structure-function relationship of flaxseed gum from

118
Structure-Function Relationship of Flaxseed Gum from Flaxseed Hulls By Ke-Ying Qian A Thesis Presented to The University of Guelph In partial fulfilment of requirements for the degree of Doctor of Philosophy In Food Science Guelph, Ontario, Canada © Ke-Ying Qian, January, 2014

Upload: others

Post on 02-Jan-2022

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Structure-Function Relationship of Flaxseed Gum from

Structure-Function Relationship of Flaxseed Gum

from Flaxseed Hulls

By

Ke-Ying Qian

A Thesis Presented to

The University of Guelph

In partial fulfilment of requirements for the degree of

Doctor of Philosophy In

Food Science

Guelph, Ontario, Canada

© Ke-Ying Qian, January, 2014

Page 2: Structure-Function Relationship of Flaxseed Gum from

ABSTRACT

STRUCTRUE-FUNCTION RELATIONSHIP OF FLAXSEED GUM

FROM FLAXSEED HULLS

Ke-Ying Qian Co-advisor: Prof. H. Douglas Goff

University of Guleph, 2014 Co-advisor: Prof. Steve W. Cui

Soluble dietary fibre with low viscosity can be included in the diet in a significant

amount to show health benefits without over-texturization. Soluble flaxseed gum has

recently been the subject of increasing research attention due to its low viscosity, which is

favored as a potential fibre fortifier. In this work, soluble flaxseed gum (SFG) extracted

from flaxseed hulls was fractionated into a neutral (NFG) and an acidic (AFG) fraction

gum using ion exchange chromatography. The physicochemical properties, structure and

conformation of both fractions were investigated. The protein and uronic acid content

were 11.8 and 38.7 % in SFG, and 8.1 and 23.0 % in AFG, respectively. NFG contained

no protein or uronic acid. NFG and AFG exhibited pseudoplastic and Newtonian flow

behavior, respectively. The ranking of intrinsic viscosities (mL g-1) in a decreasing order

was: SFG (446.0) >NFG (377.5) >AFG (332.5). AFG was expected to have higher chain

flexibility due to its lower value of Huggins constant (0.16) compared to that of NFG (0.54)

and SFG (0.48). AFG had a highly branched rhamnogalacturonan-I backbone substituted

Page 3: Structure-Function Relationship of Flaxseed Gum from

iii

at O-3 of →2)-α-L-Rhap-(1→ which was mostly terminated by a monomeric sugar unit

(19.6 % T-α/β-D-Galp, 4.5 % T-α-L-Fucp, and 3.1 % T-β-D-Xylp) or occasionally by a

longer side chain with a similar structure as its backbone. NFG contained β-1,4-linked

xylose backbone being mono-, di- or unsubstituted at O-2 and/or O-3 positions by one to

three linearly- linked sugar residues (24.8 % arabinose, 15.5 % xylose and 9.4 %

galactose/glucose). Static and dynamic light scattering results showed that NFG exhibited

random coil conformation with a molecular weight of 616±19 kDa. The star-like

conformation of AFG (with a molecular weight of 285±25 kDa) also coincided with the

presence of longer side chains along its rather flexible RG-I backbone. The random coil

and star-like conformation of NFG and AFG, as well as their relatively small molecular

sizes, explained the overall low viscosity of flaxseed gum extracted from flaxseed hulls.

Page 4: Structure-Function Relationship of Flaxseed Gum from

iv

Acknowledgements

Firstly, I would like to thank my advisor Prof. H. Douglas Goff and my co-advisor

Prof. Steve. W. Cui for providing me the opportunity to work on this project. They both

inspired me with their constant enthusiasm and dedication to academia and their

willingness to deliver knowledge. I would like to thank them for their intellectual advices

throughout the project, and their sincerity, great tolerance and patience in helping me

overcome obstacles in experiments and writing. Special thanks are due to Prof. H. Doug

Goff for his effort on final editing of all my manuscripts, and to Prof. Steve Cui for his

insightful suggestions to my personal growth.

Sincere appreciations also extend to my committee members Prof. Milena Corredig

and Prof. Rickey Yada, who generously provided insightful suggestions and comments on

my project and this thesis, and supported me to pass my oral qualification.

I have a number of people to thank in Agriculture and Agri-Food Canada. I want

acknowledge Dr. Qi Wang, Dr. Ying Wu and Prof. Shao-Ping Nie for their knowledge and

sincere advice based on their knowledge and first-hand experience. I owe my special

thanks to technician John Nikiforuk, who put the most effort in obtaining the optimum

condition and all NMR data of my samples in this thesis, technician Ben Huang, Cathy

Wang for their technical support and willingness to share their experience, and technician

Edita Verespej (in Food Science Department, University of Guelph) for her technical

assistance. I extend my heartfelt gratitude to my friends and lab-mates, Dr. Qing-Bin Guo

Page 5: Structure-Function Relationship of Flaxseed Gum from

v

and Hui-Huang Ding. With heavy burden on their own shoulders, they took me as a

“brother” and constantly encourage and supported me in all ways throughout all adversity

regardless of my response. I wish I could have returned more. I am grateful to my

lab-mates Dr. Ji Kang, Dr. Xiao-Hui Xing and Dr. Jun-Yi Yin, and my close friend Dr.

Gui-Ying Mei for her trust and support and the memorable time we spent together.

Sincere acknowledgements also go to Prof. Fei Song, Prof. C. Bram Cadsby and their

adorable kids Mia and Benjamin, who treated me as a family member with caring respect

and trust. I thank my dear friend Jennifer Teng for her comfort and deep understanding of

my life, and my ex-roommates Xingyao Ling, Liang Li, Chen Ma, Yaning Shi for their

care and company. My heartfelt gratitude also turns to Yu Liu for sharing her genius sense

of humor and charming experiences of graduate study in her book Send You a Bullet, and

Steve Flowers and Dr. Bob Stahl for their effective treatments in regaining my confidence

and content provided in their book: Living with Your Heart Wide Open.

Endless acknowledgements extend to my kind and caring parents and elder brother

for providing me with the most freedom to push my boundaries with their unconditional

love and support.

Page 6: Structure-Function Relationship of Flaxseed Gum from

vi

Table of Content ABSTRACT ....................................................................................................................... ii

Acknowledgements .......................................................................................................... iii

Table of Content ............................................................................................................... vi

List of Tables ..................................................................................................................... ix

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

Chapter 1. Introduction ................................................................................................... 1

1.1 Significance of soluble dietary fibre and research focus ........................................... 1

1.2 Overall objectives ...................................................................................................... 3

Chapter 2. Literature Review .......................................................................................... 5

2.1 Flaxseed, soluble flaxseed gum and their health benefit ........................................... 5

2.2 Natural occurrence and extraction of flaxseed gum .................................................. 6

2.3 Fractionation of flaxseed gum and the structure of two fractions ............................. 8

2.4 Rheological properties of flaxseed gum .................................................................... 9

2.5 Structure-function relationship ................................................................................ 10

Chapter 3. Flaxseed Gum from Flaxseed Hulls: Extraction, Fractionation and

Characterization ............................................................................................................. 14

3.1. Introduction ............................................................................................................ 14

3.2. Methodology .......................................................................................................... 16

3.2.1 Materials ............................................................................................................ 16

3.2.2 Extraction of gum .............................................................................................. 16

3.2.3 Fractionation and purification of gum ............................................................... 17

3.2.4 Physico-chemical analysis ................................................................................. 19

Page 7: Structure-Function Relationship of Flaxseed Gum from

vii

3.2.5 Functional properties ......................................................................................... 22

3.3. Results and discussions .......................................................................................... 23

3.3.1 Extraction, fractionation and chemical composition ......................................... 23

3.3.2 Physical characterization ................................................................................... 26

3.3.3 Rheological properties ....................................................................................... 28

3.3.4 Functional properties ......................................................................................... 34

3.4. Conclusions ............................................................................................................ 39

Chapter 4. Structural Elucidation of the Acidic Fraction Gum ................................. 40

4.1 Introduction ............................................................................................................. 40

4.2 Experimental ........................................................................................................... 41

4.2.1 Materials ............................................................................................................ 41

4.2.3 Methylation analysis .......................................................................................... 42

4.2.3 NMR analysis ..................................................................................................... 42

4.3 Results and discussion ............................................................................................. 43

4.3.1 Methylation and GC-MS of partially methylated alditol acetate (PMAA) of the

acidic fraction gum ..................................................................................................... 43

4.3.2 1D and 2D NMR analysis of the acidic fraction gum ........................................ 45

4.4 Conclusions ............................................................................................................. 59

Chapter 5. Structure Elucidation of the Neutral Fraction Gum ................................ 60

5.1 Introduction ............................................................................................................. 60

5.2 Experimental ........................................................................................................... 61

5.2.1 Materials ............................................................................................................ 61

5.2.2 Methylation analysis .......................................................................................... 61

Page 8: Structure-Function Relationship of Flaxseed Gum from

viii

5.2.3 NMR analysis ..................................................................................................... 61

5.3 Results and discussion ............................................................................................. 62

5.3.1 Methylation and GC-MS of partially methylated alditol acetate (PMAA) of the

neutral fraction gum .................................................................................................... 62

5.3.2 1D and 2D NMR analysis of the neutral fraction gum ...................................... 64

5.4 Conclusions ............................................................................................................. 75

Chapter 6. Conformation of Neutral and Acidic Fraction Gums ............................... 76

6.1. Introduction ............................................................................................................ 76

6.2. Experimental .......................................................................................................... 78

6.2.1. Materials ........................................................................................................... 78

6.2.2. Solution preparation .......................................................................................... 78

6.2.3. Light scattering measurements ......................................................................... 79

6.3. Results and discussion ............................................................................................ 79

6.3.1 The presence and elimination of aggregates ...................................................... 79

6.3.2 Static light scattering .......................................................................................... 81

6.3.3 Dynamic light scattering .................................................................................... 83

6.4. Conclusions ............................................................................................................ 86

Chapter 7. Concluding Discussion ................................................................................. 88

Reference.......................................................................................................................... 91

Page 9: Structure-Function Relationship of Flaxseed Gum from

ix

List of Tables

Table 3.1. Yield and chemical components of flaxseed gum and its fractions (%) ........... 24

Table 3.2. Neutral monosaccharide in flaxseed gum and its fractions (%) ........................ 25

Table 3.3. The intrinsic viscosity ([η]), Huggins constant ( HK ), critical concentrations

(ccr) and coil overlap parameter values ((c[η])cr) at critical concentrations of SFG, NFG

and AFG ............................................................................................................................. 31

Table 4.1. Partially Methylated Aditol Acetate (PMAA) sugar residue derivatives of the

acidic fraction gum ............................................................................................................ 44

Table 4.2. 1H/13C NMR chemical shifts for the acidic fraction gum in D2O at 70oC (ppm)53

Table 4.3 Heteronuclear (HMBC) and homonuclear (TOCSY & NOESY) connectivities

in the acidic fraction gum (ppm) ........................................................................................ 54

Table 5.1. Partially Methylated Aditol Acetate (PMAA) sugar residue derivatives of the

neutral fraction gum ........................................................................................................... 63

Table 5.2. 1H/13C NMR Chemical Shifts for the neutral fraction gum (ppm) ................... 72

Table 5.3. Heteronuclear (HMBC) and homonuclear (NOESY) connectivities in the

neutral fraction gum (ppm) ................................................................................................ 73

Table 6.1. Molecular characteristics of the neutral (NFG) and acidic (AFG) fractions

obtained from static and dynamic light scattering ............................................................. 83

Page 10: Structure-Function Relationship of Flaxseed Gum from

x

List of Figures

Fig. 1.1. Structure of flaxseed .............................................................................................. 7

Fig. 1.2. The occurrence of mucilage in flaxseed ................................................................ 7

Fig. 3.1. Flow chart for extraction, fractionation and characterization of soluble flaxseed

gum .................................................................................................................................... 17

Fig. 3.2. Molecular weight distributions of polysaccharides (a) and their protein fractions

(b) ....................................................................................................................................... 27

Fig. 3.3. The steady shear flow curves (a) and dynamic rheological properties (b) of

soluble flaxseed gum and its two fractions ........................................................................ 28

Fig. 3.4. The intrinsic viscosity (a) and the dependence of viscosity on concentration (b-d)

of SFG, NFG and AFG ....................................................................................................... 32

Fig. 3.5. Reduction of surface tension of dispersions by soluble flaxseed gum and its

fractions .............................................................................................................................. 36

Fig. 3.6. The shifts of mean diameter (D[4,3]) and particle size distribution profiles of

emulsions stabilized by soluble flaxseed gum and its fractions within 24h after

emulsification with 2 % wt of canola oil ........................................................................... 38

Fig. 4.1. 1H (a) and 13C (b) NMR spectra of the acidic fraction gum (ppm) ..................... 45

Fig. 4.2. Key fragments of COSY spectrum of the acidic fraction gum ............................ 46

Fig. 4.3. Key fragments of TOCSY spectrum of the acidic fraction gum ......................... 47

Fig. 4.4. Key fragments of HMQC spectrum of the acidic fraction gum .......................... 48

Fig. 4.5. Key fragment of HMBC spectrum of the acidic fraction gum ............................ 49

Page 11: Structure-Function Relationship of Flaxseed Gum from

xi

Fig. 4.6. Key fragment of NOESY spectrum of the acidic fraction gum ........................... 50

Fig. 4.7. Schematic representations of the conventional (a) and recently proposed

alternative (b) structures of pectin ..................................................................................... 58

Fig. 4.8. Proposed repeating unit of the acidic fraction gum ............................................. 58

Fig. 5.1. 1H (a) and 13C (b) spectra of the neutral fraction gum ......................................... 65

Fig. 5.2. COSY spectrum of the neutral fraction gum ....................................................... 66

Fig. 5.3. TOCSY spectrum of the neutral fraction gum ..................................................... 67

Fig. 5.4. HMQC spectrum of the neutral fraction gum ...................................................... 68

Fig. 5.5. HMBC spectrum of the neutral fraction gum ...................................................... 69

Fig. 5.6. NOESY spectrum of the neutral fraction gum .................................................... 70

Fig. 5.7. Proposed repeating unit of the neutral fraction gum ........................................... 71

Fig. 6.1. Molecular size distribution of acidic fraction gums in different solvents ........... 80

Fig. 6.2. Static light scattering data presented in Zimm plot obtained from the neutral

(NFG, a) and acidic (AFG, b) fraction gums in 0.5 M NaOH at 25 °C ............................. 82

Fig. 6.3. The angular dependence (a: NFG, 0.2 g/L; c: AFG, 0.2 g/L) and concentration

dependence (at 90o, b: NFG; d: AFG) of the hydrodynamic radius (Rh) of neutral (NFG)

and acidic (AFG) fractions determined in 0.5 M NaOH. .................................................. 84

Fig. 6.4. Schematic representation of molecular structure (a & b) and conformation (c & d)

of neutral (NFG) and acidic (AFG) fraction gums ............................................................ 85

Page 12: Structure-Function Relationship of Flaxseed Gum from

1

Chapter 1. Introduction

1.1 Significance of soluble dietary fibre and research focus

The latest renewal of dietary fibre (DF) definition in the 2009 meeting of Codex

Alimentarius included natural and synthetic carbohydrate polymers that are resistant to

human gastrointestinal enzymes (Ciucanu et al., 1984; Eastwood et al., 1992; Kendall et

al., 2010). DF can be sub-divided into soluble (SDF) and insoluble (IDF) fractions based

on its hot water extractability. SDFs (except low molecular weight (MW) SDFs) exhibit

viscous and/or gelling properties, and consequently can attenuate postprandial blood

glucose and insulin levels and lower serum cholesterol (Dikeman et al., 2006; Wood,

2007) which are the leading preventive parameters against diabetes (Brennan, 2005) and

cardiovascular diseases (CVD) (Theuwissen et al., 2008).

Despite large variation in molecular structures of SDF, the common mechanism of

controlling postprandial serum glucose has been underlined: the hydration of

viscosity-altering SDFs (at reasonable intake levels) influences the digesta viscosity and

thus the ease of mixing, diffusion and adsorption, hence mediating glycemic and

insulinemic responses linked to the risks of type 2 diabetes (Dikeman et al., 2006). In

contrast, the cholesterol-lowering effects are primarily based on the capacity of SDF to

form a thick unstirred water layer that reduces the (re)absorption of intestinal fats

(cholesterol and bile acids) and/or binding bile acids during formation of micelles. The

latter will lead to an increased fecal excretion of bile acid which, in turn, may drive

Page 13: Structure-Function Relationship of Flaxseed Gum from

2

hepatic conversion of cholesterol into bile acids and absorption of serum low density lipid

(LDL)-cholesterol to compensate the hepatic free cholesterol pools (Denis et al., 2007;

Theuwissen et al., 2008). As a result, depletion of the body’s serum LDL-cholesterol

occurs. Another common hypothesis is that the increased luminal viscosity by the

hydration of SDF may impede the movement of cholesterol, bile acid and other lipid,

resulting in a lower absorption and higher excretion of these components (Eastwood et al.,

1992; Viuda-Martos et al., 2010).

Along with the above two beneficial physiological effects, short-chain fatty acids

(SCFA) produced by colonic fermentation of SDF (including low MW SDF) have drawn

increasing attention for their diverse roles in promoting health. SCFA (acetate, propionate,

and butyrate) acidify the colon environment, which may increase the development of

beneficial bacteria such as bifidobacteria and lactobacilli, reduce the pathogenic bacteria

(Swennen et al., 2006), as well as increase binding of bile acid and absorption of mineral.

Individually, propionate has been suggested to exhibit cholesterol-lowering properties by

offsetting the hyperlipidemic effect of acetate (Wong et al., 2006). Butyrate has

preventive effects on colon cancer and inflammation via regulating cell proliferation and

differentiation, although the mechanisms have not been clearly defined (Hamer et al.,

2008). The SCFA profile varies depending on both structure (Glitsø et al., 1999; Glitsø et

al., 2000; Rose et al., 2009) and average degree of polymerization (avDP) (Hughes et al.,

2007; Van Craeyveld et al., 2008), and also the essential enzymes existing in

corresponding bacterial varieties. Moreover, slowly fermented dietary fibres with

Page 14: Structure-Function Relationship of Flaxseed Gum from

3

sustaining SCFA production are desirable as most fermentation happens in the proximal

colon, yet most colonic disease occurs in the distal colon (Ferguson et al., 2000).

Increasing the intake of SDF has become recognized to prevent CVD, which remains

the leading cause of mortality in the world, especially in the lower socio-economic status

groups. Approximate one third (17.5 million) of people died from CVD in 2005

(Viuda-Martos et al., 2010). However, the challenge of increasing fibre consumption in

the western diet has been the balance between nutritional functionality and sensory

satisfaction. As SDF has to be incorporated at a significant concentration to be

physiologically effective (Giacco et al., 1998), meanwhile such high level of SDF, as a

thickener or gelling agent, will inevitably influence the sensory properties and induce an

over-texturization, high viscosity or phase separation. Various mechanical and

(bio)chemical approaches have been developed to improve the physiological and sensory

properties of SDFs via modifying their granule diameter and/or MW (Yu et al., 2008).

Due to the close relation between the possible mechanisms of each beneficial

physiological effect of soluble dietary fibre and its physicochemical characteristics, the

relationship between the structure of soluble fibres and their rheological behaviour or

physiological benefits has drawn increasing attention in both academic and industrial

fields for better explorations of novel fibres and their potential in applications.

1.2 Overall objectives

The renewed interest in flaxseed as a food source is due to its health benefits

Page 15: Structure-Function Relationship of Flaxseed Gum from

4

attributed to its components including lignans, α-linolenic acid, and flaxseed gum (SFG)

(Hall Iii et al., 2006). Our interests in flaxseed gum are primarily due to its low viscosity,

which is favoured for incorporation into food without an over-texturization, as well as the

high availability of sources and the ease of extraction. Differing from previous gum

extraction out of the whole seed, we extracted SFG from the mucilage enriched

by-product, flaxseed hulls, for lower costs and higher yield. To date, several previous

reports have been published on the structure of flaxseed gum (Naran et al., 2008) or its

two fractions(Cui et al., 1994a; Warrand et al., 2005c), yet their fine structures and

conformation are still not clear.

The overall objective of the current study is to investigate the structure-function

relationship of flaxseed gum and its fractions. The milestones of this work are:

1) To extract the gum from flaxseed hulls and to separate the soluble gum (SFG)

into neutral and acidic fractions using ion exchange chromatography

2) To examine the chemical, physicochemical, rheological, and functional

characteristics of SFG and its two fractions

3) To elucidate the linkage patterns and fine structure of neutral and acidic fractions

using methylation analysis and 1D/2D NMR spectroscopy

4) To investigate the conformation of neutral and acidic fractions using light

scattering

5) To reveal the relationship between the molecular structure, conformation and the

rheological behavior of neutral and acidic fractions

Page 16: Structure-Function Relationship of Flaxseed Gum from

5

Chapter 2. Literature Review

2.1 Flaxseed, soluble flaxseed gum and their health benefit

Flax (Linum usitatissiumum L.) was first introduced to North America as a fibre crop,

but its value and importance as an oil source quickly became apparent (Cunnane et al.,

1995). Canada is the leading country in producing and exporting oil-type flaxseed.

Western Canadian flaxseed is composed of 45% oil and 23% protein (1990-2008 means)

(www.grainscanada.gc.ca). The renewed interest in flaxseed as a food source is due to its

health benefits attributed to its components including lignans (secoisolariciresinol

diglucoside (SDG) being the predominant form), α-linolenic acid (ALA), and soluble

flaxseed gum (SFG) (Hall Iii et al., 2006). Flaxseed gum exhibited much lower viscosity

over a range of shear rate from 10 to 1,000 s-1, in comparison to locust bean gum, guar

gum and xanthan gum at the same concentration (0.3%) (Mazza et al., 1989). Its low

viscosity is favoured for fortification of fibre in food without leading to

over-texturization.

Although health benefits of flaxseed and its food application have been extensively

investigated (Bassett et al., 2009; Bloedon et al., 2004; Mentes et al., 2008; Prasad, 2009;

Rendon-Villalobas et al., 2009; Warrand et al., 2005b), very few investigations have

specifically focused on the role of soluble gum in promoting human health (Fodje et al.,

2009; Thakur et al., 2009). Based on in vitro test results, soluble flaxseed gum showed

higher bile acid binding capacity and generated higher amount of acetate and propionate

Page 17: Structure-Function Relationship of Flaxseed Gum from

6

compared with that of flax meal, wheat and rye bran, due to its significantly higher

soluble fibre content and solubility (Fodje et al., 2009). Inclusion of flaxseed in broiler

chicken diet was also found to be able to significantly increase the viscosity of ileal

digesta and the number of lactobacilli (Alzueta et al., 2003).

Despite the strong gel-forming ability of psyllium (Yu et al., 2003), the similarity in

molecular structure of the neutral fraction of flaxseed gum (NFG) and psyllium (Edwards

et al., 2003) implies that SFG may exhibit slow fermentability as psyllium does..

2.2 Natural occurrence and extraction of flaxseed gum

Flax mucilage (soluble flaxseed gum, SFG) occurs mainly at the outermost layer of

hull as shown in Fig. 1.1 (Thompson et al., 2003) & 1.2 (Cunnane et al., 1995).

This fibre-rich hull is able to release mucilaginous material (gum) easily when soaked

in water. A patented dehulling process (Cui & Han, 2006) could be performed to separate

flaxseed into a kernel (63%) fraction and a hull (37%) fraction (Cui, 2000). Therefore, to

reduce the costs and increase the gum yield, it is better to extract gum from the

by-product, hulls, but not from the whole flaxseed.

Page 18: Structure-Function Relationship of Flaxseed Gum from

7

Fig. 1.1. Structure of flaxseed

(A-C, three views of a flaxseed (approx.7x): A, seed; B, transverse section; C, longitudinal section. D,

transverse section (x400); E, longitudinal section (x400), showing testa, endosperm and cotyledon

cells. Abbreviations: cots, cotyledons; es, endosperm; g, germ; t, testa. Figure is from Thompson &

Cunnane, 2003)

Fig. 1.2. The occurrence of mucilage in flaxseed

(The line drawing of flaxseed is from Cunnane & Thompson, 1995)

Kernel

Mucilage Round cells Fiber cells Cross cells Pigment cells Endosperm

Seed Hull

Page 19: Structure-Function Relationship of Flaxseed Gum from

8

The composition and yield of flaxseed gum varies with agronomic practices,

climate, genotype, and extraction condition. A broad variation in extraction yield (3.6~8%)

from flaxseed was found due to different geographical regions and cultivars (Cui et al.,

1996b; Oomah et al., 1995). Soaking flaxseed with a water: seed ratio of 13 at 85 to 95°C

and pH from 6.5 to 7.0 for 3 h was found to be the optimum gum extraction condition to

achieve higher yield and quality, using response surface methodology (Cui et al., 1994b).

However, higher extraction temperature induced higher protein content and a brown color

(Mazza et al., 1989).

2.3 Fractionation of flaxseed gum and the structure of two fractions

Flaxseed gum is reported to be composed of a neutral (NFG) and an acidic (AFG)

fraction, with some proteinaceous contaminants. Fractionation of NFG and AFG from

SFG was performed using ion exchange chromatography (Cui et al., 1994a; Warrand et

al., 2003). NFG is composed of a high MW arabinoxylan (AX), whereas AFG is mainly

composed of a low MW pectin-like rhamnogalacturonan. The structure of SFG (Naran et

al., 2008) and its two fractions (Cui et al., 1994a) were revealed via methylation analysis.

In AFG, nonreducing terminal L-galactose is attached at the O-3 position of the

rhamnosyl residues instead of at the typical O-4 position, whereas the NFG has a highly

branched xylan backbone with nonreducing terminal L-arabinosyl units attached at the

O-2 and/or O-3 positions.

An alternative method has been published recently to separate neutral fraction from

Page 20: Structure-Function Relationship of Flaxseed Gum from

9

SFG via depectination using pectinase (Guilloux et al., 2009). The removal of acidic

fraction was not complete as indicated from the 1D NMR spectrum of the resultant

neutral fraction. However, compared with this method, separation with ion exchange

chromatography is tedious and time-consuming. Each method has its own advantages. To

evaluate the efficiency in separation of neutral from acidic fraction, the discrepancy in

composition, MW distributions and rheological properties of each fraction from different

methods needs to be compared. A fast and efficient fractionation will be favored if a

solution is available to improve either of the two methods.

2.4 Rheological properties of flaxseed gum

Flaxseed gums exhibit various rheological properties due to different cultivars,

extraction methods and measuring conditions. The viscosity of gum solutions exhibits

Newtonian-like behaviour at concentrations below 0.2% and pseudoplasticity at

concentrations above 0.2% (Mazza et al., 1989). Similar behaviour was also found in

crude (CFG), dialysed (DFG) and neutral (NFG) fractions. However, acidic (AFG)

fraction shows much lower viscosity and Newtonian-like behaviour at concentrations

from 0.3 to 2.0%. Small deformation oscillation rheological measurements of 2 % gum

solutions showed that CFG and DFG dispersions exhibited “weak gel” properties,

whereas NFG showed a viscoelastic fluid behaviour comparable to that of guar gum (Cui

et al., 1994a).

Intrinsic viscosity is a parameter reflecting the hydrodynamic volume occupied by the

Page 21: Structure-Function Relationship of Flaxseed Gum from

10

polymer, the value of which depends primarily on the molecular size and chain

conformation of the polymer, as well as the solution quality. Previous work (Cui et al.,

1996c) showed that the intrinsic viscosity (in 1 M NaCl) of flaxseed gums from five

genotypes varied from 434.0 to 657.8 mL/g. Among these cultivars, crude gum (CFG)

from the cultivar, Norman, was further fractionated into NFG and AFG. The values of

intrinsic viscosity (in 1 M NaCl) of CFG, NFG and AFG were 483.0, 530.4 and 248.4

mL/g, respectively. In contrast, the values of gum Arabic, guar gum and xanthan gum

were 14.4, 1135, and 1355 mL/g, respectively (Cui et al., 1996c). The intrinsic viscosity

of flaxseed gum decreases with the increase of ionic strength, and higher value (1030±20

mL/g) of intrinsic viscosity of flaxseed gum from flaxseed meals have been reported

when it was dissolved in MilliQ water (Goh et al., 2006).

2.5 Structure-function relationship

Polysaccharides have great diversity in physical properties including solubility,

water-holding capacity, flow behaviour, gelling potential, and surface or interfacial

properties. They provide various functions in food applications as stabilizers, thickening

and gelling agents, crystallization inhibitors, and encapsulating agents, etc. The

functional properties exhibited by food polysaccharides depend on their molecular

structures and shapes, molecular weight and molecular weight distributions, as well as

their concentrations (Wang et al., 2005a). The discussions below focus mainly on the

overall relationships between rheological properties of the polysaccharides and their four

Page 22: Structure-Function Relationship of Flaxseed Gum from

11

levels of structure.

Physical properties of a polysaccharide can be attributed to the various levels of

molecular organizations. The primary structure of a polysaccharide consists of its

monosaccharide composition, the ring size and anomeric configuration of each monomer,

the covalent linkage pattern and sequence, and substitution (Cui, 2005). The primary

structure is essentially rigid and is the basis for polysaccharide classification and their

three-dimensional shapes, in solid state and in solution. (Lapasin et al., 1995a; Morris et

al., 1979) In a homopolysaccharide chain, which contains only one type of monomeric

unit, four characteristic conformations (or secondary structures) can be recognized: type

A-extended ribbon-like chains, type B-hollow helices, type C-rigid and crumpled ribbons,

and type D-flexible coils or loosely joined chains. However, for most polysaccharides

composed of more than one sugar and/or more than one type of glycosidic bond, the

sequence can be periodic, interrupted or aperiodic and is likely to have an irregular

corresponding conformation. Energetically favorable non-covalent interactions, such as

van der Waals attractive forces, hydrogen bonding, charge and hydrophobic interactions

(Wang et al., 2005a), between shaped chains restrict chain flexibility and result in ordered

tertiary structures such as nested ribbons, multiple helices or aggregates of mixtures.

These tertiary structures exist in solution as the primary structural elements to form

aggregates and cause high viscosity or network formaiton. Tertiary structures may further

associate to establish higher molecular organization, often referred to as quaternary

Page 23: Structure-Function Relationship of Flaxseed Gum from

12

structures.

Molecular weight and molecular weight distribution influence the shape and size of

polysaccharide chains and their behaviour in solution (Launay et al., 1985). In dilute

solution, the intrinsic viscosity, a measure of the hydrodynamic volume occupied by an

isolated polysaccharide, depends primarily on molecular weight, chain rigidity and

solvent quality. On the other hand, the viscosity of a concentrated solution mainly derives

from the interactions or fluctuating entanglement between chains. Hence, the low

shear-rate Newtonian viscosity η0 of undiluted solution is higher and increases more

drastically with concentration for high-molecular-weight polysaccharides than

low-molecular polysaccharides.

The aforementioned four main regular chain-shape types were also found somehow to

correlate with particular biological functions (Rees et al., 1971). Skeletal polysaccharides

capable of forming fibrous aggregates by tight packing are usually of type A-extended

ribbon-like shaped chains. They are mainly derived from 1,4-linkages, which are

prevalent in natural materials such as cellulose, chitin and xylans. Storage and network

polysaccharides are often of type B-hollow helices mainly composed of α-1,4-glucans

such as amylase, amylopectin, and β-1,3-glucans. The β-glucans from cereal grains

contain both periodic chain portions of type A (β-1,4-linkages) and type B (β-1,3-linkages)

and can serve structural and reserve functions during dormancy and germination,

respectively. Chains of type C shape are unnatural and rare in natural polysaccharides.

Page 24: Structure-Function Relationship of Flaxseed Gum from

13

Loosely joined polysaccharides with type D shapes are mainly composed of

1,6/5-linkages and widely found in branched polysaccharides such as amylopectin and

various plant gums. These 1,5/6-linked residues introduce “flexible joints” into bush-like

molecules and result in a sponge-like instead of stiff textures.

Page 25: Structure-Function Relationship of Flaxseed Gum from

14

Chapter 3. Flaxseed Gum from Flaxseed Hulls: Extraction,

Fractionation and Characterization

3.1. Introduction

Flax (Linum usitatissiumum L.) was first introduced to North America as a crop for

structural fibres, but its value and importance as an oil source quickly became apparent

(Cunnane et al., 1995). Canada is the leading country in producing and exporting oil-type

flaxseed. Western Canadian flaxseed is composed of 45% oil and 23% protein (1990-2008

means) (www.grainscanada.gc.ca). The renewed interest in flaxseed as a food source is

due to its health benefits attributed to its components including lignans

(secoisolariciresinol diglucoside (SDG) being the predominant form), α-linolenic acid,

and soluble flaxseed gum (SFG) (Hall Iii et al., 2006). In comparison to locust bean gum,

guar gum and xanthan gum at a concentration of 0.3 % (w/v), SFG exhibited much lower

viscosity over a range of shear rate from 10 to 1,000 s-1 (Mazza et al., 1989). Its low

viscosity might be favoured in dietary fibre fortification in food without leading to an

over-texturization, when a significant concentration of fibre is required to show health

benefits.

In vitro fermentation results showed that SFG exhibited higher bile acid binding

capacity and generated higher amount of acetate and propionate compared with that of

equivalent amount of flax meal, wheat and rye bran, due to its significantly higher soluble

fibre content and solubility (Fodje et al., 2009). Inclusion of flaxseed in the diet of broiler

Page 26: Structure-Function Relationship of Flaxseed Gum from

15

chickens significantly increased the viscosity of ileal digesta and the number of lactobacilli

(Alzueta et al., 2003). The high bile acid binding capacity will lead to an increased fecal

excretion of bile acid, which may lower serum cholesterol (Denis et al., 2007;

Theuwissen et al., 2008). The high short-chain fatty acids (acetate, propionate and

butyrate) productivity and selective stimulation of growth and/or activity of probiotics

(Gibson et al., 2004) of SFG also showed its potential as a good source of prebiotics. The

abundance of two predominant bacteria divisions in the human gut, the Bacteroidetes and

the Firmicutes, were found to have positive and negative correlation with percentage loss

of body weight (Ley et al., 2006).

Flax mucilage (soluble flaxseed gum, SFG) occurs mainly at the outermost layer of

hull. This fibre-rich hull is able to release mucilaginous material (soluble gum) easily

when soaked in water. Earlier research analyses were based on the gum extracted from

the whole seed (Cui et al., 1994a; Diederichsen et al., 2006; Mazza et al., 1989;

Muralikrishna et al., 1987; Naran et al., 2008; Oomah et al., 1995; Warrand et al., 2003,

2005a, 2005c) or flax meal (Fedeniuk et al., 1994; Mueller et al., 2010). With the success

of a patented dehulling process (Cui et al., 2006), the whole flaxseed could be efficiently

separated into a kernel (~63%) fraction and a hull (~37%) fraction (Cui, 2000) in large

scale. Investigating the composition, structure, physicochemical and rheological

properties of SFG from hulls will assist in exploring its potential for the food industry.

In this current study, the gum (SFG) extracted from flaxseed hulls was further

separated into neutral (NFG) and acidic (AFG) fraction gums using ion exchange

Page 27: Structure-Function Relationship of Flaxseed Gum from

16

chromatography (IEC). The chemical (neutral monosaccharide composition, uronic acid

and protein content), physical (molecular weight distribution and heat stability),

rheological (viscosity, viscoelasticity, intrinsic viscosity and critical concentration,) and

functional (surface tension and emulsification) characteristics of SFG and its two

fractions are reported.

3.2. Methodology

3.2.1 Materials

Flaxseed hull (variety Bethune) was supplied by Natunola Health Biosciences

(Winchester, Ontario, Canada).

3.2.2 Extraction of gum

The extraction of the gum was conducted at room temperature as shown on the flow

chart in Fig. 3.1.

A batch of 500 g of flaxseed hull was soaked in 6 L of distilled water overnight

under gentle stirring. A coarse filtration followed using cheesecloth to separate the hull.

The filtered mucilage was collected and centrifuged (Beckman Coulter, Mississauga,

Ontario) at 27,000 g and 25°C for 25 min, to eliminate insoluble particles. The

supernatant was then thoroughly mixed with one volume of 100% ethanol to precipitate

the polysaccharide. The precipitate was recovered by centrifugation of the mixture at

3,000 g for 10 min and redissolved in distilled water at approximately 4-5 (w/v) % before

lyophilization.

Page 28: Structure-Function Relationship of Flaxseed Gum from

17

Fig. 3.1. Flow chart for extraction, fractionation and characterization of soluble

flaxseed gum

(SFG: soluble flaxseed gum; NFG and AFG: neutral and acidic fraction gum; IEC: ion-exchange

chromatography; Buffer 1: 20 mM Tris/HCl, pH 8).

3.2.3 Fractionation and purification of gum

3.2.3.1 Fractionation

The procedure of fractionation followed that of Warrand et al. (2003) with minor

modification: the washing solution of 0.1 M NaOH was replaced by 2 M NaCl and 1M

NaOH. A flow chart of the fractionation procedure is also shown in Fig. 3.1. For each

fractionation, 0.35 g of SFG was dissolved in 1 L of buffer 1 (20 mM Tris/HCl, pH 8) at

80°C for 1 h. After cooling down to room temperature, the SFG solution was loaded onto

the pre-equilibrated column (XK 50 Column, GE Healthcare, packed with 900 mL of

Q-Sepharose fast flow as matrix) at a flow rate of 10 mL min-1 for 100 min and flushed

Concentration (vacuum, 50 oC)

Functional Properties (Surface property; Emulsification)

Page 29: Structure-Function Relationship of Flaxseed Gum from

18

continuously with 2 L of buffer 1 for 200 min. The eluate (denoted as NFG) was collected

between 25 and 150 min after the sample loading, when ~90 % of NFG was eluted out.

The column was successively washed with 1 L of buffer 2 (20 mM Tris/HCl + 1 M NaCl,

pH 8) at the same flow rate (10mL min-1). The eluate (denoted as AFG) was collected

between 55 and 110 min after buffer 2 was loaded, during which period most of the acidic

fraction (AFG) was eluted out. The column was sequentially flushed with 1 L of each of

the following three washing solutions: 2 M NaCl, 1 M NaOH and 2 M NaCl, and

re-equilibrated with 2-3 L of buffer 1 for next use. The acidic and neutral fractions were

then concentrated by rotary vacuum evaporator at 50°C, dialyzed (MW cut-off 3.5 kDa,

Fisher Scientific) against distilled water at 25°C for 72 h, and freeze-dried.

3.2.3.2 Purification

The protein fractions in both SFG and AFG were removed via protease hydrolysis,

and the two resultant fractions were denoted as SFGnP and AFGnP. Aqueous solutions of

SFG or AFG (1%, wt) were prepared by dissolving the polysaccharide in phosphate

buffer (80 mM, pH 7.5) at 80°C for 1 h with constant stirring before cooling down to

60°C. Stock solution of purified protease ( Megazyme cat. no. E-BSPRT, 350 tyrosine U

mL-1) was mixed thoroughly (0.2 mL g-1 gum or 70 U g-1 gum) with the gum solution and

incubated at 60°C for 30 min. The mixture was heated at 80°C for 10 min to inactivate

the enzyme. The resultant solution was dialyzed (MW cut-off 3.5 kDa, Fisher Scientific)

against distilled water at 25°C for 72 h, and freeze-dried.

Page 30: Structure-Function Relationship of Flaxseed Gum from

19

3.2.4 Physico-chemical analysis

Monosaccharide analysis of SFG and its fractions was conducted on a high

performance anion exchange chromatography system (Dionex, DX500 Sunnyvale, CA,

USA) coupled with a pulsed amperometric detector (HPAEC-PAD) (Cui et al., 2000).

Hydrolysis was conducted as previously published (Wu et al., 2009). Uronic acid analysis

was conducted using the m-hydroxyphenyl colorimetric method (Blumenkrantz et al.,

1973). Protein analysis was conducted using NA2100 Nitrogen and Protein Analyzer

(Thermo Quest, Milan, Italy). The protein content (%) was obtained by multiplying the

nitrogen content (%) by 6.25. All chemical analyses were completed in triplicate.

The molecular weight (MW) distribution and heat stability of SFG, NFG and AFG

were measured using an HPSEC system consisting of an HPLC system with degasser,

autosampler and refractive index detectors (Optilab Rex, Santa Barbara, CA, USA). To

eliminate baseline variations, the mobile phase (100 mM sodium nitrate + 5 mM Sodium

azide) was degassed and filtered through 0.2 µm filters (Millipore, Fisher Scientific).

Before analysis, the gums were dissolved in the mobile phase at 1 g L-1, and heated at

80°C for 1 h under gentle stirring. To measure the heat stability of gums, heating duration

was extended to 3 h at 80°C. Samples were filtered through disposable 0.45 µm filters

(Millipore, Fisher Scientific) before injection, and aliquots (100 µL) of samples were

injected into the column (Shodex OHpak SB-806M HQ, 7.8 mm ID×300 mm, column

temperature at 40 oC) and eluted at a flow rate of 0.6 mL min-1 for 60 min. Pullulan 800

(MW 788,000), 400 (MW 404,000), 100 (MW 112,000), 50 (MW 47,300), 20 (MW

Page 31: Structure-Function Relationship of Flaxseed Gum from

20

22,800) were used as MW markers to calibrate the column.

To investigate the association of protein and polysaccharides, aliquots (20 µL) of 1g

L-1 gum solutions prepared at the same condition as for MW distribution were injected into

a HPLC system (Waters 600E, Waters Ltd., Toronto, ON, Canada) with refractive index

detector (Waters 410) coupled with UV detector (Waters 486) and with the same column

for MW distribution but at column temperature of 25 oC. Separation was carried out at a

flow rate of 1.0 mL/min in the mobile phase (100 mM sodium nitrate + 5 mM sodium

azide).

Both steady shear and oscillation measurement of flaxseed gum and its fractions

were carried out on a strain-controlled rheometer (ARES, TA Instruments, New Castle,

DE, USA) using parallel-plate geometry (50 mm diameter, 1 mm gap size). The viscosity

of sample solutions of six concentrations (1, 5, 10, 15, 20, 25 g L-1) was measured at

25°C in shear rates ranging from 0.01 to100 s-1. Viscoelastic properties of samples were

measured at 20 g L-1 and 25 °C.

Intrinsic viscosity and critical concentration measurement followed that of (Doyle et

al., 2009). Gum dispersions were heated at 80°C for 1h and cooled down. Salt solutions

(0.2 M) were prepared individually before mixing with gum solutions under constant

stirring at ambient temperature, and salt concentration in each final sample was adjusted

to 0.1 M. Viscosity of solutions was measured using double wall couette geometry (inner

and outer cup radii were 27.94 and 34 mm; inner and outer bob radii were 29.5 and 32

mm, and 1 mm gap) on a strain-controlled rheometer (ARES, TA Instruments, New

Page 32: Structure-Function Relationship of Flaxseed Gum from

21

Castle, DE, USA). The sample (7.5 mL) was left unperturbed for 15 min before each

measurement. All measurements were made in duplicate at 25 °C.

Intrinsic viscosity ( ][η ) was calculated using the following relationships (Harding,

1997).

srel ηηη /= (3.1)

1−= relsp ηη (3.2)

cKc Hspred ⋅+== 2][][ ηηηη (3.3)

cKc Krelinh ⋅+== 2][][)ln( ηηηη (3.4)

where η and sη are the zero shear viscosities (the constant apparent viscosities in a

limited Newtonian region at very low shear rates) of the solution and solvent (MilliQ

H2O); relη and spη are two dimensionless parameters of relative and specific viscosity;

and redη and inhη represent reduced viscosity and inherent viscosity, respectively.

HK (generally positive) and KK (generally negative) are Huggins and Kraemer constants

in Huggins and Kraemer equations (Eqs 3.3 and 3.4). Both equations are valid only for

solution viscosity up to twice of the solvent viscosity (i.e. 2≤relη ). Beyond this region,

with concentration increase polymer-polymer interactions will progressively become

significant, thus higher-order terms (c2, c3, etc.) will no longer be negligible. A lower

limit of 2.1≥relη is also required to eliminate the increasing errors in spη as spη

approaches 0. Concentrations of each gum fraction adopted were confined to this region

( 22.1 ≤≤ relη ). The value of intrinsic viscosity ( ][η ) was reported as the mean of both

Page 33: Structure-Function Relationship of Flaxseed Gum from

22

intercepts in linear functions (3) and (4) by plotting cspη and

crelηln against c and

extrapolating each linear trendline to zero concentration.

3.2.5 Functional properties

3.2.5.1 Surface activity

The surface tension of the air-water interface was measured by the Du Nouy ring

method using the Fisher Surface Tensiomat (model 21, Fisher Scientific, Nepean, ON)

(Izydorczyk et al., 1991). SFG and its fractions were dissolved in 20 mL of distilled water

at various concentrations, heated at 80ºC for 1 h under gentle stirring and then left

unperturbed for 2 h before measurements. The surface tension (dynes cm-1) of distilled

water and gum solutions were measured every 5 min for 6 times at 25ºC.

3.2.5.2 Emulsifying properties

Emulsions were prepared by mixing canola oil (2 % wt) with SFG or its fraction

dispersions (preheated at 80°C for 1 h, final gum concentrations were adjusted to 0.44 %

wt) using Polytron (Kinematica GmbH, Brinkmann homogenizer, Switzerland) at medium

speed (level 5) for 5 min, followed by homogenization (Nano DeBEE Electric Bench-top

Lab Homogenizer, BEE International, South Easton, MA, USA) with 2 passes at 5,000

psi.

The particle size distribution of the emulsions was measured using integrated light

scattering (Mastersizer X, Malvern Southborough, MA) (Khalloufi et al., 2008). The

measurements were performed within one day after emulsion preparation as phase

Page 34: Structure-Function Relationship of Flaxseed Gum from

23

separation occurred in all the emulsions within 24 h. The emulsifying stability was

determined by comparing the shifts of the oil droplet size distributions.

Duncan’s multiple-range tests were conducted using SAS software to examine the

changes in monosaccharide composition of soluble flaxseed gum or the acidic fraction

before and after protein removal. Thus the influence of protein removal processing on

monosaccharide composition of each gum fraction could be evaluated.

3.3. Results and discussions

3.3.1 Extraction, fractionation and chemical composition

The composition and yield of flaxseed gum has been reported to vary with extraction

conditions, as well as culture environment and genotype. Gum yield from flaxseed

increased from 4 to 9.4 % when extraction temperature increased from 25 to 80 oC

(Fedeniuk & Biliaderis, 1994). Broad variations of 3.6~8.0 % among 109 cultivars (80 oC,

2h) and 5.4~7.9 % among 12 genotypes (85 oC, 3h) were also found by Oomah et al.

(1995) and Cui et al. (1996), respectively. Soaking flaxseed with a water: seed ratio of 13

at 85 to 95°C and pH from 6.5 to 7.0 for 3 h was found to be the optimum gum extraction

condition to achieve higher yield and quality, using response surface methodology (Cui et

al., 1994b). However, higher extraction temperature induced higher protein content and a

brown color (Fedeniuk and Billiaderis, 1994). Considering the browning may affect the

properties of the polysaccharides, in the present work, the extraction of the gum was

conducted at room temperature. The yield and chemical composition of SFG, NFG and

Page 35: Structure-Function Relationship of Flaxseed Gum from

24

AFG are shown in Table 3.1.

Table 3.1. Yield and chemical components of flaxseed gum and its fractions (%)

SFG NFG AFG

Yield 9.7 ± 0.3 a 23.2 ± 0.9 b 27.3 ± 3.0 b

Protein 11.8 ± 1.0 nd 8.1 ± 0.4

Uronic Acid 23.0 ± 0.1 1.8 ± 1.0 38.7 ± 1.0

(SFG: soluble flaxseed gum; NFG and AFG: neutral and acidic fraction gum; a: yield was based on

flaxseed hull mass; b: yield was based on SFG mass; nd: not detectable; data were on a dry basis)

The yield of SFG is 9.7 % of the hull mass. The recovery of the chromatographic

fractionation was approximately 50 %; the loss of half of the SFG may partially be due to

physical blockage of large particles in the column and/or insufficient unbinding process

while washing AFG out. The uronic acid content in AFG was about 20 fold of that in

NFG, suggesting an efficient separation of NFG from AFG. NFG was free of protein,

while AFG (8.1%) contained less protein than SFG (11.8%). After protein removal from

SFG and AFG, SFGnP and AFGnP were obtained with their protein content decreased to

an undetectable level.

Relative neutral monosaccharide composition of SFG and its fractions is shown in

Table 3.2.

Page 36: Structure-Function Relationship of Flaxseed Gum from

25

Table 3.2. Neutral monosaccharide in flaxseed gum and its fractions (%)

SFG SFGnP NFG AFG AFGnP

Fucose 7.0 ± 0.2c 7.0 ± 0.4c nd 14.7 ± 0.3b 16.5 ± 0.1a

Rhamnose 16.5 ± 0.6c 14.5 ± 0.3c nd 38.3 ± 2.3a 32.8 ± 0.7b

Arabinose 12.7 ± 0.1b 12.6 ± 0.5b 20.2 ± 0.1a 2.9 ± 0.4c 3.0 ± 0.1c

Galactose 22.4 ± 1.0c 22.5 ± 0.7c 7.9 ± 0.1d 35.2 ± 0.4b 39.7 ± 0.4a

Glucose 2.7 ± 0.1b 3.1 ± 0.3ab 3.7 ± 0.4a nd nd

Xylose 38.6 ± 1.2b 40.9 ± 2.1b 68.2 ± 0.6a 8.9 ± 1.3c 7.9 ± 0.3c

(SFG: soluble flaxseed gum; NFG and AFG: neutral and acidic fraction gum; data are given as

“mean±SD”, n=3; data in the same row followed with the same superscript letter are not significantly

different by Duncan’s multiple-range test (p≤0.05); nd: not detectable.)

No significant differences were found between SFG and SFGnP, implying the

removal of protein via enzymatic hydrolysis from these two fractions did not modify their

neutral sugar composition. However, the same enzymatic hydrolysis procedure changed

the three main components in AFG: small portion of rhamnose was possibly removed

causing the relative increase in the other two (galactose and fucose). These changes in

SFG and SFGnP were not significant, since AFG occupied only ~27 % of SFG. The

neutral monosaccharide composition of SFG, NFG and AFG was comparable with

previous results (Cui et al., 1994a; Fedeniuk et al., 1994). NFG was mainly composed of

Page 37: Structure-Function Relationship of Flaxseed Gum from

26

xylose (68.2 %) and arabinose (20.2 %) therefore being identified as arabinoxylans

(Warrand et al., 2003). Its minor components included galactose (7.9 %) and glucose

(3.7 %), but no rhamnose or fucose. In contrast, AFG mainly consisted of rhamnose

(38.3 %), galactose (35.2 %) and fucose (14.7 %). With most galactose and all fucose

existing as terminals in side chains (Naran et al., 2008), AFG was referred to as

rhamonogalacturonans for its high content of rhamnose and galacturonic acid (38.7 %,

Table 1.). The presence of small amounts of xylose (8.9 %) and arabinose (2.9 %) in AFG

might be due to the residual fraction of NFG (Cui et al., 1994), or might be derived from

side chain components covalently linked to its backbone (Naran et al., 2008).

3.3.2 Physical characterization

The molecular weight (MW) distribution of SFG, NFG and AFG are shown in Fig.

3.2(a). NFG consisted of one fraction with high MW of approximately 1,470 kDa. By

contrast, AFG consisted mainly of 3 portions of MW: 1,510, 341 and 6.6 kDa, respectively.

When subjected to heat at 80°C from 1 to 3 h, the MW of NFG decreased from 1,470 to

850 kDa. Similarly, a significant reduction of intermediate MW (340 kDa) polymers and

the appearance of additional small MW molecules (2.6 kDa) also implied the degradation

of AFG.

Page 38: Structure-Function Relationship of Flaxseed Gum from

27

Fig. 3.2. Molecular weight distributions of polysaccharides (a) and their protein

fractions (b)

(SFG: soluble flaxseed gum; NFG and AFG: neutral and acidic fraction gum; the DRI and UV curves

exhibited the molecular weight distribution of polysaccharide and protein, respectively; the unit of

numbers in (a) is kDa; the column temperatures were 25 and 40 oC in (a) and (b), respectively.)

The UV and DRI response in Fig. 3.2(b) depicts the association of protein and

polysaccharides. No protein was detected in NFG. The major peak of high-MW protein

eluted at around 32 min, substantially later than two major high-MW polysaccharide

peaks (at ~26 and ~30 min, respectively) in both SFG and AFG. The differences in elution

time between these fractions of protein and polysaccharides suggested that most of the

protein fractions were not covalently linked to the polysaccharides in both SFG and AFG,

although trace amount of low-MW protein and polysaccharides were eluted out

simultaneously at around ~37min, making it difficult to reveal their association. That the

Page 39: Structure-Function Relationship of Flaxseed Gum from

28

protein could be completely removed by protease hydrolysis (in Section 3.3.1) also

supported this conclusion, as it was reported that inherent protein could be very difficult

to remove (Garti et al., 1994).

3.3.3 Rheological properties

3.3.3.1 Viscosity and viscoelasticity

Flaxseed gum exhibited relatively low viscosity. At a concentration of 0.3 % (w/v),

the viscosity of flaxseed gum was only around half of that of guar gum and locust bean

gum (Mazza et al., 1989). The steady shear flow curves and dynamic rheological

properties of SFG and its fraction solutions at 25°C from this study are shown in Fig. 3.3.

Fig. 3.3. The steady shear flow curves (a) and dynamic rheological properties (b) of

soluble flaxseed gum and its two fractions

(SFG: soluble flaxseed gum; NFG and AFG: neutral and acidic fraction gum.)

SFG and NFG exhibited pseudoplastic flow behaviour at concentrations above 0.5

and 1.0 %, respectively, over a wide range of shear rate; whereas the steady shear flow

Page 40: Structure-Function Relationship of Flaxseed Gum from

29

curves of AFG were more Newtonian-like at all concentrations examined. This was in

good agreement with those reported earlier by Cui et al. (1994). Mazza et al. (1989) also

found that flaxseed gum solutions exhibited pseudoplastic flow behaviour and

Newtonian-like behaviour at concentrations below and above 0.2 %, respectively.

All three dispersions showed liquid-like behaviour, as the loss modulus (G″)

exceeded the storage modulus (G') over the entire frequency range investigated (Fig. 3.3b).

In previous work by Cui et al. (1994), dispersions of crude, dialyzed flaxseed gums and

neutral fractions all exhibited “weak gel” properties. The discrepancy may be due to the

different sources of raw materials and extraction procedures.

3.3.3.2 Intrinsic viscosity

Intrinsic viscosity, denoted as [η], is a parameter reflecting the hydrodynamic volume

occupied by the polymer. The value of [η] primarily depends on the molecular size and

chain rigidity of the polymer, as well as the solution quality (Lapasin et al., 1995b).

Intrinsic viscosity increases with molecular weight (MW) according to the

Mark-Houwink relationship,

αη KM=][ (3.5)

where K is a constant, M is the average MW and, α is an exponent constant related to

the chain flexibility. For polyelectrolytes, the presence of counterions in the solution will

affect their intrinsic viscosity primarily through the reduction of intramolecular repulsion

among groups with the same charges. With increasing concentration of salts (i.e.

increasing ionic strength), the [η] of both SFG and AFG will decrease significantly as

Page 41: Structure-Function Relationship of Flaxseed Gum from

30

they contain acidic groups (see uronic acid content in Table 1). The relation between [η]

and ionic strength has been proposed by Pals and Hermans as follows (Harding, 1997):

5.0][][ −∞ += SIηη (3.6)

where ∞][η is the intrinsic viscosity at infinite ionic strength, I is the ionic strength, and

S is a criterion only for the comparison of stiffness among polymers with the same MW,

as well as in the same solvent counterion environment. The “Smidsrod” stiffness

parameter, B, is defined by

υη )]([ 1.0=⋅= IBS (3.7)

where 1.0][ =Iη is the intrinsic viscosity at 0.1 M ionic strength, and 1.03.1 ±=υ . This

modified parameter B could be used to compare the relative stiffness of polymers even

without knowing their MW. It has been reported that flaxseed polysaccharide was

between a flexible and semi-flexible polymer with a B value of ~0.018, which lies

between ~0.005 (for xanthan gum with stiff conformation) and ~0.045-0.065 (for

carboxymethylcellulose with a flexible chain) (Goh et al., 2006). As a rough guide,

Huggins constant, HK , also reveals the general conformation of a polymer. The value of

HK can be as high as ~2 for uncharged spheres. Lower value is expected for more

extended biopolymer, such as ~0.35 for flexible biomolecules (Harding, 1997). Goh et al.

(2006) reported the HK value for flax meal polysaccharide equal to 0.34±0.05, which

also implied its flexible conformation. The values of HK for SFG, NFG and AFG are

shown in Table 3.3. NFGHK , (0.54) was slightly higher than SFGHK , (0.48) and both were

Page 42: Structure-Function Relationship of Flaxseed Gum from

31

much higher than AFGHK , (0.16), implying the chain flexibility of SFG was between that

of NFG and AFG, as SFG is composed of both fractions.

Table 3.3. The intrinsic viscosity ([ŋ]), Huggins constant ( HK ), critical

concentrations (ccr) and coil overlap parameter values ((c[ŋ])cr

) at critical

concentrations of SFG, NFG and AFG

(SFG: soluble flaxseed gum; NFG and AFG: neutral and acidic fraction gum)

The Huggins and Kraemer plots of SFG, NFG and AFG are shown in Fig. 3.4(a) and

the intrinsic viscosities ([η], in 0.1 M NaCl) of SFG, NFG and AFG are 446.0, 377.5 and

332.5 mL g-1, respectively (Table 3.3).

NFG had a larger intrinsic viscosity than AFG, probably due to its higher molecular

rigidity and average MW (Fig. 3.2). Although SFG had similar molecular rigidity to NFG

and a similar MW distribution to AFG (both contained low-MW fraction), it had larger

intrinsic viscosity than both. This might be due to the presence of the synergistic

SFG NFG AFG

[ŋ] (mL g-1) 446.0 ± 3.4 377.5 ± 5.7 332.5 ± 6.8

HK / 0.48 0.54 0.16

ccr (g L-1) 6.61 6.68 6.43

(c[ŋ])cr / 2.95 2.52 2.14

Page 43: Structure-Function Relationship of Flaxseed Gum from

32

interaction between NFG and AFG. Higher intrinsic viscosities of deacetylated

xanthan-guar mixtures have been reported in comparison with those expected on the

assumption of no synergistic interactions (Khouryieh et al., 2007). Goh et al. (2006)

reported a linear relation between [η] of flaxseed gum and 5.0−I as follows:

5.05.0 )9.05.52()4289(][][ −−∞ ⋅±+±=+= ISIηη , according to which the value of [η] in

0.1 and 1 M NaCl would be approximately 455 and 342 mL g-1, respectively. The former

value was comparable to the [η] of SFG (446.0 mL g-1 in 0.1 M NaCl), but higher values

of [η] (in 1 M NaCl) ranged from 434.0 to 657.8 mL g-1 for five genotypes of flax were

also reported (Cui et al., 1996c).

Fig. 3.4. The intrinsic viscosity (a) and the dependence of viscosity on concentration

(b-d) of SFG, NFG and AFG

(SFG: soluble flaxseed gum; NFG and AFG: neutral and acidic fraction gum; viscosity was measured in 0.1

M NaCl (a-d). Reduced viscosity (ηred = ηsp/c) and inherent viscosity ( ηrel = (ln ηrel)/c) versus concentration

are plotted with empty and solid symbols (a), respectively; the common intercept gives [η], and the positive

and negative slopes are KH[η]2 and KK[η]2, where KH and KK are Huggins and Kraemer constant,

respectively (a). n1 and n2 represent the slope of each region; the gum concentration for the circles within

the squares in figure (b-d) equaled 5 g L-1) .

Page 44: Structure-Function Relationship of Flaxseed Gum from

33

3.3.3.3 Critical concentrations

The dependence of viscosity on concentration and intrinsic viscosity could be

expressed by a power law relationship (Morris et al., 1981a):

nsp ck ])[( ηη ⋅⋅= (3.8)

where spη is the “zero-shear” specific viscosity and can be obtained by Eq (3.1) and

(3.2); the degree of space occupancy (c[η]) is also called the coil overlap parameter, in

which c is the concentration and [η] is the intrinsic viscosity; the value of (lg k) and n

equal the extrapolated interception and the slope of the double logarithmic plots (Fig. 3.4

(b-d)) for SFG, NFG and AFG. A drastic change of slope was found in each of the three

fractions (Fig. 3.4 (b-d)) at a specific degree of space occupancy ((c[η])cr). The

corresponding concentration at this point is called the critical concentration (ccr).

Solutions with concentrations below and above ccr are referred to as dilute and

semi-dilute solutions, respectively. In dilute solution, the increase in viscosity is caused

by interfering the flow of solvent by the separated polymer chains and the increment is

proportional to the increment in volume fraction occupied by these polymers; rheological

behaviour in semi-dilute solutions, however, primarily depends on the polymer-polymer

interaction, as the overlap and interpenetration of neighbouring polymer chains are no

longer negligible (Morris et al., 1981b). In practice, there is a transition zone around the

cross of the two linear trendlines for dilution and semi-dilution regions, where the data

points of viscosity obtained from measurements slightly depart from the expected values

Page 45: Structure-Function Relationship of Flaxseed Gum from

34

on each trendline. Therefore critical concentration is an approximate measure for the

onset of overlap and entanglement between polymer chains. The critical concentrations

(ccr, mL g-1) for SFG (~6.61), NFG (~6.68) and AFG (~6.43) were similar to each other,

irrespective of their significant differences in intrinsic viscosities (Table 3.3), and the

corresponding values of (c[η])cr fell within a range from 2.14 to 2.95, not far away from

the values reported between 2.5 and 4 (Morris et al., 1981a). The values of n1 were

confined to 1.08-1.21 (Fig. 3.4) and were comparable with the value (1.1~1.4) reported

(Doyle et al., 2009; Morris et al., 1981; (Wang et al., 2005b)). The n2 for SFG (3.28) and

NFG (3.51) were close to the typical value (~3.3) for various disordered polysaccharides

(Morris et al, 1981), though higher value (n2 = ~5) was also found in the chain of several

galactomannans (Doyle et al., 2009) due to the “hyperentanglement” of unsubstituted

mannan regions. Compared with SFG and AFG, a much lower value of n2 (2.17) detected

for AFG may probably also be due to its relatively higher flexibility and thus be less

likely to form an entanglement network, just as implied from its lower value of

AFGHK , (0.16).

3.3.4 Functional properties

Most water soluble polysaccharides are known to act as stabilizers in oil-in-water

emulsions, only a few can act as emulsifiers mainly due to the presence of hydrophobic

moieties, such as proteinaceous materials. Gum arabic is the most used polysaccharide as

an emulsifier because of the presence of proteinaceous moieties in its structure. The

Page 46: Structure-Function Relationship of Flaxseed Gum from

35

polysaccharide chains can then absorb to the oil phase via the protein structure (Garti et

al., 2001). By contrast, a protein-rich fraction, which was separated from crude guar gum,

contained higher protein (10.6 %) but showed much lower surface activity and

emulsifying stability than both the protein-depleted fraction (0.8 %) and crude gum

(3.0 %), indicating the protein did not play any significant role in their stabilizing action

(Garti et al., 1994). The adsorption mechanism, however, for either protein or the gums

with trace protein left are still not clear.

3.3.4.1 Surface activity

The protein fraction in both SFG and AFG were completely removed via protease

hydrolysis. Two resultant fractions were denoted as SFGnP and AFGnP, respectively. The

surface tension of dispersions decreased slightly with the addition of SFGnP, AFGnP and

NFG (0.01~ 0.5 %, w/v) (Fig. 3.5). SFG and AFG reduced the surface tension to

0.055±0.001and 0.056±0.00 N/m 1, respectively at a concentration of 0.5 %, which is

comparable with the value of the solution containing 0.5 % of guar gum (Huang et al.,

2001). After protein removal, SFGnP (0.062±0.001 N/m) and AFGnP (0.065±0.002 N/m)

reduced the surface tension to the same level as NFG (0.063±0.003 N/m) did at an

equivalent concentration of 0.5 %, implying that the protein fractions in both SFG and

AFG were responsible for their higher surface activity.

Page 47: Structure-Function Relationship of Flaxseed Gum from

36

Fig. 3.5. Reduction of surface tension of dispersions by soluble flaxseed gum and its

fractions

(SFG and SFGnP: soluble flaxseed gum with and without protein; NFG: neutral fraction gum; AFG and

AFGnP: acidic fraction gum with and without protein; surface tension of distilled water: 0.072±0.001 N/m.

Surface tension of each fraction at each concentration was measured for six times and presented as the

mean value with error bars showing the standard deviations.)

3.3.4.2 Emulsifying stability

The volume–length diameter, D[4,3], is sensitive to the presence of large particles

(i.e., sensitive to flocculation and coalescence), thus is commonly used in expressing the

mean particle size of a polydisperse emulsion to indicate the stability of an emulsion. It is

the sum of the volume ratio of droplets in each size-class multiplied by the mid-point

diameter of the size-class (McClements, 2005):

∑∑∑===

==11

3

1

4]3,4[i

iii

iii

ii ddndnD φ (3.9)

where D[4,3] is the volume fraction-length mean diamerter, ni, di and φ represent the

number, mid-point diameter, and the volume fraction of each size-class, respectively.

The top five factors affecting the rate of creaming, thus the stability of an emulsion,

0.075

0.070

0.065

0.060

0.055

0.050

Surfa

ce T

ensi

on

(N/m

)

Page 48: Structure-Function Relationship of Flaxseed Gum from

37

are rheology of the continuous phase, droplet volume fractions, the density difference

between phases, droplet size and its distribution (Hill, 1998). In a preliminary experiment,

an emulsion containing 10 % wt oil and 1.0 % wt SFG showed larger D[4,3] (2.4 µm) but

was stable for more than 30 days, whereas in another emulsion containing 5 % wt oil and

0.5 % wt of SFG, though smaller D[4,3] (2.0 µm) was found, phase separation occurred

within 4 days after emulsification, indicating the higher viscosity by addition of 1% SFG

retarded the rate of flocculation and coalescence. To avoid the viscosity effect, 0.50, 0.48

and 0.44 % wt of SFG, AFG and protein-free gum fractions (SFGnP, AFGnP and NFG)

(containing identical amount (0.44 % wt) of pure gum for each fraction, which is below

their critical concentrations (6.4-6.7 g L-1)) were included to prepare dilute gum solutions.

Before mixing with 2 % wt of canola oil, the zero-shear viscosity (mPa.s) of the

continuous phase of emulsions with SFG (7.4±0.1) and SFGnP (7.5±0.1) were two-fold

of that of NFG (3.7±0.1), AFG (3.6±0.1) and AFGnP (3.5±0.1). The shifts of droplet size

distributions and D[4,3] values of emulsions with the above five fractions are shown in

Fig. 3.6. Monodispersed droplet size distribution and stable D[4,3] values were found in

emulsions containing SFG and AFG from 3 to 24 h after emulsification, whereas the

volume fraction (%) of the second peak (larger size particles), as well as the D[4,3]

values, increased obviously within 24 h for all three protein-free fractions (SFGnP, NFG

and AFGnP), which again confirmed that the protein fractions in SFG and AFG

contributed to their better emulsifying stabilities than the other three protein-free fractions,

in spite of their distinctions in viscosity.

Page 49: Structure-Function Relationship of Flaxseed Gum from

38

Fig. 3.6. The shifts of mean diameter (D[4,3]) and particle size distribution profiles of

emulsions stabilized by soluble flaxseed gum and its fractions within 24h after

emulsification with 2 % wt of canola oil

(SFG and SFGnP: soluble flaxseed gum with and without protein; NFG: neutral fraction gum; AFG and

AFGnP: acidic fraction gum with and without protein. Each mean diameter was measured in triplicate and

presented as the mean value with error bars showing the standard deviations.)

Page 50: Structure-Function Relationship of Flaxseed Gum from

39

3.4. Conclusions

Flaxseed gum may become a significant source of soluble fibre for both its

availability and low viscosity. Two fractions, neutral (NFG) and acidic (AFG), of soluble

flaxseed gum (SFG) from hulls were effectively separated using ion exchange

chromatography. NFG, consisting of only one polymer with high MW, was free of protein

and contained <2 % of uronic acid. AFG mainly consisted of polymers with different

MW associated with 8 % of protein which was not covalently linked. Both fractions were

stable at 80°C for 1 h, however, extended heating resulted in degradations of the

polymers. The reduced surface activity and emulsifying stability after complete removal

of protein content from SFG (11.8 %) and AFG (8.1 %) using protease revealed the

contribution of the protein fractions to their emulsification, regardless of their distinctions

in molecular conformation (molecular mass and chain flexibility) and rheological

properties. However, the differences between these two fractions in MW distribution and

chain stiffness affected their intrinsic viscosity and viscosity dependence on the

concentration in both dilute and semi-dilute solutions.

Page 51: Structure-Function Relationship of Flaxseed Gum from

40

Chapter 4. Structural Elucidation of the Acidic Fraction Gum

4.1 Introduction

In our previous study, soluble flaxseed gum extracted from flaxseed hulls was

separated into a neutral and an acidic fraction using anion exchange chromatography

(Qian et al., 2012b). The acidic fraction gum (AFG) might be favored for its low viscous

properties as a potential dietary fibre fortifier to be included into food systems in a large

amount without over-texturization. Methylation analysis in previous studies showed the

main linkages of the AFG to include →4) -α-D-GalpA-(1→, and linear or branched

→2)-α-L-Rhap-(1→ at O-3 position (Cui et al., 1994a; Naran et al., 2008).

Rhamnogalacturonan-I (RG-I) backbone features its repeating units of alternatively

distributed →2)-α-L-Rhap-(1→ and →4)-α-D-GalpA-(1→. RG-I, RG II and

homogalacturonan (HG) comprise the three main structural elements of pectin. In general,

pectins encompass hetero-polysaccharides containing at least 65% of galacturonic

acid-based units according to the FAO and EU stipulation(Willats et al., 2006). Arabinose

or galactose is the second most abundant neutral sugar component in pectins with no

more than 4% of rhamnose and fucose typically (Stephen, 1995). Although AFG was

referred to as pectic polysaccharides, RG-I (Naran et al., 2008), this fraction differed

from pectins in its higher rhamnose content and lower galacturonic acid content.

Interest in polysaccharides as potential sources for anti-tumor agents increased in

recent decades due to their fewer side effects. Modified pectin was reported to show (in

Page 52: Structure-Function Relationship of Flaxseed Gum from

41

vitro) anti-tumor bioactivities probably due to the existence of galactans from side chains

in the hairy region (Glinsky et al., 2009; Morris et al., 2011). Revealing the fine structure

of polysaccharides assists in further tracing their potential functionalities and/or

bioactivities. The detailed structure of AFG has not been reported. This study focused on

the structure of AFG using methylation-GC-MS analysis and 1D/2D NMR spectroscopy

including homonuclear 1H/1H correlation spectroscopy (COSY, TOCSY) and nuclear

overhauser effect spectroscopy (NOESY), heteronuclear 1H/13C multiple-quantum

coherence spectroscopy (HMQC) and heteronuclear 1H/13C multiple bond correlation

spectroscopy (HMBC).

4.2 Experimental

4.2.1 Materials

Soluble flaxseed gum was isolated by aqueous extraction from flaxseed hulls

(variety Bethune) supplied by Natunola Health Biosciences (Winchester, Ontario,

Canada). The acidic fraction gum (AFG) was separated from soluble flaxseed gum using

anion-exchange chromatography and purified by protease hydrolysis to remove

proteinaceous contaminant (Qian et al., 2012b). All chemicals were of reagent grade

unless otherwise specified.

Page 53: Structure-Function Relationship of Flaxseed Gum from

42

4.2.3 Methylation analysis

The acidic fraction gum (AFG) was reduced into neutral polysaccharides before

methylation. The procedure of reduction and methylation followed that of previous work

(Kang et al., 2011). The resulting partially methylated alditol acetates (PMAAs) were

injected into a GC-MS system (ThermoQuest Finnigan, San Diego, CA) with an SP-2330

(Supelco, Bellefonte, Pa) column (30 m×0.25 mm, 0.2 mm film thickness, 160-210 oC at

2 oC /min, and then 210-240 oC at 5 oC /min) equipped with an ion trap MS detector.

4.2.3 NMR analysis

Sample was dissolved in deuterium oxide (D2O, 80 oC, 1 h) and lyophilized for three

times to replace the exchangeable protons with deuterons before being finally redissolved

in D2O (3 %) for NMR analysis. High-resolution 1H and 13C NMR spectra were recorded

in D2O at 500.13 and 125.78 MHz, respectively, on a Bruker ARX500 NMR

spectrometer operating at 25 oC. A 5 mm inverse geometry 1H/13C/15N probe was used.

Chemical shifts are reported relative to external standards Trimethylsilyl propionate (TSP

in D2O, 4.76 ppm, for 1H) and 1,4-Dioxane (in D2O, 66.5 ppm, for 13C). Homonuclear

1H/1H correlation spectroscopy (COSY, TOCSY) and nuclear overhauser effect

spectroscopy (NOESY), heteronuclear 1H /13C multiple-quantum coherence spectroscopy

(HMQC) experiments, and heteronuclear multiple bond correlation spectroscopy (HMBC)

were run using the standard Bruker pulse sequence at 70 oC.

Page 54: Structure-Function Relationship of Flaxseed Gum from

43

4.3 Results and discussion

4.3.1 Methylation and GC-MS of partially methylated alditol acetate (PMAA) of the

acidic fraction gum

Methylation-GC-MS analysis of the acidic fraction gum (AFG) showed that this

fraction was composed of four hexosyl residues (rhamnose (38.1 %), galactouronic acid

(24.8 %), galactose (20.8 %), and fucose (4.5 %)) and two pentosyl residues (xylose

(7.7 %) and arabinose (2.6 %)) as listed in Table 4.1. Four major linkage patterns

constituted 78.5 % of the total sugar residues include: →2,3) -L-Rhap-(1→ (19.5 %),

→2)-L-Rhap-(1→ (16.5 %), →4)-D-GalpA-(1→ (22.9 %) and D-Galp-(1→ (19.6 %).

The molar ratio of total non-reducing terminal residues and total branching points were

29.2 and 25.8 %, respectively.

The degree of branching (DB) of AFG was equal to 0.55 as calculated according to

the equation below (Hawker et al., 1991; Tao et al., 2007):

DB=(NT+NB)/(NT+NB+NL)=(29.2+24.5)/(29.2+24.5+44.8)=0.55 (4.1)

where NT, NB and NL are the molar percentage of the terminal, branched and linear

residues, respectively. The DB value of a linear chain equals to 0, whereas that of a fully

branched polymer is 1. The DB value (0.55) of AFG indicated it was highly branched.

Page 55: Structure-Function Relationship of Flaxseed Gum from

44

Table 4.1. Partially Methylated Aditol Acetate (PMAA) sugar residue derivatives of

the acidic fraction gum

Deduced Residue RT* Mol%** PMAA

R3 →2,3)-L-Rhap-(1→ 1.307 19.5 4-Me Rhap

R →2)-L-Rhap-(1→ 0.957 16.5 3,4-Me2 Rhap

RT L-Rhap-(1→ 0.603 1.2 2,3,4-Me3 Rhap

R34/4 →2,3,4)/→2,4)-L-Rhap-(1→ 1.379 0.9 Rhap-(OAC)5 /3-Me Rhap

Rhamnose 38.1

GA →4)-D-GalpA-(1→ 1.543 22.9 2,3,6-Me3 GalpA

GA2 →2,4)-D-GalpA-(1→ 1.900 1.1 3,6-Me2 GalpA

GA3 →3,4)-D-GalpA-(1→ 1.764 0.8 2,6-Me2 GalpA

Galacturonic Acid 24.8

GT D-Galp-(1→ 1.117 19.6 2,3,4,6-Me4 Galp

G' →6)-D-Galp-(1→ 1.703 1.2 2,3,4-Me3 Galp

Galactose 20.8

FT L-Fucp-(1→ 0.754 4.5 2,3,4-Me3 Fucp

Fucose 4.5

XT D-Xylp-(1→ 0.784 3.1 2,3,4-Me3 Xylp

X →4)-D-Xylp-(1→ 1.250 2.4 2,3,-Me2 Xylp

X34 →2,3,4)-D-Xylp-(1→ 2.054 1.3 Xylp-(OAC)5

X2/3 →2,4)/3,4)-D-Xylp-(1→ 1.646 0.9 3/2-Me Xylp

Xylose 7.7

A →3)-L-Araf-(1→ 1.054 1.8 2,5-Me2 Araf

AT L-Araf-(1→ 0.642 0.8 2,3,5-Me3 Araf

Arabinose 2.6

(RT*: Retention time is relative to 2,3,4,6-Me4 Glc (14.591 min); Mol%**: molar ratio of each sugar residue

is based on the percentage of its peak area; Letter labeled by a superscript T refers to a terminal residue;

The superscript numbers follow the letters indicate the branching sites in these residues; Data for sugar

residues less than 0.6 % are not shown; Linkages below 2% were not included in the proposed repeating

unit for AFG in Fig. 4.7 )

Page 56: Structure-Function Relationship of Flaxseed Gum from

45

4.3.2 1D and 2D NMR analysis of the acidic fraction gum

More than six peaks are shown in the anomeric region (4.3~5.2 ppm) of the 1H NMR

spectrum (Fig. 4.1a). Two peaks at 1.08 and 1.11 ppm arose from the proton resonances

of methyl groups in fucose and rhamnose, respectively; their corresponding resonances in

the 13C spectrum (Figure 4.1b) are at 16.2 and 17.6 ppm, respectively.

Fig. 4.1. 1H (a) and 13

(R3: →2,3)-α-L-Rhap-(1→; R: →2)-α-L-Rhap -(1→; GA: →4)-α-D-GalpA -(1→; βGT:β-D-Galp-(1→; αGT:

α-D-Galp-(1→; FT: α-L-Fucp-(1→; XT: β-D-Xylp-(1→; X: →4)-β-D-Xylp-(1→.)

C (b) NMR spectra of the acidic fraction gum (ppm)

Page 57: Structure-Function Relationship of Flaxseed Gum from

46

Fig. 4.2. Key fragments of COSY spectrum of the acidic fraction gum

(Intra-ring COSY correlations are labeled by residue abbreviations and two unseparareted numbers of correlating

protons; R3: →2,3)-α-L-Rhap-(1→; R: →2)-α-L-Rhap -(1→; GA: →4)-α-D-GalpA -(1→; α/βGT: α/β-D-Galp-(1→; FT:

α-L-Fucp-(1→; XT: β-D-Xylp-(1→; X: →4)-β-D-Xylp-(1→; A: →3)-α-Araf-(1→.)

Page 58: Structure-Function Relationship of Flaxseed Gum from

47

Fig. 4.3. Key fragments of TOCSY spectrum of the acidic fraction gum

(Intra-ring TOCSY correlations are labeled by residue abbreviations and two unseparated numbers of correlating

protons; R3: →2,3)-α-L-Rhap-(1→; R: →2)-α-L-Rhap -(1→; GA: →4)-α-D-GalpA -(1→; α/βGT: α/β-D-Galp-(1→; FT:

α-L-Fucp-(1→.)

Page 59: Structure-Function Relationship of Flaxseed Gum from

48

Fig. 4.4. Key fragments of HMQC spectrum of the acidic fraction gum

(Intra-ring H/C correlations are labeled by residue abbreviation and the number of correlating proton/carbon; R3:

→2,3)-α-L-Rhap-(1→; R: →2)-α-L-Rhap -(1→; GA: →4)-α-D-GalpA -(1→; α/βGT: α/β-D-Galp-(1→; FT:

α-L-Fucp-(1→.)

Page 60: Structure-Function Relationship of Flaxseed Gum from

49

Fig. 4.5. Key fragment of HMBC spectrum of the acidic fraction gum

(Inter-ring H/C correlations are labeled by correlating proton or carbon and its number followed by the residue

abbreviation; R*: R & R3; R3: →2,3)-α-L-Rhap-(1→; R: →2)-α-L-Rhap -(1→; GA: →4)-α-D-GalpA -(1→; α/βGT:

α/β-D-Galp-(1→; FT: α-L-Fucp-(1→; XT: β-D-Xylp-(1→.)

Page 61: Structure-Function Relationship of Flaxseed Gum from

50

Fig. 4.6. Key fragment of NOESY spectrum of the acidic fraction gum

(Inter-ring H/H correlations are labeled by the residue abbreviation/its proton number; R*: R & R3; R3:

→2,3)-α-L-Rhap-(1→; R: →2)-α-L-Rhap -(1→; GA: →4)-α-D-GalpA -(1→; α/βGT: α/β-D-Galp-(1→; FT:

α-L-Fucp-(1→; XT: β-D-Xylp-(1→; X: →4)-β-D-Xylp-(1→.)

Page 62: Structure-Function Relationship of Flaxseed Gum from

51

The spin system for each sugar residue of AFG was assigned according to the COSY

spectrum (Fig. 4.2), with assistant/confirmative information from the TOCSY (Fig. 4.3)

and HMQC spectra (Fig. 4.4). The sequence of the linkages of sugar residues were

inferred from the HMBC (Fig. 4.5) and NOESY spectra (Fig 4.6). The assignment of

each sugar residue and their relative sequences are discussed in the following sections

before a possible structure of the repeating unit of AFG is proposed.

4.3.2.1 Rhamnose residues

Rhamnosyl residues (38.2 %) in AFG include four linkages as listed in decreasing

order of predominance in Table 4.1: →2,3) -α-L-Rhap-(1→ (R3, 19.5 %),

→2)-α-L-Rhap-(1→ (R, 16.5 %), and minor amount of T-α-Rhap-(1→ (RT, 1.2 %) and

→2,3,4)/→2,4)-α-L-Rhap-(1→ (R34/4, 0.9 %). The α-configuration of all Rhap units was

inferred from their common chemical shifts of anomeric carbon (99.6 ppm) and proton

(5.11 ppm) as shown in the anomeric region of HMQC spectrum (Fig. 4.4a).

The spin units for residue →2) -α-L-Rhap-(1→ (R) and →2,3)-α-L-Rhap-(1→ (R3)

are fully assigned in COSY, TOCSY and HMQC spectra (Fig. 4.2~4.4 & Table 4.2). The

H-2 (4.05 ppm), H-3 (3.83/3.85 ppm) and H-4 (3.50 ppm) chemical shifts of the R3

moved to downfield by 0.1~0.2 ppm, due to the glycosylation at O-3 position (Colquhoun

et al., 1990; Pawan, 1992). The substitution at O-3 positions of rhamnosyl residues

instead of the typical O-4 positions, as found in pectin and soluble soybean

polysaccharides, was also reported based on methylation analysis of flaxseed mucilage in

earlier studies (Muralikrishna et al., 1987; Naran et al., 2008). Both cross-peaks at

Page 63: Structure-Function Relationship of Flaxseed Gum from

52

3.83/72.2 and 3.85/77.9 ppm in HMQC (Fig. 4.4) were tentatively assigned to H-3/C-3

correlation of R3. This splitting may be due to the substitution by different sugar units,

according to HMBC (Fig. 4.5) and NOESY (Fig. 4.6) spectra. The chemical shifts of

glycosylated carbon (C-3) moved downfield by 1.7 and 7.5 ppm (Table 4.2), respectively.

The former value was far smaller than the value of 7~10 ppm reported for rhamosyl

residues branched at O-3 position(Choma et al., 2009; Fedonenko et al., 2008; MacLean

et al., 2010; Ojha et al., 2008); other than this, the provisional assignment was in good

agreement with literature data(Cui et al., 1996a; Mikshina et al., 2012; Ojha et al., 2008;

Sengkhamparn et al., 2009).

4.3.2.2 Galactouronic acid residue

The presence of three peaks at 173.6-174.5 ppm from carboxyl group in 13C 1D NMR

spectrum (Fig. 4.1b) is the evidence for the existence of galactouronic acid (GA, 24.8 %).

Linear →4) -D-GalpA-(1→ (GA, 22.9 %) accounted for 92 % of this residue, with the

rest singly substituted at O-2 or O-3 position (Table 4.1). A full assignment for

→4)-α-D-GalpA-(1→ (Table 4.2) was in good agreement with previous reports

(Fedonenko et al., 2008; MacLean et al., 2010; Mikshina et al., 2012; Ojha et al., 2008).

Page 64: Structure-Function Relationship of Flaxseed Gum from

53

Table 4.2. 1H/13C NMR chemical shifts for the acidic fraction gum in D2O at 70o

Deduced Residues

C

(ppm)

H-1/C-1 H-2/C-2 H-3/C-3 H-4/C-4 H-5/C-5 H-6/C-6

R3 →2,3)-α-L-Rhap-(1→ 5.11/99.6 4.05/78.0 3.83/72.2 3.50/74.1 3.74/70.5 1.16/17.5

3.85/77.9

R →2)-α-L-Rhap-(1→ 5.12/99.6 3.97/77.5 3.73/70.5 3.30/73.1 3.62/70.5 1.13/17.5

GA →4)-α-D-GalpA-(1→ 4.89/99.0 3.76/69.0 3.97/71.1 4.32/78.3 4.63/72.1 /173.6-174.5

αGt α-D-Galp-(1→ 5.12/102.1 3.67/69.5 3.69/69.6 3.87/70.5 3.77/70.5 3.59/62.5

βGt β-D-Galp-(1→ 4.40/97.4 3.32/- 3.48/- 3.80/- 3.59/- -/-

4.44/97.2 3.38/- 3.48/- 3.80/- 3.59/- -/-

FT α-Fucp-(1→ 5.02/102.3 3.64/- 3.60/- -/- 3.90/- 1.08/16.2

XT β-D-Xylp-(1→ 4.44/97.2 3.20/- -/- -/- -/- -/-

X →4)-β-D-Xylp-(1→ 4.31/97.4 3.20/- -/- -/- -/- -/-

A →3)-Araf-(→ 5.12/- 4.17/- -/- -/- -/- -/-

Page 65: Structure-Function Relationship of Flaxseed Gum from

54

Table 4.3 Heteronuclear (HMBC) and homonuclear (TOCSY & NOESY)

connectivities in the acidic fraction gum (ppm)

Inter-residue connectivities Intra-residue connectivities

HMBC NOESY HMBC

Atom δ Atom δ Atom δ Atom δ Atom δ Atom δ

H-1 R* 5.12 C-4 GA 78.3 H-1 GA 4.90 H-2 R 3.97 H-1 R* 5.12 C-2 78.0

C-2 R* 77.5-78.0 H-2 R3 4.05 H-1 R 5.12 C-3 70.9

C-1 R* 99.7 H-2 R 3.99 H-1 FT 5.03 H-2 R 3.97 H-1 R3 5.12 C-3 72.2

99.9 H-4 GA 4.32 H-1 R* 5.12 H-4 GA 4.32 H-4 GA 4.32 C-2 69.4

C-1 GA 99.0 GA H-4 4.32 H-1 GA 4.90 H-4 GA C-3 71.3

99.0 H-2 R 3.97 H-1 αGT 5.12 C-2/3 69.8

(Appendants linked to C-3 R3 ) (Appendants linked to C-3 R3 ) TOCSY

Atom δ Atom

H-1 αGT 5.12 C-3 R3 72.1/77.9 H-1 αGT 5.12 H-3 R3 3.83 H-1 αGT 5.12 H-2,3,4,5,6

H-1 βGT 4.44 C-3 R3 72.2 H-1 βGT 4.46 H-3 R3 3.85 H-1 βGT 4.44 H-3,4

H-1 XT 4.44 C-3 R3 72.2 H-1 XT 4.46 H-3 R3 3.85 H-1 R* 5.12 H-2,3,4,5

H-1 R 5.12 C-3 R3 72.1 H-1 R 5.12 H-3 R3 3.85 H-6 R* 1.11 H-1,2,3,4,5

H-1 GA 4.90 C-3 R3 71.7 H-1 GA 4.90 H-3 R3 3.85 H-1 GA 4.89 H-2,3,4

H-1 FT 5.03 C-3 R3 70.8 H-1 FT 5.03 H-3 R3 3.83 H-6 FT 1.08 H-5,6

(R*: R and R3; R3: →2,3)-α-L-Rhap-(1→; R: →2)-α-L-Rhap -(1→; GA: →4)-α-D-GalpA -(1→;

βGT:β-D-Galp-(1→; αGT: α-D-Galp-(1→; FT: α-L-Fucp-(1→; XT: β-D-Xylp-(1→.)

4.3.2.3 Galactose residues

The second most abundant neutral sugar residue is Galactose (20.8 %, Table 4.1),

94 % of which is (GT, 19.6 %). The cross-peak at 5.12/102.1 ppm in HMQC (Fig. 4.4a)

Page 66: Structure-Function Relationship of Flaxseed Gum from

55

arose from H-1/C-1 correlation of α-D-Galp-(1→ (αGT), and its full assignment

completed was in accordance with literature data (Ahrazem et al., 2002; Galbraith et al.,

1999; Sengkhamparn et al., 2009). Two cross-peaks at 4.40/97.4 and 4.44/97.2 ppm in

HMQC (Fig. 4.4a) may arise from H-1/C-1 connectivity of β-D-Galp-(1→ (βGT)(Deng et

al., 2006; Sengkhamparn et al., 2009; Vidal et al., 2000) though both could be only

partially assigned according to COSY and TOCSY spectra (Fig. 4.2 & 4.3) due to the low

intensity in the rest of spectra. The α-configuration was predominant, however, the ratio

of both configurations could not be estimated due to the overplayed anomeric proton

chemical shifts of α-D-Galp-(1→ and Rhap (5.12 ppm), or of β-D-Galp-(1→ and

β-D-Xylp-(1→(4.44 ppm) as shown in Fig. 4.1.

4.3.2.4 Fucose residue

Terminal α-L-Fucp-(1→ (FT, 4.5 %) was the only linkage for Fucosyl residue.

Correlations of H-1/H-2 and H-5/H-6 (Fig. 4.2 & 4.3) could be assigned according to

literature data (Shashkov et al., 1993).

4.3.2.5 Minor residues

Xylose (7.7 %) was constituted of 3.1 % β-D-Xylp-(1→ (XT) and 2.4 % of

→4)-β-D-Xylp-(1→ (X). The rest were →4)-β-D-Xylp-(1→ branched at O-2, O-3 or

both. Two group of cross-peaks at 4.44-4.46/3.20 and 4.31-4.33/3.20 ppm in COSY (Fig.

4.2) may arise from XT and X, respectively (Höije et al., 2006; Pastell et al., 2008; Sun et

Page 67: Structure-Function Relationship of Flaxseed Gum from

56

al., 2011). However, full assignment was impossible due to the low intensity of signals.

A small portion (2.6 %) of Arabinofuranose existed in two linkages:

→3)-α-L-Araf-(1→ (A, 1.8 %) or α-L-Araf-(1→ (AT, 0.8 %). The chemical shift for

H-1/H-2 correlation of α-L-Araf-(1→ was reported to be close to that of rhamnose (Höije

et al., 2006; Pastell et al., 2008; Sun et al., 2011), thus could not be assigned. The

cross-peak at 5.12/4.17 ppm in COSY (Fig. 4.2) may arise from H-1/H-2 correlation of

→3)-α-L-Araf-(1→ as the H-2 chemical shift moved to downfield by 0.1-0.2 ppm due to

the glycosylation at O-3(Westphal et al., 2010).

Two cross-peaks at 4.79/3.42 and 4.98/3.42 may be from the H-1/H-2 correlation of

→4)-α-D-Glcp-(1→ and →6)-α-D-Glcp-(1→, respectively. However, the glucose content

in this acidic fraction was too low to be detected as indicated in either methylation (Table

4.1) or monosaccharide analysis (Qian et al., 2012b).

4.3.2.6 Linkage sequence of AFG

Heteronuclear multiple bond correlation (HMBC) spectrum shows correlation

between protons and carbons 2-3 bonds away. Nuclear overhauser effect spectroscopy

(NOESY) correlates nuclei through space (distance smaller than 5Å). Both methods are

helpful to reveal the glycosylic linkages between sugar residues although intra-residue

connectivities also included. The inter-residue and intra-residue correlations observed are

listed in Table 4.3.

The most intense inter-residue connectivities assigned in Fig 4.5 were C-1R*/H-4GA

Page 68: Structure-Function Relationship of Flaxseed Gum from

57

(99.9/4.32 ppm, R*: R or R3) and H-1R*/C-4GA (5.12/78.3 ppm), indicating the

abundance of diglycosyl repeating unit →2)-α-L-Rhap-(1→4)-α-D-GalpA-(1→. Links

between →4) -α-D-GalpA-(1→ itself, i.e. homogalacturonan (HG) regions, was evident

by H-1/H-4 correlation of →4) -α-D-GalpA-(1→ (4.89/4.32 ppm, Fig. 4.6), though its

evidence in HMBC (C-1/H-4GA, 99.0/4.32 ppm, Fig. 4.5) overplayed with

C-1R*/H-4GA, (99.9/4.32 ppm). However, the amount of HG should be limited due to

the low GalpA content and its predominant occurrence in repeating unit

→2)-α-L-Rhap-(1→4)-α-D-GalpA-(1→. Strong correlations of H-1/C-2R*

(5.12/77.5-78.0 ppm) and C-1/H-2R (99.7/3.99 ppm) observed in Fig. 4.5 may derive

from both intra- and inter-residue connectivites between rhamnosyl residues. However,

the existence of homorhamnan (HR) region could still be confirmed due to the abundance

of rhamnosyl residues. Given the above evidence, a backbone consisting of RG-I possibly

intervened by small amount of HG and HR could be drawn.

The branching site of R3 (19.5 %) at O-3 were mostly substituted by monosaccharides,

e.g., α/β-D-Galp-(1→ (α/βGT, 19.6 %), α-L-Fucp-(1→ (FT, 4.5 %) and β-D-Xylp-(1→ (XT

,

3.1 %) (Table 4.1). The C-3 chemical shifts of R3 varied with substitution by different

sugar units (Table 4.3, Fig. 4.5 & 4.6). However, when it was substituted with

→2)-α-L-Rhap-(1→ or →4)-α-D-GalpA-(1→, a longer side chain consisting of more

than two residues may induced.

Page 69: Structure-Function Relationship of Flaxseed Gum from

58

Fig. 4.7. Schematic representations of the conventional (a) and recently proposed

alternative (b) structures of pectin

(RG-I: rhamnogalacturonan-I; RG-II: rhamnogalacturonan-II; HG: homogalacturonan;

schematics are adapted from Willats & Knox et al. 2006.)

[→2)-α-L-Rhap-(1→]m[→2)-α-L-Rhap-(1→4)-α-D-GalpA-(1→]n[→4)-α-D-GalpA-(1→]i

Fig. 4.8. Proposed repeating unit of the acidic fraction gum

(HR, RG-I and HG refer to homorhamnan, rhamnogalacturonan-I and homogalacturonan, respectively.

The locations of HR, RG-I and HG are interchangeable; (m+n)/(n+i)≈1.5. The substitution rate of R1 is

~54 %. R1 is mostly monosaccharide (α/β-D-Galp-(1→, α-L-Fucp-(1→ or β-D-Xylp-(1→). R1 may also

occasionally be a longer side chain with more than two residues beginning with →4) -α-GalpA-(1→ or

→2)-α-L-Rhap-(1→, wherein the side-chain structure may be similar to part of the main chain.)

3

R1

3

R1

HR RG-I HG

HG

(a)

(b)

RG-II

b

a

RG-I

Page 70: Structure-Function Relationship of Flaxseed Gum from

59

The fine structure of pectin is still in dispute due to its complexity and heterogeneity

between plants and tissues (Willats et al., 2006). The controversy between conventional

and recently proposed alternative structural models of pectin is whether the HG domain is

inserted into its backbone or linked as a side chain (Fig. 4.7). A similar question also

exists in revealing the structure of AFG. A combination of both models is tentatively

adopted. Given the above evidence, a possible repeating unit of AFG (Fig. 4.8) is

proposed.

4.4 Conclusions

The acidic fraction gum (AFG) from flaxseed hulls was a pectic polysaccharide as

investigated by methylation analysis and 2D NMR spectroscopy. AFG contained a

backbone consisting of rhamnogalacturonan-I which might be intervened by small

amount of homorhamnan or homogalacturonan. However, AFG differed from pectin in a

much higher rhamnose (38.2 %) and much lower galacturonic acid content (25.4 %), thus

had limited amount of HG region. This fraction also featured its high amount of

mono-galactosyl branches 3-linked to half of 1,2-linked rhamnosyl residues. However

longer side chains with more than two residues also existed when substitution with

→4)-D-GalAp-(1→ or →2)-α-L-Rhap-(1→ was 3-linked to 1,2-linked rhamnosyl

residues.

Page 71: Structure-Function Relationship of Flaxseed Gum from

60

Chapter 5. Structure Elucidation of the Neutral Fraction Gum

5.1 Introduction

In chapter 3, soluble flaxseed gum extracted from flaxseed hulls was separated into a

neutral and an acidic fraction using anion-exchange chromatography. The neutral fraction

gum (NFG) was referred to as arabinoxylans for its high xylose (70 %) and arabinose

(20 %) content (Qian et al., 2012b). Earlier results of methylation analysis showed the

main linkages of the NFG included T-α-D-Xylp, 1,4-linked α-D-Xylp unsubstituted or

substituted at O-2 or O-3 or both positions and 1,3-linked or 1,5-linked α-L-Araf (Cui et

al., 1994a; Naran et al., 2008). Arabinoxylan is one of the most abundant components in

plant cell wall polysaccharide, along with cellulose and pectins. The structural features of

arabinoxylans including the ratio of arabinose to xylose residues, the relative amount and

sequence of mono-, di- and unsubstituted Xylp, and the possible presence of other

substituents (Izydorczyk et al, 2008). The structural diversity and complexity are likely

related to their functionality and the source of material, the genotype and cellular origin,

as well as the growth stage of the plants (Izydorczyk, 2009).

Revealing the fine structure of arabinoxylans provides the basis for further tracing

their potential functional and/or physiological properties. Detailed structure of NFG has

not been reported. This study focused on the structure of NFG using methylation-GC-MS

analysis and 1D/2D NMR spectroscopy including homonuclear 1H/1H correlations

spectroscopy (COSY, TOCSY), heteronuclear 13C/1H multiple-quantum coherence

Page 72: Structure-Function Relationship of Flaxseed Gum from

61

spectroscopy (HMQC), heteronuclear multiple bond correlation spectroscopy (HMBC)

and nuclear overhauser effect spectroscopy (NOESY).

5.2 Experimental

5.2.1 Materials

Soluble flaxseed gum was isolated by aqueous extraction from flaxseed hulls

(variety Bethune) supplied by Natunola Health Biosciences (Winchester, Ontario,

Canada). The protein-free neutral fraction gum (NFG) was separated from soluble

flaxseed gum using ion exchange chromatography (Qian et al., 2012b). All chemicals

were of reagent grade unless otherwise specified.

5.2.2 Methylation analysis

The methylation procedure of neutral fraction (NFG) followed that of previous work

(Kang, 2011). The resultant partially methylated alditol acetates (PMAAs) were injected

into a GC-MS system (ThermoQuest Finnigan, San Diego, CA) with an SP-2330

(Supelco, Bellefonte, Pa) column (30 m×0.25 mm, 0.2 mm film thickness, 160-210 oC at

2 oC/min, and then 210-240 oC at 5 oC /min) equipped with an ion trap MS detector.

5.2.3 NMR analysis

Sample was dissolved in deuterium oxide (D2O, 3 %, 80 oC, 1 h) and lyophilized for

three times to replace the exchangeable protons with deuterons before being finally

redissolved in D2O (2 %) for NMR analysis. High-resolution 1H and 13C NMR spectra

were recorded in D2O at 500.13 and 125.78 MHz, respectively, on a Bruker ARX500

Page 73: Structure-Function Relationship of Flaxseed Gum from

62

NMR spectrometer operating at 70 oC. A 5 mm inverse geometry 1H/13C/15N probe was

used. Chemical shifts are reported relative to external standards Trimethylsilyl propionate

(TSP in D2O, 4.76 ppm, for 1H) and 1,4-Dioxane (in D2O, 66.5 ppm, for 13C).

Homonuclear 1H/1H correlation spectroscopy (COSY, TOCSY and nuclear overhauser

effect spectroscopy (NOESY)), heteronuclear 1H /13C multiple-quantum coherence

spectroscopy (HMQC) and heteronuclear multiple bond correlation spectroscopy (HMBC)

experiments were conducted using the standard Bruker pulse sequence. To increase the

signal-to-noise ratio(s/n), all experiments were run at 70 oC.

5.3 Results and discussion

5.3.1 Methylation and GC-MS of partially methylated alditol acetate (PMAA) of the

neutral fraction gum

Methylation-GC-MS analysis of the neutral fraction gum (NFG) showed this fraction

was composed of 65.1 % of Xylp and 24.8 % of Araf,and the rest (9.5 %) originated

from Galp or Glcp (Table 5.1). Four major linkage patterns constituted 77.4 % of the total

sugar residues and they are: linear →4) -D-Xylp-(1→ (25.0 %) or double substituted at

O-2 and O-3 (19.6 %), T-D-Xylp-(1→ (15.5 %), and →1)-L-Araf-(5→ (17.3 %).

Page 74: Structure-Function Relationship of Flaxseed Gum from

63

Table 5.1. Partially Methylated Aditol Acetate (PMAA) sugar residue derivatives of

the neutral fraction gum

Abbr. Deduced Linkage RT* Mol%** PMAA

X4 →4)-β-D-Xylp-(→ 1.255 25.0 2,3-Me2 Xylp

X234 →2,3,4)-β-D-Xylp-(1→ 2.066 19.6 Xylp-(OAc)5

XT T-β-D-Xylp-(1→ 0.790 15.5 2,3,4-Me3 Xylp

X24/34 →2/3,4)-β-D-Xylp-(1→ 1.648 5.1 3/2-Me Xylp

Xylose 65.1

A5 →5)-α-Araf-(l→ 1.158 17.3 2,3-Me2 -Araf

A3 →3)-α-Araf-(l→ 1.045 4.9 2,5-Me2 Araf

AT T-α-Araf-(l→ 0.646 2.6 2,3,5-Me3 Araf

Arabinose 24.8

Pentose 89.9

GT T-α-D-Galp-(1→ 1.114 4.5 2,3,4,6-Me4 Galp

G4/C4 →4)-α-D-Galp/Glcp -(1→ 1.576 3.5 2,3,6-Me3 Galp/Glcp

C6 →6)-α-D-Glcp-(1→ 1.517 1.4 2,3,4-Me3 Glcp

Hextose 9.4

(RT*: Retention time is relative to 2,3,4,6-Me4 Glc (14.591 min); Mol%**: molar ratio of each sugar residue

is based on the percentage of its peak area; Letter labeled by a superscript T refers to a terminal residue;

The superscript numbers follow the letters indicate the branching sites in these residues; Data for sugar

residues less than 0.3 % are not shown.)

Page 75: Structure-Function Relationship of Flaxseed Gum from

64

Molar ratio of total non-reducing terminal residues and total branching points were

22.6 % and 44.3 %, respectively. The degradation of T-Araf-(1→ may be the main reason

causing the low estimation of terminal residues. The degree of branching (DB) of NFG

was equal to 0.55 as calculated according to the equation below:

DB=(NT+NB)/(NT+NB+NL)=(22.6+24.7)/(22.6+24.7+52.1)=0.48 (5.1)

where NT, NB and NL are the molar percentage of the terminal, branched and linear

residues, respectively. The DB value varies from 0 (a linear chain) to 1 (a fully branched

polymer). The DB value (0.48) of NFG indicated it was highly branched.

5.3.2 1D and 2D NMR analysis of the neutral fraction gum

As shown from the 1H NMR spectrum (Fig. 5.1a), eleven peaks in the anomeric

region (δ 4.3~5.3 ppm) can be divided into three groups of sugar residues: three peaks

from α-Araf (5.17-5.26 ppm), three peaks from α-Glcp or α-Galp (4.77-4.97 ppm) and

five peaks from β-Xylp (4.29-4.55 ppm). Their corresponding 13C peaks were also

assigned (Fig. 5.1b). The spin system for each sugar residue of NFG was assigned

according to the COSY spectrum (Fig. 5.2), with the assistant/confirmative information

from the TOCSY (Fig. 5.3) and HMQC spectra (Fig. 5.4). The linkage sites and sequence

of sugar residues were inferred from HMBC (Fig. 5.5) and NOESY spectra (Fig. 5.6).

The assignment of each sugar residue and their relative sequences are discussed in the

following sections before the repeating unit structure (Fig. 5.7) of NFG is proposed.

Page 76: Structure-Function Relationship of Flaxseed Gum from

65

Fig. 5.1. 1H (a) and 13

C (b) spectra of the neutral fraction gum

(X4: →4)-β-D-Xylp-(1→; X234: →2,3,4)-β-D-Xylp-(1→; XT: T-β-D-Xylp-(1→; X24:

→2,4)-β-D-Xylp-(1→; X34: →3,4)-β-D-Xylp-(1→; A5: →5)-α-Araf-(l→; AT: T-α-Araf-(l→; A3:

→3)-α-Araf-(l→; C4: →4)-D-α-Glcp-(1→; C6: →6)-D-α-Glcp-(1→; GT: T-D-α-Galp-(1→.)

Page 77: Structure-Function Relationship of Flaxseed Gum from

66

Fig. 5.2. COSY spectrum of the neutral fraction gum

(X4: →4)-â-D-Xylp-(1→; X234: →2,3,4)-â-D-Xylp-(1→; XT: T-â-D-Xylp-(1→; X24:

→2,4)-â-D-Xylp-(1→; X34: →3,4)-â-D-Xylp-(1→; A5: →5)-α-Araf-(l→; AT: T-α-Araf-(l→; A3:

→3)-α-Araf-(l→; C4: →4)-D-Glcp-(1→; C6: →6)-D-Glcp-(1→; GT: T-D-Galp-(1→.)

Page 78: Structure-Function Relationship of Flaxseed Gum from

67

Fig. 5.3. TOCSY spectrum of the neutral fraction gum

(X4: →4)-β-D-Xylp-(1→; X234: →2,3,4)-β-D-Xylp-(1→; XT: T-β-D-Xylp-(1→; X24:

→2,4)-β-D-Xylp-(1→; X34: →3,4)-β-D-Xylp-(1→; A5: →5)-α-Araf-(l→; AT: T-α-Araf-(l→; A3:

→3)-α-Araf-(l→; C4: →4)-D-α-Glcp-(1→; C6: →6)-D-α-Glcp-(1→; GT: T-D-α-Galp-(1→.)

Page 79: Structure-Function Relationship of Flaxseed Gum from

68

Fig. 5.4. HMQC spectrum of the neutral fraction gum

(X4: →4)-β-D-Xylp-(1→; X234: →2,3,4)-β-D-Xylp-(1→; XT: T-β-D-Xylp-(1→; X24:

→2,4)-β-D-Xylp-(1→; X34: →3,4)-β-D-Xylp-(1→; A5: →5)-α-Araf-(l→; AT: T-α-Araf-(l→; A3:

→3)-α-Araf-(l→; C4: →4)-D-α-Glcp-(1→; C6: →6)-D-α-Glcp-(1→; GT: T-D-α-Galp-(1→.)

Page 80: Structure-Function Relationship of Flaxseed Gum from

69

Fig. 5.5. HMBC spectrum of the neutral fraction gum

(X4: →4)-β-D-Xylp-(1→; X234: →2,3,4)-β-D-Xylp-(1→; XT: T-β-D-Xylp-(1→; X24:

→2,4)-β-D-Xylp-(1→; X34: →3,4)-β-D-Xylp-(1→; A5: →5)-α-Araf-(l→; AT: T-α-Araf-(l→; A3:

→3)-α-Araf-(l→; C4: →4)-D-α-Glcp-(1→; C6: →6)-D-α-Glcp-(1→; GT: T-D-α-Galp-(1→.)

Page 81: Structure-Function Relationship of Flaxseed Gum from

70

Fig. 5.6. NOESY spectrum of the neutral fraction gum

(X4: →4)-β-D-Xylp-(1→; X234: →2,3,4)-β-D-Xylp-(1→; XT: T-β-D-Xylp-(1→; X24:

→2,4)-β-D-Xylp-(1→; X34: →3,4)-β-D-Xylp-(1→; A5: →5)-α-Araf-(l→; AT: T-α-Araf-(l→; A3:

→3)-α-Araf-(l→; C4: →4)-D-α-Glcp-(1→; C6: →6)-D-α-Glcp-(1→; GT: T-D-α-Galp-(1→.)

Page 82: Structure-Function Relationship of Flaxseed Gum from

71

Fig. 5.7. Proposed repeating unit of the neutral fraction gum

5.3.2.1 Xylose residues

Xylose (65.1 %) is the most abundant component in NFG including five linkage

patterns: linear →4) -β-D-Xylp-(1→ (X, 25.0 %) and substituted at either (X24/34, 5.1 %)

both of O-2 and O-3 (19.6 %), as well as T-β-D-Xylp-(1→ (XT, 15.5 %). The integration

labeled under each anomeric proton peak (Fig 5.1a) also showed the single substituted

→4)-β-D-Xylp-(1→ at O-2 or O-3 was the minority compared with the other three

linkages. However, there is no good correlation between the relative ratio of each linkage

from integration calculation and its percentage from methylation analysis. All five

linkage patterns were fully assigned according to the literature data (Fischer et al., 2004;

Höije et al., 2006; Mazumder et al., 2010; Pastell et al., 2008; Pastell et al., 2009; Skendi

et al., 2011; Sun et al., 2011). The chemical shift of C-2 shifted downfield by 4-6 ppm in

the C13 NMR spectrum due to the glycosylation at O-2 (Table 5.2). The chemical shift of

Page 83: Structure-Function Relationship of Flaxseed Gum from

72

C-4 moved downfield by ~4 or 7-8 ppm depending on whether its neighboring C-3 was

also glycosylated or not.

Table 5.2. 1H/13

Sugar Residues

C NMR Chemical Shifts for the neutral fraction gum (ppm)

H-1/C-1 H-2/C-2 H-3/C-3 H-4/C-4 H-5/C-5 H-6/C-6

X4 →4)-β-D-Xylp-(0→ 4.30/102.1 3.15/74.4 3.40/74.4 3.58/76.8 3.92/63.5 3.20/63.5

X234 →2,3,4)-β-D-Xylp-(1→ 4.54/100.7 3.55/80.1 3.73/77.5 3.74/73.7 4.00/62.7 3.36/62.7

Xt T-β-D-Xylp-(1→ 4.43/104.0 3.19/73.8 3.29/76.5 3.48/70.1 3.83/65.9 3.14/65.9

X24 →2,4)-β-D-Xylp-(1→ 4.40/103.9 3.48/78.3 3.65/75.2 3.78/77.8 -/- 3.37/62.8

X34 →3,4)-β-D-Xylp-(1→ 4.46/103.8 3.23/73.8 3.52/78.3 3.54/74.1 3.98/64.8 3.3/64.8

A5/T →5)-α-Araf-(l→ 5.26/108.7 4.02/81.7 3.88/78.3 4.24/83.8 3.70/70.3 -/-

A3 →3)-α-Araf-(l→ 5.16/108.7 4.17/80.9 4.02/85.2 4.19/83.6 3.61/62.0 3.67/61.3

AT T-α-Araf-(l→ 5.20/108.6 4.19/80.9 4.00/81.2 4.06/85.0 3.62/61.3 3.58/61.3

C4 →4)-α-D-Glcp-(1→ 4.78/99.6 3.39/74.5 3.48/74.1 3.59/76.8 -/- -/-

C6 →6)-α-D-Glcp-(1→ 4.97/100.5 3.37/72.4 3.48/74.1 3.48/70.2 3.56/70.2 3.76/68.8

GT T-α-D-Galp-(1→ 4.89/99.9 3.70/69.1 3.87/70.1 -/- -/- 3.63/61.9

5.3.2.2 Arabinose residues

The second abundant component in NFG, arabinose (24.8 %), existed in three linkage

patterns: →1) -L-Araf-(5→ (17.3 %), →1)-L-Araf-(3→ (4.9 %) and T-L-Araf-(1→

(2.6 %). The percentages of these three anomeric protons from methylation analysis are

in the same decreasing order as the relative ratios from integration calculation (Fig. 5.1a).

Page 84: Structure-Function Relationship of Flaxseed Gum from

73

Table 5.3. Heteronuclear (HMBC) and homonuclear (NOESY) connectivities in the

neutral fraction gum (ppm)

Inter-residue connectivities

HMBC NOESY

Atom δ Atom δ Atom δ Atom δ

A5H-1 5.26 X234C-3 77.4 A5H-1 5.26 X234H-3 3.73

X34C-3 78.3 X34H-3 3.52

X24C-2 78.3 X24H-2 3.48

GTH-1 4.88 A5C-5 70.3 GTH-1 4.88 A5H-5 3.7

C6H-1 4.97 X34C-3 78.3 C6H-1 4.97 X34H-3 3.52

X24C-2 78.3 X24H-2 3.48

C4C-1 99.6 C4H-4 3.59 C4C-1 4.78 C6H-4 3.76

X34H-3 3.52 X34 H-3 3.52

X234H-2 3.55 X234H-2 /

Additional Inter-residue connectivities in NOESY

Atom δ Atom δ Atom δ Atom δ

XTH-1 4.43 A5H-5 3.70 ATH-1 5.20 X234H-3 3.73

A3H-3 4.02 A5H-1 5.26 X234H-2 3.55

X234H-3 3.73 A3H-1 5.16 X234H-2 3.55

X234H-2 3.55 X234H-3 3.73

X34H-3 3.52 X34H-3 3.52

X24H-2 3.48

(X4: →4)-β-D-Xylp-(1→; X234: →2,3,4)-β-D-Xylp-(1→; XT: T-β-D-Xylp-(1→; X24:

→2,4)-β-D-Xylp-(1→; X34: →3,4)-β-D-Xylp-(1→; A5: →5)-α-Araf-(l→; AT: T-α-Araf-(l→; A3:

→3)-α-Araf-(l→; C4: →4)-D-α-Glcp-(1→; C6: →6)-D-α-Glcp-(1→; GT: T-D-α-Galp-(1→.)

Page 85: Structure-Function Relationship of Flaxseed Gum from

74

A full assignment for each linkage was in good agreement with previous data (Ferreira et

al., 2006; Habibi et al., 2004; Westphal et al., 2010). The chemical shifts in the C13 NMR

spectrum of glycosylated C-5 of →1) -L-Araf-(5→ and C-3 of→1) -L-Araf-(3→ moved

downfield by 4-7 and 8-10 ppm, respectively.

5.3.2.3 Minor residues

A minor portion of glucose residue in NFG mainly existed as →1) -D-Glcp-(6→

(1.4 %) and →1) -D-Glcp-(4→ (~3.5 % or less). The α-configurations of

→1)-D-Glcp-(6→ (De Castro et al., 2005; Leone et al., 2006; Muldoon et al., 2002) and

→1)-D-Glcp-(4→ (Chakraborty et al., 2004; Ghosh et al., 2008; Perepelov et al., 2011)

were evident by the cross-peaks at 4.78/99.6 and 4.97/100.5 ppm in the anomeric region

of HMQC, respectively.

Galactose residues were appended to NFG mainly in forms of T-α-D-Galp-(1→

(4.5 %) as assigned in COSY (Fig. 5.2) and TOCSY (Fig. 5.3) spectra according to the

literature (Ahrazem et al., 2002; Galbraith et al., 1999; Sengkhamparn et al., 2009).

5.3.2.4 Sequence of sugar residues

The β-1,4-linked xylose backbone was evident by the cross-peaks confined in the

dashed square in the NOESY spectrum (Fig. 5.6), where the H-1/H-4 connectivities of

linear or branched →4) -β-D-Xylp-(1→ could be observed. The evidence in HMBC (Fig.

5.5) was listed in Table 5.3: 1) the →5)-L-Araf-(1→ and →4)-D-Glcp-(1→ was appended

Page 86: Structure-Function Relationship of Flaxseed Gum from

75

to the backbone at O-3 and O-2, respectively; 2) mono-substitution with →5)-L-Araf-(l→

or →6)-D-Glcp-(1 →at either O-2 or O-3, or with →4)-D-Glcp-(1→ at O-2 were also

evident; 3) the appendence of T-D-Galp-(1→ to →5)-L-Araf-(l→ and links between

→4)-D-Glcp-(1→ also existed. All above links found in HMBC were also confirmed in

NOESY, except that links between →4) -D-Glcp-(1→ was replaced by

→6)-D-Glcp-(1→4)-D-Glcp-(1→ and that the appendence of →4) -D-Glcp-(1→

1.2-linked to →2,3,4)-D-Xylp-(1→ was missing. More evidence shown in NOESY was

listed in Table 5.1. Given the above evidence, a proposed structure of the repeating unit

of NFG is given (Fig. 5.7).

5.4 Conclusions

The neutral fraction gum (NFG) from flaxseed hulls was an arabinoxylan as indicated

by methylation analysis and 1D/2D NMR spectroscopy. NFG contained a β-1,4-linked

xylose backbone being mono-, di- or unsubstituted at O-2 and/or O-3 positions by short

branches mostly consisting of one to three sugar residues. These branches are appended

to the backbone through linear linkages like →5) -L-Araf-(1→ (17.3 %),

→3)-L-Araf-(1→ (4.9 %), →4)-D-Glcp/Galp-(1→ (3.5 %) and →6)-D-Glcp-(1→(1.4 %),

and ended with three terminals: T-D-Glcp-(1→(15.5 %), T-L-Araf-(1→(2.6 %) and

T-D-Galp-(1→(4.5%).

Page 87: Structure-Function Relationship of Flaxseed Gum from

76

Chapter 6. Conformation of Neutral and Acidic Fraction Gums

6.1. Introduction

In our previous studies, soluble flaxseed gum extracted from flaxseed hulls was

separated into a neutral and an acidic fraction using anion-exchange chromatography

(Qian et al., 2012b). Both neutral (DB=0.48) and acidic (DB=0.55) fractions were highly

branched as indicated by their high degree of branching (DB) (Qian et al., 2012a). The

neutral fraction (NFG) was an arabinoxylan and contained a β-1,4-linked xylose

backbone being mono-, di- or unsubstituted at O-2 and/or O-3 positions by short branches

consisting mainly of one to three sugar residues. In contrast, the acidic fraction (AFG)

consisted of a rhamnogalacturonan-I (RG-I) backbone intervened by limited amount of

homogalacturonan (HG) or homorhamnan (HR). Half of its most abundant neutral

residues (→2) -α-L-Rhap-(1→, 38.2%) were monosubstitued at O-3 mostly by

monosaccharides and occasionally by longer side chains consisting of HR, or RG-I, or a

blend of both.

Molecular conformation varies with the linkage type in the backbone, as well as the

appendance of side chains. Soybean soluble polysaccharide (SSPS) is a pectin-like

polysaccharide and contains a backbone consisting of rhamnogalacturanon-I (RG-I)

intervened by homogalacturonan (HG) (Akihiro et al., 2001). This backbone was highly

branched by long side chains of α-(1,3/5)-arabinans and β-(1,4)-galactans, which

constituted 21 and 50 % of total sugar respectively. With 80 % of galactan side chains

Page 88: Structure-Function Relationship of Flaxseed Gum from

77

removed by β-(1,4)-D-galactosidase (GPase), SSPS changed from an overall globular

shape (ρ=1.1) to a more linear random coil conformation (ρ=1.3), as indicated by the

increase of structure parameter ρ, but still differed from a linear random coil (ρ=1.7-2.0)

(Wang et al., 2005c). Knowledge of the conformation of both neutral and acidic fractions

from flaxseed hulls may assist in confirming their characteristic structure, and in

correlating the functions with their structures and shapes.

Light scattering is one of few techniques bringing a deep insight into molecular

structure by directly measuring the molecular parameters (Burchard, 2005). There are two

types of light scattering measurements, static and dynamic scattering. Static light

scattering measures the average total scattering intensity over a relatively long period

(1~2 s) and provides information on weight average molecular weight (Mw), the radius

of gyration (Rg), and the inter-particle interaction (the second viral coefficient, A2).

Dynamic light scattering measures the fluctuation of the intensity over time by detecting

the time-dependent autocorrelation function, and derives the hydrodynamic radius (Rh).

The objectives of the present study were to determine the molecular characteristics of

both neutral and acidic fraction gums from flaxseed hulls in aqueous solution, using static

and dynamic light scattering, and to correlate their conformational features with the

corresponding structural characteristics and functions.

Page 89: Structure-Function Relationship of Flaxseed Gum from

78

6.2. Experimental

6.2.1. Materials

A neutral (NFG) and an acidic fraction gum (AFG) were separated from soluble gum

extracted from flaxseed hulls, using anion-exchange chromatography (Qian et al., 2012b).

6.2.2. Solution preparation

Dust should be first removed from sample solutions for all light scattering

measurements by consecutive filtration four times through a 0.2 μm (for solvents) or a

0.45 μm (for sample solutions) nylon filter to prepare dust-free solutions. Consecutive

filtration was also reported to be capable of eliminating macromolecular aggregates in

solutions (Wang et al., 2005c). Stock solutions of NFG and AFG were separately

prepared by dissolving each freeze dried fraction into MilliQ water at 80℃ for 1 h under

constant stirring. After cooling down to room temperature, each stock solution was

filtered through a 0.45 μm nylon filter 4 times to remove dust and diluted by MilliQ

water and/or 1 M NaOH, or by MilliQ water and/or 1M NaCl, into various concentrations

of solutions for both static and dynamic light scattering measurements. Each sample

solution was filtered directly into a cylindrical quartz cell (25 mm in diameter), which

was immersed in a decalin bath for each measurement. Samples were fully recovered

after filtration as indicated by the unreduced concentration of total sugar content in NFG

solutions or uronic acid content in AFG solutions.

Page 90: Structure-Function Relationship of Flaxseed Gum from

79

6.2.3. Light scattering measurements

Both static and dynamic light scattering measurements were carried out on a

BI-200SM Brookhaven light scattering instrument (Brookhaven Instruments, Holtsvile,

New York, USA) with a He-Ne laser (637nm) as the monochromatic light source. This

instrument includes a precision Goniometer, a photomultiplier, and a 128-channel

BI-9000AT digital autocorrelator. Instrument alignment was conducted before every use

with well-filtered dust-free toluene. All measurements were carried out over the angular

range of 40°-150° at a room temperature of 25°C.

For static light scattering measurements, toluene (Rayleigh ratio: 1.40×10-5cm -1) was

also used as a reference. Data were presented by the Zimm plot method.

The results of dynamic light scattering measurement or the particle size distributions

were calculated by either the constrained regularization (CONTIN) method or

Non-Negatively Constrained Least Squares (NNLS) method using Brookhaven dynamic

light scattering software.

6.3. Results and discussion

6.3.1 The presence and elimination of aggregates

Various non-covalent interactions, such as van der Waals attraction, hydrogen

bonding, ionic and hydrophobic interactions (Wang et al., 2005a), may contribute

molecular aggregates to obtain a generally lower energy in solution systems. Elimination

of aggregates is critical to guarantee a precise measurement of molecular size and

Page 91: Structure-Function Relationship of Flaxseed Gum from

80

conformation.

Molecular distribution was pre-measured by dynamic light scattering (at 90o) to check

the presence or elimination of aggregates in sample solutions. NaCl or NaOH solutions (0.

1 M and 0.5 M) were applied instead of water as solvents to help reduce or retard

aggregates by interrupting the hydrogen bonding and/or charge interactions (Li et al.,

2006). The molecular size distributions of AFG (0.5 M) in both solvents are shown in Fig.

6.1.

Fig. 6.1. Molecular size distribution of acidic fraction gums in different solvents

(a: 0.1 M NaCl; b: 0.5 M NaCl; c: 0.1 M NaOH; d: 0.5 M NaOH)

Fairly broad molecular distribution ranging from 20 to 600 nm was found (Fig. 6.1 a

& b) in either 0.1M or 0.5 M NaCl solutions and the mean hydrodynamic radius Rh

reduced from 77.3 to 55.6 nm with the concentration of NaCl increased from 0.1 M to 0.5

Diameter (nm)

a

b

c

d

Page 92: Structure-Function Relationship of Flaxseed Gum from

81

M. The molecular distribution was narrowed to a range from 20 to 120 nm by 0.1 M

NaOH solutions, which was similar as in 0.5 M NaOH (20-88nm). However, the

molecular size turned to a broader distribution and the mean value of Rh also increased

with time gradually. Even in 0.5 M NaOH solution (Fig. 6.1 c & d), gradual increases in

the mean value of Rh from 29 to 63 nm were also detected due to a higher concentration

(0.8 mg/ml) of AFG within 6 h after filtration, indicating that aggregates were prone to

form gradually with time. The interval to form detectable aggregates reduced to 3 h for

1.0 mg/ml AFG in 0.5 M NaOH. Similarly, although NFG in both 0.1 and 0.5 M NaOH

solution had shown stable mean value of Rh at the beginning after filtration, yet 0.5 M

NaOH exhibited better prevention of aggregates with time. Therefore, both AFG and

NFG samples were detected for static and dynamic light scattering measurements

immediately after filtration before an aggregate initialization. Within a measuring period

of 24 h, no obvious degradation of NFG or AFG was detected.

6.3.2 Static light scattering

Static light scattering (SLS) determined weight average molecular weight (Mw),

radius of gyration (Rg), and the second virial coefficient (A2) using the Debye equation

(Debye, 1947):

cAqMRMRKc wgw 222 2)/(3/1/1/ ++=θ (6.1)

Where K is an optical contrast factor, c the polymer concentration, Rθ is the Rayleith

ration, and the scattering vector q is defined as

Page 93: Structure-Function Relationship of Flaxseed Gum from

82

00 /)2/sin(4 λθπnq = (6.2)

With n0 is the refractive index and λ0 is the wavelength in a vacuum.

The SLS data of NFG and AFG are presented with the Zimm plot (Zimm, 1948) by

plotting Kc/∆Rθ against q2+kc (Fig. 6.2), where the k is a freely chosen constant. The

slope of angular dependence at c=0 referred to the z-average mean square radius of

gyration (Rg)2 . The second virial coefficient, A2, was obtained from the slope of

concentration dependence at θ=0. Weight average molecular weight Mw equaled the value

at the interception of two extrapolated lines at θ=0 and c=0.

Fig. 6.2. Static light scattering data presented in Zimm plot obtained from the

neutral (NFG, a) and acidic (AFG, b) fraction gums in 0.5 M NaOH at 25 °C

(mol/g)

0.0 1.0

7.339e-0

Sin2(θ/2)+32C

Kc/∆Rθ

3.126e-0

b C (mg/mL) 1.3 1.0 0.6 0.3 0.1

Kc/∆Rθ

1.322e-0

(mol/g)

4.932e-0

a C (mg/mL) 1.0 0.8

0.6 0.4 0.2

0.0 1.0 Sin2(θ/2)+56C

Page 94: Structure-Function Relationship of Flaxseed Gum from

83

The parameters from obtained SLS, including Mw, Rg and A2, are summarized in Table.

6.1. The positive value of A2 for both fractions indicated the molecular-solvent

interactions were more favorable than molecular interactions, indicating that 0.5 M

NaOH was a good solvent.

Table 6.1. Molecular characteristics of the neutral (NFG) and acidic (AFG) fractions

obtained from static and dynamic light scattering

Mw

/kDa

Rg

/nm

A2

/(cm3mol/g2)

Rh

/nm ρ

NFG 616±19 49.2±3.4 9.8×10-4 27.1±1.5 1.8

AFG 285±25 38.2±2.7 9.3×10-4 29.6±2.3 1.3

(Where Mw: weight average molecular weight; Rg: the z-average radius of gyration; A2: the second

virial coefficient; Rh: the hydrodynamic radius; ρ: structural parameter.)

6.3.3 Dynamic light scattering

The dynamic light scattering measurements were applied for various concentrations

of both NFG and AFG in 0.5 M NaOH at different detecting angles (Fig. 6.3). The

dynamic radius Rh of both NFG and AFG in 0.5 M NaOH were independent of either

detection angles or sample concentrations, as no obvious correlation between Rh and

detection angle or sample concentration was found (Fig. 6.3).

The structural parameter ρ, which is equal to the ratio of Rg/Rh, is an indicator of

Page 95: Structure-Function Relationship of Flaxseed Gum from

84

macromolecular conformation and depends on the molecular structure and polydispersity,

regardless of molecular mass (Burchard, 2004): a ρ value between 1.5-2.05 indicates a

typical random coil conformation, whereas lower ρ values between 1.0-1.3 suggest a

star-like conformation.

Fig. 6.3. The angular dependence (a: NFG, 0.2 g/L; c: AFG, 0.2 g/L) and

concentration dependence (at 90o, b: NFG; d: AFG) of the hydrodynamic radius (Rh

The structural parameter ρ of NFG was 1.8 (Table 6.1), indicating a random coil

conformation (Fig. 6.4c). This was in good agreement with the presence of longer side

chains in AFG (

)

of neutral (NFG) and acidic (AFG) fractions determined in 0.5 M NaOH.

Qian et al., 2012a) which increased the segment density. Whereas NFG

has a β-1,4-xylan backbone mono-, di- or un-substituted at O-2 and/or O-3 by short

branches mostly consisting of one to three sugar residues.

a b

c d

Page 96: Structure-Function Relationship of Flaxseed Gum from

85

Fig. 6.4. Schematic representation of molecular structure (a & b) and conformation

(c & d) of neutral (NFG) and acidic (AFG) fraction gums

Lower value of ρ (1.3) for AFG (Table 6.1) suggested its star- like conformation (Fig.

6.3d) and relatively higher segment density. AFG, like soybean soluble polysaccharide

(SSPS) (Akihiro et al., 2001), was a pectin-like polysaccharide due to its

rhamnogalacturanon-I (RG-I) backbone. Thanks to the presence of high amount of

α-1,3/5-arabinans (21% of total sugar) and β-1,4-galactans (50% of total sugar) as side

chains, SSPS features a star-like conformation with a typical ρ value (1.1) (Wang et al.,

2005c). With removal of 80 % galactan branches, β-1,4-D-galactosidase treated SSPS

(GPase/SSPS) was changed into to a more linear conformation (ρ=1.3), but still varied

from a linear random coil (ρ=1.7-2.0). As the ρ value indicated, AFG seemed to exhibit a

more linear shape than natural SSPS, and to be more like GPase/SSPS. This could be

(a) NFG (b) AFG

(d) AFG

Xylose Glucose Arabinose Galactose

(c) NFG

Page 97: Structure-Function Relationship of Flaxseed Gum from

86

traced to its high amount (19.6 %) of mono-galactosyl branches (Qian et al., 2012a),

which overall shortened its average side chain length compared to that of SSPs. Yet, its

non-linear-random-coil conformation was supposed to be attributed to the existence of

longer side chains, which imparted more substantial steric hindrance than

monosaccharide side chains.

The fine structure of pectin is still in dispute due to its complexity and heterogeneity

between plants and tissues (Willats et al., 2006). The controversy between conventional

and recently proposed alternative structural models of pectin is whether the HG domain is

inserted into its backbone or linked as a side chain. A similar question also exists in

revealing the structure of AFG. Although a small portion of HG exists in AFG, we

propose its presence in both backbone and side chains. A schematic of AFG representing

its proposed structure and conformation properties is proposed (Fig. 6.4b).

6.4. Conclusions

Sodium hydroxide (0.5 M) was capable of eliminating aggregates from both neutral

and acidic fraction solutions. To retard aggregates initialization with time at higher

concentrations (0.8-1.0 mg/mL), filtration through 0.45μm filter was carried out before

each measurement. The neutral fraction exhibited random coil conformation with a

molecular weight of 616±19 kDa. Whereas acidic fraction had a star-like conformation

with a molecular weight of 285±25 kDa due to the presence of longer side chains along

Page 98: Structure-Function Relationship of Flaxseed Gum from

87

its backbone. The structural and conformational features of both fractions may account

for the distinctions in their flow behavior.

Page 99: Structure-Function Relationship of Flaxseed Gum from

88

Chapter 7. Concluding Discussion

Soluble dietary fibre with low viscosity may be a solution to increase daily dietary

fiber consumption, as it can be included in the diet in a significant amount without

over-texturization. Research interests in low viscous polysaccharide have boosted the

potential use of soluble flaxseed gum in both academic research and industry

applications.

Flaxseed hulls, as a byproduct from the dehulling process of flaxseeds, are available

in large amount in Canada since it is the leading country in producing and exporting

oil-type flaxseed. The patented separation method and equipment applied to fractionate

hulls (occupy ~37 % in whole seed mass) from flaxseed kernels facilitated the aqueous

extraction of soluble flaxseed gum (SFG) from the gum-enriched hull fraction rather than

from the whole seed by reducing the costs and raising the extraction yields.

Two fractions, a neutral (NFG) and an acidic (AFG) fraction gums, distinct in

molecular structure, conformation and flow behavior, were obtained by efficient

fractionation of SFG using ion-exchange chromatography. For the first time, the detailed

primary structure of each fraction was revealed using methylation analysis and 1D/2D

NMR, and its corresponding conformation was also investigated with static and dynamic

light scattering.

Correlating macromolecular structure and conformation with its function is one of the

most challenging topics in polysaccharide studies. The flow behavior of each fraction

Page 100: Structure-Function Relationship of Flaxseed Gum from

89

may be traced back to its primary structure and conformation. The NFG had β-1,4-linked

xylose backbone, which was mono-, di- or unsubstituted at O-2 and/or O-3 positions by

short side chains composed of one to three linearly linked sugar residues (24.8 %

arabinose, 15.5 % xylose and 9.4 % galactose/glucose). This primary structure of NFG

coincided with its random coil conformation. The limited flexibility of this random coil

chain was evident by a higher intrinsic viscosity (377.5 mL g-1), a higher Huggins

constant (0.54), and its shear-thinning behavior and relatively higher viscosity compared

with those of AFG. AFG had a highly-branched rhamnogalacturonan-I backbone

substituted at O-3 of 1,2-Rhap mostly by monomeric sugar units (total terminal sugar

units: 27.2 %) or occasionally by longer side chains with a similar structure as its

backbone. The attachment of longer side chains along the backbone of AFG was

confirmed by its star like conformation. The Newtonian flow behavior and low viscosity

arose from its low molecular weight (285±25 kDa) and star like conformation, and the

high flexibility indicated by a fairly low value of Huggins constant (0.16), as well as a

lower intrinsic viscosity (332.5 mL g-1).

With the knowledge of the flow behavior distinction between NFG and AFG,

adjusting the viscosity of soluble flaxseed gum to favored levels may be obtained by

targeting the mass ratio of NFG/AFG, which varies with flaxseed breeds. With

acknowledgement of the correlation between primary structure and flow behavior of each

fraction, targeting the mass ratio of NFG/AFG can be simplified into a detection of

monosaccharide composition.

Page 101: Structure-Function Relationship of Flaxseed Gum from

90

Functionality analyses of soluble flaxseed gum were mainly restricted to its viscosity,

viscoelastisity, surface properties and emulsification in our research. For possible food

and nonfood application of soluble flaxseed gum, other functionalities like gel forming

ability, and potential physiological functions in the gastro-intestinal tract associated with

decreasing glycemic and insulemic response, reducing serum cholesterol, and

fermentability also deserve further investigations in the near future. When the gelling

potential and mechanism of soluble flaxseed gum is considered, the possible presence of

ferulic acid and its association with the neutral fraction needs to be examined, due to its

significant contribution to the gelation of various arabinoxylans from cereal cell wall

materials. The heterogeneity of the acidic fraction in soluble gums as indicated by the

heterogeneous Mw distribution by high performance size exclusion chromatography

(HPSEC), also need some attention especially when the molecular structure is concerned

with possible mechanisms related to its potential physiological properties.

Page 102: Structure-Function Relationship of Flaxseed Gum from

91

Reference

Ahrazem, O., Prieto, A., Leal, J. A., Jiménez-Barbero, J., & Bernabé, M. (2002). Fungal cell-wall

galactomannans isolated from Geotrichum spp. and their teleomorphs, Dipodascus and

Galactomyces. Carbohydrate Research, 337(21-23), 2347-2351.

Akihiro, N., Hitoshi, F., Hirokazu, M., Yasunori, N., & Akihiro, Y. (2001). Analysis of structural

components and molecular construction of soybean soluble polysaccharides by stepwise

enzymatic degradation. Bioscience, biotechnology, and biochemistry, 65(10), 2249-2258.

Alzueta, C., Rodriguez, M. L., Cutuli, M. T., Rebole, A., Ortiz, L. T., Centeno, C., & Trevino, J.

(2003). Effect of whole and demucilaged linseed in broiler chicken diets on digesta

viscosity, nutrient utilisation and intestinal microflora. British Poultry Science, 44(1),

67-74.

Bassett, C. M. C., Rodriguez-Leyva, D., & Pierce, G. N. (2009). Experimental and clinical

research findings on the cardiovascular benefits of consuming flaxseed. Applied

Physiology Nutrition and Metabolism-Physiologie Appliquee Nutrition Et Metabolisme,

34(5), 965-974.

Bloedon, L. T., & Szapary, P. O. (2004). Flaxseed and cardiovascular risk. Nutrition Reviews,

62(1), 18-27.

Blumenkrantz, N., & Asboe-Hansen, G. (1973). New method for quantitative determination of

uronic acids. Analytical Biochemistry, 54(2), 484-489.

Brennan, C. S. (2005). Dietary fibre, glycaemic response, and diabetes. Molecular Nutrition &

Food Research, 49(6), 560-570.

Page 103: Structure-Function Relationship of Flaxseed Gum from

92

Burchard, W. (2004). Light Scattering from Polysaccharides. In Polysaccharides: CRC Press.

Burchard, W. (2005). Light Scattering from Polysaccharides. In S. Dumitriu (Ed.),

Polysaccharides: Structural Diversity and Functional Versatility (2nd ed., pp. 189-236).

New York: CRC Press.

Chakraborty, I., Mondal, S., Pramanik, M., Rout, D., & Islam, S. S. (2004). Structural

investigation of a water-soluble glucan from an edible mushroom, Astraeus

hygrometricus. Carbohydrate Research, 339(13), 2249-2254.

Choma, A., Komaniecka, I., & Sowinski, P. (2009). Revised structure of the repeating unit of the

O-specific polysaccharide from Azospirillum lipoferum strain SpBr17. Carbohydrate

Research, 344(7), 936-939.

Ciucanu, I., & Kerek, F. (1984). A simple and rapid method for the permethylation of

carbohydrates. Carbohydrate Research, 131(2), 209-217.

Colquhoun, I. J., de Ruiter, G. A., Schols, H. A., & Voragen, A. G. J. (1990). Identification by

n.m.r. spectroscopy of oligosaccharides obtained by treatment of the hairy regions of

apple pectin with rhamnogalacturonase. Carbohydrate Research, 206(1), 131-144.

Cui, S. W. (2000). Flaxseed Gum In S. W. Cui (Ed.), Polysaccharide Gums from Agricultural

Products: Processing, Structures and Functionality (pp. 59-101). Lancaster: Technomic

Publishing Company, Inc.

Cui, S. W. (2005). Structural Analysis of Polysaccharides. In S. W. Cui (Ed.), Food

Carbohydrates: Chemistry, Physical Properties, and Applications (pp. 105-160). Boca

Raton: CRC Press.

Page 104: Structure-Function Relationship of Flaxseed Gum from

93

Cui, W., Eskin, M. N. A., Biliaderis, C. G., & Marat, K. (1996a). NMR characterization of a

4-O-methyl-β-d-glucuronic acid-containing rhamnogalacturonan from yellow mustard

(Sinapis alba L.) mucilage. Carbohydrate Research, 292(0), 173-183.

Cui, W., & Han, N. F. (2006). Process and apparatus for flaxseed component separation In (Vol.

U.S. Patent 7,022,363). Guelph, CA.

Cui, W., Kenaschuk, E., & Mazza, G. (1996b). Influence of genotype on chemical composition

and rheological properties of flaxseed gums. Food Hydrocolloids, 10(2), 221-227.

Cui, W., & Mazza, G. (1996c). Physicochemical characteristics of flaxseed gum. Food Research

International, 29(3-4), 397-402.

Cui, W., Mazza, G., & Biliaderis, C. G. (1994a). Chemical structure, molecular size distributions,

and rheological properties of flaxseed gum. Journal of Agricultural and Food Chemistry,

42(9), 1891-1895.

Cui, W., Mazza, G., Oomah, B. D., & Biliaderis, C. G. (1994b). Optimization of an aqueous

extraction process for flaxseed gum by response surface methodology. Lebensmittel

Wissenschaft und Technologie, 27(4), 363-369.

Cui, W., Wood, P. J., Blackwell, B., & Nikiforuk, J. (2000). Physicochemical properties and

structural characterization by two-dimensional NMR spectroscopy of wheat

[beta]-D-glucan--comparison with other cereal [beta]-D-glucans. Carbohydrate Polymers,

41(3), 249-258.

Cunnane, S. C., & Thompson, L. U. (1995). Flaxseed in human nutrition (1 ed.). Champaign,

Illinois: AOCS Press.

Page 105: Structure-Function Relationship of Flaxseed Gum from

94

De Castro, C., Molinaro, A., Lanzetta, R., Holst, O., & Parrilli, M. (2005). The linkage between

O-specific caryan and core region in the lipopolysaccharide of Burkholderia caryophylli

is furnished by a primer monosaccharide. Carbohydrate Research, 340(11), 1802-1807.

Debye, P. (1947). Molecular-weight determination by light scattering. Journal of Physical and

Colloid Chemistry, 51(1), 18-32.

Deng, C., O’Neill, M. A., & York, W. S. (2006). Selective chemical depolymerization of

rhamnogalacturonans. Carbohydrate Research, 341(4), 474-484.

Denis, L., Barbara, P., & Dominique, J.-R. (2007). Digestible and indigestible carbohydrates:

interactions with postprandial lipid metabolism. The Journal of nutritional biochemistry,

18(4), 217-227.

Diederichsen, A., Raney, J. P., & Duguid, S. D. (2006). Variation of mucilage in flax seed and its

relationship with other seed characters. Crop Science, 46(1), 365-371.

Dikeman, C. L., & Fahey, G. C. (2006). Viscosity as related to dietary fiber: A review. Critical

Reviews in Food Science and Nutrition, 46(8), 649-663.

Doyle, J. P., Lyons, G., & Morris, E. R. (2009). New proposals on "hyperentanglement" of

galactomannans: Solution viscosity of fenugreek gum under neutral and alkaline

conditions. Food Hydrocolloids, 23(6), 1501-1510.

Eastwood, M. A., & Morris, E. R. (1992). Physical properties of dietary fiber that influence

physiological function: a model for polymers along the gastrointestinal tract. The

American journal of clinical nutrition, 55(2), 436-442.

Edwards, S., Chaplin, M. F., Blackwood, A. D., & Dettmar, P. W. (2003). Primary structure of

Page 106: Structure-Function Relationship of Flaxseed Gum from

95

arabinoxylans of ispaghula husk and wheat bran. Proceedings of the Nutrition Society,

62(01), 217-222.

Fedeniuk, R. W., & Biliaderis, C. G. (1994). Composition and Physicochemical Properties of

Linseed (Linum usitatissimum L.) Mucilage. Journal of Agricultural and Food Chemistry,

42(2), 240-247.

Fedonenko, Y. P., Konnova, O. N., Zdorovenko, E. L., Konnova, S. A., Zatonsky, G. V., Shashkov,

A. S., Ignatov, V. V., & Knirel, Y. A. (2008). Structural analysis of the O-polysaccharide

from the lipopolysaccharide of Azospirillum brasilense S17. Carbohydrate Research,

343(4), 810-816.

Ferguson, M. J., & Jones, G. P. (2000). Production of short-chain fatty acids following in vitro

fermentation of saccharides, saccharide esters, fructo-oligosaccharides, starches,

modified starches and non-starch polysaccharides. Journal of the Science of Food and

Agriculture, 80(1), 166-170.

Ferreira, J. A., Mafra, I., Soares, M. R., Evtuguin, D. V., & Coimbra, M. A. (2006). Dimeric

calcium complexes of arabinan-rich pectic polysaccharides from Olea europaea L. cell

walls. Carbohydrate Polymers, 65(4), 535-543.

Fischer, M. H., Yu, N., Gray, G. R., Ralph, J., Anderson, L., & Marlett, J. A. (2004). The

gel-forming polysaccharide of psyllium husk (Plantago ovata Forsk). Carbohydrate

Research, 339(11), 2009-2017.

Fodje, A. M. L., Chang, P. R., & Leterme, P. (2009). In vitro bile acid binding and short-chain

fatty acid profile of flax fiber and ethanol co-products. Journal of Medicinal Food, 12(5),

Page 107: Structure-Function Relationship of Flaxseed Gum from

96

1065-1073.

Galbraith, L., Sharples, J. L., & Wilkinson, S. G. (1999). Structure of the O-specific

polysaccharide for Acinetobacter baumannii serogroup O1. Carbohydrate Research,

319(1-4), 204-208.

Garti, N., & Leser, M. E. (2001). Emulsification properties of hydrocolloids. Polymers for

Adwanced Technologies(12), 123-135.

Garti, N., & Reichman, D. (1994). Surface properties and emulsification activity of

galactomannans. Food Hydrocolloids, 8(2), 155-173.

Ghosh, K., Chandra, K., Ojha, A. K., & Islam, S. S. (2008). NMR and MALDI-TOF analysis of a

water-soluble glucan from an edible mushroom, Volvariella diplasia. Carbohydrate

Research, 343(16), 2834-2840.

Giacco, R., Brighenti, F., Parillo, M., Ciardullo, A. V., Capuano, M., Rivieccio, A. M., Rivellese,

A. A., & Riccardi, G. (1998). Are the beneficial metabolic effects of fibre enriched foods

preserved when consumed within a composite meal? Diabetes Nutrition & Metabolism,

11(2), 130-135.

Gibson, G. R., Probert, H. M., Van Loo, J., Rastall, R. A., & Roberfroid, M. B. (2004). Dietary

modulation of the human colonic microbiota: updating the concept of prebiotics.

Nutrition Research Reviews, 17(2), 259-275.

Glinsky, V. V., & Raz, A. (2009). Modified citrus pectin anti-metastatic properties: one bullet,

multiple targets. Carbohydrate Research, 344(14), 1788-1791.

Glitsø, L. V., Gruppen, H., Schols, H., Højsgaard, S., Sandström, B., & Knudsen, K. E. B. (1999).

Page 108: Structure-Function Relationship of Flaxseed Gum from

97

Degradation of rye arabinoxylans in the large intestine of pigs. Journal of the Science of

Food and Agriculture, 79(7), 961-969.

Glitsø, L. V., Jensen, B. B., & Knudsen, K. E. B. (2000). In vitro fermentation of rye

carbohydrates including arabinoxylans of different structure. Journal of the Science of

Food and Agriculture, 80(8), 1211-1218.

Goh, K. K. T., Pinder, D. N., Hall, C. E., & Hemar, Y. (2006). Rheological and light scattering

properties of flaxseed polysaccharide aqueous solutions. Biomacromolecules, 7(11),

3098-3103.

Guilloux, K., Gaillard, I., Courtois, J., Courtois, B., & Petit, E. (2009). Production of

arabinoxylan-oligosaccharides from flaxseed (Linum usitatissimum). Journal of

Agricultural and Food Chemistry, 57(23), 11308-11313.

Höije, A., Sandström, C., Roubroeks, J. P., Andersson, R., Gohil, S., & Gatenholm, P. (2006).

Evidence of the presence of 2-O-β-d-xylopyranosyl-α-l-arabinofuranose side chains in

barley husk arabinoxylan. Carbohydrate Research, 341(18), 2959-2966.

Habibi, Y., Heyraud, A., Mahrouz, M., & Vignon, M. R. (2004). Structural features of pectic

polysaccharides from the skin of Opuntia ficus-indica prickly pear fruits. Carbohydrate

Research, 339(6), 1119-1127.

Hall Iii, C., Tulbek, M. C., & Xu, Y. (2006). Flaxseed. Advances in food and nutrition research,

Volume 51, 1-97.

Hamer, H. M., Jonkers, D., Venema, K., Vanhoutvin, S., Troost, F. J., & Brummer, R. J. (2008).

Review article: the role of butyrate on colonic function. Alimentary Pharmacology &

Page 109: Structure-Function Relationship of Flaxseed Gum from

98

Therapeutics, 27(2), 104-119.

Harding, S. E. (1997). The intrinsic viscosity of biological macromolecules. Progress in

measurement, interpretation and application to structure in dilute solution. Progress in

Biophysics and Molecular Biology, 68(2-3), 207-262.

Hawker, C. J., Lee, R., & Frechet, J. M. J. (1991). One-step synthesis of hyperbranched dendritic

polyesters. Journal of the American Chemical Society, 113(12), 4583-4588.

Hill, S. E. (1998). Emulsions and foams. In S. E. Hill, D. A. Ledward & J. R. Mitchell (Eds.),

Functional properties of food macromolecues (2nd edition) (pp. 303-334): Aspen

Publishers.

Huang, X., Kakuda, Y., & Cui, W. (2001). Hydrocolloids in emulsions: particle size distribution

and interfacial activity. food Hydrocolloids(15), 533-542.

Hughes, S. A., Shewry, P. R., Li, L., Gibson, G. R., Sanz, M. L., & Rastall, R. A. (2007). In vitro

fermentation by human fecal microflora of wheat arabinoxylans. Journal of Agricultural

and Food Chemistry, 55(11), 4589-4595.

Izydorczyk, M. S. (2009). Arabinoxylans. In G. O. Phillips & P. A. Williams (Eds.), Handbook of

hydrocolloids (2nd ed., pp. 653-692). New Delhi: Woodhead Press.

Izydorczyk, M. S., Biliaderis, C. G., & Bushuk, W. (1991). Physical properties of water-soluble

pentosans from different wheat varieties. . Cereal Chemistry, 68, 145-150.

Kang, J., Cui, S. W., Chen, J., Phillips, G. O., Wu, Y., & Wang, Q. (2011). New studies on gum

ghatti (Anogeissus latifolia) part I. Fractionation, chemical and physical characterization

of the gum. Food Hydrocolloids, 25(8), 1984-1990.

Page 110: Structure-Function Relationship of Flaxseed Gum from

99

Kendall, C. W. C., Esfahani, A., & Jenkins, D. J. A. (2010). The link between dietary fibre and

human health. Food Hydrocolloids, 24(1), 42-48.

Khalloufi, S., Alexander, M., Khalloufi, S., Alexander, M., Goff, H. D., & Corredig, M. (2008).

Physicochemical properties of whey protein isolate stabilized oil-in-water emulsions

when mixed with flaxseed gum at neutral pH. Food Research International, 41(10),

964-972.

Khouryieh, H. A., Herald, T. J., Aramouni, F., & Alavi, S. (2007). Intrinsic viscosity and

viscoelastic properties of xanthan/guar mixtures in dilute solutions: Effect of salt

concentration on the polymer interactions. Food Research International, 40(7), 883-893.

Lapasin, R., & Pricl, S. (1995a). The Polysaccharides: Sources and Structures. In Rheology of

Industrial Polysaccharides: Theory and Applications (pp. 1-134). London: Springer.

Lapasin, R., & Pricl, S. (1995b). Rheology of polysaccharide systems. In Rheology of Industrial

Polysaccharides: Theory and Applications (pp. 267-296). London: Springer.

Launay, B., Doublier, J. L., & Guvelier, G. (1985). Flow Properties of Aqueous Solutions and

Dispersions of Polysaccharides. In J. R. Mitchell & D. A. Ledward (Eds.), Functional

Properties of Food Macromolecules (1st Edition ed., pp. 1-78). New York: Elsevier

Applied Science Publishers LTD.

Leone, S., Molinaro, A., Gerber, I. B., Dubery, I. A., Lanzetta, R., & Parrilli, M. (2006). The

O-chain structure from the LPS of the endophytic bacterium Burkholderia cepacia strain

ASP B 2D. Carbohydrate Research, 341(18), 2954-2958.

Ley, R. E., Turnbaugh, P. J., Klein, S., & Gordon, J. I. (2006). Microbial ecology: Human gut

Page 111: Structure-Function Relationship of Flaxseed Gum from

100

microbes associated with obesity. Nature, 444(7122), 1022-1023.

Li, W., Wang, Q., Cui, S. W., Huang, X., & Kakuda, Y. (2006). Elimination of aggregates of

(1→3) (1→4)-β-D-glucan in dilute solutions for light scattering and size exclusion

chromatography study. Food Hydrocolloids, 20(2–3), 361-368.

MacLean, L. L., Vinogradov, E., Pagotto, F., Farber, J. M., & Perry, M. B. (2010). The structure

of the O-antigen of Cronobacter sakazakii HPB 2855 isolate involved in a neonatal

infection. Carbohydrate Research, 345(13), 1932-1937.

Mazumder, K., & York, W. S. (2010). Structural analysis of arabinoxylans isolated from

ball-milled switchgrass biomass. Carbohydrate Research, 345(15), 2183-2193.

Mazza, G., & Biliaderis, C. G. (1989). Functional properties of flax seed mucilage. Journal of

Food Science, 54(5), 1302-1305.

McClements, D. J. (2005). Molecular characteristics. In Food emulsions: principles, practices,

and techniques: CRC Press.

Mentes, O., Bakkalbasi, E., & Ercan, R. (2008). Effect of the use of ground flaxseed on quality

and chemical composition of bread. Food Science and Technology International, 14(4),

299-306.

Mikshina, P. V., Gurjanov, O. P., Mukhitova, F. K., Petrova, A. A., Shashkov, A. S., & Gorshkova,

T. A. (2012). Structural details of pectic galactan from the secondary cell walls of flax

(Linum usitatissimum L.) phloem fibres. Carbohydrate Polymers, 87(1), 853-861.

Morris, & Edwin, R. E. (1979). Polysaccharide Conformation and Interactions in Solutions and

Gels. In W. J. Gettins & E. Wyn-Jones (Eds.), Techniques and Applications of Fast

Page 112: Structure-Function Relationship of Flaxseed Gum from

101

Reactions in Solution (Vol. 50, pp. 379-388): Springer Netherlands.

Morris, E. R., Cutler, A. N., Ross-Murphy, S. B., Rees, D. A., & Price, J. (1981a). Concentration

and shear rate dependence of viscosity in random coil polysaccharide solutions.

Carbohydrate Polymers, 1(1), 5-21.

Morris, E. R., & Ross-Murphy, S. B. (1981b). Chain flexibility of polysaccharides and

glycoproteins from viscosity measurements. Techniques in Carbohydrate Metabolism,

B310, 1-46.

Morris, V., Gromer, A., Kirby, A., Bongaerts, R., & Patrick Gunning, A. (2011). Using AFM and

force spectroscopy to determine pectin structure and (bio) functionality. Food

Hydrocolloids, 25(2), 230-237.

Mueller, K., Eisner, P., Yoshie-Stark, Y., Nakada, R., & Kirchhoff, E. (2010). Functional

properties and chemical composition of fractionated brown and yellow linseed meal

(Linum usitatissimum L.). Journal of Food Engineering, 98(4), 453-460.

Muldoon, J., Shashkov, A. S., Moran, A. P., Ferris, J. A., Senchenkova, S. y. N., & Savage, A. V.

(2002). Structures of two polysaccharides of Campylobacter jejuni 81116. Carbohydrate

Research, 337(21–23), 2223-2229.

Muralikrishna, G., Salimath, P. V., & Tharanathan, R. N. (1987). Structural features of an

arabinoxylan and a rhamno-galacturonan derived from linseed mucilage. Carbohydrate

Research, 161(2), 265-271.

Naran, R., Chen, G. B., & Carpita, N. C. (2008). Novel rhamnogalacturonan I and arabinoxylan

polysaccharides of flax seed mucilage. Plant physiology, 148(1), 132-141.

Page 113: Structure-Function Relationship of Flaxseed Gum from

102

Ojha, A. K., Maiti, D., Chandra, K., Mondal, S., Roy, D. D. S. K., Ghosh, K., & Islam, S. S.

(2008). Structural assignment of a heteropolysaccharide isolated from the gum of

Cochlospermum religiosum (Katira gum). Carbohydrate Research, 343(7), 1222-1231.

Oomah, B. D., Kenaschuk, E. O., Cui, W., & Mazza, G. (1995). Variation in the composition of

water-soluble polysaccharides in flaxseed. Journal of Agricultural and Food Chemistry,

43(6), 1484-1488.

Pastell, H., Tuomainen, P., Virkki, L., & Tenkanen, M. (2008). Step-wise enzymatic preparation

and structural characterization of singly and doubly substituted

arabinoxylo-oligosaccharides with non-reducing end terminal branches. Carbohydrate

Research, 343(18), 3049-3057.

Pastell, H., Virkki, L., Harju, E., Tuomainen, P., & Tenkanen, M. (2009). Presence of 1 →3-linked

2-O-β-d-xylopyranosyl-α-l-arabinofuranosyl side chains in cereal arabinoxylans.

Carbohydrate Research, 344(18), 2480-2488.

Pawan, K. A. (1992). NMR Spectroscopy in the structural elucidation of oligosaccharides and

glycosides. Phytochemistry, 31(10), 3307-3330.

Perepelov, A. V., Liu, B., Senchenkova, S. y. N., Guo, D., Shevelev, S. D., Feng, L., Shashkov, A.

S., Wang, L., & Knirel, Y. A. (2011). O-antigen structure and gene clusters of Escherichia

coli O51 and Salmonella enterica O57; another instance of identical O-antigens in the

two species. Carbohydrate Research, 346(6), 828-832.

Prasad, K. (2009). Flaxseed and cardiovascular health. Journal of Cardiovascular Pharmacology,

54(5), 369-377.

Page 114: Structure-Function Relationship of Flaxseed Gum from

103

Qian, K.-Y., Cui, S. W., Nikiforuk, J., & Goff, H. D. (2012a). Structural elucidation of

rhamnogalacturonans from flaxseed hulls. Carbohydrate Research, 362(0), 47-55.

Qian, K. Y., Cui, S. W., Wu, Y., & Goff, H. D. (2012b). Flaxseed gum from flaxseed hulls:

Extraction, fractionation, and characterization. Food Hydrocolloids, 28(2), 275-283.

Rees, D. A., & Scott, W. E. (1971). Polysaccharide conformation. Part VI. Computer

model-building for linear and branched pyranoglycans. Correlations with biological

function. Preliminary assessment of inter-residue forces in aqueous solution. Further

interpretation of optical rotation in terms of chain conformation. Journal of the Chemical

Society B: Physical Organic(0), 469-479.

Rendon-Villalobas, J. R., Bello-Perez, L. A., Agama-Acevedo, E., Islas-Hernandez, J. J.,

Osorio-Diaz, P., & Tovar, J. (2009). Composition and characteristics of oil extracted from

flaxseed-added corn tortilla. Food Chemistry, 117(1), 83-87.

Rose, D. J., Patterson, J. A., & Hamaker, B. R. (2009). Structural Differences among

Alkali-Soluble Arabinoxylans from Maize (Zea mays), Rice (Oryza sativa), and Wheat

(Triticum aestivum) Brans Influence Human Fecal Fermentation Profiles. Journal of

Agricultural and Food Chemistry, 58(1), 493-499.

Sengkhamparn, N., Bakx, E. J., Verhoef, R., Schols, H. A., Sajjaanantakul, T., & Voragen, A. G. J.

(2009). Okra pectin contains an unusual substitution of its rhamnosyl residues with acetyl

and alpha-linked galactosyl groups. Carbohydrate Research, 344(14), 1842-1851.

Shashkov, A. S., Vinogradov, E. V., Knirel, Y. A., Nifant'ev, N. E., Kochetkov, N. K., Dabrowski,

J., Kholodkova, E. V., & Stanislavsky, E. S. (1993). Structure of the O-specific

Page 115: Structure-Function Relationship of Flaxseed Gum from

104

polysaccharide of Salmonella arizonae O45. Carbohydrate Research, 241(0), 177-188.

Skendi, A., Biliaderis, C. G., Izydorczyk, M. S., Zervou, M., & Zoumpoulakis, P. (2011).

Structural variation and rheological properties of water-extractable arabinoxylans from

six Greek wheat cultivars. Food Chemistry, 126(2), 526-536.

Stephen, E. (1995). Rheology of industrial polysaccharides. Theory and applications: Edited by

Romano Lapasin and Sabrina Pricl, Blackie Academic and Professional, London 1995.

ISBN 0-7514-0211-7, 620 pp, £99.00. Carbohydrate Research, 276(1), C1-C3.

Sun, Y., Cui, S. W., Gu, X., & Zhang, J. (2011). Isolation and structural characterization of water

unextractable arabinoxylans from Chinese black-grained wheat bran. Carbohydrate

Polymers, 85(3), 615-621.

Swennen, K., Courtin, C., & Delcour, J. (2006). Non-digestible Oligosaccharides with Prebiotic

Properties. Critical Reviews in Food Science & Nutrition, 46(6), 459-471.

Tao, Y., Zhang, L., Yan, F., & Wu, X. (2007). Chain Conformation of Water-Insoluble

Hyperbranched Polysaccharide from Fungus. Biomacromolecules, 8(7), 2321-2328.

Thakur, G., Mitra, A., Pal, K., & Rousseau, D. (2009). Effect of flaxseed gum on reduction of

blood glucose and cholesterol in type 2 diabetic patients. International Journal of Food

Sciences and Nutrition, 60, 126-136.

Theuwissen, E., & Mensink, R. P. (2008). Water-soluble dietary fibers and cardiovascular disease.

Physiology & Behavior, 94(2), 285-292.

Thompson, L. U., & Cunnane, S. C. (2003). Flaxseed in human nutrition (2 ed.). Champaign,

Illinois: AOCS Press.

Page 116: Structure-Function Relationship of Flaxseed Gum from

105

Van Craeyveld, V., Swennen, K., Dornez, E., Van de Wiele, T., Marzorati, M., Verstraete, W.,

Delaedt, Y., Onagbesan, O., Decuypere, E., Buyse, J., De Ketelaere, B., Broekaert, W. F.,

Delcour, J. A., & Courtin, C. M. (2008). Structurally Different Wheat-Derived

Arabinoxylooligosaccharides Have Different Prebiotic and Fermentation Properties in

Rats. J. Nutr., 138(12), 2348-2355.

Vidal, S., Doco, T., Williams, P., Pellerin, P., York, W. S., O’Neill, M. A., Glushka, J., Darvill, A.

G., & Albersheim, P. (2000). Structural characterization of the pectic polysaccharide

rhamnogalacturonan II: evidence for the backbone location of the aceric acid-containing

oligoglycosyl side chain. Carbohydrate Research, 326(4), 277-294.

Viuda-Martos, M., Lopez-Marcos, M. C., Fernandez-Lopez, J., Sendra, E., Lopez-Vargas, J. H., &

Perez-Alvarez, J. A. (2010). Role of fiber in cardiovascular diseases: a review.

Comprehensive Reviews in Food Science and Food Safety, 9(2), 240-258.

Wang, Q., & Cui, S. W. (2005a). Understanding the Physical Properties of Food Polysaccharides.

In S. W. Cui (Ed.), Food Carbohydrates: Chemistry, Physical Properties, and

Applications (pp. 161-217). Boca Raton: CRC Press.

Wang, Q., & Cui, W. S. (2005b). Understanding the physical properties of food polysaccharides.

In S. W. Cui (Ed.), Food carbohydrates: chemistry, physical properties, and applications

(pp. 187). Boca Raton: Taylor & Francis

Wang, Q., Huang, X., Nakamura, A., Burchard, W., & Hallett, F. R. (2005c). Molecular

characterisation of soybean polysaccharides: an approach by size exclusion

chromatography, dynamic and static light scattering methods. Carbohydrate Research,

Page 117: Structure-Function Relationship of Flaxseed Gum from

106

340(17), 2637-2644.

Warrand, J., Michaud, P., Picton, L., Muller, G., Courtois, B., Ralainirina, R., & Courtois, J.

(2003). Large-scale purification of water-soluble polysaccharides from flaxseed mucilage,

and isolation of a new anionic polymer. Chromatographia, 58(5), 331-335.

Warrand, J., Michaud, P., Picton, L., Muller, G., Courtois, B., Ralainirina, R., & Courtois, J.

(2005a). Contributions of Intermolecular Interactions between Constitutive

Arabinoxylans to the Flaxseeds Mucilage Properties. Biomacromolecules, 6(4),

1871-1876.

Warrand, J., Michaud, P., Picton, L., Muller, G., Courtois, B., Ralainirina, R., & Courtois, J.

(2005b). Flax (Linum usitatissimum) seed cake: A potential source of high molecular

weight arabinoxylans? Journal of Agricultural and Food Chemistry, 53(5), 1449-1452.

Warrand, J., Michaud, P., Picton, L., Muller, G., Courtois, B., Ralainirina, R., & Courtois, J.

(2005c). Structural investigations of the neutral polysaccharide of Linum usitatissimum L.

seeds mucilage. International Journal of Biological Macromolecules, 35(3-4), 121-125.

Westphal, Y., Kühnel, S., de Waard, P., Hinz, S. W. A., Schols, H. A., Voragen, A. G. J., &

Gruppen, H. (2010). Branched arabino-oligosaccharides isolated from sugar beet

arabinan. Carbohydrate Research, 345(9), 1180-1189.

Willats, W. G. T., Knox, J. P., & Mikkelsen, J. D. (2006). Pectin: new insights into an old polymer

are starting to gel. Trends in Food Science & Technology, 17(3), 97-104.

Wong, J. M. W., de Souza, R., Kendall, C. W. C., Emam, A., & Jenkins, D. J. A. (2006). Colonic

Health: Fermentation and Short Chain Fatty Acids. Journal of Clinical Gastroenterology,

Page 118: Structure-Function Relationship of Flaxseed Gum from

107

40(3), 235-243.

Wood, P. J. (2007). Rheology and physiology of soluble fibres: what are the relationships and

what use can be made of them? In H. Salovaara, F. Gates & M. Tenkanen (Eds.), Dietary

Fibre Components and Functions (pp. 113-125). Wageningen: Wageningen Academic

Publishers.

Wu, Y., Cui, W., Eskin, N. A. M., & Goff, H. D. (2009). Fractionation and partial characterization

of non-pectic polysaccharides from yellow mustard mucilage. Food Hydrocolloids, 23(6),

1535-1541.

Yu, L., Lutterodt, H., & Cheng, Z. (2008). Chapter 4 Beneficial health properties of psyllium and

approaches to improve its functionalities. Advances in food and nutrition research,

Volume 55, 193-220.

Yu, L. G., & Perret, J. (2003). Effects of xylanase treatments on gelling and water-uptaking

properties of psyllium. Journal of Agricultural and Food Chemistry, 51(2), 492-495.