The resins include a wide variety of

non-volatile compounds, including fats, fatty acids, alcohols, resin acids, phytosterols,

and less known neutral compounds in small amounts. The phenols consist of a large

Page 35: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

20

number of compounds, such as tannins, heartwood phenol, and related substances.

Additionally, low-molecular-weight carbohydrates, alkaloids, and soluble lignin are

extractable as well. The non-cell-wall substances, such as starch, pectin and protein are

not extractable. Generally, extractives do not have significant effects on the

bioconversion process. In gasification or pyrolysis, nitrogen and sulfur from amino acids

will form NH3 and H2S, which can deactivate catalysts in downstream catalytic

processes.

Inorganic components in biomass are usually termed ash, the name given to the

non-aqueous residual components of biomass that remain after it is burned. The

dominating components are alkali and alkali earth carbonates, and oxalates. Silica

deposited as crystals is especially abundant in straws. The ash component of biomass can

forms precipitants such as CaSO4, Ca(OH)2, and Mg(OH)2 in acid or alkali pretreatments.

Additionally, in alkaline pretreatment, silicon dioxide found in high concentration in

grassy plants, can react with alkali to form sodium silicate, which could affect

downstream operations. In gasification, alkali metals also have significant effects on the

syngas composition and might cause slagging, deposition, corrosion and fluidized bed

agglomeration (Hald, 1995; Hedjazi et al., 2009; Lv et al., 2010; Muangrat et al., 2010).

Page 36: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

21

1.3 Technical issues in biomass conversion

Biomass from plants offers a potentially abundant source of sugars for ethanol

fermentation , but its complex laminate structure, consisting of cellulose, hemicelluloses

and lignin, often collectively termed lignocellulose, is difficult to disrupt and break down

into fermentable sugars. Pretreatment is required to separate the components, detach

lignin, and discompose the cellulose fibers for efficient conversion into fermentable

sugars. Therefore, the processing required to break down biomass into fermentable

sugars is more energy-intensive and expensive than obtaining sugar from grain starch or

pressing juice from sugar cane.

Figure 1.9 Flowchart of biorefining.

Sugarcane

Starch

Cellulose

Oil

Pretreatment and

Enzymatic hydrolysis

Gasification

Microbial fermentation

Acid hydrolysis

Pyrolysis

Catalytic fuel synthesis

-FT process

-Aldol condensation

-Oligomerization of butene

Catalytic cracking

Transesterification

Methanol

Ethanol

Butanol

……

Gasoline

Jet fuel

Diesel

Biodiesel

l

H2

CH4

DME

……

Page 37: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

22

Primary approaches of converting biomass into biofuels are summarized in

Figure 1.9. Fermentation of sugars into ethanol is a very mature technique in the brewing

industry, therefore attracting more attention than other types of liquid biofuel production

processes. Starch, which can be easily hydrolyzed chemically or enzymatically into

sugars at a relatively low cost, was the first feedstock used to produce commercial

bioethanol, but the relatively high price of starch and its potential threat to food security

limit the use of starch as the feedstock to meet the increasing demand for renewable fuels.

Therefore, studies on bioethanol production from cheap and abundant lignocelluloses

have been increasing readily in recent years (Smith, 2008). However, cellulose in

lignocelluloses is enclosed with hemicelluloses and lignin, which makes it more

challenging to be hydrolyzed into glucose than starch; severe conditions, such as high

temperature, and high pressure, are needed to hydrolyze it or activate it (Demirbas,

2005). It is already widely studied that sugar can be produced through swelling of

crystalline cellulose in ground lignocelluloses with concentrated acid, including sulfuric

acid, hydrochloric acid, and phosphoric acid, followed by the hydrolysis of the swelled

cellulose at dilute acid concentration (Miller and Hester, 2007; Zhu et al., 2009).

Page 38: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

23

Unfortunately, the acid hydrolysis process has issue with equipment corrosion and the

difficulty of sugar-acid separation, thereby limiting their applicability.

Currently, the two-step process, pretreatment followed by enzymatic hydrolysis,

is the most studied one for extracting sugars from lignocellulose for cellulose ethanol

production. Pretreatment aims at removing recalcitrance of lignocelluloses and

simultaneously decrystallizing and/or prehydrolyzing cellulose. A good pretreatment

method should retain as much cellulose as possible for the subsequent enzymatic

hydrolysis step while improving the enzymatic hydrolysis ability of the pretreated

materials (Piccolo and Bezzo, 2009; Tomas-Pejo et al., 2008; Xu et al., 2009). Generally,

pretreatment conditions are conducted under temperatures of 150-190 °C indicating that

pretreatments are energy-intensive processes and need high-quality facilities. High

temperatures also cause degradation of sugars into byproducts, which not only decreases

sugar yield, but also inhibits organisms in the following fermentation steps. Enzymatic

hydrolysis is to saccharify cellulose with cellulases, unfortunately, which needs costly

enzymes, large amounts of buffer solution, large hydrolysis tank, and long hydrolysis

time (Banerjee et al., 2010). Besides, substrate consistency in enzymatic hydrolysis is

typically lower than 15-20% (w/v) because of stirring and mixing issues, which limits

Page 39: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

24

the end sugar concentration to 6-15% (w/v) after hydrolysis, and thereby the ethanol

concentration to 2.5-6% (w/v) after fermentation. Ethanol concentration of 4% (w/v) is

deemed as the minimum for the distillation process to be economical, as the energy

required for distillation is significantly reduced for ethanol concentrations above 4%

(w/v). Therefore, low ethanol concentration will significantly increase the ethanol

distillation cost (Modenbach and Nokes, 2012). In a typical cellulosic bioethanol

production scenario, feedstock, pretreatment, and enzymatic hydrolysis each accounts

for around 1/4 of the total cost and at least 1/8 of total cost is attributed to ethanol

distillation (Aden et al., 2002). Furthermore, pretreatment and enzymatic hydrolysis

require the post-treatment of large volume of waste water from pretreatment and

fermentation (Lingaraju et al., 2012). In addition, a pretreatment method may be

unsuitable to all types of lignocelluloses. For example, AFEX (Ammonia Fiber

Expansion) pretreatment does not work well with woody biomass. These issues together

result in the high production cost of current cellulose ethanol, which is the key factor

retarding the commercialization of cellulose ethanol. In summary, the challenges of

cellulose ethanol include (1) developing effective pretreatment technology to produce

readily digestible lignocellulose substrate, (2) lowering the cost and energy consumption

Page 40: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

25

for pretreatment, (3) developing high-activity and low-cost cellulases, (4) reducing

enzyme loading, and (5) developing process and equipment for high-consistency

enzymatic hydrolysis of lignocellulosic substrate to raise the end concentration of sugar

stream for ethanol fermentation.

In order to overcome the corrosion problem of acid and avoid high cost of

cellulose hydrolytic enzymes, solid acids such as zeolite (Zhang and Zhao, 2009),

silica/carbon nanocomposites (Van de Vyver et al., 2011), sulfonated carbon materials

such as graphene (Hara et al., 2009) and CMK-3 (a type of mesoporous carbon)

(Kobayashi et al., 2010; Pang et al., 2010), and layered niobium molybdate (HNbMoO6)

(Takagaki et al., 2008; Takagaki et al., 2010), have been investigated to hydrolyze

cellulose. However, the reported yields of cellulose hydrolysis to glucose were not

satisfactory as the solid acids have limited contact with cellulose, which significantly

slows down the hydrolysis rate.

Due to the limitations and disadvantages of bioethanol, as discussed above,

interests in converting biomass into hydrocarbons is increasing as the hydrocarbons have

the same physicochemical properties as traditional transportation fuels from petroleum

(Elliott and Schiefelbein, 1989; West et al., 2009). Hydrocarbons can be produced from

Page 41: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

26

biomass through either biological or chemical processes. In biological processes, sugars

can be fermented into hydrocarbons, but this process still requires a sugar source and

efficient organisms. Bypassing the sugar platform, biomass could be chemically

converted into a series of precursors, such as syngas by gasification of biomass, bio-oil

by pyrolysis of biomass, and HMF or levulinic acid by dehydration of sugars. Although

thermochemical processes seem simple, the purification and upgrading of the precursors

are quite complicated and challenging. In gasification, the yield and composition of

syngas are greatly dependent on feedstock composition such as carbohydrate, lignin, ash,

and extractives. Carbohydrate and lignin content will directly affect how much oxygen is

needed to supply into reactor. Metal ions in ash can catalyze different decomposition

pathways of biomass. Protein in extractives can lead to the formation of H2S and NH3,

which should be removed to avoid deactivating catalysts. For pyrolysis of biomass, the

intermediate bio-oil is a very complicated and instable mixture with more than 300

compounds, of which most are carboxylic acids, aldehydes, and aromatic compounds.

Because limited oxygen is supplied in the pyrolysis process, carbonized byproducts such

as charcoal and tar form. Because of the strong acidity and high oxygen content of bio-

oil, bio-oil needs to be purified, upgraded, and deoxygenated for advanced applications.

Page 42: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

27

The mostly investigated upgrading method is cracking bio-oil on zeolite catalysts where

oxygen of bio-oil is removed with formation of aromatic hydrocarbons. However,

because of the strong acidity of zeolite and absence of extra hydrogen supply, a lot of

charcoal caused by strong dehydration will form in the upgrading process, which not

only decreases the recovery yield of product, but also deactivates catalyst. In summary,

thermochemical methods for converting biomass into hydrocarbons still face a lot of

challenges, such as low product selectivity, product purification and upgrading, and

catalyst deactivation (Balfanz et al., 1993; Kirubakaran et al., 2009; Zhang and Wyman,

2011; Zhao et al., 2007).

Extensive work has been done to produce hydrocarbons from lignocelluloses

based on hydroxymethylfurfural (HMF) or levulinic acid platforms. Direct conversion of

glucose into hexane on bifunctional catalyst was reported through hydrodeoxygenation

(Huber et al., 2004). However, hexane was too volatile to be used as transportation fuel.

Currently, one of the potential ways to produce extended hydrocarbon chain from sugars

was through aldol-condensation between furfural (or HMF) and acetone (Chheda et al.,

2007). This process produced HMF from glucose in DMSO (dimethyl sulfoxide); the

resulting HMF was extracted from the reaction mixture with MIBK; and the third step

Page 43: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

28

was to conduct aldol condensation of HMF and acetone. The process involved multiple

steps and was very complex. In addition, this process could only use soluble sugars as

feedstock, and efforts were still needed to overcome the barrier of producing sugars from

lignocelluloses.

Furthermore, the intermediates (HMF and furfural) could not be cost-effectively

produced in high yield as HMF tends to polymerize and form insoluble humin in acidic

aqueous solution (Vandam et al., 1986). Many studies were conducted to improve the

selectivity of sugars to furfural/HMF. For example, organic solvent of dimethyl

sulfoxide (DMSO) was used to replace water, and the water-free environment promoted

the dehydration of glucose into HMF (Amarasekara et al., 2008). In another study,

methyl isobutyl ketone (MIBK) was added as extraction solvent to collect HMF

(Roman-Leshkov et al., 2006). The HMF formed was immediately extracted into upper

MIBK layer, which largely reduced the opportunity of the polymerization/condensation

of HMF into insoluble humin. However, it is difficult to separate HMF from the solvents

(DMSO and MIBK) because of their similar boiling points.

It was reported that HMF could be produced in high yield and with high

selectivity using ionic liquids as solvent and chromium halide as catalyst (Zhao et al.,

Page 44: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

29

2007). The issues with this process were that ionic liquid is very expensive and difficult

to recycle, and chromium halide is potentially toxic to environment. Partial replacement

of ionic liquid with traditional cellulose solvent was also investigated (Binder and Raines,

2009). High conversions up to 90% were achieved from monosaccharide and pure

cellulose, whereas the yield was only 50% from corn stover powder. This complicated

system involved two expensive organic solvents (dimethylacetamide (DMAC) and ionic

liquid) as well as inorganic catalysts, which made them extremely difficult to recycle. In

addition, the ionic liquid has high viscosity and works well only in water-free

environment, which requires the feedstock to be finely ground and completely dry.

Unfortunately, both grinding and drying of biomass are energy-intensive.

In summary, currently there is still no cost-effective way to produce sugars and

sugar derivatives such as HMF and levulinic acid from lignocelluloses. There are still

strong demands for efficient ways to convert lignocelluloses into sugars, HMF and

levulinic acid selectively.

Page 45: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

30

1.4 Project description

After reviewing existing techniques of producing biofuels from lignocelluloses,

sugar is an important platform compound for producing both bioethanol and

hydrocarbons because of its simple and stable chemical state, which makes subsequent

conversion processes more controllable. According to present technologies, sugars could

be fermented or catalytically converted into not only fuels but also important chemicals,

such as lactic acid, diols, and H2 (Davda and Dumesic, 2004; Onda et al., 2008;

Palkovits et al., 2010).

Therefore, a cost-effective way of producing sugars from lignocelluloses is still

crucial to the development of a biorefining industry. Direct transformation of

lignocelluloses to advanced biofuels like hydrocarbons is badly needed. Therefore, in

response to these issues, my dissertation work will focus on four projects:

(1) Investigating the effectiveness of SPORL and dilute acid pretreatments for

enzymatic saccharification of spruce for bioethanol production. (Chapter 2)

(2) Synthesizing cellulase-mimetic solid acids for cellulose hydrolysis. (Chapter 3)

(3) Developing a new process to directly saccharify lignocelluloses into sugars

without pretreatment and enzymatic hydrolysis. (Chapter 4)

Page 46: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

31

(4) Developing a new process to directly convert lignocelluloses into hydrocarbon

precursors without prior pretreatment or saccharification. (Chapter 5)

It is the author’s hope that the present work will provide cost-effective ways of

utilizing lignocelluloses, accelerating the development of biorefinery industry.

Page 47: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

32

Chapter 2: Comparative Study of Dilute Acid and

SPORL Pretreatments of Spruce for

Enzymatic Saccharification

2.1 Introduction

Bioethanol, produced from lignocelluloses, is a promising alternative to fossil

fuel for vehicular transportation (Carroll and Somerville, 2009; Ragauskas et al., 2006).

The benefits of cellulosic ethanol include, but are not limited to, reduced greenhouse

gases emission, value-added utilization of agricultural and forest residues, enhancement

of the rural economy, and improved national energy independence and security (Farrell

et al., 2006; Scharlemann and Laurance, 2008). A typical process for cellulosic ethanol

production consists of three steps: feedstock pretreatment to enhance cellulose

accessibility to the cellulases, enzymatic saccharification of the cellulose to glucose,

followed by fermentation of the glucose to ethanol. One of the barriers to the

commercialization of cellulosic ethanol is the lack of economical and effective

technologies for the feedstock pretreatment. Pretreatment is a necessary operation

required to achieve optimal bioconversion for all forms and types of lignocellulosic

Page 48: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

33

feedstocks to ethanol, but is particularly important for the more recalcitrant softwoods.

The significance and importance of the pretreatment step cannot be overemphasized, as

the effectiveness of the pretreatment affects the upstream selection of biomass, the yield

of fermentable sugars, and the chemical and morphological characteristics of the

pretreated substrate which in turn govern downstream hydrolysis/saccharification.

An effective pretreatment method should be economical (in terms of both capital

and operating costs) and effective for a variety of lignocelluloses. Specifically, it should

require minimal feedstock preparation and preprocessing prior to pretreatment,

maximally recover all lignocellulosic components in usable forms with minimal

formation of fermentation inhibitors, and produce a readily digestible cellulosic substrate

that can be easily hydrolyzed with a low loading of enzymes. In the last several decades,

research and development efforts have made significant progress in pretreatment

technologies for lignocellulosic feedstocks (Chandra et al., 2007; Hendriks and Zeeman,

2009; Lloyd and Wyman, 2005; Lynd et al., 2002; Mosier et al., 2005). Many

pretreatment technologies, such as lime, dilute acid, hot water, ammonia, steam

explosion, SPORL and organosolv pretreatments have achieved varying levels of success

Page 49: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

34

(Chandra et al., 2007; Jeong et al., 2009; Marbe et al., 2006; Monavari et al., 2009;

Nguyen et al., 2000; Sendich et al., 2008; Sierra et al., 2009; Wingren et al., 2008).

Dilute-acid pretreatment (DA) is one of the most investigated pretreatment

methods, typically using sulfuric acid at high temperature (160-200 ºC) (Lloyd and

Wyman, 2005; Schell et al., 2003). DA enhances the digestibility of lignocellulose

mainly by hydrolyzing hemicelluloses and partially hydrolyzing cellulose. DA can

achieve satisfactory levels of cellulose saccharification for agricultural residues and

some hardwood species, but is not effective for softwoods. The low pH value of the

dilute acid process causes serious equipment corrosion problem. In addition, high

temperature and low pH lead to formation of significant amounts of fermentation

inhibitors, notably furfurals from hemicellulosic sugars. Furthermore, dilute acid

pretreatment causes extensive condensation of lignin, which diminishes the commercial

value of lignin for co-products development.

Sulfur dioxide (SO2)-catalyzed steam explosion is another acidic pretreatment

using milder (higher pH) condition, which reduces equipment corrosion and minimizes

the generation of fermentation inhibitors. The process has been extensively studied by

Zacchi’s and Saddler’s groups (Chandra et al., 2007; Galbe and Zacchi, 2002).

Page 50: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

35

Feedstock is first treated using gaseous SO2 and then subjected to steam explosion. The

process works well with agricultural residues and some hardwoods, but performs less

satisfactorily with softwoods. Two-stage steam explosion can improve the enzymatic

digestibility and overall sugar recovery from softwoods, but the advantages are partially

outweighed by increased energy consumption and operation cost (Galbe and Zacchi,

2002). The toxicity of gaseous SO2 is another concern in application. Furthermore, the

mild SO2 treatment is unable to sulfonate or dissolve lignin significantly. The lignin

extensively condensed during the subsequent steam explosion (Shevchenko, 1999),

leading to recalcitrance of steam exploded softwood substrate (Marbe et al., 2006).

Recently, we developed and reported a novel pretreatment method, Sulfite

Pretreatment to Overcome Recalcitrance of Lignocelluloses (SPORL), for robust

conversion of woody biomass including softwood to sugars (Lu et al., 2009; Melillo et

al., 2009). The pretreatment consists of a short chemical treatment of feedstock followed

by mechanical size reduction (fiberization). Wood chips or other feedstocks first react

with a solution of a sulfite salt (e.g., Na, Mg, or Ca) at 160-180 ºC and pH 2-4 for about

30 min, and are then fiberized (size-reduced) using a disk mill to generate fibrous

substrate for subsequent saccharification and fermentation. SPORL treatments produced

Page 51: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

36

readily digestible substrates due to partial removal of lignin and lignin and partial

hydrolysis of cellulose. Because of the decreased acidity of SPORL liquor, most of

hydrolyzed cellulose and hemicellulose sugars were recovered with low amount of

fermentation inhibitors. Energy consumption for post-SPORL size-reduction of wood

chips was about 30 Wh/kg, equivalent to those consumed for size-reduction of

agricultural biomass. In addition, direct pretreatment of commercial wood chips afforded

a low liquid-to-wood ratio (<3) (Lu et al., 2009), which can lead to considerable thermal

energy savings and produce a concentrated hemicelluloses sugar stream. Because the

lignin dissolved in SPORL pretreatment hydrolysate is sulfonated (lignosulfonate), it has

a variety of commercial applications within the established market. It should be noted

that SPORL pretreatment is compatible with SO2-catalyzed steam explosion. One can

conduct sulfite-catalyzed steam explosion by using sulfite in addition to SO2 as catalysts

to significantly improve cellulose enzymatic saccharification, especially for softwoods.

To further understand the fundamentals of the SPORL pretreatment, in this study, we

systematically compared SPORL with the DA for the pretreatment of a softwood, spruce.

In order to investigate the effects of sulfite addition on the pretreatment process, the

same reaction conditions were used for DA and SPORL pretreatments, excepting that

Page 52: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

37

sulfite was added in SPORL pretreatment. Softwoods are notoriously more recalcitrant

than other types of biomass and was chosen as feedstock in this experiment. The

experimental flowchart is summarized in Figure 2.1.

Figure 2.1 Flowchart of experiment and analysis.

Air-dried Spruce

Powder (40-mesh)

Extractive-free

Spruce

Pretreated

Spruce

NMR

Analysis

Chemical

Analysis

Saccharide analysis

Acid-insoluble lignin

Acid-soluble lignin

Inhibitor and

Sugar Analysis

DA/SPORL

Pretreatment

Enzymatic

Hydrolysis

Spent Pretreatment

Liquor

Water/Ethanol

Extraction

In-vitro Ruminal

Fermentation

Page 53: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

38

2.2 Experimental

2.2.1 Materials

Fresh spruce chips were generously provided by the Wisconsin Rapids mill of

Stora Enso North America (now New Page Corporation, Miamisburg, Ohio). After being

air-dried, the chips were ground passing a 40-mesh screen using a Wiley mill. Chemical

composition of the spruce wood is presented in Table 2.1. Commercial enzymes,

Celluclast and β-glucosidase produced by Novozymes, were purchased from Sigma-

Aldrich (St. Louis, MO) and used as received. All the chemical reagents used in this

study were purchased from Fisher Scientific (Pittsburgh, PA) and used as received.

2.2.2 Pretreatments

Chemical pretreatments were conducted in batch mode using a lab-scale rotatable

reactor, as described previously (Lu et al., 2009; Melillo et al., 2009). Both SPORL and

dilute acid pretreatment were carried out in triplicate; the average results of the three

runs were reported. In general, ground spruce (100 g oven-dry material) was loaded into

a 1 L stainless steel vessel. Prepared pretreatment solution (500 mL 1% H2SO4 for DA

pretreatment or 500 mL 1% H2SO4 + 9 g Na2SO3 for SPORL pretreatment) was then

poured into the vessel. Three sealed 1 L vessels were mounted inside a 23 L stainless

Page 54: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

39

steel. The system was heated via the external steam jacket and rotated at a speed of 2

rpm to provide mixing during pretreatments. The temperature was raised to 180 ºC in

about 7 min and maintained for an additional 30 min. At the end of the pretreatment, the

pretreatment spent liquor was separated from the solid (pretreated substrate) by filtration

and stored for fermentation study and chemicals analysis. The solid substrate was

collected in a Buchner funnel on filter paper and washed thoroughly with water.

Substrate yield was determined from the measured wet weight and moisture content of

the washed solid substrate.

2.2.3 Enzymatic Hydrolysis

Enzymatic hydrolysis of the pretreated substrates and original ground spruce was

conducted as described previously (Ahring and Langvad, 2008; Weimer et al., 2005).

Briefly, the hydrolysis was carried out at 50 ºC on a shaking incubator (Thermo Fisher

Scientific, Model 4450, Waltham, MA) at 150 rev/min. Substrate equivalent to 2 g

glucan was loaded into a 250 mL Erlenmeyer flask with 100 mL of 0.05 M sodium

acetate buffer (pH 4.8). Approximately 4 mg of tetracycline chloride was used to control

the growth of microorganisms and prevent consumption of liberated sugars. Cellulase (5

or 15 FPU, Filter Paper Units, per gram glucan) and β-glucosidase (10 or 30 IU,

Page 55: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

40

International Units, per gram glucan) were loaded into the flask. Hydrolysates were

sampled periodically and subjected to glucose analysis. The hydrolysis was conducted in

duplicates for each substrate; the average is reported here.

2.2.4 Analytical Methods

Acid-insoluble lignin of spruce and pretreated SPORL and DA substrates was

determined according to National Renewable Energy Laboratory (NREL) Analytical

Procedure with modifications. Acid-soluble lignin was determined by UV at 205 nm

using an extinction coefficient of 110 L·g-1

·cm-1

(Dence, 1992).

Saccharides analysis was conducted using a Dionex High Performance Ion

Chromatography (HPIC) system (ICS-3000) equipped with integrated amperometric

detector and CarbopacTM

PA1 guard and analytical columns at 20 ºC. Eluent was

provided at a rate of 0.7 mL/min, according to the following gradient: 0→25 min, 100%

water; 25.1→35 min, 30% water and 70% 0.1 M NaOH; 35.1→40 min, 100% water. To

provide a stable baseline and detector sensitivity, 0.5 M NaOH at a rate of 0.3 mL/min

was used as post-column eluent.

Page 56: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

41

Fermentation inhibitors generated in pretreatment including acetic acid, formic

acid, furfural, levulinic acid and 5-hydroxylmethylfurural (HMF) were analyzed using

the Dionex ICS-3000 equipped with a Supelcogel C-610H column at temperature 30 ºC

and UV detector at 210 nm. Eluent was 0.1% phosphoric acid at a rate of 0.7 mL/min.

Cellulase activity was determined using the filter paper assay recommended by

the International Union of Pure and Applied Chemists (Ghose, 1987) and is expressed in

terms of filter paper units (FPUs). β-Glucosidase activity was determined using p-

nitrophenyl-β-D-glucoside as the substrate (Cyr et al., 1988) and is expressed in terms of

International Units (IUs).

2.2.5 Whole Cell-Wall NMR of substrates

Whole cell-wall NMR of pretreated substrate and original spruce was conducted

according to (Boraston et al., 2003). In brief, 1.5 g of extractive-free spruce powder or

pretreated substrate was loaded into a 50 mL ZrO2 jar and ball-milled on a Retsch PM-

100 ball mill for 10 h (20 min on and 10 min off). The ball-milled sample (600 mg) was

dissolved in 10 mL dimethyl sulfoxide (DMSO) and 5 mL N-methylimidazole (NMI). A

clear solution was formed in approximately 3 h, depending on the sample. Excess acetic

anhydride (3 mL) was added to the solution, and the mixture was stirred for 1.5 h. The

Page 57: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

42

resulting clear brown solution was dropped into 2000 mL water by a glass pipette, and

the mixture was allowed to stand overnight. The precipitate was recovered by filtration

through a nylon membrane (0.2 μm). The product was washed with water (250 mL) and

freeze-dried. About 60 mg dried powder was dissolved in CDCl3 for NMR spectrometry.

NMR spectra were acquired on a Bruker DRX-360 instrument fitted with a 5 mm

1H/broadband gradient probe with inverse geometry.

2.2.6 Degree of Polymerization of Cellulose

The degree of polymerization of cellulose was indirectly determined by a

viscometry method. The viscosity of cellulose in cupriethylenediamine solution was

measured using a KV 3000 kinematic viscosity bath (Koehler Instrument Company) with

a Cannon Ubbelohde capillary viscometer, according to TAPPI (Technical Association

of Pulp and Paper Industry) Standard Method T230-om-99. Cellulose samples for

viscosity measurement were prepared from original spruce and SPORL and DA

substrates by delignification using sodium chlorite according to the PPTAC (Pulp and

Paper Technical Association of Canada) Useful Method G.10U.

2.2.7 Fermentability of SPORL and DA Pretreatment Liquors

Fermentability of the pretreatment spent liquors was evaluated according to an in

Page 58: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

43

vitro ruminal fermentation assay (Weimer et al., 2005). In vitro ruminal experiments

were conducted at 39 ºC using a single replicate of each sample, and replication was

achieved through a second in vitro run. Triplicates of a standard ryegrass were included

in each run. Incubations were conducted in nominal 60 mL serum bottles (volume was

calibrated to 0.01 mL) and that contained the equivalent of 100 mg (weighed to 0.1 mg)

of biomass material, 6.7 mL of Goering and Van Soest buffer, 0.3 mL of cysteine-sulfide

reducing agent (6.25 g/L each of cysteine HCl and Na2S·9H2O) and a CO2 gas phase.

Gas pressure readings were made at 24 h and at 96 h, using a SenSym digital pressure

gauge modified to accept a 20 gauge hypodermic needle.

2.3 Results and Discussion

2.3.1 Changes in Cell-Wall Components after pretreatments

As described in the Experimental section, ground spruce wood (40 mesh) rather

than wood chips was used as the feedstock, and mechanical size reduction was not

included in either SPORL or DA pretreatment. The consideration of doing so is that a

refining step adds difficulties to mass balance evaluations. In addition, only physical size

changes and no (significant) chemical reactions are expected during the mechanical size

Page 59: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

44

reduction. Both SPORL and DA chemical pretreatments were carried out at the same

temperature (180 ºC), pretreatment time (7 min to attain the desired temperature and 30

min at that temperature), ratio of liquor to wood (5:1), and acid loading (5% on a dry

wood basis). The only difference is that 9% (on dry wood) sodium sulfite was added in

SPORL pretreatment.

Chemical composition of the spruce wood is presented in Table 2.1. It had a

composition of 29.0% acid-insoluble lignin and 46.7% glucose (or 42.1% glucan). The

majority of the hemicellulosic sugars were hexoses (10.8% mannose and 2.6% galactose),

accompanied by 5.5% xylose and 1.2% arabinose. The chemical compositions of

SPORL and DA pretreated substrates are compared in Table 2.1. The DA pretreatment

apparently hydrolyzed all hemicellulose from the feedstock spruce, leaving only

cellulose and lignin in the DA substrate. Essentially no delignification (but lignin

condensation) occurred during the DA pretreatment, so the lignin was enriched in the

substrate (48.5% acid-insoluble lignin) after the hemicelluloses and part of the cellulose

were dissolved. On the other hand, SPORL pretreatment retained more cellulose

(SPORL 86.3% vs. DA 71.3%, as shown in Figure 2.3) and a low level of mannan, but

less lignin (32.9% acid-insoluble lignin), in the substrate. The extent of delignification

Page 60: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

45

(percentage of original wood lignin dissolved) during the SPORL pretreatment was

approximately 32%, calculated from the total lignin contents of spruce wood and the

SPORL substrate (Table 2.1) and the substrate yield (Figure 2.3). The differences

between DA and SPORL substrates were caused by the addition of the sulfite. First,

sodium sulfite is alkaline, buffering the pH value from 1.2 in the DA pretreatment liquor

to 2.7 in the SPORL liquor, which protected the cellulose and hemicelluloses from

extensive hydrolysis and further degradation at elevated temperature to inhibitors

(Tables 2.1 and 2.2) and prevented lignin from extensive condensation. Second, the

sulfite introduced sulfonic groups at the lignin benzylic carbons, which may (1) partially

depolymerize and dissolve the lignin and (2) increase the hydrophilicity of the residual

lignin that is retained in the pretreated substrate. These two effects are important to the

enzymatic digestibility of the SPORL substrate, as discussed in section 2.3.3.

Page 61: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

Table 2.1 Chemical analyses of original spruce, pretreated spruce substrates and spent pretreatment liquors

Note: (1) % on oven-dry material; (2) concentration in the liquor, g/L; (3) ND-not detected; and (4) NA-not applicable.

Arabinose Galactose Glucose Xylose Mannose Acid-insoluble lignin Acid-soluble lignin

Original spruce1 (%) 1.2±0.0 2.6±0.1 46.7±0.9 5.5±0.2 10.8±0.4 29.0±1.1 1.1±0.1

DA substrate1 (%) ND

3 ND 51.9±2.1 ND ND 48.5±1.2 0.8±0.0

SPORL substrate1 (%) ND ND 66.6±3.1 ND 1.2±0.1 32.9±0.8 1.2±0.0

DA liquor2 (g/L) 0.2±0.0 0.8±0.0 6.0±0.2 0.4±0.0 1.7±0.1 NA

4 4.8±0.2

SPORL liquor2 (g/L) 0.8±0.0 2.6±0.0 5.7±0.4 4.3±0.1 9.0±0.1 NA 16.6±0.4

46

Page 62: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

47

The concentrations of sugars and lignin in the SPORL and DA liquors, listed in

Table 2.1, provided insight into the effects of sulfite on the reactions of carbohydrates

and lignin, discussed above. The SPORL liquor contained more sugars, except for

glucose, than the DA liquor. The reason is that the higher pH value of the SPORL liquor

limited sugar degradation at high temperature. This is supported by the level of inhibitors

derived from the sugars (Table 2.2) in the spent liquors. The slightly higher content of

glucose in the DA liquor was due to the enhanced acidic hydrolysis of cellulose in the

DA pretreatment, which has been reflected by the low cellulose content in the DA

substrate. The SPORL liquor contained significantly more soluble lignin detected by UV

method than did the DA liquor. The lignin in the SPORL liquor was in the form of

lignosulfonate with a yield of 8.3% (on dry wood, calculated from the amount of the

dissolved lignin in the SPORL liquor estimated by a UV method). Lignosulfonate has

been the most successful lignin product in the market. It has been widely used as

dispersants for carbon black, pesticides, dyestuffs and pigments; emulsifiers for soils,

asphalt, waxes and oil in water; additives for drilling mud and concrete; and adhesives

and binders for animal-feed pellets, minerals, and laminates (Fengel and Wegener, 1984).

Page 63: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

48

The SPORL pretreatment therefore has potential for high-value lignin co-products

development.

Table 2.2 Concentrations of major fermentation inhibitors in SPORL and DA pretreatment spent

liquors

Note: ―+‖ and ―-― represent standard deviation.

To investigate the behavior of cellulose during the SPORL and DA pretreatments,

viscometry was used to provide an indication of cellulose depolymerization (Table 2.3).

The viscosity of the cellulose solution from the substrates was only approximately one

tenth of that from original spruce wood, implying that the cellulose was significantly

depolymerized (hydrolyzed) during both pretreatments. This is one of the reasons why

the pretreated substrate had better enzymatic digestibility than the untreated wood (Figure

2.4). No significant difference was found between SPORL and DA substrates in cellulose

viscosity. DA cellulose solution showed a slightly lower viscosity because the lower pH

value of DA pretreatment enhanced hydrolysis of cellulose.

Spent liquor

Inhibitors (g/L)

Acid-soluble

lignin

Formic

acid Acetic acid Furfural HMF

Levulinic

acid

DA liquor 4.8±0.2 7.4±0.3 5.3±0.4 2.9±0.1 4.7±0.1 11.4±0.6

SPORL liquor 16.6±0.4 1.9±0.2 2.7±0.7 1.3±0.2 2.0±0.5 3.2±0.5

Page 64: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

49

Table 2.3 Viscosity of cellulose solutions from untreated- and pretreated-spruces

Note: ―+‖ and ―-― represent standard deviation.

To further understand the changes of cell-wall components, in particular lignin,

during the pretreatments, whole cell-wall solution-state two-dimensional HSQC NMR

methods (Boraston et al., 2003) were used to compare the SPORL and DA pretreated

substrates with the original spruce. As shown in Figure 2.2, the cell wall HSQC NMR

spectrum from untreated spruce is typical for a softwood, showing the dominant C-H

correlations (colored in green) from cellulose and hemicelluloses (colored in black) along

with correlations from major lignin substructures (β-aryl ethers A and phenylcoumarans

B). The C-H correlations in the aromatic region of the spectrum also show typical

softwood lignin correlations from (almost) solely guaiacyl (G) units. After pretreatment

with SPORL and DA, the spruce residues contain mainly cellulose and lignin-derived

material as shown by their HSQC NMR spectra, which are consistent with the chemical

analysis results (Table 2.1). For the residual lignin, although the methoxyl signal is

readily seen, the absence of recognizable lignin subunits A and B suggests that

considerable degradation of the side chain has occurred. Although no detectable

Material Viscosity at 25 ºC (mPa·s)

Untreated spruce 25.6±0.3

SPORL substrate 2.7±0.0

DA substrate 2.4±0.0

Page 65: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

50

monomeric sugars are in the DA pretreated sample, HSQC NMR shows some

correlations from non-cellulosic carbohydrates (colored in black). These may be

condensed with the residual lignin preventing their release by acid hydrolysis (Yasuda

and Murase, 1995), and thus would be unavailable for further conversion to ethanol.

Meanwhile lignin itself can undergo condensation reactions under acidic conditions. The

most common condensation reactions occur between benzylic α-positions of lignin

sidechains and 5-positions of other aromatic rings, producing diarylmethane structures

(Gierer, 1985). The C-H correlations (colored in pink) in the aromatic regions of these

spectra from pretreated spruce suggest that such condensation reactions indeed occurred

during both pretreatment processes. During SPORL treatment lignin was partially

sulfonated producing lignosulfonate soluble in water so that wood substrates were

partially delignified. That explains why the SPORL treated substrate has less lignin than

the DA treated material. Sulfonated lignin structures in the SPORL treated spruce

residues were not detected by NMR; model work is needed to confirm the presence. The

reason for such observations is that the number of lignin units on the substrate surface

only accounts for very small part of total lignin units and not every lignin unit is

sulfonated.

Page 66: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

51

Figure 2.2 HSQC NMR spectra of untreated and pretreated spruce cell walls.

Page 67: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

52

2.3.2 Mass balance of sugars after pretreatments

An ideal pretreatment is not only able to produce a readily digestible substrate,

but also to maximize the recovery of sugars in a fermentable form with limited formation

of inhibitors. As shown in Figure 2.3, the feedstock spruce wood was separated into two

fractions after the pretreatments, solid substrate and a liquid stream (spent pretreatment

liquor) containing dissolved sugars, lignin and sugar degradation products. The SPORL

substrate appeared lighter in color than the DA substrate because of its lower lignin

content and less-condensed lignin. SPORL liquor was darker than DA liquor, probably

because the former contained more soluble lignin (Table 2.1) than did the latter. To

compare the performance of SPORL and DA in sugar recovery during the pretreatment, a

mass balance of sugars during the two pretreatments was performed (Figure 2.3). From

100 g of oven-dry spruce wood (containing 46.7 g glucose, 2.6 g galactose, 10.8 g

mannose, 1.2 g arabinose, and 5.5 g xylose), 64.1 g (containing 33.3 g glucose) and 60.5

g (containing 40.3 g glucose and 7.1 g mannose) substrates were obtained from DA and

SPORL pretreatments, respectively. High substrate yield from DA pretreatment was due

in part to the retention of almost all of the original lignin. Total detected sugars in the

pretreatment liquors were 4.6 g (3.0 g glucose, 0.4 g galactose, 0.9 g mannose, 0.1 g

Page 68: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

53

arabinose, and 0.2 g xylose) for DA pretreatment and 11.3 g (2.9 g glucose, 1.3 g

galactose, 4.5 g mannose, 0.4 g arabinose, and 2.2 g xylose) for SPORL, respectively.

The calculations from these data indicated that total sugar recovery was 56.7% (DA) and

87.9% (SPORL), and glucose recovery was 77.7% (DA) and 92.5% (SPORL).

Comparing the recovery of hexoses and pentoses, it was found that the recovery of

hexoses was much higher than that of pentoses, 62.8% vs. 4.5% for DA and 93.7% vs.

38.8% for SPORL, respectively, indicating the pentoses were easier to further degrade at

high temperature and low pH than were the hexoses. The results above clearly indicate

that under the same acid loading, temperature and reaction time, SPORL is superior to

DA for the recovery of total sugars–hexoses and pentoses. As discussed previously, this

is likely because of the higher pH value of the SPORL pretreatment liquor formed by the

addition of sulfite. High sugar recovery in SPORL process implies that fewer sugars were

degraded and limited inhibitors were generated, which will benefit the fermentation of

the liquors.

Page 69: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

54

Figure 2.3 Mass balance of saccharides during the DA and SPORL pretreatments.

2.3.3 Enzymatic digestibility of pretreated spruce

The substrate characteristics that affect enzymatic hydrolysis of cellulose include

hemicellulose content; lignin structure, distribution and content; cellulose crystallinity

and degree of polymerization; and surface area, pore size and particle size of the substrate

(Lu et al., 2009; Mansfield et al., 1999). Generally speaking, removing hemicellulose and

lignin; swelling cellulose to destroy crystallinity; hydrolyzing cellulose to shorten chain

length (increasing the number of reducing ends for enzymes to attack); and increasing

DA substrate: 64.1 g

Glucose 33.3 g

DA spent liquor

Glucose 3.0 g

Galactose 0.4 g

Mannose 0.9 g

Arabinose 0.1 g

Xylose 0.2 g

SPORL substrate: 60.5g

Glucose 40.3 g

Mannose 7.1 g

SPORL spent liquor

Glucose 2.9 g

Galactose 1.3 g

Mannose 4.5 g

Arabinose 0.4 g

Xylose 2.2 g

100g oven-dry spruce powder

Glucose 46.7 g

Galactose 2.6 g

Mannose 10.8 g

Arabinose 1.2 g

Xylose 5.5 g

DA SPORL

Page 70: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

55

surface area or decreasing particle size, are favorable for enzymatic digestibility of

cellulosic substrates.

(a)

(b) Figure 2.4 Comparison of time-dependent enzymatic hydrolysability of SPORL and DA

pretreated spruce at different levels of enzyme loading. (a) 15 FPU cellulase+30 IU β-glucosidase

per gram of cellulose; (b) 5 FPU cellulase+10 IU β-glucosidase per gram of cellulose), 50 ºC, pH

4.8 and on a 250 rpm shaker. CGCY: Cellulose-to-Glucose Conversion

0

20

40

60

80

100

0 10 20 30 40 50

CG

CY

(%

)

Enzymatic hydrolysis time (h)

SPORL

Untreated

DA

0

20

40

60

80

100

0 10 20 30 40 50

CG

CY

(%

)

Enzymatic hydrolysis time (h)

SPORL

Untreated

DA

Page 71: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

56

The enzymatic digestibility of SPORL- and DA-pretreated and original spruce

wood was compared in Figure 2.4. At enzyme loadings of 15 FPU (Filter Paper Units)

cellulase and 30 IU (International Units) -glucosidase per gram cellulose (Figure 2.4 a),

the SPORL pretreated spruce displayed much greater hydrolysis than did the DA

pretreated spruce and untreated spruce. For example, only 25% and 55% of the cellulose

in untreated spruce and the DA substrate, respectively, was hydrolyzed to glucose after a

2-day hydrolysis, whereas 93% of the cellulose in SPORL substrates was saccharified

within the same time. When the enzyme loadings were reduced to 5 FPU cellulase and 10

IU -glucosidase per gram cellulose (Figure 2.4 b), the cellulose-to-glucose conversion

after a 2-day hydrolysis of the SPORL substrate (71%) was still substantially higher than

those of the DA substrate and untreated spruce (49 and 17%, respectively). The results

clearly indicated that the SPORL substrate had substantially better enzymatic digestibility

than did the DA substrate under the same hydrolysis conditions. As discussed above, the

DA substrate contained no hemicellulose and had slightly lower viscosity (lower

cellulose degree of polymerization), compared to the SPORL substrate. Based on these

factors alone, the DA substrate might be assumed to have a better digestibility. However,

lignin (both content and nature), and other factors, resulted in quite the opposite situation.

Page 72: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

57

The DA substrate contained approximately 50% lignin, and the lignin was highly

condensed, extremely hydrophobic, and covered the surface of the substrate, all of which

enhanced the effect of lignin as physical barrier and non-productive enzyme adsorbent.

On the other hand, the SPORL substrate contained less lignin, and the possible

sulfonation of lignin in the substrate made the lignin less hydrophobic, which may have

reduced the non-productive hydrophobic adsorption of enzymes onto the lignin. Similar

report has shown that addition of surfactants can significantly decrease the non-

productive adsorption of enzymes on lignin because the hydrophobic end of surfactants

was adsorbed onto hydrophobic lignin and the hydrophilic end of surfactants decreased

the adsorption of enzyme onto lignin surface (Qing et al., 2010). Approximately one

third (~33%, Table 2.1) of the SPORL substrate was lignin, but this lignin retarded the

hydrolysis little. The observation implies that the residual lignin in the SPORL substrate

was enzyme-friendly and behaved differently from the acid-condensed DA lignin. This

suggests that costly delignification is not the only way to remove the recalcitrance

(attributable to lignin) to enzymatic degradation of cellulose; less expensive lignin

modification may be more promising. The results also suggest that the action of lignin as

a purely physical barrier played a less important role in its impacts on enzymes,

Page 73: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

58

compared to other interactions such as non-productive adsorption, which agrees with our

previous results (Marbe et al., 2006).

2.3.4 Fermentability of spent pretreatment liquors

Potential fermentation inhibitors formed during the DA and SPORL pretreatments

are listed in Table 3, including the soluble lignin, acetic acid released from acetyl groups

on hemicelluloses, furfural derived from pentoses, HMF from degradation of hexoses,

and levulinic and formic acids from successive decomposition of HMF. The formation of

the furfurals and their degradation products (levulinic and formic acids) is shown in

Figure 2.5.

Figure 2.5 Inhibitors formation from cellulose and hemicellulose during pretreatment.

Page 74: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

59

The data in Table 2.2 clearly indicated that fewer inhibitors were formed from

degradation of saccharides during the SPORL pretreatment than the DA pretreatment.

Due to the interferences of unknown impurities in spent liquors, the amounts of formic

acid and acetic acid could not be determined accurately. However, this did not affect the

evaluation on these two methods as acetic acid is not derived from saccharide portion of

the carbohydrates and the molar amount of formic acid corresponds equivalently to the

molar amount of levulinic acid. The total of known inhibitors (furfural, 5-hydroxy-

methylfurfural, and formic, acetic and levulinic acids) formed in SPORL were only 35%

of those formed in DA pretreatment. As discussed above, this was owing to the addition

of sulfite in the SPORL pretreatment, which increased the pH value of the pretreatment

liquor, limiting extensive degradation of saccharides. The high lignin concentration in

SPORL liquor is due to the formation of water-soluble lignosulfonate. Low traditional

inhibitor (Table 2.2) and high sugar (Table 2.1) concentrations suggest that good

fermentability can be expected for SPORL liquor.

Page 75: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

60

Figure 2.6 Fermentability of SPORL and DA pretreatment spent liquors by in vitro ruminal

fermentation assay. (DM: dry matter)

The fermentability of SPORL and DA liquors was evaluated by an in vitro

ruminal gas production assay, as shown in Figure 2.6. Surprisingly, the two liquors did

not show difference in net gas yield within the first 24 h, while the gas yield of SPORL

liquor decreased beyond this point. The loss of gas was not clear at this time and will be

further investigated in future work. In tests with various perennial grasses, in vitro ruminal

gas production has been shown to provide a reasonable surrogate estimate of ethanol

production potential by an enzyme/yeast SSF (Simultaneous Saccharification and

Fermentation) system (Weimer et al., 2005). In addition, the in vitro ruminal gas

production assay has displayed less inhibition than SSF for certain forages that contain

0

100

200

300

400

500

0 20 40 60 80 100

Ne

t g

as y

ield

(m

L/g

DM

)

Fermentation time (h)

SPORL

DA

Page 76: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

61

natural fermentation inhibitors (e.g., some varieties of switchgrass). It was thus a surprise

that in vitro ruminal gas production from the SPORL-pretreatment liquor was less than

that of DA-pretreatment liquor that apparently had higher concentrations of classical

pretreatment-derived inhibitors (e.g., furfural and HMF). This inhibition may have been

due to lignosulfonate (present in much higher concentrations in the SPORL liquor).

Alternatively, inhibition could have been due to the presence of other metabolites that

may be released from wood upon pretreatment.

2.4 Conclusion and recommendations

SPORL pretreatment significantly decreased the recalcitrance of spruce wood and

allowed nearly complete enzymatic hydrolysis (>90%) within 24 h with a cellulase

loading of 15 FPU/g cellulose. SPORL reduced the recalcitrance not only by dissolving

hemicellulose and depolymerizing cellulose, but also by partially (32%) dissolving lignin

and sulfonating the residual lignin in substrate. The latter presumably made lignin more

hydrophilic and thereby reduced the hydrophobic interactions between lignin and the

enzymes. SPORL pretreatment achieved a significantly higher sugar recovery and

produced much lower levels of traditional fermentation inhibitors than dilute acid

Page 77: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

62

pretreatment. The overall saccharides (hexoses and pentoses) recovery of SPORL was

87.9%, compared to 56.7% with dilute acid.

Page 78: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

63

Chapter 3: Synthesis of Cellulase Mimetic Solid

Acid for Cellulose Hydrolysis

3.1 Introduction

Energy-efficiently and cost-effectively hydrolyzing cellulose to glucose is the

bottleneck to converting lignocelluloses to value-added chemicals and liquid fuels

through the sugar platform (Ragauskas et al., 2006). Acids and enzymes are the most

common catalysts for hydrolyzing cellulose into glucose. Sulfuric acid is an inexpensive

and effective acid catalyst. For example, concentrated sulfuric acid (>65%) can

effectively swell and hydrolyze crystalline cellulose into glucose at moderate temperature

through disrupting strong hydrogen bonds in crystalline cellulose, but acid recovery is

difficult. Dilute acid (0.5-10%) is also able to hydrolyze cellulose, but needs higher

temperature (150-220 °C). The drawback of high-temperature acid hydrolysis is the

undesirable dehydration of glucose to hydroxymethylfurfural (HMF) and further

degradation to levulinic and formic acids, which are fermentation inhibitors. These side

reactions not only reduce glucose yield but also generate fermentation inhibitors. Other

issues with sulfuric acid hydrolysis include process effluent treatment and gypsum

Page 79: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

64

handling from acid neutralization (Bienkowski et al., 1984; Harris and Lang, 1947; Li et

al., 2008; Taherzadeh and Karimi, 2007; Taherzadeh and Karimi, 2008). Cellulase is

highly selective and able to hydrolyze cellulose at lower temperature (~50 °C). However,

enzymatic hydrolysis of cellulose is a slow and incomplete reaction, typically taking days

to achieve satisfactory yields. In addition, prior to enzymatic hydrolysis of

lignocelluloses, an energy-intensive pretreatment is necessary to remove the recalcitrance

to enzymes. Finally, cellulases are still expensive enzymes today, making it unaffordable

to pursue high hydrolysis yield by increasing enzyme loading (Chandra et al., 2007;

Huang, 1975; Medve et al., 1998).

The difference in the temperature required for acidic and enzymatic hydrolysis of

cellulose is related to their reaction activation energies. It was reported that the activation

energies for cellulose hydrolysis and glucose dehydration with dilute acid ranged

between 170-180 and 135-145 kJ·mol-1

, respectively, no matter what temperature and

acid concentration were used (Girisuta et al., 2006; Heeres et al., 2007). On the other

hand, the activation energy of enzymatic hydrolysis was only 3-50 kJ·mol-1

(Bravo et al.,

2000; Hsuanyu and Laidler, 1985), which is the reason why enzymatic hydrolysis of

cellulose can be conducted at ~50 °C (Schall et al., 2006; Shuai et al., 2010). However,

Page 80: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

65

enzymatic hydrolysis of cellulose is slow and may take days to complete. Because of the

thermal instability of cellulases, the hydrolysis rate cannot be accelerated through

increasing temperature because cellulase will denature at high temperature. An ideal

catalyst for cellulose hydrolysis should lower the activation energy of cellulose

hydrolysis, which will lower hydrolysis temperature, shorten reaction time, and reduce

the undesirable glucose degradation at elevated temperature.

Solid acid catalysts for cellulose hydrolysis have recently drawn a lot of attention

because, on the one hand, the solid acid is less expensive and easier to recover/reuse than

the cellulases, and on the other hand, the solid acid lowers the activation energy of

cellulose hydrolysis thus allows lower hydrolysis temperature than homogenous acids. In

addition, because of the thermal stability of the solid acid, increasing temperature could

be an option to speed up the reaction rate. Sulfuric acid-carbonized biomass is one of the

mostly studied solid acids for hydrolyzing cellulose (Suganuma et al., 2008; Yamaguchi

et al., 2009). Cellulose hydrolysis catalyzed by this carbonized-biomass solid acid had

lower activation energy (110 kJ·mol-1

vs. 170 kJ·mol-1

) than the one catalyzed by sulfuric

acid. Other solid acids studied include zeolite (Zhang and Zhao, 2009), silica/carbon

nanocomposites (Van de Vyver et al., 2011), sulfonated carbon materials (graphene (Hara

Page 81: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

66

et al., 2009) and CMK-3 (Kobayashi et al., 2010; Pang et al., 2010)), and layered niobium

molybdate (HNbMoO6) (Takagaki et al., 2008; Takagaki et al., 2010). Recently,

magnetic solid catalysts bearing –SO3H were synthesized for hydrolyzing cellulose. The

uniqueness of the catalysts is that they can be magnetically recovered from the reaction

system (Bond et al., 2010; Fu et al., 2011; He et al., 2010; Lai et al., 2011). The

performance of the solid acid catalysts above for the hydrolysis of cellulose to glucose

varied, with the overall glucose yield of 30-75% (Pang et al., 2010; Van de Vyver et al.,

2011). To achieve a reasonable glucose yield, high temperature (>150 °C) is necessary.

Another issue of the solid acids is the need for very high catalyst/substrate mass ratio

(Van de Vyver et al., 2011). One of the reasons is that the current solid acids, unlike

cellulases, do not have dedicated substrate-binding sites for associating cellulose onto

catalyst surface.

Structurally, most cellulases have two domains, a catalytic domain and a cellulose

binding domain linked by a peptide (Boraston et al., 2003; Coutinho et al., 1993; Gilkes

et al., 1992; Gilkes et al., 1993). The binding domain is responsible for associating the

cellulases to cellulose, while the catalytic domain is to hydrolyze glycosidic linkages of

cellulose chain. In cellulases, acid groups such as carboxylic and phenolic hydroxyl

Page 82: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

67

groups of amino acids in the catalytic domain function as proton sources for cleaving

glycosidic bonds (Berti and Tanaka, 2002; Mccarter and Withers, 1994). It is believed

that enzyme-cellulose binding is via ion attraction, hydrophobic affiliation, or hydrogen

bonding between cellulose and aromatic amino acids, such as tryptophan and tyrosine, on

the enzyme (Creagh et al., 1996; Tomme et al., 1996). Mosier et al. (Mosier et al., 2004)

screened potential cellulose binding compounds for designing mimetic catalysts and

found that aromatic, hydrophobic, and planar structures, such as indole and tryphan blue,

have high affinity to cellulose. They also pointed out that hydrophobic interactions and

hydrogen bonding are the most important two contributors to the enzyme adsorption onto

cellulose.

Chemically, it is possible to synthesize a cellulase-mimetic catalyst containing

both catalytic and cellulose-binding domains through molecular design, as shown in

Figure 3.1. For example, acidic groups, such as sulfonic acid (-SO3H), can be introduced

onto the surface of the catalyst and function as catalytic domains. As hydroxyl groups on

cellulose form hydrogen bonds with electronegative groups such as hydroxyl, amine, and

halide, the electronegative groups anchored on the surface of synthesized catalyst can be

expected to function as cellulose binding domains. In addition, the formation of hydrogen

Page 83: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

68

bonding between the electronegative groups of catalyst and hydroxyl group proton helps

to disrupt the intra- and inter-molecular hydrogen bonds of cellulose and thereby

enhances the swelling of cellulose (Dawsey and Mccormick, 1990; Turbak et al., 1977).

Figure 3.1 A proposed model of cellulase-mimetic solid acid and cellulose interaction.

So far, most of the studies on solid acids for cellulose hydrolysis have focused

just on the catalytic activity of the catalysts (Hara et al., 2009; Onda et al., 2008;

Suganuma et al., 2008). Little attention has been paid to the interaction/binding between

the catalyst and cellulose and, to our knowledge no research has been reported on

introducing cellulose-binding sites onto synthesized solid acids. In this paper, we report a

simple method to synthesize a solid acid catalyst containing both cellulose binding and

hydrolyzing sites to mimic cellulase for cellulose hydrolysis.

Cl ClClClClCl ClClClCl SO3H

O

H

HO

H

HO

H

HOHH

O

OH

O

H

H

HO

H

HOHH

O

OH

O

H

H

HO

H

HOHH

O

OH

O

H

H

HO

H

HOHH

OH

OH

Cellulose chain

Mimetic cellulase

Binding domain Catalytic domain

Hydrogen bond

Binding domain

Page 84: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

69

3.2 Experimental

3.2.1 Chemicals and materials

Chemical reagents used in this study were purchased from Fisher Scientific

(Pittsburgh, PA) or Sigma-Aldrich (St. Louis, MO) and used as received. Chloromethyl-

polystyrene (CP) (100 mesh) with 3.5-4.5 mmol chlorine per gram resin and Amberlyst-

15 (100 mesh) with 4.5 mmol -SO3H per gram resin were dried at 105 ºC overnight prior

to use.

3.2.2 Synthesis of cellulase mimetic solid acid

(1) Synthesis of solid acid with –Cl binding domain (CP-SO3H)

The synthesis of CP-SO3H is schematically shown in Figure 3.2. The oven-dried

CP resin (3 g) and DMF (50 mL) were added into a 100-mL flask and stirred at 120 ºC

for 30 min in an oil bath. Then 2 g of sulfanilic acid was added into the flask and reacted

with the resin at 120 ºC for 48 h. After the reaction, the resin was recovered by filtration

(10 μm retention) followed by washing with 100 mL ethanol and 200 mL water

sequentially. The resin prepared was in the form of a green powder after being dried at

105 ºC overnight.

Page 85: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

70

Figure 3.2 Synthesis of CP-SO3H.

(2) Synthesis of solid acid with –NH2 binding domain (NP-SO3H)

Prepared CP-SO3H (2 g) was suspended in 10 mL water and reacted with 1 mL 30%

NH3·H2O at 100 ºC for 2 h. The resulting resin was filtered and acidified with 200 mL of

5% HCl and then washed with deionized water until neutral pH was achieved.

(3) Synthesis of solid acid with –OH binding domain (OP-SO3H)

Prepared CP-SO3H (2 g) was suspended in 10 mL water and reacted with 1 mL of

50% NaOH at 100 ºC for 2 h. The resulting resin was filtered and acidified with 200 mL

of 5% HCl and then washed with deionized water until a neutral pH was achieved.

(4) Synthesis of solid acid without binding domain (HP-SO3H)

Prepared CP-SO3H (2 g) was suspended in 10 mL toluene and reacted with 1 g

NaH at room temperature for 24 h. The resulting resin was filtered and acidified with 200

mL of 5% HCl and then washed with deionized water until a neutral pH was achieved.

Page 86: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

71

3.2.3 Hydrolysis of biomass with solid acids

Figure 3.3 A model for cellulose hydrolysis on solid acid (CP-SO3H) surface.

The proposed mechanism of SO3H-catalyzed hydrolysis of the glycosidic bond is

shown in Figure 3.3. A predetermined amount of solid acid and sulfuric acid were mixed

with 100 mg substrate (cellobiose, starch, or Avicel) into 1 mL water in 20-mL glass

vials and mixed well. Vials were heated in oil bath at different temperatures for varied

durations (reaction conditions list in the notes under Figures or Tables). Supernatant

samples were taken to determine the glucose yield.

Page 87: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

72

3.2.4 Adsorption of glucose and cellobiose on CP-SO3H

CP-SO3H (600 mg) was added into 12 mL water and vortexed to mix well. The

mixture was divided into 12 equal portions in microtubes. Then 100 mg glucose or 90 mg

cellobiose was loaded into each microtube and mixed well. All microtubes were kept at

room temperature and vortexed every 5 min. One microtube for each was taken and

centrifuged at a time of 10, 20, 30, 60, 90, and 120 min. Glucose and cellobiose in the

supernatant were determined with High Performance Ion Chromatography, as described

below, to calculate the adsorption capacity of CP-SO3H.

3.2.5 Determination of glucose

Procedure is described in section 2.2.6.

3.2.6 FT-IR spectra of prepared solids

Fourier transform infrared (FT-IR) spectra were recorded on a PerkinElmer

Spectrum 100 FT-IR spectrophotometer with a universal attenuated-total-reflection (ATR)

sampling accessory (Waltham, MA). ATR allows samples to be examined directly in the

solid state without further preparation.

Page 88: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

73

3.3 Results and discussion

3.3.1 Screening of binding groups

Four types of solid acids with different binding domains were synthesized and

tested for their catalyst activities. The proposed mechanism of cellulose hydrolysis

catalyzed by CP-SO3H is schematically shown in Figure 3.3. The sulfonation of CP was

verified with FT-IR. The spectra of CP and CP-SO3H resins are shown in Figure 3.4.

Wavenumbers (cm-1)

Figure 3.4 FT-IR spectra of (a) CP resin and (b) CP-SO3H resin.

The peak at 1260 cm-1

results from the adsorption of the -CH2Cl groups (Zhang et

al., 2011), while those at 1150 and 1100 cm-1

are from the vibration and stretching of

a

b

-CH2Cl

-SO3H

3

1

Page 89: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

74

sulfonic acid group (-SO3H) (Panicker et al., 2006). It is apparent that the intensity of the

-CH2Cl band decreased after the sulfonation, indicating the substitution of -Cl via

sulfanilic acid. The increased intensities at 1150 and 1100 cm-1

further verified the

introduction of the -SO3H group. The content of the acid group (-SO3H) introduced was

approximately 0.0067 mmol per gram of the resin, as determined by electro-conductivity

titration. Considering the fact that the sulfonation reaction occurred primarily on the resin

surface, the surface concentration of -SO3H would be higher than the value above

calculated based on total resin. Surface concentration of -SO3H was further estimated

from the spectra of ATR-FTIR. As shown in Figure 3.4, it was estimated one third of

surface chloride groups were substituted with -SO3H. Further study is needed to the

distribution of -SO3H groups inside and on the surface of the resin. The FT-IR spectra of

NP-SO3H, OP-SO3H and HP-SO3H in Figure 3.5 show that the chloride group was

successfully replaced. The peak representing the -CH2Cl band significantly decreased or

almost disappeared, indicating that solid acid with different binding groups were

synthesized successfully. It was hard to quantitatively and accurately estimate the degree

of substitution as the styrene units only accounted for very small part of the total styrene

Page 90: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

75

units. However, as –Cl was easily substituted by alkalis, the substitutions of –Cl by OH-

and NH3 on resin surface were assumed to be complete.

Figure 3.5 FT-IR spectra of solid acids with different binding domains. (a) CP resin, (b) NH2

resin, (c) OH resin, and (d) H resin.

Then effect of binding groups on the hydrolysis of cellulose was tested at

temperatures of 100 and 120 ºC. It can be seen from Figure 3.6 that at both temperatures,

the hydrolysis rate of cellobiose followed the order of Cl>NH2>OH>H. According to

the Arrhenius equation, the apparent activation energies of cellobiose hydrolysis

catalyzed by the four types of solid acids were estimated, as shown in Table 3.1.

Page 91: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

76

(a)

(b)

Figure 3.6 Cellobiose hydrolysis catalyzed by four types of solid acids as a function of time.

0

10

20

30

40

50

60

70

80

0 100 200 300 400 500 600

Co

nve

rsio

n(%

)

Reaction time (min)

Cl

NH2

OH

H

0

10

20

30

40

50

60

0 20 40 60 80 100 120

Co

nve

rsio

n(%

)

Reaction time (min)

Cl

NH2

OH

H

Page 92: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

77

(a) 100 ºC, (b) 120 ºC.

It can be seen that -Cl lowered apparent activation energy of cellobiose

hydrolysis and therefore was chosen for further investigation for its ability to hydrolyze

cellulose. The reason why the chloride group (-Cl) showed the lowest activation energy

in cellobiose hydrolysis was presumably that chlorine groups could form stronger

hydrogen bonds with hydroxyl groups of cellulose but also disrupt intra- and inter-

hydrogen bonds of cellulose, which would enhance the swelling and dissolution of

cellulose. This is consistent with the wide involvement of chloride in cellulose solvents,

such as DMAC/LiCl and chloride-containing ionic liquids (Kosan et al., 2008;

Mccormick et al., 1985; Remsing et al., 2006). The evidence that the polymers of vinyl

chloride and vinylidene chloride had higher adhesion to cellulose than those of ethylene,

propylene and vinyl acetate (McLaren, 1948) supports the hypothesis that chloride has

cellulose-binding capacity.

Table 3.1 The apparent activation energies of cellobiose hydrolysis catalyzed by four types of

solid acids

Binding group Cl H OH NH2

Apparent activation energy (kJmol-1

) 77.7 88.5 89.5 104.7

Page 93: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

78

3.3.2 Mechanism study

When cellobiose is hydrolyzed with a homogeneous acid, such as H2SO4, the

reaction follows collision theory, i.e. the reaction rate depends on the concentration of

both the acid and cellobiose. Lowering acid or cellulose concentration will slow the

reaction rate (Bobletera et al., 1986; Lee et al., 2003). However, the situation is different

when solid acid is used as catalyst. The reaction rate is not only dependent on the

concentrations of cellobiose and acid sites/groups on the solid acids, but also on the

collision probability between cellobiose and the acid sites. Therefore, if cellobiose can be

selectively adsorbed (concentrated) onto the catalyst surface, the hydrolysis rate would be

enhanced. Meanwhile, the resulting glucose molecules need to be desorbed (released)

from the catalyst surface as soon as they are generated to free the binding sites on the

catalyst surface for new cellobiose molecules. It is hypothesized that polysaccharide and

oligosaccharide molecules would be more preferably adsorbed onto the CP-SO3H surface

compared to monosaccharides because the former has more binding locations (hydroxyl

groups) than the latter.

Page 94: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

79

In order to verify the above hypothesis, the adsorption of glucose and cellobiose

onto the CP-SO3H surface was studied in aqueous solution at room temperature. As

shown in Figure 3.7, CP-SO3H adsorbed both glucose and cellobiose, which is different

from the solid acid from carbonized biomass that only adsorbed cellobiose but not

glucose (Creagh et al., 1996; Hara et al., 2009; Suganuma et al., 2008). Similarly,

sulfonated activated carbon was able to adsorb oligosaccharides but not monosaccharides,

and the adsorption was dependent on the size of the oligosaccharides (Hara et al., 2009).

It was proposed that the two solid acids above adsorbed the sugars through the interaction

(hydrogen binding) between phenolic hydroxyl groups on the activated carbon and the

oxygen in β-1,4-glycosidic bond, which explained why carbonized biomass only

adsorbed cellobiose but not glucose, and the activated carbon had greater affinity for

longer oligosaccharides. However, this mechanism is unable to explain the adsorption of

glucose on CP-SO3H since there is no glycosidic oxygen in glucose. We believe that CP-

SO3H adsorbs the sugars through the hydrogen bonding between the chloride groups on

the catalyst and the hydroxyl groups of the sugars, as will discussed further below.

When the same amount of glucose units (glucose/cellobiose, 2:1, mol/mol) was

loaded, cellobiose was adsorbed more onto the catalyst surface than glucose (Figure 3.7),

Page 95: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

80

suggesting that CP-SO3H has a higher affinity for cellobiose than glucose. This is

presumably because more hydrogen bonds are formed between the catalyst and

cellobiose due to the higher number of hydroxyl groups available in cellobiose. In order

to further verify the higher affinity of cellobiose than glucose to CP-SO3H, the

competitive adsorption of cellobiose and glucose on CP-SO3H was investigated. It was

observed (data not shown here) that when cellobiose and glucose were mixed with CP-

SO3H at equivalent loadings of glucose units, cellobiose was selectively absorbed by CP-

SO3H. Certainly, the preferable adsorption of cellobiose over glucose is desirable and

critical for cellobiose hydrolysis to proceed, which ensures the selective adsorption of

cellobiose to the catalyst and prompt desorption of glucose from the catalyst.

85

90

95

100

105

0 20 40 60 80 100 120 140

Re

sid

ua

l sa

ccha

ride

s(%

)

Time (min)

a

d

c

b

Page 96: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

81

Figure 3.7 Timecourse for adsorption of glucose and cellobiose onto resins in aqueous solution.

(a) glucose on PS-SO3H; (b) cellobiose on PS-SO3H; (c) glucose on CP-SO3H; (d) cellobiose on

CP-SO3H.

To demonstrate the importance of the -Cl groups on CP-SO3H in adsorbing

saccharides through hydrogen bonding, the adsorption of glucose and cellobiose onto

sulfonated polystyrene resin (Amberlyst-15, abbreviated as PS-SO3H) was studied. As

expected, PS-SO3H did not adsorb glucose or cellobiose at all because it did not have the

sites (-Cl) for binding the sugars, as indicated in Figure 3.7. This is presumably the

reason why solid Bronsted acids, such as inorganic oxide solid acids, H-mordenite, and

SO3H-bearing polymers like Nafion NR50 and Amberlyst-15 are unable to hydrolyze

cellulose because they do not have the binding capacity to associate the substrate onto the

catalyst surface (Hara et al., 2009; Suganuma et al., 2008).

Page 97: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

82

Figure 3.8 Comparison of cellobiose hydrolysis catalyzed by (a) CP-SO3H and (b) PS-SO3H.

(note: 500 mg resins, 100 mg cellobiose, 100 ºC, 1 mL water, 2 h).

The catalytic capacity of CP-SO3H and PS-SO3H was compared for hydrolyzing

cellobiose at the same catalyst loading of 500 mg resin per 100 mg cellobiose. As shown

in Figure 3.8, the two catalysts had comparable glucose yield. More precisely, PS-SO3H

had a slightly higher glucose yield than CP-SO3H. As every styrene unit contains a –

SO3H group in PS-SO3H, PS-SO3H has much higher acid site density on the resin surface

than CP-SO3H. With the similar glucose yields the hydrolysis of cellobiose, CP-SO3H

should have much greater specific catalytic capacity than PS-SO3H, which is probably a

result of the superior cellobiose-binding ability due to -Cl groups on CP-SO3H.

3.3.3 Hydrolysis of cellulose with CP-SO3H

The catalytic capacity of CP-SO3H was first evaluated and compared with sulfuric

acid using a soluble substrate cellobiose. As shown in Figure 3.9, at an equivalent acid

loading of 0.0017 mmol (0.17 mg H2SO4 or 250 mg CP-SO3H) per 100 mg cellobiose,

approximately 73% cellobiose was hydrolyzed into glucose with CP-SO3H in 2 h at 120 ºC,

while only 4% was hydrolyzed with sulfuric acid under the same conditions. When the

acid loading was doubled to 0.0033 mmol (0.34 mg H2SO4 or 500 mg CP-SO3H),

Page 98: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

83

cellobiose was completely hydrolyzed with CP-SO3H within 2 h, whereas only 8% was

hydrolyzed by sulfuric acid.

Figure 3.9 Hydrolysis of cellobiose catalyzed by CP-SO3H and sulfuric acid. (a) 0.0033 mmol

CP-SO3H, 120 °C; (b) 0.0017 mmol CP-SO3H, 120 °C; (c) 0.0033 mmol H2SO4, 120 °C; (d)

0.0017 mmol H2SO4, 120 °C.

Table 3.2 Hydrolysis of starch and Avicel cellulose catalyzed by CP-SO3H and sulfuric acid

Entry Feedstock Temp. (°C) Time (h) Yield (%)

Sulfuric acid CP-SO3H

1 Cellobiose 100 2 NA 62

2 Cellobiose 100 4 NA 99

3 Cellobiose 120 2 7.9 100

4 Starch 120 2 3 100

5 Avicel cellulose 120 2 0.5 32

6 Avicel cellulose 120 10 NA 93

7 Starch 120 20 NA 25

8 Avicel cellulose 120 20 NA 27

9 Cellobiose 140 0.5 7 98

10 Cellobiose 160 0.5 15 NA

Note: ―NA‖refers to ― not applicable‖; entry 1 to 8 (100 mg carbohydrate; 1 mL water; 0.0033 mmol

catalyst); entry 9 to 10 (100 mg carbohydrate; 1 mL water; 0.0017 mmol catalyst)

Page 99: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

84

The performance of CP-SO3H at different temperatures was also investigated. It

worked well at 100 ºC, though the yield of cellobiose hydrolysis was lowered to 62.1%

after a 2-h reaction. If the reaction was extended to 4 h, complete hydrolysis of cellobiose

(99.3%) was achieved. Low temperature slowed down the reaction, but had the advantage

of avoiding some further degradation of glucose. When temperature was elevated to 140

ºC, the time for complete hydrolysis was shortened to 30 min (as presented in Table 3.2).

These results suggest that CP-SO3H performed better than sulfuric acid in catalyzing

cellobiose hydrolysis. This is presumably attributed to the capacity of adsorbing

cellobiose and high surface acid concentration, as will discussed below.

The durability of CP-SO3H was examined by recycling the catalyst three times in

cellobiose hydrolysis. The catalyst was recovered by filtration after the hydrolysis,

washed, dried at 105 ºC overnight, and reused in the next batch of hydrolysis reactions.

As shown in Figure 3.10, the catalyst activity did not decline after recycling three times,

indicating that CP-SO3H resin had good stability in catalytic activity. The chemically

bonded -Cl and -SO3H on CP-SO3H did not have the leaching problem that happened to

the carbon based-solid acid (Van de Vyver et al., 2010).

Page 100: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

85

Figure 3.10 Cellobiose hydrolysis catalyzed by recycled CP-SO3H resin. (Note: 100 mg

cellobiose, 120 ºC, 1 mL water, 2 h).

The catalytic activity of CP-SO3H in hydrolyzing polysaccharides was also

investigated. As shown in Table 3.2, starch and crystalline cellulose (Avicel) were

hydrolyzed with CP-SO3H and sulfuric acid at an equivalent acid loading. It was found

that starch was completely hydrolyzed to glucose by CP-SO3H in 2 h (entry 4), while

only 32% of Avicel cellulose was hydrolyzed into glucose under the same conditions

(entry 5). The crystalline structure of cellulose is considered as the major reason for the

slow hydrolysis and low yield of cellulose hydrolysis to glucose. When hydrolysis was

extended to 10 h, 93% of Avicel cellulose could be hydrolyzed. By contrast, with sulfuric

Page 101: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

86

acid at an equivalent acid loading at the same temperature and reaction time, starch was

only 3% hydrolyzed (entry 4), and almost no Avicel cellulose was hydrolyzed (entry 5).

This is because cellulose hydrolysis catalyzed by homogeneous acid hydrolysis has a

higher activation energy than the one catalyzed by CP-SO3H. The capability of CP-SO3H

to hydrolyze Avicel at moderate temperature can be attributed to the presence of -Cl. The

chloride groups, in addition to binding cellulose, are possibly able to decrystallize

cellulose by breaking intra- and inter-molecular hydrogen bonds between the hydroxyl

groups of cellulose through forming stronger hydrogen bonds with the hydroxyl groups.

The mechanism needs further investigation, and is outside the scope of this thesis.

To further understand why CP-SO3H can hydrolyze cellobiose and cellulose at a

decent rate at moderate temperature, the activation energies of cellobiose and cellulose

hydrolysis catalyzed by CP-SO3H were estimated. The activation energy of cellobiose

hydrolysis on two types of synthesized resins was calculated from the Arrhenius equation:

(k = e−Ea RT ). In order to estimate the apparent activation energy, the reactions were

assumed to be pseudo-first order reaction. The rate constant was obtained from equation:

k=ln(2)/t1/2 (t1/2 is half-life time). Half-life time at different temperatures was recorded

and plotted, as shown in Figure 3.11.

Page 102: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

87

Figure 3.11 Arrhenius plot for cellulose hydrolysis catalyzed by CP-SO3H.

y = -9.969x + 19.436R² = 0.9995

-7.5

-7

-6.5

-6

-5.5

-5

-4.5

2.4 2.45 2.5 2.55 2.6 2.65 2.7

lnk

1/T (10-3 K-1)

Ea=9.4142×8.31=78 kJ mol-1

Ea=9.969×8.31=83 kJ mol-1

Cellobiose

Crystalline cellulose

Page 103: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

88

As shown in Table 3.3, the apparent activation energies for hydrolysis of

cellobiose and Avicel are 78 and 83 kJmol-1

, respectively, at 373-413 K. The similar

activation energies of the two reactions indicate that they can proceed at the same

temperature, which has been verified by the results above. Although hydrolysis of

cellobiose and Avicel by CP-SO3H had a similar activation energy as mentioned above,

the hydrolysis of Avicel was much slower than that of cellobiose under the same

conditions, as shown in Table 3.2 (entries 3 and 5). This is attributed to the solid state of

Avicel, which has fewer collision chances with the catalyst, compared to soluble

cellobiose. The values of the activation energy are significantly smaller than those of the

hydrolysis reactions catalyzed by sulfuric acid (170 kJmol-1

for crystalline cellulose and

130 kJmol-1

for cellobiose) (Bobletera et al., 1986) and sulfonated active carbon (AC-

SO3H) (110 kJmol-1

) (Suganu ma et al., 2008). This is the reason why CP-SO3H was

able to hydrolyze cellobiose and cellulose at a lower temperature, which will minimize

energy consumption and undesirable sugar degradation. Compared with cellulase (3-50

kJmol-1

, depending on source of the enzyme) (Paljevac et al., 2007; Xiao et al., 2002),

CP-SO3H had a higher activation energy and therefore needed higher reaction

temperatures. On the other hand, the high temperature allowed the hydrolysis to be

Page 104: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

89

completed in a short time, as discussed previously. The lower activation energy of

hydrolysis catalyzed by CP-SO3H was attributed to the ability of the solid acid to adsorb

cellobiose or cellulose, which lowers the energy barrier for the hydrolysis to proceed.

Table 3.3 Apparent activation energy of cellulose hydrolysis catalyzed by different catalysts

Low activation energy and cellobiose/cellulose binding ability give CP-SO3H

high specific catalyst activity. It was reported that sulfonated active carbon (AC-SO3H)

(110 kJmol-1

) could hydrolyze cellobiose by 12% in 2 h with an acid loading of 54%

(molar/molar, based on cellobiose) (Hara et al., 2010), whereas CP-SO3H achieved a

hydrolysis yield of 62% in 2 h with an acid loading of only 0.59% (molar/molar, based on

cellobiose). It is noteworthy that although CP-SO3H performed well in catalyzing the

hydrolysis of cellulose, the density of acid sites on CP-SO3H was relatively low (0.0033

mmol/g). If the density of acid sites could be increased, the cellulose hydrolysis rate

would be further enhanced from the current level.

Catalyst Apparent activation energy (kJ·mol

-1)

Cellobiose Crystalline cellulose

Sulfuric acid 133 170

Carbon-SO3H / 110

CP-SO3H 78 83

Cellulase 3-50 3-50

Page 105: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

90

3.4 Conclusion and recommendations

A cellulase-mimetic solid acid catalyst was synthesized by sulfonating

commercial chloromethyl polystyrene resin. Chloride and sulfonic groups on the

resulting solid acid served as cellulose-binding domains and catalytic domains,

respectively. The catalyst was able to hydrolyze cellobiose, starch, and crystalline

cellulose. Because of its substrate-adsorbing ability, the synthesized resin showed

significantly higher catalytic activity than the homogeneous sulfuric acid or other solid

acids at equivalent acid loadings. The low activation energy of the CP-SO3H-catalyzed

reaction allowed the hydrolysis to proceed at moderate temperature. Preliminary results

indicated that the solid acid catalyst had good stability and could be recycled/reused

without activity loss. Future study is needed to increase the density of catalytic domain (-

SO3H) to reduce the catalyst loading. It is also desirable to develop a strategy to

separate/recover the catalyst from hydrolysis residue, in particular when applied to real

biomass where the presence of lignin would make the catalyst recovery more difficult.

For example, incorporating magnetic material in the solid acid catalyst could make the

catalyst magnetically recoverable.

Page 106: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

91

Chapter 4: Saccharification of Lignocellulose in

Concentrated Salt Solution

4.1 Introduction

Currently bioethanol is produced from cornstarch or sugarcane. These processes

are not sustainable and are unable to meet the increasing demand for renewable fuels.

Sustainable production of biofuel needs to rely on abundant, inexpensive, and non-food

lignocelluloses. A core bottleneck of the conversion processes based on the sugar

platform is the effective release of sugars from lignocelluloses at low cost and low energy

input (Smith, 2008). Cellulose in lignocelluloses is wrapped by hemicelluloses and

especially lignin, making cellulose far more difficult to hydrolyze to glucose than starch.

In addition cellulose has a crystalline structure. As a consequence, relatively harsh

conditions such as high temperature and high chemical loadings are needed for

hydrolyzing cellulose (Demirbas, 2005). The primary methods that have been extensively

investigated for saccharification of lignocelluloses include concentrated acid, dilute acid,

ionic liquids, and enzymatic processes.

Page 107: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

92

Concentrated acid saccharification is an extensively studied cellulose hydrolysis

process. This process is conducted at relatively mild temperature and can lead to a nearly

theoretical yield of sugars. In this process, cellulose in lignocelluloses is first swollen at

room temperature with concentrated acid (typically sulfuric acid), and then the swelled

cellulose is hydrolyzed in dilute acid at elevated temperature (50-120 C) (Miller and

Hester, 2007; Zhu et al., 2009). However, acid corrosion of equipment and the difficulty

of recycling concentrated sulfuric acid have restricted the development of this technology.

Although ion-exclusion chromatography can be used to separate sugars and sulfuric acid,

the method is costly and energy-intensive. In addition, the acid is extensively diluted

during the sugar-acid separation, and of the recovered sulfuric acid has to be

reconcentrated to 70-80% prior to reuse (Cuzens, 1998; Russo, 1999).

In order to avoid the use of concentrated acid, a saccharification method using

dilute acid at higher temperatures (160-190 ºC) was developed. Unfortunately, the dilute

acid process gives a sugar yield of only about 50% due to the incomplete hydrolysis of

cellulose and sugar degradation at high temperature. Additionally, the sugar degradation

products, such as furfural, hydroxymethylfurfural (HMF), and levulinic acid, can inhibit

the downstream fermentation of the sugars, for example, to produce ethanol. In order to

Page 108: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

93

reduce the degradation of sugars, in particular of pentoses, a two-stage process was

developed. In the first stage, hemicelluloses were first extracted at moderate temperature;

and in the second stage, the temperature was elevated to hydrolyze the cellulose to

glucose. Even so, the total yield of sugars was only 60-70%, depending on feedstock and

processing conditions (Harris, 1985; Zerbe et al., 1988; Zhu et al., 2009). Further, the

need to use a two-stage process adds complexity and cost.

In summary, problems encountered with acid processes include low sugar yield

due to the incomplete hydrolysis of cellulose and undesirable degradation of the sugars,

formation of fermentation inhibitors (furfural, HMF, and levulinic acid, etc.), extensively

condensed lignin (which limits the coproducts production potential of the lignin),

equipment corrosion, acid recovery issues, and wastewater treatment concerns.

The enzymatic saccharification of lignocellulose using cellulose and

hemicellulose hydrolytic enzymes is another popular method used to break down

cellulose and hemicelluloses into monosaccharides. Enzymatic saccharification itself is

inexpensive and less hazardous than acid hydrolysis because of the use of mild process

conditions (~50 ºC and pH 4-5). However, enzymatic saccharification of lignocelluloses

is economically less attractive which limits its commercialization. A major obstacle is the

Page 109: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

94

unavailability of high-activity and low-cost enzymes (both cellulases and hemicellulases).

Although significant progress has been made in recent decades in improving enzyme

activity and reducing enzyme production cost, enzyme is still a considerable contributor

to the high cost of the sugars from lignocellulose platform (Sukumaran et al., 2005;

Zhang et al., 2006). Additionally, because of the natural recalcitrance of lignocelluloses

to the enzymes, enzymatic saccharification of untreated raw biomass is very difficult and

very slow. In order to achieve a satisfactory level of cellulose hydrolysis, an energy-

intensive pretreatment operation is required. Such pretreatment functions to remove

lignin and/or hemicelluloses thereby exposing cellulose. Pretreatment can result in

destruction of the physical matrix by mechanically grinding or milling to reduce particle

size (and thereby increasing accessible surface area to enzymes), enhancing cellulose

hydrolysis by decrystallization and depolymerization, or combinations thereof.

Representative pretreatment technologies include, for example, acid treatment (e.g., with

dilute acid, concentrated phosphoric acid, etc.); the organosolv process (e.g., U.S. patent

3,585,104), ammonia fiber expansion (AFEX), treatment with ionic liquids, treatment

with alkali, and sulfite processes (Balan et al., 2009; Dadi et al., 2006; McIntosh and

Vancov, 2010; Mosier et al., 2005; Pan et al., 2007; Sathitsuksanoh et al., 2010; Wyman

Page 110: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

95

et al., 2005; Yang and Wyman, 2008; Yang et al., 2009). However, due to technical

and/or economic barriers, none of these technologies has yet commercially succeeded. In

addition, unlike chemical reactions, enzymatic hydrolysis is a time-consuming process

and typically takes days to complete. Finally, as high consistency (substrate solid content)

hydrolysis is an engineering challenge, enzymatic hydrolysis typically generates a dilute

sugar stream (5-10%, w/w).

Recently, direct hydrolysis of lignocelluloses in ionic liquids has been reported

from pure cellulose to real biomass, such as untreated corn stover, wheat straws, and

wood powder (Ahring and Langvad, 2008; Binder and Raines, 2010; Li and Zhao, 2007).

The use of ionic liquids can be problematic due to the generally higher cost of these

materials, the poor cellulose dissolution ability of ionic liquids in the presence of water,

and the complexity that can be encountered in separation of the ionic liquids from sugars

and the recycling of ionic liquids.

Hydrolysis of cellulose in salt solutions were also reported. U.S. patent 4,018,620

(Penque) relates to a method of hydrolyzing cellulose to monosaccharides by treating

cellulose with 55% calcium chloride in the presence of acid to hydrolyze newsprint

(newspaper). An overall saccharification yield of 50% was reported, but cellulose was

Page 111: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

96

only 20% hydrolyzed (Penque, 1977). Because of its capability of swelling and

dissolving cellulose, ZnCl2 was widely used in cellulose solvent systems (Cao et al., 1994;

Fischer et al., 2003). A two-step process was reported to swell cellulose at high ZnCl2

concentration followed by hydrolyzing cellulose to glucose in dilute ZnCl2 solution in the

presence of acid (Chen, 1984). It was reported that over 90% of pure cellulose could be

saccharified to glucose with the process. However, the process was less effective when

applied to real lignocelluloses where an overall saccharification yield of polysaccharides

(cellulose and hemicellulose) was 60-70%, but that of cellulose was only 30-50%.

U.S. patents 4,713,118 and 4,787,939 relate to a process for modification,

solubilization and/or hydrolysis of a glycosidically linked carbohydrate having reducing

groups. The process used a mixture of water, an inorganic acid, and a halide of lithium,

magnesium or calcium. However, these processes used significant amounts of acid (1-10

M), which elevated the corrosion issue. The method to separate sugars and salt was not

mentioned in these patents, which makes the process inapplicable due to the high salt

concentration.

Whereas processes are known in the art for hydrolyzing lignocelluloses there is

still a significant need for efficient and low-cost processes that hydrolyze lignocelluloses,

Page 112: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

97

particularly wood-based materials that are hard to hydrolyze, predominantly to

monosaccharides, with minimal losses to undesired coproducts and preferably without

the need for pretreatment. In response to this need, a new cost-effective way to hydrolyze

cellulose into sugars using acidic concentrated salt solutions at moderate temperature was

developed. This process eliminated the use of energy-intensive pretreatment and costly

enzyme. The sugar yield of over 95% was achieved with limited sugar degradation into

furan compounds. Additionally, several methods of separating salt and sugars that could

maximize sugar stream concentrations and further decrease downstream processing cost

were developed.

4.2 Experimental

4.2.1 Materials and Chemicals

Chemical reagents used in this study were purchased from Fisher Scientific

(Pittsburgh, PA) or Sigma-Aldrich (St. Louis, MO) and used as received. Air-dried

lignocelluloses, ground using a Wiley mill to pass a 40-mesh screen, were used in the

present study. Chemical composition of different types of lignocelluloses is presented in

Page 113: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

98

Table 4.1. All the chemical reagents and solvents used in this study were purchased from

Fisher Scientific or sigma-Aldrich and used as received.

4.2.2 Liquefaction of lignocellulose in concentrated LiBr solution

LiBr (700, 1000, 1300, 1400, 1500, 1700, 2000 or 2500 mg) was dissolved in 1

mL water in a 15-mL vial, as shown in Figure 4.3. To the solution, 100 mg 40 mesh

spruce powder was added and then vortexed to mix well. Vials were heated at 100-160 ºC

in an oil bath and stirred with a magnetic stir bar for 2 h. After the reaction, the

hydrolysate was filtered and the residue was washed with water. Filtrate and washings

were collected for glucose and HMF or furfural analysis by HPLC.

4.2.3 Hydrolysis of lignocellulose in acidic concentrated LiBr

solution

(1) Optimization of reaction condition

The chemical reagents list in Table 4.5 were added into 5 mL water and stirred to

dissolve well. To this solution, 0.5 g spruce powder was added and stirred to react for

certain time period list in each entry.

Page 114: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

99

(2) Batch feeding to enhance biomass loading

For batch feeding, 7.5 g LiBr and 50 mg concentrated HCl were dissolved into 5

mL water and stirred until clear solution was obtained. To this solution, 0.5 g biomass

was added each batch at an interval of 5 min with stirring.

After the addition of feedstock, hydrolysis was allowed to proceed at a total time

(from the first feedstock addition to the end of reaction) listed in the Table 4.6. For

entries 4 to 9, after the sixth feeding, additional 20 mg HCl were added and reactions

were allowed to proceed for 10-30 min before the next addition. The composition of

original poplar and corn stover is shown in Table 4.1.

4.2.4 Extraction of LiBr by organic solvents

Firstly, the extraction condition was optimized, as shown in Table 4.9 and 4.10.

Then real lignocellulose hydrolysate was subjected to the extraction with butanol/hexane

for defined times in 50 mL Falcon tube. In each reaction, after adding organic solvent, the

tube was vortexed for 1 min to complete the extraction and then was centrifuged to

separate the organic phase and water phases. The organic phase was removed and fresh

organic solvent was added to repeat this extraction procedure. Details are shown in the

note under Table 4.9.

Page 115: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

Table 4.1 Composition analysis of different feedstocks

Note: ND, not determined.

Composition (%) Spruce Corn Stover Poplar Switch grass Newspaper Printpaper

Moisture 10.00 4.50 9.56 9.86 ND ND

Extractive 5.00 15.50 6.32 16.58 ND ND

Saccharide

Arabinose 0.98 2.71 0.29 1.93 0.35 0.19

Galactose 2.34 1.05 0.53 0.71 0.40 0.08

Glucose 42.03 35.28 43.45 29.66 63.20 71.28

Xylose 5.18 18.40 13.41 18.01 12.13 14.74

Mannose 10.18 0 2.34 0.33 4.86 3.41

Total lignin 26.88 16.78 19.42 15.51 13.08 3.22

100

Page 116: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

101

4.2.5 Removal of salt by ion-exchange chromatography

Two small columns (1 cm in diameter and 5 cm in length) were packed, one with

anion exchange resin (DOWEX-2, Cl- form, 100-200 mesh), and one with cation

exchange resin (Amberlyst 15, 25-50 mesh), as shown in Figure 4.13. Prior to operation,

the anion exchange column was washed with 10 mL of 5% sodium hydroxide solution,

and the cation exchange column was washed with 10 mL of 5% sulfuric acid. The

columns were then washed with water until neutral pH. Both columns were de-watered

by injecting air to prevent the water from diluting the sugar solution. The diluted syrup of

sugar concentration of about 50% (0.5 mL) was loaded onto the cation exchange column

first to remove Li+. The loaded sample was pushed through the column by injecting air.

The recovered solution was then loaded onto the anion exchange column to remove Br-.

Again, air was used to push the sample through the column. The recovered solution was

analyzed for LiBr. The whole process scheme is shown below.

(1) Removal of LiBr

Resin-H+ + LiBr Resin-Li

+ + HBr

Resin-OH- + HBr Resin-Br

- + H2O

Page 117: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

102

(2) Regeneration of resins

Resin-Li+ + 1/2 H2SO4 Resin-H

+ + ½ Li2SO4

Resin-Br- + 1/2 Ca(OH)2 Resin-OH

- + 1/2 CaBr2

(3) Recovery of residual LiBr

1/2 Li2SO4 + 1/2 CaBr2 LiBr + 1/2 CaSO4

4.2.6 Quantification of sugars and sugar derivatives

Sugar analysis was conducted using a Dionex HPIC system (ICS-3000) equipped

with integrated amperometric detector and Carbopac™ PA1 guard and analytical

columns at 20 ºC. Eluent was provided at a rate of 0.7 mL/min, according to the

following gradient: 0-25 min, 100% water; 25.1~30 min, 30% water and 70% 0.1 M

NaOH; 30.1-35 min, 100% water. To provide a stable baseline and detector sensitivity,

0.5 M NaOH at a rate of 0.3 mL/min was used as post-column eluent.

5-hydroxylmethylfurural (HMF) and furfural were analyzed using the Dionex

ICS-3000 equipped with a Supelcogel C-610H column at temperature 30 ºC and UV

detector at 210 nm. Eluent was 0.1% phosphoric acid at a rate of 0.7 mL/min.

Page 118: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

103

4.2.7 Determination of LiBr amount

The pH value of the solution to be determined was adjusted by 2 M NaOH to 7-9

before titration. The amount of LiBr was determined by the Mohr method (Doughty,

1924). Content of Br- and Cl

- was determined by titration of AgNO3 with KCrO4 as

indicator. The mechanism of titration is shown below.

Br- + Ag

+ AgBr (yellow precipitant)

Cl- + Ag

+ AgCl (white precipitant)

Ag+ + CrO4

- AgCrO4 (red precipitant)

4.3 Result and discussion

4.3.1 Description of whole process

As schematically shown in Figure 4.1, real biomass (such as corn stover,

switchgrass, hardwood, and softwood) was mixed with LiBr, a small amount of acid, and

water at a desired LiBr concentration (e.g., 40-70%). The mixture was heated at elevated

temperature (100-160 ºC) for typically several minutes to several hours (dependent on

acid concentration, salt concentration, biomass species and particle size) with stirring.

Page 119: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

104

Both cellulose and hemicelluloses of the biomass were completely hydrolyzed

(saccharified), whereas lignin (up to 30% of the biomass) remained as an insoluble

residue. By filtration or centrifugation, lignin was separated from the solution of sugars

and LiBr. The sugars from cellulose and hemicelluloses and LiBr could be further

separated by ion exchange, extraction, or crystallization methods based on their

differences in ionization and solubility in water and organic solvents. The recovered LiBr

can be reused in the process, and the sugars can be converted into biofuels and chemicals

by biological or chemical approaches.

Figure 4.1 Process flow chart of biomass saccharification in concentrated salt solution.

Page 120: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

105

The fast chemical saccharification method was able to simultaneously hydrolyze

cellulose and hemicelluloses without the need for extensive size reduction or any

chemical pretreatment of the biomass, as illustrated in Figure 4.2,. The process could

produce a concentrated sugar solution (>30%, w/w) in high yield with limited

degradation of sugars to furans (furfural and hydroxymethylfurfural). Sugars produced

were predominantly monosaccharides. In addition, the process could directly handle

small sized wood chips without requiring extensive size reduction or any other

pretreatment, which significantly simplified operation and reduced processing cost and

energy consumption. This process will be discussed in detail below.

Figure 4.2 Hydrolysis way of biomass in concentrated salt solution.

4.3.2 Liquefaction of lignocellulose in concentrated LiBr solution

We initially studied the liquefaction of various types of lignocelluloses in a series

of metal halide salts and found that concentrated LiCl, LiBr and CaBr2 solutions (60%,

Page 121: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

106

w/w) had very good capabilities to liquefy lignocellulose through the dissolution of

cellulose and the partial hydrolysis of cellulose into oligosaccharides and

monosaccharides. In this thesis, the focus is on the applicability of concentrated LiBr as a

solvent for hydrolysis of lignocellulose. Experimental results show that lignocellulose

could be liquefied/hydrolyzed in concentrated LiBr solution to release saccharides (both

oligosaccharides and monosaccharides) at moderate temperatures (120-160 C), as shown

in Tables 4.2, 4.3 and 4.4. The effect of temperature and LiBr concentration on the

hydrolysis of spruce powder in LiBr solution is summarized in Figure 4.3.

Table 4.2 LiBr-hydrolysis of various types of feedstock

Note: reaction condition: 1 mL H2O, 1500 mg LiBr, 100 mg lignocelluloses powder, 140 ºC, and 2 h.

Pure cellulose such as microcrystalline cellulose (e.g., Avicel® microcrystalline

cellulose, FMC Biopolymer) and dissolving pulp (bleached wood pulp typically

having >95% cellulose) were tested for solubilization and hydrolysis in concentrated

LiBr. The results showed that the pure cellulose dissolved faster (5-30 min) under the

Feedstock Lignin content in biomass (%) Residue left (%)

Switchgrass 16 21

Spruce 27 25.7

Poplar 20 22

Corn stover 17 22.5

Page 122: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

107

same conditions (60% LiBr solution at 140 ºC) than real biomass, presumably because of

the absence of blocking from hemicellulose and lignin. It was found that the molecular

weight of cellulose slightly decreased during the dissolution in concentrated LiBr

solution, indicating that some cellulose hydrolysis occurred during the dissolution

process. Increasing temperature or extending time elevated the degree of cellulose

hydrolysis (or depolymerization) as was expected.

Figure 4.3 Hydrolysis of spruce powder in LiBr solution (no acid) as a function of LiBr

concentration at different conditions.

Note: experimental procedures: Specific amounts of LiBr (700, 1000, 1300, 1400, 1500, 1700, 2000 or

2500 mg) were dissolved in 1 mL water in a 15-mL vial. To the solution, 100 mg 40 mesh spruce powder

was added and then vortexed to mix well. Vials were heated at 100-160 ºC in an oil bath and stirred using a

magnetic stir bar to react for 2 h. After the completion of the reaction, hydrolysate was filtered and the

residue was washed with water. Filtrate and washings were collected for glucose and HMF or furfural

analysis by HPLC.

0

25

50

75

100

40 50 60 70 80

Co

ve

rsio

n(m

ola

r%)

LiBr concentration(w%)

Saccharides, 120℃, 2h

Saccharides, 140℃, 2h

Saccharides, 160℃, 2h

HMF, 120℃, 2h

HMF, 140℃, 2h

HMF, 160℃, 2h

LiBr.4H2O

LiBr.3H2O

Page 123: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

108

Table 4.3 Monosaccharides in the hydrolysates from LiBr hydrolysis of different feedstocks

without acid

Note: reaction condition: 1 mL H2O, 1500 mg LiBr, 100 mg lignocelluloses powder, 140 ºC, and 2 h.

In light of the excellent dissolution/hydrolysis ability of 60% LiBr solution,

hydrolysis of different feedstocks in this solution at 140 ºC was investigated. The

comparison of lignin content in original biomass and the amount of residue after

hydrolysis is shown in Table 4.2. The results indicate that almost all of the carbohydrates

were dissolved. The carbohydrate composition in the hydrolysates before and after

autoclaving at 120 ºC with 3% dilute acid for one hour was shown in Tables 4.3 and 4.4.

Autoclaving oligosaccharides at 120 ºC with 3% acid can completely hydrolyze

oligosaccharides into monosaccharides. Comparing the results in Tables 4.3 and 4.4, one

can deduce that autoclaved hydrolysates contained much more monosaccharides than

original hydrolysates, implying that most of the carbohydrates dissolved from

lignocelluloses in concentrated LiBr solution without acid existed in the form of water-

soluble oligosaccharides. The hydrolysis of cellulose into oligosaccharides in

Sugar type Corn stover (%) Switchgrass (%) Poplar (%) Spruce (%)

Arabinose 0.66 0.86 0.25 0.59

Galactose 0.39 0.55 0.39 1.31

Glucose 2.079 5.11 8.08 9.00

Xylose 7.059 9.42 7.37 2.00

Mannose 0 0 2.80 6.04

Page 124: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

109

concentrated LiBr solution without acid is predicted to be caused by the Lewis acidity of

Li+, which will be further discussed later.

Table 4.4 Monosaccharides in hydrolysate from LiBr hydrolysis of different feedstocks after

autoclaving at 120 °C for 1 h

Note: autoclave procedure: 1 mL hydrolysate was mixed with 1 mL 6% sulfuric acid solution into 2 mL

hydrolysate with 3% acid concentration. This solution was autoclaved at 120 °C for 1 h to hydrolyze

oligosaccharides into monosaccharides.

4.3.3 Dissolution mechanism of cellulose in concentrated LiBr

solution

The conversion of cellulose in spruce powder starts to increase readily when the

concentration of LiBr was larger than 55% at the reaction temperature of 140 ºC. To

explain this phenomenon, mechanism for cellulose dissolution in concentrated LiBr

solution is proposed in Figure 4.4. It is interestingly found that 55% and 60% LiBr

solutions roughly correspond to LiBr·4H2O and LiBr·3H2O, respectively. It can be seen

from Figure 4.3 that either increasing one coordination water or decreasing the reaction

temperature down to 120 ºC caused significantly decreased conversion of cellulose. It is

Sugar type Corn stover (%) Switchgrass (%) Poplar (%) Spruce (%)

Arabinose 0.73 1.07 0.21 0.67

Galactose 0.55 1.09 0.57 2.70

Glucose 30.84 26.43 39.53 40.32

Xylose 16.26 14.06 9.63 3.34

Mannose 0 0 4.72 7.56

Page 125: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

110

understandable that either addition of water or decreasing temperature could cause an

increase of coordinated water around ions. Therefore, the number of coordination water

around Li+ and Br

- was thought to play a key role in affecting the cellulose dissolution

ability of LiBr solutions. A lot of research on the coordination number of Li+ in aqueous

solution has been reported and the coordination number of 4 or 6 is disputable (Wolfram

Rudolph, 1995). The coordination number 6 can probably explain our observations. In

the case of LiBr·3H2O (61% LiBr), as shown in Figure 4.4a, six H2O molecules form a

hexahedral structure with Li+ through Li···O bond, and each H2O molecule is shared by

two Li+ ions. It seems that all Li

+ ions are saturated with water molecules in LiBr·3H2O.

However, as the solvation of Br- likely occupies some of the water molecules, some of

the Li+ ions may not be completely solvated. In other words, Li

+ has potential water-

deficit sites. In particular, when the system is heated, Li···O bond could be broken to

form more water-deficit sites. The Li+ with water-deficit spots can coordinate the O of

hydroxyl groups in cellulose, and meanwhile, the ―naked‖ Br- ions in the solution tend to

form hydrogen bonds with hydrogen of hydroxyl groups. As a result, the Li···O and

Br···H bonds replace the inter- and intra-molecular O···H bonds in crystalline cellulose,

and therefore disrupt cellulose crystal structure and enhance the dissolution of cellulose.

Page 126: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

111

The situation is very different in the case of LiBr·4H2O (55% concentration). As shown

in Figure 4.4b, because of the presence of an extra water molecule, Li+ ions would be

more solvated than those in LiBr·3H2O. Similarly, Br- ions are more solvated than those

in LiBr·3H2O. Therefore, the dissolution ability of LiBr·4H2O is significantly decreased.

Besides, empty Li+ coordination sites can act as Lewis acid to hydrolyze glycosidic bonds.

Because of the hydrolysis ability of Li+, dissolved polysaccharides can be further broken

down into small units and dissolved away from the lignin.

(a)

(b)

Figure 4.4 Models for cellulose dissolution in salt solutions of varied concentrations. (a)

LiBr·3H2O (61% LiBr concentration) and (b) LiBr·4H2O (55% LiBr concentration).

Page 127: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

112

In the experiment, a temperature of 140 ºC (sufficient to overcome the activation

energy) was thought to be the minimum temperature to break Li···O bonds to create

water-deficit sites. Temperatures lower than 120 ºC, as tested in our experiment, could

not provide enough energy to break Li···O bond and create empty sites for dissolution

and hydrolysis of cellulose, so consistently neither dissolution nor hydrolysis was

observed in our experiment. The decreased conversion of sugars observed at higher

temperatures (160 ºC) and higher LiBr concentration (>60%) may be attributed to the

significantly increased Lewis acid sites, which leads to significant condensation of lignin

and degradation of sugars into humins, as indicated by the recovered black residues,

rather than brown residues (lignin). As shown in Figure 4.3, concentrations of LiBr

between 55% and 60% were ideal to dissolve cellulose, as is consistent with the proposed

mechanism. In this range of concentrations, LiBr·4H2O and LiBr·3H2O structures coexist

in solution. With increasing of concentration, LiBr·3H2O increased and thereby

dissolution was enhanced accordingly. The conversion of ~20% in LiBr concentration

lower than 55% is attributable to the dissolution of amorphous cellulose faction in the

spruce powder. In summary, 60% LiBr solution at 140 ºC is the optimum condition to

dissolve/hydrolyze the polysaccharides in spruce.

Page 128: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

113

4.3.4 Hydrolysis of lignocellulose in acidic concentrated LiBr

solution

The experimental results in Figure 4.3 indicates that concentrated LiBr was

capable of dissolving and hydrolyzing the polysaccharides in lignocellulose, but the

hydrolysis was incomplete and slow, and most of the hydrolysis products were in the

form of oligosaccharides (Tables 4.2, 4.3, and 4.4). To accelerate the hydrolysis of

polysaccharides, acid was added to the LiBr solution. The process conditions were further

optimized within the range of LiBr concentrations of 44-60% (w/w), temperatures of

110-160 ºC, acid loading of 1-10% (on spruce powder), and reaction time of 5-60 min, as

summarized in Table 4.5. Entries 1 to 5 indicate that the LiBr concentration is vital to the

hydrolysis of cellulose, which is consistent with the conclusion derived from Figure 4.3.

When LiBr concentration was equal to or above 55% (for example, 60, 56, and 55% in

entries 1, 2, and 3), the conversion/hydrolysis of polysaccharides was complete. Lower

LiBr concentration (for example, 50 and 44% in entries 4 and 5) led to a decreased

conversion of polysaccharides. It should be noted that the degradation of sugars to

furfural and HMF was lower at low LiBr concentrations. Although acid addition and high

temperature can enhance the hydrolysis, LiBr concentration remains the most critical

Page 129: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

114

parameter. When LiBr concentration was lower than 55%, increasing temperature or acid

loading was unable to achieve a satisfactory rate of hydrolysis in decent time, as shown in

the results of entries 4, 5, and 8.

Acid addition and high temperature can enhance the hydrolysis of biomass in

LiBr solution. For example, compared to the results in Figure 4.3, entries 1 and 6 in Table

4.5 show that addition of 2% hydrochloric acid (w/w on spruce powder) could completely

hydrolyze spruce in 60% LiBr solution in less than 10 min. As shown in entries 10 to 12,

satisfactory conversion could be achieved at high temperatures even with low acid

loading (1% on biomass) in 56% LiBr solution. From entries 1 to 14, it can be seen that

100% conversion/dissolution of polysaccharides was achieved in some entries, but the

yield of sugars was lower than theoretical, because either the polysaccharides were not

completely hydrolyzed to monosaccharides (partially still in the form of oligosaccharides)

or part of the sugars were further degraded to furfural and HMF. It is noteworthy that

extending the reaction time or elevating the temperature generally generated more HMF

and furfural. It is not wise to apply higher concentration of acid and higher temperature

simultaneously. At high LiBr concentration, for example 60%, two options are available

to achieve both high conversion and high sugar yield with a low yield of furan

Page 130: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

115

compounds in a short time: (1) high temperature (150-160 ºC) with low acid loading

(1%), such as in entries 10 and 11; (2) high acid loading (5-10% on biomass) at low

temperature (110 ºC), such as in entries 16 and 17. As shown in Table 4.5, up to 96%

sugar yield was obtained under the conditions used in entries 16 and 17 with very limited

degradation of sugars. As low temperature saves energy, the conditions of entries 16 and

17 will be used in our following research.

Page 131: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

Table 4.5. Hydrolysis of spruce powder under various conditions

Note: ―Tem.‖ denotes temperature; ―Ara.‖ denotes arabinose; ―Gal.‖ denotes galactose; ―Glu.‖ denotes glucose; ―Xyl.‖ denotes xylose; ―Man.‖ denotes

mannose; ―HMF.‖ denotes hydroxymethylfurfural; ―Fur.‖ denotes furfural; ―Res‖ denote residues (lignin and unhydrolyzed polysaccharides) after reaction;

―Con.‖ denotes conversion; ―TSY‖ denotes total monosaccharide yield; ―TFY‖ denotes total furan compounds (HMF and furfural) yield; Reaction procedure:

the chemical reagents list in each entry in Table 5 were added into 5 mL water and stirred to dissolve well. To this solution, 0.5 g spruce powder was added

and stirred to react for certain time period list in each entry. ―*‖ the weight of pure hydrogen chloride. ―Ara.‖, ―Gal.‖, ―Glu.‖, ―Xyl.‖, ―Man.‖, ―HMF.‖

―Fur.‖, and ―Res‖ (%) are weight percentage based on original feedstocks. ―con‖, ―TSY‖, and ―TFY‖(%) are molar percentage based on total sugar content

in feedstocks.

Entry LiBr

(g)

HCl*

(mg)

Tem.

(°C)

Time

(min)

Ara.

(%)

Gal.

(%)

Glu.

(%)

Xyl.

(%)

Man.

(%)

HMF

(%)

Fur.

(%)

Res.

(%)

Con.

(%)

TSY

(%)

TFY

(%)

0 / / / / 0.98 2.34 42.03 5.18 10.18 / / 26.88 / / /

1 7.50 10 140 10 0.80 1.79 30.51 2.55 6.29 5.29 2.45 24.85 100 67 12

2 6.25 10 140 10 0.80 1.81 30.22 2.44 6.01 6.00 3.25 25.65 100 66 15

3 6.00 10 140 10 0.83 1.84 30.96 2.84 6.39 4.50 2.53 26.12 100 69 11

4 5.00 10 140 10 1.05 2.17 13.01 3.81 7.49 1.86 2.27 42.31 76 44 7

5 4.00 50 140 20 0.18 0.90 13.98 0 1.95 1.62 1.99 49.63 64 27 6

6 7.50 10 140 5 0.89 1.95 34.43 3.73 7.05 3.2 2.02 24.78 100 77 8

7 6.00 50 120 10 0.66 1.72 27.61 1.64 4.79 4.05 2.78 32.31 91 58 11

8 5.00 50 120 10 0.96 2.30 18.26 2.95 7.39 1.69 3.04 36.54 85 51 8

9 7.50 10 120 25 0.81 1.82 31.69 3.00 6.32 3.28 2.24 30.12 95 70 9

10 6.25 5 160 5 1.01 2.08 36.68 4.41 8.05 2.98 1.58 22.65 100 84 7

11 6.25 5 150 5 1.03 2.05 35.20 4.68 8.21 1.65 0.88 23.21 100 82 4

12 6.25 5 140 5 1.10 1.99 27.76 4.94 7.72 0.66 0.57 23.82 100 70 2

13 7.50 10 100 60 0.98 2.02 36.59 4.17 8.48 1.31 1.01 26.32 100 84 4

14 7.50 50 100 10 0.90 1.96 34.23 3.69 7.71 1.68 1.46 25.65 100 78 5

15 5.00 25 140 15 0.32 1.17 17.52 0.06 2.41 2.68 3.06 41.01 78 34 9

16 7.50 50 110 5 1.16 2.36 43.38 5.06 9.97 1.41 1.15 27.12 100 99 4

17 7.50 25 110 5 1.16 2.38 42.46 5.18 9.71 0.98 0.86 26.93 100 98 3

116

3

Page 132: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

117

4.3.5 Hydrolysis of lignocellulose through batch feeding

In order to increase production efficiency and decrease LiBr/sugars separation cost

in downstream processing, a high feedstock loading (a low ratio of LiBr solution to

feedstock solid) is desirable, as it will lead to high concentration final sugar solution.

However, addition of too much feedstock caused mixing and mass transfer problems and

therefore affected the hydrolysis efficiency and yield. In order to overcome these problems,

feedstock was fed in multiple steps to ensure that enough liquid is available to wet

biomass and hydrolyze the cellulose at all times. In our experiment, biomass feedstock

was fed into the reactor as soon as the last batch was closed to completely liquefied and

hydrolyzed. From entries 1 to 5 in Table 4.6, 1, 2, 3, 4 and 5 g spruce powder were added

into a 60% LiBr solution (7.5 g LiBr + 5 mL water) (feeding procedure is described in the

notes under Table 4.6). Results showed that when total biomass feeding was below 4 g

(entries 1 to 4), almost all of the polysaccharides in spruce powder was hydrolyzed into

monosaccharides with high selectivity. Figure 4.5 shows the picture of the separated

hydrolysate after centrifugation.

Page 133: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

Table 4.6 Hydrolysis of spruce powder in batch-feed mode

Entry LiBr

(g)

HCl*

(mg)

Tem.

(°C)

Bio.

(g)

Time

(min)

Ara.

(%)

Gal.

(%)

Glu.

(%)

Xyl.

(%)

Man.

(%)

HMF

(%)

Fur.

(%)

Res.

(%)

Con.

(%)

TSY

(%)

TFY

(%)

0 / / / / / 0.98 2.34 42.03 5.18 10.18 / / 26.88

1 7.5 50 110 1 20 1.14 2.31 42.89 5.11 10.02 1.22 0.88 26.92 100 98 3

2 7.5 50 110 2 40 1.07 2.36 42.22 5.13 9.69 1.12 1.03 26.34 100 97 3

3 7.5 50 110 3 60 0.98 2.03 41.30 5.04 9.78 1.06 1.25 25.78 100 95 4

4 7.5 70 110 4 80 1.21 2.11 41.75 5.09 9.82 0.94 1.31 25.19 100 95 4

5 7.5 70 110 5 100 0.94 2.05 41.93 5.11 9.78 0.97 1.25 27.07 92 86 3

6 7.5 50 110 4 80 0.24 0.45 42.81 12.24 2.27 0.87 1.01 19.11 100 99 2

7 7.5 50 110 4 80 2.01 0.86 35.22 17.34 0 0.67 1.55 16.24 100 99 5

8 7.5 50 110 4 80 0.25 0.44 39.32 11.35 2.06 0.77 1.23 25.25 90 90 4

9 7.5 50 110 4 80 2.12 0.99 32.21 16.12 0 0.69 1.61 23.27 91 89 4

Note: Entries 1 to 5: spruce powder; entry 6: poplar powder; entry 7: corn stover powder; entry 8: poplar chips (0.2-0.5 cm); entry 9: corn stover chips (0.2-0.5

cm). ―Tem.‖ denotes temperature; ―Bio‖ denotes biomass; ―Ara.‖ denotes arabinose; ―Gal.‖ denotes galactose; ―Glu.‖ denotes glucose; ―Xyl.‖ denotes xylose;

―Man.‖ denotes mannose; ―HMF.‖ denotes hydroxymethylfurfural; ―Fur.‖ denotes furfural; ―Res‖ denote residues (lignin and unhydrolyzed polysaccharide)

after reaction; ―Con.‖ denotes conversion; ―TSY.‖ denotes total monosaccharides yield; ―TFY.‖ denotes total furan compounds (HMF and furfural) yield. ―*‖

the weight of pure hydrogen chloride. Reaction procedure: 7.5 g LiBr and 50 mg concentrated HCl were dissolved into 5 mL water and stirred until clear

solution was obtained. To this solution, 0.5 g biomass was added each batch at an interval of 5 min with stirring. After the addition of feedstock, hydrolysis

was allowed to proceed at a total time (from the first feedstock addition to the end of reaction) listed in the Table. For entries 4 to 8, after the sixth feeding,

additional 20 mg HCl were added and reactions were allowed to proceed for 10-30 min before next addition. The composition of original poplar and corn

stover is shown in Table 1. ―Ara.‖, ―Gal.‖, ―Glu.‖, ―Xyl.‖, ―Man.‖, ―HMF.‖ ―Fur.‖, and ―Res‖ (%) are weight percentage based on original feedstocks. ―con‖,

―TSY‖, and ―TFY‖(%) are molar percentage based on total sugar content in feedstocks.

118

Page 134: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

119

However, when 5 g spruce powder was fed, the conversion decreased compared

to previous lower feeding because the high solid loading and the accumulated lignin

residue in the hydrolysate made the stirring and mass transport difficult. If efficient

mixing is provided, the upper limit of total biomass loading may be increased, and the

hydrolysis yield at high feedstock loading may be improved.

Figure 4.5 The picture of the separated hydrolysate.

In entries 6 and 7, polysaccharides in poplar and corn stover powder were also

completely hydrolyzed into monosaccharides using the same conditions and procedure

for as entry 4. Because the hydrolysis was a heterogeneous reaction process, biomass

Page 135: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

120

size had a significant influence on the hydrolysis rate. For example, when poplar and

corn stover chips particle size of 2-5 mm was used, they were still hydrolyzed but

hydrolysis yield decreased to around 90%, as shown in entries 8 and 9; this was lower

than those in entries 6 and 7 where 40 mesh biomass powder was used. Extension of

reaction time or efficient mixing could possibly achieve a complete hydrolysis.

4.3.6 Saccharification of lignocellulose in concentrated solution

of different salts

Hydrolysis of biomass in different metal halide solutions was investgiated.

Results in Table 4.7 show that when the concentration of salt was 60%, LiCl, LiBr and

CaBr2 could give complete conversion/hydrolysis of polysaccharides. CaBr2 is as

effective as LiBr in this experiment. In addition, experiments with separation methods

discussed below indicate that the methods for separating LiBr from sugars, also work

with CaBr2. Furthermore, CaBr2 is currently much cheaper than LiBr and as such is

currently preferred. Although LiCl also gave a complete conversion, the solubility of

LiCl in water is significantly lower than that of LiBr. For example, the solubility of LiBr

in water at 90 ºC is about 254 g/100 mL, whereas that of LiCl at 100 ºC is about 128

g/100 mL. As a consequence, 60% LiCl starts crystalizing out at 120 ºC, which makes

Page 136: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

121

downstream separation operations difficult. ZnCl2 and CaCl2, consistant with the reports

by Penque and Chen, only gave a conversions of 50~70% (Chen, 1984; Penque, 1977).

Table 4.7 Effect of different salts on hydrlysis of biomass

Note: reaction condition: 1 mL H2O, 1500 mg salt, 100 mg spruce powder, 2 mg HCl, 140 ºC, and 30 min.

It is also instetrestingly found that LiCl concentration of 60% has significantly

higher molar concentration than LiBr concentration of 60%, indicating that the structure

of LiCl and LiBr in aqueous soltuion are different. As we discussed previously,

LiBr·3H2O (60% LiBr concentration) could dissolve cellulose. This is not the case for

LiCl·3H2O (44% LiCl concentration) which could not dissolve cellulose. The possible

explanation is that they have different struture caused by different properties of anions

(Cl- and Br

-). Because Cl

- has higher charge density than Br

- , it can form stronger ionic

bonds with Li+. Additionally, Cl

- and water has comparable affinity to Li

+ and can

compete for coordination sites, which lead to some coordination sites occupied by Cl-. A

Salt Conversion (%)

LiBr 100

LiCl 100

CaBr2 100

ZnBr2 66

ZnCl2 60

MgBr2 75

CaCl2 70

NaBr 68

Page 137: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

122

possible struture is shown in Figure 4.6. As we disussed previously, applying heat could

break Li···O bonds to create water-deficit sites. However, in LiCl·3H2O, the situation is

same to that of LiBr·4H2O where the created empty site will be occupied by water first.

Therefore, LiCl·3H2O has very poor dissolution ability. Considering LiCl·2H2O, the

empty site could be occuped by a cellulose hydroxyl group. However, it is observed that

under a LiCl concentration of 54% (LiCl·2H2O), the dissoluton is also very poor even

though sinificant swelling was observed. This might by attribted to the inactive Cl-

which was restricted by Li+ and has decreased activity. Dissoluton of cellulose could be

completed at a LiCl conentration of 60% which is supposed to contain both LiCl·2H2O

and LiCl·H2O and have much more empty sites for cellulose hydroxyl groups, as shown

below. The structure of LiCl·3H2O ,LiCl·2H2O, and LiCl·H2O are shown below in

Figure 4.6. The proposed structure of LiCl·2H2O is consistent with reported crystal

structure of LiCl·2H2O (Brendler et al., 2002). Therefore, we think that the dissolution of

cellulose in 60% LiCl solution should be mostly attributed to rich empty coordination

sites on Li+.

Page 138: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

123

LiBr·3H2O LiCl·3H2O LiCl·2H2O LiCl·H2O

Figure 4.6 The proposed structures of LiCl hydrates.

Why concentrated solutions of LiBr, LiCl and CaBr2 are more efficient at

hydrolyzing cellulose than other metal halides. Research has concluded that the

following characteristics mainly determine the dissolution power towards cellulose: (1)

the acidity, (2) the water content of the melts, and (3) the properties of the coordination

sphere of the cations. However, there is no report to explain how these factors affect

dissolution of cellulose. Based on the hypothesis discussed above, these factors can have

a reasonable explanation. Higher acidity metal ions have a higher ability to coordinate

with cellulose hydroxyl group and hydrolyze cellulose. Lower water content can lead to

more empty coordination sites for coordination cellulose hydroxyl group and for

hydrolysis of polysaccharides. Anion will significantly affect the coordination structure

of cation. High charge density, such as Cl-, could allow anion to form strong cation-

anion association and water-anion interactions, leading to inactive anions. Besides, the

Page 139: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

124

relaxation time (desolvation kinetics) of salvation shell might also significantly affect the

binding of other ligands such as hydroxyl group of cellulose. Fast and frequent

desolvation of metal ion can automatically create empty coordination sites for cellulose

hydroxyl group. For example, the lifetime of Ca2+

-water coordination is about 18 ps, 10

times longer than the relaxation time previously reported for K+ or Na

+. Even though

Mg2+

is slightly smaller than Ca2+

, the lifetime of water molecules around Mg2+

is on the

order of a few hundreds of picoseconds (Jiao, 2006). This might explain why calcium

halides are better than Magnesium halides. Different salts have different structures

because of the properties of cations and anions and these factors discussed above work

together to determine the dissolution ability of salts.

4.3.7 Hydrolysis of lignocellulose in concentrated LiBr solution

with different acids

It was already found that acid could significantly enhance the hydrolysis of

lignocelluloses in concentrated LiBr solution, as disscused above. Here different acids

were tested for their hydrolysis efficacy on spruce powder in 60% LiBr solution. It can

be seen from Table 4.8 that strong acids such as sulfuric acid and nitric acid are as

effective as hydrochloric acid in catalyzing the hydrolysis of spruce, while weak acids

Page 140: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

125

gave a decreased hydrolysis yield. For example, acetic acid had a yield of 90% at the

same conditions.

Table 4.8 Effect of different acids on hydrlysis of biomass

Note: reaction condition: 1 mL H2O, 1500 mg LiBr, 100 mg spruce powder, 2 mg acid, 140 ºC, 30 min.

4.3.8 Separation of LiBr and sugars by different methods

The chemical saccharification process presented here uses concentrated LiBr

solution as a medium. It is obvious that effective separation of sugars and LiBr after the

saccharification is critical to the industrial application of the process. Fortunately, LiBr is

quite different from sugars in physical and chemical properties, which can be used to

separate LiBr from sugars. Three methods including boronic acid extraction, ion

exclusion chromatography, and solvent extraction, were used to separate LiBr from

sugars in this study.

4.3.8.1 Extraction of sugar from LiBr-sugar solution with boronic acid

Boronic acid can react with cis-diol in sugars in the presence of lipophilic

Acid Conversion (%)

HCl 100

H2SO4 100

HNO3 100

H3PO4 95

HCOOH 96

CH3COOH 90

Page 141: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

126

quaternary alkyl amines to form a stable boronic acid-sugar complex under alkaline

conditions. The complex is soluble in organic solvent and can be extracted with hexane,

by which sugars can be separated from LiBr. The mechanism of this method is shown in

Figure 4.7. The boronic acid-sugar complex is stable under alkaline conditions but

unstable under acidic condition. Therefore, the sugar extracted to the organic phase can

be recovered by stripping the organic phase with acidic water. The boronic acid remains

in the organic phase and can be used in next batch extraction.

Figure 4.7 Mechanism for extraction of glucose with boronic acid.

Note: Naphthalene-2-boronic acid (N2B) was used in this study. A glucose-LiBr solution (100 mg glucose

and 500 mg LiBr in 1 mL buffer solution at pH = 11) was tested. The organic phase was prepared by

dissolving 100 mg N2B and 200 mg Aliquat 336 into 5 mL hexane/octanol (85:15, v/v). The sugar-LiBr

solution (1mL) was mixed with the organic solution (5 mL) and vortexed for 30 min to reach equilibrium.

The mixture was then centrifuged to separate the aqueous phase and organic phase. The organic phase was

then stripped with 5 mL 1% HCl solution. Analysis of the stripping solution indicated that approximately

10 mg sugar was recovered through the single pass of extraction.

O

HOHO

OH

B

HO

HOB

HO

HO

OH

B

OO

OH

OH- glucose

glucose

aqueous phase

organic phase

OHO

HO

OH

OHOH

aqueous phase

organic phase

OHO

HO

OH

B

OO OH

organic phase

NH4

Alkyl

H+

OHO

HO

OH

B

OO

organic phase

B

HO

HO

LiBr

NH4

Alkyl

LiBr

H2O

NH4

Alkyl

LiBr

organic phase

not stable stable

Page 142: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

127

The preliminary results indicate that the boronic acid method was able to separate

LiBr and sugars, but the selectivity and efficiency were not satisfactory. In addition, the

method was expensive and had less potential in industrial applications.

4.3.8.2 Separation of LiBr and sugars with ion-exclusion chromatography

Ion-exclusion chromatography has been applied to separate salts and sugars. The

process does not consume acid or alkaline as ion exchange chromatography does. Ion-

exclusion chromatography uses the resin containing the same anion or cation as that in a

salt as its stationary phase. Because of the exclusion force between the resin and the salt,

and penetration of sugars into the micropores of the resin, the salt will elute faster than

sugars. The mechanism is shown in Figure 2.8.

The experimental procedure was shown in the note under Figure 4.9. A synthetic

LiBr-sugars solution (10 mg arabinose, 10 mg galactose, 80 mg glucose, 20 mg xylose,

20 mg mannose, 50 mg LiBr, and 0.5 mL water) was loaded to a column and eluted with

de-ionized water at a rate of 1.5 mL/min. The profile in Figure 4.9 shows that LiBr and

sugars separated very well, indicating that this method is an efficient way of separating

sugars and LiBr.

Page 143: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

128

Figure 4.8 Schematic diagram for ion-exclusion chromatography for separating sugar and salt.

The hydrolysate of spruce was tested with ion-exclusion method as well. When

ethanol is removed/recovered by evaporation, the hydrolysate from section 4.3.9.2 (~80

mg glucose, ~15 mg xylose, ~60 mg mannose, and ~50 mg LiBr) was diluted with water

into a solution of 0.5 mL. This solution was loaded on the column and eluted with de-

ionized water at a rate of 1.5 mL/min. The result in Figure 4.10 shows that LiBr and

sugars could be separated very well, and the recovered sugar and LiBr concentrations

were ~1% and ~1%, respectively.

Page 144: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

129

Figure 4.9 Separation of LiBr from sugar solution using ion-exclusion chromatography.

Note:Separation test was performed on a glass column (diameter: 2 cm; length: 50 cm), packed with

anion exchange resin (DOWEX 18-400, Cl- form) at room temperature. Prior to test, the column was

fully converted into Br- form by eluenting with 400 mL of 0.2 N NaBr at a flow rate of 1.5 mL/min,

followed by a thorough rinse with 2000 mL of deionized water, after which it was ready for sugar

separation test. Sugar solution was injected at the top of the column. The sample was then eluted with

deionized water at a flow rate of 1.5 mL/min. AgNO3 solution was used to monitor the elution out of LiBr,

after which 10 fractions of 2.5 mL were collected from the exit of the column, and their sugar profiles

were determined off-line with a HPIC. LiBr amount in each fraction was determined through titration.

The raw hydrolysate from spruce saccharification with LiBr without any

extraction was tested as well. It was possible to separate concentrated LiBr and sugars

directly using ion-exclusion chromatography. However, because of the high

concentration of LiBr, longer column and more water were needed for a good separation;

thus the significant dilution of LiBr and sugars is the major drawback of the method.

15 16 17 18 19 20 21 22 23 24 25 26

Collected fractions

LiBr glucose xylose mannose arabinose galactose

Page 145: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

130

Figure 4.10 Separation of residual LiBr from sugar solution using ion-exclusion chromatography.

(~80 mg glucose, ~15 mg xylose, ~ 60 mg mannose, and ~50 mg LiBr),

4.3.8.3 Extraction of LiBr from LiBr-sugar solution by organic solvents

Direct separation of LiBr and sugars in an aqueous solution is very difficult

because they are both highly soluble in water. However, LiBr and sugars have very

different solubility in organic solvent (such as alcohol, ketone and ether) where LiBr is

still highly soluble but sugars are not, making it possible to separate LiBr from sugars by

extraction with water-immiscible organic solvents. Extraction of LiBr from brines by

organic solvent such as TBP (tributylphoshate) (Gou and Zhu, 1998; Huang, 1991) and

15 16 17 18 19 20 21 22 23 24 25 26

Collected fractions

LiBr

Glucose

Xylose

Mannose

Page 146: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

131

butanol (Uhlemanna, 1993) and CaBr2 from brines by amine (Grinbaum et al., 2008) has

been reported.

In the present study, n-butanol was chosen as an extraction solvent for LiBr

because of its relatively low price. LiBr is soluble in n-butanol, but sugars are not. As n-

butanol is slightly soluble in water (63.2 g/L), small amount of water will be picked up

into n-butanol layer in the process of extraction. The presence of water increases the

solubility of both salt and sugars in organic phase. As a result, small amount of sugars

will be extracted into the organic phase. To reduce the solubility of sugars in n-butanol,

hexane was used as a phase modifier. However, the inclusion of hexane inevitably

decreases the solubility of LiBr in n-butanol due to the decreased polarity of the organic

phase. Because of strong interaction between LiBr and sugars in thick sugar syrup, it is

very difficult and costly to completely remove LiBr from sugars by extraction alone.

Small amount of LiBr left over in the sugar stream after the extraction should be

removed by other means, such as ion-exchange resin, crystallization of sugars, anti-

solvent precipitation of sugars, precipitation of salt by sodium carbonate, as shown in

Figure 4.11.

Page 147: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

132

The results in Tables 4.9 and 4.10 show that not only LiBr but also glucose were

extracted into butanol. It was also observed that small amount of water was also

extracted into the butanol phase. As glucose is insoluble in pure butanol, the extraction

of glucose into butanol should be attributed to the presence of water and LiBr in butanol.

In order to reduce the extraction of glucose into butanol phase, hexane was added as a

phase modifier.

Figure 4.11 Flowchart for separating LiBr and sugars by solvent extraction.

Page 148: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

133

Table 4.9 Separation of glucose and LiBr by extraction with a mixture of butanol and hexane

Note: Glucose and LiBr solution (200 mg glucose and 1500 mg LiBr in 1mL water) was extracted with

butanol/hexane in a 10-mL screw capped vial. In each extraction, 1 mL organic solvent was added, and

the vial was vortexed for 1 min for extraction. Organic phase was separated from aqueous phase by

centrifugation. When organic phase was removed, 1 mL fresh organic solvent was added for next

extraction.

Table 4.10 Separation efficiency of LiBr and glucose by solvent extraction

Note: calculation is based on composition analysis of the organic phase from the 1st time extraction.

As shown in Table 4.9, when the ratio of hexane in butanol increased gradually,

less glucose was extracted and more extraction times were needed to achieve satisfactory

Butanol/hexane

(v/v) Extraction times

Glucose retained after

extraction (mg)

Glucose retained

/extraction times

10:0 6 95 16

9:1 7 115 16.5

8.5:1.5 8 140 17.5

8:2 9 160 17.8

7.5:2.5 9 175 19.4

7:3 10 185 18.5

6.5:3.5 12 187 16

6:4 13 189 14.5

5:5 15 190 12.6

Butanol/hexane Glucose extracted (mg) LiBr extracted (mg) LiBr/sugar

10:0 20 383 19

9:1 12 310 26

8.5:1.5 7 277 40

8:2 5.5 252 46

7.5:2.5 5 231 46

7:3 4.5 216 48

6.5:3.5 4 195 49

6:4 3 152 51

5:5 2 141 71

Page 149: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

134

extraction of LiBr because of the decreased solubility of LiBr in the polarity-decreased

solvent. When butanol to hexane ratio increased to 7:3, about 10% of the original

glucose (~30 mg out of 320 mg glucose) was extracted into butanol phase. Further

increasing hexane ratio was unable to significantly reduce the extraction of glucose,

instead, more extraction times were needed for a satisfactory extraction, as shown in

Table 4.10. Meanwhile, because of the significant extraction of both LiBr and water into

the organic phase, the volume of the hydrolysate decreased gradually, resulting in a thick

sugar syrup. For example, with the butanol to hexane ratio of 7:3, extracting 10 times

reduced the amount of LiBr in the LiBr-glucose solution (see note under Table 4.9) from

1500 mg to 250 mg. Extracting 20 times and 30 times could decrease the amount of LiBr

from 1500 mg to 50 mg and from 1500 mg to 30 mg, respectively. Further extraction

could remove more LiBr, but it was very difficult to completely remove LiBr from

glucose. This might be caused by the strong interaction between glucose and LiBr. In the

thick sugar syrup, LiBr molecules were surrounded by glucose and water molecules

through hydrogen bonding. Therefore, the organic solvent in which glucose and water

have much lower solubility was unable to penetrate into the thick sugar syrup to extract

the residual LiBr.

Page 150: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

135

Figure 4.12 The picture of extraction of LiBr from hydrolysate with butanol-hexane.

As we discussed previously, CaBr2 can work as effectively as LiBr to hydrolyze

lignocelluloses. Therefore, we tested the extraction of CaBr2 with butanol in the same

procedure. Preliminary result shows that CaBr2 could also be extracted into

butanol/hexane (extraction procedure same to that of LiBr extraction); however, lower

butanol/hexane ratio (5:5) was needed to facilitate the separation of CaBr2 and glucose.

When butanol/hexane (7:3) was used to extract CaBr2-glucose solution (1500 mg CaBr2,

1 mL water, and 200 mg glucose), ~95 mg glucose was left in the aqueous phase after 10

times’ extraction. When butanol/hexane (5:5) was used to extract CaBr2-glucose solution

(1500 mg CaBr2, 1 mL water, and 200 mg glucose), extracting 30 times decreased the

Page 151: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

136

amount of CaBr2 from 1500 mg to 70 mg, and glucose from 200 mg to only 190 mg. It

can be seen that more extraction times were needed to further decrease residual CaBr2. In

real application, the extraction process could be continuous and the amount of solvent

will be significantly decreased because of fresh extraction solvent in later stage could be

reused in previous stage. It might be more effective and economic to remove the residual

LiBr in other ways.

Based on the result of extracting LiBr from glucose-LiBr solution above, it was

tested to separate LiBr from real biomass hydrolysate with n-butanol. Spruce powder (4

g) was hydrolyzed using the conditions of entry 4 in Table 4.6. After the hydrolysis, the

hydrolysate was centrifuged to separate supernatant and precipitate (lignin and

unhydrolyzed spruce, if any). Precipitate was washed with 5 mL of n-butanol twice to

recover LiBr in the precipitate. The butanol washings were mixed with hexane

(butanol/hexane, 7:3, v/v) and used as extraction solvent. The collected supernatant

(containing 0.048 g arabinose, 0.084 g galactose, 1.64 g glucose, 0.2 g xylose, and 0.4 g

mannose) was shown in Figure 4.5 and extracted with butanol/hexane (7:3) for 20 times

in a 50-mL screw capped bottle. Specifically, 5 mL of organic solvent was added in the

first extraction, as shown in Figure 4.12. After vortexed for 1 min, the bottle was

Page 152: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

137

centrifuged to separate organic phase and aqueous phase. When the organic phase was

removed, 5 mL of fresh organic solvent was added. The extraction was repeated 20 times.

After the extraction, the resulting syrup-like sugar mixture was analyzed, which

contained 1.4 g glucose, 0.075 g xylose, 0.3 g mannose, 0.260 g LiBr, and 1 mL water,

as shown in Figure 4.13. From the sugar content of spruce shown in Table 4.6 and initial

LiBr usage, it was calculated that the recovery yields of glucose, xylose and mannose

were 83, 36, and 74%, respectively, and that 96.5% of initially loaded LiBr was

recovered. The unrecovered sugars could be found in two parts. Small amount of the

sugars was degraded to furfural and HMF during the hydrolysis, and the rest was

extracted to the organic layer, which could be recovered in next cycle.

Figure 4.13 Formed sugar syrup after butanol extraction of hydrolysate.

Page 153: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

138

4.3.9 Removal of residual LiBr from sugar stream

4.3.9.1 Precipitation of sugars from sugar-LiBr solution by anti-solvent

Because sugar has lower solubility in organic solvent than in water, it is expected

that the sugars in the sugar syrup obtained after n-butanol extraction would precipitate

out and LiBr would retain in the mother liquor if water-miscible anti-solvent such as

methanol, ethanol or acetone, in which sugars are insoluble and LiBr is soluble, is added.

By doing this, sugars and LiBr might be completely separated.

However, it was found that the direct addition of methanol or ethanol did not

work, and no sugars precipitated, because the alcohols and the syrup are miscible. The

presence of water and LiBr in the syrup was thought to increase the solubility of sugars

in the water-alcohol mixture. On the other hand, addition of less polar acetone into the

syrup did not work either. The sugar syrup was partially dissolved in acetone, and the

undissolved part still contained a significant amount of LiBr, indicating that acetone is

not suitable to extract LiBr and precipitate out sugars. The dissolution of sugars in the

organic solvent is attributed to the presence of water in the syrup.

In order to reduce the dissolution of sugars in organic solvent, we tried to remove

the water by evaporation before the precipitation. However, the presence of LiBr made

Page 154: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

139

the complete removal of water in the sugar syrup very difficult through vaporization

because of the extremely low vapor pressure of LiBr-water. It was found that after the

water was evaporated, the syrup was still soluble in methanol, but became only partially

soluble in ethanol, and completely insoluble in acetone.

From the observation above, we proposed that dissolving the syrup into methanol

or ethanol first followed by pouring the solution in acetone to precipitate out glucose and

retain LiBr in the solution. Pure glucose syrup with a very small amount of water was

tested first. For example, 320 mg glucose, 50 mg LiBr, and 50 L water were dissolved

in 1 mL methanol at 80 ºC. The resulting solution was then added into 10 mL acetone

dropwise with constant stirring, which precipitated the glucose that could be separated

by centrifugation. The results showed that 240 mg glucose was recovered as solid and

only 4.5 mg LiBr was carried over into the solid.

Then the real sugar syrup from spruce hydrolysis was tested with this method.

The sugar syrup (containing ~280 mg glucose, ~15 mg xylose, ~60 mg mannose, 50 mg

LiBr, and 0.2 mL water) was first evaporated to reduce the water content from 0.2 mL to

0.05 mL. Then the concentrated syrup was dissolved in 1 mL methanol at 80 °C. This

solution was then added into 10 mL acetone dropwise with severe stirring to allow the

Page 155: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

140

formation of sugar precipitate. The sugar precipitate were collected by centrifugation or

filtration, leaving LiBr in the solution. Approximately 220 mg glucose and 4 mg LiBr

were found in the precipitate. In other words, 79% glucose in the syrup was recovered

and only 8% LiBr in the syrup was carried over with the sugar.

In summary, dissolving sugar-LiBr mixture in methanol followed by dropping

the solution into acetone avoided the strong interaction among sugars, water, and LiBr

and thereby separate LiBr and sugars. Addition of methanol weakened the interaction of

LiBr, sugar, and water and increased the solubility of LiBr in acetone. It turned out that

most of the residual LiBr in the syrup was dissolved into acetone and most of the

glucose in the syrup were precipitated out. Small amounts of glucose and xylose were

dissolved into organic solvent with LiBr, which still need further separation. In addition,

a large amount of acetone was needed to precipitate sugars in this method. However, it is

worthy pointing out that the dissolved sugar and LiBr in organic solvent can be directly

recycled together and used in next hydrolysis cycle without further separation.

Page 156: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

141

4.3.9.2 Crystallization of sugars in anti-solvent

Vaporizing water from the sugar syrup significantly decreased the solubility of

sugar in ethanol, making the crystallization of the sugar possible. When the syrup was

dissolved in ethanol, ethanol molecules broke the strong interactions between sugars,

water, and LiBr. Because of the high solubility of LiBr in ethanol and the low solubility

of glucose in ethanol, glucose is supposed to crystallize or be precipitated out.

Table 4.11 Effects of LiBr and water on the precipitation of glucose by ethanol

Note: Specific amount of LiBr and water (as shown in the Table), 400 mg glucose, and 2 mL ethanol were

mixed and heated to 120 C under stirring until a clear solution formed. Then solution was cooled down to

room temperature naturally under stirring and stirred for additional 20 min. Sugar precipitated out and was

filtrated and then washed with 5 mL ethanol. The precipitate then was dried at 105 C.

To verify this assumption, specific amounts of glucose, LiBr and water, as shown

in Table 4.11, was dissolved into ethanol by heating to 120 °C to facilitate the

dissolution of glucose in ethanol and thereby form a clear solution and then cooled to

Entry LiBr (mg) Water (µL) Precipitated glucose (mg)

1 50 30 330

2 50 50 330

3 50 100 330

4 50 150 330

5 100 150 320

6 150 150 300

7 200 150 250

Page 157: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

142

room temperature naturally under stirring for sugar crystallization (or precipitation).

Precipitated glucose was filtered and washed with ethanol. However, the presence of

water and LiBr increased the solubility of glucose in ethanol, and therefore glucose still

could not be fully recovered in this way. Effects of LiBr and water on the precipitation of

glucose by ethanol were investigated, and the results are shown in Table 4.11.

Figure 4.14 Sugars precipitated from solvent.

Then the sugar syrup from spruce hydrolysis was tested. The sample of the sugar

syrup containing ~280 mg glucose, ~15 mg xylose, ~60 mg mannose, ~50 mg LiBr, and

Page 158: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

143

0.2 mL water was vaporized to reduce water content from 0.2 mL to 0.05 mL. The same

operation procedure was carried out as did for the glucose-LiBr mixture above. It was

found that approximately 200 mg glucose was precipitated, as shown in Figure 4.14.

Almost all other sugars and LiBr retained in ethanol. It was also found that hemicellulose

sugars such as xylose were very difficult to crystallize. In addition, xylose may

negatively affect the crystallization of glucose.

In summary, precipitation or crystallization in anti-solvent was not an efficient

way to completely separate sugars and LiBr. Some sugars, in particular hemicellulose

sugars, retained unrecovered.

4.3.9.3 Separation of LiBr and sugars by Ion exchange chromatography

Although sugars and LiBr could be separated by ion exclusion chromatography,

as discussed above, recovered LiBr and sugars were extremely diluted (both at

concentration of ~1%), which will consume a lot of energy for reconcentration. The ion

exclusion method may be not feasible in industrial application.

To keep the sugar concentration as high as possible, ion exchange

chromatography might be an alternative method. Ion exchange resin has been widely

used to remove salt by exchanging the cation and anion in salt with the H+ and OH

- in

Page 159: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

144

resins. The disadvantage of ion exchange resin is that when H+ and OH

- are consumed,

the resin needs regeneration by flushing cation exchange resin with acid and anion

exchange resin with alkali. Obviously, directly applying ion exchange resin to raw

hydrolysate from biomass saccharification to separate LiBr from sugar is almost

impossible because of the very high concentration of LiBr, which would need huge

amounts of ion exchange resins and very large columns, and consume tremendous

amount of acid and alkali to regenerate the resins. However, when the majority of LiBr is

extracted with organic solvent, as discussed above, LiBr concentration in the sugar syrup

becomes much lower, for example ~5% of original feedstock loading (50 mg residual

LiBr/0.8 g lignocelluloses), which theoretically only needs ~3.5% sulfuric acid (28 mg)

and ~2.6% calcium hydroxide (21 mg) to regenerate the cation and anion resins. The

mechanism of whole process is shown in experimental section.

The consumption of the acid and alkali for resin regeneration would be

comparable to that in the dilute acid hydrolysis of biomass (including the neutralization

of acid). Typically, 1-5% acid loading is common in dilute acid hydrolysis or

pretreatment, and accordingly calcium hydroxide is needed to neutralize the acid to form

gypsum. In our experiment, after extraction of spruce hydrolysate with butanol/hexane,

Page 160: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

145

concentration of sugar syrup reached to 67% (1.4 g glucose, 0.075 g xylose, 0.3 g

mannose, 0.260 g LiBr, and 1 mL water), which is too thick to pass the columns packed

with the ion exchange resins. Therefore, the sugar syrup needed to be diluted to ~50%

before added onto columns.

Figure 4.15 The pictures of columns packed with cation and anion exchange resins.

Two small columns (diameter: 1 cm and length: 5 cm) were packed, one with

anion exchange resin (DOWEX-2, Cl- form, 100-200 mesh), and another with cation

exchange resin anion exchange (Amberlyst 15, 25-50 mesh), as shown is Figure 4.15.

Prior to operation, column was washed with 10 mL 5% NaOH solution followed by fresh

Page 161: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

146

water until neutral pH. Both columns were de-watered by injecting air. The syrup after

LiBr was extracted with butanol-hexane (sugar concentration: ~67%) was diluted with

water into sugar concentration of 50%. And then 0.5 mL the dilute sugar solution was

loaded into cation (or anion) exchange column and was pushed through the column by

injecting air. Recovered solution was subsequently loaded onto anion (or cation)

exchange column. Analysis of recovered sugars solution showed that LiBr was

completely removed. The sugar stream maintained the original sugar concentration of

~50% because air was used as eluent and no dilution occurred. If necessary, dry solid

sugar products could also be produced by removing water. The picture of purified

concentrated sugar solution was shown in Figure 4.16.

Figure 4.16 The picture of purified concentrated sugar solution.

Page 162: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

147

4.3.9.4 Removal of LiBr from sugar solution through precipitation

Another method to remove LiBr from sugar stream is the precipitation of salt

with sodium carbonate according to the following reaction.

2 LiBr + Na2CO3 Li2CO3 + 2NaBr

This is an easy and effective way to remove small amount of LiBr in the sugar

stream after the majority of LiBr is extracted with butanol and hexane. Since lithium

carbonate is insoluble in water, it is very easy to separate it by filtration or centrifugation.

If necessary, lithium carbonate can be converted to lithium bromide for next batch of

saccharification, according to the reaction below.

Li2CO3 + 2HBr 2LiBr + H2O + CO2

4.4 Conclusion and recommendations

Concentrated solutions of varied halide salts were tested for their cellulose

dissolution ability. It was found that LiCl, LiBr and CaBr2 solutions at a concentration of

over 60% had superior cellulose dissolution ability over other salts. It is proposed that

solvation structure/pattern of the salts significantly affects their cellulose dissolution

ability. In concentrated solutions, the cation empty coordination sites were thought to

Page 163: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

148

play an important role in coordinating cellulose hydroxyl group, thereby disrupting

hydrogen bonding. In addition, cation could act as Lewis acids to hydrolyze cellulose

and allow depolymerized oligosaccharides to be dissolved away from lignin.

Addition of acid into the concentrated salt solutions accelerated the hydrolysis of

polysaccharides and oligosaccharides into monosaccharides. It was found that the

lignocelluloses could be completely and selectively hydrolyzed into sugars quickly with

limited formation of fermentation inhibitors. Batch-feeding could reduce the ratio of salt

solution to biomass to 1:1, and thereby produce a concentrated sugar stream.

Several methods were investigated for separation of sugars and salt. Ion-

exclusion chromatography could separate sugars from the salt, but the resultant sugar

and salt streams were extremely diluted. Solvent extraction was able to separate LiBr

from sugars, but it was very difficult and costly to completely separate them. It turned

out that the combination of organic solvent extraction and ion-exchange chromatography

was a successful method to separate sugars from salt, for example, extracting 95% of the

salt from sugar solution with butanol-hexane first and removing the residual 5% salt

using ion-exchange resins.

Page 164: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

149

The results suggested that saccharification of biomass in concentrated salt

solution at moderate temperature might be a cost-effective way to produce sugars

without pretreatment and enzymatic hydrolysis. The process needs further, for example,

to create mass balance of the process, elucidate and confirm the hydrolysis mechanism

of lignocellulose in concentrated salt solutions and the role of halide salt, modify and

optimize the sugar-salt separation methods, and investigate the changes of lignin during

the saccharification. Appropriate equipment needs to be selected and evaluated for the

saccharification process to improve efficiency and reduce cost, for example, a twin-

screw extruder as the reactor for saccharification and a continuous liquid-liquid extractor

for separating sugars and salt.

Page 165: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

150

Chapter 5: Conversion of Lignocelluloses into

Hydrocarbons

5.1 Introduction

Cellulosic ethanol has been receiving a lot of attentions for years. However, its

production still faces a lot of issues. Costly pretreatment of the feedstock under severe

conditions such as high temperature and high pressure is required to remove the

recalcitrance caused by lignin and hemicelluloses prior to enzymatic saccharification of

cellulose (Demirbas, 2005). In addition, the high cost of cellulases, the low efficiency of

fermentation of pentoses, the high energy consumption for ethanol distillation, as well as

long production cycle make cellulosic ethanol economically incomparable to fossil fuels

at this stage. Furthermore, the low heating value and water-absorbent property of ethanol

make it not an ideal substitute for gasoline (Yoon et al., 2009). These issues of fuel

ethanol have been the driving force of developing the next generation liquid biofuels

from biomass. For example, converting biomass into liquid hydrocarbons which have the

same physiochemical properties as the traditional fossil fuels, attracting more and more

attention (Elliott and Schiefelbein, 1989; West et al., 2009). Hydrocarbons (gasoline,

Page 166: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

151

diesel and jet fuel) from biomass have great advantages and promise. For example,

hydrocarbons have an overall energy efficiency of 2.1 (ratio of the heating value of

alkanes to the energy required to produce the alkanes), compared to 1.1-1.3 of bioethanol

(Huber and Dumesic, 2006). In addition, limited water use for processing, short

production cycle, and the elimination of energy-intensive distillation allow low-cost

production of hydrocarbons.

There are five main pathways under investigation to convert biomass into liquid

hydrocarbon fuels: (1) biomass gasification (to syngas) followed by Fischer-Tropsch

synthesis (Kirubakaran et al., 2009); (2) biomass pyrolysis (to bio-oil) followed by

cracking and upgrading (Balfanz et al., 1993); (3) dehydration of oxygenates from

biomass under multifunctional heterogeneous catalysts followed by hydrogenation

(Chheda et al., 2007); (4) depolymerization followed by hydrodeoxgenation of lignin

(Pandey and Kim, 2011); and (5) decarboxylation of alkyl carboxyl acids such as

levulinic acid derived from hexoses followed by chain extension (Bond et al., 2010).

Of the approaches above, dehydration and hydrogenation of oxygenates derived

from biomass saccharides have garnered considerable interest for hydrocarbon

production. First, the resulting hydrocarbon fuels are the same as traditional fossil fuels

Page 167: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

152

so that modification of existing distribution infrastructure and vehicle engine is

unnecessary. Second, hydrocarbons from biomass have comparable heating value and

gas mileage as gasoline. Third, hydrocarbons are immiscible with water; therefore, the

expensive distillation step is eliminated. Fourth, as the bio-hydrocarbons are produced in

chemical ways, it allows a much shorter production period and higher feedstock

concentration than biological ethanol. Fifth, because heterogeneous catalysts can work at

higher feedstock concentration, processing water needed in hydrocarbon production can

be significantly reduced, compared to the low ethanol concentration caused by limited

ethanol tolerance of yeast. Sixth, catalysts in chemical conversion can be recycled by

simple filtration and reused for months or years, whereas the recovery of enzyme and

yeast in bioethanol production is expensive and incomplete. Besides, as most of sugar

and sugar derivatives are water soluble, the reactions can be conducted in aqueous

solution, which allows automatic separation of the final hydrophobic products from

water (Huber and Dumesic, 2006).

Extensive work has been done on saccharides-derived hydrocarbons via

hydrogenation/dehydration of sugar derivatives, such as hydroxymethylfurfural (HMF)

and furfural. A method of directly converting six-carbon sugars into hexane under a

Page 168: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

153

bifunctional catalyst was reported (Huber et al., 2004). Hydrocarbons with carbon chains

longer than five or six carbons could be prepared through aldol-condensation of

furfural/HMF and acetone, resulting in hydrocarbons with up to 20 carbons, which was

very similar to the composition of gasoline and jet fuel (Chheda et al., 2007). However,

the intermediates (HMF and furfural) could not be easily produced in high yield and at

low cost as HMF tended to polymerize and form insoluble humin in acidic aqueous

solutions (Vandam et al., 1986). Many studies were conducted to improve the selectivity

of sugar conversion into furfural/HMF. For example, dimethyl sulfoxide (DMSO) was

used to replace water, and the water-free environment promoted the dehydration of

glucose into HMF (Amarasekara et al., 2008). In another study, methyl isobutyl ketone

(MIBK) was added as an extraction solvent to collect HMF in situ when it was produced

(Roman-Leshkov et al., 2006). The HMF formed was immediately extracted into the

upper MIBK layer, which largely reduced the opportunity for the

polymerization/condensation of HMF into insoluble humin. However, the low solubility

of sugars in the organic solvents decreased the feedstock concentration and therefore the

final product concentration. In addition, it was difficult to separate HMF from the

solvent because of their similar boiling points.

Page 169: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

154

It was reported (Zhao et al., 2007) that HMF could be produced in high yields

and with high selectivity using ionic liquids as solvents with chromium halide as catalyst.

The issues related to this process are that ionic liquid is very expensive and difficult to

recover, and chromium halide is potentially toxic to environment. Partial replacement of

ionic liquid with traditional cellulose solvents was also investigated (Binder and Raines,

2009). High conversions of up to 90% were achieved from monosaccharides and pure

cellulose, whereas the yield was only 50% from corn stover powder. In addition, this

complicated system involved two expensive organic solvents (dimethylacetamide and

ionic liquid) as well as inorganic catalysts, which made them extremely difficult to

recycle. Also, the ionic liquid had high viscosity and worked well only in water-free

environment, which required the feedstock to be finely ground and completely dry.

Unfortunately, both grinding and drying of biomass are energy intensive. Currently, one

of the most potentially useful ways to produce extended hydrocarbon chain from sugars

was through aldol-condensation of furfural compounds and acetone, as reported by

Chheda et al (Chheda et al., 2007). This process used DMSO as solvent and MIBK as

extraction solvent to produce HMF and required another step to conduct the aldol-

condensation of HMF and acetone; thus the whole process was very complex. In

Page 170: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

155

addition, this process could only use soluble sugars as feedstock, which did not break the

bottleneck of utilizing lignocelluloses for cheap production of sugars.

In this chapter, a process of directly converting lignocelluloses into hydrocarbon

precursors (furfural-/HMF-acetone adducts) for liquid hydrocarbon production under mild

conditions was demonstrated. Using acetone as a solvent in presence of lithium bromide

and a small amount of hydrochloric acid and water, the process integrated the hydrolysis

of polysaccharides (both cellulose and hemicelluloses) to monosaccharides, dehydration of

the monosaccharides to HMF (from hexoses) and furfural (from pentoses), and aldol

condensation of HMF and furfural with acetone into a single step to produce hydrocarbon

precursors. The process was abbreviated as HDA (Hydrolysis-Dehydration-Aldol

condensation). The hydrocarbon precursors from the process could be easily hydrogenated

to hydrocarbons with chain lengths in the range of C5-C21. Meanwhile, lignin in the

biomass was depolymerized and dissolved in acetone in the process. The dissolved lignin

could be easily separated when the acetone was evaporated for recycling. The lignin was

expected to have considerable potential for co-products development. The acetone and

LiBr could be recycled and reused in next batch.

Page 171: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

156

5.2 Experimental

5.2.1 Chemicals and materials

Various types of biomass feedstocks were used in the present research, including

softwood spruce, hardwood poplar, crop residue corn stover, energy crop switchgrass, and

paper (newspaper and print paper). After being air-dried, the biomass samples were

ground to pass a 40-mesh screen using a Wiley mill. Sources and chemical composition of

the biomass samples are presented in Table 4.1. Chemical reagents used in this study were

purchased from Fisher Scientific (Pittsburgh, PA) or Sigma-Aldrich (St. Louis, MO) and

used as received.

5.2.2 Production of hydrocarbon precursors from biomass

Exact amounts (see Table 5.1) of LiBr and hydrochloric acid, 1 mL acetone and

100 μL water were loaded into a 15-mL vial and vortexed to mix well. To this solution

100 mg 40 mesh biomass powder or sugar was added and the mixture was vortexed again.

The vials was heated at 80-120 ºC in an oil bath and stirred with magnetic stir bar at 500

rpm for 2 h. At the end of reaction, acetone in the mixture was carefully removed at low

temperature (50 C) using a rotary evaporator. The residue was extracted with CH2Cl2 to

collect hydrocarbon precursor products. The residue after the extraction, consisting of

Page 172: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

157

liquid (water + LiBr) and solid (lignin and unreacted biomass), was filtered and washed

with water. The filtrate was collected for analysis of saccharides, HMF, furfural, and

residual LiBr and HCl. The solid residue was redissolved in acetone and filtered. The

filtrate was dried to produce acetone-soluble lignin, and the residue after filtration was

dried for composition analysis.

5.2.3 Determination of residual LiBr

Procedure is shown in Section 2.2.7.

5.2.4 Determination of sugars and sugar derivatives

Procedure is shown in Section 2.2.6.

5.2.5 Qualitative analysis of hydrocarbon precursors using GC-

MS

GC-MS was used to identify components in the products. Approximately 5 mg

product was dissolved into 1 mL CH2Cl2 for GC-MS analysis. GC-MS was performed on

a GC-MS-QP 2010 instrument (Shimadzu Co., Addison, IL) equipped with a 30 m × 0.25

mm i.d., 0.25 m film, SHR5XLB capillary column. Helium was used as carrier gas at a

flow rate of 1 mL/min. GC conditions were as follows: initial column temperature, 100 °C,

Page 173: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

158

held for 1 min, ramped at 2 °C/min to 310 °C, and then held for 4 min; injector

temperature, 250 °C; split ratio, 1/20, EI mode at 20 kV for ionization.

5.2.6 Quantitative analysis of hydrocarbon precursors using ESI-

MS and oxidation methods

(1) ESI-MS analysis

ESI-MS in positive-mode was used to quantify the products on an Applied

Biosystems 3200 QTRAP instrument. Product (10 mg) was dissolved in 10 mL

methanol/CHCl3 (1:1, v/v) to prepare 1 mg/mL product solution. Furfural-acetone (1 mg)

was dissolved in 10 mL methanol/ CHCl3 (1:1, v/v) to prepare 0.1 mg/mL standard

solution, which was further diluted into 0.01 mg/mL. Combining 0.5 mL the product

solution and the 0.5 mL 0.1 mg/mL standard solution led to a sample with product

concentration of 0.5 mg/mL and standard concentration of 0.05 mg/mL.

(2) Oxidation analysis

i. Preparation of dimethyldioxirane in acetone: A 250 mL two-necked round-

bottomed flask was connected by a U tube to a receiving flask. The receiving flask was

connected to a vacuum pump. Receiving flask was cooled at -80 °C by means of liquid

nitrogen/ethanol bath. The reaction flask was loaded with a mixture of water (50 mL),

Page 174: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

159

acetone (40 mL), and NaHCO3 (15 g) and cooled to 5-10 °C with the help of an

ice/water bath. With vigorously stirring and cooling, solid oxone (40 g) was added in

five portions at 3-min intervals. Waiting for 3 min after the last addition, the ice/water

bath was removed and a moderate vacuum (80-100 Torr) was applied.

Dimethyldioxirane/acetone solution was distilled and collected in the cooled receiving

flask. The volume of dimethyldioxirane (5% yield) in acetone was approximately 40 mL.

ii. Oxidation of the furan-derived hydrocarbon precursors with dimethyldioxirane:

to avoid the oxidation of hydroxyl in the hydrocarbon precursors into carbonyl group,

the hydrocarbon precursors was acetylated with acetyl bromide in CH2Cl2/pyridine prior

to oxidation. After removing CH2Cl2/pyridine, the acetylated hydrocarbon precursors

(from 100 mg sugars) was dissolved in acetone and reacted with excessive

dimethyldioxirane in acetone for 30 min at room temperature. After the completion of

the reaction, acetone was removed; the oxidized hydrocarbon precursors were dissolved

in 50 mL ethanol and ready for coloration assay.

iii. Determination of carbonyl content in the hydrocarbon precursors: The original

or oxidized hydrocarbon precursors in ethanol (1 mL) was reacted with 1 mL 5% (w/v)

dinitrophenylhydrozine in acidic ethanol at 55 °C for 20 min and then cooled in an ice

Page 175: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

160

bath. Then 8 mL of 5% (w/v) KOH in ethanol was added and the mixture was

centrifuged to separate precipitant (KCl) and red wine colored supernatant. The

supernatant was taken to determine the carbonyl content using UV at 432 nm.

5.2.7 Estimation of lignin molecular weight

Gel permeation chromatography (GPC) was used to estimate the molecular weight

of lignin. GPC was performed with a Viscotek GPCmax VE-2001 chromatograph using

two columns (VARIAN 5M-POLY-008-27 and VARIAN 5M-POLY-008-32). THF was

used as an eluent with a flow rate of 1 mL/min at 30 ºC. Monodispersed Polystyrene

standards were used for calibration.

5.2.8 Characterization of LiBr/acetone solvent systems

FT-IR was used to investigate the bonding changes caused by the interactions

between acetone/water and LiBr. Spectra were recorded on a PerkinElmer Spectrum 100

FT-IR spectrophotometer with universal attenuated-total-reflection (ATR) sampling

accessory (Waltham, MA).

5.3 Results and Discussion

5.3.1 Description of the HDA process

Page 176: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

161

The proposed reaction pathway for direct conversion of biomass to hydrocarbon

precursors is shown in Figure 5.1.

Figure 5.1 Reaction pathway of biomass to hydrocarbon precursors in HDA process (Structures

of R2 are shown in Table 5.6).

The experimental flowchart of the HDA process of converting biomass into

hydrocarbons is shown in Figure 5.2. Biomass feedstock was fed with acetone, water,

salt and acid, into reactor. Reaction was conducted under 100-140 ºC for 1-4 h. After

reaction, unreacted acetone in final liquor was recycled through vaporization. Left

residues including hydrocarbon precursors, salt, acid and water were suspended into

certain amount of CH2Cl2. Extra fresh water was added to facilitate the separation of

Page 177: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

162

LiBr and wash CH2Cl2 and lignin. Suspension was centrifuged to separate CH2Cl2 layer

from insoluble lignin and water. CH2Cl2 was recovered through vaporization, and the

hydrocarbon precursors was obtained as thick liquid or solid. The LiBr/water phase was

concentrated and reused. Hydrocarbon precursor from CH2Cl2 was converted into

hydrocarbon through hydrodeoxygenation on bifunctional catalysts.

Biomass

Fuel

1 23

45

6

7 8

a b

c

de

fg

h

i

jk

1

2

3

4

56

7

8

9 Lignin

11

10

4

l

mn

Figure 5.2 Process flow chart of converting lignocelluloses into hydrocarbons. Main facilities:1.

Hydrocarbon precursor synthesis reactor; 2. Acetone evaporator; 3. Centrifuge; 4. Water

evaporator; 5. CH2Cl2 evaporator; 6. Hydrodeoxygenation reactor; 7.Products tank; 8. Hydrogen

gas tank; 9. Acetone tank; 10. Extraction solvent (CH2Cl2) tank; 11. Water tank. Mass flow: a.

Hydrocarbon precursors, acetone, LiBr, water and acid; b. Hydrocarbon precursors, LiBr, water

and acid; c. CH2Cl2 and hydrocarbon precursors; d. Hydrocarbon precursors; e. Hydrocarbons; f.

Acetone; g. Acetone; h. CH2Cl2; i. Water; j. Diluted acidic LiBr solution; k. Concentrated acidic

LiBr solution; l. Hydrogen gas; m. CH2Cl2; n. Water.

Page 178: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

163

5.3.2 LiBr/water and LiBr/acetone systems

The interaction of LiBr with different solvents is illustrated in Figure 5.3. In

aqueous solution, polarized -OH bond of water induces partial positive charge on

hydrogen atom that has affinity to Br- and partial negative charge on oxygen that attacks

Li+. Therefore, both Li

+ and Br

- are solvated by water molecules (Figure 5.3a), which not

only separates Li+ and Br

-, but also limits Br

- from attacking positively charged positions.

From mechanism of traditional cellulose solvents of LiBr/DMSO (Figure 5.3b) and

LiBr/DMF (Figure 5.3c), it can be seen how free halide ion is crucial to dissolving

cellulose by destroying hydrogen bonds in cellulose. Different from water, only Li+ is

able to coordinate closely with electronegative O in DMSO or DMF, while Br- is

relatively free in the solvent. The reason is that the sterically hindered positive position

(C or S atom) by neighboring groups cannot interact with Br-. Free halide ion plays a

significant role in destroying the hydrogen bonds and promotes the dissolution of

cellulose in ionic liquid. However, these solvents are expensive, toxic, and hard to

recover due to high boiling point, which excludes or limits their industrial potential as

solvents in directly hydrolyzing lignocelluloses into monosaccharides.

Page 179: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

164

Figure 5.3 Interaction of LiBr with different solvents. (a) LiBr in H2O; (b) LiBr in DMSO; (c)

LiBr in DMF or DMAC; (d) LiBr in acetone; (e) ionic liquid (1-Butyl-3-methylimidazolium

bromide).

In addition to these extensively investigated polar aprotic solvents, acetone has

similar structure, as illustrated in Figure 5.3d, and theoretically can coordinate Li+ and

leave Br- free. Besides, acetone has other advantages over DMSO and DMF. First, the

low boiling point makes acetone very easy to separate and recover from the reaction

HO

H

Li+

Br- Br-

O

H

H

H

H

H

H

Li+

S

O

H

H

H

H

H

H

Li+

R N

O

Li+

a: LiBr/H2O

d: LiBr/acetone

c: LiBr/DMF or LiBr/DMACb: LiBr/DMSO

NN Br-

Br- Br-

Br-

f: ionic liquid(1-Butyl-3-methylimidazolium bromide)e:

Page 180: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

165

system. The most important one is that acetone can react in situ with newly generated

furfural and HMF through aldol condensation reactions to form the precursors of liquid

hydrocarbons with extended carbon numbers (Figure 5.3e). In addition, the aldol

condensation reduces or prevents the self-condensation of furfural and HMF to humin.

Furthermore, compared to DMSO and DMF, acetone is inexpensive, less toxic, and of

low viscosity. It is also a good solvent for lignin.

165016701690171017301750

wavenumber (cm-1)

acetone+0

acetone+100

acetone+200

acetone+250

acetone+300

acetone+400

acetone+500

acetone+600

Figure 5.4 FT-IR spectra of LiBr/acetone solution.

In order to verify the proposed interaction between LiBr and acetone and the

formation of loosely attached Br-, solutions of LiBr in acetone were studied by FT-IR. In

Page 181: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

166

Figure 5.4, the band at 1715 cm-1

represents the stretching vibration of pure acetone.

Because of the interaction of acetone carbonyl group with Li+, partial electron on C=O

bond was donated to Li+, which weakened the C=O bond. The reflection of this on FT-

IR spectra is that a red shift of the peak to low wavenumber region occurred. Reduced

positive charge on Li+ caused by the electron donation from C=O bond weakened the

attraction force between Li+ and Br

-. Thus, loosely attracted or free Br

- was formed in

this situation. The free Br- could effectively catalyze xylose and glucose to furfural and

HMF, respectively, which will be discussed later. It can be seen from Figure 5.4 that

with the increasing of LiBr, peak red shifted more, indicating formation of more loosely

attached Br-.

5.3.3 One-step conversion of biomass into hydrocarbon

precursors

Different types of feedstock, including monomeric saccharides (glucose, xylose,

arabinose, galactose and mannose), polysaccharides (cellulose and starch), papers (filter

paper, newspaper, and print paper), and lignocelluloses (spruce, poplar, corn stover, and

switchgrass), were treated with the HDA process, and the results are summarized in

Table 5.1.

Page 182: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

167

The first 14 experiments were conducted with glucose as feedstock to investigate

the effect of LiBr and acid dosage on the conversion of glucose. With the increase of

LiBr or acid dosage, glucose conversion increased. Addition of 300 mg LiBr appeared to

be a turn point. The conversions of 70-80% might be caused by the shortage of effective

free Br- because the amount of LiBr was below 300 mg. However, the conversion

decreased when LiBr dosage increased beyond 300 mg, which was probably caused by

the high viscosity of reaction solvent and the formation of humin. With the same amount

of LiBr, Entry 13 had a higher conversion than Entry 3 because more acid was added.

Entry 4 and Entry 14 had the same conversion though Entry 14 had a higher acid dosage

than Entry 4, implying that 0.25% acid was sufficient to complete a high conversion

without the need of more acid loading in Entry 14. Conversions of Entry 2 and Entry 3

were far lower than that of Entry 4, which might indicate that LiBr dosage less than 200

mg was unable to produce enough free Br- as catalyst (as discussed above, in the

presence of water, most of Br- ions were surrounded and solvated by water). It is worth

pointing out that ~70% conversion was observed even without any addition of LiBr

(Entry 1) whereas no HMF was detected in the final liquor, which was attributable to the

formation of ketal through reaction of acetone with sugar (Pfaff, 1987).

Page 183: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

Table 5.1 Conversion of lignocelluloses or sugars into hydrocarbon precursors in acetone/LiBr system

Note: (1) Other reaction conditions: 1 mL acetone, 100 μl water, 120 ºC, 2 h. (2) Acid loading, w% based on weight of solvent. (3) Glu-glucose; Ara-

arabinose; Man-mannose; Xyl-xylose; Cell-cellulose; FP-filter paper; PP-print paper; Pop-poplar; NP-newspaper; CS-corn stover; and SG-switchgrass. (4)

Soluble solid-CH2Cl2 soluble hydrocarbon precursor; insoluble solid-acetone-insoluble residues; residual sugars-sugars left in solution after reaction. (5) A/F

ratio-The molar ratio of reacted acetone to converted sugars.

Entry LiBr

(mg)

Acid

(%,

w/w)

Feedstock

(mg)

Soluble

solid (mg)

Insoluble

solid (mg)

Residual

sugars (mg)

Reacted

acetone (mg)

Sugar

conversion (%)

A/F

ratio

1 0 1 100 (Glu) 46 0 30 13 70 0.6

2 100 0.25 100 (Glu) 50 0 25 21 75 0.9

3 200 0.25 100 (Glu) 64 0 21 25 79 1.0

4 300 0.25 100 (Glu) 83 0 3 39 97 1.3

5 400 0.25 100 (Glu) 92 0 5 45 95 1.5

6 500 0.25 100 (Glu) 105 0 10 55 90 1.9

7 500 0.20 100 (Glu) 99 0 10 51 90 1.8

8 500 0.15 100 (Glu) 87 0 16 46 84 1.7

9 500 0.10 100 (Glu) 63 0 20 31 80 1.2

10 500 0.05 100 (Glu) 51 0 30 28 70 1.3

11 750 0.00 100 (Glu) 43 0 36 4 64 0.2

12 100 0.50 100 (Glu) 67 0 26 34 74 1.5

13 200 0.50 100 (Glu) 135 0 8 97 92 3.3

14 300 0.50 100 (Glu) 155 0 2 136 98 4.4

15 300 0.25 100 (Ara) 110 0 4.3 81 96 2.3

16 300 0.25 100 (Gal) 96 0 14 63 86 2.3

17 300 0.25 100 (Man) 93 0 11.5 57 88 2.1

18 300 0.25 100 (Xyl) 99 0 4.5 68 96 1.9

19 300 0. 50 100 (Cell) 156 0 0 121 98 3.0

20 300 0.75 100 (Xyl) 104 7 0 113 90 2.7

21 300 0.25 100 (Star) 98 0 0 46 98 1.3

22 300 0.5 100 (FP) 141 0 0 92 99 2.6

23 300 2 100 (PP) 50 2 0 34 95 1.1

24 300 1 100 (NP) 113 8 0 53 96 2.0

25 300 0.5 100 (pop)) 95 1 0 76 99 3.8

26 300 0.5 100 (CS) 62 1 0 55 99 2.8

27 300 0.5 100 (spr) 87 1 0 72 99 3.6

28 300 0.5 100 (SG) 74 2 0 61 99 3.5

168

Page 184: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

169

Acid promoted the formation of acetone self-condensation product.

Comparing Entry 2-4 and Entry 12-14, with the same LiBr dosage, the later had much

higher glucose conversion than the former, implying that the acid promoted/catalyzed

the reactions. In addition, higher acid loading enhanced the consumption of acetone

caused by self-condensation. In Entry 11, acetone was hardly consumed without the

addition of the acid. In summary, it appeared that 300 mg LiBr with 0.25% acid was

enough to ensure efficient conversions of sugars.

Other saccharides occurring in lignocellulose, including arabinose, galactose,

xylose, and mannose, were studied at the same conditions (300 mg LiBr and 0.25%

acid). The results indicated that these sugars could be converted as easily as glucose.

In addition to the monomeric saccharides above, the HDA process was applied

to polysaccharides and real lignocelluloses at the similar conditions studied above.

Considering that different feedstocks had varied recalcitrance to the reactions, the acid

loading varied with different feedstock. Conversion of 99% was achieved when 0.5%

acid was used for cellulose micron and filter paper. Starch, because of its amorphous

structure, only needed 0.25% acid for a high conversion. However, xylan needed

more acid to get a high conversion probably due to the dry and tough feature of

extracted xylan. Because the inorganic fillers in print paper and newspaper neutralize

acid, they needed more acid to achieve a high conversion (1 and 2% for newspaper

Page 185: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

170

and print paper, respectively). For the lignocelluloses investigated (softwood spruce,

hardwood poplar, agricultural residue corn stover, and energy crop switchgrass), 0.5%

acid was enough to break down lignin and convert cellulose and hemicellulose to

sugars for further reaction. A/F ratio (molar ratio of condensed acetone to condensed

furfural or HMF) was used to estimate the average carbon number of hydrocarbon

precursors. High A/F ratio indicates more acetone connected to furfural or HMF, thus

longer chain in final hydrocarbon precursors; low A/F ratio indicates less acetone

connected to furfural or HMF, thus shorter chain hydrocarbon. Although this ratio

was highly dependent on the conversion and acid concentration, it could be used as an

indicator of the extent of acetone self-condensation. For example, higher acid

concentration (>0.5%) gave an A/F ratio larger than 2, indicating a severe acetone

self-condensation, while lower acid concentration (0.25%) resulted in an A/F ratio

less than 2, indicating that acetone condensation was significantly reduced.

The pictures of the product solutions from different feedstocks are shown in

Figure 5.5. The control experiment (acetone and LiBr only) gave a light brown liquid.

The color was supposed to be from the acetone self-condensation products. On the

other hand, the experiments with xylose, glucose, and spruce powder generated

solutions in dark brown or black. The color was probably from the derivatives or

condensation products of furfural or HMF. In all cases, the feedstocks dissolved and

Page 186: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

171

formed homogeneous solutions at the end of reaction. After vaporizing acetone, the

hydrocarbon precursors could be easily extracted by methylene dichloride to recycle

catalyst (in aqueous phase), as show in Figure 5.5e.

Figure 5.5 The hydrocarbon precursors from HDA process in acetone. (a) Acetone self-

condensed products; (b) Hydrocarbon precursors from xylose; (c) Hydrocarbon precursors

from glucose; (d) Hydrocarbon precursors from softwood spruce; (e) Hydrocarbon precursors

extracted by CH2Cl2 (bottom layer was water).

5.3.4 Effect of temperature and time on conversion of spruce

powder

Effect of temperature on conversion of spruce powder was investigated in the

range of 80-120 ºC. The results in Table 5.2 indicated that the conversion of glucose

decreased with the decreasing of temperature; however, high conversion could be

achieved by extending the reaction time at lower temperature, which can be seen from

Table 5.3. For example, the conversion of spruce powder was only 67% at 100 ºC for

a b c d e

Page 187: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

172

2 h, compared to the 100% conversion at 120 ºC for 2 h. The conversion at 100 ºC

increased to 88% at 3 h and 100% at 4 h, respectively.

Table 5.2 Effect of temperature on conversion of spruce powder

Note: other reaction conditions: 300 mg LiBr, 100 mL water, 1 mL acetone, 100 μL water, 100 mg

spruce powder, 2 h.

Table 5.3 Effect of time on conversion of spruce powder

Note: other reaction conditions: 300 mg LiBr, 100 mL water, 1 mL acetone, 100 μL water, 100 mg

spruce powder, 100 ºC.

5.3.5 Effect of different salts on conversion of spruce powder

The catalysis abilities of different common and inexpensive halide salts on the

conversion were investigated. The results in Table 5.4 shows that CaBr2, NaBr, LiBr

and LiCl had almost the same catalysis abilities in the same reaction time. On the

other hand, transitional metal halide including CuCl2, FeCl3 and ZnCl2 did not give as

good results as alkali metal halide did. The reason probably is that d orbital of

transition metals ion still formed bonds with halide ion in spite of the presence of the

interaction with C=O, which reduced the releasing of free halide ions. Malfunction of

Temperature (°C) Conversion (%)

80 58

100 67

120 99

Time (h) Conversion (%)

1 58

2 67

3 88

4 99

Page 188: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

173

LiI might be attributed to its instability, as LiI tends to decompose into LiOH and

iodine. AlCl3 is not an ionic compound that cannot form free Cl- in solution, thus no

obvious change was observed in reaction process.

Table 5.4 Effect of different halogen salts on conversion of spruce powder

Salt Left solid (mg) Carbohydrates Conversion (%)

CaBr2 15 99

NaBr 26 97

NaBr (200mg water) 20 99

LiCl (200mg water) 25 99

ZnBr2 52 65

FeBr3 52 65

MgCl2 40 81

CuCl2 25 /

FeCl3 29 /

CaCl2 41 80

ZnCl2 59 50

LiI Doesn’t work

AlCl3 Doesn’t work

KBr Doesn’t work

Note: other reaction conditions: 300 mg salt, 100 μL water, 1 mL acetone, 100 mg spruce powder, 0.25%

(w/w) HCl (pure hydrogen chloride, based on weight of solvent), 2 h, 120 ºC.

5.3.6 Effect of different acids on conversion of spruce powder

The catalysis abilities of different common mineral acids were investigated.

The results are shown in Table 5.5. It was observed that the conversion of glucose

decreased with decreasing acidity. The conversion reaction proceeded faster when

strong acids were added. Formic acid could only convert 54% spruce powder, but it is

expected that longer reaction time might be able to get a higher conversion. Acetic

acid as weak acid did not result in significant change in the conversion process.

Page 189: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

174

Table 5.5 Effect of different acids on conversion of spruce powder

Salt Left solid (mg) Carbohydrates conversion (%)

H2SO4 17 99

HNO3 25 99

H3PO4 12 99

HCOOH 54 54

CH3COOH / /

Note: other reaction conditions: 300 mg LiBr, 100 μL water, 1 mL acetone, 100 mg spruce powder, 2 h,

0.25 % (w/w) acid (pure acid, based on weight of solvent), 120 ºC.

5.3.7 Identification of products

GC-MS was used to identify the products and verify the validity of the

proposed reaction pathway. The proposed mechanisms of acetone self-condensation

and aldol-condensation between furfural and acetone are shown in Figures 5.6 and 5.7,

respectively. From the pathway shown in Figure 5.6, it is expected that the molecular

weight of acetone self-condensation products should be n1 40 + n2 58 (n1, n2=0, 1,

2…, n1 + n2 is the number of condensed acetone) because of possible intramolecular

dehydration. Similarly, molecular weight of HMF-acetone condensation products

should be 126 + n1 40 + n2 58 (n1, n2 = 0, 1, 2…, n1 + n2 is the number of

condensed acetone), while molecular weight of furfural-acetone condensation product

should be 96 + n1 40 + n2 58 (n1, n2=0, 1, 2…, n1 + n2 is the number of condensed

acetone). The products would be a mixture of the products from acetone self-

condensation and HMF- and furfural-acetone condensation with varied molecular

weight. The proposed structures and their molecular weights from these proposed

Page 190: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

175

mechanisms is shown in Table 5.6. The GC-MS and positive ESI-MS of acetone self-

condensation products with peaks in black, HMF-acetone condensation products in

red, furfural-acetone condensation products in blue, are shown in Figures 5.8 and 5.9,

respectively. As expected, molecular weight of most peaks observed from GC-MS

agreed with the number calculated from the formulas proposed above, which verified

the proposed reaction pathways. In positive ESI-MS graphics, as an ion was produced

by the addition of a proton, so the observed m/z was the value of molecular weight

plus one. Based on this, most peaks had the m/z fitting the proposed formulas.

Figure 5.6 Proposed self-condensation mechanism of acetone.

Figure 5.7 Proposed condensation mechanism of furfural with acetone.

Page 191: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

176

Table 5.6 Proposed structure of products and corresponding molecular weight

Note: ‖R‖ represents side chain of furan ring (see Figure 5.1).

In Figure 5.8a, MW of acetone self-condensation products ranged from 98 to

298. The former was the MW of the dehydration products of 2 acetone molecules, and

the latter was the MW of the dehydration products of 7 acetone molecules. Between

them, the peaks of 138, 178, 218, 258, and 298 were attributed to dehydration

products of 3, 4, 5, and 6 acetone molecules, while those of 120, 160, 200, 240, and

280 were the MWs of the ring structures formed from the compounds of 138, 178,

218, 258, and 298 after further dehydration (Figure 5.8 and Table 5.6).

Condensed

acetone

R (double bond could be

added by water)

MW of HMF-acetone

adduct

MW of furfural-acetone

adduct

1

166, 184, 136, 154,

2

206, 224, 242, 176, 194, 212

3

246,264 216,234

4

286,304 256,274

O

O

O

O

O

O

Page 192: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

177

a) Acetone + acetone

c) Furfural + acetone

b) HMF + acetone

d) Glucose + acetone

Page 193: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

178

Figure 5.8 GC-MS spectra of hydrocarbon precursors in CH2Cl2.

For HMF-acetone condensation products, Figure 5.8d shows that the

molecular weights were from 126 for HMF to 246 for HMF + 3 acetone - 3 H2O. The

peaks of 166 and 206 were the MWs for HMF + 1 acetone – 1 H2O and HMF + 2

acetone – 2 H2O, respectively.

For xylose-acetone condensation products that are shown in Figure 5.8e, 136

was the MW for furfural + 1 acetone – 1 H2O, followed by 176, 216, and 256 that

were attributed to the MWs for furfural + 2 acetone – 2 H2O, furfural + 3 acetone – 3

H2O, and furfural + 4 acetone – 4 H2O, respectively. Products with incomplete

f) Spruce + acetone

e) Xylose + acetone

Page 194: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

179

dehydration were also observed, which were 234 (furfural + 4 acetone – 3 H2O), and

274 (furfural + 5 acetone – 4 H2O), respectively.

For spruce-acetone condensation products, only one condensation product

derived from pentose (furfural-acetone) was observed at 136, the MW of dehydration

product of 1 acetone and 1 furfural. Peaks for hexose-derived products observed were

the condensation products of 1 HMF with 1, 2, and 3 acetone. The peak at 184 was

the condensation product of 1 HMF and 1 acetone without dehydration.

The difference between GC-MS and ESI-MS spectra was that peaks with

higher m/z in ESI-MS spectra were observed, but they were weak and not clear when

m/z was above 300. With m/z below 300, most peaks were from condensation of 1

furfural with a few acetones. Most of the peaks with m/z above 300 observed were

from the condensation of 2 furfural with acetone. Statistically, this happened with

very low possibility since the concentration of acetone was much higher than that of

furfural, and each furfural molecule was surrounded by a large amount of acetone

molecules, as indicated by the very small amount of products with molecular weight

above 300.

Page 195: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

180

Figure 5.9 ESI-MS spectra of hydrocarbon precursors in CH2Cl2.

When molecular weight was below 300, carbon number was estimated to be 1-

20. Since these components had the same response to detector, their intensities in

spectra reflected their relative contents (see next section). From Figure 5.9, it can be

seen that intensity of components showed a normal distribution, statistically, which

Page 196: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

181

matched the reaction chances of furfural with acetone. In this reaction, the longer the

chain was, the smaller the chance of production would be. Thus, the average carbon

number would be around 12, which was qualitatively given by the largest peak in

ESI-MS graphics, which was the product of 1 HMF + 2 acetone.

In order to verify the proposed furfural acetone condensation pathway, GC-

MS of pure HMF and furfural as feedstock were investigated. When 100 mg HMF

reacted with 1 mL acetone, we found that large amount of humin formed in 30 min at

120 C. Thus, loading of HMF and furfural was decreased to 10 mg, and the reaction

was conducted at 100 C for 30 min. The GC-MS results are show in Figures 5.8b and

5.8c. Furfural showed the peak of 1 furfural + 1 acetone with MW of 136. HMF also

showed the peak of 1 HMF + 1 acetone product. These peaks verified the formulas for

calculating products MW, proposed above.

5.3.8 Quantification of conversion by ESI-MS

The complexity and diversity of the hydrocarbon precursors formed in HDA

process made the quantification of the yield and selectivity of the process very

difficult. Fortunately, the furan rings in furfural or HMF kept unchanged during the

aldol-condensation; therefore the selectivity of sugars to the hydrocarbon precursors

could be estimated through quantifying the furan rings. Two methods (2D NMR and

Page 197: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

182

oxidation, respectively) were developed for estimating reaction selectivity, as shown

in Figures 5.10 and 5.11, respectively.

(a)

(b)

Page 198: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

183

(c)

Figure 5.10 Quantification of the hydrocarbon precursors from lignocelluloses using 2D-

NMR method (internal standard: pyrazine; (a) glucose; (b) xylose; and (c) spruce (softwood)).

The results of the sugar conversion and selectivity of sugar to the hydrocarbon

precursors of the HDA process are summarized in Table 5.7. Both glucose and xylose

had a conversion of ~95% and selectivity of 85-90%, respectively, which resulted in

an overall yield of ~80% from the sugars to the furan-derived hydrocarbon precursors.

When real biomass (softwood spruce) was used, ~85% sugar conversion and ~85%

selectivity were obtained, respectively, leading to an overall hydrocarbon precursor

yield of approximately 72%. The two methods gave comparable results, whereas the

2D NMR gave slightly higher values than the oxidation method, which might be

attributed to the over-integration of hydrogen peaks that did not belong to the furan

ring in the rectangle area of integration (Figure 5.10).

Page 199: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

184

Figure 5.11 Mechanism of quantifying hydrocarbon precursors by oxidation method

Table 5.7 Quantitation of hydrocarbon precursors derived from carbohydrates

Feedstock Conversion (%)a

Selectivity (%)b

2D NMR Oxidation

Glucose 95 90 86

Xylose 93 91 83

Spruce 85 88 84 Note: (a) Conversion represents the molar percentage of converted (consumed) carbohydrates based on

initial carbohydrate; (b) Selectivity represents the molar percentage of the furans formed from the

converted carbohydrates.

Ooxone, NaHCO3

5~15oC

o o

O O

HO

O O

OAc

o

o

O

OAc

O O

NO2

O2N

NHH2N

OAc

N N

NO2

O2N

NH

NO2

O2N

HN

AcBr

N

NO2

O2N

NH

Page 200: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

185

5.3.9 Carbon number distribution of hydrocarbon precursors

The carbon number and distribution of the hydrocarbon precursors were

estimated from positive ESI-MS spectra. Carbonyl group is easier to be protonated

than the hydroxyl or oxygen on furan in the hydrocarbon precursors because of its

larger proton affinity, as shown in Table 5.8. In particular, when a double bond

existed in the hydrocarbon precursors, the ketone would have a substantially higher

proton affinity than regular ketone because the formed positively-charged center

([M+1]+) could be stabilized through delocalization. As every hydrocarbon precursor

molecule was a ketone or double bond conjugated ketone (after dehydration), all the

hydrocarbon precursors were ionized in the same way and had the same response to

detector. Therefore, the ion intensity of each peak on the spectra reflected the relative

content of the responding hydrocarbon precursor compound. According to the ESI-

MS spectra shown in Figure 5.9, the carbon numbers of the identified hydrocarbon

precursors from glucose, xylose and spruce were estimated and summarized in Figure

5.12. It can be seen that the carbon numbers of most hydrocarbon precursors were

between 5 and 21, indicating that 0-3 acetone molecules were condensed to furan

rings. The precursor with more than 3 condensed acetone molecules were not detected.

The carbon numbers of the hydrocarbon precursors fell into the range of

Page 201: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

186

transportation fuels. In addition, the carbon numbers seemed to follow a normal

distribution with C11/C12 the highest frequency.

Table 5.8 Proton affinity of different functional groups in hydrocarbon precursors

Functional groups Proton affinity (kJ·mol-1

)

R OH

760~800

(Jolly, 1991) O

~820

(Jolly, 1991)

O

~803

(Pan et al., 2010)

O

~870

(Bouchoux et al., 1988)

Note: Definition of proton affinity: the proton affinity, Epa, of an anion or of a neutral atom or

molecule is a measure of its gas-phase basicity. It is the energy released in the following

reactions: B + H+ → BH

+ (Jolly, 1991).

Figure 5.12 Carbon number distributions of hydrocarbon precursors from different

feedstocks.

C6

126

C9

126

C12

126

C15

126

C18

126

C21

126

C8

126

C11

126

C14

126

C17

126

C20

126

C9

126

C12

126

C15

126

C18

126

C5

126

Page 202: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

187

5.3.10 Decomposition of lignin during HDA process

In many small-scale experiments above various types of lignocelluloses were

used as feedstock, no insoluble residue was observed at the end of reaction, implying

that lignin was completely soluble in acetone. In addition, the residue after CH2Cl2

extraction was far less than theoretical lignin content, indicating that part of lignin

was soluble in CH2Cl2. The evidence above suggests that the lignin be significantly

degraded during the HDA process as natural lignin is insoluble in acetone or CH2Cl2.

The acetone-soluble part of lignin was investigated for its molecular weight by

GPC. As seen from Figure 5.13, lignin was severely depolymerized into small

fragments with weight average molecular weight around 1000. In other words, the

lignin consisted of approximately only 4-5 monomeric units. The number molecular

weight of the lignin was around 400. Considering the mild condition (120-140 C) of

HDA process, there must have been a unique delignification mechanism involving

LiBr, mineral acid and acetone. This will be investigated in the future research. In

addition, the fraction soluble in CH2Cl2 are supposed to have even smaller molecular

weight than the acetone-soluble fraction. Further study will be conducted to try to

characterize the CH2Cl2-soluble lignin as well.

Page 203: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

188

Elution time (min)

Figure 5.13 Gel permeation chromatograph of HDA lignin from spruce. Note: Mn represents number-average molecular weight; Mw represents weight-average molecular

weight

5.3.11 Reaction mechanism of glucose to HMF

The results above (Table 5.1) shows that both acid and Br- played a very

important role in catalyzing sugars to furfurals, which can also be seen from Figure

5.14. It was reported (Aida et al., 2007; Qian et al., 2005) that six-member ring of

glucose experienced a ring opening followed by a ring closing to form five-member

ring of fructose as intermediate in the dehydration process to form HMF. In order to

investigate whether Br- was involved in the dehydration of fructose, converting

fructose to HMF under different conditions was studied. Several test experiments

showed that fructose was dehydrated very quickly so that the reaction temperature

was lowered down to 75 C in 30 min to get a decent data set, as shown in Figure 5.15.

Retention time 21 min

Mw 1000

Mn 400

Page 204: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

189

Note:other reaction condtion: 100 ºC, 1 mL acetone, 100 μL water, 100 mg glucose.

Figure 5.14 Dehydration of glucose to HMF derivative at different LiBr concentrations.

Note:other reaction condtion: 75 ºC, 1 mL acetone, 100 μL water, 100 mg fructose.

Figure 5.15 Dehydration of fructose to HMF derivative at different LiBr concentrations.

According to Arrhenius equation ( k = e−Ea RT ), the dehydration rate of

fructose was 100 times faster than dehydration of glucose. In addition, the dosage of

0

20

40

60

80

100

120

0 10 20 30 40 50 60

Le

ft g

luco

se

(%

)

Time (min)

300mg LiBr+0.25% acid

300mg LiBr+0.5% acid

600mg LiBr+0.25% acid

0

20

40

60

80

100

0 5 10 15 20 25 30

Left

fr

ucto

se

(%

)

Time (min)

300mg LiBr+0.25% acid

600mg LiBr+0.25% acid

Page 205: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

190

LiBr did not have any significant effect on the dehydration of fructose. Therefore, in

the all reaction steps from glucose to HMF, the effect of the step from fructose to

HMF could be ignored. In other words, the formation of fructose intermediate from

glucose was the rate-determining step where Br- and H

+ were both involved. In kinetic

study, the rate of glucose dehydration was affected by concentrations of both acid and

LiBr.

According to these experiments, the mechanism of acetone-LiBr promoted

conversion of glucose to HMF was proposed in Figure 5.16. Glucose reacted with

acetone first to form a cyclic acetal. Since the O-C-O angle in the cyclic acetal was

not optimum for tetrahedron geometry, the tension force favored the opening of the

ring and would prevent the reverse reaction. In addition, Br- as a strong neucleophile

attacked the positively charged anomeric carbon, resulting in the ring opening to form

enol, which was the key intermediate in the isomerization of glucose. The enol then

closed ring and formed fructose, which was readily and rapidly dehydrated to HMF.

The cyclic acetal always formed at C1 and C2 positions, because the cyclic acetal

formed at other positions was unable to give a stable ring-opening structure due to the

tension force of formed cyclic acetal. Similar mechanism was proposed for boric acid

catalyzed glucose conversion into HMF where boric acid formed a cyclic intermediate

Page 206: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

191

to facilitate the ring opening of glucose and stabilize the resultant ring-opening

structure (Stahlberg et al., 2011).

Figure 5.16 The proposed mechanism of glucose to HMF in acetone/LiBr system.

It is well known that furans (furfural and HMF) are instable and tend to

condense and form insoluble humins in acidic environment. It was reported that using

transition metals (e.g. CrClx) (Zhao et al., 2007) as catalyst and DMF, DMSO or ionic

liquid as solvent (Binder and Raines, 2009; Roman-Leshkov et al., 2006; Zhao et al.,

2007) instead of water could prevent HMF from the condensation (polymerization). In

this process, as excessive acetone was used in HDA process, newly formed

furfural/HMF molecules were isolated and surrounded by acetone molecules, which

prevented them from self-condensation. Therefore, the HDA environment favored the

aldol-condensation between furfural/HMF and acetone to form the hydrocarbon

precursors. In summary, acetone in HDA process had three functions: (1) as polar

Page 207: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

192

aprotic reagent to promote the formation of free Br-, (2) as reactive reagent involved

in the aldol-condensation reaction with furans, and (3) as solvent to dissolve the

hydrocarbon precursors and lignin degradation products.

5.3.12 Recycling/recovery of solvents and LiBr

Recycling of solvents and LiBr is crucial to the success of the HDA process.

After the reaction, acetone was in a mixture along with hydrocarbon precursors, LiBr,

water, lignin, and, if any, unreacted/residual biomass/sugars. The unreacted acetone

could be easily recovered by evaporation after the reaction. Then water was added to

the mixture to facilitate the separation of LiBr from lignin, followed by the extraction

of the precursors with CH2Cl2. Fresh water could be further added to wash CH2Cl2

and lignin, if necessary. The addition of water allowed the formation of three phases:

LiBr/water aqueous phase, CH2Cl2/precursors organic phase, and lignin solid phase.

After centrifugation, three phases could be easily separated. The LiBr solution

(including HCl) could be reused in the next batch directly or after concentration, if

necessary. CH2Cl2 was recycled through vaporization, and hydrocarbon precursor was

left as thick liquid or solid. The recovery yields of LiBr and HCl were estimated by

titration with Ag+. When water/acetone (50:50) was used as reaction solvent, titration

indicated that 97.5% LiBr was reserved in final LiBr solution after extraction with

CH2Cl2. Washing CH2Cl2 and lignin fraction with 5 mL water could recover

Page 208: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

193

additional 1.5% LiBr. Only approximately 1% of LiBr was non-recoverable, which

was probably dissolved in CH2Cl2 or absorbed on the precursors or lignin.

5.3.13 Hydrodeoxygenation of the precursors and lignin into

hydrocarbons

The hydrodeoxygenation of furan-derived hydrocarbon precursors and the

CH2Cl2-soluble low molecular weight lignin was conducted in ethanol with Pd/C and

SiO2-Al2O3 as catalyst. It might be unnecessary to achieve complete hydrogenation to

deoxygenate and saturate the hydrocarbon precursors, which is costly and needs a lot

of hydrogen gas. In other words, partial hydrodeoxygenation is probably enough to

make the hydrocarbon precursors miscible in hydrocarbon fuels. Preliminary results

indicated that both the hydrocarbon precursors and lignin were converted into hexane-

soluble products in 5 h at 250 C with initial hydrogen pressure of 6-7.5 MPa. The

images of the products in ethanol and suspended above water is shown in Figure 5.17.

For comparison, the precursor in ethanol before hydrodeoxygenation was given as

well. Because of the saturation of furan rings and double bonds in hydrocarbon

precursors, the color of products faded. The water-immiscibility and hexane-

miscibility of products indicated that the products were miscible with gasoline and

can be directly used as drop-in transportation fuel. It is anticipated that the

Page 209: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

194

hydrodeoxygenated hydrocarbon precursors could be blended in gasoline even

without the removal of solvent ethanol.

Figure 5.17 Images of products before and after hydrodeoxygenation. (a) Hydrocarbon

precursors dissolved in ethanol before hydrodeoxygenation; (b) Hydrodeoxygenated

hydrocarbon precursors dissolved in ethanol after hydrodeoxygenation; (c) Hydrophobic

products were separated from ethanol by adding water; (d) Separated hydrophobic products

(upper layer in C) were miscible in hexane.

Further investigations are required into the optimization of

hydrodeoxygenation, characterization of the deoxygenated hydrocarbon precursors,

and their engine performance when blended with gasoline. It is worthy pointing out

that no coke formed during the hydrodeoxygenation. This is different from the

aqueous phase hydrodeoxygenation where two steps were needed and some coke

formed in the aqueous phase (Huber et al., 2005).

a b c d

Page 210: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

195

5.4 Conclusion and recommendations

One-pot process (HDA) for converting lignocelluloses into hydrocarbon

precursors to fuel-grade hydrocarbons was developed. This process was carried out

under mild temperature (~120 °C) with inexpensive and recyclable catalyst (halide

salt) and could directly use real lignocelluloses as feedstock without pretreatment or

fractionation. Carbohydrates in the biomass can be readily converted into furan-

derived hydrocarbon precursors of C5-C21 to hydrocarbons in high yield and with

high selectivity (e.g., 72% for spruce). Lignin was extensively depolymerized during

the HDA process. Because of the very low molecular weight, the resultant HDA

lignin, in particular that from hardwood and herbage, had good potential to be

converted to chemicals and fuels. Not only the furan-derived hydrocarbon precursors

but also the low-molecular-weight fractions of the HDA lignin could be

hydrodeoxygenated (separately or jointly) into hydrocarbon fuels (or fuel additives).

As acetone is currently produced from petroleum and natural gas, renewable source of

acetone or other ketones should be pursued for HDA process in the future, for

example, from biomass through fermentation or pyrolysis process. The

hydrodeoxygenation of the hydrocarbon precursors and engine performance of the

resulting hydrocarbons need further investigation.

Page 211: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

196

Reference

Achyuthan, K. E., Achyuthan, A. M., Adams, P. D., Dirk, S. M., Harper, J. C., Simmons, B.

A., and Singh, A. K. (2010). Supramolecular self-assembled chaos: polyphenolic

lignin's barrier to cost-Effective lignocellulosic biofuels. Molecules, 15(12), 8641-

8688.

Aden, A., Ruth, M., Ibsen, K., Jechura, J., Neeves, K., Sheehan, J., Lukas, J. (2002).

Lignocellulosic biomass to ethanol process design and economic utilizing co-current

dilute acid prehydrolysis and enzymatic hydrolysis for corn Stover, National

Renewable Energy Laboratory, Golden, CO, NREL/TP-510-32438.

Ahring, B. K., and Langvad, N. (2008). Sustainable low cost production of lignocellulosic

bioethanol - "The carbon slaughterhouse". A process concept developed by BioGasol.

International Sugar Journal, 110(1311), 184-190.

Aida, T. M., Sato, Y., Watanabe, M., Tajima, K., Nonaka, T., Hattori, H., and Arai, K. (2007).

Dehydration of D-glucose in high temperature water at pressures up to 80 MPa.

Journal of Supercritical Fluids, 40(3), 381-388.

Amarasekara, A. S., Williams, L. D., and Ebede, C. C. (2008). Mechanism of the dehydration

of D-fructose to 5-hydroxymethylfurfural in dimethyl sulfoxide at 150 degrees C: an

NMR study. Carbohydrate Research, 343(18), 3021-3024.

Balan, V., Bals, B., Chundawat, S. P., Marshall, D., and Dale, B. E. (2009). Lignocellulosic

biomass pretreatment using AFEX. Methods in Molecular Biology, 581, 61-77.

Balfanz, U., Rupp, M., and Baldauf, W. (1993). Upgrading of pyrolysis oil from biomass to

petroleum-products. Abstracts of Papers of the American Chemical Society, 205, 56-

Cell.

Banerjee, S., Mudliar, S., Sen, R., Giri, B., Satpute, D., Chakrabarti, T., and Pandey, R. A.

(2010). Commercializing lignocellulosic bioethanol: technology bottlenecks and

possible remedies. Biofuels Bioproducts and Biorefining, 4(1), 77-93.

Berti, P. J., and Tanaka, K. S. E. (2002). Transition state analysis using multiple kinetic

isotope effects: Mechanisms of enzymatic and non-enzymatic glycoside hydrolysis

and transfer. Advances in Physical Organic Chemistry, 37, 239-314.

Bienkowski, P. R., Ladisch, M. R., Voloch, M., and Tsao, G. T. (1984). Acid-hydrolysis of

pretreated lignocellulose from corn residue. Biotechnology and Bioengineering

Symposium Series, 14, 511-524.

Page 212: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

197

Binder, J. B., and Raines, R. T. (2009). Simple chemical transformation of lignocellulosic

biomass into furans for fuels and chemicals. Journal of the American Chemical

Society, 131(5), 1979-1985.

Binder, J. B., and Raines, R. T. (2010). Fermentable sugars by chemical hydrolysis of

biomass. Proceedings of National Academy of Sciences of the United States of

America, 107(10), 4516-4521.

Bobletera, O., Schwalda, W., Concina, R., and Bindera, H. (1986). Hydrolysis of Cellobiose

in dilute sulpuric acid and under hydrothermal conditions. Journal of Carbohydrate

Chemistry, 5(3), 387-399.

Bond, J. Q., Alonso, D. M., Wang, D., West, R. M., and Dumesic, J. A. (2010). Integrated

catalytic conversion of gamma-valerolactone to liquid alkenes for transportation fuels.

Science, 327(5969), 1110-1114.

Boraston, A. B., Kwan, E., Chiu, P., Warren, R. A. J., and Kilburn, D. G. (2003). Recognition

and hydrolysis of noncrystalline cellulose. Journal of Biological Chemistry, 278(8),

6120-6127.

Bouchoux, G., Djazi, F., Houriet, R., and Rolli, E. (1988). Gas-phase basicity of olefinic C-5

and C-6 carbonyl-compounds. Journal of Organic Chemistry, 53(15), 3498-3501.

Bravo, V., Paez, M. P., Aoulad, M., and Reyes, A. (2000). The influence of temperature upon

the hydrolysis of cellobiose by b-1,4-glucosidases from Aspergillus niger. Enzyme

and Microbial Technology, 26, 614-620.

Brendler, E., Fisher, S., and Leipner, H. (2002). 7Li NMR as probe for solvent-cellulose

interactions in cellulose dissolution. Cellulose, 8, 283-288.

Cao, N. J., Xu, Q., Chen, C. S., Gong, C. S., and Chen, L. F. (1994). Cellulose hydrolysis

using zinc chloride as a solvent and catalyst. Applied Biochemistry and Biotechnology,

45(6), 521-530.

Carroll, A., and Somerville, C. (2009). Cellulosic Biofuels. Annual Review of Plant Biology,

60, 165-182.

Chandra, R. P., Bura, R., Mabee, W. E., Berlin, A., Pan, X., and Saddler, J. N. (2007).

Substrate pretreatment: The key to effective enzymatic hydrolysis of lignocellulosics.

Biofuels, 108, 67-93.

Chen, L. Y., and Yang, C. (1984). U.S patent Patent No. 4452640.

Chheda, J. N., Huber, G. W., and Dumesic, J. A. (2007). Liquid-phase catalytic processing of

biomass-derived oxygenated hydrocarbons to fuels and chemicals. Angewandte

Chemie-International Edition, 46(38), 7164-7183.

Coutinho, J. B., Gilkes, N. R., Kilburn, D. G., Warren, R. A. J., and Miller, R. C. (1993). The

nature of the cellulose-binding domain affects the activities of a bacterial

Page 213: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

198

endoglucanase on different forms of cellulose. Fems Microbiology Letters, 113(2),

211-218.

Creagh, A. L., Ong, E., Jervis, E., Kilburn, D. G., and Haynes, C. A. (1996). Binding of the

cellulose-binding domain of exoglucanase Cex from Cellulomonas fimi to insoluble

microcrystalline cellulose is entropically driven. Proceedings of the National

Academy of Sciences of the United States of America, 93(22), 12229-12234.

Cuzens, W. A. F. a. J. (1998). U.S. patent Patent No. 5820687.

Cyr, N., Elofson, R. M., Ripmeester, J. A., and Mathison, G. W. (1988). Study of lignin in

forages and wood by C-13 CP/MAS NMR. 1. Some evidence of polymerization and

depolymerization. Journal of Agricultural and Food Chemistry, 36(6), 1197-1201.

Dadi, A. P., Varanasi, S., and Schall, C. A. (2006). Enhancement of cellulose saccharification

kinetics using an ionic liquid pretreatment step. Biotechnology and Bioengineering,

95(5), 904-910.

Davda, R. R., and Dumesic, J. A. (2004). Renewable hydrogen by aqueous-phase reforming

of glucose. Chemical Communications, (1), 36-37.

Dawsey, T. R., and Mccormick, C. L. (1990). The lithium lhloride/dimethylacetamide solvent

for cellulose-a literature review. Journal of Macromolecular Science-Reviews in

Macromolecular Chemistry and Physics, 30(3-4), 405-440.

Demirbas, A. (2005). Bioethanol from cellulosic materials: A renewable motor fuel from

biomass. Energy Sources, 27(4), 327-337.

Demirbas, M. F. (2009). Biorefineries for biofuel upgrading: A critical review. Applied

Energy, 86, 151-161.

Dence, C. W. (1992). The determination of lignin. Methods in lignin chemistry, 33-61.

Jiao, D., King, C., Grossfield, A., Darden, T. A., and Ren, P. (2006). Simulation of Ca2+

and

Mg2+

solvation using polarizable atomic multipole potential. Journal of Physical

Chemistry B, 110, 18553-18559.

Doughty, H. W. (1924). Mohr's method for the determination of silver and halogens in other

than neutral solutions. Journal of American Chemical Society, 46(12), 2707-2709.

Elliott, D. C., and Schiefelbein, G. F. (1989). Liquid hydrocarbon fuels from biomass.

Abstracts of Papers of the American Chemical Society, 34(4), 1160.

Fan, L., Lee, Y. H., and Gharpuray, M. (1982). The nature of lignocellulosics and their

pretreatments for enzymatic hydrolysis. Microbial Reactions, 23, 157-187.

Farrell, A. E., Plevin, R. J., Turner, B. T., Jones, A. D., O'Hare, M., and Kammen, D. M.

(2006). Ethanol can contribute to energy and environmental goals. Science,

311(5760), 506-508.

Fengel, D., and Wegener, G. (1984). Wood : chemistry, ultrastructure, reactions. Berlin ;

New York: W. de Gruyter.

Page 214: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

199

Fischer, S., Leipner, H., Thummler, K., Brendler, E., and Peters, J. (2003). Inorganic molten

salts as solvents for cellulose. Cellulose, 10(3), 227-236.

Fu, Y., Lai, D. M., Deng, L., Li, J. A., Liao, B., and Guo, Q. X. (2011). Hydrolysis of

cellulose into glucose by magnetic solid acid. Chemsuschem, 4(1), 55-58.

Galbe, M., and Zacchi, G. (2002). A review of the production of ethanol from softwood.

Applied Microbiology and Biotechnology, 59, 618-628.

Ghose, T. K. (1987). Measurement of cellulase activities. Pure and Applied Chemistry, 59,

257-268.

Gierer, J. (1985). Chemistry of delignification. 1. General concept and reactions during

pulping. Wood Science and Technology, 19(4), 289-312.

Gilkes, N. R., Jervis, E., Henrissat, B., Tekant, B., Miller, R. C., Warren, R. A. J., and Kilburn,

D. G. (1992). The adsorption of a bacterial cellulase and its two isolated domains to

crystalline cellulose. Journal of Biological Chemistry, 267(10), 6743-6749.

Gilkes, N. R., Kilburn, D. G., Miller, R. C., Warren, R. A. J., Sugiyama, J., Chanzy, H., and

Henrissat, B. (1993). Visualization of the adsorption of a bacterial endo-beta-1,4-

glucanase and its isolated cellulose-binding domain to crystalline cellulose.

International Journal of Biological Macromolecules, 15(6), 347-351.

Girisuta, B., Janssen, L. P. B. M., and Heeres, H. J. (2006). A kinetic study on the

decomposition of 5-hydroxymethylfurfural into levulinic acid. Green Chemistry, 8(8),

701-709.

Grinbaum, B., Kogan, L., Barnea, E., Harel, G., Semiat, R., and Wahrmann, S. (2008). USA

Patent No. 7452520.

Gupta, R. C., and Sehgal, V. K. (1979). Effect of viscosity and molecular weight of lignin

phenol formaldehyde resin on the glue adhesion Strength of plywood. Holzforschung

Und Holzverwertung, 31(1), 7-9.

Hald, P. (1995). The behavior of alkali-metals in biomass conversion systems. Abstracts of

Papers of the American Chemical Society, 40(3), 753-757.

Hara, M., Kitano, M., Yamaguchi, D., Suganuma, S., Nakajima, K., Kato, H., and Hayashi, S.

(2009). Adsorption-enhanced hydrolysis of beta-1,4-glucan on graphene-based

amorphous carbon bearing SO3H, COOH, and OH groups. Langmuir, 25(9), 5068-

5075.

Hara, M., Suganuma, S., Nakajima, K., Kitano, M., Yamaguchi, D., Kato, H., and Hayashi, S.

(2010). Synthesis and acid catalysis of cellulose-derived carbon-based solid acid.

Solid State Sciences, 12(6), 1029-1034.

Harris, E. E., and Lang, B. G. (1947). Hydrolysis of wood cellulose and decomposition of

sugar in dilute phosphoric acid. Journal of Physical and Colloid Chemistry, 51(6),

1430-1441.

Page 215: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

200

Harris, J. F. (1985). Two-stage, dilute sulfuric acid hydrolysis of wood: an investigation of

fundamentals. [Madison, Wis.]: U.S. Dept. of Agriculture, Forest Service, Forest

Products Laboratory.

He, H., Wang, H. Y., Zhang, C. B., and Wang, L. A. (2010). Preparation of magnetic

sulfonated carbon-based solid acid catalysts for the hydrolysis of cellulose. Acta

Physico-Chimica Sinica, 26(7), 1873-1878.

Hedjazi, S., Kordsachia, O., Patt, R., Latibari, A. J., and Tschirner, U. (2009). Alkaline

sulfite-anthraquinone (AS/AQ) pulping of wheat straw and totally chlorine free (TCF)

bleaching of pulps. Industrial Crops and Products, 29(1), 27-36.

Heeres, H. J., Girisuta, B., and Janssen, L. P. B. M. (2007). Kinetic study on the acid-

catalyzed hydrolysis of cellulose to levulinic acid. Industrial and Engineering

Chemistry Research, 46(6), 1696-1708.

Hendriks, A. T., and Zeeman, G. (2009). Pretreatments to enhance the digestibility of

lignocellulosic biomass. Bioresource Technology, 100(1), 10-18.

Hsuanyu, Y., and Laidler, K. J. (1985). Kinetics of the hydrolysis of cellobiose catalysed by

P-glucosidase. Canadian Journal of Biochemistry and Cell Biology, 63, 167-175.

Huang, A. A. (1975). Enzymatic hydrolysis of cellulose to sugar. Biotechnology and

Bioengineering(5), 245-252.

Huber, G. W., Chheda, J. N., Barrett, C. J., and Dumesic, J. A. (2005). Production of liquid

alkanes by aqueous-phase processing of biomass-derived carbohydrates. Science,

308(5727), 1446-1450.

Huber, G. W., Cortright, R. D., and Dumesic, J. A. (2004). Renewable alkanes by aqueous-

phase reforming of biomass-derived oxygenates. Angewandte Chemie-International

Edition, 43(12), 1549-1551.

Huber, G. W., and Dumesic, J. A. (2006). An overview of aqueous-phase catalytic processes

for production of hydrogen and alkanes in a biorefinery. Catalysis Today, 111(1-2),

119-132.

Jeong, S., Joo, S., Kim, H., Kim, S., and Jeong, H. (2009). Effect on two-step polishing

process of electrochemical mechanical planarization and chemical-mechanical

planarization on planarization. Japanese Journal of Applied Physics, 48(6).

Jolly, W. L. (1991). Modern Inorganic Chemistry (2nd ed.). New York: McGraw-Hill.

Kim, Y., Mosier, N. S., and Ladisch, M. R. (2009). Enzymatic digestion of liquid hot water

pretreated hybrid poplar. Biotechnology Progress, 25(2), 340-348.

Kirubakaran, V., Sivaramakrishnan, V., Nalini, R., Sekar, T., Premalatha, M., and

Subramanian, P. (2009). A review on gasification of biomass. Renewable and

Sustainable Energy Reviews, 13(1), 168-175.

Page 216: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

201

Kobayashi, H., Komanoya, T., hara, K., and Fukuoka, A. (2010). Water-tolerant mesoporous-

carbon-supported ruthenium catalysts for the hydrolysis of cellulose to glucose.

Chemsuschem, 3, 440-443.

Kosan, B., Michels, C., and Meister, F. (2008). Dissolution and forming of cellulose with

ionic liquids. Cellulose, 15(1), 59-66.

Krassig, H., and Schurz, J. (2002). Ullmann's Encyclopedia of Industrial Chemistry (6 ed.).

Weinheim, Germany, Wiley-VCH.

Lai, D. M., Deng, L., Guo, Q., and Fu, Y. (2011). Hydrolysis of biomass by magnetic solid

acid. Energy and Environmental Science, 4, 55-58.

Lawrence J. Russo, J. (1999). U.S. patent Patent No. 5968362.

Lee, Y. Y., Xiang, Q., Pettersson, P. O., and Torget, R. (2003). Heterogeneous aspects of acid

hydrolysis of alpha-cellulose. Applied Biochemistry and Biotechnology, 105, 505-514.

Li, C. Z., Wang, Q., and Zhao, Z. K. (2008). Acid in ionic liquid: an efficient system for

hydrolysis of lignocellulose. Green Chemistry, 10(2), 177-182.

Li, C. Z., and Zhao, Z. K. B. (2007). Efficient acid-catalyzed hydrolysis of cellulose in ionic

liquid. Advanced Synthesis and Catalysis, 349(11-12), 1847-1850.

Limayem, A., and Ricke, S. C. (2012). Lignocellulosic biomass for bioethanol production:

Current perspectives, potential issues and future prospects. Progress in Energy and

Combustion Science, doi:10.1016/j.pecs.2012.03.002.

Lingaraju, B. P., Lee, J. Y., and Yang, J. Y. (2012). Process and utility water requirements for

cellulosic ethanol production processes via fermentation pathway. Environmental

Progress and Sustainable Energy. doi: DOI 10.1002/ep.11604

Lloyd, T. A., and Wyman, C. E. (2005). Combined sugar yields for dilute sulfuric acid

pretreatment of corn stover followed by enzymatic hydrolysis of the remaining solids.

Bioresource Technology, 96(18), 1967-1977.

Lu, Q., Zhu, X. F., Li, W. Z., Zhang, Y., and Chen, D. Y. (2009). On-line catalytic upgrading

of biomass fast pyrolysis products. Chinese Science Bulletin, 54(11), 1941-1948.

Lv, D. Z., Xu, M. H., Liu, X. W., Zhan, Z. H., Li, Z. Y., and Yao, H. (2010). Effect of

cellulose, lignin, alkali and alkaline earth metallic species on biomass pyrolysis and

gasification. Fuel Processing Technology, 91(8), 903-909.

Lynd, L. R., Weimer, P. J., van Zyl, W. H., and Pretorius, I. S. (2002). Microbial cellulose

utilization: Fundamentals and biotechnology. Microbiology and Molecular Biology

Reviews, 66(4), 739-739.

Mansfield, S. D., Mooney, C., and Saddler, J. N. (1999). Substrate and enzyme characteristics

that limit cellulose hydrolysis. Biotechnology Progress, 15, 804-816.

Page 217: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

202

Marbe, A., Harvey, S., and Berntsson, T. (2006). Technical, environmental and economic

analysis of co-firing of gasified biofuel in a natural gas combined cycle (NGCC)

combined heat and power (CHP) plant. Energy, 31(10-11), 1614-1631.

Mccarter, J. D., and Withers, S. G. (1994). Mechanisms of enzymatic glycoside hydrolysis.

Current Opinion in Structural Biology, 4(6), 885-892.

Mccormick, C. L., Callais, P. A., and Hutchinson, B. H. (1985). Solution studies of cellulose

in lithium chloride and N,N-dimethylacetamide. Macromolecules, 18(12), 2394-2401.

McIntosh, S., and Vancov, T. (2010). Enhanced enzyme saccharification of Sorghum bicolor

straw using dilute alkali pretreatment. Bioresource Technology, 101(17), 6718-6727.

McLaren, A. D. (1948). Adhesion of high polymers to cellulose. Influence of structure,

polarity, and tack temperature. Journal of Polymer Science, 3, 652-662.

Medve, J., Karlsson, J., Lee, D., and Tjerneld, F. (1998). Hydrolysis of microcrystalline

cellulose by cellobiohydrolase I and endoglucanase II from Trichoderma reesei:

Adsorption, sugar production pattern, and synergism of the enzymes. Biotechnology

and Bioengineering, 59(5), 621-634.

Melillo, J. M., Reilly, J. M., Kicklighter, D. W., Gurgel, A. C., Cronin, T. W., Paltsev, S.,

Schlosser, C. A. (2009). Indirect Emissions from Biofuels: How Important?. Science,

326(5958), 1397-1399.

Miller, S., and Hester, R. (2007). Concentrated acid conversion of pine softwood to sugars.

Part I: Use of a twin-screw reactor for hydrolysis pretreatment. Chemical Engineering

Communications, 194(1), 85-102.

Modenbach, A. A., and Nokes, S. E. (2012). The use of high-solids loadings in biomass

pretreatment—A review. Biotechnology and Bioengineering, 109(6), 1430-1442.

Monavari, S., Galbe, M., and Zacchi, G. (2009). Impact of impregnation time and chip size on

sugar yield in pretreatment of softwood for ethanol production. Bioresource

Technology, 100(24), 6312-6316.

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

(2005). Features of promising technologies for pretreatment of lignocellulosic

biomass. Bioresource Technology, 96(6), 673-686.

Mosier, N. S. (2007). Cellulosic ethanol—biofuel beyond corn. Purdue extension, ID 335.

Mosier, N. S., Wilker, J. J., and Ladisch, M. R. (2004). Rapid chromatography for evaluating

adsorption characteristics of cellulase binding domain mimetics. Biotechnology and

Bioengineering, 86(7), 756-764.

Muangrat, R., Onwudili, J. A., and Williams, P. T. (2010). Alkali-promoted hydrothermal

gasification of biomass food processing waste: a parametric study. International

Journal of Hydrogen Energy, 35(14), 7405-7415.

Page 218: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

203

Nguyen, Q. A., Tucker, M. P., Keller, F. A., and Eddy, F. P. (2000). Two-stage dilute-acid

pretreatment of softwoods. Applied Biochemistry and Biotechnology, 84(6), 561-576.

Onda, A., Ochi, T., Kajiyoshi, K., and Yanagisawa, K. (2008). A new chemical process for

catalytic conversion Of D-glucose into lactic acid and gluconic acid. Applied

Catalysis A-General, 343(1-2), 49-54.

Onda, A., Ochi, T., and Yanagisawa, K. (2008). Selective hydrolysis of cellulose into glucose

over solid acid catalysts. Green Chemistry, 10(10), 1033-1037.

Paljevac, M., Primozic, M., Habulin, M., Novak, Z., and Knez, Z. (2007). Hydrolysis of

carboxymethyl cellulose catalyzed by cellulase immobilized on silica gels at low and

high pressures. Journal of Supercritical Fluids, 43(1), 74-80.

Palkovits, R., Tajvidi, K., Procelewska, J., Rinaldi, R., and Ruppert, A. (2010).

Hydrogenolysis of cellulose combining mineral acids and hydrogenation catalysts.

Green Chemistry, 12(6), 972-978.

Pan, X., Xie, D., Kang, K. Y., Yoon, S. L., and Saddler, J. N. (2007). Effect of organosolv

ethanol pretreatment variables on physical characteristics of hybrid poplar substrates.

Applied Biochemistry Biotechnology, 137(12), 367-377.

Pan, X. J., Arato, C., Gilkes, N., Gregg, D., Mabee, W., Pye, K., Saddler, J. (2005).

Biorefining of softwoods using ethanol organosolv pulping: Preliminary evaluation of

process streams for manufacture of fuel-grade ethanol and co-products.

Biotechnology and Bioengineering, 90(4), 473-481.

Pan, Y. J., Liu, P. Y., Hu, N., and Tu, Y. P. (2010). Ion-neutral complexes resulting from

dissociative protonation: fragmentation of alpha-furanylmethyl benzyl ethers and 4-

N,N-dimethylbenzyl benzyl ethers. Journal of the American Society for Mass

Spectrometry, 21(4), 626-634.

Pandey, M. P., and Kim, C. S. (2011). Lignin depolymerization and conversion: a review of

thermochemical methods. Chemical Engineering and Technology, 34(1), 29-41.

Pang, J., Wang, A., Zheng, M., and Zhang, T. (2010). Hydrolysis of cellulose into glucose

over carbons sulfonated at elevated temperatures Chemical Communications, 46,

6935-6937.

Pang, J. F., Wang, A. Q., Zheng, M. Y., and Zhang, T. (2010). Hydrolysis of cellulose into

glucose over carbons sulfonated at elevated temperatures. Chemical Communications,

46(37), 6935-6937.

Panicker, C. Y., Yarghese, H. T., Anto, P. L., and Philip, D. (2006). Potential dependent

SERS profile of sulfanilic acid on silver electrode. Journal of Raman Spectroscopy.,

37, 853-857.

Penque, R. A. (1977). U.S. patent Patent No. 4018620.

Pfaff, K. P., Paust, J., Hartmann, H. (1987). United States Patent No. 4681936.

Page 219: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

204

Piccolo, C., and Bezzo, F. (2009). A techno-economic comparison between two technologies

for bioethanol production from lignocellulose. Biomass and Bioenergy, 33(3), 478-

491.

Qian, X. H., Nimlos, M. R., Johnson, D. K., and Himmel, M. E. (2005). Acidic sugar

degradation pathways. Applied Biochemistry and Biotechnology, 121, 989-997.

Qing, Q., Yang, B., & Wyman, C. E. (2010). Impact of surfactants on pretreatment of corn.

Bioresource Technology, 101(15), 5941-5951.

Ragauskas, A. J., Williams, C. K., Davison, B. H., Britovsek, G., Cairney, J., Eckert, C. A.,

Tschaplinski, T. (2006). The path forward for biofuels and biomaterials. Science,

311(5760), 484-489.

Ralph, J., and Landucci, L. L. (2010). Lignin and Lignans Advances in Chemistry, 137-243.

Boca Raton, FL: CRC press (Taylor and Francis group).

Ralph, S. (2005). Conundrums regarding 5-O-4-linkages in softwood MWLs. paper presented

at the 59th appita annual conference and exhibition : incorporating the 13th ISWFPC

(International Symposium on Wood, Fibre and Pulping Chemistry) Auckland, New

Zealand.

Remsing, R. C., Swatloski, R. P., Rogers, R. D., and Moyna, G. (2006). Mechanism of

cellulose dissolution in the ionic liquid 1-n-butyl-3-methylimidazolium chloride: a C-

13 and Cl-35/37 NMR relaxation study on model systems. Chemical Communications,

(12), 1271-1273.

Roman-Leshkov, Y., Chheda, J. N., and Dumesic, J. A. (2006). Phase modifiers promote

efficient production of hydroxymethylfurfural from fructose. Science, 312(5782),

1933-1937.

Sathitsuksanoh, N., Zhu, Z., Ho, T. J., Bai, M. D., and Zhang, Y. H. (2010). Bamboo

saccharification through cellulose solvent-based biomass pretreatment followed by

enzymatic hydrolysis at ultra-low cellulase loadings. Bioresource Technology,

101(13), 4926-4929.

Schall, C. A., Dadi, A. P., and Varanasi, S. (2006). Enhancement of cellulose saccharification

kinetics using an ionic liquid pretreatment step. Biotechnology and Bioengineering,

95(5), 904-910.

Scharlemann, J. P. W., and Laurance, W. F. (2008). Environmental science - How green are

biofuels?. Science, 319(5859), 43-44.

Schell, D. J., Farmer, J., Newman, M., and McMillan, J. D. (2003). Dilute sulfuric acid

pretreatment of corn stover in pilot-scale reactor - Investigation of yields, kinetics,

and enzymatic digestibilities of solids. Applied Biochemistry and Biotechnology, 105,

69-85.

Page 220: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

205

Scheller, H. V., and Ulvskov, P. (2010). Hemicelluloses. Annual Review of Plant Biology, 61,

263-289.

Sendich, E., Laser, M., Kim, S., Alizadeh, H., Laureano-Perez, L., Dale, B., and Lynd, L.

(2008). Recent process improvements for the ammonia fiber expansion (AFEX)

process and resulting reductions in minimum ethanol selling price. Bioresource

Technology, 99(17), 8429-8435.

Gou, Z. and Zhu, S. (1998). Extraction of lithium from brine containing high concentration of

magnesium by tri-n-butyl phosphate dissolved in kerosene. Chinese Journal of

Chemical Engineering, 6, 124-129.

Shevchenko, S. M. (2000). Depolymerization of lignin in wood with molecular hydrogen

Iodide. Croatica Chemica Acta, 73(3), 831-841.

Shuai, L., Yang, Q., Zhu, J. Y., Lu, F. C., Weimer, P. J., Ralph, J., and Pan, X. J. (2010).

Comparative study of SPORL and dilute-acid pretreatments of spruce for cellulosic

ethanol production. Bioresource Technology, 101(9), 3106-3114.

Sierra, R., Granda, C., and Holtzapple, M. T. (2009). Short-term lime pretreatment of poplar

wood. Biotechnology Progress, 25(2), 323-332.

Smith, A. M. (2008). Prospects for increasing starch and sucrose yields for bioethanol

production. Plant Journal, 54(4), 546-558.

Stahlberg, T., Rodriguez-Rodriguez, S., Fristrup, P., and Riisager, A. (2011). Metal-free

dehydration of glucose to 5-(hydroxymethyl)furfural in ionic liquids with boric acid

as a promoter. Chemistry-a European Journal, 17(5), 1456-1464.

Suganuma, S., Nakajima, K., Kitano, M., Yamaguchi, D., Kato, H., Hayashi, S., and Hara, M.

(2008). Hydrolysis of cellulose by amorphous carbon bearing SO3H, COOH, and OH

groups. Journal of the American Chemical Society, 130(38), 12787-12793.

Sukumaran, R. K., Singhania, R. R., and Pandey, A. (2005). Microbial cellulases - Production,

applications and challenges. Journal of Scientific and Industrial Research, 64(11),

832-844.

Sun, Y., and Cheng, J. Y. (2002). Hydrolysis of lignocellulosic materials for ethanol

production: a review. Bioresource Technology, 83(1), 1-11.

Taherzadeh, M., and Karimi, K. (2007). Acid-based hydrolysis processes for ethanol from

lignocellulosic materials: a review. Bioresources, 2, 472-499.

Taherzadeh, M. J., and Karimi, K. (2008). Pretreatment of lignocellulosic wastes to improve

ethanol and biogas production: a review. International Journal of Molecular Sciences,

9, 1621-1651.

Takagaki, A., Tagusagawa, C., and Domen, K. (2008). Glucose production from saccharides

using layered transition metal oxide and exfoliated nanosheets as a water-tolerant

solid acid catalyst. Chemical Communications, (42), 5363-5365.

Page 221: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

206

Takagaki, A., Tagusagawa, C., Hayashi, S., Hara, M., and Domen, K. (2010). Nanosheets as

highly active solid acid catalysts for green chemical syntheses. Energy and

Enviromental Sciences, 3, 82-93.

Ting Chia Huang, T. H. T. (1991). Study of extraction behavior of lithium chloride by tri-n-

butyl phosphate dissolved in kerosene. Journal of Chemical Engineering Data, 36,

231-235.

Tomas-Pejo, E., Oliva, J. M., & Ballesteros, M. (2008). Realistic approach for full-scale

bioethanol production from lignocellulose: a review. Journal of Scientific and

Industrial Research, 67(11), 874-884.

Tomme, P., Creagh, A. L., Kilburn, D. G., and Haynes, C. A. (1996). Interaction of

polysaccharides with the N-terminal cellulose-binding domain of Cellulomonas fimi

CenC. 1. Binding specificity and calorimetric analysis. Biochemistry, 35(44), 13885-

13894.

Turbak, A. F., Hammer, R. B., Portnoy, N. A., and Davies, R. E. (1977). Critical review of

cellulose solvent systems. Abstracts of Papers of the American Chemical Society, 173,

6.

Bukowskya H. and Uhlemanna E. (1993). Selective extraction of lithium chloride from brines.

Separation Science and Technology, 28, 1357-1360.

Van de Vyver, S., Geboers, J., Jacobs, P. A., and Sels, B. F. (2011). Recent advances in the

catalytic conversion of cellulose. Chemcatchem, 3(1), 82-94.

Van de Vyver, S., Peng, L., Geboers, J., Schepers, H., de Clippel, F., Gommes, C. J., Sels, B.

F. (2010). Sulfonated silica/carbon nanocomposites as novel catalysts for hydrolysis

of cellulose to glucose. Green Chemistry, 12(9), 1560-1563.

Vandam, H. E., Kieboom, A. P. G., and Vanbekkum, H. (1986). The conversion of fructose

and glucose in acidic media - formation of hydroxymethylfurfural. Starch-Starke,

38(3), 95-101.

Weimer, P. J., Dien, B. S., Springer, T. L., and Vogel, K. P. (2005). In vitro gas production as

a surrogate measure of the fermentability of cellulosic biomass to ethanol. Applied

Microbiology and Biotechnology, 67(1), 52-58.

West, R. M., Kunkes, E. L., Simonetti, D. A., and Dumesic, J. A. (2009). Catalytic conversion

of biomass-derived carbohydrates to fuels and chemicals by formation and upgrading

of mono-functional hydrocarbon intermediates. Catalysis Today, 147(2), 115-125.

Wingren, A., Galbe, M., and Zacchi, G. (2008). Energy considerations for a SSF-based

softwood ethanol plant. Bioresource Technology, 99(7), 2121-2131.

Wolfram Rudolph, M. H. B., and Cory C. Pye. (1995). Hydration of lithium ion in aqueous

solution. Journal of Physical Chemistry, 99, 3793-3797.

Page 222: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

207

Wyman, C. E., Dale, B. E., Elander, R. T., Holtzapple, M., Ladisch, M. R., and Lee, Y. Y.

(2005). Coordinated development of leading biomass pretreatment technologies.

Bioresource Technology, 96(18), 1959-1966.

Xiao, Z. Z., Wang, P., Qu, Y. B., Gao, P. J., and Wang, T. H. (2002). Cold adaptation of a

mesophilic cellulase, EG III from Trichoderma reesei, by directed evolution. Science

in China Series C-Life Sciences, 45(4), 337-343.

Xu, Q., Singh, A., and Himmel, M. E. (2009). Perspectives and new directions for the

production of bioethanol using consolidated bioprocessing of lignocellulose. Current

Opinion in Biotechnology, 20(3), 364-371.

Yamaguchi, D., Kitano, M., Suganuma, S., Nakajima, K., Kato, H., and Hara, M. (2009).

Hydrolysis of cellulose by a solid acid catalyst under optimal reaction conditions.

Journal of Physical Chemistry C, 113(8), 3181-3188.

Yang, B., and Wyman, C. E. (2008). Pretreatment: the key to unlocking low-cost cellulosic

ethanol. Biofuels Bioproducts and Biorefining, 2(1), 26-40.

Yang, Y., Sharma-Shivappa, R., Burns, J. C., and Cheng, J. J. (2009). Dilute acid

pretreatment of oven-dried switchgrass germplasms for bioethanol Production.

Energy and Fuels, 23(7), 3759-3766.

Yasuda, S., and Murase, N. (1995). Chemical structures of sulfuric-acid Lignin. 12. Reaction

of lignin models with carbohydrates in 72-percent H2SO4. Holzforschung, 49(5), 418-

422.

Yoon, S. H., Ha, S. Y., Roh, H. G., and Lee, C. S. (2009). Effect of bioethanol as an

alternative fuel on the emissions reduction characteristics and combustion stability in

a spark ignition engine. Proceedings of the Institution of Mechanical Engineers Part

D-Journal of Automobile Engineering, 223(7), 941-951.

Zakzeski, J., Bruijnincx, P. C. A., Jongerius, A. L., and Weckhuysen, B. M. (2010). The

catalytic valorization of lignin for the production of renewable chemicals. Chemical

Reviews, 110(6), 3552-3599.

Zerbe, J. I., and Baker, A. J. (1988). Investigation of fundamentals of two-stage, dilute

sulfuric acid hydrolysis of wood. Madison, Wis.: Forest Products Laboratory.

Zhang, T. Y., and Wyman, C. E. (2011). Integrated production of levulinic acid and furfural

from cellulosic biomass. Abstracts of Papers of the American Chemical Society, 241.

Zhang, X., Jin, Q., Dai, L., and Yuan, S. (2011). Facile synthesis of hypercrosslinked resins

via chloromethylation and continuous condensation of simple aryl molecules. Bulletin

of Material Science, 34, 735-738.

Zhang, Y. H. P., Himmel, M. E., and Mielenz, J. R. (2006). Outlook for cellulase

improvement: Screening and selection strategies. Biotechnology Advances, 24(5),

452-481.

Page 223: Transforming Lignocelluloses to Sugars and Liquid Fuels By LI SHUAI A dissertation submitted in

208

Zhang, Z. H., and Zhao, Z. B. K. (2009). Solid acid and microwave-assisted hydrolysis of

cellulose in ionic liquid. Carbohydrate Research, 344(15), 2069-2072.

Zhao, H. B., Holladay, J. E., Brown, H., and Zhang, Z. C. (2007). Metal chlorides in ionic

liquid solvents convert sugars to 5-hydroxymethylfurfural. Science, 316(5831), 1597-

1600.

Zhu, Z. G., Sathitsuksanoh, N., Vinzant, T., Schell, D. J., McMillan, J. D., and Zhang, Y. H. P.

(2009). Comparative study of corn stover pretreated by dilute acid and cellulose

solvent-based lignocellulose fractionation: enzymatic hydrolysis, supramolecular

structure, and substrate accessibility. Biotechnology and Bioengineering, 103(4), 715-

724.