71
Xyloglucan-based polymers and nanocomposites – modification, properties and barrier film applications Joby Kochumalayil Jose Doctoral Thesis KTH, Royal Institute of Technology School of Chemical Science and Engineering Department of Fibre and Polymer Technology Division of Biocomposites

Xyloglucan-based polymers and nanocomposites

  • Upload
    others

  • View
    9

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Xyloglucan-based polymers and nanocomposites

Xyloglucan-based polymers and nanocomposites –

modification, properties and barrier film applications

Joby Kochumalayil Jose

Doctoral Thesis

KTH, Royal Institute of Technology

School of Chemical Science and Engineering

Department of Fibre and Polymer Technology

Division of Biocomposites

Page 2: Xyloglucan-based polymers and nanocomposites

TRITA-CHE Report 2012:53 KTH School of Chemistry

ISSN 1654-1081 SE-100 44 Stockholm

ISBN 978-91-7501-528-6 SWEDEN

AKADEMISK AVHANDLING

Som med tillstånd av Kungliga Tekniska Högskolan framläggs till offentlig granskning för avläggande av teknologie doktorsexamen i polymerteknologi fredagen den 21 december 2012 kl 10:00 i sal F3, Lindstedtsvägen 26, KTH, Stockholm. Avhandlingen försvaras på engelska.

© Joby Kochumalayil Jose, 2012

Tryck: Eprint AB 2012

Page 3: Xyloglucan-based polymers and nanocomposites

iii

ABSTRACTBiopolymers from renewable resources are of interest for packaging applications as an alternative to conventional petroleum-based polymers. One of the major application areas for biopolymers is food packaging, where a candidate polymer should meet critical requirements such as mechanical and oxygen barrier performance, also in humid conditions. Starch has long been used in certain packaging applications, either in plasticized state or blended with other polymers. However, native starch has high sensitivity to water and low mechanical and barrier performance. Recently, wood-derived hemicelluloses have been extensively studied as oxygen barrier films, but suffer from low film-forming ability and mechanical performance. In the present study, xyloglucan (XG) from tamarind seed waste is explored as an alternative high-performance biopolymer in packaging applications. The obstacles of polysaccharides in terms of moisture sensitivity and processability are addressed in this thesis.

In Paper I, film properties of XG were studied. XG has a cellulose backbone, but unlike cellulose, it is mostly soluble in water forming highly robust films. Moisture sorption isotherms, tensile tests and dynamic mechanical thermal analysis were performed. Enzymatic modification (partial removal of galactose in side chains of XG) was performed to study the effect of galactose on solubility and film-forming characteristics. XG films showed lower moisture sorption than starch. Stiffness and tensile strength were very high of the order of 4 GPa and 70 MPa respectively, with considerable ductility and toughness. The thermomechanical performance was very high with a softening temperature near 260 ºC.

In Paper II, several plasticizers were studied in order to facilitate thermal processing of XG films: sorbitol, urea, glycerol and polyethylene oxide. Films of different compositions were prepared and studied for thermomechanical and tensile properties. Highly favorable characteristics were found with XG/sorbitol system. A large drop in glass transition temperature (Tg) of XG of the order of 100 ºC with 20 - 30 wt% sorbitol was observed with an attractive combination of increased toughness.

In Paper III, XG was chemically modified and the structure-property relationship of modified XG studied. XG modification was performed using an approach involving periodate oxidation followed by reduction. The oxidation is highly regioselective, where the side chains of XG are mostly affected with the cellulose backbone well-preserved as noticed from MALDI-TOF-MS and carbohydrate analysis. Films were cast from water and characterized by dynamic mechanical thermal analysis, dynamic water vapor sorption, oxygen transmission analysis and tensile tests. Property changes were interpreted from structural changes. The regioselective modification results in new types of cellulose derivatives without the need for harmful solvents.

In Paper IV, moisture durability of XG was addressed by dispersing montmorillonite (MTM) platelets in water suspension. Oriented bionanocomposite coatings with strong in-plane orientation of clay platelets were prepared. A continuous water-based processing approach was adopted in view of easy scaling up. The resulting nanocomposites were characterized by FE-SEM, TEM, and XRD. XG adsorption on MTM was measured by quartz crystal microbalance analysis. Mechanical and gas barrier properties were measured, also at high relative humidity. The reinforcement in mechanical properties and effects on barrier properties were remarkable, also in humid conditions.

In Paper V, cross-linked XG/MTM composite was prepared with high clay content (ca. 45 vol%) by an industrially scalable “paper-making” method. Instead of using cross-linking molecules, cross-linking sites were created on the XG chain by selective oxidation of side chains. The in-plane orientation of MTM platelets were studied using XRD and FE-SEM. The mechanical properties and barrier performance were evaluated for the resulting 'nacre-mimetic' nanocomposites. The elastic modulus of cross-linked nanocomposites is as high as 30 GPa, one of the stiffest bionanocomposites reported.

Page 4: Xyloglucan-based polymers and nanocomposites

iv

SAMMANFATTNINGBiopolymerer är av stort intresse i förpackningstillämpningar. När det gäller livsmedel, så bör ett lämpligt material ha goda mekaniska egenskaper men också bra barriäregenskaper. Stärkelse används flitigt, antingen med mjukgörare eller i kombination med andra polymerer. Ett av de stora problemen med stärkelse är en hög fuktkänslighet men också låga barriär- och mekaniska egenskaper. Hemicellulosor från ved är också intressanta men dessa material har dåliga filmbildnings och mekaniska egenskaper. Den här avhandlingen är den första där xyloglukan (XG) från tamarindfrö har använts som en högpresterande biopolymer för förpackningar. Studien angriper de klassiska fukt- och processproblemen för polysackarider genom att kemiskt eller enzymatiskt modifiera XG. Modifiering av XG genomförs också med hjälp av mjukgörare eller genom att tillsätta nanopartiklar.

I artikel I studeras filmegenskaperna hos XG. Polymeren har en huvudkedja av cellulosa med substituerade sidokedjor. XG har acceptabel vattenlöslighet och kan bilda högpresterande filmer. Fuktadsorption, dragprovning, och dynamisk mekanisk analys har utförts för att karakterisera de tillverkade filmerna. Enzymatisk modifiering (galaktosgrupper avlägsnas) har genomförts och effekterna på löslighet och filmbildning har utvärderats. fuktadsorptionen är mycket lägre än för stärkelse och det har dessutom klarlagts att XG har hög E-modul och draghållfasthet, men även brott-töjning och brottarbete är höga. Termomekaniska prestanda är mycket goda med en glasomvandlingstemperatur nära 260°C.

I artikel II studeras möjligheten att använda mjukgörare för att underlätta för termiska tillverkningsprocesser. Materialets sammansättning har varierats och de termomekaniska egenskaperna studerats. XG/sorbitol visar goda egenskaper. Glasomvandlingstemperaturen sjunker med ca 100°C vid 20-30% tillsats av sorbitol, samtidigt som brottarbetet vid dragprovning ökar.

I artikel III genomfördes en kontrollerad kemisk modifiering av XG och grundläggande struktur-egenskapsrelationer etablerades. Den kemiska modifieringen bestod av periodatoxidering följd av reduktion. Oxidationen är starkt regioselektiv på så sätt att endast sidogrupperna påverkas. Förändringarna i kemisk struktur har studerats med hjälp av MALDI-TOF-MS och kolhydratanalys. Filmer har tillverkats genom gjutning från vattenlösning och studerats med termogravimetri, DMA, vattensorption, barriäregenskapsmätningar och dragprovning. Den valda modifieringsmetoden är regioselektiv, kan kontrolleras med stor precision och kan användas för att framställa nya cellulosaderivat utan användning av organiska lösningsmedel.

I artikel IV har olika metoder att förbättra fuktresistensen hos XG undersökts genom att använda montmorrilonit (MTM) som tillsats i XG/MTM kompositer. Orienterade ytbeläggningar av XG/MTM har tillverkats där MTM är starkt ordnat i planet. Den använda tillverkningsprocessen är kontinuerlig och vattenbaserad. Nanokompositerna har karakteriserats med FE-SEM, TEM, och XRD. Adsorption av XG på MTM har också mätts med QCM (en mikrovåg baserad på kvartskristallteknik). Filmernas barriär och mekaniska egenskaper har utvärderats och resultaten visar att användningen av MTM resulterar i en kraftig förbättring av filmernas mekaniska egenskaper och också i förbättrade barriäregenskaper. Egenskaperna bevaras också i fuktig miljö.

Artikel V behandlar en ny typ av tvärbundna XG/MTM kompositer som framställts med en hög volymsandel silikatlera (45 volymsprocent). De har framställts med en teknik som liknar papperstillverkningsprocessen vilket innebär att en mer storskalig produktion kan vara möjlig. Med hjälp av de selektivt oxiderade sidokedjorna på XG har materialen tvärbundits kemiskt till tredimensionella nätverk. Silikatlerans (MTM) orienteringsgrad har studerats med XRD och FE-SEM. Mekaniska egenskaper och barriäregenskaper har utvärderats för dessa “pärlemorliknande” material och resultaten visar att E-modulen hos de tvärbundna nanokompositerna uppgår till 30 GPa, vilket är bland de högsta värden som rapporterats för bionanokompositer.

Page 5: Xyloglucan-based polymers and nanocomposites

v

LISTOFPUBLICATIONS

Thisthesisisbasedonthefollowingmanuscripts:

I. Kochumalayil, J. J., Sehaqui, H., Zhou, Q., & Berglund, L. A. (2010). Tamarind seed xyloglucan - a thermostable high-performance biopolymer from non-food feedstock. Journal of Materials Chemistry. 20(21), 4321-4327.

II. Bergström, E. M., Salmén, L., Kochumalayil, J. J., & Berglund, L. A. (2011). Plasticised xyloglucan for improved toughness – thermal and mechanical behavior. Carbohydrate Polymers. 87(4), 2532–2537.

III. Kochumalayil, J. J., Zhou, Q., Kasai, W., and Berglund, L. A. (2012). Regioselective modification of a xyloglucan hemicellulose for high-performance biopolymer barrier films. Carbohydrate Polymers. Accepted.

IV. Kochumalayil, J. J., Bergenstråhle-Wohlert, M., Utsel, S., Wågberg, L., Zhou, Q., and Berglund, L. A. (2012). Bioinspired and highly oriented clay nanocomposites with xyloglucan biopolymer matrix – extending the range of mechanical and barrier properties. Biomacromolecules, Accepted.

V. Kochumalayil, J. J., Morimune, S., Nishino, T., Walther, A., Ikkala, O., and Berglund, L. A. (2012). Nacre-mimetic xyloglucan/clay bionanocomposites prepared from hydrocolloidal suspension – a chemical modification route for preserved performance at high humidity, Manuscript.

The author’s contribution to the appended papers is as follows:

In paper I, I have been involved in the planning and performed all experiments. I have also performed most of the data analysis and written the major part of the manuscript.

In paper II, I have been involved in planning and supervision of the master project. I have performed the material purification and demonstration experiment. I have also done part of data evaluation and the preparation of the manuscript.

In paper III, I have done the planning and performed all experiments. I have also performed most of the data analysis and written the major part of the manuscript.

In paper IV, I have done the planning and performed most of the experiments. I have also performed most of the data analysis and written the major part of the manuscript.

In paper V, I have done the planning and performed most of the experiments. I have also performed most of the data analysis and written the major part of the manuscript.

Page 6: Xyloglucan-based polymers and nanocomposites

vi

PATENTS

I. Kochumalayil, J. J., Sehaqui, H., Zhou, Q., & Berglund, L. A. (2011). Xyloglucan films. WO Patent WO/2011/040,871.

II. Kochumalayil, J. J., Zhou, Q., and Berglund, L. A. (2012). Oxygen barrier for packaging applications. International application: PCT/SE2012/050470

III. Kochumalayil, J. J., and Berglund, L. A. (2012). Regioselective modification of xyloglucan hemicellulose for high-performance biopolymer films. US provisional: 61715643

Otherrelevantpublicationsnotincludedinthethesis:

I. Marais A., Kochumalayil J. J., Nilsson C., Fogelström L., Gamstedt E. K. (2012). Toward an alternative compatibilizer for PLA/cellulose composites: grafting of xyloglucan with PLA. Carbohydrate Polymers, 89(4),1038–1043.

II. Sehaqui, H., Kochumalayil,J. J., Liu, A., Zimmermann, T., and Berglund, L. A. (2012). Highly ductile and flexible nacre-mimetic structured clay composites via paper-making process: mechanical, thermomechanical and barrier properties. Manuscript.

Page 7: Xyloglucan-based polymers and nanocomposites

vii

TABLEOFCONTENTS

1. INTRODUCTION.......................................................................................................... 1

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

1.2. Objective ....................................................................................................................... 3

1.3. Xyloglucan .................................................................................................................... 3

1.3.1. Xyloglucan structure and solution properties ........................................................ 4

1.3.2. Xyloglucan modifications ...................................................................................... 7

1.4. Oxygen barrier films in packaging ............................................................................... 7

1.4.1. Oxygen permeability of polymers .......................................................................... 8

1.4.2. New barrier films and challenges .......................................................................... 9

1.5. Clay nanocomposites .................................................................................................... 9

1.5.1. Montmorillonite structure .................................................................................... 10

1.5.2. Bionanocomposites .............................................................................................. 11

1.5.3. Recent advances (nacre mimetic nanocomposites) .............................................. 11

2. GENERAL MATERIALS AND METHODS ........................................................... 13

2.1. Xyloglucan purification .............................................................................................. 13

2.2. Modification of xyloglucan ........................................................................................ 13

2.2.1. Chemical modification of XG (Paper III & V) .................................................... 13

2.2.2. Enzymatic modification of XG (Paper I & IV) .................................................... 15

2.3. Film Preparation (Paper I, II & III) ............................................................................. 15

2.4. Nanocomposites preparation (Paper IV& V) .............................................................. 15

2.4.1. MTM exfoliation .................................................................................................. 15

2.4.2. Nanocomposite preparation ................................................................................. 16

2.5. Quartz crystal microbalance analysis (Paper IV) ....................................................... 18

2.6. Characterization methods ........................................................................................... 19

3. RESULTS AND DISCUSSION .................................................................................. 25

3.1. Xyloglucan film properties ......................................................................................... 25

3.1.1. Film-formation ..................................................................................................... 25

3.1.2. Hygrothermal properties ...................................................................................... 26

3.1.3. Hygromechanical properties ................................................................................ 27

3.1.4. Oxygen permeability ............................................................................................ 29

3.2. Effect of plasticization ................................................................................................ 30

3.2.1 Thermomechanical properties ............................................................................... 30

Page 8: Xyloglucan-based polymers and nanocomposites

viii

3.2.2 Tensile properties .................................................................................................. 31

3.3. Effect of chemical modification ................................................................................. 32

3.3.1. Periodate oxidation and subsequent reduction ..................................................... 32

3.3.2. Properties of modified XG films .......................................................................... 37

3.4. Xyloglucan/MTM bionanocomposites ....................................................................... 39

3.4.1. Adsorption of XG to MTM .................................................................................. 40

3.4.2. Nanocomposite film characterization .................................................................. 41

3.4.2. Film properties ..................................................................................................... 42

3.5. Nacre-mimetic nanocomposites based on xyloglucan and clay ................................. 45

3.5.1. Composite preparation ......................................................................................... 45

3.5.2. Nanocomposite film characterization .................................................................. 46

3.5.3. Film properties ..................................................................................................... 49

4. CONCLUSIONS .......................................................................................................... 53

5. ACKNOWLEDGEMENTS ........................................................................................ 55

6. REFERENCES ............................................................................................................. 57

Page 9: Xyloglucan-based polymers and nanocomposites

ix

ABBREVIATIONS

AFM

AcGGM

DMTA

DSP

DVS

FE-SEM

FTIR

Galp

Glcp

MALDI-TOF-MS

MTM

Mw

m/Z

NFC

OTR

OPET

PDI

PVOH

QCM

RH

TEM

Tg

TGA

XG

XGO

XRD

Xylp

Atomic Force Microscopy

O-Acetyl Galactoglucomannan

Dynamic Mechanical Thermal Analysis

Digital Speckle Photography

Dynamic Vapour Sorption

Field Emission – Scanning Electron Microscope

Fourier Transform Infrared Spectroscopy

Galactopyranose

Glucopyranose

Matrix-assisted Laser Desorption/Ionization time-of-flight mass spectrometry

Montmorillonite

Weight average molecular weight

Mass / charge ratio

Nanofibrillated Cellulose

Oxygen Transmission Rate

Oriented polyester

Polydispersity index

Polyvinyl alcohol

Quartz Crystal Microbalance

Relative Humidity

Transmission Electron Microscope

Glass transition temperature

Thermogravimetric Analysis

Xyloglucan

Xyloglucan Oligosaccharide

X-Ray Diffraction

Xylopyranose

Page 10: Xyloglucan-based polymers and nanocomposites

x

Page 11: Xyloglucan-based polymers and nanocomposites

1

1.INTRODUCTION

1.1.Introduction

Utilization of biopolymers from renewable resources in order to replace petroleum-

derived polymers is of increasing interest both in industry and in the scientific research

community. The packaging industry is a major consumer of the global plastic production,

where food packaging is a particularly important area of application. In packaging

applications, polymer materials provide flexibility, processability, optical transparency

and gas barrier performance. Several polysaccharide-based biopolymers such as starch,

cellulose, chitin etc. are investigated as possible alternatives to conventional polymers in

packaging applications.1, 2 Starch has high potential for packaging because of low cost,

renewability, biodegradability and processability in conventional thermoplastics

processing equipment. Despite of these advantages, more wide-spread applications of

starch-derived polymers have not been realized. The most important limitations are that

the material is very sensitive to moisture and that the mechanical and gas barrier

properties are much lower compared to many petroleum-derived polymers used in

packaging. More recently, nanofibrillated cellulose (NFC) films with high cellulose purity

are suggested for packaging applications.3-5 NFC films show very high modulus and

strength.6 Thin films of NFC have demonstrated both optical transparency and low

oxygen permeability.4 However, the commercial prospects are marred by the sharp

decline in mechanical and barrier performance of unmodified NFC under humid

conditions. Furthermore, the cost of commercially available NFC is still high, and

industrial production methods need to be developed. Recently, hemicelluloses have

become a biopolymer of commercial interest, especially wood hemicelluloses that have

been extensively studied as oxygen barrier films.7-9 However, the commercial prospects of

hemicelluloses are still uncertain owing to limitations in film-forming characteristics and

challenges in extraction procedures from plant biomass. In the present thesis, xyloglucan

Page 12: Xyloglucan-based polymers and nanocomposites

2

(XG), a hemicellulose derived from an industrial waste product – tamarind seed – is

studied as an engineering polymer of interest for packaging applications.

Moisture durability is a critical parameter in food packaging applications. This can be a

hindrance for polysaccharides, including XG, which show significantly lowered

mechanical and barrier properties at high humidity.10 Chemical modification, including

esterification and several grafting approaches were tried in order to address the problem.11

However, critical mechanical and barrier properties were sacrificed.11 To this end,

nanocomposites based on polysaccharides and layered silicates were also studied, but

with limited success.12, 13 The lack of success is related to poor dispersion and

uncontrolled platelet orientation in the matrix polymer. Thus, the problem of moisture

sensitivity to mechanical and barrier performance still needs to be addressed in order to

successfully use polysaccharides in packaging applications. In the present study, moisture

absorption in XG materials and its effects on physical properties are investigated.

Another important parameter is processability. Thermal processing of thermoplastics is

the most widely used technique in packaging industry. However, most polysaccharides

degrade before reaching the glass transition temperature (Tg). The reduction of Tg to a

sufficiently low temperature is one possibility. The other potential strategies are “paper-

making” procedures and new coating methods. These approaches are explored in the

present study.

Page 13: Xyloglucan-based polymers and nanocomposites

3

1.2.Objective

The engineering objective of this work was to investigate the potential of xyloglucan

(XG) as an engineering polymer in packaging applications. To this end, XG was modified

or combined with layered silicates, and primarily evaluated with respect to mechanical

properties and barrier performance, also under humid conditions. The scientific objective

was to study the effect of modification or layered silicate addition on material structures

as well as to analyse structure-property relationships in the resulting materials. The

following projects were carried out:

Basic hygromechanical and thermomechanical properties of tamarind seed XG

films were analysed (Paper I).

XG was plasticized in order to reduce Tg and facilitate thermal processing; the

thermomechanical and tensile properties were analysed (Paper II).

XG was subjected to regioselective chemical modification of the side chains in

order to reduce Tg, but maintain the cellulosic nature of the backbone (Paper III).

XG was combined with layered silicate reinforcement of up to 12 % by volume;

the mechanical and barrier properties were evaluated, also in moist condition

(Paper IV).

Cross-linked nanocomposite films of up to 45 % by volume of layered silicates

were prepared by a “paper-making” method; the mechanical and barrier properties

were evaluated (Paper V).

1.3.Xyloglucan

Xyloglucan (XG) is a major hemicellulose component present in the primary cell wall

of plants, where it links the cellulose microfibrils and thus contributes to the load-bearing

function of the cell wall.14, 15 Xyloglucan makes up to 20 – 30 % of the dry weight of the

primary cell wall.16 However, the primary cell wall is a thin component in wood tissue

and extraction of XG for commercial exploitation is unfeasible. On the other hand, the

seeds of a few plants such as tamarind (tamarindus indica L.), detarium senegalense,

afzelia africana, and Jatoba have abundant deposits of XG.17 The only seed XG that is

commercially exploited at present is of the tamarind tree. It grows in more than 50

Page 14: Xyloglucan-based polymers and nanocomposites

4

countries in the tropics and subtropics, and produces brown pod-like fruits, which contain

fruit pulp and many hard-coated seeds (Figure 1.1).18, 19 The decorticated seeds contain a

large proportion (ca. 60 %) of XG.20 Compared to wood-based hemicelluloses, the

extraction of XG from seed powder is straightforward, also in a commercial perspective.

Methods such as air separation of pulverized tamarind seed powder using a centrifugal

vortex air classifier21, 22 or alcohol precipitation from aqueous solution of the

polysaccharide are available.23, 24

Figure 1.1. Tamarind fruit and seeds from tamarind tree (tamarindus indica)

The tamarind seed is a by-product of the tamarind industry and the fruit pulp is the chief

souring agent for curries, sauces and certain beverages.18 The decorticated flour, known as

tamarind kernel powder (TKP), is a major industrial product widely used as a sizing

material in textile, paper, and jute industries.16, 18 Other high-end applications of XG

derived from tamarind kernel powder are recently reviewed.19 However, there were no

publications on XG being studied as an “engineering” biopolymer prior to Paper I. Even

in highly comprehensive reviews on XG applications, no comment is made about the

potential applications of XG as a biopolymer for thermoplastic mouldings, packaging

materials, or as a biocomposite matrix.18, 19

1.3.1.Xyloglucanstructureandsolutionproperties

Xyloglucans, or more generally galactoxyloglucans, are polysaccharides with the main

chain identical to that of cellulose, i.e. β (1→4) linked D-glucan. Xyloglucan is the only

Page 15: Xyloglucan-based polymers and nanocomposites

5

known hemicellulose with the main-chain structure as cellulose (Figure 1.2.). In tamarind

seed XG, up to 75 % of the glucose residues are substituted with α (1→6) linked xylose

side residues, where some of these xyloses have an additional β (1→2) linked galactose

residue.17, 25, 26 The XG present in wood cell wall has an additional sugar moiety –

fucose.27 Tamarind seed XG has a high molecular weight with reported weight average

molecular weight (Mw) values in the range 650,000 to 2,500,000.28-30 This is a marked

difference compared with wood-derived hemicelluloses, which have low molecular

weights in the order of 10,000 – 50,000 Da after extraction processes.7

The basic repeat unit of tamarind seed XG can be described in terms of four

oligosaccharides, which differ in the number and distribution of galactose residues. They

are conveniently represented as XXXG, XLXG, XXLG, and XLLG and are reported to be

in the molar ratio of 1:0.42:2.08:6.20.17, 31 X represents a Xylp(α1→6)-Glcp unit, L

represents a Galp(β1→2)Xylp(α1→6)Glcp unit, and G represents a Glcp residue. When

written sequentially, a β(1→4) linkage between the Glcp residues is implied, with the

reducing end on the right.31

Xyloglucan is water-soluble, but the individual chains are not fully hydrated and remain

aggregated even in very dilute solutions. Lateral interaction of XG chains is widely

documented17, but glucose – glucose ring stacking interactions can also be expected as

has been shown in the context of studies performed on XG/cellulose network in the

primary plant cell wall.32 The side chain sugars, galactose and xylose, have an important

role in promoting solubility and counteracting gelation of XG in water.17 The aqueous

solution of XG shows unique properties such as Newtonian fluidity, thermostability, and

stability towards acid hydrolysis.16 The molar mass of XG is also highly stable in

mechanical force-fields.16 The details of the rheological characteristics have been

comprehensively reviewed.16, 18

Page 16: Xyloglucan-based polymers and nanocomposites

6

Figure 1.2. “Average” chemical structure of xyloglucan in comparison with typical hemicelluloses and starch macromolecules. (A) Hardwood glucuronoxylan, (B) Softwood galactoglucomannan, (C) Starch macromolecules – amylose and amylopectin, and (D) Xyloglucan.

Page 17: Xyloglucan-based polymers and nanocomposites

7

1.3.2.Xyloglucanmodifications

The galactose residues in side chains have a significant role in the solution properties of

XG. Specific enzymes such as β – galactosidase, have been used to remove the galactose

and the solution properties of the resulting modified XG were studied.16, 33, 34 In vitro

analysis of the interaction between modified XG and cellulose was also reported.35

Enzymatic oxidation of galactose residues was employed to prepare carboxylated

derivatives of XG.36 The objective was to create individual charged XG macromolecules

for structural elucidation by light scattering.

A range of carboxylated, sulphated, and alkylaminated derivatives of XG were prepared

by oxidation of terminal side chain galactose residues with the use of galactosidase

enzymes and subsequent chemical derivatization steps.37

Chemical modification procedures for XG are scarce. In contrast, many derivatization

protocols are available for starch and cellulose.38-41 A well-known product is cationic

starch, widely used in the paper industry as a wet-end additive.42

1.4.Oxygenbarrierfilmsinpackaging

Transmission of oxygen through food packages is the main cause for food deterioration

and has a strong effect on the shelf-life of such products.43 Oxygen barrier performance is

therefore a critical parameter in the design of food packages. The traditional barrier layer

has been aluminium,44 but non-renewability is one concern for its continued use. Over the

past decades, synthetic polymers such as polyvinyl alcohol (PVOH), ethylene vinyl

alcohol copolymer (EVOH), and polyvinylidene chloride (PVDC) have successfully

replaced part of aluminium-based packaging applications.45 Although packaging

technology has gone through many development stages, virtually all oxygen barrier

polymers in commercial use today were present already in the late 1970s.45, 46 However,

due to depleting oil reserves, the continued use of petroleum derived plastics in packaging

is challenged. Moreover, new legislations and increased social awareness for sustainable

development gathers momentum in favour of bio-based materials in packaging.

Page 18: Xyloglucan-based polymers and nanocomposites

8

1.4.1.Oxygenpermeabilityofpolymers

The permeability of polymer films is defined as:47

P = DS 1.1

P = Permeability coefficient

D = Diffusion coefficient

S = Solubility coefficient (the amount of transferring molecules retained or dissolved in

the film at equilibrium conditions)

D is the kinetic term and S is the thermodynamic term incorporated during gas

transmission.

Permeability at steady state with D and S constant becomes:47

P= (dM dt)steady state L⁄

A∆p 1.2

where M is the quantity of permeant (in mass or volume), t is time, L is thickness of the

film, A is cross-sectional area of the film, Δp is the partial pressure difference across the

film, and P is the permeability coefficient.

Note that P is expressed as a function of film thickness. In contrast, transmission rate is

independent of film thickness. The most widely used unit for OTR is cc[m2-day]atm and

unit for permeability is cc.µm/[m2-day]kPa.

Oxygen transmission through a membrane film depends on many factors such as

material structure, temperature, and humidity. An aluminium foil is, theoretically, a total

barrier to oxygen; however O2 molecules in reality can permeate since aluminium foil

may contain pinholes. A polymer membrane is permeable even if there are no visible

pinholes. The mechanism of gas diffusion through polymer membrane is usually

explained with models based on either free volume48 or energetics49. Polar polymers such

as polyols and polysaccharides have good chain packing (high energy density). This gives

rise to favourable properties (low free volume, relatively high energy required to separate

polymer chains) for high oxygen barrier performance. Although gas diffusion below Tg is

not well understood,50 an important aspect to consider is the low mobility of the polymer

chains. In the time scale of gas diffusion, the redistribution of free volume for the passage

Page 19: Xyloglucan-based polymers and nanocomposites

9

of O2 molecules in polymer membrane is rather restricted in the glassy state. At high RH,

water molecules might be able to disrupt the polymer chain packing, thereby lowering the

activation energies for oxygen diffusional jumps. At the same time, water is much more

mobile, which makes diffusion of oxygen in the water phase much easier from a free

volume point of view.

1.4.2.Newbarrierfilmsandchallenges

New bio-based materials have recently been explored for the development of barrier

films.1, 47, 51 Wood hemicelluloses have also been studied extensively as oxygen barrier

films.7, 9, 52

The oxygen gas barrier performance of highly polar polymers such as PVOH or

hemicelluloses is reduced when exposed to high relative humidity. This is a significant

hindrance in food packaging applications. Another problem is the complex extraction

processes used for wood hemicelluloses. Most hemicelluloses are extracted under alkaline

conditions from wood tissue. A challenge is to preserve high molar mass in the final

product.

Initial lab trials showed that XG in the dry state has good oxygen barrier performance,

which is comparable or even superior to commercial barrier polymer films. However,

oxygen permeability of XG increases at high relative humidity. The following sections

will discuss further the strategies adopted in the present work to solve this problem.

1.5.Claynanocomposites

The development of MTM-based (montmorillonite) nanocomposites by Toyota

researchers was a ground-breaking achievement in the field of polymer composites with

good mechanical properties, and has been thoroughly reviewed.53 In the first study, MTM

was combined with polyamide-6 by polymerizing ϵ-caprolactam in the galleries of

MTM.54 Discrete MTM platelets of only 1 nm thickness and high aspect ratio provided

extremely high reinforcement efficiency at low concentrations (1 - 3 wt%). Since then,

the field of clay nanocomposites has grown dramatically and has inspired the

development of new classes of hybrid materials.55-57 Later studies showed that, in addition

Page 20: Xyloglucan-based polymers and nanocomposites

10

to reinforcement effects on mechanical properties, oriented platelet structures in polymer

matrices can function as barriers to gas and moisture.58, 59

1.5.1.Montmorillonitestructure

Natural montmorillonite (MTM) is a 2:1 phyllosilicate mineral.60 Silicate layer

thickness is 1 nm and lateral dimensions can vary from few nm to several µm. Figure 1.3

(A) shows two neighbouring montmorillonite layers (according to the widely accepted

Hofmann-Endell-Wilm structure61), truncated in X-Y plane.

Figure 1.3. (A) Idealized structure of 2:1 layered silicate with one dioctahedral sheet sandwiched between two tetrahedral sheets. Si4+ is found in tetrahedral coordination to 4 O2- in the basal tetrahedral sheets. Al3+ is positioned in the central dioctahedral sheet in octahedral coordination to 4 O (shared with silicate sheet) and 2 OH groups.62 (B) Molecular structure of a single layer of montmorillonite.

Isomorphous substitution of Si4+ and Al3+, primarily by Al3+ and Mg2+ or Fe2+,

respectively, results in a net negative charge of the montmorillonite layer surfaces.63 This

excess charge is balanced by exchangeable cations, usually Li+, Na+, K+, Ca2+, and Mg2+,

in the interlayer voids. A representative structural formula of Na-montmorillonite is

(technical report TR-06-30 of SKB, Swedish Nuclear Fuel and Waste Management

Company, at www.skb.se)62:

(Al3.11Fe0.37Ti0.01Mg0.48)(Si7.92Al0.08)O20(OH)4Na0.65

Page 21: Xyloglucan-based polymers and nanocomposites

11

Natural MTM can form a stable colloidal suspension in water by exfoliation of the

layered structure. The efficiency of the process depends on the counterions and details of

the exfoliation process. In general, the swelling decreases depending on the counterions

present: Li+ > Na+ > Ca2+ > Fe2+ > K+. In the present study, Na-MTM was used.

1.5.2.Bionanocomposites

Most clay nanocomposite preparations were based on petroleum-derived polymers and

monomers. Recently, biopolymer matrices from renewable resources have been used to

prepare MTM bionanocomposites.13 This includes the use of water-processed starch,

chitosan and pectin biopolymer matrices.12 However, the results in terms of

nanocomposite properties are disappointing. The reason is poor dispersion of MTM, lack

of nanostructural order and an MTM content of typically only 5 wt% or lower. The

preparation of structures with well-dispersed inorganic nanoparticles is a challenge for

polymer nanocomposites in general.64

1.5.3.Recentadvances(nacre‐mimeticnanocomposites)

The difficulties encountered with dispersion and nanostructural control in man-made

nanocomposites have motivated attempts to prepare bio-inspired composites. In biological

structures, most load-bearing materials are indeed nanostructured composites. Several

contain inorganic particles and examples such as bone, dentin and nacre have

sophisticated nanoscale organization.65 Even in highly hydrated environment, they have

substantial mechanical function. The work by Tang et al.66 on artificial nacre was thus a

seminal study in that the layered and parallel orientation of inorganic platelets in nacre

was mimicked and it was possible to reach high inorganic content (Vf ≈ 50 %), a modulus

of 11 GPa and a tensile strength of 100 MPa. MTM clay platelets were combined with a

charged, water-soluble polymer in an elaborate layer-by-layer (LbL) deposition process.

The clay platelets overlap each other and are glued together by the deposited polymer, see

Figure 1.4 (A). This process results in in-plane orientation and provides nanostructural

control of the platelet distribution, but the entire process is time-consuming and most

likely difficult to extend from laboratory practice to large scale industrial processing.

Page 22: Xyloglucan-based polymers and nanocomposites

12

Walther et. al. showed that nacre-mimetic oriented composites can also be produced

using a simple and industrially scalable water-based processing approach akin to paper-

making.67, 68 MTM in hydrocolloid form was combined with a water-soluble

polyelectrolyte or polyvinyl alcohol (PVOH). The dissolved polymer adsorbed to

exfoliated clay platelets. Excess polymer was removed and the polymer-coated platelets

were subjected to filtration. An oriented nanocomposite structure of nacre-like “brick-

and-mortar” organization and high inorganic content was achieved. The most important

difference compared to nanocomposites prepared by LbL process is that the clay platelets

are individually separated by the polymer coating (Figure 1.4 (B)). In a subsequent study,

Yao et al. applied the same procedure, but used chitosan as biopolymer matrix.69 Liu et al.

combined clay with nanofibrillated cellulose (NFC) and prepared clay nanopaper.70, 71 All

these material designs are classified as ‘nacre-mimetic’ owing to the high content of

inorganic phase.

Figure 1.4. Schematics of (A) Nacre mimetic structures prepared by layer-by-layer (LbL) assembly of water soluble polymer solution with an MTM water suspension. Platelet aggregation is visible (B) MTM platelets pre-coated with water soluble polymer fabricated in a “paper making” method. The individual platelets are separated by the polymer coatings.

Since polymers used so far for fabrication of nacre-mimetic composites are water

soluble and hydrophilic, the properties depend on moisture content. The polymer-polymer

interaction is also important in determining the properties under humid conditions.

Podsiadlo et. al. used gluteraldehyde to cross-link polyvinyl alcohol/MTM systems to

improve substantially the mechanical strength and stiffness.72 In addition, the cross-

linking preserved the mechanical properties in humid conditions.

Page 23: Xyloglucan-based polymers and nanocomposites

13

2.GENERALMATERIALSANDMETHODS

2.1.Xyloglucanpurification

The industrially available xyloglucan (Innovassynth technologies Ltd., India) contains

nearly 10 % of impurities, including a fraction of proteinous matter. The weight average

molecular mass was 1.5 - 2.5 MDa determined by SEC using Pullulan standard. About 0.5

wt% of XG solution was prepared in distilled water by mechanically stirring at 60 °C for

1 h. Most part of the water insoluble protein fraction was removed by centrifugation at

4500 rpm for 45 min., which was followed by filtration with a Büchner funnel using glass

microfiber filters (Whatman GF/A, pore size, 1.6 μm). The filtrate solution was freeze-

dried to obtain purified XG for further experiments.

2.2.Modificationofxyloglucan

2.2.1.ChemicalmodificationofXG(PaperIII&V)

The chemical modification uses periodate oxidation of vicinal hydroxyl groups present

on XG to form dialdehyde products with ring opening and subsequent reduction to

dialcohols. The general modification steps are schematically presented in Scheme 2.1.

with the oxidation and reduction reactions of a representative sugar moiety. Periodate

oxidation of cellulose and other polysaccharides are well known in organic chemistry and

the oxidation reaction is commercially utilized to prepare dialdehyde starch.73, 74 The

oxidation and reduction steps for XG were carried out in accordance with an earlier

protocol reported for cellulose.75

2 g of purified XG was dissolved in 100 ml water by heating at 60 ˚C for 1h with

magnetic stirring. 3.6 g of sodium meta-periodate (Sigma-Aldrich) (an equimolar amount

of periodate for complete oxidation of carbohydrate rings present in XG) in 35 ml water

was added to three different glass beakers containing the XG solution. The reaction was

continued for 30 min., 1 h and 2 h for different solutions in dark with magnetic stirring at

Page 24: Xyloglucan-based polymers and nanocomposites

14

room temperature. Gelation was observed even after 20 min. of reaction. The reactions

were terminated by adding 2.5 ml of ethylene glycol to decompose the excess periodate.

The solution was transferred to 1.2 l of methanol in a beaker to precipitate the dialdehyde

XG. The precipitate was washed with methanol/water and vacuum dried overnight. The

dry precipitate was ground and transferred to a beaker containing 80 ml of water.

Scheme 2.1. Periodate oxidation and subsequent reduction of pyranose rings with vicinal diol groups

4 g of sodium borohydride (Sigma Aldrich) was dissolved in 20 ml water and added

drop wise to the beaker containing dialdehyde XG solution with magnetic stirring in an

ice bath. The precipitate was completely dissolved after 1 h of stirring. The excess sodium

borohydride was neutralized with 30 % acetic acid. The solution was transferred to 1.4 l

ice-cold methanol in a beaker. The precipitate was washed and dried under vacuum

overnight. To purify further, the dried mass was dissolved in 50 ml distilled water and

dialyzed in running water for three days. It was then precipitated again in methanol and

dried under vacuum.

In Paper V, periodate oxidation was carried out in a 0.5 wt% XG solution: 10 , 20 and

30 % of the stoichiometric amount of sodium meta-periodate (Sigma Aldrich) needed for

complete oxidation of carbohydrate rings present in XG was added to 200 ml of 0.5 wt%

XG solution. The reaction was allowed to continue for 13 hours in dark conditions under

magnetic stirring at room temperature. The solutions were dialyzed under running water

for three days. These dialdehyde XGs are designated as ddXG10, ddXG20 and ddXG30,

corresponding to the amount of periodate added.

Page 25: Xyloglucan-based polymers and nanocomposites

15

2.2.2.EnzymaticmodificationofXG(PaperI&IV)

In order to study the effect of galactose residues on material properties of XG, the

galactose residues in the side chains of XG were partially removed by using β-

galactosidase (from Aspergillus oryzae, Sigma-Aldrich).34 2 g of freeze-dried XG was

dissolved in 80 ml of 50 mM sodium acetate buffer (pH = 4.5). 50 mg of β-galactosidase

enzyme dissolved in 5 ml of MilliQ water was added to the XG solution. The reaction

mixtures were kept at 30 °C and the reactions were terminated after specified times by

heating at 90 °C for 30 min. to deactivate the enzyme. The partially degalactosed XGs

(modified XGs) were purified by precipitation in ethanol and dried in a vacuum oven at

50 °C for 24 h. The dried mass was re-dissolved in 100 ml water and filtered through a

Büchner funnel using glass microfiber filters (Whatman GF/A, pore size, 1.6 μm). The

filtrate solution was freeze-dried to obtain dried XG for further experiments. The amount

of galactose removed in 100 mg of material was estimated by carbohydrate analysis.

2.3.FilmPreparation(PaperI,II&III)

A general method for the film casting of XG: 1g of purified XG was dissolved in 80 ml

deionized water by heating at 60 °C for 60 min. with magnetic stirring. The solution was

further mixed using an Ultra-Turrax® blender (IKA, DI25 Basic), for 15 min. at 9500

rpm. The air bubbles in the solution were removed by keeping in a desiccator under

vacuum for 2 h. The solution was then spread over a Teflon-coated Petri dish and placed

in oven at 35 °C. The dried films were peeled-off, and conditioned at desired humidity.

2.4.Nanocompositespreparation(PaperIV&V)

2.4.1.MTMexfoliation

An important step in the nanocomposites preparation is the exfoliation of MTM

platelets in water suspension. A general method used for the exfoliation in the present

study was: 1 wt% MTM solution was prepared by mixing MTM powder (Cloisite Na+,

density of 2.86 g / cc, Southern Clay Products, Inc.) in deionized water using Ultra-

Turrax® blender (IKA, DI25 Basic) at 25000 rpm for 15 min. followed by sonication

Page 26: Xyloglucan-based polymers and nanocomposites

16

using Vibra-Cell (Sonics & Materials, Inc.) ultrasonic processor. This procedure was

repeated several times. The resultant MTM suspension was kept undisturbed for one week

and any clay aggregates were then removed. The AFM height image with a typical MTM

platelet dimension is presented in Figure 2.1. The thickness of particles was in the range

of a typical single layer of MTM, ca. 1 nm.

Figure 2.1. Atomic force microscope image (Height image) of a MTM suspension coated over a silica surface (scanned area = 2 µ X 2 µ). The dimension of a typical MTM platelet marked in (A) is presented in (B).

2.4.2.Nanocompositepreparation

0.5 wt% XG solution prepared from purified XG and 1 wt% exfoliated MTM

suspension were used for nanocomposite preparation. The MTM concentrations in XG-

MTM suspensions were 1.0, 2.5, 5.0, 10.0, and 20.0 % (wt / wt). The suspensions were

mixed with Ultra-Turrax® at 13500 rpm for 15 min. and kept under magnetic stirring

overnight. The resulting solutions were centrifuged at 4000 rpm for 20 min. to remove

micro bubbles and any clay aggregates. The final solid contents of different

nanocomposite solutions were in the range of 4 - 5 % after rotoevaporation to desired

viscosity. The resulting solutions were evenly spread over a Teflon mould and dried under

constrained condition in an oven at 40 °C overnight. The films were peeled off from the

Teflon surface for further characterization. The thickness of the films was in the range of

20 - 30 µm.

Page 27: Xyloglucan-based polymers and nanocomposites

17

Different nanocomposite suspensions were coated on an oriented polyester film (OPET)

in a comma coater (Hirano Tecseed Co., Ltd., Japan), where the OPET film was rolling at

a speed of 0.5 meter / min (see Scheme 2.2). The film with wet coating was immediately

dried in a heating chamber at 120 °C. The thicknesses of the wet coatings were adjusted

in such a way that the final thickness of the dried films was 4 µm.

Scheme 2.2. Schematic representation of xyloglucan/MTM nanocomposite preparation and film formation

In Paper V, a different approach was adopted: 100 ml each of 0.5 wt% XG / oxidized

XG solution was thoroughly mixed with 100 ml 0.2 wt% exfoliated MTM suspension.

The excess polymer was removed by centrifugation at 4000 rpm for 90 min. The resulting

precipitates were washed with water and centrifuged further at 4000 rpm for 45 min. The

precipitates were re-dispersed in 100 ml water and magnetically stirred overnight. The

resulting suspensions were filtered through a vacuum filter set up using a 0.65 µm PTFE

filter paper to form self-assembled XG / oxidized XG-MTM composite films (Scheme

2.2). The films were dried at 92 °C for half-an-hour under vacuum. A covalent cross-link

of acetaldehyde is expected to be formed between the aldehyde groups on oxidized XG

Page 28: Xyloglucan-based polymers and nanocomposites

18

and hydroxyl groups on XG chains or MTM surface.72 76, 77 A possible mechanism for the

cross-linking reaction is schematically presented in Scheme 2.3.

Scheme 2.3. Schematic representation of a possible cross-linking reaction of dialdehyde XG and hydroxyl groups on XG chains or MTM surface

2.5.Quartzcrystalmicrobalanceanalysis(PaperIV)

A Quartz crystal microbalance (QCM) E4 from Q-Sense AB (Västra Frölunda, Sweden)

was used to study the adsorption of xyloglucan on clay model surfaces with a continuous

flow of 100 µl/min.78 Native XG, enzymatically modified XG (40 % galactose removed)

and a synthetic polymer PVOH were studied. All samples had a concentration of 100

mg/L. The QCM crystals used were AT-cut quartz crystals (5 MHz resonant frequency)

with an active surface of sputtered silica which were plasma treated using an air plasma

cleaner (Model PDC 002, Harrick Scientific Corporation, NY, USA) under reduced air

pressure for 120 s at high effect (30 W). The samples were adsorbed on model surface

and the change in frequency in quartz crystal was recorded.

To convert a change in frequency into an adsorbed mass per area unit, the Sauerbrey

model was used:79

m CΔ

2.1

where, m = Adsorbed mass per unit area [mg/m2], C = Sensitivity constant, -0.177

[mg/(m2·Hz)], Δf = Change in resonant frequency [Hz] and n = Overtone number

Page 29: Xyloglucan-based polymers and nanocomposites

19

Clay model surface preparation: Clay model surfaces were prepared in-situ by

consecutively adsorbing three bilayers of Poly (diallyl Dimethyl Ammonium Chloride)

(pDADMAC) and exfoliated MTM using the LbL-technique. Both samples had a

concentration of 100 mg/L.

2.6.Characterizationmethods

2.6.1. Carbohydrate analysis: This was to determine the sugar composition in XG and

modified XG samples. After acid hydrolysis to individual sugar molecules using 70 %

sulphuric acid80, the hydrolyzates were analysed using high performance anion exchange

chromatography equipped with Pulsed Amperometric Detector (HPAEC-PAD, Dionex

ICS-3000). Standard solutions of glucose, xylose and galactose were used for calibration.

2.6.2. MALDI- TOF.-MS: Matrix-assisted laser desorption/ionization time-of-flight mass

spectrometry (MALDI-TOF-MS) was performed on Ultraflex MALDI-TOF workstation

(Bruker Daltonics, Bremen, Germany) equipped with a nitrogen laser (337 nm) and

operated in positive reflector mode. XG has very high molecular weight, which makes it

difficult to be analysed in MALDI. Thus it was enzymatically hydrolyzed to

oligosaccharides.

Enzymatic hydrolysis of XG: This was performed using cellulase enzyme from

Trichoderma reesei (EC 3.2.1.4) (Sigma-Aldrich). About 80 mg of XG and modified XG

samples were dissolved in 8 ml of sodium acetate buffer (pH = 5.0) in glass vials. 8 mg of

enzyme was added to different solutions and stirred for 24 h at 37 °C. The enzyme

reaction was terminated by increasing the temperature to 90 °C and stirring for half an

hour. The resulting solutions were filtered through glass fibre filter paper (Whatman

GF/A) and the filtrate was further passed through 0.45 µ PTFE syringe filters. The final

filtrate was freeze dried.

About 2 mg of freeze-dried oligosaccharides obtained from enzyme hydrolysis were

dissolved in 1 ml of MilliQ water. 3 µL of sample solution was mixed with 6 µL of

matrix (5 mg of 2, 5-dihydroxy benzoic acid in 0.5 ml acetone) and spread over a

MALDI-TOF plate (Bruker Daltonics) and air-dried. The ions were accelerated with a

laser power of 21 kV. The mass/charge peak (m/Z peak) corresponding to the molecular

Page 30: Xyloglucan-based polymers and nanocomposites

20

weight of oligosaccharides were noticed. Note that the m/Z value obtained from MALDI-

TOF-MS includes the mass for the counterion present in matrix mixture, Na+.

2.6.3. Mechanical testing: The mechanical properties of conditioned films were measured

using an Instron tensile testing machine in tensile mode with a 50 N or 100 N load cell

(Paper I-III). The specimens were thin rectangular strips (60 X 5 mm2) and the gauge

length was 40 mm. The thickness of the film strips were accurately measured using a

thickness meter and were typically 20 - 30 µm. The stress–strain curves of specimen

samples were recorded at room temperature and 50 % RH at a strain rate of 10 % min-1.

At least six specimens were tested from each sample and results were reported for those

specimens that do not show premature failure at the jaw face. Young’s modulus (E) was

determined from the slope of the low strain region in the vicinity of 0.05 % strain. Yield

stress was calculated as stress value corresponding to the intersection of the tangent

between the initial elastic region and the following plastic region.

Tensile testing of nanocomposites (Paper IV-V) were performed on a Deben microtester

with a load cell of 200 N in order to make a comparison with other ‘nacre-mimetic’

materials67, 68 and due to the limited amount of material. The films were cut in rectangular

strips of dimensions 5 mm in width and 20 mm in length. The gauge length was 10 mm

and the extension rate was 0.5 mm / min. The samples were conditioned for one week at

50 % RH and 90 % RH prior to testing.

In paper V, digital speckle photography (DSP) was used to determine strain. A non-

contact technique was applied to retrieve the strain fields. These were monitored using a

commercial 2D Digital Image Correlation System Vic 2D (Limess Messtechnik and

Software GmbH). A CCD camera with resolution of 5.0 mega pixels was used. The

camera was aligned perpendicular to the surface of the samples. One picture per

millisecond of the whole specimen’s area was recorded during the tensile loading process

of each sample. Post-processing of the images was done with Vic2D software. The area of

the sample that was subjected to pure tensile strain was masked with so called facets of 35

X 35 pixels. Each image was converted to grey scale, which was identified, in order to

compute displacements during loading with a digital image correlation algorithm. A

rectangle of about 5 mm X 3 mm in the centre of the sample was masked. The measured

Page 31: Xyloglucan-based polymers and nanocomposites

21

strain data points from the DSP system consist of the mean strain in a region

corresponding to a square with sides approximately the double spatial distance between

two data points.

The volume fractions of the inorganic content in composite films were determined from

weight fraction of inorganic phase obtained from thermogravimetric analysis. The MTM

volume fraction was calculated as:

2.2

where, = Volume fraction of clay, = density of XG (1.5 g/cm3), = density of

clay (2.86 g/cm3), and are weight fraction of MTM and XG respectively. It is

assumed that modified XGs have similar density in the composite film. It is also assumed

that the composites have negligible porosity and swelling in humid conditions.

2.6.4. Thermo-mechanical analysis: Dynamic mechanical thermal analysis (DMTA)

measurements were performed on a dynamic mechanical analyser (TA Instruments Q800)

operating in tensile mode. Typical sample dimensions were 15 5 0.03 mm3. The

measurement frequency was kept at 1 Hz. At a nominal strain of 0.02 %, temperature scan

was made in the range 25 – 300 ºC at a heating rate of 3 ºC min-1 in air. All samples were

pre-heated to 105 ºC for 20 min. to remove any residual moisture in the material. The

glass transition temperature was determined by noting the temperature of the tan δ peak.

2.6.5. Thermal analysis: Thermogravimetric analysis (TGA) was conducted on a Mettler

Toledo TGA/SDTA851 instrument. The samples were heated from 25 to 800 °C at a

heating rate of 10 °C / min. with an O2 or N2 flow of 50 mL / min. The thermogram

recorded in O2 atmosphere was used to find the inorganic content in the nanocomposites

whereas the thermogram recorded in N2 atmosphere was used to study the degradation

behaviour of polymer and nanocomposites.

Differential scanning calorimetric analysis (DSC) was conducted on a Mettler Toledo

DSC instrument. The samples were heated from room temperature to 120 °C, followed by

quenching to -40 °C at a heating/cooling rate of 5 °C / min., in a N2 flow of 50 mL / min.

This was followed by another temperature scan from -40 °C to 360 °C at the same

condition and this scan data was used for comparative analysis of different samples.

Page 32: Xyloglucan-based polymers and nanocomposites

22

2.6.6. X-ray Diffraction: XRD provides the most important parameter for interlayer

distance of MTM platelets in polymer matrix - the spacing between diffractional lattice

planes. The original so-called interlayer or gallery distance between the stacked layers for

Na-MMT is reported to be around 10 Å.81 For an exfoliated nanocomposite, the extensive

delamination of the silicate layers in the polymer matrix results in the eventual

disappearance of any coherent X-ray diffraction from the distributed silicate layers. For

intercalated nanocomposites, the finite layer expansion associated with the polymer

intercalation results in the appearance of a new basal reflection corresponding to the

gallery height.82

X-ray diffraction profile scans were recorded in reflection mode in the angular range of

0.5 - 70 ° (2θ). The measurements were performed with an X’Pert Pro diffractometer

(model PW 3040/60). The CuKα radiation (1.5418 Å) generated with a tension of 45 kV

and current 35 mA was monochromatized using a 20 μm Ni filter. An increment step of

0.05 º and a rate of 1 step per 10 sec. were used. Samples were dried prior to experiment.

XRD was also used to determine the orientation degree of the platelets in the matrix

polymer. For this, XRD images were obtained by using an X-ray diffractometer (Rigaku,

RINT 2100) equipped with Ni filtered Cu Kα radiation. The X-ray beam was operated at

40 kV and 20 mA. The samples were irradiated in parallel (cross-sectional) direction of

the composite films. From azimuthal intensity distribution profiles for the (060) equatorial

reflection, the degree of orientation (Π), and Herman’s orientation parameter (f), were

calculated according to eqs 1−3:

Degree of orientation, Π

Π (180-FWHM)

180 2.3

Herman’s orientation factor, f

cos ϕ ∑ I ϕ cos ϕ/

∑ I ϕ sinϕ/ 2.4

3 cos ϕ 1

2 2.5

Page 33: Xyloglucan-based polymers and nanocomposites

23

where FWHM is the full width at half maximum of the peak in the azimuthal profile of

MTM, ϕ represents the azimuthal angle and I(ϕ) is the intensity along the Debye−Scherrer

ring. f = 1 corresponds to a maximum orientation of the platelets in the plane of the film,

whereas f = 0 indicates random orientation of the platelets and -1/2 corresponds to

perpendicular orientation.

2.6.7. Oxygen Permeability: The oxygen transmission rate (OTR) measurements were

performed with Oxygen Permeation Analyser (Systech 8001, Systech Instruments Ltd.,

UK) at 23 °C using 100 % oxygen as test gas and nitrogen as carrier gas. Tests were done

at dry condition, 50 % RH and 80 % RH in both N2 and O2 flow paths. The active area of

measurement was 5 cm2 by using a steel mask.

To render the oxygen barrier data for coatings more comparable, the oxygen

permeability for each coating was calculated using the relationship:83, 84:

1

t

t

(2.6)

where Ptotal is total permeability of the laminate and Ps and Pc are the permeabilities of

the substrate and coating respectively. The thickness of the coating and the substrate films

are tc and ts respectively so that the total thickness of the laminate is t. In this model, it is

assumed that substrate/coating interface has no significant effect on oxygen permeation.

Nielsen model: For dry films, the effect of MTM on permeability can be compared with

predictions based on Nielsen’s model based on the tortuous path which a penetrant gas is

forced to diffuse along due to the in-plane oriented platelet organization.85 The effect of

tortuosity on the permeability is expressed as a function of the length (L) and thickness of

the sheets (W), as well as their volume fraction in the polymer matrix (φs):

PXG

= 1

(1+(L

2W ) ) (2.7)

where Pc and PXG represent the permeability of the XG/MTM nanocomposites and pure

polymer, respectively. An average length of 100 nm and thickness 1 nm is taken for

MTM. The model assumes that platelets are oriented normal to the direction of diffusion.

Page 34: Xyloglucan-based polymers and nanocomposites

24

2.6.8. Electron microscopy analysis: The samples for transmission electron microscopy

(TEM) study were prepared by embedding in epoxy resin. The cured epoxies containing

nanocomposite film strips were cut with a LKB Bromma 2088 ultramicrotome into slices

of 50 – 100 nm in thickness. These slices were placed on a 200 mesh copper grid and

observed using a Philips Tecnai 10 electron microscope operated at 80 kV.

An ultra-high resolution field emission-scanning electron microscope (FE-SEM) (Hitachi

S-4800) employing a semi-in-lens design and a cold field emission electron source was

used for micro-structural analysis. The cross-sectional samples prepared by using a

microtome for TEM were also used for FE-SEM analysis. Prior to FE-SEM observation,

samples were vacuum dried. In order to suppress specimen charging during analysis, the

specimens were coated with gold/palladium (3 nm thickness) using an Agar HR sputter

coater.

2.6.9. Water sorption: A Dynamic Vapour Sorption (DVS) instrument from Surface

Measurement Systems was used to determine the water sorption isotherm under different

humidity atmospheres. XG and modified XG films were first dried in the DVS cell, and

then the relative humidity (RH) in the DVS cell was increased in steps from dry state up

to 95 %. The samples were weighed in different humidity atmospheres when the steady

state point was reached. The moisture content (M) at a particular RH level was calculated

on a dry (or total weight) basis, as indicated by the formula below:

M 100% 2.8

where M is the moisture content (% by weight) in the material, Ww is the weight of the

sample in the DVS cell when the water content has reached steady state condition, and Wd

is the weight of the dried sample.

Page 35: Xyloglucan-based polymers and nanocomposites

25

3.RESULTSANDDISCUSSION

3.1.Xyloglucanfilmproperties

3.1.1.Film‐formation

Film preparation of polysaccharides is often challenging. For most hemicelluloses, an

undesirable amount of plasticizer addition is necessary to form films of acceptable

quality.7, 86, 87 This is most likely due to the low molar mass of the materials and their

limited solubility in water. In contrast, XG in the present study showed excellent film-

forming characteristics that could be attributed to its high molar mass of around 1.5 - 2.5

MDa. Despite the high molar mass, XG can still be dissolved in water up to a

concentration of ca. 4 %, forming a viscous solution. The cast films were mechanically

robust, optically transparent and smooth. A comparison between the XG film and the film

casted from hard wood xylan in lab trials is presented in Figure 3.1. Enzymatically

modified XG (with 25 - 30 % galactose removed) showed similar film-forming

properties, but the preparation was more challenging and extensive mechanical treatment

was necessary to dissolve the material in water. Thus, it is concluded that the presence of

the galactose has a strong influence on the solubility of XG in water.

Figure 3.1. (A) Xyloglucan and (B) hardwood xylan films without adding plasticizer.

Page 36: Xyloglucan-based polymers and nanocomposites

26

3.1.2.Hygrothermalproperties

XG absorbs water and the moisture sorption isotherm for XG is shown in Figure 3.2. in

comparison with waxy maize starch and potato starch.88 The equilibrium moisture content

of XG films is much lower than that for the starches in the practically very important

relative humidity range 0 – 70 %. The cellulose backbone may be responsible for this

property since cellulose is considered amphiphilic with a very low water solubility due to

the interactions between the non-polar part of the cellulose backbone.89, 90 XG inherits this

nature of cellulose due to its backbone structure, whereas the presence of side sugars

results in a more hydrophilic nature compared to cellulose. The increased water uptake at

high RH indicates capillary condensation.91

The storage modulus (E') and tan δ plots versus temperature for XG films is presented

in Figure 3.3 (A). It is interesting to note that XG is thermomechanically stable until the

large drop in E' versus temperature at around 275 °C associated with the glass transition

of amorphous polymers in contrast to the Tg reported for cellulose (220 ºC).92 On the

other hand, many other polysaccharides including starch components (amylose and

amylopectin) degrade prior to glass transition.93 XG is thermally stable until around 300

°C, where it starts to degrade with maximum degradation above 310 °C, see the TGA

results in Figure 3.3 (B). This is similar to the observations made for cellulose, a pyrolysis

temperature in the range 300 – 400 °C.94, 95 Hemicelluloses generally show lower thermal

stability than cellulose.95 The thermal degradation of hemicelluloses starts at much lower

temperature of the order of less than 250 °C. The high thermal stability of xyloglucan is

related to its ‘cellulosic nature’, which is not the case for other hemicelluloses.

Page 37: Xyloglucan-based polymers and nanocomposites

27

Figure 3.2. Water sorption isotherm for Xyloglucan, waxy maize starch88, and potato starch films88 as a function of relative humidity.

Figure 3.3. (A) Storage modulus and tan δ of XG films at 1Hz as a function of temperature, (B) Thermogravimetric analysis and first derivative thermogram for native XG in N2 atmosphere.

3.1.3.Hygromechanicalproperties

The use of XG in industrial bioplastics and biocomposites applications would imply

exposure to ambient and moist conditions. In Figure 3.4., stress-strain curves are

presented for XG-films subjected to different relative humidities at room temperature and

the data are summarized in Table 3.1. In the dry state, XG shows a Young’s modulus of

ca. 6 GPa, and a tensile strength of 114 MPa. The modulus is significantly higher than

other hemicelluloses reported in the literature7, supposedly due to the strong

intermolecular interaction and low free volume at the molecular scale.96 Predominant in-

plane orientation of the macromolecular chains is also possible. As moisture is absorbed,

Page 38: Xyloglucan-based polymers and nanocomposites

28

XG becomes more ductile (strain-to-failure increases), but strength and modulus are

reduced. However, even at severe humidity environment, XG films preserve a substantial

amount of mechanical function (modulus of ca. 2 GPa and strength of 45 MPa at 92 %

RH). High molar mass and the associated physical entanglements are advantages96, and

possibly intermolecular interactions between XG molecules are still quite high despite the

presence of absorbed moisture.

Figure 3.4. Typical tensile stress-strain curves for xyloglucan films conditioned under room temperature and different relative humidity atmospheres (relative humidity values and corresponding moisture uptake are indicated).

Table 3.1. Mechanical properties of XG films conditioned at different RH using data obtained from tensile tests (values in parentheses are standard errors).

Sample condition Tensile strength, MPa

Strain to failure, % Elastic modulus, GPa

Vacuum dried 114 (12.0) 2.7 (0.6) 5.95 (0.83)

RH 15 % (4 % H2O) 64 (5.6) 6.6 (3.6) 5.07 (0.24)

RH 50 % (15 % H2O) 61 (8.6) 4.8 (3.6) 4.30 (0.2)

RH 74 % (17 % H2O) 56 (3.1) 6.4 (0.5) 3.67 (0.23)

RH 92 % (31 % H2O) 45 (4.7) 9.2 (0.7) 2.36 (0.71)

RH 99 % (ca. 40 % H2O)

28 (6.4) 7.5 (1.1) 1.63 (0.31)

Page 39: Xyloglucan-based polymers and nanocomposites

29

3.1.4.Oxygenpermeability

Low oxygen permeability is an important target property for food packaging films.

Xyloglucan has very low oxygen permeability and is comparable to the commercial

barrier polymers such as PVOH and biopolymers such as wood hemicelluloses (see Table

3.2). The oxygen permeability of the xyloglucan film under dry condition is 0.41

cc.µm/[m2-day]kPa at 23 °C and the average oxygen permeability at 50 % RH and 23 °C

is 0.5 - 2.0 cc.µm/[m2-day]kPa. An explanation for this is the strong intermolecular

interactions that, together with the high chain stiffness, could have helped to reduce the

free volume in the polymer.

Table 3.2. Oxygen permeability of packaging materials measured at 23 C, 50 % RH

Material Plasticizer Oxygen permeability

(cc.µm/[m2-day]kPa)

Reference & thickness

AcGGM sorbitol (21 wt %)

2.0 52 thickness=30 – 60 µm

AcGGM-CMC - 1.3 52 thickness=30 – 60 µm

amylopectin glycerol (40 wt %)

14 97 thickness=70 – 100 µm

poly(lactic acid)

- 160 Present study, thickness=100 µm

Poly(vinyl alcohol)

- 0.21 86 thickness=35 µm

High density polyethylene

- 427 Miller et al.47

Polyester . 15.6 Miller et al.47

xylan sorbitol (35 wt %)

0.21 86 thickness=35 µm

xyloglucan - 0.5 - 2.0 Present study, thickness=20-30 µm

Page 40: Xyloglucan-based polymers and nanocomposites

30

However, similar to other polysaccharides and polyols, oxygen permeability increases

in high humid atmosphere. For instance, at 80 % RH, XG films show oxygen

permeability of the order of more than 10 cc.µm/[m2-day]kPa. Improved oxygen barrier

performance at high RH becomes a technical goal in the present thesis and will be

discussed in coming sections.

3.2.Effectofplasticization

Thermoplastic starch is well established in research and industry. Plasticized starch is

used for compression moulding or extrusion of articles. The Tg of XG is very high, in the

order of 260 °C, and this is an obstacle in the context of thermal processing. Plasticizing

effects were studied using four different candidate plasticizers: urea, glycerol, sorbitol,

and polyethylene glycol (PEG). Film casting from different solution compositions

resulted in good quality and more or less transparent films. However phase separation

occurred in films containing urea higher than 20 wt%. This was noted as visible patches

of white urea crystals on the film surface.

3.2.1.Thermomechanicalproperties

Urea and PEG were not effective in decreasing the Tg of the XG films. On the other

hand, both sorbitol and glycerol decreased the Tg of XG films. Only the sample

containing 10 wt% glycerol showed a clear and large stiffness drop of the film

corresponding to the expected glass transition temperature of the glycerol/XG system. A

complex softening behaviour was noticed with increased glycerol content, most likely due

to the migration of the plasticizer from the bulk to the surface. For the films containing

sorbitol, a clear softening effect was evident with a progressive lowering of the softening

temperature with increasing content of sorbitol (Figure 3.5). This softening corresponds to

the Tg of the sorbitol/XG mixture as evident from the tan δ peak in Figure 3.5. With 40

wt% of sorbitol added, a Tg decrease of more than 130 ºC is observed. Thus with sorbitol

as a plasticizer, the softening of the XG film may be lowered to temperatures suitable for

commercial scale thermal processing.

Page 41: Xyloglucan-based polymers and nanocomposites

31

Figure 3.5. Storage modulus and examples of tan at 1 Hz versus temperature for neat XG (A) as well as XG films containing 10 (B), 20 (C), 30 (D) and 40 (E) wt% sorbitol.

3.2.2Tensileproperties

In Table 3.3. the stress–strain properties of XG films with 20 and 30 wt% sorbitol are

compared with a neat XG film. The neat XG film has a low strain-to-failure of the order

of less than 5 % at 50 % RH and 23 ºC. In contrast, the plasticized XG film (20 wt%

sorbitol) shows ductile behaviour with a strain to failure above 25 %. The films preserve

strength and modulus in a satisfactory manner as apparent from Table 3.3. From a

practical point of view, the data shows an attractive combination of modulus, strain-to-

failure and tensile strength with a decreased Tg of more than 100 °C obtained with

plasticized XG containing 20 – 30 wt% sorbitol.

Table 3.3. Tensile properties of XG and sorbitol plasticized XG films measured at 23 °C and 50 % RH (values in parentheses are standard errors).

Strain to failure, ε

(%)

Tensile Strength, σ

(MPa)

Young's Modulus, E

(GPa)

XG 4.8 (3.6) 61 (8.6) 4.3 (0.2)

XG + 20 wt% sorbitol 27 (7.0) 50 (6.5) 1.17 (0.09)

XG + 30 wt% sorbitol 50 (13) 43 (10) 0.65 (0. 11)

Page 42: Xyloglucan-based polymers and nanocomposites

32

3.3.Effectofchemicalmodification

The previous study showed that at least 20 - 30 wt% of plasticizer is needed to reduce

the Tg of XG to a level where thermal processing is an attractive alternative. In the present

work xyloglucan is instead chemically modified in order to reduce the Tg. The chemical

modification uses periodate oxidation of vicinal hydroxyl groups present on XG to form

dialdehyde products with ring opening and subsequent reduction to hydroxyl groups (see

section 2.2.1).

3.3.1.Periodateoxidationandsubsequentreduction

XG was rapidly oxidized by periodate at room temperature to form dialdehyde XG. The

resulting aldehyde groups were reduced to hydroxyl groups by the use of NaBH4. The

reaction steps are discussed in section 2.2.1. The dialdehyde formation and subsequent

reduction were monitored with the characteristic carbonyl peak at 1720 cm-1 in FTIR. The

reduction of the aldehyde groups was complete as inferred from the disappearance of

carbonyl absorption peak at 1720 cm-1 in the modified XG samples.

The effect of oxidation on carbohydrate composition of XG and modified XG samples

is presented in the carbohydrate analysis, see Table 3.4. The amount of glucose in the

dXG30 sample remains constant, whereas a slight decrease in the amount of glucose is

observed for dXG60 and dXG120 samples. The reduction in the amount of glucose in

dXG60 and dXG120 samples could be the results of the oxidation of few glucopyranose

rings during periodate oxidation. The data reveals that the galactose and xylose residues

on the side chains are favoured for oxidation compared to the glucose residues on the

main chain. The selectivity mechanism is not completely understood. One reason for the

preferential oxidation is the availability of the pendant side sugar residues compared to

the glucose residues on the main chain. Moreover, the vicinal –OH groups of the β-

(1→4)- linked D-glucopyranoses main chain are in trans state, which is kinetically less

favourable for periodate oxidation than the predominant cis conformations observed in

galactose and xylose residues.

Page 43: Xyloglucan-based polymers and nanocomposites

33

Table 3.4. Carbohydrate composition of XG and modified XG samples. Data are reported as mmols of carbohydrates / 100 mg of sample. Designations dXG30, dXG60 and dXG120, refer to time in minutes for periodate treatment.

Samples Glucose Galactose Xylose

XG 0.269 0.088 0.203

dXG30 0.271 0.0008 0.023

dXG60 0.263 Nil* 0.004

dXG120 0.231 Nil* 0.005

*The amount was below the detectable limit of the instrument, ca. 0.1 mg/L.

The effect of oxidation on the structure of xyloglucan was characterized with MALDI-

TOF MS analysis. Xyloglucan oligosaccharides (XGOs) were obtained from enzymatic

hydrolysis of neat XG. The mass / charge ratio values (m/Z) recorded in the mass

spectrum for XGOs from neat XG are presented in Table 3.5. Similarly, the modified

XGs were enzymatic hydrolyzed to form oligosaccharides. The m/Z values of these

modified XGOs as well as the anticipated chemical structures are presented in Table 3.6.

Page 44: Xyloglucan-based polymers and nanocomposites

34

Table 3.5. Four basic repeat units of neat XG oligosaccharides obtained from m/Z peaks recorded in MALDI-TOF MS (The exact mass (EM) includes atomic weight of Na+ counter ion)

Page 45: Xyloglucan-based polymers and nanocomposites

35

Table 3.6. The modified XG oligosaccharide structures interpreted from m/Z values obtained from MALDI-TOF MS (The exact mass (EM) includes atomic weight of Na+ counter ion)

modified

XXXG O O

O

O

O O

O

O

O

O

O

O

HO

O

OH

HOHO

OH

HO

OH

OH

HOOH

OH

HO

HO

HO

HO

HOOH

OH

Na

HOHO

EM = 1061.4

EM = 1091.4

O O

O

O

O O

O

O

O

O

O

O

HO

O

OH

HOHO

OH

HO

OH

OH

HOOH

OH

HO

HO

HO

HO

HOOH

OH

Na

HO

EM = 1063.4

EM = 1093.4

modified

XXLG or XLXG

EM = 1225.4 EM = 1195.4

O O

O

O

O O

O

O

O

O

O

O

HO

O

O

OOH

HOHO

OH

HO

OH

OH

HOOH

OH

HO

OHHO

HOHO

HOOH

Na

HO

OH

EM = 1165.4

modified

XLLG

EM = 1389.5 EM = 1359.5 EM = 1329.5

Page 46: Xyloglucan-based polymers and nanocomposites

36

The mass peaks at m/Z values of 1409.2, 1247.2 and 1085.2 correspond to the XGO

structures of XLLG, XXLG or XLXG and XXXG, which have molecular mass of 1386.5,

1224.4 and 1062.4 Da, respectively. All XGOs were oxidized during periodate treatment

with none of the original XGOs present in the final product for all oxidation conditions.

The modified XGOs are identical in all samples except for the products obtained for

dXG30 sample where the m/Z value of modified XXXG oligosaccharides are different

from the corresponding values for dXG60 and dXG120 samples. The m/Z value of 1063.4

and 1093.4 for modified XXXG of dXG60 and dXG120 samples points to the fact that

oxidation also occurred at one of the four glucose units, which results in an increase of the

m/Z value by 2 units. On the other hand, for the dXG30 sample, the m/Z values for

modified XXXG were predominantly recorded at 1061.4 and 1091.4. These results

suggest that the glucopyranose rings in modified XXXG for dXG30 samples were not

likely affected by periodate treatment compared to dXG60 and dXG120 samples. This is

in accordance with the carbohydrate analysis discussed earlier, where the amount of

glucose remained constant for dXG30 sample, and a slight decrease in glucose content

was observed for dXG60 and dXG120 samples (Table 3.4). The results point to the fact

that the cellulose backbone is largely preserved in the modified XG samples. The

schematic structures deduced from MALDI-TOF MS data and sugar analysis are

presented in Figure 3.6.

Figure 3.6. Schematic representations of the modified XG structures, as deduced from MALDI-TOF MS and sugar analysis. Designations dXG30, dXG60 and dXG120, refer to time in minutes for periodate treatment.

Page 47: Xyloglucan-based polymers and nanocomposites

37

3.3.2.PropertiesofmodifiedXGfilms

The typical stress-strain curves in tension for modified XG samples at 50 % RH and 23

°C are presented in Figure 3.7.(A) and property data are presented in Table 3.7. The

dXG30 sample shows very high toughness (work to fracture, i.e. area under stress-strain

curve). The modulus and yield strength decrease with increased oxidation time. The

oxidation opens the side chain sugars and this result in differences in intermolecular

interaction and spatial packing of macromolecules compared with native XG. The

molecular weights of all modified XG substrates are fairly high due to the short oxidation

time used (see Table 3.7.). In contrast, periodate oxidation of cellulosic materials require

longer oxidation time in ambient conditions and is always accompanied by severe

depolymerization so that the final product has inferior properties compared to commercial

polymers.98

Figure 3.7. (A) Typical tensile stress–strain curves for xyloglucan films conditioned at room temperature and 50 % RH (B) Moisture sorption isotherms for native XG and modified XGs. Designations dXG30, dXG60 and dXG120, refer to time in minutes for periodate treatment.

In section 3.1.3., the plasticizing effect of water for XG films are discussed. The

moisture sorption isotherms (Figure 3.7. (B)) show that modified XGs are less

hygroscopic than XG, at least at RH < 80 %. The moisture uptake at 50 % RH is up to 60

% lower than native XG. Side sugar opening reduces moisture absorption at the molecular

level. Although the precise mechanism is unclear, it can be speculated that opening of

side chain sugar residues to aliphatic chains enhances the ‘cellulosic nature’ of the

backbone, that might give the polymer a less hydrophilic character. Increased strain to

Page 48: Xyloglucan-based polymers and nanocomposites

38

failure for modified XGs can then best explained by the internal plasticization effects

from the opened side chains rather than moisture plasticization.

Table 3.7. Tensile properties of modified XG using data obtained from tensile tests at 23 °C and 50 % RH (values in parentheses are standard errors). Designations dXG30, dXG60 and dXG120, refer to time in minutes for periodate treatment. PDI refers to polydispersity index, Mw / Mn.

Sample Mw from SEC, MDa

Tensile strength,

MPa

Strain to failure

Elastic modulus, MPa

Yield stress, MPa

XG 1.2 (PDI = 1.62) 78 (8.6) 5.7 (3.6) 4597 (164) 72 (5.4)

dXG30 0.77 (PDI = 2.61) 37 (3.9) 33 (11.2) 1084 (141) 28 (0.8)

dXG60 0.31 (PDI = 2.11) 17 (1.7) 50 (4.2) 333 (24.9) 7.6 (0.3)

dXG120 0.57 (PDI = 2.09) 20 (2.2) 46 (14) 291 (68.3) 6.9 (2.5)

Figure 3.8. Storage modulus (E’) and tan δ at 1 Hz as a function of temperature for XG and modified XG with different oxidation times. Designations dXG30, dXG60 and dXG120, refer to time in minutes for periodate treatment.

The dynamic mechanical thermal properties of XG and modified XGs as a function of

temperature are presented in Figure 3.8. The native XG has a Tg of the order of 275 °C.

The modification reduces the glass transition temperature of native XG to 167 °C for

dXG30. The Tg is further reduced to around 145 °C for dXG60 and dXG120 samples

(Figure 3.8). The Tg of modified XGs is in the range of water-soluble cellulose

derivatives such as hydroxyethyl cellulose, HEC (120 °C).99 Reductions in Tg correlate

qualitatively with reductions in modulus and yield stress observed in the tensile tests. The

Page 49: Xyloglucan-based polymers and nanocomposites

39

decrease in Tg is a consequence of the opening of pendant side chains of XG. It results in

perturbations in the spatial packing of modified XG chains, changes in intermolecular

interactions and increased molecular mobility. The reduction in Tg is significant from a

thermal processing point of view.

Oxygen permeability of modified XG, 0.12 - 0.15 cc.µm/[m2-day]kPa, is slightly higher

than that for native XG, 0.08 cc.µm/[m2-day]kPa in the dry state. However, it is

significantly lower than that for commercial barriers in use today, such as PVOH (0.38

cc.µm/[m2-day]kPa). The chain packing in XG is significantly altered as a result of the

opening of side chains. This is reflected in increased oxygen permeability of modified

XGs compared to XG. On the other hand, the oxygen barrier performance of modified

XGs is markedly improved at 50 % RH. When the permeability of XG is recorded at 2.16

cc.µm/[m2-day]kPa, the dXG120 sample shows permeability values in the order of 0.5

cc.µm/[m2-day]kPa. The main reason could be that the modified XGs have a reduced

moisture uptake compared with native XG, and therefore has a larger effect in the oxygen

barrier performance at 50 % RH.

In summary, the highly advantageous oxygen permeability of XG is improved at 50 %

RH for modified XG samples, making this novel biopolymer an interesting candidate for

packaging applications. The largely preserved cellulose backbone is important as is the

reduced moisture absorption compared with XG. The modification results in water-

soluble “cellulose derivatives”, structurally similar to cellulose derivatives such as

hydroxyethyl cellulose (HEC) and hydroxypropyl cellulose (HPC) etc. Furthermore, the

modification is carried out without the use of organic solvents, as is used in the

preparation of conventional cellulose derivatives.

3.4.Xyloglucan/MTMbionanocomposites

Similar to other polysaccharides, the material properties of XG are reduced at high

humidity. The present subproject focuses on improving the properties by dispersing MTM

platelets in XG matrix.

Page 50: Xyloglucan-based polymers and nanocomposites

40

3.4.1.AdsorptionofXGtoMTM

The adsorption of polymer on MTM surface in aqueous environment is important in the

final properties of the nanocomposites. Even though the adsorption of polysaccharides on

clay surface has been studied in soil science, 100, 101 it has never been thoroughly

investigated in a materials science context, where several polysaccharide/clay

nanocomposites already exists.

Figure 3.9. Frequency changes in quartz crystal microbalance analysis (QCM) as function of time for PVOH, xyloglucan and modified xyloglucan adsorbed on MTM model surfaces. Model surfaces were prepared by consecutive deposition of three bilayers of MTM and polyDADMAC. The adsorbed mass per unit area (m), including water, calculated from the frequency difference (Δf) is presented alongside.

The adsorption of XG and enzymatically modified XG from water solution to MTM

was compared with the adsorption of a commercial polymer, PVOH, using a quartz

crystal microbalance (QCM). Modified XG was prepared by enzymatically removing part

of the galactose side chains (see section 2.2.2). The results presented in Figure 3.9. show

that the amount of XG adsorbed on MTM is 6.5 mg / m2, and this is higher than that for

modified-XG (5.1 mg / m2) and significantly higher than the amount of adsorbed PVOH

(2.8 mg / m2). Note that these adsorption values include immobilized water. The

adsorption of uncharged polymers to clays is largely driven by the entropy gained by the

desorption of water molecules from the clay surface.100 XG is considered water-soluble,

but the individual chains are not fully hydrated, showing an amphiphilic character even in

Page 51: Xyloglucan-based polymers and nanocomposites

41

dilute solutions.28 Modified XG has even lower water solubility, since the density of

galactose side chains has been reduced. If the difference in associated water is taken into

account, the mechanism for the adsorption of XG and modified XG may partly be due to

hydrophobic interaction with the MTM surface. The data indicate that van der Waal’s

attractive forces and hydrophobic interactions may contribute to the formation of an

XG/MTM interface in a polar water environment.

3.4.2.Nanocompositefilmcharacterization

Films and coatings with MTM content as high as 20 % by dry weight (ca. 12 vol %)

were successfully prepared from XG/MTM aqueous colloidal suspensions.

The MTM orientation in solution cast XG/MTM nanocomposites was studied using X-

ray diffraction (XRD). In the cross-sectional plane, strong in-plane orientation of MTM

platelets can be observed as shown by strong reflections from (020) planes, whereas the

orientation in the plane of the film is completely random (Figure 3.10 A & B). The degree

of orientation is estimated from the azimuthal intensity distribution (Figure 3.10(C)) and

is as high as 74 % in the cross-sectional plane. This reveals a layered structure similar in

nature to the high clay content multi-layered structures prepared by LbL assembly66 or

papermaking approaches.67, 70 The molecular rigidity of XG is related to its cellulose

backbone and may be significant for the in-plane orientation of MTM platelets.

Figure 3.10. X-ray diffractograms of a nanocomposite film containing 20 wt% MTM in perpendicular (A) and parallel (B) to the film surface. (C) Azimuthal intensity distributions of XG/MTM nanocomposites containing 20 wt% MTM derived from XRD image presented in (B)

Page 52: Xyloglucan-based polymers and nanocomposites

42

The X-ray diffraction spectra, Figure 3.11(A), indicate that for MTM addition of 1 wt%

and 2.5 wt%, the MTM platelets are completely exfoliated in the XG matrix polymer,

since the XRD peak associated with d001= 9.8 Å for stacked layers of MTM disappears

completely. At 5 wt% MTM or more, intercalated MTM structures are predominantly

formed and the average interlayer spacing is 26 Å. Interestingly, the interlayer gallery

spacing for the XG/MTM nanocomposites is independent of MTM content. TEM

micrographs of a representative nanocomposite film with 10 wt% MTM are presented in

Figure 3.11 (B) and (C). Intercalated MTM platelets with parallel orientation are apparent

from Figure 3.11(B), while the TEM reveals also that part of MTM platelets are

exfoliated, Figure 3.11(C). XRD and TEM point to a fairly uniform structure at the

nanoscale for XG/MTM nanocomposites.

Figuer 3.11. (A) X-ray diffraction pattern of XG-MTM nanocomposite. The interlayer or gallery distance between the stacked layers for MTM is 9.8 Å. (B) and (C) TEM image of the cross-section of XG-MTM nanocomposite containing 10 wt% of MTM showing silicate layers as dark lines.

3.4.2.Filmproperties

The tensile properties show remarkable improvements for XG/MTM nanocomposites,

see Figure 3.12. The tensile strength at 50 % RH and 23 ºC is increased from 92 MPa for

native XG to 123 MPa at 20 wt% MTM content. There is a remarkable three-fold increase

in XG/MTM modulus for the same composition, 4 GPa for neat XG to 11.6 GPa for

nanocomposites containing 20 wt% MTM. At less than half the MTM content, the

Page 53: Xyloglucan-based polymers and nanocomposites

43

modulus reaches the same level as nanocomposites with more than 50 wt% MTM in

PVOH or polyelectrolyte matrices prepared by LbL technique.66, 69, 72 This means that the

present XG/MTM bionanocomposite apparently has high reinforcement efficiency, due to

high in-plane orientation of MTM platelets and favourable XG-MTM interfacial

adhesion.

The high matrix content in the present nanocomposites compared with nacre-mimetic

composites improves ductility. The high ductility also indicates good dispersion of the

MTM platelets, since agglomerates would initiate failure at low strains. The

nanocomposites with MTM content of up to 2.5 wt% show remarkable strain to failure of

more than 10 %. Even though the precise mechanism is unknown, one may note that the

clay platelets are completely exfoliated at this MTM content (see Figure 3.11 (A)).

Figure 3.12. (A) Typical stress- strain curves of XG-MTM films conditioned at 50 % RH and 23 °C (B) Modulus and tensile strength data for XG/MTM nanocomposites as a function of MTM content.

The oxygen transmission rate was measured on coatings of XG/MTM on OPET film.

The cross-sectional FE-SEM images (Figure 3.13(A)) reveal that the MTM platelets are

predominantly oriented parallel to the substrate film similar to cast films previously

discussed.

The apparent oxygen permeability of the nanocomposite coatings is presented in Table

3.8. There are two main observations: increasing MTM content strongly improves barrier

properties, and even at high relative humidity, the oxygen permeability is very low for

high MTM content nanocomposite coatings. The relative permeability of XG/MTM with

Page 54: Xyloglucan-based polymers and nanocomposites

44

respect to native XG in the dry state is plotted as a function of MTM content in Figure

3.13(B), as well as predictions based on the Nielsen model. The MTM effect in the

Nielsen model is based on the more tortuous path which a penetrant gas is forced to

diffuse along due to the in-plane oriented MTM platelets. The agreement between

predictions and data are very good up to an MTM content of 3 % by volume, in support of

the tortuosity mechanism. The experimental data, however, reaches a plateau value earlier

than predicted. Considering the nature of the model and the data, it is difficult to ascribe

this observation to a specific mechanism.

Figure 3.14. (A) Cross-sectional view of xyloglucan-MTM composites (8.9 wt% MTM) coated on OPET film as observed by FE-SEM. (B) Relative oxygen permeability for XG/MTM nanocomposites with respect to neat XG in dry condition as a function of MTM content (volume fraction), model predictions and experimental data.

Table 3.8. Average oxygen permeability of XG/MTM nanocomposite films (cc.µm/[m2-day]kPa)

An interesting application of XG/MTM films could be as replacement of aluminium

barriers in liquid packaging. The oxygen permeability at 80 % RH is then of particular

interest, since polysaccharides typically fail to perform under these conditions. In Table

3.8., it is apparent that the XG/MTM composition with 20 wt% MTM has an oxygen

permeability of only 1.44 cc.µm/[m2-day]kPa at 80 %RH.

Condition XG XG+4.3 wt% MTM

XG+8.9 wt% MTM

XG+20 wt% MTM

0 % RH, 23 °C 0.02 0.01 0.00 0.00

50 % RH, 23 °C 0.45 0.18 0.04 0.05

80 % RH, 23 °C - 30.1 6.03 1.44

Page 55: Xyloglucan-based polymers and nanocomposites

45

3.5.Nacre‐mimeticnanocompositesbasedonxyloglucanandclay

Cross-linking has been used for nacre-mimetic materials based on water-soluble

polymers in order to improve substantially the stiffness, tensile strength and

hygromechanical performance.72 In the present study, nanocomposites with high

inorganic content were prepared by creating cross-linking sites on XG by regioselective

oxidation of side chains followed by adsorption on MTM platelets. A covalent cross-link

of acetaldehyde is expected to be formed between the highly reactive aldehyde groups on

oxidized XG and hydroxyl groups on XG chains or MTM surface (see Scheme 2.3.). The

mechanical and oxygen barrier performance of the resulting nanocomposites under humid

conditions are studied.

3.5.1.Compositepreparation

XG was oxidized by periodate to form dialdehyde XG as noticed by the strong

absorption peak at 1720 cm-1 in FTIR. The carbohydrate analysis of the sugars present in

XG and oxidized XGs is presented in Table 3.8. The glucose content is similar for all

samples, which means that the cellulose backbone is preserved also for the oxidized

samples. The periodate oxidation of XG is, therefore, regio-selective for the present

oxidation condition so that only the side chains are oxidized. The side chains, galactose

and xylose residues, are increasingly oxidized as the amount of periodate increased.

Table 3.8. Carbohydrate composition of XG and oxidised XG samples. Data are reported as mmols of carbohydrates / 100 mg of sample. Designations ddXG10, ddXG20 and ddXG30, refer to the stoichiometric amount of periodate used during oxidation of XG.

Samples Galactose Glucose Xylose

XG 0.086 0.256 0.183

ddXG10 0.042 0.269 0.180

ddXG20 0.017 0.265 0.159

ddXG30 0.003 0.252 0.121

One important feature of XG is its strong adsorption to the MTM surface (section

3.4.1). The TGA results showed that ca. 40 wt% of XG is adsorbed to MTM platelets and

Page 56: Xyloglucan-based polymers and nanocomposites

46

the amount of adsorption is more or less similar for oxidized XGs. Thus oxidation of XG

does not strongly affect the affinity of XG to the MTM surface.

Colloidal suspensions were prepared with XG adsorbed MTM platelets (reference

material) and also suspensions with MTM platelets to which oxidized XG had been

adsorbed. The coated MTM suspensions were vacuum filtered and dried to form

nanocomposite films. The cross-linking reaction demonstrated in Scheme 2.3 is likely to

take place during drying, and covalent bonds are formed both between XG chains and

between XG chains and hydroxyls at MTM surfaces.

3.5.2.Nanocompositefilmcharacterization

The X-ray diffraction (XRD) analysis, Figure 3.15, shows that the XRD peak associated

with (001) reflection for MTM is disappeared or shifted to lower 2θ for the

nanocomposites. This indicates an exfoliated and/or intercalated platelet structure. Due to

the preparation procedure (see section 2.3.2.) the clay platelets are to a substantial extent

individually separated by a coating of XG.

Characterization of nanocomposite film cross-sections with FE-SEM shows a layered

architecture with strong in-plane orientation of the MTM platelets, see Figure 3.16. This

is conceptually similar to the hybrid materials fabricated by layer-by-layer assembly.72

Figure 3.15. X-ray diffraction patterns of nacre-mimetic XG-MTM nanocomposites and Na-MTM. Designation ddXG20 refers to the stoichiometric amount of periodate used during oxidation of XG.

In the cross-sectional plane, in-plane orientation of MTM platelets is observed as

noticed from XRD images in Figure 3.17 (A) and (B) for XG/MTM and oxidized

Page 57: Xyloglucan-based polymers and nanocomposites

47

XG/MTM respectively. Interestingly, the degree of orientation for nacre-mimetic

XG/MTM and oxidized XG/MTM are in the same range and are as high as 82 %, Table

3.9. Furthermore, Herman’s orientation parameter is calculated to be as high as 0.4,

confirming the high degree of in-plane orientation. The processing method adopted in the

present case results in a more oriented platelet distribution in the composites compared to

free standing films, where the degree of orientation was 74 % (section 3.4.2). Highly

oriented MTM platelets have important implications for mechanical and barrier properties

of the nanocomposites.

Figure 3.16. Scanning electron microscopy images of the cross-section of nacre-mimetic XG/MTM nanocomposites.

Table 3.9. Degree of orientation (Π) and Herman’s orientation parameter (f) of nanocomposite samples. Designation ddXG20 refers to the stoichiometric amount of periodate used during oxidation of XG.

Sample FWHM (deg) Degree of orientation (Π), %

Herman’s orientation

parameter (f)

Nacre XG/MTM 31.8 82.3 0.39

Nacre ddXG20/MTM 31.7 82.4 0.41

Page 58: Xyloglucan-based polymers and nanocomposites

48

Figure 3.17. (A) and (B) cross-sectional X-ray diffraction images of nacre-mimetic XG/MTM and ddXG20/MTM nanocomposites respectively, (C) and (D) the corresponding azimuthal intensity distribution profiles of the equatorial reflection (060) of MTM. Designation ddXG20 refers to the stoichiometric amount of periodate used during oxidation of XG.

Thermal analysis of the nanocomposites gives insight into XG/MTM interaction at the

molecular scale. The DSC thermogram (Figure 3.17(A)) for native XG has shown a glass

transition temperature at about 270 °C. The oxidized XG has not shown any glass

transition since cross-linking most likely increases the Tg above the temperature for onset

of thermal degradation. The DSC curve shows distinct thermal change at around 196 °C

for nacre-mimetic XG/MTM composite. The thermal changes are related to composite

degradation as evidenced by thermogravimetric analysis, where the onset of thermal

degradation occurred at around 200 °C. Naturally occurring montmorillonite is a complex

material with many different ions present, including Fe3+ and Mg2+ ions.102 These ions can

form weak complexes with polysaccharides and polyols,103 resulting in reduced thermal

Page 59: Xyloglucan-based polymers and nanocomposites

49

stability. For oxidized XG/MTM, thermal changes are noticed at even lower temperature.

Moreover, there are two distinct thermal changes noted for oxidized XG/MTM samples

(Figure 3.17(B)).This confirms two distinct chemical environments in the composite, one

associated with the interaction of sugar residues with MTM and the other likely arise from

covalent cross-links between aldehyde groups and hydroxyl groups as discussed

previously.

Figure 3.18. (A) Differential scanning calorimetry analysis of XG, oxidized XG and corresponding nacre-mimetic nanocomposites with montmorillonite (MTM). (B) Closer view of the thermal changes associated with nacre-mimetic nanocomposites made from native XG and oxidized XG samples. Designation ddXG10, ddXG30 refer to the stoichiometric amount of periodate used during oxidation of XG

3.5.3.Filmproperties

The tensile properties of nanocomposite films at 50 % RH and 23 °C are reported in

Table 3.10. and typical tensile plots are given in Figure 3.19. The modulus and strength of

nacre-mimetic XG/ MTM are increased significantly compared to native XG. The elastic

modulus of the native XG/MTM nanocomposite is 13.7 GPa and tensile strength is ~ 100

MPa. Nacre-mimetic chitosan/MTM bionanocomposites showed a maximum of tensile

strength 76 MPa and elastic modulus 10.7 GPa at 50 % RH and 23 °C.69 Differences at

the same MTM content may be due to MTM orientation differences and polymer-MTM

adhesion differences. For oxidized XG/MTM, the elastic modulus is twice as high at 50 %

RH and is 30 GPa, whereas the tensile strength is increased to 160 MPa. The high

mechanical performance of oxidized XG/MTM nanocomposites is related to the

Page 60: Xyloglucan-based polymers and nanocomposites

50

formation of covalent cross-links as discussed in section 2.3.2. Since the cross-linking

sites are pre-formed on the polymer, the cross-linking is probably homogeneous

throughout the composite. Podsiadlo et.al. added gluteraldehyde to the PVOH/MTM

composites for crosslinking purposes.72 The gluteraldehyde needs to diffuse into the

structure, and as a consequence, the method was inefficient for thicker structures.67

Figure 3.19. Typical stress- strain curves of nacre-mimetic XG/MTM films conditioned at 50 % RH and 23 °C. The tensile strain of all samples, except XG was recorded using digital speckle photography. Designations ddXG10, ddXG20 and ddXG30, refer to the stoichiometric amount of periodate used during oxidation of XG.

The mechanical properties of the nanocomposites at 90 % RH are presented in Table

3.10. XG has shown an elastic modulus of 2.4 GPa at this humidity (section 3.1.2). With

ca. 45 vol% of inorganic content, the modulus is 3.6 GPa for nacre mimetic XG/MTM

nanocomposites at this humidity condition. Thus at high humidity, the modulus of the

nanocomposites is significantly reduced. The presence of moisture is likely to reduce

XG/MTM interfacial adhesion so that stress transfer at the interface is impaired and this

leads to reduced modulus. At similar humidity condition, oxidized XG/MTM shows a

modulus as high as 20 GPa, a nearly 6-fold increase compared to native XG/MTM

nanocomposites. This indicates that the oxidized XG is covalently linked to the MTM as

reported for nacre-mimetic PVOH/MTM nanocomposites cross-linked with

gluteraldehyde.72

Page 61: Xyloglucan-based polymers and nanocomposites

51

Table 3.10. Tensile properties of nacre-mimetic XG/MTM nanocomposites. Designations ddXG10, ddXG20 and ddXG30, refer to the stoichiometric amount of periodate used during oxidation of XG.

Samples conditioned at 50 % RH and 23 °C

Sample Volume fraction of MTM (%)

Elastic modulus,

GPa

Tensile strength,

MPa

Tensile strain to

failure, %

Xyloglucan - 5.4 ± 0.1 76.0 ± 4.4 10.7 ± 2.8

Nacre XG/MTM 44 13.7 ± 1.1 100.7 ± 3.6 1.1 ± 0.3

Nacre ddXG10/MTM 44 27.2 ± 0.4 140.6 ± 3.1 0.6 ± 0.1

Nacre ddXG20/MTM 45 30.1 ± 0.3 147.5 ± 5.7 0.5 ± 0.0

Nacre ddXG30/MTM 47 30.0 ± 0.4 158.8 ± 3.0 0.5 ± 0.0

Chitosan/MTM69 45* 10.7 ± 1.7 76 ± 10.0 0.9 ± 0.1

Samples conditioned at 90 % RH and 23 °C

Sample Volume fraction of MTM (%)

Elastic modulus,

GPa

Tensile strength,

MPa

Tensile strain to

failure, %

Nacre XG/MTM 44 3.6 ± 0.5 23.7 ± 3.4 5.8 ± 0.6

Nacre ddXG10/MTM 44 8.4 ± 2.5 51.1 ± 5.0 2.0 ± 0.4

Nacre ddXG20/MTM 45 20.3 ± 1.8 81.9 ± 12.9 0.8 ± 0.3

* Density of MTM = 2 g/cm3 and chitosan = 0.9 g/cm3

The oxygen permeability data for nacre-mimetic nanocomposites are presented in Table

3.11. Interestingly, the effect of moisture uptake on oxygen transmission is smaller

compared to the effect on mechanical properties. In dry state, the OTR values are less

than the detectable limit of the instrument, 0.008 cc/[m2-day]. At 80 % RH (Table 3.11),

the oxygen transmission rate of nanocomposites made from native XG and oxidized XG

are nearly in the same range (1 – 2 cc/[m2-day]), indicating the importance of in-plane

orientation of platelets. As discussed previously, the degree of in-plane orientation of

XG/MTM and oxidized XG/MTM nanocomposites are in the same range. The in-plane

orientation of MTM creates a longer tortuous path for the diffusion of oxygen molecules

Page 62: Xyloglucan-based polymers and nanocomposites

52

through the membrane.85 In food packaging applications, mechanical and barrier

performance in moist environments are critical. The oxygen barrier performances of the

present nanocomposites combined with favourable mechanical properties make these

materials a promising “green” alternative to conventional barriers.

Table 3.11. Oxygen barrier properties of nacre-mimetic XG/MTM nanocomposites. Designations ddXG10, ddXG20 and ddXG30, refer to the stoichiometric amount of periodate used during oxidation of XG.

Sample Oxygen transmission

rate, cc/[m2-day]

(80 % RH, 23 °C)

Oxygen permeability

cc.µm/[m2-day]kPa

(80 % RH, 23 °C)

XG 40.0 11.5

Nacre XG/MTM 2.0 0.58

Nacre ddXG10/MTM 2.24 0.72

Nacre ddXG20/MTM 1.86 0.65

Nacre ddXG30/MTM 1.13 0.37

Page 63: Xyloglucan-based polymers and nanocomposites

53

4.CONCLUSIONSIn the primary plant cell wall, the role of xyloglucan (XG) is to link cellulose

microfibrils so that a stiff network with substantial mechanical function is formed. This

function of XG in nature should be kept in mind when we attempt to use it in industrial

applications. The XG studied in the present thesis has a different origin. It is a storage

polysaccharide in tamarind seeds. However, its strong adsorption to cellulose, the subject

of many previous studies, shows that the basic characteristics of tamarind seed XG are

similar to those of plant cell wall XG. Furthermore, tamarind XG has a very high molar

mass, and this is advantageous from the point of view of mechanical properties.

The study shows that XG films cast from water are highly thermostable with a glass

transition temperature (Tg) of ~260 °C, and also have strong resistance to thermal

degradation. Unlike other polysaccharides, such as starch, XG is less sensitive to moisture

and XG films tend to be more ductile. XG films have a modulus as high as 4 GPa, a

tensile strength of 80 MPa, and a strain to failure of 4 – 5 % at room temperature and 50

% RH. The key structural reason may be that XG has a rigid cellulose main chain with

moderate amphiphilic characteristics. XG is “soluble” in water, but individual chains are

less hydrated and can also form local aggregates in the form of ordered domains of

parallel stacked chains. The strong intermolecular interaction explains also the highly

favourable oxygen barrier performance of XG films.

A larger part of the thesis is devoted to address two major problems. One is the high Tg

of XG, which is a disadvantage if thermal processing (i.e. thermoforming) is considered.

The other problem is to preserve the properties also under humid conditions.

Tg of XG is reduced by two routes, the first is plasticization where sorbitol can be used

to reduce Tg by about 130 °C and preserve fairly good mechanical properties. The other

approach is regioselective chemical modification of side chains, where periodate

oxidation is combined with subsequent reduction. This creates a form of internal

Page 64: Xyloglucan-based polymers and nanocomposites

54

plasticization. The cellulosic character of the main chain is preserved. An added benefit to

Tg reduction is that the barrier properties at 50 % RH are improved.

XG by itself preserves mechanical and barrier properties surprisingly well at elevated

relative humidity. The lower solubility of XG in water is one explanation. In addition,

even at moisture content as high as 40 %, XG is shown to preserve its integrity and it has

substantial strength and stiffness. The properties at high humidity are further enhanced by

combining XG with MTM clay platelets aiming for biocomposite substitution of the

traditional oxygen barriers in packaging. An industrially scalable continuous coating

process is used, and the clay platelets are strongly oriented. Both mechanical properties

and barrier performance are exceptionally good for these nanocomposites, which have a

clay content of around 12 percent by volume. The performance is further improved in a

study of “nacre-mimetic” composites based on 45 percent by volume of clay. Here the

XG is adsorbed to the clay platelets in aqueous colloidal suspension and excess polymer

is removed. It then becomes possible to run “paper-making” processes with filtration of

the aqueous phase. Crosslinking of the XG based on the previously developed

regioselective periodate oxidation procedure is performed. It has a very strong effect on

the mechanical performance and moisture durability of the material. The elastic moduli of

the cross-linked composites are as high as 30 GPa at 50 % RH and 23 ºC. Most likely, this

is related to improved XG-MTM interfacial adhesion and stress transfer under moist

conditions.

Future work on hemicelluloses for coating and film applications is encouraged.

Modification of hemicelluloses by chemical or enzymatic routes and their composites can

improve the performance under moist conditions. Clay nanocomposites are of

considerable interest, since barrier performance can be combined with good mechanical

properties also in humid conditions. In summary, XG is a hemicellulose with particularly

good properties for the applications considered, due to its cellulosic main chain and

limited water solubility.

Page 65: Xyloglucan-based polymers and nanocomposites

55

5.ACKNOWLEDGEMENTSI gratefully acknowledge Prof. Lars Berglund for giving me this opportunity to work in

Biocomposite group, and for his enthusiastic encouragement and useful critiques during

the course of this work. I am particularly indebted to him for suggesting a 'spicy' material

for my thesis work! I also thank my co-supervisor Assoc. Prof. Qi Zhou for his support

and fruitful discussions at every stage of the present work. All my co-authors are

thankfully acknowledged for the wonderful collaborative efforts.

I am also grateful to all former and present colleagues in Biocomposite group, and

Wallenberg Wood Science Centre for their invaluable help and support. The kind help

from Ramiro in scrutinizing thesis draft and invaluable discussions with Ivon, Gisela,

Jakob, Martin, Bipinbal, Per, Ingela, and Ngesa in lab work or data interpretation are

specially acknowledged. Jakob, Seira and Simon are acknowledged for providing the clay

images included in the thesis. The discussions with Michael Holmgren of Nuclear

Chemistry on XRD were really enlightening. Thanks are due to colleagues and friends at

FPT and the help from Brita, Mona, Inga, and Mia in many practical issues are

particularly acknowledged. I would also like to thank many teachers, especially Prof. Lars

Wågberg, and Prof. Ulf Gedde at FPT, Assoc. Prof. Andreas Fischer at Oorganisk Kemi,

Assoc. Prof. Magnus Bergström at Yki, Dr. Lennart Salmén at Innventia, and Prof.

Morooka of Kyoto University, Japan, for many invaluable discussions. Prof. Lars

Wågberg is specially thanked for reviewing the thesis.

My due respect to former teachers and gurus in moulding my personality and research

interests, especially mentioned are Paul Maliekal and late K. V. George.

I am also grateful to Vinnova Center of Excellence program BiMac Innovation for their

generous financial support. The encouraging stance of BiMac Director, Prof. Tom

Lindström, on many research projects was greatly appreciated. I also thank Prof. Kristofer

Gamstedt for being my mentor in BiMac and Biocomposite group during the early years

of my thesis work.

Page 66: Xyloglucan-based polymers and nanocomposites

56

Swedish Centre for Biomimetic Fiber Engineering (BioMime) and Wallenberg Wood

Science Centre (WWSC) are thanked for providing lab facilities during the course of this

work. The laboratory access from industrial collaborators Innventia, Tetrapak, and Stora

Enso are also thankfully acknowledged.

Last, but not least, I would like to thank my family for their patience and support that

they have shown during this time.

Page 67: Xyloglucan-based polymers and nanocomposites

57

6.REFERENCES1. Gandini, A., Polymers from Renewable Resources: A Challenge for the Future of Macromolecular Materials. Macromolecules 2008, 41, (24), 9491-9504. 2. Johansson, C.; Bras, J.; Mondragon, I.; Nechita, P.; Plackett, D.; Simon, P.; Svetec, D. G.; Virtanen, S.; Baschetti, M. G.; Breen, C.; Clegg, F.; Aucejo, S., Renewable fibers and bio-based materials for packaging applications - A review of recent developments. Bioresources 2012, 7, (2), 2506-2552. 3. Klemm, D.; Kramer, F.; Moritz, S.; Lindström, T.; Ankerfors, M.; Gray, D.; Dorris, A., Nanocelluloses: A New Family of Nature‐Based Materials. Angewandte Chemie International Edition 2011, 50, (24), 5438-5466. 4. Fukuzumi, H.; Saito, T.; Wata, T.; Kumamoto, Y.; Isogai, A., Transparent and High Gas Barrier Films of Cellulose Nanofibers Prepared by TEMPO-Mediated Oxidation. Biomacromolecules 2009, 10, (1), 162-165. 5. Berglund, L. A.; Peijs, T., Cellulose Biocomposites—From Bulk Moldings to Nanostructured Systems. MRS Bulletin 2010, 35, (03), 201-207. 6. Henriksson, M.; Berglund, L. A.; Isaksson, P.; Lindstro m, T.; Nishino, T., Cellulose nanopaper structures of high toughness. Biomacromolecules 2008, 9, (6), 1579-1585. 7. Hansen, N. M. L.; Plackett, D., Sustainable films and coatings from hemicelluloses: A review. Biomacromolecules 2008, 9, (6), 1493-1505. 8. Hartman, J.; Albertsson, A. C.; Sjoberg, J., Surface- and bulk-modified galactoglucomannan hemicellulose films and film laminates for versatile oxygen barriers. Biomacromolecules 2006, 7, (6), 1983-1989. 9. Edlund, U.; Ryberg, Y. Z.; Albertsson, A. C., Barrier Films from Renewable Forestry Waste. Biomacromolecules 2010, 11, (9), 2532-2538. 10. Myllarinen, P.; Partanen, R.; Seppala, J.; Forssell, P., Effect of glycerol on behaviour of amylose and amylopectin films. Carbohydrate Polymers 2002, 50, (4), 355-361. 11. Moad, G., Chemical modification of starch by reactive extrusion. Progress in Polymer Science 2011, 36, (2), 218-237. 12. Chivrac, F.; Pollet, E.; Averous, L., Progress in nano-biocomposites based on polysaccharides and nanoclays. Materials Science & Engineering R-Reports 2009, 67, (1), 1-17. 13. Sinha Ray, S.; Bousmina, M., Biodegradable polymers and their layered silicate nanocomposites: in greening the 21st century materials world. Progress in Materials Science 2005, 50, (8), 962-1079. 14. Cosgrove, D. J., Growth of the plant cell wall. Nature Reviews Molecular Cell Biology 2005, 6, (11), 850-861. 15. Chanliaud, E.; Burrows, K. M.; Jeronimidis, G.; Gidley, M. J., Mechanical properties of primary plant cell wall analogues. Planta 2002, 215, (6), 989-996. 16. Yamatoya, K.; Shirakawa, M., Xyloglucan: Structure, rheological properties, biological functions and enzymatic modification. Curr. Trends Polym. Sci 2003, 8, 27-71. 17. Urakawa, H.; Mimura, M.; Kajiwara, K., Diversity and versatility of plant seed xyloglucan. Trends in Glycoscience and Glycotechnology 2002, 14, (80), 355-376. 18. Kumar, C. S.; Bhattacharya, S., Tamarind seed: properties, processing and utilization. Critical Reviews in Food Science and Nutrition 2008, 48, (1), 1-20. 19. Mishra, A.; Malhotra, A. V., Tamarind xyloglucan: a polysaccharide with versatile application potential. Journal of Materials Chemistry 2009, 19, (45), 8528-8536.

Page 68: Xyloglucan-based polymers and nanocomposites

58

20. Savur, G. R.; Sreenivasan, A., Isolation and characterization of tamarind seed (tamarindus indica L.) polysaccharide. The Journal of Biological Chemistry 1947, 11, 501- 509. 21. Teraoka, T.; Iriguchi, A.; Ebie, K.; Mizuhashi, N. Process for separating polysaccharides from tamarind seeds. 1990. 22. Jones, D. A.; Jordan, W. A. Purification of tamarind gum. 1978. 23. Rao, P.; Srivastava, H., Tamarind. RL Whistler, Academic Press, 2nd Ed, New York 1973, 369-411. 24. Edwards, M.; Dea, I. C. M.; Bulpin, P. V.; Reid, J. S. G., Xyloglucan (amyloid) mobilisation in the cotyledons of Tropaeolum-majus L. seeds following germination. Planta 1985, 163, (1), 133-140. 25. Fry, S. C., The structure and functions of xyloglucan. Journal of Experimental Botany 1989, 40, (1), 1. 26. Gidley, M. J.; Lillford, P. J.; Rowlands, D. W.; Lang, P.; Dentini, M.; Crescenzi, V.; Edwards, M.; Fanutti, C.; Reid, J. S. G., Structure and solution properties of tamarind-seed polysaccharide. Carbohydrate Research 1991, 214, (2), 299-314. 27. Hayashi, T.; Kaida, R., Functions of xyloglucan in plant cells. Molecular Plant 2011, 4, (1), 17-24. 28. Picout, D. R.; Ross-Murphy, S. B.; Errington, N.; Harding, S. E., Pressure cell assisted solubilization of xyloglucans: Tamarind seed polysaccharide and detarium gum. Biomacromolecules 2003, 4, (3), 799-807. 29. Lang, P.; Kajiwara, K., Investigations of the architecture of tamarind seed polysaccharide in aqueous solution by different scattering techniques. Journal of Biomaterials Science, Polymer Edition 1993, 4, (5), 517-528. 30. Muller, F.; Manet, S.; Jean, B.; Chambat, G.; Bou , F.; Heux, L.; Cousin, F., SANS Measurements of Semi-Flexible Xyloglucan Polysaccharide Chains in Water Reveal Their Self-avoiding Statistics. Biomacromolecules 2011, 12, (9), 3330–3336. 31. Fry, S. C.; York, W. S.; Albersheim, P.; Darvill, A.; Hayashi, T.; Joseleau, J. P.; Kato, Y.; Lorences, E. P.; Maclachlan, G. A.; McNeil, M., An unambiguous nomenclature for xyloglucan derived oligosaccharides. Physiologia Plantarum 1993, 89, (1), 1-3. 32. Zhou, Q.; Rutland, M. W.; Teeri, T. T.; Brumer, H., Xyloglucan in cellulose modification. Cellulose 2007, 14, 625-641. 33. Edwards, M.; Bowman, Y.; Dea, I.; Reid, J., A beta-D-galactosidase from nasturtium (Tropaeolum majus L.) cotyledons. Purification, properties, and demonstration that xyloglucan is the natural substrate. Journal of Biological Chemistry 1988, 263, (9), 4333-4337. 34. Shirakawa, M.; Yamatoya, K.; Nishinari, K., Tailoring of xyloglucan properties using an enzyme. Food Hydrocolloids 1998, 12, (1), 25-28. 35. Whitney, S. E. C.; Wilson, E.; Webster, J.; Bacic, A.; Reid, J. S. G.; Gidley, M. J., Effects of structural variation in xyloglucan polymers on interactions with bacterial cellulose. American journal of botany 2006, 93, (10), 1402-1414. 36. Lang, P.; Kajiwara, K.; Burchard, W., Investigations on the Solution Architecture of Carboxylated Tamarind Seed Polysaccharide by Static and Dynamic Light Scattering. Macromolecules 1993, 26, (15), 3992-3998. 37. Lang, P.; Masci, G.; Dentini, M.; Crescenzi, V.; Cooke, D.; Gidley, M.; Fanutti, C.; Reid, J., Tamarind seed polysaccharide: preparation, characterisation and solution properties of carboxylated, sulphated and alkylaminated derivatives. Carbohydrate Polymers 1992, 17, (3), 185-198. 38. Fang, J.; Fowler, P.; Tomkinson, J.; Hill, C., The preparation and characterisation of a series of chemically modified potato starches. Carbohydrate Polymers 2002, 47, (3), 245-252. 39. Singh, J.; Kaur, L.; McCarthy, O., Factors influencing the physico-chemical, morphological, thermal and rheological properties of some chemically modified starches for food applications—A review. Food Hydrocolloids 2007, 21, (1), 1-22.

Page 69: Xyloglucan-based polymers and nanocomposites

59

40. Bemiller, J. N., Starch modification: challenges and prospects. Starch‐Stärke 2006, 49, (4), 127-131. 41. Cunha, A. G.; Gandini, A., Turning polysaccharides into hydrophobic materials: a critical review. Part 1. Cellulose. Cellulose 2010, 17, (5), 875-889. 42. Nachtergaele, W., The benefits of cationic starches for the paper industry. Starch‐Stärke 1989, 41, (1), 27-31. 43. Brody, A. L.; Bugusu, B.; Han, J. H.; Sand, C. K.; McHugh, T. H., Innovative food packaging solutions. Journal of Food Science 2008, 73, (8), R107-R116. 44. Leterrier, Y., Durability of nanosized oxygen-barrier coatings on polymers - Internal stresses. Progress in Materials Science 2003, 48, (1), 1-55. 45. Strupinsky, G.; Brody, A. L. In A Twenty-Year Retrospective on Plastics: Oxygen Barrier Packaging Materials, 1998 TAPPI Proceedings: Ploymers, Laminations & Coatings Conference, San Francisco, CA, 1998; TAPPI Press: San Francisco, CA, 1998; pp 119-140. 46. Lange, J.; Wyser, Y., Recent innovations in barrier technologies for plastic packaging—a review. Packaging Technology and Science 2003, 16, (4), 149-158. 47. Miller, K. S.; Krochta, J. M., Oxygen and aroma barrier properties of edible films: A review. Trends in Food Science & Technology 1997, 8, (7), 228-237. 48. Fujita, H., Diffusion in polymer-diluent systems. Fortschritte der Hochpolymeren-Forschung 1961, 1-47. 49. Brandt, W. W., Model calculation of the temperature dependence of small molecule diffusion in high polymers. The Journal of Physical Chemistry 1959, 63, (7), 1080-1085. 50. Klopffer, M.; Flaconneche, B., Transport Properdines of Gases in Polymers: Bibliographic Review. Oil & Gas Science and Technology 2001, 56, (3), 223-244. 51. Petersen, K.; Nielsen, P. V.; Bertelsen, G.; Lawther, M.; Olsen, M. B.; Nilsson, N. H.; Mortensen, G., Potential of biobased materials for food packaging. Trends in Food Science & Technology 1999, 10, (2), 52-68. 52. Hartman, J.; Albertsson, A. C.; Lindblad, M. S.; Sjoberg, J., Oxygen barrier materials from renewable sources: Material properties of softwood hemicellulose-based films. Journal of Applied Polymer Science 2006, 100, (4), 2985-2991. 53. Okada, A.; Usuki, A., Twenty years of polymer-clay nanocomposites. Macromolecular Materials and Engineering 2006, 291, (12), 1449-1476. 54. Usuki, A.; Kojima, Y.; Kawasumi, M.; Okada, A.; Fukushima, Y.; Kurauchi, T.; Kamigaito, O., Synthesis of Nylon 6-Clay hybrid. Journal of Materials Research 1993, 8, (5), 1179-1184. 55. Pavlidou, S.; Papaspyrides, C., A review on polymer-layered silicate nanocomposites. Progress in Polymer Science 2008, 33, (12), 1119-1198. 56. Ray, S.; Okamoto, M., Polymer/layered silicate nanocomposites: a review from preparation to processing. Progress in Polymer Science 2003, 28, (11), 1539-1641. 57. Giannelis, E. P., Polymer layered silicate nanocomposites. Advanced Materials 1996, 8, (1), 29-35. 58. Silvestre, C.; Duraccio, D.; Cimmino, S., Food packaging based on polymer nanomaterials. Progress in Polymer Science 2011, 36, (12), 1766–1782. 59. Bharadwaj, R. K., Modeling the barrier properties of polymer-layered silicate nanocomposites. Macromolecules 2001, 34, (26), 9189-9192. 60. Bergaya, F.; Theng, B. K. G., Handbook of clay science. Elsevier Science: 2006; Vol. 1. 61. Hofmann, U.; Endell, K.; Wilm, D., Röntgenographische und kolloidchemische Untersuchungen über Ton. Angewandte Chemie 2006, 47, (30), 539-547. 62. Holmboe, M. The Bentonite Barrier : Microstructural properties and the influence of γ-radiation. PhD Thesis, KTH Royal Institute of Technology, Stockholm, 2011. 63. Tournassat, C.; Ferrage, E.; Poinsignon, C.; Charlet, L., The titration of clay minerals: II. Structure-based model and implications for clay reactivity. Journal of Colloid and Interface Science 2004, 273, (1), 234-246.

Page 70: Xyloglucan-based polymers and nanocomposites

60

64. Dzenis, Y. A., Structural nanocomposites. Science 2008, 319, 419-420. 65. Mayer, G., Rigid Biological Systems as Models for Synthetic Composites. Science 2005, 310, (5751), 1144-1147. 66. Tang, Z. Y.; Kotov, N. A.; Magonov, S.; Ozturk, B., Nanostructured artificial nacre. Nature Materials 2003, 2, (6), 413-U8. 67. Walther, A.; Bjurhager, I.; Malho, J. M.; Pere, J.; Ruokolainen, J.; Berglund, L. A.; Ikkala, O., Large-Area, Lightweight and Thick Biomimetic Composites with Superior Material Properties via Fast, Economic, and Green Pathways. Nano Letters 2010, 10, (8), 2742-2748. 68. Walther, A.; Bjurhager, I.; Malho, J.-M.; Ruokolainen, J.; Berglund, L.; Ikkala, O., Supramolecular Control of Stiffness and Strength in Lightweight High-Performance Nacre-Mimetic Paper with Fire-Shielding Properties. Angewandte Chemie International Edition 2010, 49, (36), 6448-6453. 69. Yao, H. B.; Tan, Z. H.; Fang, H. Y.; Yu, S. H., Artificial Nacre‐like Bionanocomposite Films from the Self‐Assembly of Chitosan–Montmorillonite Hybrid Building Blocks. Angewandte Chemie International Edition 2010, 49, (52), 10127-10131. 70. Liu, A.; Walther, A.; Ikkala, O.; Belova, L.; Berglund, L. A., Clay Nanopaper with Tough Cellulose Nanofiber Matrix for Fire Retardancy and Gas Barrier Functions. Biomacromolecules 2011, 12, (3), 633-641. 71. Liu, A.; Berglund, L. A., Clay nanopaper composites of nacre-like structure based on montmorrilonite and cellulose nanofibers—Improvements due to chitosan addition. Carbohydrate Polymers 2012, 87, (1), 53-60. 72. Podsiadlo, P.; Kaushik, A. K.; Arruda, E. M.; Waas, A. M.; Shim, B. S.; Xu, J.; Nandivada, H.; Pumplin, B. G.; Lahann, J.; Ramamoorthy, A., Ultrastrong and stiff layered polymer nanocomposites. Science 2007, 318, (5847), 80. 73. Jackson, E. L.; Hudson, C. S., Application of the Cleavage Type of Oxidation by Periodic Acid to Starch and Cellulose1. Journal of the American Chemical Society 1937, 59, (10), 2049-2050. 74. Kim, U. J.; Kuga, S.; Wada, M.; Okano, T.; Kondo, T., Periodate oxidation of crystalline cellulose. Biomacromolecules 2000, 1, (3), 488-492. 75. Morooka, T.; Norimoto, M.; Yamada, T., Periodate oxidation of cellulose by homogeneous reaction. Journal of Applied Polymer Science 1989, 38, (5), 849-858. 76. Back, E. L., Thermal auto-crosslinking in cellulose material. Pulp and paper magazine of Canada 1967, 68, (4), 165-171. 77. Larsson, P. A.; Gimåker, M.; Wågberg, L., The influence of periodate oxidation on the moisture sorptivity and dimensional stability of paper. Cellulose 2008, 15, (6), 837-847. 78. Marx, K. A., Quartz crystal microbalance: a useful tool for studying thin polymer films and complex biomolecular systems at the solution-surface interface. Biomacromolecules 2003, 4, (5), 1099-1120. 79. Sauerbrey, G., Verwendung von Schwingquarzen zur Wägung dünner Schichten und zur Mikrowägung. Zeitschrift für Physik A Hadrons and Nuclei 1959, 155, (2), 206-222. 80. ASTM, E., 01 (2003) Determination of carbohydrates in biomass by high performance liquid chromatography. Annual Book of ASTM Standards 2003, 11. 81. Giannelis, E. P., Polymer layered silicate nanocomposites. Advanced Materials 1996, 8, (1), 29-&. 82. Ray, S. S.; Bousmina, M., Biodegradable polymers and their layered silicate nano composites: In greening the 21st century materials world. Progress in Materials Science 2005, 50, (8), 962-1079. 83. Comyn, J., Polymer permeability. Elsevier Applied Science, London: 1985. 84. Lange, J.; Stenroos, E.; Johansson, M.; Malmstrom, E., Barrier coatings for flexible packaging based on hyperbranched resins. Polymer 2001, 42, (17), 7403-7410. 85. Nielsen, L. E., Models for the permeability of filled polymer systems. Journal of Macromolecular science (Chem.) 1967, A1, (5), 929-942.

Page 71: Xyloglucan-based polymers and nanocomposites

61

86. Grondahl, M.; Eriksson, L.; Gatenholm, P., Material properties of plasticized hardwood Xylans for potential application as oxygen barrier films. Biomacromolecules 2004, 5, (4), 1528-1535. 87. Höije, A.; Gröndahl, M.; Tømmeraas, K.; Gatenholm, P., Isolation and characterization of physicochemical and material properties of arabinoxylans from barley husks. Carbohydrate Polymers 2005, 61, (3), 266-275. 88. Jarowenko, W., Starch. Encyclopedia of Polymer Science and Technology 1970, 12, 787-862. 89. Lindman, B.; Karlström, G.; Stigsson, L., On the mechanism of dissolution of cellulose. Journal of Molecular Liquids 2010, 156, (1), 76-81. 90. Glasser, W. G.; Atalla, R. H.; Blackwell, J.; Malcolm Brown, R.; Burchard, W.; French, A. D.; Klemm, D. O.; Nishiyama, Y., About the structure of cellulose: debating the Lindman hypothesis. Cellulose 2012, 1-10. 91. Sudhamani, S.; Raj, A. E.; Raj, B.; Prasad, M., Moisture Sorption Characteristics of Microbially Produced Polysaccharide and Polyvinyl Alcohol Blends. Journal of Polymers and the Environment 2005, 13, (3), 253-260. 92. Szcześniak, L.; Rachocki, A.; Tritt-Goc, J., Glass transition temperature and thermal decomposition of cellulose powder. Cellulose 2008, 15, (3), 445-451. 93. Orford, P. D.; Parker, R.; Ring, S. G.; Smith, A. C., Effect of water as a diluent on the glass transition behaviour of malto-oligosaccharides, amylose and amylopectin. 1989, 11, (2), 91-96. 94. Alen, R.; Kuoppala, E.; Oesch, P., Formation of the main degradation compound groups from wood and its components during pyrolysis. Journal of Analytical and Applied Pyrolysis 1996, 36, (2), 137-148. 95. Yang, H.; Yan, R.; Chen, H.; Lee, D. H.; Zheng, C., Characteristics of hemicellulose, cellulose and lignin pyrolysis. Fuel 2007, 86, (12-13), 1781-1788. 96. Ashby, M. F.; Shercliff, H.; Cebon, D., Materials: engineering, science, processing and design. Butterworth-Heinemann: 2009. 97. Rindlav-Westling, A.; Stading, M.; Hermansson, A. M.; Gatenholm, P., Structure, mechanical and barrier properties of amylose and amylopectin films. Carbohydrate Polymers 1998, 36, (2-3), 217-224. 98. Painter, T. J., Control of depolymerisation during the preparation of reduced dialdehyde cellulose. Carbohydrate Research 1988, 179, 259-268. 99. Kararli, T. T.; Hurlbut, J. B.; Needham, T. E., Glass–rubber transitions of cellulosic polymers by dynamic mechanical analysis. Journal of Pharmaceutical Sciences 1990, 79, (9), 845-848. 100. Theng, B., Clay-polymer interactions; summary and perspectives. Clays and Clay Minerals 1982, 30, (1), 1-10. 101. Greenland, D., Adsorption of polyvinyl alcohols by montmorillonite. Journal of Colloid Science 1963, 18, (7), 647-664. 102. Anthony, J. W.; Bideaux, R.; Bladh, K.; Nichols, M., Handbook of mineralogy. Mineral Data Publishing Tucson,, Arizona: 1990; Vol. 1. 103. Rendleman, J., Metal-polysaccharide complexes—Part I. Food Chemistry 1978, 3, (1), 47-79.