185
FUNDAMENTALS AND CHARACTERIZATION OF FUNGALLY MODIFIED POLYSACCHARIDES FOR THE PRODUCTION OF BIO-PLASTICS by Arturo Rodriguez Uribe A thesis submitted in conformity with the requirements for the degree of Doctor in Philosophy Graduate Department of Forestry University of Toronto © Copyright by Arturo Rodriguez 2010

FUNDAMENTALS AND CHARACTERIZATION OF · PDF fileFUNDAMENTALS AND CHARACTERIZATION OF FUNGALLY MODIFIED POLYSACCHARIDES FOR THE PRODUCTION OF ... Objectives and approach ... Oligo and

  • Upload
    lambao

  • View
    220

  • Download
    2

Embed Size (px)

Citation preview

FUNDAMENTALS AND CHARACTERIZATION OF FUNGALLY

MODIFIED POLYSACCHARIDES FOR THE PRODUCTION OF

BIO-PLASTICS

by

Arturo Rodriguez Uribe

A thesis submitted in conformity with the requirements for the

degree of Doctor in Philosophy

Graduate Department of Forestry

University of Toronto

© Copyright by Arturo Rodriguez

2010

FUNDAMENTALS AND CHARACTERIZATION OF FUNGALLY

MODIFIED POLYSACCHARIDES FOR THE PRODUCTION OF

BIO-PLASTICS

Arturo Rodriguez Uribe

Ph D thesis, Graduate Department of Forestry

University of Toronto, 2010

Abstract

Starch and microbial exo-polysaccharides produced by prokaryotes (i.e. Eubacteria

and Archaebacteria) and eukaryotes (i.e. phytoplankton, fungi, and algae) are

recognized as a permanent source of biopolymers for the packaging industry.

However, the unsuitable mechanical properties for thermoplastic applications and/or

high cost of production have restricted their generalized use.

Fungal isolates of the genus Ophiostoma are able to produce exo-polysaccharides or

protein-like compounds in a medium containing starch as the substrate. Various

analytical techniques were used as an approach to investigate the interaction

between starch and the fungal extracellular metabolites and the effect of the

molecular-structural modifications on the functional properties of the materials.

Native starches were used as control in all experiments.

Analyses performed by dynamic mechanical thermal analysis (DMTA), which provides

information related to the viscoelastic properties, showed that the storage modulus

(E') increased substantially after the modification of the starch showing a process of

chain stiffness. The determination of the glass transition temperature (Tg) by tan

and loss modulus (E'') peaks showed various thermal transitions indicating a complex

molecular aggregation due to the potential presence of dissimilar amorphous

polymers. Experiments performed in DSC confirmed the presence of the various

thermal transitions associated to the Tg of these materials. The first derivative of

mass loss with respect to temperature during the thermogravimetric (TG) analysis

was slightly lower compared with native starches (at ~630 and 650°C). However,

II

modified starches can withstand high temperatures showing residues up to 20% at

1000°C.

Studies on the characterization of the flow properties of the polymers by capillary

rheology showed in both samples a shear thinning behavior. The double logarithmic

plot of the shear rate vs. shear viscosity produced a straight line and in consequence

a power law equation was used to describe the rheological behavior (( = K'n). The

results showed that in order to achieve the same shear rate (') in both samples

(modified and native starches) it is necessary to apply a higher shear stress () in

the fungal treated materials. As a result, the consistency power law index (n)

decreased and the consistency value increased (K). The practical consequence is that

the melting point of these polysaccharides shifted to higher temperatures.

By using various analytical techniques (including chromatography, spectroscopy,

spectrometry) it was found that these phenomena may be due to the interaction of

starch with protein-like or exo-polysaccharides or both which may influence the

viscosity, bind adjacent molecules (i.e. network-like) and restrict the molecular

motion. Evidences of the presence of pendant groups attached to high molecular

weight compounds were also found. This information will give guidance to further

structural studies and it is intended to pave the way for a variety of industrial

applications.

III

Acknowledgment

This thesis could not have been completed without the constant encouragement, advice,

financial support, and unconditional help of my supervisor Dr. M. Sain from which I am

very greatly indebted.

I would also like to thank to Mexican Council of Science and Technology and BIOCAR

Canada Foundation and an NSERC strategic grant for financial support.

I am grateful to my advisory committee members: Dr. D. N. Roy, Dr. Martin Hubbes, Dr.

Charles Q. Jia, and Dr. Sally Krigstin for all support and invaluable comments to the

drafts of this thesis.

I would also like to thank the professors who shaped my academy formation: Dr. Ning

Yan, Dr. Sandy Smith, Dr. Andy Kenney, Dr. T. Blake. I am especially grateful to Dr.

Robert Jeng for his support in laboratory matters.

Also my especial thanks to Deborah Paes, because not everything can be solved in the

laboratory. Thanks to the University of Toronto, great institutions are made of great

people. I would like to thank to all the people in the international school of graduate

studies (SGS). There were many people which directly or indirectly played important

roles in my personal and professional development as well as in the performance of my

experiments. Thanks to the Dean T. Smith, Mary-Rose Naudi, Ian Kennedy, John

McCarron, Shiang Law, Tony Ung, my classmates and coworkers and all the staff in the

laboratories.

This work is also dedicated with all my love to my wife Sofia Ocana Alonso and my

daughter Sofia Rodriguez Ocana. I need to thank all their support, comprehension, and

company during these years, both are a blessing for me.

IV

Table of contents

Abstract …………………………………………………………………………………………………………….……… ii

Acknowledgment ……………………………………………………………………………………………..……….iv

Table of contents………………………………………………………………………………………………………..v

List of figures …………………………………………………………………………………………………..……. iX

List of tables ……………………………………………………………………………………………………………xvii

List of abbreviations ………………………………………………………………………………………………xviii

List of symbols ………………………………………………………………………………………………………….xx

1. Introduction ……………………………… …………………………………………………………..…………..1

1.1. Motivation of the study ……………. ……………………………………………………………..…1

1.2. Molecular, physical, and functional properties of starch ………………………….. 4

1.2.1. Molecular structure of starch …………………………………………………..……… 4

1.2.1.1. Minor components of starch ……………………………………………...6

1.2.1.2. Comparison between cereal and tuber starches ………………...8

1.3. Exopolysaccharides and other fungal metabolites ……………………………..…...9

1.3.1. Microbial metabolites and other industrial uses ……………………..….….9

1.3.1.1. Enzymatic conversion of starch ……………………………………….… 9

1.3.1.2. Microbial metabolites: the case of Ophiostoma spp. ………..11

1.4. Objectives and approach ……………………………………………………………………… …19

1.4.1. Objectives ………………………………………………………………………………..19

1.4.2. Approach ……………………………………………………………………………………... 21

1.4.3. Structure of thesis ………………………………………………………………………..22

2. Experimental …………………………………………………………………………………………………….24

2.1. Production of polymers …………………………………………………………………………….24

2.2. Protein determination ……………………………………………………………………………….25

3. Results and analysis ………………………………………………………………………………………..26

3.1. Morphology and chemical analyses.………………………………………………………….26

3.1.1. Morphology (SEM and FT-Raman confocal analysis) ……………..…..26

3.1.1.1. Introduction …………………………………….…………………………………26

3.1.2. Materials and methods ………………………………………………………………….27

3.1.3. Results and discussion ………………………………………………………………….27

3.1.4. Conclusions to the section.…………………………………………………………… 32

3.2. XRD-analysis ………………………………………………………………………………………….32

3.2.1. Abstract …………………………………………………………………………………………32

V

3.2.2. Introduction ……………………………………………………………………………… .32

3.2.3. Materials and methods …………………………………………………………………..34

3.2.4. Results and discussion …………………………………………………………… .….34

3.2.5. Conclusions …………………………………………………………………………………...35

3.3. FT-IR (ATR) ……………………………………………………………………………………………..35

3.3.1. Abstract ……………………………..…………………………………………………………..35

3.3.2. Introduction ……………………………………..…………………………………………..36

3.3.3. Materials and methods ………………………..…………………………………………37

3.3.4. Results and discussion ……………………..………………………………….……….38

3.3.4.1. FTIR (ATR) ……………………………..………………………………………….38

3.3.4.2. FTIR (ATR) ……………………………………………………………………………43

3.3.5. Conclusions ………………………………………..……………………………………………46

3.4. FT-Raman ……………………………………………………………………….………………………….47

3.4.1. Abstract ………………………………………………………………………………………….47

3.4.2. Introduction …………………………………………………………………………………….48

3.4.3. Materials and methods ……………………………………………………………………48

3.4.4. Results and discussion ……………………………………………………………………49

3.4.5. Conclusions ……………………………………………………………………………………..58

3.5. Liquid state NMR……………………………………………………………………………………… ..59

3.5.1. Abstract …………………………………………………………………………………………..59

3.5.2. Introduction ……………………………………………………………………………………..59

3.5.3. Materials and methods ……………………………………………………………….… 60

3.5.4. Results and discussion ………………………………………………………………….. .60

3.5.5. Conclusions …………………………………………………………………………… …..….63

3.6. Solid state and liquid state NMR ……………………………………………………… ..…… 63

3.6.1. Abstract ……………………………………………………………………………………… ….63

3.6.2. Introduction ……………………………………………………………………………………..64

3.6.3. Materials and methods ……………………………………………………………… ….64

3.6.4. Results and discussion …………………………………………………………………….65

3.6.5. Conclusions …………………………………………………………………………………….68

3.7. MALDI-TOF MS ……………………………………………………………………………………… … 68

3.7.1. Abstract ……………………………………………………………………………………………68

3.7.2. Introduction …………………………………………………………………………… ………68

3.7.3. Materials and methods ………………………………………………………………….69

3.7.3.1. Sample preparation …………………………………………………………….69

VI

3.7.3.2. Instrumental conditions ……………………………………………… … 69

3.7.4. Results and discussion …………………………………………………………………..69

3.7.5. Conclusions …………………………………………………………………………………… 71

3.8. HPAEC-PAD …………………………………………………………………………………………….. 72

3.8.1. Abstract ……………………………….………………………………………………………..72

3.8.2. Introduction ………………………………………………………………………………….72

3.8.3. Materials and methods ……………………………………………………………….. .73

3.8.3.1. Polymer production and sample preparation …………………..73

3.8.3.2. Instrumental conditions …………………………………………………..74

3.8.4. Results and discussion ………………………………………………………………….75

3.8.4.1. Oligo and polysaccharides …………………………………………………75

3.8.4.2. Sugars- composition ………………………………………………………….80

3.8.5. Conclusions ……………………………………………………………………………………84

3.9. Viscoelastic properties ………………………………………………………………………………85

3.9.1. Abstract ……………………………………………………………………………………….…85

3.9.2. Introduction …………………………………………………………………………………. 86

3.9.2.1. Dynamic mechanical thermal analysis of polymers .………..86

3.9.2.2. Basic definitions……………………………………. ……………..…………86

3.9.2.3. DMTA basic principles……………….…………………………………………93

3.9.2.4. DMTA of starch …………………………………………………………………..93

3.9.3. Materials and methods ………………………….……………………………………..94

3.9.3.1. Formation of films by casting method…………………………….. .94

3.9.3.2. Extrusion of materials.………………………………………………………. 95

3.9.3.3. DMTA conditions …………………………..………………………………….. 96

3.9.4. Results and discussion…… ………………………………………………………….. 97

3.9.4.1. Samples produced by film casting method ……………………….97

3.9.4.1.1. Determination of the linear viscoelatic region (LVR)…. 97

3.9.4.1.2. Creep compliance test…………………………………………………100

3.9.4.2. DMTA-Samples produced by casting method…………………..103

3.9.4.3. DMTA- Samples produced by extrusion……………………………105

3.9.4.4. DMTA-starch/clay/glycerol samples ……….. …………………107

3.9.5. Conclusions

3.10. Thermal properties …………………………………………………………..111

3.10.1. TG (thermogravimetry).…………………………………………………… 111

3.10.1.1. Abstract …………………………………………………………………………… 111

VII

3.10.1.2. Introduction………………………………………………………………………. 111

3.10.1.3. Fundamentals of TG ………………………………………………………… 112

3.10.1.3.1. General reaction of decomposition.……………………….113

3.10.1.3.2. Definitions ………………………………………………………………114

3.10.1.3.3. Theory …………………………………………………………………….115

3.10.1.3.3.1. Reaction rate and extent of decomposition.116

3.10.1.3.3.2. Reaction rate and temperature ………………….120

3.10.1.4. Materials and methods ……………………………………………………. 120

3.10.1.5. Results and discussion……………………………………………………… 120

3.10.1.6. Conclusions ………………………………………………………….…………….125

3.10.2. DSC……………………………………………………………………………………..126

3.10.2.1. Abstract ………………………………………………………………………………126

3.10.2.2. Introduction ………………………………………………………………………..126

3.10.2.3. Materials and methods ……………………………………………………….129

3.10.2.4. Results and discussion ……………………………………………………….130

3.10.2.5. Conclusions ………………………………………………………………………..136

3.11. Rheology …………………………………………………………………………………………137

3.11.1. Introduction ……………………………………………………………………….137

3.11.2. Materials and methods ……………………………………………………….137

3.11.2.1. Sample preparation …………………………………………………………….137

3.11.2.2. Instrumental conditions ……………………………………………………. 138

3.11.3. Results and discussion …………………………………………………………138

3.11.4. Conclusions ……………………………………………………………………… 142

3.12. Mechanical properties ……………………………………………………………………..142

3.12.1. Abstract …………………………………………………………………………….. 142

3.12.2. Introduction …………………………………….………………………………….143

3.12.3. Materials and methods ……………………………………………………….150

3.12.3.1. Sample preparation ……………………………………………………………. 150

3.12.3.2. Extrusion ……………………………………………………………………………. 151

3.12.3.3. Injection molding …………………………………………………………………152

3.12.4. Results and discussion ……………………………….………………………152

3.12.5. Conclusions ……………………………………………………………..…….……154

4. General conclusions ………………………………………………………………………….…………….154

5. Future work ……………………………………………………………………………………………………….156

6. References ………………………………………………………………………………………….…………….157

VIII

List of Figures

Figure 1— Confocal FT-Raman microscope observations of a variety of used commercial potato starch; Scale= 10 micros Figure 2- SEM image showing the porosity at the surface of the granules Scale=6.1 m p……………………………………………………………………………………………………………………………………28

Figure 3- SEM micrographs showing granular aggregation in native starches Scale 20 m

Figure 4- Confocal microscope FT-Raman- amylose-iodine complexing denoting the amylose fractions within the granules-Scale 20 m

p…………………………………………………………………………………………29 Figure 5- Confocal microscope FT-Raman- amylose-iodine complexing denoting thick layers of amylose fractions within the granules-Scale 20 m

p. ………………………………………………………………………………………………………………………………….30

Figure 6- Optical images of modified starch granules

Figure 7-SEM images of modified starch granules

p…………………………………………………………………………………………………………….……………………..31

Figure 8--PXRD patterns of granular native starches (GNS), native gelatinized starches (NS), modified gelatinized starches (DMS), and granular modified starches (GMS)

p…………………………………………………………………………………………………………………………………….35

IX

Figure 9- FT-IR spectrum of modified starches- (detection of the peak associated to double bonds probably in C=O vibrations) p…………………………………………………………………………………………………………………………………….40

Figure 10- FT-IR spectrum of exopolysaccharides (EPSs) produced in absence

of substrate (detection of the peak associated to double bonds probably in

C=O vibrations)

Figure 11--FT-IR spectrum of granular modified starches (G-MS) (detection of the peaks at ~800 and 1240 cm-1)

p…………………………………………………………………………………………………………………………………….41

Figure 12- FT-IR spectrum of exopolysaccharides (EPSs) produced in absence of substrate (detection of peaks at 800 and 1240 cm-1) p…………………………………………………………………………………………………………………………………… 42

Figure 13-FT-IR spectrum of modified starches- separation of water-like and water insoluble fractions

p…………………………………………………………………………………………………………………………………….43

Figure 14---Attenuated total reflectance (ATR) spectrum of native-starch/glycerol/clay composites (top) and clay spectra (below) Figure 15--Attenuated total reflectance (ATR) spectrum of two different samples of modified starch clay glycerol composites showing complementary information related to new molecular interactions p…………………………………………………………………………………………………………………………………….45

Figure 16 -FT-Raman spectrum of the substrate, modified starches, and exopolysaccharides produced by the microorganisms in absence of substrate

p…………………………………………………………………………………………………………………………………….51

X

Figure 17-A-B –Oostergetel and Van Bruggen model of the amylopectin clusters, branching and molecular pattern (A); the left-handed three-dimensional helical structure of amylopectin (B). It’s been explained by the authors of this model [53] that neighboring helices are shifted relative to each other by half the helical pitch (indicated by 0 and ½). Figure 18-- Substitutions occurring in amorphous regions of the amylopectin molecules near the branching points [53]

p…………………………………………………………………………………………………………………………………….52

Figure 19- FT-Raman spectrum of Polyplast® samples- laser source 532 nm 20 mW; spectral range 70, 1555 -1525, 2740-2710, 3700 cm-1; integration time 20 sec Figure 20- Spectra of native and modified granular starches (cd2c/MSP): laser source 532 nm 20 mW; spectral range 1525-2740 cm-1; integration time 30 sec p…………………………………………………………………………………………………………………………………….55

Figure 21-FT-Raman scanning of the surface (3D) of the substrate (native starch) p…………………………………………………………………………………………………………………………………….57

Figure 22-FT-Raman scanning of the surface (3D) of the modified starch

p…………………………………………………………………………………………………………………………………….58

Figure 23-- 300 MHz 1H NMR spectrum of Polyplast® polymers showing solvated, probably pendant groups, in D2O p…………………………………………………………………………………………………………………………………….62

Figure 24-300 MHz 1H NMR spectrum of EPS produce by the fungi in absence of starch salvation in D2O p…………………………………………………………………………………………………………………………………….63

XI

Figure 25-MALDI-TOF MS spectrum of native starch- p…………………………………………………………………………………………………………………………………….70

Figure 26- MALDI-TOF MS spectrum of modified starches p…………………………………………………………………………………………………………………………………….71

Figure 27- Chromatographic profiles of modified starches synthesized from

tapioca starch at the 3rd day of modification (CarboPac PA1)

p…………………………………………………………………………………………………………………………………….77

Figure 28-Chromatographic profiles of modified starches synthesized from potato starch at the 3rd day of modification (CarboPac PA1) Figure 29- Chromatographic profiles of modified starches synthesized form corn starch at the 3rd day of modification (CarboPac PA1) Figure 30- Chromatogram profile of modified starches- detail of peak separation performed with a CarboPac PA200 column (peak separation corresponding to peak no. 4 in Fig. 55) p…………………………………………………………………………………………………………………………………….78

Figure 31-- Chromatographic profiles of modified starches synthesized from corn starch after the 3rd day of modification (CarboPac PA1) Figure 32- Chromatographic profile of modified starch synthesized from PDB

(CarboPac PA1)

Figure 33-Exo-polysaccharides (EPSs) produced by the fungi in yeast extract

(CarboPac PA1) –no substrate involved

p…………………………………………………………………………………………………………………………………….79

XII

Figure 34-- Chromatographic profiles obtained for fermented starches (corn, tapioca, or potato) with increase in the spore concentration Figure 35-- Chromatographic profiles modified starch (from tapioca, potato, or corn) produced in Na+NO2

-, Na+NO3-, HPO4

-2(NH4)2, or NH4+NO3

-. The effect of the different nitrogen sources was similar. Figure 36-Chromatogram of one of the substrates (the example native starch (Carbo Pac PA1) p…………………………………………………………………………………………………………………………………….80

Figure 37- Chart showing the retention time of the various used standards

p…………………………………………………………………………………………………………………………………….81

Figure 38-Chromatogram showing the sugar separation of hydrolyzed mod. starch from tapioca starch. Separation by CarboPac PA1 Figure 39-Chromatogram showing the sugar separation of hydrolyzed mod starch from potato starch. Separation by CarboPac PA1

p…………………………………………………………………………………………………………………………………….82

Figure 40- Chromatogram showing the sugar separation of hydrolyzed

modified corn starch. Separation by CarboPac PA1

Figure 41-Chromatogram of hydrolyzed fungal exo-polysaccharides (EPSs)

produced in absence of substrates. Separation by CarboPac PA1

Figure 42--Chromatogram of hydrolyzed mod. starch from amylopectin. Separation by CarboPac PA1

p…………………………………………………………………………………………………………………………………….83

XIII

Figure 43-Chromatogram of hydrolyzed modified starches from PDB.

Separation by CarboPac PA1

Figure 44-Chromatogram of one of the standards -D-Glucose. Separation by CarboPac PA1 p…………………………………………………………………………………………………………………………………….84

Figure 45--Linear viscoelastic region (LVR) determined by DMA in films produced by casting method with modified starches Figure 46- Linear viscoelastic region (LVR) determined by DMA in films produced by casting method with native starches p…………………………………………………………………………………………………………………………………….99

Figure 47-Creep compliance determined during the LVR test in native starch films

Figure 48- Creep compliance determined during the LVR test in modified

starch films

p………………………………………………………………………………………………………………………………….101

Figure 49-Stress relaxation curves for native starch films

Figure 50-Stress relaxation modulus in modified starch films

p………………………………………………………………………………………………………………………………….102

Figure 51- DMTA spectrum of native starch films produced by casting method

Figure 52-DMTA spectrum of modified starch films produced by casting method p………………………………………………………………………………………………………………………………….104

XIV

Figure 53- DMTA curve profiles of native starch glycerol composites produced

by extrusion

p………………………………………………………………………………………………………………………………….105

Figure 54-DMTA curve profiles of modified starch glycerol composites

produced by extrusion

Figure 55-DMTA curve profiles of modified starch-glycerol composites produced after extrusion

p………………………………………………………………………………………………………………………………….106

Figure 56---DMTA curve profiles of native starch-glycerol-clay composites

p………………………………………………………………………………………………………………………………….108

Figure 57-DMTA curve profiles of modified starch-glycerol- clay composites p………………………………………………………………………………………………………………………………….109

Figure 58- TG-DTG plot of native starches showing the degradation point at the 1st derivative Figure 59- TG-DTG plot of modified starches showing the degradation point at the 1st derivative

p………………………………………………………………………………………………………………………………….123

Figure 60-TG-DTG plots showing successive derivatives for modified starches showing a clear double thermal transition peak

Figure 61-TG-DTG-successive derivatives obtained by TG in native starches showing the lack of thermal transitions p………………………………………………………………………………………………………………………………….124

Figure 62- Successive derivatives modified starches showing high energy consumption during the transitions at the point of thermal degradation

p………………………………………………………………………………………………………………………………….125

XV

Figure 63- DSC thermograms of the unmodified substrate (native starch). Peaks induced with 0.6 M KCl (granular starch)

Figure 64- DSC thermograms of modified polysaccharides

p………………………………………………………………………………………………………………………………….133

Figure 65- DSC thermograms of native starch-films Figure 66- DSC thermograms of modified starch films

p………………………………………………………………………………………………………………………………….134

Figure 67-DSC thermograms of native starch films Figure 68-DSC thermograms of modified starch films

p………………………………………………………………………………………………………………………………….135

Figure 69- Plots of shear viscosity vs. shear rate (TP-native starch; TP-EPSs-Modified starches)

p………………………………………………………………………………………………………………………………….140

Figure 70-Shear stress vs. shear rate (TP-native starch; TP-EPSs-Modified starches)

p………………………………………………………………………………………………………………………………….141

Figure 71- Tensile strength (MPa) and elongation at break (%) of modified

and native starch glycerol and clay composites

p………………………………………………………………………………………………………………………………….154

XVI

List of Tables

Table 1

Chemical composition and physical characteristics of two different sources of common starches

type A (cereal-corn) and type B (tuber-potato) ……………………………………………………………………… 9

Table 2 FT-IR data: GMS (granular modified starches); DMS (gelatinized modified starches); GNS/NS (granular or gelatinized native starches)…………………………………………………….…………………………….33 Table 3

Assignment of the most important Raman bands of the native and modified starches

………………………………………………………………………………………………………………………………………….………….53

Table 4

Solid state NMR, chemical shifts for the different carbons of native and modified

starches……………………………………………………………..………………………………………………………………………..67

Table 5

Solid state NMR averaged associated area by carbon type ………………………………………………………67

Table 6-

Extrusion temperature profiles for samples finally tested ………………………………………………………..96

TABLE 7

Commonly Used Kinetic Equations …………………………………………………………..………………………………119

Table 8

DSC melting parameters for the substrate and modified

starches………………………………………………………………………………………………….………………………………….131

Table 8

Calculated K values for modified and native starches (Capillary rheometer) from the power law

equation …………………………………………………………………………………………………………..……………………….141

XVII

List of abbreviations

DMTA- dynamic mechanical thermal analysis

DMA- dynamic mechanical analysis

TA- thermal analysis

TG- thermogravimetry

TG-DTG- thermogravimetry and successive derivatives

DSC- differential scanning calorimetry

FTIR- Fourier transform infrared

FTIR-ATR- attenuated total reflectance

FT-Raman- Fourier transformed Raman spectroscopy

NMR- nuclear magnetic resonance 1H NMR- proton nuclear magnetic resonance

SS CP/MAS 13C NMR- solid state cross polarization magic angle spinning 13C NMR

MALDI-TOF MS- matrix assisted laser induced time of flight mass spectrometry

HPAEC-PAD- high performance anion exchange chromatography-PAD

PAD-pulsed amperometric detection

DSC- differential scanning calorimetry

SEM—scanning electron microscopy

DP- Degree of polymerization

LVR- linear viscoelastic region

Mw—Molecular weight

XRD— X-Ray diffraction

CL—crystalline lamellae

AL—amorphous lamellae

AV—amylose lipid complex

mc—moisture content

EC—enzyme commission numbers (http://www.brenda-enzymes.info/index.php4)

EPS(s)—exopolysaccharide (s)

LPLs—lysophospholipids

FFAs—free fatty acids

PHAs—polyhydroxyalkanoaes

TPS—thermoplastic starch by using a suitable plasticizer (e.g. glycerol)

RH—relative humidity

mp—melting point

XVIII

TS—tensile strength

EM—elastic modulus

ASTM— American standards of testing materials

XIX

XX

List of symbols

E— Young modulus

h— Inelastic behavior

L— elongation

Lo— initial length

L— increment in elongation

F— Force

E'— storage modulus

E''— loss modulus

tan —loss factor defined as tan = E''/ E'

' — strain rate

— shear stress

— strain

ŋ — viscosity

—shear stress (engineering stress)

n— power low index in a power law relation = K'n

K— consistency index in = K'n

tan —damping factor (tan =E”/E’)

J(t)—creep compliance

E(t)—relaxation modulus

Tg — glass transition temperature

Cp—heat capacity

g mol-1— grams per mole

Da—Daltons

ppm—parts per million

m—microns

%E—percentage of elongation at break

MPa—mega Pascal

A— Ampers

Vf— free volume

E* or G*—complex modulus

* or*—complex stress

*or*—complex strain

k — constant in the Hook’s relation (F= - k*displacement)

1. Introduction

1.1. Motivation of the study

The increased release and accumulation of synthetic plastics—especially

packaging—into the waste stream around the world, has driven the demand

for bio-degradable polymers [1, 2]. These materials are mainly targeted to

single use, disposable packaging, consumer goods, pharmaceutical capsules,

disposable nonwovens, coatings for paper and paperboard, and some non-

packaging markets [3]. The approach of composting as an ecological

alternative to manage most of these materials is currently supported by most

researches around the world, industry, international markets, and

municipal/national facilities [4]. Moreover, their importance increases since

not all synthetic plastics are recovered and not all of those bearing recycling

symbols are recycled for economic or technical reasons [5].

Starch has become one of the most promising candidates among the various

alternatives to substitute synthetic plastics, especially for packaging because

it is an inexpensive material and behaves as a thermoplastic [2, 6].

Therefore, numerous studies have been conducted to optimize the

performance of the starch-based polymers [7-16]. However, starch

thermoplastics have not come into a practical and widespread use mainly

because their susceptibility to water and low compatibility with most

polymers—synthetic, photo- or bio-degradable. The alternatives to improve

the properties of native starches include chemical and physical modifications,

but also starches may be converted to more useful forms by using enzymes;

however, usually related with the starch fragmentation [17]. Overall, the

final goals in the area of biodegradable polymers based on starch are related

to the improvement of the processability (i.e., extrusion, injection molding)

and compatibility with other thermoplastic polymers, as well as the reduction

of the water intake by using cost-effective and environmentally safe

methodologies.

1

The chemical conversion of starch has widely been explored (esters, ethers,

or grafted starches). For example, a number of authors have reported the

preparation of esterified starches of high degree of substitution (DS), in the

presence of organic solvents, or systems of solvents, used to achieve

homogeneous modification of the starch. Such modifications produce

thermoplastic starches, but the treatments are not economic and/or

environmentally efficient due to the toxic and/or expensive solvents used

under high alkaline conditions and temperatures, conditions which are

unsuitable for industrial scale [18-25]. Therefore, physical or enzymatic

treatments of starch become more attractive alternatives in pursuing such

objectives.

To date there are two ways in which starch is modified by using physical

means to produce thermoplastic derivatives: starch is used as a filler in

blends with synthetic or biodegradable polymers or starch is extruded in the

presence of a suitable plasticizer to form a thermoplastic mass [26, 27]. The

use of starch as a filler is one of the most investigated processes to decrease

the use of synthetic plastics or to lower the price of biodegradable polymers

[28-35]. In this method, the starch and the polymers are extruded or

injection molded to produce thermoplastic composites. The disadvantages of

this method are: the inferior properties of the materials when the amount of

starch exceeds 10%, the low interfacial affinity with most polymers

(synthetic or biodegradable, i.e., PLA), and the inaccessibility of the

encapsulated starch to biodegrading agents such as water, light, air, and

microorganisms. This problem may be solved in part by reducing the particle

size of the components in the composite by a strong destructurization

occurring with the gelatinization of the starch (thermomechanical input and

water content) during the extrusion which may lead to a more or less

continuous phase more susceptible to biodegradation [26].

The second alternative and one of the most studied methods is to process

starch by extrusion or injection molding in the presence of inexpensive

2

plasticizers (normally water or glycerol) to produce a thermoplastic mass.

These materials are relatively inexpensive, totally safe and biodegradable.

Their properties can be improved or modified by reactive extrusion or by

extruding the starch with fillers like mineral clay [36-38].

Microbial exo-polysaccharides (EPSs) produced by both prokaryotes

(Eubacteria and Archaebacteria) and eukaryotes (phytoplankton, fungi, and

algae) which are rich sources of enzymes, are also being intensively

investigated as a permanent source of polysaccharides for industrial

applications. They present a wide rage of chemical structures, but with

exception of polyhydroxyalkanoaes (PHAs), most of them have not yet

acquired appreciable significance in packaging or similar applications [28, 39-

41].

A recent study, published by Jeng et al. [42], is of critical significance. The

study basically showed that it is possible to modify and enhance the

properties of starch by using bio-catalysis or fermentation. The particular

fungal species used by these authors were Ophiostoma spp. It was reported

that these fungi have the ability to produce polysaccharides of high molecular

weight in culture media. When the medium is supplied with any source of

starch the recovered polysaccharides exhibit better functional properties than

native starches. It was also observed the lag or null degradation of the starch

source at prolonged reactions times. These observations may be attributed in

a first instance to the lack of degrading enzymes. The reported increase in

the molecular weight may be due to the production of EPSs and protein-like

compounds [42, 43]. However, the overall influence of the fungal treatment

on the functional properties of the recovered polysaccharides has not been

sufficiently explored.

This work forms an essential part of a larger study aimed at producing bio-

plastics using the biosynthetic pathway as described previously [44, 45]. The

specific purpose of this study was to investigate some of the functional

3

properties of these materials such as the viscoelasticity and rheology. The

description of certain chemical properties is used to support and explain the

dynamic mechanical and rheological observations. Since the study of the

details of the bio-synthetic pathway is ongoing, the information elucidated in

this work can be used to give insights related to the process by which these

polysaccharides are produced.

Some of the analytical techniques used during this study include dynamic

mechanical thermal analysis (DMTA), differential scanning calorimetry (DSC),

thermogravimetric analysis (TG), Fourier transform infrared (FTIR), FT-

Raman, capillary rheology, solid and liquid state NMR, etc.

1.2. Molecular, physical, and functional properties of starch

1.2.1. Molecular structure of starch

The molecular composition and architecture of the starch granules are the

main properties influencing the processing conditions, final products, and the

performance of starch-derived materials. These features also determine the

interaction of the enzymes with the substrate (chemical reactions, whether

enzyme-catalyzed or not, proceed mainly through the formation and

cleavage of chemical bonds). In order to discuss these relationships and

bring them forward it is necessary to briefly review these topics.

Starch is a natural polymer easily isolated in huge quantities from

agricultural staples such as corn, potato, or tapioca roots. It occurs naturally

in semi-crystalline granules of different sizes, size distribution, and shapes

mainly composed of two -D-gluco-polysaccharides with different

architectures; amylose and amylopectin, in a ratio of ~30% to 70%

respectively. This ratio can be altered substantially by selective breeding or

by biotechnological methods, e.g., waxy maize is 99% amylopectin, and the

different amylomaize varieties can be classified according to the percentage

4

of amylose; i.e., amylomaize V contains 50-60%, VI 60-70%, and VIII 70-

80% [2, 46].

Amylose is basically a helical (non- or slightly branched) polymer with

molecular weights (Mw) around 1X105 g mol-1 and it is found within the

granules in amorphous regions. The chains show spiral-shaped single or

double helixes with a rotation in the (1-4)-link and with six glucose units per

rotation. Amylopectin is a highly branched polymer with Mw in the range of

1X107 to 1X109 g mol-1. The branches occur at C-6 hydroxyl group of a given

anhydroglucose unit, and 4 to 6% of the CH2OH groups are substituted. The

average degreed of polymerization (DP) of the branches is ~30 glucose units.

The branches are localized every 20 to 70 glucose units giving the

appearance of a grape-branched like structure called ‘cluster’. The branching

molecular aggregation is responsible for the two different X-ray diffraction

(XRD) patterns reported for cereals (i.e. corn type-A) and tubers (i.e. potato

type B). The crystalline structure is normally determined by the length and

density of the branches of the amylopectin molecules which are part of the so

called cluster-structure [46].

Granules (~20 to 100 m) are composed of alternating semi-crystalline rings

or shells (~1200-4000 Å). Crystalline layers are about 50 Å and increases up

to 70 to 80 Å at the end of the growth ring. The amorphous layers of

amylopectin regions are probably less than 40 Å. Amorphous regions may be

formed by -1-6 branching regions. The rings are visible by atomic force

microscopy (AFM) or by optical microscopes with a resolution in micrometer

(m). The smallest visible structures by AFM are the so called ‘blockets’

which have been associated to the crystalline fractions forming the ‘clusters’.

The size of the blockets is reported up to 4000 Å (the size of one growth

ring). Blockets have been defined as ‘semi-crystalline globular structures’

surrounded by a soft matrix (amorphous amylose) and disordered regions of

amylopectin fractions.

5

The ‘blockets’, are formed by ‘stacks’. Within each crystalline stack, there are

arrays of amylopectin or ‘clusters’ arranged in the form of crystalline lamellae

(CL) which are double helices zones (5-6 nm), and amorphous lamellae (AL)

formed by branching zones (4nm), giving a total of a ~9 nm periodicity. The

gaps between neighboring clusters (~5 nm) are filled with amylose and in

some cases amylose-lipid complexes (AV). Neighbor clusters merge together

to form a three-dimensional structure—the super three-dimensional helix

model [47].

The lamellae responsible for the crystalline regions are formed by three

discrete components: the backbone which support the double helices, parallel

‘rigid’ double-helical units (~5-6 nm), and amorphous regions (more flexible

un-branched regions, also called ‘spacers’ or side chains) with sizes of ~4 nm

[48, 49]. The size of the crystalline lamellae is ~9 nm. It has been by

observed simultaneous appearance and disappearance of the 9 nm and 1.6

nm reflections in small angle X-ray scattering under hydration and

dehydration experiments. The hydration produces the 9 nm reflection due to

the smectic periodicity. With a ~10% mc (moisture content) solid state NMR

spectra show a set of sharp and strong peaks at ~100 ppm (three peaks in

starches type A, and two in type B) associated to the crystalline regions.

Neither dehydrated native granules (<5% mc) nor the amorphous

dehydrated starches show these signals. The same phenomenon occur for

highly hydrated starches (~20% mc <). These authors suggested that under

dry conditions the starches may be in a pure glassy form (<5% mc), while

mc of ~10 % allow the formation of crystalline regions. These particular

structures of intermediate order are known as liquid crystals; SCLCP (side

chain liquid crystal polymer). The degree of mobility of these three

components, coupled with the helix-coil transition, may be used to explain

physicochemical and structural properties of starch such as gelatinization,

dehydration or molecular composition [48-53].

6

1.2.1.1. Minor components of starch

Minor components may also be present in different proportions: lipids

(~1.0% in cereal endosperm and 0.1 % in potato tuber), proteins (~0.25%

average; 0.5% in cereal endosperm and 0.05% in potato tuber), and silica

and phosphates are also present in low concentrations (potato starch

granules are highly phosphorylated). Phosphate groups in potato starch are

located in the center of the granules [2, 27, 54].

Cereal starches contain integral lipids in the form of lysophospholipids (LPLs)

and free fatty acids (FFAs) which have been found in association with the

amylose fraction. The lipids form a hydrophobic core in the helical molecule

of amylose. LPL can be as high as 2% by weight in high amylose starches.

The surface of starch granules can also present lipids such as triglycerides,

glycolipids, phospholipids and free fatty acids. These materials are generated

from the amyloplast membrane and non-starchy sources. The presence of

LPL and FFA depend also on the starch source [46].

The chemical signal appearing at 25-35 ppm in the 13C CP/MAS NMR (Cross

Polarization Magic Angle Spinning NMR) spectrum for native corn starch has

been associated directly with the presence of amylose-lipid complexes. The

presence of these complexes in the starch granules is also shown by the

lower iodine binding capacity of defatted amylose helices than the

corresponding lipid-extracted material. Moreover, lipids could be localized in

specific zones known as V-type starch structures. The content of lipids based

on amylose content can be as high as 50% or higher which is the case of oat

starches. These fractions are highly susceptible to enzymatic (fungal or

microbial) attack, and they will be removed firstly during the process or

modification [55].

In the granules, proteins have been reported to be localized either in the

surface or in the core of the granules, and mostly near the hilum. Isolated

7

starch granules may contain up to ~0.6% protein. Regardless of its origin, it

seems that proteins are located in the surface of the granules and at the

matrix of the granules formed by the amylose-amylopectin. Proteins seem to

affect the functionality of the granules, i.e., the grain hardness in wheat

starch is probably due to the friabilin. The molecular weight of the proteins

located at the surface is less (~15-30 Da) than those located at the core of

the granules (~50-150 Da). Proteins with higher molecular weight are

probably located at the hilum of the granules. Proteins include the enzymes

of starch bio-synthesis which may contribute to the flavor of the starch such

as starch synthase involved in the starch synthesis [54, 56].

Starches can also contain minerals such as calcium, magnesium, phosphorus,

potassium and sodium (in percentage less than 0.4%). Phosphorous may be

present in form of phosphate monoesters, phospholipids and inorganic

phosphates. Phosphate monoester is present in potato starch in quantities

not exceeding the 0.1%. Although in low concentrations, proteins, inorganic

materials and lipids can influence at different degrees the technical properties

of the starch [46, 55].

1.2.1.2. Comparison between cereal and tuber starches

The general properties of two different starch sources are provided in Table

1. Between these two starches there are clear differences in the size of the

granules and chemical composition which can directly affect the processing

conditions[2, 27, 46].

8

Table 1

Chemical composition and physical characteristics dry weight basis of two

different sources of common starches type A (cereal-corn) and type B (tuber-

potato) (11, 47, 54).

Corn Potato

Amylose (%) 27 ±1 23 ±2

Amylopectin (%) 72±1 76±3

Lipid content (%) 0.63 0.03

Protein content (%) 0.30 0.05

Phosphorous content (%) 0.02 0.08

Moisture content (%) 12-13 18-19

Granular size (m) 15 30-100

Crystallinity (%) 40 25

1.3. Exopolysaccharides and other fungal metabolites

1.3.1. Microbial metabolites and industrial uses

Microorganisms such as bacteria and fungi are a rich source of internal and

external metabolites such as enzymes, polysaccharides and/or protein-like

polymers. Since this work is based on the production of polysaccharides from

starch by specific fungal isolates, it is of interest for this work to briefly

discuss the industrial use of some of these microbial metabolites— and

particularly those already identified in the genus Ophiostoma spp. It is also of

special interest is the process by which these metabolites modify or convert

the different substrates. In general, enzymes have been used to degrade,

but also to produce thermoplastic starches. Microorganisms have also been

used to improve the properties of these substrates.

9

1.3.1.1. Enzymatic conversion of starch

Hydrolases such as - (EC 3.2.1.1; enzyme commission numbers)

(http://www.brenda-enzymes.info/index.php4) and -amylase (EC 3.2.1.2),

glucoamylase (EC 3.2.1.3 - glucan 1,4-alpha-glucosidase), pullulanase (EC

3.2.1.41) and isoamylase (EC 3.2.1.68) are the industrial enzymes used for

the production of a wide range of low molecular weight derivatives from

starch such as dextrose, maltodextrins, glucose, and maltose syrups, as well

as substrates for culture media [17].

In general, -amylase randomly hydrolyzes the glycosidic linkages along the

starch backbone, -amylase produces the equivalent to maltose units from

the end of a starch molecule, and glucoamylase produce one glucose unit at

a time. Glucoamylase attacks the starch molecules from the non-reducing

end-groups. At 37oC, with limited water content this enzyme converts a mass

of 10-50% of the starch granule to glucose. On the other hand, pullulanase

and iso-amylase are debranching enzymes which attack the 1, 6-linkage—the

size of the molecules obtained by fragmentation with these enzymes from

the amylopectin molecules correspond to the length of the branches. The

temperature at which bioconversion of the starch is conducted depends on

the source and the type of enzyme. For example, reactions in -amyloases

from bacteria are performed at 90-100oC, -amylase from fungi normally at

50-60oC, pullulanase 50-60oC [57].

Enzymatic activity seems to change with the starch source, morphology

(crystallinity), and the methods used during the starch conversion. Enzymes

such as glucoamylase and isoamylase first attach to the active sites of the

substrate before product formation. The enzymes can penetrate through the

pores of the starch granules and then bind to the internal starch molecules.

The physical damage due to these enzymes can be seen as pin-holes on the

surface of the starch granules by scanning electron micrographs (SEM) [58].

During this process, the enzymes release glucose from internal molecules.

10

Therefore, the process of hydrolysis depends not only on the porosity

pattern, but also on the chemical structure of the granules [59, 60].

The different amylolitic patterns among the dissimilar crystalline types may

be due to the variation in the location of their amylopectin branch points. The

A-chains (DP 6-12) (therefore the branch linkages in the crystalline lamellae)

of the A-type starches may be more susceptible to enzyme hydrolysis than

B-chains in B-type starches. In B-type starches, more branch points may be

found merged in amorphous regions providing an apparent crystalline

structure more resistant to hydrolysis.

Other related phenomena occur at the branching points. Based on the

degradation of hydroxyl propyl di-tapioca, hydroxyl propyl potato,

methylated potato, cationic waxy corn, and cationic potato starches by -

amylase and pullulanase, it has been found that the substitutions are located

near or at the branching points in the amorphous regions of the amylopectin

molecules (branching points are amorphous, more flexible un-branched

regions, also called ‘spacers’ or side chains, with sizes of ~4 nm which are in

alternating order with the crystalline regions composed of parallel double

helices with the size of ~5-6 nm). It has also been demonstrated that the

amylase fractions are easily accessible for hydroxypropylation, and these

amorphous regions are easily accessed by acid or enzymatic activities.

-amylolysis is affected by the size, type and arrangement of starch

molecules in the amorphous and crystalline lamellae and their interactions

with non-starch components. Crystalline regions may be formed by chain

association after initial hydrolysis hindering the further accessibility of -

amylase to the glucosidic bonds. It is common to find reports of a fast initial

hydrolysis followed by a lag enzymatic degradation. In potato starch (type B)

the size and location of the blockets; bigger than in A-type starches and

located mostly at the surface, may influence the enzymatic pattern [54].

Enzymes (i.e., lipases) are also being used to produce esterification of starch

11

by using long chain fatty acids and by using new methods based on

microwave radiation [41].

1.3.1.2. Microbial metabolites: the case of Ophiostoma spp.

Polysaccharides (PSs) (exopolysaccharides, EPSs; encapsulated

polysaccharides, ECPSs; and structural polysaccharides; EPs) from

prokaryotes (Eubacteria and Archeabacteria) or eukaryotes (algae,

phytoplankton, and fungi) are important biological products of growing

interest for a variety of food and non-food industrial uses such as rheological

modifiers as well as bioactive molecules for therapeutic uses among others

[1].

Structural polysaccharides provide support and give coherence to the

microbial cells. The term exopolysaccharides (EPSs) is used to describe

polysaccharides produced during the growth of the microorganisms and may

occur as ECPSs or as slimy substance surrounding the medium of the cell

which are used by microorganisms to propagate, avoid desiccation,

protection, and as a fixation mechanism, some of them may also have

pathogenic activities [61-63]. In bacteria, EPSs are a protective barrier

against bacteriophage and are produced to resist desiccation and survive

under dry conditions [64]. There is one more possibility in which EPSs are

used as the food source.

The chemical structure of EPSs may be study to support theories associated

to biosynthesis and functionality [65]. EPSs according to their molecular

structure can be divided in homopolymers (i.e., cellulose, dextran, levan,

curdlan, pullulan) and heteropolymers (i.e., gellan, xanthan).

Heteropolysaccharides produced by lactic acid bacteria may be branched and

constituted by different ratios of D-glucose, D-galactose, and L-rhamnose,

and in some cases by glucoronic acid, acetylated amino sugars moieties, and

non-carbohydrate substituents like phosphate or acetate groups [65, 66].

12

Some techniques used to study the chemical structure of microbial

polysaccharides are UV, 1H and 13C solid and liquid state NMR, GC-MS, FTIR,

FT-Raman [67].

Polysaccharides from microbial origin may find or have found applications in

pharmaceuticals, cancer therapy, drug delivery, oil and metal recovery in the

mining industry, waste recovery, detergents, textiles, adhesives, paper,

paint, food, and beverage industries. Alginates, a filamentous or granular

polymer produced by Pseudomonas aeruginosa and Azotobacter vinelandii

bacteria, is a viscous gum that occur in the cell walls of brown algae.

Emulsans produced by Acinetobacter calcoaceticus have found applications as

substrates to produce enzymatic reactions, encapsulate fertilizers, pesticides,

and nutrients, as coatings of roots of seedlings and plants to prevent

desiccation, and as hypo-allergic wound-healing tissue. There are also some

reports in which these materials are used for their metal binding properties

[68]. Gellan gum, used in the food industry, is a water-soluble

polysaccharide produced by bacterium Sphingomonas elodea or S.

paucimobilis, it is used as immobilizing (solidifying) agent of microorganisms

propagated in culture media. Hyaluronic acid (from Streptococus equii and S.

zooepidermicus) is used as ocular, skin, and wound protectant (i.e.,

lubricants), as synovial fluid, and cosmetics. Xanthan (from Xanthomonas

campestris bacterium) (E 415) is widely used as additive in the food industry

as a rheology modifier: as thickening, stabilizing agent and in free gluten

formulations, also as a tertiary crude-oil recovery, in paints, pesticide and

detergent formulations, cosmetics, pharmaceuticals, printing inks, is also

used in combination with guar gum. Cellulose from bacteria (Acetobacter

spp.) is used as temporary skin to heal burns or surgical wounds, in dietary

formulations in combination with vitamins and minerals, as micro membranes

for filtration, as acoustic membranes in audio-visual equipment. Curdlan

secreted by Bacteria Alcalinenes faecalis var. myxogenes, Rhizobium meliloti

and Agrobacterium radiobacter is a polymer used for biodegradable materials

for medical and other important uses. Curdlan is also a gelling and

13

immobilizing agent, it is being tested in combination with zidovudine (AZT)

as antiretroviral (anti AIDS-drug). Succinoglycans (acid glucans) produced by

bacteria of the genera Pseudomonas, Rhizobium, and others have been found

similar applications to curdlan. Dextran (from bacteria Leuconostoc

mesenteroides, Leuconostoc dextranicum, Lactobacillus brevis, and

Lactobacillus hilgardii) has been used as antithrombotic and blood reducer

viscosity and to lower the cholesterol, also in sieving technologies, and as a

micro-carrier in tissue culture (cross-linked dextran) [69].

“Bioplastics” (poly 3-hydroxyalkanoates; PHAs) natural polyesters produced

by bacteria have been intended for plastic bottles, fibers, latex, and in

general for packaging (Biopol® by ICI Ltd). However, the high production

costs prohibits such uses, and PHAs are being used for the manufacture of

medical devices and for therapeutic applications such as thread for sutures,

implants, urological stents, neutral- and cardiovascular-tissue engineering,

fracture fixation devices, in the treatment of narcolepsy and alcohol

addiction, drug-delivery vehicles, cell microencapsulation, support of

hypophyseal cells, or as precursors of molecules with anti-rheumatic,

analgesic, radiopotentiator, chemopreventive, antihelmintic or anti-tumoral

properties (those containing aromatic monomers or those linked to

nucleosides) [70].

Screening of thermophilic and hyperthermophilic bacteria from deep-sea

hydrothermal basins has produced bacteria like Pseudomonas, Alteromonas,

and Vibrio which have been used so far to produce extracellular polymers

(exopolysaccharides; EPSs) in aerobic-carbohydrate-based media. Some

properties studied so far are metal binding capabilities and biological

activities like antitumor, immunostimulatory, and anticoagulant activities.

The anticoagulant activity has been linked to the high sulfate content of

some of these polysaccharides. The uronic acid content in these

polysaccharides varies from 10 to 40%, and posses molecular weights up to

14

1X106 g mol-1. These bacteria may be also a rich source of thermostable

enzymes.

Alteromonas strain 1545 (bacteria), isolated from the epidermis of the

polychaete Alvinella pompejana found in the hydrothermal vents of the

Pacific Ocean (one of the most heat tolerant organisms on earth) produces

under laboratory conditions an anionic EPS, which may consist of glucose,

galactose, glucoronic, and galacturonic acids along with a 4, 6-O-(1-

carboxyethylidene)-galactose residue. Some studies performed on this new

polysaccharide include its rheology, and it has been proposed as a thickening

agent. A polysaccharide secreted by a bacterium (Alteromonoas strain 1644)

isolated from Alvinellide—the polychaete Paralvinella sulfincola also from the

hydrothermal vents of the East Pacific— may be composed of four neutral

sugars and four acidic sugars. Three of the uronic acid residues form a

trisaccharide unit and the last one carries a lactate group. In solution, the a

gel is formed which shows strong selectivity between monovalent and

divalent ions, as well as a great affinity for the divalent ions, higher than

predicted by electrostatic theories, with the exception of Mg2+.

A polymer produced by Alteromonas macleodii subsp. fijiensis, which is an

aerobic, mesophilic, heterotrophic bacterium isolated from a diluted

hydrothermal vent fluid at a depth of 2600 m in a rift system of the North Fiji

basin, is a hexasaccharide with three linked uronic acids and with a side

chain ended by a 4, 6-O-(1-carboxyethylidene)-mannose residue. This

polymer may be used as thickening material and present a shear-thinning

behavior. Gelation properties observed in the presence of calcium can be

explained on the basis of intermolecular Ca2+ bridges formed between

carboxyl oxygen atoms of the glucuronosyl and galacturonosyl residues. A

high metal-binding maximum capacity (up to 316 mg Pb(II)/g polymer) was

observed in a single metal system, indicating that this polymer may have

potential for use in applications in wastewater treatment and

biodetoxification of heavy metal-polluted water. This polymer is hydrophobic,

15

and has been intended to encourage bone healing by adhering onto the

osteoblastic cells. Alteromonas infernus can produce water soluble EPS in

presence of glucose with high attraction for heavy metals such as lead,

cadmium, and zinc. Sulfated and depolymerized of materials may be used as

anticoagulant agents. The EPS produced by Pseudoalteromonas strain 721 is

an octasaccharide with two side chains. This material produces gelatinization

after thermal treatment. This material after gelatinized with NaCl exhibits

viscoelastic behavior.

An aerobic, heterotropic, and mesophilic bacterium (Vibrio diabolicus)

isolated from the polychaete Alv. pompejana found in the deep-sea

hydrothermal field of the East Pacific produces EPS in presence of glucose

characterized by equal amounts of uronic acid and hexosamine (N-acetyl

glucosamine and N-acetyl galactosamine. Structural studies recently

conducted on this polymer demonstrated that it consists of a linear

tetrasaccharide repeating unit. The role of this novel bacteria polysaccharide

in bone regeneration has recently been successfully investigated [71, 72].

Extremophiles are organisms that live and evolve in extreme environments of

temperature, alkalinity, salt concentration, pressure, etc., and fall into a

number of different classes and domini belonging to Archea as well as

Bacteria. They include thermophiles, acidophiles, alkaliphiles, psychrophiles,

barophiles, radiophiles, etc. Extremophile organisms are studied as new

sources of bioactive molecules, and industrial interest. Thermophilic bacteria

— which grow between 60°C and 80°C — belong to the genus Bacillus,

Clostridium, Thermoanaerobacters, Thermus, Fervidobacter, Thermotoga,

and Aquifex. Hyperthemophiles—which can withstand temperatures up to

110°C belong to the Archaea consisting in two major kingdoms and short

phylogenetic branches. Lipids from archeal membranes are known to be

extremely stable materials, and have been proposed for use in drug delivery.

It has been also thought that these materials may be intended for the

production of biodegradable materials. Of special interest is the production of

16

enzymes from these microorganisms. Many materials produced by

thermophilic strains are being studied and characterized like Bacillus spp. a

thermophilic bacterium or Halophilic archaea (e.g., various strains of

Haloarcula japonica)[73, 74]. A thermophilic strain reported by Moriello et al.

[62] produce up to 90 mg/l of EPSs at temperatures of 60°C at pH of 7.0.

Metal and non-metal substrates (teflon, nylon, polycarbonate, and

polyacrylate) have been used as substratum for propagating marine

microorganisms. The complete characterization of the biofilms generated

during this process is in progress, but some researchers have already

generated some rheological and chemical information related to these

materials and the microorganisms. Many bacteria (like lactic acid bacteria)

are characterized by their ability to convert large proportions of their carbon

feed, fermentable sugars, to lactic acid and EPSs. One interesting property is

that many of these EPSs are water soluble or when suspended or dissolved in

aqueous solution provide thickening and gelling properties which may be of

enormous importance in the food and other industries.

EPSs occur in fungi as a fixation system to the substrate, and some of them

may have biological activities or pathogenic effects over the host. Some

species like Pleurotus produce up to 28 g dry weight of EPSs per liter of

culture media. These EPSs have been used for biomedical applications [75].

Selbmann et al. [76] studied the ability of one-hundred and five fungal

strains from 46 species to produce EPSs. They found the highest yields in

Botryosphaeria rhodina DABAC-P82 which in optimal growth conditions of

nitrogen sources (NaNO3) and pH (3.7) produced 2.0 g l-1 after 24 h of

fermentation. The polysaccharides were characterized as

homopolysaccharides of glucose with molecular weights of 4.87X106 Da, and

the potential presence of -1-3 and -1-6 linkages.

17

It has been found that modifying the culture conditions of the fungus

Antrodia cinnamomea — which is used for therapeutic purposes — the

production of biomass and EPSs can be controlled up to certain point. It has

been reported by Lin and Sung [77] a maximum EPSs production of 0.49 g/l.

Ophiostoma spp. is known to produce exo-polysaccharides (EPS) and various

enzymes which have been linked to the pathogenic activity of the fungus. In

synthetic media, different biological activities have been reported depending

on the source of carbon [42, 78, 79]. For various substrates, Przybył et al.

[78] reported the presence of different ‘hydrolytic’ (cellulolytic) endo- and

exo-enzymes in Ascomycetes fungi of the genus Ophisotoma novo-ulmi, O.

ulmi and fast-waxy such as cellulase, polygalacturonase, xylanase, pectinase,

endogluconase, -glucosidase, exo-galactase, exo-glucanase, -galactosidase

and -galactosidase. It is interesting to note that Binz and Canevascini [79],

tested these microorganisms among other substrates for the production of

extra-cellular enzymes in the presence of starch. Their research focused

more on the pathogenic activity of the enzymes, and they reported higher -

glucosidase activity.

Recently, Jeng et al. [42] reported the production of EPSs and or protein-like

polymers in culture media by these fungi in synthetic media containing

various sources of starch. Specifically, they found that the potato starch

contained in the potato dextrose broth (PDB) was not consumed by the fungi.

Instead, they recovered a mass of polymer partially soluble in water, and

with relatively high molecular weight (average molecular weight 1.5X106 Da).

It has been reported that the enzymes used in industry to induce the starch

hydrolysis (-amylase, -amylase, glucoamylase and pullulanase) are

different from those reported in these fungal species [80, 81].

The paper of Jeng et al. [42] also presents results showing that the partial

hydrolysis of EPSs from PDB by 1,4--glucosidase (glucoamylase) and -

18

glucosidase, while -glucosidase partially hydrolyses the yeast polymer, but

the hydrolysis was not present when they used glucosidase.

Sain and Jeng [44] and Huang et al. [43] have reported the synthesis of

starch-like polymers by biosynthetic pathways and the production of films

based on these materials were shown to be highly hydrophobic. They also

reported that this films were stronger compared with films made with non-

modified starches. Specifically, water absorbance was reported for modified

potato, corn, tapioca or rice starches as low as 1 g(H2O)/ g (polymer), while

for unmodified starches this value increased to 8 g(H2O)/ g (polymer). For

films made with the starch-like polymers, the values shown for peak stress

and elongation at break were up to 8 MPa and 10 mm, respectively, while for

films prepared with native starches the values found were 0.3 MPa and 50

mm respectively. The elongation at break was lower in modified starches,

after the modification, the starch gained in rigidity. These authors also

reported the following results for films based on pure amylopectin: water

absorbance, around 5 g(H2O)/ g (polymer); elongation at break of 20 mm;

and peak stress around 1 MPa (values located between the results obtained

for materials made with modified and unmodified starches). The elongation

at break was also higher in amylopectin films compared to that showed for

the starch-like polymers.

1.4. Objectives and approach

1.4.1. Objectives

The literature survey showed that one of the best alternatives to future

biodegradable materials for short-time applications is starch. However, there

are significant drawbacks, such as high water absorption that limits the

current widespread adoption of starch-derived polymers. In addition, it was

found that there is a requirement to develop non-chemical and inexpensive

modifications to enhance the functional properties of starch.

19

It has also been shown in the literature review that microorganisms,

including bacteria and fungi, are able to produce various enzymes and

exopolysaccharides (EPSs) which may be used to induce modifications to

various substrates. The modification of the starch by using isolates of

Ophisotoma spp. is a new research alternative for starch modification.

However, there is a lack and the need of information related to the functional

and chemical properties of these materials.

Since the study of the biosynthetic pathway that control starch modification

by these fungal isolates is an ongoing process, the focus of this work is to

investigate links between the mechanical and structural properties of this

novel material. The results are used to give insights to the process by which

these specific fungal isolates produce the modification. With a better

description of the molecular structure of these materials a better

understanding of the mechanisms of the biosynthesis can be achieved. The

main objectives were as follow:

Study of the dynamic mechanical and thermal properties of modified

starches (known under the commercial name of Polyplast®) to detect

changes in the glass transition temperature (Tg) and susceptibility to

degradation by heat as compared to unmodified starch.

Study of the rheological properties of modified starches (Polyplast®)

to explain their behavior under extrusion and/or injection molding

conditions and to further improve these processes.

Study of the chemical properties by spectroscopic, spectrometric, and

chromatographic methods to provide insights related to the chemical

structure of these polysaccharides and to support the results obtained

by thermal and rheological analyses.

20

Study of the influence of the chemical structure on the mechanical

properties of modified starches.

1.4.2. Approach

To determine the chemical, physical, and mechanical properties of these

materials (Polyplast®) polymers, the following analyses were performed:

Thermal properties (TA) were determined by dynamic mechanical

thermal analysis (DMTA), thermogravimetry (TG), and differential

scanning calorimetry (DSC). These analyses were performed in order

to study important parameters such as the determination of the glass

transition temperature (Tg). The Tg value determined via DMA or DSC,

is one of the most important parameters in the chemical

characterization of the polymers, since the Tg links the chemical and

functional properties. The degradation temperature and other

degradation properties of the biopolymers were studied by TG.

o In order to measure the Tg by DMTA spectra, two different

samples were used: films plasticized with glycerol and produced

by the casting method, and films produced by hot press method

after the extrusion of Polyplast®-glycerol to produce the

respective thermoplastic polymer. In all cases native starch was

used as control. The mechanical properties of these materials

were corroborated by using a universal mechanical testing

instrument and the appropriate standard (ASTM D638, Type I

dog-bone specimens).

Composites produced with micro-clay were produced in

order to investigate its interaction with the modified

starch polymers (Polyplast®). The composites were also

analyzed by DMTA. The mechanical properties of these

21

The flow properties were studied by using a capillary rehometer.

Information related to the shear viscosity and shear rate is generated

using this instrument. The determination of these parameters is

important since the capillary rehometer mimics the process by which

the material is forced through a nozzle similar to an extrusion process.

Spectroscopic analyses (FT-IR, FT-Raman, 1H NMR, 13NMR, as well as

XRD) were performed in order to study the molecular structure of the

the materials. To determine the sugar and polymer composition the

following techniques are used: high performance anion exchange

chromatography with pulsed amperometric detection (HPAEC-PAD)

and matrix assisted laser desorption/ionization time of flight mass

spectrometry (MALDI-TOF MS).

Important morphological features of modified starches were

determined using scanning electron micrographs and optical

microscopy.

22

1.4.3. Structure of thesis

According to this brief introductory chapter, the thesis is separated into five

chapters:

Chapter one reviews important topics related to the molecular

structure of starch and its influence on the enzymatic and microbial

modes of action. It gives also a quick introduction to this project,

including its objectives and significance.

Chapter 2 explains the process followed to produce the

polysaccharides which are used through this work.

The third chapter show the results obtained through the various

analytical techniques starting from the chemical and microscopic

analyses, followed by viscoelastic and thermal properties, and finally

the mechanical properties.

General conclusions are reported in section 4.

The future work is briefly described in section 5.

The sources used to support this investigation are listed in section 6.

23

2. Experimental

2.1. Production of bio-polymers

Native potato, and tapioca starches were obtained from Jack Hua Company

Limited (Thailand) and Wind Mill imported by Western Rice Mills LTD,

Canada), respectively. Potato dextrose broth (PDB) was from Sigma-Aldrich.

Corn starch was from CASCO Co. Canada. Production of modified starches

were carried out according to the methods followed by Jeng et al. [42] and

Huang et al. [43]. Modified starches were produced by fermentation with

isolates of the fungus Ophiostoma.

The process consisted in the production of the stock culture and further

inoculation the culture media containing the starch. The basal culture

medium used for preculture and culture production, was with minor

modifications, the method described by Selbmann et al. [76]. The growth

medium contained 2 g of yeast extract, 10 g of glucose, 1 g of KH2PO4, 0.1 g

of MgSO4.H2O, 1 g of ZnSO4 solution (36 mg in 100 mL of distilled water),

and 1 g of FeCl3.6 H2O solution (48 mg in 100 mL of distilled water), per liter

of distilled water. The pH of the medium was 4.5± 0.5. This stock solution

was incubated for 72 hours (or until mycelia dry weight of 7-8 g per liter was

obtained) in a shaker at 150 rpm at room temperature.

For the production of the polysaccharides (bio-plastics), objective of this

study, 20 g (dry weight) of corn, potato, or tapioca starch were feed to one

liter of culture growth (prepared with the same formulation as stated

previously). Potato dextrose broth (PDB) was used as specified by the

provider. The growth medium was sterilized by autoclaving for 20 minutes at

121oC (250oF). The process was subsequently improved to reach a mass

production of ~300 g of polysaccharides per liter of culture media. The

recovery of polysaccharides was performed either by using ethanol (99%) or

24

by ultracentrifugation (see the referred articles). Spores were removed from

the solution by using ultracentrifugation at 16000 X g for 20 min. The

polysaccharides were recovered by precipitation with two volumes of ethanol

(98%) at room temperature from the cell free solutions. Polymers were

freeze-dried and weighed for yield determination.

Since the processing conditions have been constantly changed to improve the

mass production, the overall results shown in this work represent an average

of a minimum of 10 different batches all with the same properties. To ensure

quality control, the batches were analyzed by spectroscopic and thermal

properties. These polysaccharides were used with minor changes according

to the requirements of the different techniques during the chemical analyses.

Details when necessary are provided for each particular technique.

2.2. Protein determination

The presence of protein (~0.2%) was confirmed from free cell starch-

polymer samples by using crystallized bovine albumin (Sigma Co., U.S.A.) as

standard and Coomassive brilliant blue R-250 as the dye reagent, method

from Bio-Rad Bradfor Protein Assay with BSA (bovine serum almbumin; Bio

Rad Co., Technical note 1069) at a UV adsorption of 595 nm. Therefore,

further analyses were performed with relatively pure polysaccharides.

25

3. Results and analysis

3.1. Morphology and chemical analyses

3.1.1. Morphology (SEM and FT-Raman confocal analysis)

3.1.1.1. Introduction

In the structure of the starch granules, pores, channels, or voids can be

observed by using scanning electron microscopy. These structures may allow

the dispersion or flow of water, solvents, chemical or enzymatic reagents in

starch. The removal of channel-associated proteins, would also allow for new

and important starch derivatives [82-84]. Proteins located in the channels

may interfere with the flow of the different solutions. The removal of proteins

with enzymes e.g. proteases, may allow for better internal reactions. During

enzymatic treatments of the starch granules pin-holes on the surface and

strong internal degradation are produced [82]. The crystalline areas are less

accessible to chemical or enzymatic substitutions, but they may be affected

in first instance near to the branching points as noted above. Cross-linked

corn starch hydroxypropyl ethers, for example, were found to be substituted

at these points. Substitutions may introduce bulky molecules which interfere

with the tendency of the amylopectin to retrograde in solution. Therefore,

substituted amylopectins may remain more time in solution. Methyl

substituents in potato starch have also been reported in the amorphous

regions of the amylopectin clusters. In 2-nitropropyl starch the susbstituents

are almost exclusively bound to the amylose molecule. This could indicate a

relationship with initial complex-building between the starch and the reagent.

Enzymatic and acid hydrolysis of cationic waxy corn starch showed that

cationization in an aqueous slurry predominantly occurs in the amorphous

areas of the granule, and especially near the branching points of the

amylopectin molecules, on the surface, and inside of the channels.

26

3.1.2. Materials and methods

The morphology of the starch granules after fungal treatment was examined

by using a Hitachi S800 scanning electron microscope at an accelerating

voltage of 10kV. The samples were coated with platinum-palladium for the

analyses. The confocal optical microscope Senterra (Bruker Optics USA) was

used for the visual inspection of the samples.

3.1.3. Results and discussion

Starch occurs as discrete semi-crystalline granules of various sizes, size

distribution and shapes. The size of the granules varies from ~20 to 80 m.

Fig. 1 shows an optical microscopic image (confocal microscope FT-Raman)

of some granules (potato from Maple Smell, Summer Star Trading Co. LTD

Toronto, Canada). The pith of the granule can be observed at the centre and

the alternating growth rings in dark and white contours. The width of the

alternating rings or shells is ~1200-4000 Å/~0.1 to 0.5m.. Scanning

electron micrographs (SEM) showed the presence of pores in the surface

(Fig.2).

The granular aggregation is shown in Fig. 3. It is interesting to note in this

micrograph that the granules adopt the shape according the neighboring

granules saving room within the main structure that holds the granules (i.e.,

kernels in corn). Figs. 4 and 5 show the complexing of amylose-iodine and

therefore the regions of amylose distribution within the starch granules.

These images were obtained by using an optical microscope with a resolution

of 1 m (FT-Raman confocal optical instrument). The smallest structures

detected by AFM were observed in small round structures of 300 nm (3000

Å). These structures are probably the so called “blockets”. Each blocket is

formed by concentric layers of 40-70 nm (400-700 Å) [51, 52].

27

Figure 1- Confocal FT-Raman microscope observations of a variety of used commercial potato starch; Scale= 10 microns

Figure 2- SEM image showing the porosity at the surface of the granules Scale=6.1 m

28

Figure 3- SEM micrographs showing granular aggregation in native starches Scale 20 m

Figure 4- Confocal microscope FT-Raman- amylose-iodine complexing denoting the amylose fractions within the granules-Scale 20 m

29

Figure 5- Confocal microscope FT-Raman- amylose-iodine complexing denoting thick layers of amylose fractions within the granules-Scale 20 m

It is possible to observe from the micrograph of NS the porous nature of the

native starch granules as well as a smooth surface. Electronic micrographs of

native and treated starch granules are shown in Fig. 6. These images show

that the fungal treatment may result in the damage and/or blocking of the

pores as well as in a visible rough appearance of the surface, this is believed

to be caused by the deposition of some polysaccharides within the granules.

Chemical analyses of the starches suggest that the hydrolysis should occur

primarily at the amorphous areas of the granules. Some of these patterns

are coincident with a typical enzymatic degradation [58, 83]. Fig. 7 also

shows that modified starch granules are physically affected after the

modification. The granules are affected on the surface and others collapsed

or showed visible damage at the pith.

30

Figure 6- Optical images of modified starch granules

.

Figure 7-SEM images of modified starch granules

31

3.1.3.1. Conclusion to this section

Some important structural features of the starch granules can be determined

by SEM and FT-Raman confocal optical microscopy such as the crystalline

pattern and surface porosity. The amylose fraction in these particular starch

sample seems to be located in specific growth rings (and it is distributed

randomly within the granules), feature which may influence the enzymatic

activity. Although no evidences of damage to the crystalline regions of

modified starch granules were obtained by XRD analysis, the scanning of the

granular surface of the granules showed visible damage associated with

typical patterns of enzymatic activity.

3.2. XRD

3.2.1. Abstract

X-ray diffraction patterns of modified starches did not show changes in the

crystalline composition respecting native starches. Gelatinized and dried

starches showed the lack of crystalline regions due to the disruption of the

amylopectin ordering regions.

3.2.2. Introduction

By using small angle X-ray scattering Oostergetel and van Bruggen[53]

proposed a model for the organization or architecture of the amylopectin

molecules in which these molecules form a three-dimensional tetragonal

super-helix structure which was estimated to be 40 to 70 nm by using small

angle X-ray scattering. The organization of the amylopectin molecules is

complex. It is believed that one hypermolecule of amylopectin carries just

one reducing end. The branches in the amylopectin occur at the C-6 position.

They are on average 30 glucose units. There are branches which are not

32

substituted in the C-6 position (A-type branches) and substituted branches

(B-type). The branches may be occurring in pairs which twist together in a

helical fashion. Six of these double chains form an individual crystalline

structure and lead to progressive accumulation of structures up to the

formation of the three-dimensional structure. What occurs in this process has

not been totally explained and the denomination of some of the cumulated

structures were given the name of ‘stacks’.

It has been clearly observed by X-ray diffraction that there is a constant 9

nm in periodicity due to the crystalline regions of starch [48, 49]. Based on

the X-ray diffraction patterns there are three allomorphs identified with

cereal (type-A), tuber (type-B) and pea starches (type-C). B-type starches

are produced if the central cavity of the crystalline hexagon (the middle

channel) is filled with some water molecules, whereas the type-A crystals

have another double-helix within the middle channel giving rise to a densely

packed crystallite. Type-C is a combination of A- and B- allomorphs.

The lamellae responsible for the crystalline regions, are formed by three

discrete components the backbone which supports the double helices,

parallel ‘rigid’ double-helical (menogenic) units (~5-6 nm) and amorphous

regions (more flexible un-branched regions, also called ‘spacers’ or side

chains) with sizes of ~4 nm. It has been observed by Waigh et al [48] that a

simultaneous appearance and disappearance of the 9 nm and 1.6 nm

reflections in small angle X-ray scattering occurs under hydration and

dehydration experiments. The hydration produces the 9 nm reflection due to

the smectic periodicity. With a ~10% mc solid state NMR spectra show a set

of 4 sharp with a doublet or triplet at ~100 ppm (Three peaks in starches

type A, and two in type B) associated to the crystalline regions—crystallinity

in corn is ~40% and 20% in potato starch. Neither dehydrated native

granules (<5% mc) nor the amorphous dehydrated starches show these

signals. The same phenomenon occur for highly hydrated starches (~20%

mc <). These authors suggested that under dry conditions the starches may

33

be in a pure glassy form (<5% mc), while moisture contents of ~10 % allow

the formation of crystalline regions. These particular structures of

intermediate order are known as liquid crystals; SCLCP (side chain liquid

crystal polymer). The degree of mobility of these three components, coupled

with the helix-coil transition, may be used to explain physicochemical and

structural properties like gelatinization, dehydration or molecular

composition.

3.2.3. Materials and methods

Starch granules were examined by using a Hitachi S800 scanning electron

microscope. The granules were mounted on an aluminum stub using double-

sided adhesive tape. The stubs were platinum-palladium coated and the

starch was viewed at an accelerating voltage of 10kV.

X-ray diffraction patterns (PXRD) of the starch samples were obtained using

a Shimadzu S6000 diffractometer operating at 40V and 30 mA (Cu Ka

radiation of 0.154 nm). The samples were tested without prior treatment.

The intensity was measured from 5 to 40° as a function of 2and at scanning

speed of 0.5/min and step size of 0.05°.

3.2.4. Results and discussion

PXRD patterns of analyzed starch samples as shown in Fig. 8. GMS (granular

modified starch), GNS (granular native starch), DMS (gelatinized modified

starch), NS (gelatinized native starch). Gelatinizes starches showed the

pattern of amorphous materials, and the patterns of GMS and GNS were

similar showing some crystallinity. Hence, no new crystalline formation or

degradation in the amylopectin fraction occurred after the modification in the

starch granules (GMS or DMS). Moreover, both starches showed a peak at

approximately 9°C after gelatinization which may be characteristic of

amylose resistant regions of starch

34

Figure 8--PXRD patterns of granular native starches (GNS), native gelatinized starches (NS), modified gelatinized starches (DMS), and granular modified

starches (GMS)

3.2.5. Conclusions

The modification of granular starches do not affect the crystalline regions of

the starch. Therefore, the modification may be occurring on the amylose

fraction and due to the deposition of fungal exo-metabolites.

3.3. FT-IR (ATR)

3.3.1. Abstract

FT-IR spectroscopy was used to study the chemical properties of modified

starches. These polymers were produced from commercial starches as shown

in section 3 (cf. “Starch-like exopolysaccharide produced by the filamentous

fungi Ophiostoma sp.” Jeng et al., 2007, Forest Pathology, 37, 80-95).

35

In this study gelatinized as well as undisrupted starch granules exposed to

the fungal attack were analyzed by FTIR spectroscopy. The treatment before

analysis consisted in the separation of the starch mass from the fungal

spores by centrifugation, recovery of the polymers and freeze dried before

analyses. Characteristic FT-IR molecular vibrations of native starches were

compared to those found in modified starches. Specific vibrations; which

were hidden during the FT-Raman analysis by fluorescence or by the working

conditions of the instruments (lack of polarizability of the molecules),

produced strong bands associated with amide I and II, which are probably

directly related with the process of modification. This analysis showed a

potential route in the process of the starch modification in presence of these

microorganisms.

3.3.2. Introduction

Carbohydrates, particularly starch, have been widely analyzed by FT-IR or

FT-Raman spectroscopies. The assignment of the most important IR and

Raman bands present in native starches are relatively well established. Since

the main differences between these analytical techniques are in the working

physical principles of the instruments, the main advantages and drawbacks

of the instruments as well as the detection and/or intensity of the signals

such as presence of water, fluorescence and symmetry of the molecules

(generation of the dipole moment or polarizability) may be used to make the

characterization.

In general, bands at 578 and 528cm-1 belong to skeletal modes—low

frequency vibrations of the ring, etc. The band at 2850cm-1 is attributable to

CH2 groups, and the band at 1463 cm-1 to OCH and CH2 groups. If the

intensity of the vibration is reduced it may be associated to reduced mobility

of the functional groups or shifting of the peaks may be attributable to new

molecular interactions [85].

36

Carbohydrates have many strong bands in the 1160-1000 cm-1 region

involving stretching of the C-O bonds of C-O-C groups. The strong absorption

near 3350 cm-1 is associated to OH stretch. The shape of this curve is

affected mostly by hydrogen bonds. There is medium-weak CH stretch band

near 2900 cm-1 and multiple medium bands in the 1460-1200 cm-1 region

involving CH2 deformation, CH and CH2 wag, and OH in-plane deformation.

The medium weak bands in the 960-730 cm-1 region have been the most

studied resonances in the characterization of the different types of

carbohydrates. In the pyranose type sugars, bands associated to in-phase

ring stretch at 770± 14 cm-1 and resonances involving COC out-of-phase

stretch in the ring have been also reported at 917 ± 13 cm-1. and

anomers have been also reported at 844 ± 8 cm-1 and 891 ± 7 cm-1

respectively.

3.3.3. Materials and methods

Spectrophotometers used in this study were: FT-IR model Tensor 27 from

Bruker Optics, USA. KBr pellets or ATR mode were utilized for the analysis by

FT-IR. The samples were analyzed with a resolution of 4cm-1 and an average

scanning time of 1 min. The spectral resolution for samples tested in the

Tensor 27 (KBr pellets) was within the range 4000 to 400 cm-1. The spectral

resolution for samples tested in the ATR mode was within the range 4000 to

600 cm-1.

The FT-Raman spectra were collected on a Senterra from Bruker Optics

(USA) at operating wavelengths of 785 (at 100mW) and 532 (at 20 mW).

Data for Raman mapping over the starch granules were collected randomly

over an areas no bigger than 50.0X50.0 microns (the size of the potato

starch granules analyzed was around 40 m, while the starch granular size of

corn and tapioca averaged ~10 m). Resolutions allowed with this instrument

were ~9-12 cm-1 (with one spectral range of 70, 3500) and ~3-5 cm-1 (with

37

various spectral ranges). Data were processed using the integrated software

Opus version 6.0.

3.3.4. Results and discussion

3.3.4.1. FTIR

FT-IR and FT-Raman are routinely used in research and for process control

[86]. In particular, these techniques have been widely used since many

decades ago to study the structural variations of starch subjected to chemical

or physical modifications [87, 88]. Potential variations in the chemical

structure of starch after the fermentation were first studied by FT-IR

spectroscopy. The results are shown in Table 2. IR bands between 2800-

3000 cm-1 are related to stretching of CH2, and bands around 2920 and 2850

cm-1 are generated due to the asymmetric and symmetric stretching of

methylene, respectively. Bands associated to C-O stretching at 900-1260

cm-1 showed a broader absorption band range in modified starches, some

chemical interaction may also be occurring at this bonding position. Peaks at

1472-1466 cm-1 arise due to scissoring vibrations of CH2. The scissoring

bands showed variations with interchain interactions, packing arrangement

and ordering of methylene groups [89]. These bands showed drastic changes

in the position after the starch fermentation. The shifting of the band at

2927 to 2922 cm-1 showed a slight variation, and suggests a lower molecular

mobility after the fermentation. The band associated to O-H stretching [89-

95] also showed the shifting from 3447 to 3421 cm-1, indicating also a lower

molecular mobility. The band located at 1636 cm-1 in native starches shifted

within a broader spectral range from 1649 to 1623 cm-1.

38

Table 2- FT-IR data: GMS (granular modified starches); DMS (gelatinized

modified starches); GNS/NS (granular or gelatinized native starches)

W cm-1

GMS W- cm-1

DMS W cm-1

GNS/NS Assignments for the starch molecule

based on its molecular structure and comparison with general

vibrations

Assignments of functional groups for spectral peaks wavenumbers in cm-1

3421 3424 3447 O-H stretching - 2964 2963 - 2962 ± 10 C-H stretches of alkenes in CH3

asymmetric vibrations 2922 2923 2927 C-H asymmetrical stretch vibrations

in CH2 2926 ± 10 C-H asymmetrical stretch in alkanes in CH2 vibrations 2925± 5 CH3 symmetric stretch in benzene rings (low probable)

2882 - - - 2872 ± 10 C-H stretches of alkenes in CH3 symmetric vibrations

2852 2853 2853 C-H symmetric vibrations in methylene

2855±10 C-H stretches of alkenes in CH2 symmetric vibrations

1649-1623 1629-1619

1636 H2O entrained in the sample; O-H bending vibration

Also may show C=O stretching and C–N stretching

1566 - - Aromatic ring in lignin and others 1470-1460 1465 1459 Scissoring vibrations of CH2 C–H bending in alkyl groups

1441 - - C-H bending 1418 1418 1420 - C-H deformation

1400 - - C-H deformation 1366 1384 1375 - C-H vibrations 1339 - 1344

(small shoulder)

-

1262 1262 - - 1280-1185 C-O-H deformation and C-O stretching in phenolic compounds. N-H bending in amine III

- - 1256 (small

shoulder)

C-O-H deformation in starch

- 1237 - - C-O-H deformation due to new inter- and intramolecular bonding

1165-1153 1175 1162 - 1144 -

1107-1080 1086 1090

O-C stretch within the anhydroglucose ring

- - 1041 - - -

1017 (1055-984)

1023 1021 C-O stretch in the ring -

923 929 926 C-O bending in starch - 863 863 858 C-H bending 810-750 out of plane C-H bending bands in mea-

substituted benzene rings - 846 - -

799 803 805 (small shoulder)

- out-out-plane C-H bending in mono- and disubstituted benzene rings 810-750 with ring bend (690±10) 860-790 no ring bend

764 764 763 - Possible C-H rocking 752 - - 740 - -

707 717-707 712 - - 661 661 - 1350±50 and 650±10 OH bends in benzene rings 614 622 613 - Probably due to halogenated compounds, which

are found in this range and up 575 571 573 - - 533 531-518 526 - - 479 490-455 - - - 436 426 - - -

39

The Figures from 9 to 12 show the different spectra of EPSs

(exopolysaccharides) and GMS/DMS (granular or gelatinized modified

starches) which were recurrent in the multiple samples tested. New peaks

with high intensity were depicted at ~2964±5, 1745, 1262 and 800 cm-1.

The peak at ~2964±5 cm-1 was associated with to Va(C-H) vibrations in CH3

(Va; asymmetric stretching vibrations). The band at 2855±10 cm-1 was

associated to Vs(C-H) (Vs; symmetric stretching vibrations) in CH2. The band

at ~ 2925 ±5 cm-1 corresponds to Va(C-H) in CH2. Double bonds may be

associated to the chemical vibration at 1745 cm-1. A peak with high intensity

also arose at the region of C-O-C stretching (bands at ~1000 and 900 cm-1)

in GMS or DMS from tapioca starches. The spectrum corresponding to EPS

which were produced in the absence of starch displayed the peak at ~1747

cm-1 which may be associated to the presence of double bonds (Figs. 9, 10

and shoulder in 11 and 12). The peaks appearing at ~1262 and 800cm-1

were further localized in water soluble fractions of modified starches, but not

in water insoluble fractions (Fig. 13). Hence, the two different spectra may

be attributed to two different stages in the process of modification.

4000 3500 3000 2500 2000 1500 1000 500

0.00

0.01

0.02

0.03

0.04

0.05

0.06

Abs

orb

ance

uni

ts

Wavenumber cm-1

1745

2854

2925

1081

3421

1465

1261

Figure 9- FT-IR spectrum of modified starches- (detection of the peak associated to double bonds probably in C=O vibrations)

40

4000 3500 3000 2500 2000 1500 1000 500-0.01

0.00

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

Abs

orb

anc

e u

nits

Wavenumber cm-1

1747

Figure 10- FT-IR spectrum of exopolysaccharides (EPSs) produced in absence of substrate (detection of the peak associated to double bonds probably in

C=O vibrations)

4000 3500 3000 2500 2000 1500 1000 500-0.002

0.000

0.002

0.004

0.006

0.008

0.010

0.012

0.014

0.016

0.018

Abs

orb

anc

e u

nits

Wavenumber cm-1

800

1260

2855

29252964

Figure 11--FT-IR spectrum of granular modified starches (G-MS) (detection of the peaks at ~800 and 1240 cm-1)

41

4000 3500 3000 2500 2000 1500 1000 500

0.00

0.01

0.02

0.03

0.04

0.05

Abs

orb

ance

uni

ts

Wavenumber cm-1

Figure 12- FT-IR spectrum of exopolysaccharides (EPSs) produced in absence of substrate (detection of peaks at 800 and 1240 cm-1)

A clear separation of the peaks at 800 and 1240 cm-1 was achieved in two

different fractions of the starch one soluble and the water insoluble. Both

fractions are depicted in their respective FTIR spectra in Figure 13. In the

spectrum of the “water insoluble” fraction it was observed the presence of

two characteristic peaks at 800 and 1260 cm-1. The vibrations, related to CH2

and CH3 stretching, were also consistent through the process of modification.

These vibrations appeared with lower intensity or just slightly depicted in the

fractions denoted as “water soluble”. It was also possible to observe the

change profile of the vibrations related to C-O-C (1055-984 cm-1) in these

water soluble fractions. Two distinctive spectra were taken for the EPS

polymers. It was also possible to observe the two peaks at ~1260 and 800

cm-1. However, the shoulder at ~1740 showed higher intensity, and in

general, the spectral profile presented broader peaks.

42

Water insoluble fraction

1259

7972916

2849

2961

Water soluble fraction

500100015002000250030003500

Wavenumber cm-1

0.00

00.

005

0.01

00.

015

0.02

00.

025

0.03

00.

035

Abs

orba

nce

Uni

ts

Figure 13-FT-IR spectrum of modified starches- separation of water-like and water insoluble fractions

3.3.4.2. FTIR (ATR)

As it was shown before, in starch the IR bands between 2800-3000 cm-1 are

related to the stretching of CH2, and bands around 2920 and 2850 cm-1 are

generated due to asymmetric and symmetric stretching of methylene,

respectively. Peaks at 1480-1450 cm-1 arise due to scissoring vibrations of

CH2. These scissoring bands are sensitive to variations with inter-chain

interactions, packing arrangement and ordering of methylene chains. The

band at approximately 1650 cm-1 is associated with water strongly bonded to

the starch molecules. The band at 960-1100 cm-1 is related to C-O-C

stretching in the glucopyranose ring.

43

The spectrum of native-starch/clay/glycerol composites and pure clay are

shown in Fig. 14 (see section 3.9.4.4). It was observed that the band at

~1023 cm-1 remained without change from native to modified starches. It is

also interesting to note the disappearance of one of the bands of the

spectrum of clay at ~3600 cm-1 in both starch composites (native and

modified) which suggests some molecular interactions.

Fig. 15 shows the ATR spectra of two samples of modified-

starch/clay/glycerol composites. In these samples new bands were observed

in the region of CH2 at 3000-2800 cm-1 and ~1470cm-1 (branching points of

starch), at ~1590, 1545 and 1655 cm-1. The band near the 1650 cm-1

attributed to water strongly attached to the starch did not showed a drastic

change in native/starch-clay-glycerol composites, but shifted in

modified/starch-clay-glycerol composites showing also a strong band. In a

recent study, it was found by solid state 13C NMR a larger anisotropic

interaction in the chemical shift associated to C6 and the broadening of the

CH2 band by FT-Raman spectroscopy (as it was shown by solid state NMR).

To investigate the source of these bands, the analysis of exo-polysaccharides

(EPSs) and protein-like compounds were analyzed by FTIR. The spectra of

the protein fraction showed bands at ~1650, 1540 cm-1 as well as strong

bands at ~1440cm-1 corresponding probably to more complex vibrations

associated to OH groups. A comparison with a known protein showed that

these bands correspond to amide I and II respectively. However, the band at

1650 and 1450 cm-1 were also detected in EPSs (spectra not shown).

44

100020003000

0.0

0.8

100020003000

0.0

0.4

Wavenumber cm-1

AT

R U

nits

Figure 14---Attenuated total reflectance (ATR) spectrum of native-starch/glycerol/clay composites (top) and clay spectra (below)

3697

32743307

2921

1651

1544

1454

1023998

100015002000250030003500

1590

16592880

29252954

3311

1023

100015002000250030003500

Wavenumber cm-1

AT

R U

nits

Figure 15--Attenuated total reflectance (ATR) spectrum of two different samples of modified starch clay glycerol composites showing complementary information related to new molecular interactions

45

3.3.5. Conclusions

The increase in mechanical properties of starches modified by fungal isolates

of the genus Ophiostoma has been reported in previous studies [21, 24]. In

this study, some of the potential sources of the functional properties of these

polysaccharides were investigated. Analysis of the glass transition

temperature (Tg) by DMTA on various specimens of modified starches showed

that various thermal transitions and new chemical bonds are produced due to

the fungal modification of the starch. These thermal relaxations may be

attributed to exopolysaccharides or protein-like compounds produced by the

fungi during the process of modification. Moreover, while native-

starch/clay/glycerol composites displayed a separation of the glycerol phase,

the samples prepared with modified starches showed a better affinity

towards the filler (kaolin clay) which was used as a reinforcing material.

Finally, it was found that the increase in the stiffness of the modified starches

is due to the reduction of the molecular motion produced by cross-link type

bonds occurring mainly at C6 (the branching points).

The FTIR spectra of modified starches and EPSs produced by the fungi in

absence of starch overlapped indicating similar chemical composition. New

peaks found in modified starches indicate the chemical molecular

modifications. These peaks can be associated to the CH2 stretching vibrations

at the methylene groups in C6 and double bounds. Also new peaks at 800

and 1250 cm-1 were detected. However, these peaks cannot at the moment

be associated with any particular molecular vibration.

46

3.4. FT-Raman

3.4.1. Abstract

The tracking in the process of the starch modification in presence of

Ophisotoma spp. was intended by using FT-Raman spectroscopy. Some

important changes in the chemical structure of starch were observed.

However, change in some characteristic bands associated to the process of

liquefaction (480cm-1) or saccharification (910-935 cm-1 region and 1127 cm-

1) were not observed within the 3rd day of modification. However, a broad

band was detected at the CH symmetric stretching vibration suggesting

strong molecular interactions at the methylene group, but not appreciable

degradation was observed in the X-ray diffraction pattern of modified

starches. Although, some bands at the region of double and triple bonds

were detected, just the band at ~2629 cm-1 was consistent within the

variation of ±5cm-1. Therefore, it is concluded that the presence of some

chemicals are influencing the stronger molecular interactions in these

materials.

Raman scattering is inherently an inefficient process, a practical obstacle to

Raman studies is interference from fluorescence or phosphorescence.

Troublesome fluorescence can arise either from impurities or from the

intrinsic relaxation process resulting from an electronic absorption of the

sample in resonance Raman spectroscopy. Crystalline regions of starch or

packing of the material could produce delays in this process yielding a curved

spectrum (as occur with potato and tapioca or corn starches). The Raman

spectrum yielded a curved profile in corn or tapioca starch granules while the

respective spectrum taken from potato produced a flat base line. These

variables can also influence the intensity of the Raman bands when the

molecular excitation is produced with a laser at 532 nm. The effects were

reduced by using an excitation wavelength of 785 nm. However, the mapping

47

spectrum of native starches produced particular spectral deformation not

found in native starches.

It was concluded that fluorescent substances remained in the starches

samples after the modification and that the functional groups sensibly

affected the CH symmetric and asymmetric modes, particularly the CH

symmetric stretching at the C6 in the glucopyranose ring.

3.4.2. Introduction

For FT-Raman spectroscopy the most important bands found in native

starches include the strong adsorption at 3000-3500 cm-1 and ~2910 cm-1

related to OH and CH stretching modes. The relatively low intensity signals at

1460, 1380 1339 and 1262 cm-1 include respectively the CH2 symmetric

deformation, CH2 scissoring and COH deformation, CH2 twist and COH

bending, and CH2OH (side chain) related mode. A series of signals between

1130 and 1050 would include COH deformations. The band at 940 ±6 cm-1 is

related to the skeletal mode involving the -1-4-glucose linkage (COC). The

signal at 866 cm-1 is associated to CH and CH2 deformation. The strong

resonance located between 614 and 440 cm-1 belong to the skeletal mode

related to C-C stretch in the ring.

3.4.3. Materials and methods

FT-Raman spectra were recorded by using a Senterra spectrometer from

Bruker Optics. This model incorporates a dual laser Raman spectrometer

module and a confocal microscope module. The confocal capability allowed

for visual inspection of the samples as well as spectral analyses. Two lasers

are integrated to this instrument 785- and 532-nm which were used at 100

and 20 mW respectively. The time of scanning was 30 seconds for all

samples. In serial mapping, spectra were obtained sequentially from a series

of positions using point-by-point scanning through a squared grid of ~150

48

m. Two wavenumer resolutions were used 3.5-5 cm-1 and 9-12 cm-1. The

3.5-5 cm-1 resolution allowed to record within different spectral ranges: 70,

1555; 1525, 2740; 2710, 3700; 70, 1555 and 1525, 2740; 70, 1555 and

2710, 3700; 1525, 2740 and 2710, 3700; 1525, 2740 and 2710, 3700; 70,

1555 and 1525, 2740 and 2710, 3700 cm-1. And just one spectral range of

70, 4500 cm-1 for 9-12 cm-1. All data was processed using the integrated

software Opus version 6.0. Calibration of frequency and intensity was

performed automatically with the integrated method SurCal®. Spectral

shape correction was also applied to the spectra. This special feature divides

the sample spectrum with a reference spectrum. The reference sample is a

fluorescence standard which produces the typical broadband curve of light.

The spectra taken when this feature is activated in the spectrometer will be

background corrected, reducing therefore the effects of fluorescence.

Comparison of relative frequencies of peaks and respective assignments were

performed according to the literature.

3.4.4. Results and discussion

Amylose and amylopectin form the starch granules, and are made up of -(1-

4)-linked D-glucose residues. Amylose is a linear polymer, whereas

amylopectin has a dense (1-6) branching pattern. Depending on the starch

source, the granules are considered to occur in three crystalline forms,

differentiated by their X-ray diffraction patterns. The A-form is found in

cereal starch, whereas the B-form occurs in tuber starches; the C-form is

rare and has been observed in, for example, tapioca, pea, and banana

starches. It seems that the A-, B-, and C-structures are very similar, and are

probably different hydrates having the same chain conformation. Many of the

spectral similarities and/or differences probably occur due to the different

interatomic distances between the molecules and/or intensity of hydrogen

bonds produced by interacting external molecules such as water. The

molecular motion of crystalline to amorphous allomorphs or new molecular

49

interactions of the starch molecules are sensitive to spectroscopic methods

such as FT-IR and FT-Raman.

In particular for this study, the FT-Raman spectra showed clear modifications

of the O-H and C-H stretching regions of the starch molecules after the

fungal modification, and also important spectral differences between native

and modified starches were observed in the region below 1500cm-1.

The Raman spectra of carbohydrates, particularly starch, have been analyzed

by Raman spectroscopy. Since the assignment of the most important IR and

Raman bands present in native starches are relatively well established, the

variations in peak intensity and shifting of the bands can be used for in situ

characterization of the materials (Table 6). The advantages and/or

drawbacks of the instrument (fluorescence and symmetry of the molecules;

generation of the dipole moment or polarizability) may also be used for

materials characterization. For FT-Raman spectroscopy the most important

bands found in native starches include the strong adsorption at 3000-3500

cm-1 and ~2910 cm-1 related to OH and CH stretching modes. The relatively

low intensity signals at 1460, 1380 1339 and 1262 cm-1 include respectively

the CH2 symmetric deformation, CH2 scissoring and COH deformation, CH2

twist and COH bending, and CH2OH (side chain) related mode. A series of

signals between 1130 and 1050 cm-1 would include COH deformations. The

band at 940 ±6 cm-1 is related to the skeletal mode involving the -1-4-

glucose linkage (COC). The signal at 866 is associated to CH and CH2

deformation. The strong resonance located between 614 and 440 cm-1 is

generally associated to the skeletal mode related to C-C stretch in the ring

[93, 96, 97].

The use of two different laser sources at different conditions (532 nm/20 mW

and 785 nm/100 mw with resolutions between 3-4 cm-1 and 9-12 cm-1) it

was possible to obtain important information related to the molecular

composition of modified starches. By comparing the Raman spectrum of

50

modified starches GMS/DMS and NS it was observed that the band

corresponding to CH stretching vibrations (~2900 cm-1) in modified starches

displayed broader profiles (from 2975 to 2878 cm-1). It was also observed in

the Raman spectra of these bio-polymers chemical vibration at ~2630 cm-1

but not in native starches (Fig. 16, Table 3). The symmetric scissoring and

twist deformation of CH at 1380 cm-1 found in the substrate did not appear

after the starch’s modification. The spectra of EPSs produced in absence of

starch were partially coincident with the spectra of fermented starches

showing that EPSs are being produced during the fermentation. Fig. 17 A-B

shows the current model of the starch molecule. Fig. 18 shows the potential

sites of enzymatic substitutions. As it was explained in the general

introduction of this work, the molecular substitution (chemical or enzymatic

occur mainly in the CH2 groups).

Fermented starches2975

2878

Unmodified starch

2914

2624

EPSs produced in YE

2000250030003500

010

000

3000

050

000

Ram

an I

nten

sity

Figure 16 -FT-Raman spectrum of the substrate, modified starches, and exopolysaccharides produced by the microorganisms in absence of substrate

51

Figure 17-A-B –Oostergetel and Van Bruggen model of the amylopectin clusters, branching and molecular pattern (A); the left-handed three-

dimensional helical structure of amylopectin (B). It’s been explained by the authors of this model [53] that neighboring helices are shifted relative to

each other by half the helical pitch (indicated by 0 and ½).

Figure 18-- Substitutions occurring in amorphous regions of the amylopectin molecules near the branching points [53]

52

Table 3-

Assignment of the most important Raman bands of the native and modified

star

Spectral assignment Nativ ches -1

fied starches

m-1

ches

e star

cm

Modi

c

Skeletal modes of pyranose ring

CH and CH2 deformation 866

85 (not consistent)

volving 1-4-glycosidic 940

COH deformation

d mode

CH2 twist, COH bending 1380 326

ring, CH and COH 1380

407

CH2 symmetric deformation 1460

OH stretching modes 3000-3500 Broader vibration

441

478

576

614

8

Skeletal mode in

linkage (COC)

1052

1082

1126

CH2OH (side chain) relate 1262

1

CH2 scisso

deformation 1

1697

2629

CH stretching modes 2910 2878-2977

53

The spectrum of modified starches (cd2c/MSP) samples is shown in Figure

19. The main spectral feature is the presence of fluorescence (probably due

to protein-like compounds). The bands at ~3365 and 2911 cm-1 associated

with the OH and CH stretching appeared with high intensity. The band with a

relatively medium intensity at ~2626 cm-1 was consistent in various samples

of modified starches. In this sample (undisrupted modified starch granules;

cd2c) the spectral features corresponding to the starch fingerprint region

seem to remain unaltered. However, detailed spectral features of modified

versus native starches, shown in Fig. 20 suggested also important spectral

variations.

The band observed in Fig. 19 at ~3390 cm-1 was related to OH and the

vibration associated to the CH stretching (between ~2970 and 2880 cm-1)

displayed broader profiles than native starches. The vibration at ~2626 cm-1,

slightly depicted in these images, was recurrent in most samples of modified

starches. The spectral range linked to double and triple bonds (2800-1600

cm-1) is shown in Fig. 20. It was observed the presence of at least 4 intense

resonances at ~2627, 2404, 1979 and 1694 cm-1. Most of the lines occurring

at the fingerprint region shifted slightly, and new bands appeared. Important

modifications in these spectra probably occurred at the CH2 scissoring, CH

and COH deformation, as well as at the CH2 symmetric deformations.

54

3358

2911

2624

50010001500200025003000350040004500

010

000

2000

030

000

4000

050

000

6000

0

Ram

an I

nten

sity

Figure 19- FT-Raman spectrum of Polyplast® samples- laser source 532 nm 20 mW; spectral range 70, 1555 -1525, 2740-2710, 3700 cm-1; integration

time 20 sec

2404

2627

1979

1695

160018002000220024002600

02

00

40

06

00

80

01

00

0

Ra

ma

n I

nte

nsi

ty

Figure 20- Spectra of native and modified granular starches (cd2c/MSP): laser source 532 nm 20 mW; spectral range 1525-2740 cm-1; integration

time 30 sec (noisy due to fluorescence)

55

Samples of modified starches obtained from gelatinized granules (ap3d/MPS)

were separated into two fractions based on their behavior in presence of

water. The spectrum corresponding to water insoluble fractions and the

water soluble fractions were strongly affected by fluorescence. Comparable

phenomenon was observed for PDB polymers. However, it was found a

strong vibration at the CH2 stretching region and the peak at 2630 cm-1, and

also at the bands associated to OH and CH after the modification. The

presence of the peaks at ~ 3570, 3560, 3519, 3469, 3196, 2903, 2825,

2807 and 2634 cm-1 was confirmed in several samples.

Visible Raman spectroscopy was also used to prove the presence of the

different starch morphologies after the modification. In-plane the substrate

structure (xy-axis) and the strain fluctuation (zy-axis) in native as well as

modified starch granules (cd2c) are shown in Figs. 21 and 22. The three-

dimensional (3D) plot results from a spatial-resolved measurement in the

xyz-axis. The physical units of the x-axis and z-axis are length units (e.g.

wavenumber and micron) and the y-axis shows the absorption intensity

versus the space. A surface area (not higher than ~30X30 micron) was

mapped using a visible Raman spectrometer excited with a Nd-YAG (at 532

nm, 20 mW in the sample). The analysis was performed randomly in all the

cases. Plots from native starch showed the typical spectral resonances in the

xy-axis together with a low projected intensity over the y-axis. The yz- plots

displayed a regular distribution without fluctuations. Modified starches

showed significant alterations in the spectra profile. The 3D (xyz) in its yz

spectral axis displayed two bands with prominent intensity projected over the

y-axis. It is interesting to note that both samples of modified starches

showed such fluctuations, however, more pronounced in cd2c samples. A

detailed analysis of the xy profile is also shown between 1500 and 400 cm-1,

for both samples cd2c/MSP and native starches. The main features in this

direction were not lost. Some weak resonances appeared after the

modification suggesting the lack and/or the presence of difference bonds

(bands at ~ 1209, 1198, double bands at ~780-756 and 719-7190 cm-1, and

56

619 cm-1). In general the shifting of the peaks over the xy profile was

observed through all the spectral range, but mostly in the region between

800 and 600 cm-1. The Raman depth profile showed an important structural

modification near the surface of the starch granules (Figs. 21 and 22).

500 1000 1500

0.0

2.0x103

4.0x103

6.0x103

Wavenumber cm -1R

aman

inte

nsity

Figure 21-FT-Raman scanning of the surface (3D) of the substrate (native starch)

57

500 1000 1500

01x1032x1033x1034x1035x103

Wavenumber cm -1

Ra

ma

n in

ten

sity

Figure 22-FT-Raman scanning of the surface (3D) of the modified starch

3.4.5. Conclusions

In general, it was observed in samples of modified starches the strong

influence of florescence probably due to the presence of carbon double and

triple bonds. The effects of fluorescence were stronger when using the laser

at 532 nm. However, this laser intensity gave the opportunity to observe

some bands not detected with the laser at 785 nm, i.e., the bands associated

to the OH and CH stretching and the detection of vibrations at the region of

double and triple bonds. In general, there was low variation at the

wavenumber value and intensity of the peaks at the finger print region of the

starch. The signals between 1130 and 1050 including COH deformations did

not shifted considerably from samples cd2c to ap3d. The band at 940 ±6

cm-1 related to the skeletal mode involving the -1-4-glycodidic linkage

(COC) was not clearly observed in the starch samples, neither in native nor

in modified starches, instead a slight shoulder appeared in this band. The

signal at 866 associated with the CH and CH2 deformation also shifted to

both sides (left or right) in samples of modified starches. The strong

resonance located between 614 and 440 cm-1 associated to the skeletal mode

58

C-O-C stretch in the ring did not show significant variations after the

modification.

These results are relevant due to the high yields produced after the

modification without undergoing in the starch degradation. Results obtained

from the three dimensional analysis indicated that the modification can be

also carried out over the surface of the starch granules. In addition the

Raman microscope allowed for the visual inspection of modified starches and

it was observed an important structural change. Finally, the images showed a

good dispersion of modified starch granules within the PLA matrix.

3.5. Liquid state NMR

3.5.1. Abstract

The use of 1H NMR of 300 MHz produced probably partial information related

to the chemical structure of modified starches. This information may be

related to the presence of pendant groups strongly solvated in the D2O. The

attachment of these groups may occur via methylen groups.

3.5.2. Introduction

The use of 1H NMR has been restricted to the detection of functional or

solvated groups appearing in polymers of high molecular weight. In general,

most advances in the interpretation of the molecular structure of starch and

its derivatives have been made with solid state NMR. Cross polarization

magic angle spinning (CP/MAS) 13C-NMR present broad lines, which have

been associated to the different positions in the glucose units, located at

~63ppm (C6; CH2-OH); 72 ppm (C2, 3, 5; CH-OH); a smaller line in 84 ppm

(C4; CH-O); and 103 ppm (C1 anomeric; C-O-C) [98, 99].

59

CP/MAS 13C NMR is sensitive to substances at low humidity content (rigid

state with low molecular mobility). This technique has been used to obtain

structural information of granular starches. The spectrum shows the position

of each one of the resonances at the respective carbons located in the

glucopyranose ring. At low humidity content the CP/MAS 13C NMR spectra of

granular starch exhibits a triplet in the signal assigned to carbon C1 in cereal

starches and a doublet in the case of tuber starch. Although the reasons for

the detection of these multiplicities have not been explained in detail, they

have been indeed associated with the presence of crystalline regions in

granular starches, and in specific with the 9 nm periodicity of the

amylopectin branches.

3.5.3. Materials and methods

NMR measurements were carried out on a Varian Mercury 300 MHz—1H, 19F,

13C, 31—5-mm gradient probe, deuterium gradient shimming, 100 sample

autochanger (SMS). Samples were dissolved directly in deuterium oxide. FID

files were also subjected to standard Fourier transformation and phasing.

3.5.4. Results and discussion

Water soluble samples were used for NMR analyses. The 1H liquid NMR

spectrum obtained for modified starches is shown in Fig. 23. In general, it

was observed that these starches had low solubility in water. The main

features of this spectrum were the chemical shifts at ~ 5.39, 3.82, 3.64 and

the triplet centered at ~3.94 ppm. Based on the 1H NMR spectra for D-

glucose and starch it is possible to highlight that the 1H liquid NMR spectrum

obtained from these water soluble fractions of modified starches are not

related to the type of native starch or starch derivatives from which they

were produced. Going back to the solid state NMR results, they showed that

modified starches are still formed by glucose units, but they also showed that

the molecules in such modified starches are subjected to strong molecular

60

interactions, mainly at the C6 site. Two alternatives, the potential presence

of two distinctive polymers (one of them produced by the starch

modification) and the other one the incorporation of specific functional

groups to the glucose ring in the starch molecules. The first alternative

implies that modified starches are a mixture of two polymers. On the other

hand, differences in molecular weight (shown by HPAEC-PAD or MALDI-TOF)

and captured in the liquid NMR spectrum may be produced by some new

functional groups. Such functional groups were strongly solvated in D2O, and

they are probably attached or reacting at the C6 site in the glucose ring. A

further simplification of the polymer probably occurred due to the removal of

a part of the coupling protons by deuteration.

As it is well known, proton chemical shifts fall within a range from 0 to 14

ppm. For mono-substituted functional groups, i.e., saturated hydrocarbons,

resonances occur generally between 1.0 and 4.0 ppm, while resonances for

olefinic protons appear in the region of 5.0 to 6.5 ppm. These values are

never exact and should be taken within the experimental limits. Olefins (-

C=CH-) are normally seen between 4.5 and 6.5 ppm. OH groups were

probably deuterated, whereas the set of overlapped resonances between 3.5

and 4 could be associated with a broad range of functional groups. However,

some specific groups attached to particular molecules of high molecular

weight in certain solvents have shown splitting into triads, as occurs with the

carboxyl group in some solvents) [100]. The 1H liquid NMR spectrum of

carboxyl methyl proton of poly(methyl methacrylate) in C6D6 is similar to that

shown in this study in which the proton of the functional group was solvated

in D2O. This empirical interpretation allows this particular functional group to

be related to a polymer of high molecular weight with random structure, but

showing patterns in the distribution of the functional groups with steric

structures meso-meso (centered at 394 ppm), meso-racemo (3.82 ppm) and

racemo-racemo (364 ppm).

61

The spectrum of exopolysaccharides (EPSs) produced by the fungus in

absence of starch is shown in Fig. 24. It basically shows the same pattern or

sets of peaks with a better resolution due to the averaging of anisotropic

NMR interactions (spin-spin and spin-matrix interactions).

5.5 5.0 4.5 4.0 3.5

Chemical Shift (ppm)

0.005

0.010

0.015

0.020

0.025

Nor

mal

ized

Inte

nsity

5.39

3.97

3.94

3.91

3.83

3.64

5.5 5.0 4.5 4.0 3.5

Chemical Shift (ppm)

0.005

0.010

0.015

0.020

0.025

Nor

mal

ized

Inte

nsity

5.39

3.97

3.94

3.91

3.83

3.64

Figure 23-- 300 MHz 1H NMR spectrum of Polyplast® polymers showing solvated, probably pendant groups, in D2O

62

Figure 24-300 MHz 1H NMR spectrum of EPS produce by the fungi in absence of starch salvation in D2O

3.5.5. Conclusions

These analyses were not conclusive, however, it can be inferred that the

modified starch chains bear functional groups able to be solvated and

detected by NMR spectroscopy. A rough approximation may be related with

the presence of carboxylic groups. The presence of double bonds and

specially of CO double bonds was corroborated by FTIR and FT-Raman.

3.6. Solid state NMR

3.6.1. Abstract

Results obtained from solid state NMR, FT-Raman, and chromatography

suggest that the functional properties of modified starches can be attributed

either to the presence of extracellular exo-polysaccharides or intermolecular

bonds occurring at the C6 in the glucopyranose ring.

63

3.6.2. Introduction

The use of 1H NMR has been restricted to the detection of functional or

solvated groups appearing in polymers of high molecular weight. In general,

most advances in the interpretation of the molecular structure of starch and

its derivatives have been made with solid state NMR. Cross polarization

magic angle spinning (CP/MAS) 13C-NMR posses broad lines, which have been

associated to the different positions in the glucose units, located at ~63ppm

(C6; CH2-OH); 72 ppm (C2, 3, 5; CH-OH); a smaller line in 84 ppm (C4; CH-

O); and 103 ppm (C1 anomeric; C-O-C) [98, 99].

CP/MAS 13C NMR is sensitive to substances at low humidity content (rigid

state with low molecular mobility). This technique has been used to obtain

structural information of granular starches. The spectrum shows the position

of each one of the resonances at the respective carbons located in the

glucopyranose ring. At low humidity content the CP/MAS 13C NMR spectra of

granular starch exhibits a triplet in the signal assigned to carbon C1 in cereal

starches and a doublet in the case of tubers. Although the reasons for the

detection of these multiplicities have not been explained in detail, they have

been indeed associated with the presence of crystalline regions in granular

starches, and in specific with the 9 nm periodicity of the amylopectin

branches.

3.6.3. Materials and methods

Solid state 13C CP/MAS (cross polarization magic angle spinning) spectra

were collected in a Bruker advance DSX 200 MH at room temperature by

using a standard Bruker wide-band MAS probe. Resonance frequency of 200

13 mHz. 4 mm bore superconducting magnet. Dry samples were packed in 4

mm zirconia rotors, with sealed Kel-Fe caps and spun at 5 kHz. 13C CP

MAS/NMR spectra with 3 ms CP contact time, 5 s recycle delay. The free

induction decay was subjected to standard Fourier transformation and

64

phasing. The chemical shifts were externally referenced to the solid

adamantane peak at 38.56 ppm. A total of 5000 scans were averaged for

each spectrum.

3.6.4. Results and discussion

The single chemical shift found between at ~60 ppm corresponds to C6. The

wide distribution of the chemical shift centered at ~72 ppm is attributed to

the hexapyranose ring carbons (C-2, 3, 4, 5). The peak at 80-84 ppm is

related to amorphous fractions of starch associated to the C4. The chemical

shifts observed as a triplet is associated to crystalline regions of starch (A-

type), and the other resonances seen as shoulders between the interval from

90 from 108 ppm are attributed to the presence of amorphous regions of C1.

The lamellae responsible for the crystalline regions, are formed by three

discrete components the backbone which support the double helices, parallel

‘rigid’ double-helical (menogenic) units (~5-6 nm) and amorphous regions

(more flexible un-branched regions, also called ‘spacers’ or side chains) with

sizes of ~4 nm. The size of the crystalline lamellae is ~9 nm. It has been

observed by Waigh et al. [49] that a simultaneous appearance and

disappearance of the 9 nm and 1.6 nm reflections in small angle X-ray

scattering under hydration and dehydration experiments. The hydration

produces the 9 nm reflection due to the smectic periodicity. With a ~10% mc

solid state NMR spectra show a set of sharp and strong peaks at ~100 ppm

(triplet in A-type, and doublet in type B starches) associated with the

crystalline regions. Neither dehydrated native granules (<5% mc) nor the

amorphous dehydrated starches show these signals. The same phenomenon

occurs for highly hydrated starches (~20% mc; moisture content <). Under

dry conditions the starches may be in a pure glassy form (<5% mc), while

moisture contents of ~10 % allows the formation of crystalline regions.

These particular structures of intermediate order are known as liquid crystals

SCLCP (side chain liquid crystal polymer). The degree of mobility of these

three components, coupled with the helix-coil transition, may be used to

65

explain physicochemical and structural properties of starch such as

gelatinization, dehydration or molecular composition. The signals seen as a

shoulder of C-2, 3, 5 at ~76 ppm can be related to amorphous domains of

amylose residues.

It can be observed from these data that the C4 resonance shifted to a lower

field and the resonance of C6 moved to a higher energy field due to

molecular shielding. This phenomenon was observed even in the early stages

of the modification. By measuring the area of the chemical shifts by the

MestReC 4.8.6.0 for Win XPM software, the main differences were found at C

2, 3, 5 and C6 which showed dramatic reductions and increases of the

respective areas suggesting the reduction of the glucopyranose rings and the

intermolecular bonding via C6, phenomena which may be related to the

reduction of the molecular motion of the starch chains.

The solid state CP/MAS 13C NMR results are shown in Table 7. Table 8

shows the associated area of the chemical shifts before and after the

modification for tapioca starch. It can be observed from these data that the

C4 resonance shifted to a lower field and the resonance of C6 moved to a

higher energy field due to molecular shielding. This phenomenon was

observed even in early stages of the modification. By measuring the area of

the chemical shifts by the MestReC 4.8.6.0 for Win XPM software, the main

differences were found at C 2, 3, 5 and C6 which showed dramatic reductions

and increase of the respective areas suggesting the reduction of the

glucopyranose rings and the intermolecular bonding via C6, phenomenon

which may be related with the reduction of the molecular motion of the

starch chains.

66

Table 6

Solid state NMR. Chemical shifts for the different carbon of native and modified starches (three different sources produced the same results;

corn, tapioca, or potato)

Starch sample C1 C4 C3,2,5 C6

NS 101.60 80.84 72.58 62.35

MS-7 103.47

82.38

72.50 60.54

MS-3 103.36 82.18 73.30 60.36

NT=native starch, MS-3=starch at the 3rd day of modification, MS-7=starch at the 7th day of modification

Table 5

SOLID STATE NMR. AVERAGED ASSOCIATED AREA BY CARBON TYPE (THREE DIFFERENT SOURCES PRODUCED THE SAME RESULTS; CORN, TAPIOCA, OR POTATO)

Carbon type

Associated area (%) Native Modified Starch Starch

C1 16 17

C4 7 7

C2,3,4,5 70 55

C6 7 21

67

3.6.5. Conclusions

It was observed from these data that the C4 resonance shifted to a lower

field and the resonance of C6 moved to a higher energy field due to

molecular shielding. This phenomenon was observed even in the early stages

of the modification. By measuring the area of the chemical shifts by the

MestReC 4.8.6.0 for Win XPM software, the main differences were found at C

2, 3, 5 and C6 which showed dramatic reductions and increase of the

respective areas suggesting the reduction of the glucopyranose rings and the

intermolecular bonding via C6, phenomenon which may be related with the

reduction of the molecular motion of the starch chains.

3.7. MALDI-TOF MS

3.7.1. Abstract

High molecular weight derived polymers have been analyzed by SEC [43].

However, this technique does not provide details related to chemical

composition of the polysaccharides. A good approximation is the use of

matrix-assisted laser desorption ionization mass spectrometry (MALDI-TOF).

A good separation of relatively low molecular weight components was

achieved with this technique. Based on the spectrometric separation it is

possible to conclude that the starch molecules were affected in the molecular

weight and therefore in their molecular composition.

3.7.2. Introduction

Matrix-assisted laser desorption/ionization time-of-flight mass

spectrometry, MALDI-TOF MS, was used by Broberg et al. [101] to study the

chain length distribution of amylopectin. The results were comparable with

those obtained by high-performance anion-exchange chromatography with

pulsed amperometric detection (HPAEC-PAD). MALDI-TOF MS was, however,

68

reported to be a more sensitive technique and provided more detailed

information on the molecular mass of the unit chains [102].

3.7.3. Materials and methods

3.7.3.1. Sample preparation

Matrices for the separation of these polymers were prepared with 1, 8, 9-

anthracenetriol (or dithranol), 2, 5-dihydroxybenzoic acid, trans-3-

indoleacrylic acid, and 2-(4-hydroxyphenylazo) benzoic acid. Also different

alkali metal salts (LiCl, NaCl, KCl) or silver salts such as silver trifluoroacetate

(AgTFA) were used to form matrix-cationization agent mixtures. Two sample

preparation methods were used: by spotting and thin layer.

3.7.3.2. Instrumental conditions

An Applied Biosystems/MDS SCIEX 4800 MALDI TOF/TOF analyzer was used

in this study. To determine the mass spectra of the different components of

the starch matrices (native and modified), the instrument was set as MS

linear low mass positive range of 500 - 5000 dalton (low MS linear mode with

150 cm ion path length). Mass spectra were averaged over 400 shots using

an Nd:YAG 200-Hz laser at a wavelength of 355 nm. The laser firing rate was

200 Hz. After laser strikes the sample, the sample stage and sample plate

were supplied with an acceleration voltage (0 to 25 KV) at a predetermined

delay time.

3.7.4. Results and discussion

The matrix combination and sample preparation showed a good separation in

both materials (native and modified starches) were 2, 5-dihydrobenzoic-KCl

and the thin layer method [101]. Although no higher resolution was

69

achieved, the spectra were able to show the main differences between the

two samples.

In the mass spectrum of native starches, the unit chains appeared duplicated

with a difference of 18 m/z which may correspond to the loss of a water

molecule. In addition, it can be observed that as the m/z increased the peak

area decreased, but the m/z difference between pairs of peaks remained in

160 m/z (Fig. 25).

On the other hand, the difference between sets of peaks in fermented

starches was 200 m/z and the difference the pair of peaks was ~50 m/z (Fig.

26). The spectral differences also showed in this spectrum peaks (1036,

1240 m/z, etc.) possibly related to the production of EPSs or starch-derived

products. The peak at 1240 m/z, in Fig. 26 for example, may correspond to

a DP ~7. Also, the peak shown in Fig. 8, section 3.1.2.1 is probably a non-

related starch-derived product with a probable DP of 7.

800.0 1041.8 1283.6 1525.4 1767.2 2009.0

Mass (m/z)

20

40

60

80

100

20

40

60

80

100

Re

lativ

e ab

und

ance

Re

lativ

e ab

und

ance 996

10161158

117613391321

9961016

11581176

13391321

800.0 1041.8 1283.6 1525.4 1767.2 2009.0

Mass (m/z)

Figure 25-MALDI-TOF MS spectrum of native starch-

70

849 1181.2 1513.4 1845.6 2177.8 2510

Mass (m/z)

20

40

60

80

100

Rel

ativ

e ab

unda

nce

Set of peaks1131

1189

1342

13921036

1240

849 1181.2 1513.4 1845.6 2177.8 2510

Mass (m/z)

20

40

60

80

100

Rel

ativ

e ab

unda

nce

Set of peaks1131

1189

1342

13921036

1240

Figure 26- MALDI-TOF MS spectrum of modified starches

3.7.5. Conclusions

The molecular changes of native starch after its fermentation with fungal

isolates of the genus Ophiostoma were studied with MALDI-TOF. Neighboring

sets of two peaks were detected in native starches. The spectrum of modified

starches also showed sets peaks in pairs, but with important spectral

differences.

The molecular weight variation between sets of adjacent pairs of peaks found

in native starches was of 160 m/z. The difference between the peaks within

the set of two peaks was 18 m/z which fit with the theoretical production of a

water molecule. On the other hand, the molecular differences between

neighboring oligo-saccharides in fermented starches were approximately 200

m/z and the difference of the peaks forming the duplicate was ~50 m/z

which does not fit with the molecular structure of the respective starch’s

glycans. In general, the areas depicted under the peaks of modified starches

were broader, phenomenon which may be related to the process of

separation of the different polysaccharides. Moreover, important similarities

were found by comparing the results obtained from HPAEC-PAD with those of

MALDI-TOF.

71

3.8. HPAEC-PAD

3.8.1. Abstract

High molecular weight derived polymers have been analyzed by SEC[43].

However, this technique does not provide details related to chemical

composition of the polysaccharides. A good approximation is the use of high

performance anion exchange chromatography with pulsed amperometric

detection (HPAEC-PAD). A good separation of relatively low molecular weight

components was achieved with this technique. Based on the chromatographic

separation it is possible to infer the presence of two dissimilar

polysaccharides forming the matrix of modified starches.

3.8.2. Introduction

High-performance anion exchange chromatography with pulsed

amperometric detection (HPAEC-PAD) and matrix-assisted laser

desorption/ionization time-of-flight mass spectrometry (MALDI-TOF MS) are

well established methods to determine the chain length distribution of

amylopectin and short chains of amylose and related products [101, 103-

106].

Sugars or different related compounds (methylated aldoses, deoxysugars,

amino sugars, N-acetylated amino sugars, acidic sugars, etc.) are separated

from the matrix by following an acidity trend of the OH groups: 1-OH>2-

OH>6-OH>3-OH>4-OH. For example, the substitution of the anomeric

group, as in the case of 1-O-methylated glucose, produces a poor retention

time in the column, while all other derivatives exhibit higher retention times.

By the other hand, oligo- and polysaccharides are separated based on their

size, chemical composition, and linkage type [106, 107]. Therefore, sugars

with close molecular weight can be separated with great sensitivity.

72

Structural studies report the average degree of polymerization (DP) of the

branches of the amylopectin molecules in waxy rice of ~18-19 nm and of

30.7 nm for high-amylose maize VII. In general, XRD B-type starches

present longer branches than A-type starches [108]. Starches with short

averaged amylopectin branch chain lengths show a low gelatinization

temperature (corn), also the phosphate (potato) groups may induce a faster

gelatinization [109]. A-type starches (corn) present a higher proportion of

crystalline region than B-type starches (potato). For tapioca starch Wong and

Jane [110] reported three DP distributions of debranched amylopectins with

DPs of 48, 19, 12. Sanderson et al. [108] has reported percentages of the

DPs’ in the ranges of 6-15, 16-24, and 25-60 for tapioca, corn, and potato of

30-34-36, 30-38-32, and 25-37-48 respectively.

3.8.3. Materials and methods

3.8.3.1. Polymer production and sample preparation

The freeze-dried polysaccharides (see section 2) were resuspended in double

distilled water (0.1 g X 20 mL) and centrifuged at 16000 X g for 20 min to

remove coarse particles and further analyzed in the HPAEC-PAD.

Total hydrolysis for sugar composition of modified starches as well as

polysaccharides produced by the fungi in absence of starch was carried out

with 2.0 M sulfuric acid at 100oC for 3h, and neutralized with 1.0 M NaOH

before analyses.

In order to determine the influence of growth pH on the starch mass

production experiments was carried out in triplicate were run at different pH

values (4, 5, 6, 7 and 7.6). The introduced variables for this experiment

were: time of reaction, 72 h; nitrogen source, yeast extract (2 g per liter);

temperature, 20oC; and spore concentration, 0.71 g per liter. The growth pH

was controlled by buffer solutions consisting in X ml 0.1 M of citric acid and Y

73

ml 0.2 M Na2HPO. The influence of the cultivation temperature on the

production of modified starch polymers was investigated by controlling the

growth temperatures at 15, 20, 25, 30 and 35oC. For temperature

experiments the settings were: pH 4; 72 h; yeas extract, 2 g/l; spore

concentration, 0.71 g/l; at 72 h. The effect of spore concentration was

measured by performing the following experiment: spore concentrations,

0.71, 1.3, 1.8, 2.23, and 3.47 g/l; temperature, 20oC; YE 2 g/l; pH 4; and

time of reaction 72 h. The influence of various nitrogen sources over the

starch mass change was assessed by using various YE, Urea, NaNO2, NaNO3,

HPO4(NH4)2 and NH4NO3. The starch mass load for all experiments was 20 g

per liter of culture media

3.8.3.2. Instrumental conditions

HPAEC-PAD was performed on a DIONEX DX-500 (Sunnyvale, CA, USA)

equipped with ED-40 electrochemical detector, a GP-50 gradient pump and

an AS-3500 autosampler. Samples (~1.0 mg/mL deionised water) were place

in a rotary shaker for 24 h at 160 rpm., then centrifuged (16000 X g for 20

min) to remove coarse particles, dialyzed against distilled water for 72 h at

4°C (tubing Fisher brand®, wall thickness 30mm, dry cylinder diameter 25.5

mm for molecular weight separation between 6,000 and 8, 000 g/mol,

membrane with a retention capacity of 15, 000 to 20, 000 Mw) and then

injected (25mL) onto CarboPac anion-exchange columns PA1 and PA200 with

respective guard columns. A triple-potential waveform was applied using the

following settings: E1_0.01 V (volts), t1_0-480 ms; E2__0.60 V, t2_481-780

ms; E3_-0.6 V, t3_781-1020 ms. The flow rate was 1.0 mL/min. Reagents

were prepared with sodium hydroxide solution (50% w/w, low carbonate,

analytical grade) and sodium acetate (CH3COONa FW/PM 82.03). Eluents

were: (A) 100 mM sodium hydroxide; (B) 100mM sodium hydroxide, 500 mM

sodium acetate. Eluentes were prepared with degassed water, sonicated

before use and determined with helium gas during the testing. The system

was equilibrated with the eluent A for 10 min before each run. The separation

74

was performed by using a gradient elution, starting with the reactive “A”

during the 10 first min to separate lower molecular weight components and

10% reactive “B” incorporated at 10 min to separate higher molecular weight

fractions (Dionex Corporation, Application Note 67).

3.8.4. Results and discussion

3.8.4.1. Oligo- and polysaccharide composition

In the present study, HPAEC-PAD was used to study the polysaccharide

composition of starches modified with fungal isolates of the genus

Ophiostoma. The chain molecular weight of these materials has been studied

before by SEC (size exclusion chromatography) and it was reported by Huang

et al. [43] the overall increase in the molecular weight after the fungal

modification of starch. However, this technique does not provide specific

information related to the process by which the molecular weight changes

occur. Although HPEAC-PAD has a low capacity to resolve high molecular

weight components, it is able to provide fine details related to the molecular

composition of the polysaccharides. In this technique the detector response

per molecule or OH group are dependant on the DP (degree of

polymerization) of the polysaccharides under study, and therefore it may

provide fine information related to the molecular composition of the modified

starches.

Two features of major importance for the interpretation of the results were

considered: the use of standards and the consistency in the separation and

retention time of carbohydrates [107]. The anion-exchange elution profiles of

various starches are shown in Figs. 27 through 36 (modified tapioca, potato,

corn, PDB, EPSs, and native starch). All profiles were similar. However,

polysaccharides produced by PDB were slightly different in the

chromatographic profile as well as the retention times. The process of

75

production of EPSs by the fungi in the absence of starch has been reported

before [42].

The relative degree of polymerization (DP) of oligosaccharides separated in

samples of modified starches was determined by using glucose, maltotriose,

maltopenatose and maltohepatose. The average retention times of the

standards D-glucose, maltotriose, maltopentaose, and maltoheptaose were

1.9, 6.5, 10.7, and 14.8±0.01 min respectively (Fig. 37). It was observed

that there was a linear relationship between the retention time and the DP.

Oligosaccharides derived from native starches showed consistency with the

standards as it can be seen from peaks 5 through 10 (Fig. 63). Initial peaks

(shown in letters, Glc=glucose) may be due to the presence of minor

components such as low molecular weight lipids, proteins, or phosphates

[56], which were not detected after the modification of the starch.

The high sensitivity of this technique allowed the separation of the matrix of

modified starches as shown in the chromatograms. The peaks at the near 2,

6, 11, and 16 min can be associated with oligosaccharides with DPs of 3, 5,

and 7 (Fig. 27). Clearly shown, there are various peaks with random

molecular weights distributed within the chromatogram. The peak at the near

3 min showed a higher intensity compared with the one associated to glucose

units, which were associated with the peak number 1 at ~2.0 min.

The peak with an apparent DP of 4 in modified corn starches (Fig. 29) was

further split into two peaks (Fig. 30). These results show that there is a

strong molecular attraction between two different molecular fractions. In

general the chromatographic profile strongly suggests the presence of

dissimilar polysaccharides in the matrix. In order to be separated, these

polymers must show a higher degree of ionization in the alkaline medium

and/or a better affinity for the stationary phase than the starch fraction. The

chromatogram from potato dextrose broth (PDB) also showed important

variations with respect to the regular sequence of starch derivatives. It is

76

interesting to note that the peaks at the retention times 3, 12, 16, and 20

min detected in modified PDB are consistent with the peaks detected in Fig.

53.

Fig. 32 shows the chromatogram of EPSs produced by the fungi in absence of

starch. The peak at ~10 min was found in samples of modified starches

produced in excess of spores and in samples produced in a culture media

supplied with Na+NO2-, Na+NO3

-, HPO4-2(NH4)2, or NH4

+NO3-, which in general,

inhibited the process of oligosaccharide production and the chromatogram

acquired a profile similar to that found in EPSs suggesting the increase in the

production of this polymer.

0.0200

0.0400

0.0594

1.0 4.0 8.0 12.0 16.0 20.0 24.0 28.2

µC

min

1

2

3

45

6 7

8 9

100.0200

0.0400

0.0594

1.0 4.0 8.0 12.0 16.0 20.0 24.0 28.2

µC

min

1

2

3

45

6 7

8 9

10

Figure 27- Chromatographic profiles of modified starches synthesized from tapioca starch on the 3rd day of modification (CarboPac PA1)

77

0.0050

0.0100

0.0150

0.0200

0.0250

0.0326

0.8 4.0 8.0 12.0 16.0 20.0 22.4

µC min

12

34

5

67 8

0.0050

0.0100

0.0150

0.0200

0.0250

0.0326

0.8 4.0 8.0 12.0 16.0 20.0 22.4

µC min

12

34

5

67 8

Figure 28-Chromatographic profiles of modified starches synthesized from potato starch on the 3rd day of modification (CarboPac PA1)

0.0100

0.0300

0.0500

0.0645

0.6 4.0 8.0 12.0 16.0 20.0 24.0 27.7

µC

min

1

2

3

4

56

7

0.0100

0.0300

0.0500

0.0645

0.6 4.0 8.0 12.0 16.0 20.0 24.0 27.7

µC

min

1

2

3

4

56

7

Figure 29- Chromatographic profiles of modified starches synthesized form corn starch on the 3rd day of modification (CarboPac PA1)

0.0025

0.0050

0.0100

0.0140

0.0 2.5 7.5 12.5 17.5 22.3

µC

min

0.0025

0.0050

0.0100

0.0140

0.0 2.5 7.5 12.5 17.5 22.3

µC

min

Figure 30- Chromatogram profile of modified starches- detail of peak separation performed with a CarboPac PA200 column (peak separation

corresponding to peak no. 4 in Fig. 55)

78

0.0200

0.0400

0.0626

1.0 4.0 8.0 12.0 16.0 20.0 24.0 27.3

µC min

0.0200

0.0400

0.0626

1.0 4.0 8.0 12.0 16.0 20.0 24.0 27.3

µC min

Figure 31-- Chromatographic profiles of modified starches synthesized from corn starch after the 3rd day of modification (CarboPac PA1)

0.050

0.150

0.250

0.344

0.8 4.0 8.0 12.0 16.0 20.0 24.0 28.0

µC

min

7

0.050

0.150

0.250

0.344

0.8 4.0 8.0 12.0 16.0 20.0 24.0 28.0

µC

min

7

Figure 32- Chromatographic profile of modified starch synthesized from PDB (CarboPac PA1)

0.0010

0.0020

0.0035

0.0 5.0 15.0 25.0 35.0 45.0 60.1

µC

min

0.0010

0.0020

0.0035

0.0 5.0 15.0 25.0 35.0 45.0 60.1

µC

min

Figure 33-Exo-polysaccharides (EPSs) produced by the fungi in yeast extract (CarboPac PA1) –no substrate involved

79

0.020

0.060

0.100

0.140

0.0 10.0 20.0 30.0 40.0 50.0 60.1

µC

min

0.020

0.060

0.100

0.140

0.0 10.0 20.0 30.0 40.0 50.0 60.1

µC

min

Figure 34-- Chromatographic profiles obtained for fermented starches (corn, tapioca, or potato) with increase in the spore concentration

0.002

0.004

0.007

0.0 10.0 20.0 30.0 40.0 50.6

µC

min

0.002

0.004

0.007

0.0 10.0 20.0 30.0 40.0 50.6

µC

min

Figure 35-- Chromatographic profiles modified starch (from tapioca, potato, or corn) produced in Na+NO2

-, Na+NO3-, HPO4

-2(NH4)2, or NH4+NO3

-. The effect of the different nitrogen sources was similar.

0.0100

0.0200

0.0300

0.0450

0.0 5.0 10.0 15.0 20.0 25.0 30.1

µC

minA - 1.3

B - 1.5Glc -2.2

D - 3.8

E - 4.2

F - 6.2

G - 7.7

5 - 8.66 - 11.3

7 - 14.1

8 - 16.7

9 - 19.1

10 - 21.40.0100

0.0200

0.0300

0.0450

0.0 5.0 10.0 15.0 20.0 25.0 30.1

µC

minA - 1.3

B - 1.5Glc -2.2

D - 3.8

E - 4.2

F - 6.2

G - 7.7

5 - 8.66 - 11.3

7 - 14.1

8 - 16.7

9 - 19.1

10 - 21.4

Figure 36-Chromatogram of one of the substrates (the example native starch (Carbo Pac PA1)

80

0

2

4

6

8

10

12

14

16

Glucose Maltotriose Maltopentaose Maltoheptaose

Ret

enti

on

tim

e

Figure 37- Chart showing the retention time of the various used standards

3.8.4.2. Sugar composition

Exopolysaccharides (EPSs) produced by the fungi in the absence of starch

were hydrolyzed for sugar identification as well as modified tapioca or PDB

starches. The process of production of EPSs by the fungi in absence of starch

has been reported before [42]. The chromatograms for various hydrolyzed

samples of modified starches (potato, tapioca, corn, potato from PDB, and

amylopectin), exopolysaccharides (EPSs) produced by the fungi in absence of

starch, and native starch are shown in Figs. 38-44.

In general, it was found that polysaccharides produced by the fungi in the

absence of starch are formed by two different basic units, while modified PDB

and tapioca showed 4 peaks. Modified amylopectin, potato, and corn starches

showed two peaks. Moreover, peaks appearing after 2 min in PDB and

tapioca starches had similar retention times to those found in hydrolyzed

EPSs. The retention times of the peaks found in corn, potato, and

amylopectin were lower than those found in EPSs. The standard at the same

conditions appeared at ~2.5 min. However, when the various samples were

added to the samples of modified starches the standard peak of glucose

81

shifted to ~2 min. Therefore, in all samples of modified starches the peak at

~2 min can be associated with glucose.

0.0200

0.0400

0.0600

0.0900

0.0 2.5 5.0 7.5

µC

1 - 1.433

4 - 2.683

0.0200

0.0400

0.0600

0.0900

0.0 2.5 5.0 7.5

µC

1 - 1.433

4 - 2.683

Figure 38-Chromatogram showing the separation of hydrolyzed mod. starch from tapioca starch. Separation by CarboPac PA1

0.025

0.050

0.075

0.100

0.140

0.0 2.5 5.0 7.5

µC

1 - 1.517

0.025

0.050

0.075

0.100

0.140

0.0 2.5 5.0 7.5

µC

1 - 1.517

Figure 39-Chromatogram showing the sugar separation of hydrolyzed modified starch from potato starch. Separation by CarboPac PA1

82

0.0100

0.0200

0.0300

0.0400

0.0 2.0 4.0 6.0 8.0

µC

1 - 1.517

0.0100

0.0200

0.0300

0.0400

0.0 2.0 4.0 6.0 8.0

µC

1 - 1.517

Figure 40- Chromatogram showing the sugar separation of hydrolyzed modified corn starch. Separation by CarboPac PA1

0.050

0.100

0.150

0.180

0.0 2.5 5.0 7.5

µC 1 - 2.250

0.050

0.100

0.150

0.180

0.0 2.5 5.0 7.5

µC 1 - 2.250

Figure 41-Chromatogram of hydrolyzed fungal exo-polysaccharides (EPSs) produced in absence of substrates. Separation by CarboPac PA1

0.0100

0.0200

0.0300

0.0400

0.0500

0.0 5.0

µC

2 - 2.267

0.0100

0.0200

0.0300

0.0400

0.0500

0.0 5.0

µC

2 - 2.267

Figure 42--Chromatogram of hydrolyzed mod. starch from amylopectin. Separation by CarboPac PA1

83

0.050

0.100

0.150

0.200

0.0 2.5 5.0 7.5

µC

0.050

0.100

0.150

0.200

0.0 2.5 5.0 7.5

µC

Figure 43-Chromatogram of hydrolyzed modified starches from PDB. Separation by CarboPac PA1

0.0200

0.0400

0.0600

0.0900

0.0 2.0 4.0 6.0 8.0

µC

min

0.0200

0.0400

0.0600

0.0900

0.0 2.0 4.0 6.0 8.0

µC

min

Figure 44-Chromatogram of one of the standards -D-Glucose. Separation by CarboPac PA1

3.8.5. Conclusions

HPAEC-PAD allowed a clear separation of dissimilar polysaccharides by

chemical composition. The polysaccharides produced during the fermentation

of starch showed to be strongly attached to the starch-like fractions. Such

components showed a better attraction towards the column compared with

the starch-like fraction and were, therefore, separated under the alkaline

conditions. The analysis of sugars showed that modified starches are

composed of at least two basic units, one of which was related to D-glucose.

84

It has been reported before by Huang et al. [43] the increases in the

molecular weight of these polymers by SEC (size exclusion chromatography).

However, SEC does not provide details related to the process of modification.

This technique together with MALDI-TOF MS allowed a fine separation of the

polymers involved. However, the separation is performed just at the low

molecular weight range with degrees of polymerization (DP) of ~10 to 40

glucose units.

3.9. Viscoelastic and mechanical properties

3.9.1. Abstract

Three different samples of thermoplastic fungal/modified starches were

prepared. One sample was produced by using the casting method with 40-

wt% glycerol and in excess of water as the plasticizers. The other two

samples were produced by extrusion. One set was produced solely in

presence of 40wt% of glycerol, and the other set was filled with 30 wt% of

Hallocote® 466 hydrasperse as reinforcing material. Similar samples of

native starches were used as control. The materials were characterized using

dynamic mechanical thermal analysis (DMTA).

In comparison with the native starch/glycerol composites, modified starches

exhibited a considerable increase in the storage modulus showing a process

of chain stiffening and the presence of multiple thermal transitions detected

by tan and loss modulus peaks. It was also observed that the presence of

clay produced a separation of the thermal transition in native starch/glycerol

composites, but not in samples prepared with modified starches which also

showed the shifting of the glass transition (Tg) at higher temperatures

suggesting a better thermal stability.

85

3.9.2. Introduction

3.9.2.1. Dynamic mechanical thermal analysis of polymers

Polymeric materials, synthetic or natural, are used extensively because of

their properties and low cost. For most applications, the most important

information when working with polymers (and with any other material) is to

have some basic knowledge of the mechanical behavior and how the

mechanical properties can vary with temperature and time of load (rheology,

creeping). In general, the mechanical properties are considered the most

important of all physical and chemical properties of polymers. The dynamic

mechanical properties provide information of one of the most fundamentals

properties of polymers, the glass transition temperature (Tg). The Tg can be

obtained by storage modulus (E') onset, loss modulus peak (E"), or tan

peak.

3.9.2.2. Basic definitions

Continuum mechanics of solids can be studied in solids through viscoelastic

properties (dynamic mechanical analysis) and the flow properties of liquids

by their Newtonian or non-Newtonian behavior (rheology) [111]. In both

cases, the materials are subjected to a deformation by an external force. The

viscoelastic behavior of different solid materials such as synthetic polymers,

wood and its derived products, and thermoplastic starches can be described

based on systems of springs and dashpots (a damper factor), which

represent the elastic and viscous or non-elastic behaviors respectively. Some

well known models to describe the visoelastic properties are the Maxwell

(consisting of a dashpot and a spring in series), Voigt (model which is a

dashpot and a spring in parallel), and the Four-element formed by a

combination of the Maxwell and Voigt models [111].

86

The slope of any graph of stress vs. strain within the elastic region yields a

straight line, and it corresponds to the Young’s modulus (E) of the material

which physically represents the stiffness of the materials. More complex

models exist to describe the mechanical properties of materials with more

complex deformation behaviors. In the Voigt model alone, for example, it is

difficult to separate the effects of the elastic behavior (E) from those

produced by the viscous or inelastic behavior (h) represented by the dashpot

(the damping factor).

A typical engineered stress () vs. strain () curve obtained at constant

temperature during a tensile test depicts the elastic properties of the solid

sample by the linear region [112]. The strain is usually calculated by using

the Cauchy engineering strain (eq. 1), but the Henchy or true strain, Kinetic

theory of rubber strain, Kirchhoff strain, or Murnaghan strain can also be

used for more accuracy. In the Cauchy test, for example, the strain is

determined by dividing the increment of the elongation (L) by the original

length of the sample (L) (L/L) and multiplied by 100 to express the result

as a percentage. The slope taken in the elastic or Hookean region is the

modulus or the Young’s modulus (E) which is associated with the stiffness of

the material. Clearly, E is dependent on the speed of the applied stress and

temperature. Different stresses will cause very different strains in the

materials. Also, by increasing the temperature the materials will pass

through its glass transition (Tg), melting (Tm), or degradation points at which

the modulus will drop abruptly due to the molecular motion or degradation.

The general definition of modulus using the Cauchy engineering strain

definition is defined by eq. 1 [111]:

E=tensile stress/tensile strain= /=(F/Ao)/(L/Lo)=(FLo)/(AoL)-------- eq. 1

Where F is force, Lo the length of the sample, L the elongation of the sample

after applying the force, and Ao is the transversal area of the sample.

87

The Hookean theory is used to represent the elastic behavior in solid

materials with viscoelastic behavior. The Hook’s law linearly relates the

deformation or strain to the stress by a constant specific to the spring. As the

spring constant increase, the material becomes stiffer, and the slop of the

strain-stress curve increases showing the increment of the E. In other words,

E is equivalent to the slope of the linear region of the strain-stress curve, and

it is also the constant “k” in the Hook’s relation (Force= - k*displacement).

For a squared triangle taking the hypotenuse in the linear region, the tangent

of the angle is equivalent to E, which can be defined as the relation of strain

to stress. In practice, the increase in modulus in polymers may be to the

increase in density cross-linking, crystallinity, and with the molecular

orientation in the direction of the testing, as well as with the addition of

certain percentages of fillers. The modulus will decrease with an increase in

temperature and plasticizer content since these variables increase the

molecular motion and produce the slippage of the neighboring molecules

inducing lower stiffness. Also, stiffness of the sample will appear higher if the

rate of testing (the speed of the stress) increases, and will decrease if the

test speed decreases. Viscous properties are represented by the curved

region or non-linear region. This region represents the Newtonian behavior or

the material’s ability to flow (rheology properties).

The viscous fraction is represented by the dashpot in the aforementioned

models. In a Newtonian fluid, the plot of shear stress () vs. strain rate ()

results in a straight line. The strain-stress variables are directly related by

the viscosity (ŋ). Many oils and liquids are Newtonian fluids; their viscosities

do not change with increasing shear rates. However, many materials are not

Newtonian liquids, since their respective plots of shear stress vs. strain rate

are not linear. For example, polymers, food products, suspensions, and

slurries are dependent on viscosity — pseudoplastic fluids get thinner as

shear rates increase; dilatants fluids increase their viscosity as shear rates

increase; plastic fluids have a yield point with pseudoplastic behavior; and

thixotrophic and rheopectic fluids exhibit viscosity-time-dependant with

88

nonlinear behavior. At molecular level, the dashpot represents the resistance

of the chains to uncoiling, whereas the spring represents the thermal

vibration of chains segments that will tend to seek the lowest energy

arrangement.

At the glass transition (Tg) of the polymer where polymer changes from

glassy to rubbery, the chains gain enough mobility to slide by each other.

The free volume (Vf) determines the ability of the molecules to move below

the Tg, or to flow above this transition point. Below the Tg some phenomena

associated with the elastic region occur like creeping, but many times the low

motion limits the ability of the instruments to measure it. On the other hand,

the free volume theory explains most of the phenomena found in starch

including aging or annealing.

Many important and interesting phenomena occur at the Tg transition. As the

polymer approaches its Tg the molecular motion reaches its peak, and all

molecular phenomena become temperature dependent. By exceeding the Tg,

the molecular structure of amorphous polymers are highly dependent on

their molecular weights and chemical structure. Per example, the chain

length and branching pattern may limit or facilitate the degree of molecular

motion. Plasticizers are common materials used to modify the rheological and

therefore molecular behavior of the polymers.

In general, the most important function of the plasticizers in a polymer is to

lower the Tg by “diluting” the polymer. These effects also decrease the

recovery or the polymer to its original shape after releasing the force which

produces the deformation. A special case is found in cross-linked polymers

which show a very specific curve with a flat equilibrium region, because the

crosslink do not allow the polymers to flow.

As mentioned, a polymer exhibits both elastic (spring-like) and viscous

(dashpot-like) behavior, but polymers also present a time dependant

89

behavior or “memory”. The behavior of the polymers is also dependant of the

temperature (free volume theory). The free volume theory is of extreme

importance for explaining the dynamic mechanical thermal analyses (DMA)

principles since the molecular motion is studied as a function of the

temperature [111]. The main objective of this instrument is to determine the

Tg of the materials either by storage modulus (E') onset or peaks produced

by tan and loss modulus (E").

3.9.2.3. Basic principles of DMTA

In DMA an oscillatory (sinusoidal) strain is produced by a sinusoidal stress.

The force to produce the deformation must be enough to allow the recovery

of the material within its elastic region. In the DMA an elongation is specified

and the force is assigned by the instrument to produce that deformation. This

will be reproducible if the strain is kept within the viscoelastic region. The

limiting extremes in viscoelastic materials are the elastic or Hookean

behavior and viscous or Newtonian behavior. These limits apply to the DMA

analyses. The user of the DMA chooses a strain (deformation) within the

elastic region of the material and performs the test at increasing

temperatures. Since the temperature will produce the material’s internal

molecular motion the instrument adjust the force to maintain the fixed strain.

By passing the Tg the material will not recover completely within the elastic

region under the applied stress, and the instrument will record the molecular

motion as a loss of energy. The Tg will appear as a Gaussian or Lorentzian

curve depending mostly of the molecular weight distribution [111].

For a material, which obeys Hooke’s law, the resulting stress will be

proportional to the amplitude of the applied strain. The strain in this case will

be in phase with the stress, i.e., the phase shift (phase angle ) between

stress and strain will be 0°. For an ideal fluid, which obeys Newton’s law, the

stress will be proportional to the strain rate. The stress signal will lead the

90

strain signal by 90°. Viscoelastic materials exhibit a phase angle between 0°

and 90°.

The “modulus” generated during an oscillating experiment in the DMA is

referred as complex modulus (E* or G*)—which is defined based on the

complex stress (* or*) and the complex strain (*or*). The complex

modulus is a measure of the material’s resistance to deformation and it

encompasses both the viscous and the elastic properties. The elastic

modulus, or storage modulus (E' or G’) and the viscous modulus, or loss

modulus (E” or G”), are defined based on the complex stress and strain. The

tan is defined as the ratio of the loss to storage modulus (tan =E”/E' or

tan = G”/G’). A summary of the calculations are as described in eq. 2 to 8:

Complex Modulus defined for a viscoelastic material (nomenclature for

compression or bending clamps):

E* = */ or E* = E' + iE” ---------------------------------------- eq. 2

Complex Modulus (nomenclature for shear mode clamps):

G* = or G* = G'+ iG” ------------------------------------------- eq. 3

Storage Modulus (tension, Compression or Bending):

E’ = ’or E' = E* cos ------------------------------------------------ eq. 4

Storage Modulus (shear):

G’ = ’ or G’ = G* cos ---------------------------------------------- eq. 5

Loss Modulus (tension, Compression or Bending):

91

E” = ” or E” = E* sin -------------------------------------------- eq. 6

Loss Modulus (shear):

G” = ” or G” = G* sin ----------------------------------------------- eq. 7

Damping produced by the viscous region of the material:

tan =E"/E’ or tan = G”/G’ --------------------------------------------- eq. 8

When the oscillatory force in the DMA is applied to a visco-elastic material

the signal of the strain undergoes de-phasing (deformation in degrees

defined by 90°>°). By measuring both the amplitude of the deformation

at the peak of the sine wave (the wave period; stiffness) and the lag

between the stress-strain sine waves, quantities like modulus, the viscosity,

and the damping can be calculated, and described based on the material’s

response to the oscillating force. From this, it is possible to calculate

properties like the tendency to flow (viscosity), from the phase lag and the

stiffness (modulus) from the sample recovery. These properties are

associated with the ability to lose energy as heat (damping) and the ability to

recover from deformation (elasticity). It is also possible to study the

relaxation of the polymer chains and the changes in the free volume of the

polymer, which describe the changes in the sample.

One advantage of the DMA is that it is possible to obtain the modulus each

time a sine wave is applied, allowing a sweep across a temperature or

frequency range. It is possible then to run an experiment at a fixed

frequency (per example 1Hz or 1cycle/second) and record the modulus every

second. This can be done while varying the temperature at some rate like

1°C to 10°C/min, recording the modulus as a function of temperature (over a

range of 200°C). Similarly, it is possible to scan a wide frequency range or

shear rate range of 0.001 to 200 Hz. Most instruments have allowable ranges

for amplitude, force, and stress of 0.5 to 10000, 0 to 18, and 0 to 1E6 MPa

92

respectively. The DMA is a very sensitive technique which allows the

measurement of transitions not apparent in other thermal methods such as

DSC. This sensitivity allows the DMA to detect the Tg of highly crosslinked

thermosets or thin coatings.

Dynamic Mechanical Thermal Analysis (DMTA) is the most sensitive

technique to determine the glass transition temperature (Tg) of polymers.

Moreover, the Tg is the most important parameter to link the mechanical with

molecular properties of the polymers.

3.9.2.4. DMTA of starch

By using the shear mode of the DMA, Xie et al. [113] determined the glass

transition temperature (Tg) of starch plasticized with 45% of water by tan

peak at 68.6 °C; with onset at approximately 61°C and offset at ~70°C. The

storage modulus (G') didn’t show a clear onset, and the loss modulus (G")

showed also a peak at 60-70°C. At temperatures of 60°C G' remained almost

constant, but decreased abruptly above this temperature. G" initially

increased slightly with increasing temperatures, but also dropped abruptly by

passing the peak. The complex viscosity (ŋ*) decreased stately with

temperature, showing also a sharp loss of viscosity after passing the 60°C.

The softening of starch particles during the gelatinization produced the

decrease of G' and G", the slight decrease in both moduli was attributed to

initial water diffusion within the starch granules. The authors also noticed

that ŋ*viscosity remained almost the same while approaching to the Tg, and

dropped abruptly after passing this point. They also report that the best

operating conditions of the instrument was at 1Hz, block thickness 3mm, and

heating rate of 2°C/min. The tan peak and the endothermic peak obtained

at a heating rate of 2°C/min in the DSC were reported to be coincident. By

using the cantilever mode in the DMA at 1Hz and heating rate of 1.5°C/min.

Similar findings were reported by Averous et al. [114].

93

It is widely known that the Tg by DSC is highly dependent on the water

content, and the same appears to be truth for DMA. By using sealed pans

withstanding up to 30 bar, Stepto [115] report the endothermic peak

associated to the Tg for starches hydrated at 12% and 42% to be 150 and

75C respectively.

Wilhelm et al. [85] also reported the production of starch-clay composites.

They analyzed these materials by dynamic mechanical analysis (DMA) and

other techniques, and reported two relaxation processes in un-plasticized

films and three in plasticized materials for the DMA results. One relaxation

process at approximately -110°C in unplasticized films was associated with

the rotation of the hydroxymethyl groups and oscillations of the sugar rings

about the glycosidic bonds, the other to water loss. The relaxation processes

found in plasticized films were associated with two phases that originated

from partial miscibility of glycerol and starch (-78°C), other to amylose-

glycerol rich phase, and the other to the starch-rich phase-glycerol. The

shifting in the Tg by tan peak was explained based on the formation of

hydrogen bonding among the different components of the composites, i.e.:

clay-water-starch. The antiplastization of sorbitol in starch composites was

also reported as the presence of an extra thermal peak in DMA by Gaudin et

al. [116].

3.9.3. Materials and methods

3.9.3.1. Formation of films by the casting method

The modification of starch was performed according to the methods of Jeng

et al. [42]. The starch samples (modified and native; 4 g/100g of water)

were diluted in a beaker of 400 mL, then 40% (total dry basis) of glycerol

was added. The solution was placed in a heater equipped with a stirrer. The

temperature used was 70°C at the heating/stirring time of approximately 1

h. The solution was then cooled and poured on a Petri dish. Water was

94

evaporated from the moulds in a ventilated oven at 40°C overnight. Dry films

were conditioned in open polyethylene bags and stored at 20°C and a RH

50% for one week before the analysis [43, 117].

3.9.3.2. Extrusion of the materials

Two samples were extruded: modified starch/glycerol and modified

starch/glycerol/clay. The amount of glycerol and clay was 40 and 30%

respectively base on starch dry weight. Similar samples were produced with

native starch and used as control. Glycerol was from Aldrich. HALLOCOTE®

466 hydrasperse clay from HallStarch Co. was from L.V. LOMAS LTEE.

The dry fractions were pre-mixed. Then the plasticizer was added and

blended in a commercial food processor. After it was sufficiently blended, the

mixture was placed in a plastic bucket covered with a hermetic lid and

allowed to stand overnight.

The extrusion was carried out in a laboratory co-rotating twin screw extruder

ONYX TEC-25/40 designed with 10 heating zones and 3 vents. The technical

parameters of this apparatus and conditions during the extrusion are listed

below:

1. screw nominal diameter 25 mm

2. twin-screw centre distance 21.2 mm

3. speed of revolution (Max) 500 r/min

4. L/D 40 (length to diameter ratio)

5. Output 2-15 kg/h (according to the materials and formulary system)

6. whole power 11.4 kw (kilo watts)

7. driving motor power 5.5 kw

8. main motor centre high 1050 mm

9. main motor outside size 2400X600X1650

10.main motor weight 700 kg

95

11.up to 600 rpm (starch sample: 120 rpm)

12.Feeder up to 25 rpm (starch sample: 8 rpm)

13.S feeder 15 rpm (not used)

14.Pressure Motor 0.02 Mpa

15.Energy imput 4.1 A (Ampers)

The blends were extruded by using the digital feeder settings incorporated

into the extruder. Feed rates were determined for each blend by weighing

the amount of blend transported from the hopper to the barrel over a fixed

time interval (this value was ~28 g/min). The vents were maintained open to

allow the release of any humidity evaporated during the extrusion. The

extrusion was carried out by using a 2 in X 3 mm exit slit die to obtain the

product in rods which were cooled and immediately pelletized. The

temperature in the barrel varied according to the samples, but it was on

average ~140oC. The profile of temperature during the extrusion is provided

in Table. 6.

Table 6 Extrusion temperature profiles for samples finally tested MS=Modified starch, NS=native starch

Zone Temp. profile (oC) 1 2 3 4 5 6 7 8 9 10 MS 130 130 135 140 140 140 145 150 155 160 NS 130 130 130 135 135 140 140 140 140 140

3.9.3.3. DMTA conditions

The dynamic mechanical thermal analysis was conducted using a Q800

Dynamic Analyzer (TA Instruments USA) with tensile clamps at a single

frequency-scanning mode of 1 Hz and heating rate of 2°C/min over a

temperature range of 40 to 150°C. A dynamic strain of 15 m was applied to

all samples, 0.4 for Poisson ratio, 125% force track, and 5 min of soak time.

The data collected consisted of storage modulus (E'), loss modulus (E"), and

loss factor (tan .

96

A pre-fabricated die of 10 X 20 mm was used to cut the films and keep the

dimensions constant. The thickness was fixed at 0.40 mm by using metallic

prefabricated stoppers in the case of hot press mold films. The thickness in

the case of films prepared by casting methods was specified by weight during

their preparation. All films were coated with silicone oil to prevent a fast

dehydration. The tests were performed repetitively until a reproducible

values were obtained (the tests were performed at least in triplicate). Data

storage and analyses were carried out with the incorporated software

Universal Analysis 2000 TA Instruments Version 4.5A and Advantage for Q

Series Version 2.5.0.256 TA Instruments.

The generation of the stress-strain curve for the determination of the linear

viscoelastic region (LVR) was determined by using a preload force of 0.1N, at

40°C, force ramp rate of 0.25N/min, and maximum force of 5N. During this

determination the instrument was set for the determination of the creep

compliance [J(t)] and relaxation modulus [E(t)]. The creep compliance was

determined with a static force of 0.01 N for 5 min.

3.9.4. Results and discussion

3.9.4.1. Samples produced by film casting method

(starch/glycerol composites)

3.9.4.1.1. Determination of the linear viscoelastic region

(LVR)

The stress vs. strain curve of the films determined using the DMA showed in

general similar results compared to those results obtained using the Instron.

The elastic region associated to the Young’s modulus is shown in Figs. 45 and

46. These plots were obtained at constant temperature (40oC), with a ramp

force of 0.25 N/min and maximum force of 10 N. The modulus of elasticity

(E) measured from the slop of the stress-strain curve of modified starch films

97

was of ~6.7 MPa, while the same value for native starch films was less than

1.3 MPa. Modified starch films showed an elastoplastic behavior within a

short strain range, and an abrupt transition region between the Hookean

behavior and permanent deformation. Native starch films behave more like a

non-linear elastic material (like rubber).

The stress-strain curve of films produced with native starches showed a non-

linear elastic or more viscous behavior. In this case, a tangent to the baseline

was used to calculate the Young’s modulus (E). The stress-strain curve of

those films produced with modified starches clearly showed a definite LVR

and a more defined elastic behavior of the material at low strain

deformations. It can also be observed an abrupt inflexion in the transition

region of permanent deformation can also be observed. In this case E was

directly associated with the slope determined at the elastic region of the

curve.

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0

0

2

4

6

8

Str

ess

(MP

a)

Strain (%)

Figure 45--Linear viscoelastic region (LVR) determined by DMA in films produced by casting method with modified starches

98

0 2 4 6 8 1

0

1

2

3

4

5

0

Str

ess/

MP

a

Strain/%

NSF

Non-linear elastic

Figure 46- Linear viscoelastic region (LVR) determined by DMA in films

produced by casting method with native starches

3.9.4.1.2. Creep compliance test

DMA can also provide useful information related to the mechanical behavior

of the materials. As detailed in the methods, two transient modes (creep and

stress relaxation) and a multi strain experiment (stress/strain plot) were

used for the determination of creep compliance [J(t)], stress relaxation

modulus [E(t)], and evaluation of the linear visco-elastic region (LVR)

respectively.

Creep compliance and recovery behavior are shown in Figs. 47 to 50. Creep

compliance was taken with a static force of 0.01N for 5 min at 40°C (Figs. 47

and 48). For the sample of native starch films two regions (primary and

secondary) can be observed for the deformation spectrum (insert of the Fig.

99

47). The primary region shown in this graph corresponds to the early stage

of loading when the creep rate decreases rapidly with time. The secondary

region of deformation also had a linear response with time and the creep rate

decreased more slowly. The recovery region had a similar trend.

Modified starches showed non-linear response as shown in Fig. 48. These

starches showed just one region under the applied load and a lag recovery

showing the molecular resistance to comply with the initial molecular

arrangement after the load was removed. Moreover, the deformation was

higher in samples of native starches at the same applied force, showing also

the tendency of modified starches to oppose the deformation.

On the other hand the maximum relaxation modulus in native starch films

was 3X104 MPa, but only of 1100 MPa in samples of modified starches. These

values are directly related with the stiffness of the materials which was lower

in native starch films. In most cases, the plot of the relaxation modulus vs.

time showed instantaneous equilibrium in native starch films after the

applied force ceased, but in the case of modified starches didn’t reached the

equilibrium, as shown in Figs. 49 and 50.

100

0 1 2 3 4 50

1x109

2x109

3x109

4x109

1.2 1.3 1.40

1x10 9

2x10 9

Prim ary

SecondaryE lastic recovery

D eform ation

Cre

ep c

ompl

ianc

e (µ

m²/

N)

Time (min)

Figure 47-Creep compliance determined during the LVR test in native starch films

0 1 2 3 4 5

0.0

2.0x108

4.0x108

6.0x108

1.5 2.0 2.50

1x108

2x108

3x108

4x108

Region of elastic recovery

Cre

ep c

ompl

ianc

e (µ

m²/

N)

Time (min)

Figure 48- Creep compliance determined during the LVR test in modified starch films

101

6.0 6.1 6.2 6.3 6.4 6.5 6.6

0

5x103

1x104

2x104

2x104

3x104

3x104

Str

ess

rala

xatio

n (M

Pa)

Time (min)

applied

Recovery zone

Figure 49-Stress relaxation curves for native starch films

4.8 5.0 5.2 5.4

0.0

2.0x102

4.0x102

6.0x102

8.0x102

1.0x103

1.2x103

Str

ess

rela

xatio

n (M

Pa)

Time (min)

Deformation

applied

Recoverable strain

Figure 50-Stress relaxation modulus in modified starch films

102

3.9.4.2. DMTA-Samples produced by film casting

(starch/glycerol composites)

Figures 51 and 52 depict the DMTA curve profiles of the films. In general

these spectra confirmed the presence of two separated fractions showing

different Tg transitions. The first drop in the curve from 130 to 150°C may be

due to water evaporation. The differences in the thermal transitions between

native and modified starches are clearly observed. Overall, the storage

modulus (E') and the loss modulus (E") decreased continuously with

increasing temperatures (up to 160°C), but composites prepared with

modified starches showed higher values of E' compared with native starch

films showing a process of chain stiffness.

The tan curve peaked at approximately at 90°C in native starch films.

However, in modified starch films two thermal transitions were observed at

~90 and 120°C by loss factor. By comparing both spectra, the peak at 90°C

in modified starch samples can be associated to the starch-rich fraction and

the peak at 120°C to thermal resistant fractions of modified starches.

103

20 40 60 80 100 120 140 160

0

20

40

60

80

100

120

TemperatureoC

E' a

nd E

'' (M

Pa)

E'

E''

Tan

0.24

0.28

0.32

0.36

0.40

Ta

n

NS

Figure 51- DMTA spectrum of native starch films produced by casting method

20 40 60 80 100 120 140 1600

50

100

150

200

250

Temperature oC

E'a

nd E

" (M

PaP

)

0.20

0.22

0.24

0.26

0.28

0.30

Tan

Tan

E'

E"

Figure 52-DMTA spectrum of modified starch films produced by casting

method

104

3.9.4.3. DMTA- samples produced by extrusion

(starch/glycerol composites)

Samples of extruded native and modified starches are shown in Figs. 53-55.

Similar to the previous analyses, native starches showed one thermal

transition by tan peak and modified starches two main thermal transitions

by tan peak.

Moreover, the loss modulus (E") curve profile of modified starch composites

showed a peak at ~60°C. The peak at 120°C was recurrent in both samples

of modified starches suggesting the presence of thermal resistant fractions

induced during the modification of the starch. The shoulder at 120°C in the

tan curve of modified starches and the irregular transition after passing the

maximum Tg not only confirmed the molecular heterogeneity of modified

starches, but also show that the Tg of the polymers can gradually shift to

higher temperatures showing higher thermal and molecular stability.

20 40 60 80 100 120

0

100

200

300

400

Temperature ( o C)

Sto

rage

(E

') an

d lo

ss (

E")

mod

ulus

0.28

0.32

0.36

0.40

0.44

0.48

Tan

Figure 53- DMTA curve profiles of native starch glycerol composites produced by extrusion

105

40 60 80 100 120 140 1600

200

400

600

800

1000

Temperature oC

Mod

ulus

(M

Pa)

E'

E"

Tan

0.20

0.25

0.30

0.35

0.40

0.45

Tan

= E

"/E'

Modified starch

Figure 54-DMTA curve profiles of modified starch glycerol composites produced by extrusion-

20 40 60 80 100 120 140 160

0.0

5.0x102

1.0x103

1.5x103

2.0x103

Temperature ( o C)

Sto

rag

e (

E')

an

d lo

ss (

E")

mo

du

lus

0.2

0.3

0.4

0.5

Tan

E"

E'

Tan

Figure 55-DMTA curve profiles of modified starch-glycerol composites produced after extrusion

106

3.9.4.4. DMTA- samples produced by extrusion

(starch/glycerol/clay composites)

Various mineral clays have been used to modify the chemical properties of

thermoplastic starches [85, 89, 118-129]. Figures 56 and 57 depict the

DMTA curve profiles of both samples under study: modified- and native-

starch/clay/glycerol composites. Overall, the same tendency was observed in

these experiments. The storage modulus (E') and the loss modulus (E")

decreased continuously with increasing temperatures (up to 160°C), but in

general, composites prepared with modified starches showed higher values of

E' compared with native starch composites showing a process of chain

stiffness. It was also observed that the incorporation of clay increased the

temperature of the glass transition temperature (Tg) by ~10°C and also the

storage modulus (E') by approximately three times (at the same conditions

of sample preparation and testing conditions).

The tan curve in native starch composites peaked at approximately 60 and

90°C. The peak at 90°C can be associated to a starch-rich fraction in both

samples. Since the peak at ~60°C was not detected in native starch-glycerol

composites, it can be assumed that this thermal relaxation is due to a

separated glycerol-rich phase originated from the partial miscibility of starch

with this plasticizer in presence of the clay [130].

In the case of modified starch composites, two main thermal transitions were

also detected at ~90 and 130°C (Fig. 52). A similar thermal transition was

detected in composites prepared in absence of clay (Fig. 4) at 110°C showing

that this broad shoulder was produced by the fungal modification of the

starch. However, it can also be observed that the temperature of this

transition shifted toward higher temperatures indicating that the clay

restricted the chain mobility of these fractions of the modified

polysaccharides.

107

Furthermore this shoulder may be also explained due to the intrinsic

molecular properties of the materials. Since the temperature will produce the

material’s internal molecular motion, the material softens and its length

increases, the instrument then adjusts the force to maintain the fixed strain

(elongation). As the polymer approaches the Tg the curve increases showing

the molecular resistance to change. As the maximum Tg is reached the

molecules lose strength (the material softens). After passing the Tg the

phenomenon becomes thermal dependent and the shape of the curve will be

depicted based on the molecular composition of the polymers, i.e.,

crystallinity if any, branching entanglement, etc.

20 40 60 80 100 120 140 160

0

200

400

600

800

1000

Temperature/oC

E' a

nd E

" (M

Pa) E'

E"

Tg by tan peak

0.25

0.30

0.35

0.40

0.45

0.50

0.55

0.60

0.65

Ta

n

Figure 56---DMTA curve profiles of native starch-glycerol-clay composites

108

20 40 60 80 100 120 140 160

0

500

1000

1500

2000

2500

3000

Temperature/oC

E' a

nd

E"

(MP

a)

E'Tg by tan peak

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

Tan

Tg by E" peak

Figure 57-DMTA curve profiles of modified starch-glycerol- clay composites

3.9.5. Conclusions

In general, the decrease of the storage modulus results from the inability of

the material to store energy. The loss modulus decreases more slowly

depending on the ability of the molecules to lose energy. The storage

modulus (E') is a measure of the stiffness of the sample. The stiffness can be

affected by various factors, i.e., the presence of different chemical species

acting like fillers or to increasing the cross-linking density. The loss modulus

(E") is related with the energy that cannot be recovered. In thermoplastic

starches this value can be associated with the diffusion of the plasticizers

through the starch mass. Also, the shape of the tan curve change

systematically with the amorphous content or molecular weight distribution

of the polymers involved as well as with the plasticizer content and type, i.e.,

the excess of plasticizer (i.e. water or glycerol) may produce sharper peaks.

109

Furthermore the shoulder at ~120oC detected in modified starch composites

may be also explained due to the intrinsic molecular properties of the

materials. Since the temperature will produce the material’s internal

molecular motion, the material softens and its length increases, the

instrument then adjust the force to maintain the fixed strain (elongation). As

the polymer approaches the Tg the curve increases showing the molecular

resistance to change. As the maximum Tg is reached the molecules lose

strength (the material softens). After passing the Tg the phenomenon

become thermal dependent and the shape of the curve will be depicted based

on the molecular composition of the polymers, i.e., crystallinity if any,

branching entanglement, etc. Moreover, the free volume theory indicates

that the molecular segments will not move randomly, but in specific

directions when an external stress is applied [111]. The resistance to change,

therefore, is due to the strongest molecular binding.

110

3.10. Thermal properties

3.10.1. TG (thermogravimetry)

3.10.1.1. Abstract

The thermal behavior of modified starches (MS) was described based on a

comparative analysis with native starches (NS). Thermogravimetric analyses

(TG) with successive derivatives (DTG) performed under non-isothermal

conditions, in an atmosphere of flowing nitrogen were used for this study.

Results obtained from NS samples showed a single peak dominating both the

TG (DTG) plots. A double thermal transition event was detected in samples of

MS. The interval of thermal decomposition (Ti–Tf; lowest onset temperature

of initial and final mass change) was carried out within a narrow interval of

temperatures in NS (610–640°C). Residues higher than 10% were recorded

for MS at temperatures of 1000°C. The presence of the double thermal

transition in MS and the shapes of the TG and DTG are discussed based on

the fundamentals that describe this technique.

3.10.1.2. Introduction

Starch is used mostly in food applications. The thermal properties of these

materials are of extreme importance for this industry. Depending on the

nature of the substituents and the degree of substitution (DS) the properties

of modified starches can be varied extensively, i.e. viscosity, association

behavior, film forming, thermal stability, or mechanical properties. Enzymatic

modifications of granular or gelatinized starches may also lead to the partial

molecular depolymerization and/or substitutions which may produce starch

derivatives with new and interesting industrial properties.

111

The starch structural differences should also be sensitive to the TG Such

structural differences should be Since the ratio amylose to amylopectin

depends on the starch source, the thermogravimetric (TG) curves have been

also used to investigate the differences among the various starch sources,

and it has been reported that the spectra of the degradation is sensitive to

the starch’s structural properties such as the ratio of amylose to amylopectin,

morphology as well as the processing conditions and/or modifications [131-

133]. However, the results are not conclusive compared with the previous

mentioned techniques.

In this study we investigate the properties of starch modified by specific

fungal strains by thermogravimetry (TG), and we also analyze the second

and the successive derivatives as a recurse to investigate the influence of the

starch structure and/or fungal modification on the shape of the plots.

3.10.1.3. Fundamentals of thermogravimetry (TG)

The Nomenclature Committee of the International Confederation for Thermal

Analysis (ICTA) thermogravimetry (TG) is defined as a technique by which

the mass loss of the sample is measured while it is subjected to a controlled

temperature program (normally heating the sample to a specific increasing

rate).

It is implicitly understood that this technique, unlike the complementary

techniques such as differential scanning calorimetry (DSC), cannot give any

information about reactions that do not involve mass change like

polymorphic transformations and double decomposition reactions. TG,

therefore, may not be useful for identification of a substance or mixture of

substances. On the other hand, when a positive identification has been made

by X-ray powder diffraction or some other method, TG can be used to

estimate or investigate the presence of a substance in a mixture or the purity

of a single substance.

112

The thermogravimetric curve is obtained by using a thermobalance. With this

instrument, the temperature range over which a reaction involving weight

change can be determined. This range depends not only on thermodynamic

and kinetic factors, but also on procedural variables, such as heating rate,

crucible geometry, atmosphere, sample weight, and sample preparation.

The thermogravimetric curve can be presented either as the TG curve, which

is a plot of the mass against time or temperature, with the mass loss on the

ordinate plotted downward, or as the derivative thermogravimetric (DTG)

curves, which are plots of the first derivative of the rate of change of mass

with respect to time or temperature.

3.10.1.3.1. General reaction of thermal decomposition

The more detailed explanation of thermal degradation can be found in

Treatise on Analytical Chemistry [134]. The basic reaction that controls the

spectra of the thermal decomposition as explained in this treatise is as

follows:

A(s) —> B(s) + C(g) (1)

This reaction can be described as the decomposition of a solid sample (s) in

another solid and the corresponding gas fraction (g). Therefore, any given

thermo analytical method necessarily involves the control of the following

variables: reaction rate and reaction environment.

In general, the instrument settings, which may affect the results, are: the

heating rate and the furnace atmosphere and geometry, as well as the flow-

rate of the gas supplied. The results can also be affected by the geometry

and material of pan (i.e., aluminum, metal). Some variations may also arise

113

due to the intrinsic properties of the samples: the mass, particle size, sample

history or pre-treatment, molecular packing, thermal conductivity, and heat

of reaction. For example, when comparing two similar samples which differ

just in the pre-treatment most of these variables do not diminish the results,

and the thermal spectra will depict the differences between the two samples.

Overall, the shapes of the TG and DTG depend on the procedural,

thermodynamic, and kinetic factors.

3.10.1.3.2. Definitions

The following quantities are involved in the determination of the thermal

spectra:

t = time

T = temperature

R = universal gas constant = 8.31434 J mol-1 K-1

k =Boltzmann’s constant = 1.380 X 10-23 J K-1

h = Plank’s constant = 6.626 X 10-34 J s

= extent of reaction.

The fraction of gas () may be defined as the fraction of total volume of gas

evolved or a fraction of total weight loss. Where two or more gases are

evolved the extent of reaction is due to the individual contributions to the

total volume of gas, for example, CO2 and O2 can be expressed as CO2 and

O2. Normally the fraction of gas goes from zero to a total of 1. In order to

detect the nature of the gases, some instruments incorporate infrared

systems.

114

3.10.1.3.3. Theory

The relationship between rate of reaction and extent of reaction is generally

expressed in the form

d/dt=f()k (2)

where = extent of reaction

d/dt = rate of reaction

f() = some function of

k = a temperature-dependent quantity.

Reaction rate may also be influenced by the presence of other gases other

than the product gases. However, such influence can be reduced or assumed

negligible by using either vacuum or an inert gas. Therefore, this Eq. 1 is

assumed to be sufficient to describe the kinetic behavior of decomposing

solids. Thus, two things must be determined:

f(), i.e., the relationship between reaction rate and the extent of

decomposition

How k changes with temperature, i.e., the relationship between

reaction rate and the temperature.

3.10.1.3.3.1. Relationship between reaction rate and

the extent of decomposition

The shape of a plot of “”against “t” (under non-isothermal or isothermal

conditions) is determined by the integral of the extent of reaction (Eq. 3).

(3) if

115

(4) then

(5) Thus the shape of the plot of against t is determined by g(). It should be

noted that Eq. 3 can be applied equally to iso- and non-isothermal results.

However, the simplest case is the isothermal analysis (Eq. 5).

The decomposition of a solid usually starts with the formation of small

localized areas of product called nuclei, which grow larger, forming an

interface between product and reactant that proceeds into the bulk of the

solid, gradually consuming the whole of the solid particle. Nuclei normally

form on the surface of the reactant. Various steps have been described

during this process such as surface desorption, surface decomposition, and

nucleation, followed by an induction period attributable to the rate of the

nuclei formation. The reaction is then preceded by a fast rate of

decomposition which reaches a maximum showing a decelaratory trend.

The generation of growth nuclei, establishes the existence of a reaction

interface and the subsequent growth of the interface; thus the form of f(a)

and g(a), is governed by

the type of nucleation

the geometry of the reactant particle

the influence of diffusion

The type of nucleation depends upon the relative magnitudes of DGn, the

free energy for nucleation, and DGg, the free energy for the growth of nuclei.

When Gg<<Gn, the growth of existing nuclei predominates over the

formation of new ones. This type of nucleation gives rise to sigma-shaped

decomposition curve because as the nuclei grow the rate of decomposition

increases until the growing nuclei begins to overlap, after which the rate of

116

decomposition progressively decreases. One of the most important general

expressions derived for g() and which corresponds to this type of behavior

is the Avrami-Erofe’ev equations:

(6)

Where “n” can take the values of 2, 3, or 4, and it may be 1 (first order

equation) and in this case the nucleation and reactant geometry do not apply

and individual molecules may decompose at random, or individual particles

nucleate and rapidly decompose at random which may be the case of

gaseous or liquid samples.

If DGg=DGn then a large number of diffuse nuclei form, none of which grow

to a visible size. Thus, the acceleratory period is reduced or completely

absent covered with small nuclei. The interface then proceeds at a constant

speed (under isothermal conditions) into the bulk of the solid. This behavior

may be described by the equation derived by Mampel:

1-(1-)1/n=kt (7)

where n has the value 2 or 3, and is the number of dimensions in which the

interface advances.

For the simple case of one-dimension diffusion the equation

2=kt (8)

has been shown to apply. For two-dimensional diffusion out of a cylindrical

particle the equation

(1-)In (1-) + =kt (9)

can be used. For three-dimensional diffusion out of a sphere the equation:

117

[1-(1-)1/3]2=kt (10)

or

(-2/3)-(1-)2/3=kt (11)

can be used.

The g() values discussed so far are summarized in Table 1. It should be

noted that the t referred to in these equations is measured from the start of

the decomposition process, i.e., after the end of the induction period if one

exists. Errors may arise because it is not always easy to accurately estimate

the point from which t should be measured. Also, some g(a)’s in Table 1 are

multiplied by a constant. This is because the differential form of the equation,

f(a), is assumed to give the correct value for k (as for heterogeneous

kinetics); thus, integrating f() according to Eq. 5 gives rise to a constant

that must be included when analyzing data using the integral form of the

kinetic equations (Table 7) [134].

118

Table 7

Commonly Used Kinetic Equations

Sigmoid Rate

Equations

F(a)=(da/dt)/k g(a)=kt Lab

el

1 Avrami-Erofe’ve (1-)[-In(1-)]1/2 2[-In(1-)]1/2 A2

2 (1-)[-In(1-)]2/3 3[-In(1-)]1/3 A3

3 (1-)[-In(1-)]3/4 4[-In(1-)]1/4

Deceleratory

4 First order (1-) -In(1-) F1

Based on

Geometric Models

5 Contracting area (1-)1/2 2[1-(1-)1/2] R2

6 Contracting

volume

(1-)2/3 3[1-(1-)1/3] R3

Based on diffusion

mechanism

7 One-dimensional

diffusion

-1 1/22 D1

8 Two-dimensional

diffusion

[-In(1-)]-1 (1-)In(1-)+ D2

9 Three-dimensional

diffusion

[1-(1-)1/3]-1(1-)2/3 3/2[1-(1-)1/3]2 D3

10 Ginstling-

Brounshtein

[(1-)-1/3-1]-1 3/2[1-2/3-(1-)2/3] D4

There is also another three types of nucleation: constant rate of nucleation,

continuously rate of nucleation, and instantaneous nucleation, all of which

give rise to an equation of the general form

-In(1-)=ktm (12)

119

where the value of m is determined by the type of nucleation, the number of

dimensions in which nuclei growth occurs, and wheter the reaction is phase

boundary or diffusion controlled. Table 2 provides values of m for Equation 8.

3.10.1.3.3.2. Relationship between reaction rate and

temperature

This relation is based on the on the Arrhenius equation. In general, the rate

of any given chemical reaction will increase with temperature [134].

3.10.1.4. Materials and methods

The equipment used consisted of a TG-DTG (thermo-gravimetry-differential

thermal analysis) unit from TA instruments model Q500. The integrated

software provided the TG and successive DTG derivative signals. The purge

gas flow rate was set at 100 mL min-1. Rising temperature experiments were

conducted in which the heating rate was 5oC/min for all experiments. With

the aim of avoiding the oxidation of the species, the thermal decomposition

was performed in an atmosphere of flowing nitrogen at 100 ml min-1. The

maximum temperature was set to 1000oC.

3.10.1.5. Results and discussion

TG and successive DTG have been largely used to study of the mechanisms

of degradation of starch and its derivatives. Although, just one-reaction

process has been observed in different starch sources, it has been shown

that particular species can be readily differentiated by the comparison of

their respective decomposition temperature intervals; Ti-Tf, where Ti is the

lowest temperature at which the onset of a mass change can be detected for

a given set of experimental conditions and Tf is the lowest temperature at

which the onset of the mass change has been completed [132].

120

TG curves and respective first derivatives of weight against temperature

(%/oC) (DTG) for native as well as modified starches at 1000oC are shown in

Figures 58 and 59. The curve of mass loss against temperature showed

apparently a single-stage of decomposition (Ti-Tf) in both samples of native

starches and modified starches. DTG plots of modified starches showed at

this zone of weight/loss transition the maximum temperature of ~307oC;

value which was slightly higher in native starches ~318oC. The percentages

of residual mass recorded at these temperatures were ~40% and ~50% for

NS and MS respectively. The weight loss in MS did not go to 100% even at

temperatures above 1000oC. Instead, remaining masses higher than 10%

were recorded for different samples of MS, including modified tapioca

(branching type C), potato (type B) or corn (type A) starches.

In contrast, the plot in Fig. 58 shows the complete degradation of native

starches at ~600oC. It was deduced at this point that some heat resistant

chemical species are being formed in the modified starches. Some examples

of such materials potentially formed include carbonaceous char- or graphite-

like residues or ionic structures which can derive inorganic compounds during

the process of degradation. For some samples of modified starches it was

found a slight gain in weight at temperatures over the 1000oC (in air at

elevated temperatures a polymer will eventually oxidize with subsequent

chain scission and degradation giving rise purely to a 100% gaseous mass,

the weight gain during the combustion in presence of air is an erroneous

result, due to the formation of oxide-like compounds) [134].

A plot of time against temperature appears always as a straight line with the

slope of the curve indicating the heating rate employed. The variations due

to the reaction of the material under increasing temperatures cannot be

easily appreciated due that the instrument automatically adjusts the heating

rate. However, differences in exothermic or endothermic heat capacity with

respect to the constant heating rate can be observed in successive

121

derivatives of weight (%/oC and 2nd derivative %/oC2) and times (min/oC and

min/oC2) versus temperature (oC).

The rate of mass change versus the temperature is shown in Fig. 60 (1st and

2nd derivatives of weight versus the temperature). The %/oC2 displayed the

minute separation of two endothermic peak events not observable in native

starches. Successive derivatives of time versus the temperature (min/oC and

min/oC2) for modified starches as well as native starches are shown in

Figures 61 and 62. The maximum fluctuations were observed in samples of

modified starches.

The first stage of discontinuity observed before the 100oC in the TG plots of

native starches was related to the loss of water. The loss of water in a first

stage of the heating was not observed in modified starches. For some

samples of modified starches (not shown in the Figures), the lack of phasing

with respect to the initial mass weight loss was probably due to the

environment of the decomposition (i.e., the presence of volatile products of

decomposition in the sample which probably induced their initial behavior). in

contrast, it is possible at this point that modified starches do not take

excessive water from the surronding environment, phenomenon which in

principle is due to the molecular arrangement of modified starch. The

presence of pendant groups or compounds of different molecular weights

may be the reason for this phenomenon as it was shown by 1H NMR, HPAEC-

PAD, and MALDI-TOF MS (chapters related to chemical characterization).

122

0 200 400 600 800 1000

0

20

40

60

80

100

Temperature /oC

Mas

s lo

ss/%

317oC

42%

Tf ,~ 610-640oC

0

2

Der

iv.

mas

s lo

ss (

%/°

C)T

i

Figure 58- TG-DTG plot of native starches showing the degradation point at the 1st derivative

200 400 600 800 1000

0

20

40

60

80

100

Temperature /oC

Mas

s lo

ss /%

309oC~60%

0

2

Der

iv. m

ass

loss

/(%

/o C)

640-660oC~6-15% residual material

Figure 59- TG-DTG plot of modified starches showing the degradation point at the 1st derivative

123

250 300 350

0

2

Temperature/oC

Der

iv. m

ass/

(%

/oC

)

-0.20

-0.15

-0.10

-0.05

0.00

0.05

0.10

0.15

2n

d Der

iv. m

ass

(%/o C

2)

Figure 60-TG-DTG plots showing successive derivatives for modified starches showing a clear double thermal transition peak

100 200 300 400 5000.05

0.10

0.15

0.20

Temperature/oC

De

riv ti

me

/(m

in/o

C)

-0.10

-0.05

0.00

0.05

0.10 2nd D

eriv. time

/(min/ oC

2)

Figure 61-TG-DTG-successive derivatives obtained by TG in native starches showing the lack of thermal transitions

124

150 200 250 300 350 4000.09

0.10

0.11

0.12

0.13

0.14

0.15

0.16

0.17

Temperature/oC

Der

iv. t

ime

/ (m

in/o C

)

-0.025

-0.020

-0.015

-0.010

-0.005

0.000

0.005

2nd D

eriv. time/(m

in/ oC2)

Figure 62- Successive derivatives modified starches showing high energy consumption during the transitions at the point of thermal degradation

3.10.1.6. Conclusions

Modified starches contain fractions which have a high thermal resistance.

Although, most of the material of modified starches decomposed at

temperatures near 1000°C, there was a resistant fraction with variable

percentages (up to 20%). This fraction could be the result of the thermal

decomposition to a recalcitrant material such as char. However, under the

same experimental conditions this fraction was not detected in native

starches. Moreover, the second derivative of mass loss showed a double

thermal transition. Successive derivatives might therefore be a useful tool for

elucidating the molecular structure of these materials.

125

3.10.2. DSC

3.10.2.1. Abstract

The thermal behavior of modified starches (MS) produced by biosynthetic

pathway is described based on a comparative analysis with native starches

(NS). MS were produced by fermentation in presence of Ophiostoma spp.

cultures. Differential scanning calorimetry (DSC) were used for this study. NS

results showed a single peak dominating the DSC plots. A double thermal

transition event was detected in samples of MS showing the presence of a

double glass transition temperature (Tg).

3.10.2.2. Introduction

Starch is the dominant food component in most diets around the world. This

material has been also widely studied and used for the production of edible

films as well as for the production of biodegradable materials for packaging

or production of disposable materials. Starch in contrast to other abundant

polysaccharides like cellulose, can be processed by extrusion or injection

molding in presence of common and non-toxic plasticizers like water or

glycerol. The functional properties of starch are intimately connected with the

unique chemical composition and architecture present in the native starch

granules. The deposition of the elements within the granules during the

starch synthesis is complex and not totally explained. However, there are

some general features in which most of the specialized researchers agree.

According to x-ray diffraction analyses, cereal starches are described as

having A-type molecular structure and tuber starches B-type. It has been

determined that the differences arise due to the presence or absence of

central double chains within the central cavity formed by a hexagonal array

of double helices which form the crystalline structure (or lamellae) in the

amylopectin molecules.

126

In the absence of the central chains (B-type starches) the central cavity may

be filled with water molecules. The application of heat in excess of water to

fully bring about one of the most important characteristics of starch, namely

gelatinization. Almost all food preparations involve gelatinization as does the

production of thermoplastic starches by extrusion or injection. During this

phenomenon the granules swell, lose molecular order, and the viscosity of

the solution increases. In the DSC spectra, in excess of water (>40-50%)

one sharp peak is observed between 60 and 80oC, but as the water content is

lowered the peak broadens and shift to higher temperatures (between 100

and 130°C). Amylose-lipid complexes may also produce a thermal

discontinuity in the DSC curve, but unfortunately there is no a general

agreement in the association of this transition with a particular peak.

DSC has also been used to determine the specific Tm (melting point) and Tg

(glass transition temperature) transition points of both granular and

amorphous starchy materials. Many authors have reported results with wide

variations which could be attributed to the measurement conditions,

plasticizer type and content as well as to the complex behavior of starch

physical morphology. In addition, the heat conductivity of granular starch can

be also increased with salt solutions, such as KCl, in such cases there was a

shift of the detected endothermic peaks towards lower temperatures. In

general, all results are entirely dependent on the moisture content (mc), or

plasticizer as well as thermal conductivity. Both phenomena (Tg and Tm) are

thermodynamically irreversible and they are not well differentiated from each

other. For example, waxy maize tested in aluminum pans at ~13% mc has

presented an endothermic heat flow at ~60oC with onset at ~40oC. This peak

was considered as a change of heat capacity associated with the Tg. When

this sample was tightly packed in stainless steel O-ring pans by ultra-sound

bath the thermal plot showed a double transition event with peaks at 170

and 190oC. In O-ring pans loaded with ~26 mg of sample at ~56% water

content the observed endothermic transitions were 4 at ~70, 90, 110 and

127

130oC. The origin of these peaks was not explained by the respective authors

[135].

On the other hand, rice starch tested in aluminum sample pans at ~60% mc

produce a strong peak at ~70oC which was related to the gelatinization

temperature of the starch (onset temperature ~60oC and final temperature

90oC) [136]. Two peaks in waxy wheat found at ~65 and 105oC (starch

tested at mc of 50% in aluminum pans) were related to the starch

gelatinization and the dissociation of amylose-lipid complexes respectively

[137]. Blends of different starches do exhibit two different heat flow

transitions [138]. Gelatinized starches also exhibit one endothermic peak

associated to Tg [139].

Gelatinization is a complex process which includes the disruption of the

crystalline regions (Tm) within the granules, while the phenomenon of the Tg

occurs just to amorphous materials. The differences in temperatures for the

endothermic peaks among different starch samples are associated therefore

to the starch composition (amylase to amylopectin ratio), granular

architecture (crystalline to amorphous ratio) and Mw (molecular weight) as

well as polydisperisty of the chains [140-142]. In sealed DSC pans with

potato starch one peak is reported by Septo [6]. The endothermic transition

was detected at ~75oC with 45% moisture content (mc) and ~150oC at 12%

mc.

In polymer theory, Tg is the glass-rubbery transition that occurs in

amorphous polymers and Tm (melting point) occurs when the ordered

regions of a polymer fall apart upon heating. However, Tg of dry starch is

inaccessible owing to the thermal degradation of starch before reaching it.

Upon heating in excess of water, the starch granules collapse together with

the crystalline regions. This last phenomenon has been associated with the

gelatinization process. Plasticizers (water, glycerol, sorbitol, sugars, etc.) do

affect the thermal transitions upon starch heating and lower the Tg

128

temperature. Starch thermal transitions in presence of water have been

studied for a long time due to its importance in the food industry.

Theoretically, water may induce the glass-rubbery transition (Tg) of

amorphous regions of starch before the melting of crystals, but the frontier

between Tg and Tm is not well defined. Poutanen and Forssell [143] reported

that the Tg in water can be found by two methods, the free-volume approach

and the thermodynamic theory.

The free-volume approach (assuming the glassy state is an iso-free volume

state) consists of studying the plasticizing effect at different water contents

and by studying the oligomeric behavior with extrapolation to molar mass.

The applicability of the free volume approach may be limited as the

dependence of free volume on molecular weight, intermolecular forces, chain

flexibility, chain geometry and structural detail such as the bulkiness of side

groups is not taken into account. With respect to biopolymers such as starch,

intermolecular interactions and hydrogen bonding may play a significant role

in plasticization behavior. The free volume approach therefore, may be useful

in some cases. The second method associated with the Tg of the amylopectin-

water system is based on the thermodynamic theory (assuming continuity of

the excess entropy of mixing at Tg). Both methods finally depend also on the

Cp (heat capacity) measured by DSC (differential scanning calorimetry).

Therefore, the thermal transitions measured by DSC can describe in

relatively detail physical and chemical changes related to Tg or Tm of starch. A

third method not described by the thereafter authors which can provide more

detail and great accuracy is DTMA (dynamic thermal mechanical analysis).

3.10.2.3. Materials and methods

Differential scanning calorimetry (DSC) thermograms were acquired to study

the course profile of the gelatinization process in both starches; modified and

unmodified, in presence of KCl to increase thermal conductivity in the starch

samples. DSC was performed with a Q1000 unit from TA Instruments fitted

129

with a cooler system and nitrogen gas as purge. All samples were tested by

triplicate in aluminum pans with 30 l volume capacity-PE No.-BO169320,

and empty pan as reference. Samples were made up as starch slurries. The

maximum temperature reached 250oC at a heating rate of 5oC/min, with size

samples of 5-7 mg. DSC data parameters collected with the software

integrated (TA Instruments; Universal Analysis) included: onset temperature

(TON), maximum peak at melting point (TM), final melting temperature (TC),

melting temperature interval or the start and conclusion temperatures (TR)

and change heat capacity (Cp) at the measured melting point. TR is defined

as the point at which the DSC trace line first ceases in following a straight

line and finally recovers the base line. The onset temperature was taken at

the first inflection point on the trace curve line of the DSC endotherm. The

maximum endothermic peak temperature is normally used for the

determination of the enthalpy of gelatinization.

3.10.2.4. Results and discussion

DSC parameters are shown in Table 8: Onset (TON), maximum peak at

melting point (TM), final melting temperature (TC), melting temperature

interval (TR), heat capacity (Cp); the enthalpy of fusion (Hfus) can be

deduced from Cp at the recorded temperatures as shown in the Table. The

main differences observed for modified starches in powder form with respect

to the rest of the samples. However, one of the most interesting results was

recorded from films manufactured with modified starches. Before film

formation the TR was ~64 and after the manufacture of the films this value

was lower than films manufactured with native starches suggesting a better

molecular packing during film formation. Similar effect was found for TON, TM,

TC and the maximum endothermic heat flow. The Cp was as low as 8.2 J/(g oC) in modified starches in powder form, however, after film formation this

value is similar to that found for films manufactured with starch granules

being of ~70 J/(g oC). Similar results were consequently observed with the

respective melting temperatures. The melting temperatures of granular

130

native starches and films did not show such a sensitive variation. It is

possible to infer a high degree of disturbance after the modification,

however, the film formation showed a high level of molecular organization

comparable to native starches.

Table 8- DSC melting parameters for the substrate and modified starches

Samples DSC melting parameters

TR

(oC)

TON (oC) TM (oC) TC (

oC) Cp [J/(g oC)] measured at the

respective melting temperatures

in oC

Modified

starch(powder

form)

64.2 35.9 74.9

83.8

100.2 ~8.2

~9.9

Unmodified

starch in KCl

(granular form)

17.4 103.1 108.6 120.4 ~23.4

Modified starch

films in KCl

16.4 103.7 107.4

108.2

120.2 ~68.4

~66.7

Unmodified

starch films in

KCl

18.5 101.5 106.7 120.9 ~72.3

Figures 63 and 64 show the endothermic curve profiles for native as well as

modified starches in powder form gelatinized with a solution 0.6 M of KCl.

Native starches exhibited a narrower endothermic peak with a maximum

temperature of ~108oC, while modified starches displayed the two peaks at

~74 and 84oC. Onset transition temperatures were found at ~36oC in

modified starches while this same transition was located at around 103oC in

unmodified starches. The peak at the maximum endothermic heat flow for

modified starches was observed at lower intensity. In presence of DMSO,

modified starch samples also slightly displayed two maximum peak transition

points at ~73.57 and 81.20oC. The thermal transitions were smother

comparer with starches gelatinized in KCl and the melting temperature

interval was broader.

131

The endothermic profiles obtained from films manufactured with modified as

well as unmodified starches and analyzed in presence of 0.6 M KCl are shown

in Figures 65 and 66. The endothermal course profiles recorded on films

manufactured with native starches displayed an abrupt transition at ~102oC,

which is also the onset temperature for this sample, while the films from

modified starches showed a smoother transition at the onset value of

~103oC. Again, the temperature course profile in the case of modified

starches showed the two thermal transition peaks at the tip of the curve with

very close temperatures.

The water loss measured by DSC in modified and native starches films is

shown in Figures 67 and 68. The films were conditioned at constant weight at

laboratory conditions (65% RH-relative humidity-, 20oC) for 3 weeks. The

water bending curve due to water evaporation recorded for MS was much

lower compared to films fabricated with NS. In general modified starches

tend to adsorb less water at the same conditions of relative humidity.

132

20 40 60 80 100 120 140

-14

-12

-10

-8

-6

-4

-2

0

Hea

t flo

w (

mW

)

Temperature (oC)

End

othe

rmic

~108oC-14 mW

Figure 63- DSC thermograms of the unmodified substrate (native starch). Peaks induced with 0.6 M KCl (granular starch)

20 40 60 80 100 120 140-8

-6

-4

-2

0

He

at fl

ow

(m

W)

Temperature (oC)

endo

ther

mic

~83oC-6.4 mW

~74oC-7 mW

Figure 64- DSC thermograms of modified starch

133

20 40 60 80 100 120 140

-50

-40

-30

-20

-10

0

He

at fl

ow (

Mw

)

Temperature (oC)

~102oC-6.8 Mw

~107oC-52 Mw

Figure 65- DSC thermograms of native starch-films

20 40 60 80 100 120 140-40

-35

-30

-25

-20

-15

-10

-5

0

Hea

t flo

w (

mW

)

Temperature (oC)

End

othe

rmic

107oC-35 mW

108.5oC-35 mW

109 oC-34.5 mW

Figure 66- DSC thermograms of modified starch films

134

50 100 150 200

-12

-10

-8

-6

-4

-2

0

2

Hea

t flo

w (

mW

)

Temperature (oC)

~122oC-11.5 mW

Figure 67-DSC thermograms of native starch films

0 50 100 150 200

-3

-2

He

at fl

ow

(m

W)

Temperature (oC)

~106 oC-2.8 mW

~170oC-3.21 mw

Figure 68-DSC thermograms of modified starch films

135

3.10.2.5. Conclusions

Two thermal transitions were detected in samples of modified starches

and one in the respective native counterpart. Even after film formation, this

double transition was detected in the chromatograms taken from modified

starches. The analysis also showed that native starches are more susceptible

to water absorption. This technique is by far less sensitive than DMTA to

detect the glass transition temperature of the polymers. Even though, the

results support the findings of the viscoelatic tests (mentioned in chapter

3.9).

136

3.11. Rheology

3.11.1. Introduction

The rheological behavior of melted thermoplastic starches under

extrusion conditions is of great interest for the control operations and

properties’ design of extrudates. From the transport phenomena point of

view, rheological behavior represents the most important property for all

starch derived products, including bio-plastics. A difference of other

properties like thermal conductivity, heat capacity, or density, viscosity of

starches can greatly vary due to the effects of plasticizer type and content.

The production of thermoplastic starches produced at low water contents is

of increasing importance, since higher plasticizer contents cause distortions

and shrinkages which causes problems in the final products[6]. Available

data on rheological behavior of food pastes is widely available [144] as well

as comprehensive studies related to the parameters influencing the melting

behavior of thermoplastic materials intended as substitutes for petroleum

based plastics obtained by in-line and off-line capillary rheometers [145,

146].

3.11.2. Materials and methods

3.11.2.1. Sample preparation

The water content of the starches at the moment of the extrusion was ~7%

in MS and 11% in NS. It was determined by weighting a known mass in an

oven at 100°C for 2 hours prior to the mixing of starch with 40% glycerol.

The mass production was on average 1.5 kg/hour for MS and 2.3 kg for NS.

The average temperature extrusion profiles established experimentally were

by maintaining constant the speed screws (120 rpm) and the glycerol

content (40%) were: 120/120/130/130/130/140/140/140/150/160°C for

137

MS, and 110/110/120/120/120/130/130/135/150/150°C for NS. At the same

motor current energy used; ~2.7 A which is the difference in electrical

current between the loaded and unloaded barrel, the mass flow rate was

lower for extrudates of MS (0.00045 kg/s) than NS (0.0007 kg/s).

3.11.3. Instrumental conditions

The rheological properties were measured by using a twin bore capillary

rheometer Rosand RH2000— radius 15 mm and barrel length 250mm—, and

Flowmaster® software for data analysis. The apparent wall shear stress (w-

ap) was corrected for entrance and exit pressure losses by Bagleys’s

correction to give the true wall shear stress (w), and the apparent wall shear

rate (w-ap) was corrected by Rabinowitsch-Weissenberg equation to give the

true wall shear rate (w). The ENS and EMS pellets were melted, equilibrated

for 5 min, and extruded at 150°C and 160°C repectively. The differences

between the materials will be discussed based on the Power Law Model, and

used to describe the behavior of shear thinning fluids. The two parameter

data (w, w) was adjusted by the squares method with the Flowmaster®

software to a power law relation and given by:

Kn

where K is the consistency coefficient and n is the power law index.

Viscosities reported at 40% glycerol content were true apparent and true

viscosities () obtained from wall shear stresses and shear rates:

=/=Kn-1

The SME in the capillary was calculated by using the following relation:

SMEcap=Pcap Av/m’

Where A is the barrel area, v the plunger speed, m’ the mass flow[147].

3.11.4. Results and discussion

Fig. 69 shows the results for shear viscosity (Pa.s) as a function of shear

rate (s-1) for the two thermoplastic (TP) materials: modified starches (TP-

138

EPSs) and native starches (TP-native starches) (see also Fig. 70). It can be

seen that, for both samples the shear viscosity decreased and the shear

stress (kPa) increased with increasing shear rate. A strong power law

dependence of viscosity on shear rate is observed as it is normal for these

materials. The dependence was linear on double-logarithmic plots indicating

the power law model can be used to describe the rheological behavior of both

starch-based materials [147].

n=Kn-1

where n is the melt viscosity, K is the consistency, g is the shear rate, and n

is the pseudoplasitc index. The corresponding consistency and pseudoplastic

index; normally determined by the intercept and the slope of each single

straight line in the double-log plots, were determined by using the software

provided. Again, for both materials, the values obtained are between zero

and 1 showing the typical shear thinning behavior associated whit starch-

based materials. However, the shear viscosity and shear stress were sensibly

higher in TP-EPSs. Although the shear stress is similar at low shear rates, it

was observed that the shear stress increased faster with the increase in

shear rate. These results are in accordance with the higher temperatures

required for the extrusion of EPSs-glycerol mixture and melting of TP-EPSs

during the rheological tests used for thermoplastics based on EPSs.

The log-log plot of shear stress vs. shear rate showed that TP-EPSs present a

higher dependence on shear viscosity than TP-native starches. The linear

fitted logarithmic plots for TP-EPSs showed a slope of ~72o, while this value

for TP-native starches was of ~63o showing clearly a higher dependence of

TP-EPSs on shear rate (the behavior of the plot of a pure solvent; low

molecular weight, may be expressed by a straight line inclined at 45o in the

log-log coordinates, showing the typical Newtonian character), phenomenon

which is related to a more marked non-Newtonian behavior of TP-EPSs.

139

Shear-thinning curves may exhibit three distinct regions: a lower Newtonian

region where the apparent viscosity (), called limiting viscosity at zero shear

rate, is constant; a middle region where the apparent viscosity () changes

with the shear rate (decreasing for shear-thinning fluids) which can be

adjusted to the power low equation; and an upper Newtonian region where

the slope of the curve (∞) is constant with changing shear rates, also called

limiting viscosity at infinite shear rate. The power law relation also relates the

stress ( by a constant K (the consistency index) and the power low index

(n) to the shear rate () by:

= n or In () = In (K)+n In ()

The K values at different points of the flow curve are shown in Table 9. This

value was higher in EMS at low shear rates, and the power law index

decreases abruptly with the increase in shear stress.

101 102 103 104102

103

104

She

ar v

isco

sity

(P

a.s)

Shear rate /s

TP-EPSs

TP-Native starch

Figure 69- Plots of shear viscosity vs. shear rate (TP-native starch; TP-EPSs-Modified starches)

140

5.0x102 1.0x103 1.5x103 2.0x103 2.5x103 3.0x103

2.0x102

4.0x102

6.0x102

8.0x102

1.0x103

She

ar s

tres

s (k

Pa)

Shear rate /s

TP-EPSs

TP-native starch

Figure 70-Shear stress vs. shear rate

(TP-native starch; TP-EPSs-Modified starches) Table 9- Calculated K values for modified and native starches

Shear stress (kPa) Shear rate (/s) n Calculated K value

(= n) TP-Modified starch

52.519 20.000 0.611 8.45 109.71 69.314 0.588 9.01 223.21 237.98 0.567 10.0 475.72 835.39 0.545 12.1 889.27 2895.7 0.523 13.7

TP-native starch 42.888 20.000 0.428 11.9 71.991 69.314 0.487 9.13 132.19 238.01 0.546 6.65 280.01 835.36 0.606 4.75 575.19 2895.3 0.667 2.82

141

3.11.5. Conclusions

The flow properties of exopolysaccharides (EPS) produced by Ophiostoma

spp. in the presence of starch were tested and compared to native starches

(NS). Rheological behaviors of extruded native (ENS) and extruded EPS (E-

EPS) with 40% glycerol content were evaluated in an off-line twin capillary

rheometer to mimic the flow properties of the extrusion conditions. Both

starches showed shear-thinning behavior, but E-EPS showed higher

plasticity. Consistency coefficients K determined by the power law model

were significantly higher at low shear rates in E-EPS, and the power low

index decreased abruptly at higher shear rates than 500/s-1. Mechanical

properties showed that E-EPS are stiffer materials than ENS.

3.12. Mechanical properties

3.12.1. Abstract

The mechanical properties of starch composites were briefly investigated.

The materials were produced by using the standards ASTM D638, Type I and

D638-5-IMP-ASTM. The results were similar. The results from modified

starches showed a more complex behavior than native starches. However, a

tendency was followed similar to that reported in previous studies by Huang

et al. [43] in which the tensile strength (and therefore the modulus) was

higher in modified starches compared with native starches. In addition, in

this section, the interactions between starch and clay in the presence of

glycerol were analyzed. It was found that, contrary to native

starch/clay/glycerol composites, in composites prepared with modified

starches the clay did not increase the elongation at break. These results

support the information obtained for the viscoelastic as well as chemical

properties.

142

3.12.2. Introduction

By itself, starch is a poor choice as a replacement for any plastic. The

hydrophobic nature of thermoplastic starches makes them susceptible to

moisture attack and as a result the dimensional changes produce important

modifications in the mechanical properties. In addition, retrogradation and

crystallization of the mobile molecules increase the complexity of the system

and it is more difficult to control the variables during a specific process.

The ratio of amylose to amylopectin also affect profoundly the properties of

the different starch sources (corn, tapioca, potato, etc.). The long, linear

chains of amylose readily associate through extensive hydrogen bonding

facilitated by the stereoregularity of the backbone and the wealth of the

hydroxyl functionality (4 per anhydroglucose moiety) giving good film

forming properties, while the ultramolecular high molecular weight of

amylopectin is associated to extensity crosslinking network formation.

The starch molecular association based on hydroxyl groups, and the

branching pattern of amylopectin also have a great impact on the rheological

behavior of the starch. It does not flow in presence of heat like a

conventional thermoplastic polymer. Phenomenon which has been associated

as having a lower degradation point than the glass transition temperature

(Tg) or the melting point (mp). In order to reduce the Tg of starch it is

necessary the use of a plasticizer. For example, to produce films by casting

method or to process starch by extrusion or injection molding it is necessary

to induce the disruption of the granules—and the crystalline regions—in

presence of a suitable plasticizer (phenomenon widely familiar known as the

gelatinization of starch).

The addition of plasticizer to starch is accepted as the means for lowering the

glass-rubber transition temperature (Tg) below the decomposition

temperature to make it more flexible. Glycerol and water are the most widely

143

used plasticizers. Glycerol is frequently used because is a non volatile

material and can remain in the mass after the extrusion. A number of studies

on the effects of plasticizers on starch have been carried out with the aim of

enhancing the properties of the thermoplastic starches.

Various authors have used a combination of glycols to plasticize various

sources of commercial starches (corn, tapioca and potato). The influence of

plasticizer content on the Tg of the starch-based materials has been explored

i.e. by Lourdin and co-workers [148]. These authors pointed out that the

efficiency of the plasticizer is governed by its ability to form favorable

interactions (probably hydrogen bonding) with starch. Moreover, the

flexibility of thermoplastic starches depends on the plasticizer content and

type. Even though, the properties of these starches cannot match the

efficiency of synthetic plastics. Therefore, further modifications are still

needed.

The use of reinforcing materials in the starch matrix is an effective method to

obtain high-performance starch derived products, i.e. there is an increased

use of cellulosic fibers, micro and nano-particles (clay or fibers) as the load-

bearing constituent in developing new and inexpensive biodegradable

materials due to their high modulus and high aspect ratio. However, it has

been observed an uneven distribution of the glycerol in the starch matrix and

its accumulation on the reinforcing phase resulting in poor mechanical

properties.

Myllarinen et al. [117] measured the effect of water and glycerol content on

the Tg of amylose and amylopectin films. They also obtained information on

the mechanical properties of films prepared at various contents of glycerol at

constant relative humidity (RH) and temperature. To produce the films,

amylose and amylopectin with different percentages of glycerol (10, 15, 20

or 30% starch dry basis) were dissolved in water at 140oC under pressure for

30 minutes and constant stirring. Films were produced by casting method,

144

stored at 20oC at RH 50% for one week before testing. Tg was measured with

DSC and was taken as the midpoint of the change in heat capacity (Cp).

Mechanical properties of stress and strain of the films (20X80 mm) were

measured by using the standard method ISO 1184-1983. The results showed

a slow decrease of Tg in presence of glycerol, and water was a better

plasticizer. Brittleness increased (reduction of %E) with the reduction of

glycerol. Overall, amylose films were stronger than amylopectin films.

Lawton [149] reported the production by the casting method of starch-

poly(vinyl-alcohol) films with the following formulation: 41% starch, 41%

PVA, 15% glycerol and 3% poly(ethylene-co-acrylic acid). The use of

different starches was reported: native corn starch, high amylose corn starch

(50 and 70% amylose), wheat starch, potato starch and tapioca starch. The

films were aged between 7 and 168 days before testing or stored at a

relative humidity (RH) of ~93% for 7 days. Tensile strength (TS), percent

elongation at break (%E), tear resistance and impact strength were tested

for the characterization of these materials. This author reported the following

results: an increase in percentage of %E and a decrease in TS as RH

increased, higher amylose films showed the greatest stability in %E, the

larger decrease of TS occurred at RHs between ~15 and 30% followed by a

linear decrease of TS as RH increased, tear resistance was low for all almost

all the range of RHs (from 15 to 93%), results for the impact strength of the

films were similar (waxy corn starch films presented lower properties), aging

for 28 days did not affected significantly the impact strength, however all

films (except high amylose starches) showed a significant decrease in %E

after aging for 168 days, TS increased with aging (~35MPa). A general

conclusion was that films prepared with high amylose starches were more

consistent over the entire range of conditions. Loss of tear resistance at

higher RHs was due to the molecular mobility, increase of TS and the

decrease of %E with aging is due to the loss of plasticizer content.

145

The behavior of the glycerol varies according to the composite formulation.

Starch-clay-glycerol films produced by the hot press method were recently

reported by Zhang et al. [123]. These authors used modified and pristine

clays. Micrographs showed an irregular distribution of the materials in the

composite. They suggested that the spacing among the materials in modified

clay films was due to the aggregation among the starch molecules, while the

spacing found in films with pristine clays was due to the presence of glycerol.

In spite of the starch source the same trend is shown for other films

prepared by the casting method. Films produced with cassava starch showed

and increase in strength in presence of higher amounts of amylose. With the

increase in glycerol concentration an observed an increase in the fracture

stress and water vapor permeability, and the stiffness decreased were

observed [150].

Godbillot et al. [151] report a maximum of 20% glycerol at a ~44% RH in

plasticized wheat starch films also produced by casting method. They

reported phase separation above this percentage and attachment of

excessive moisture to the starch and to the free glycerol molecules.

Laohakunjit and Noomhom [152] reported critical values of plasticizer at

which rice starch can be dissolved to produce starch films with improved

properties, i.e., 35% glycerol, 45% sorbitol; polyethylene glycol was reported

not suitable for film formation. In overall, the properties of films are similar

to those reported in the general literature. The TS was reported lower for

higher concentrations of plasticizer and the %E resulted higher. TS was lower

in glycerol plasticized films compared with sorbitol, but the elongation was

reported higher. The water and oxygen transmission rates increased with

plasticizer concentrations. Similar results related to the effects of glycerol

content on the TS, %E and water transmission rate were reported by Lopez

et al. [153] for films prepared with acetylated corn starch. Doungiai and

146

Sanguansr [154] also presented similar results for plasticized sorbitol tapioca

starches.

Montaño-Leyva et al. [155] also reported similar results for films prepared by

casting method by using wheat starch (durum). TS, %E, EM (elastic

modulus), solubilities were reported for two different glycerol concentrations

(25 and 40%). The films were transparent (amorphous starches). With 25%

glycerol the films turned brittle and the reported values were: TS=42-50

MPa, E=1.4-2.7%, EM 31-34 MPa). Films with 40% glycerol presented TS of

11-17 MPa, E=4-41%, EM=4-11.3 MPa). It can be observed that these

properties are affected by the glycerol concentration, in spite of the starch

source this trend is observed in all starches. By XRD analysis, these authors

reported a semi-crystalline structure. In general, the preparation of the

samples with 25% glycerol of TPS results very difficult. And the manipulation

of starch based materials plasticized with glycerol-water is in fact also very

difficult due to the sticky surface of the films, insufficient tenacity and

foaming [156].

Bertuzzi et al. [157] reported the production of films with high amylose corn

starches. The films were produced at low temperatures. The gelatinization

temperature was reduced by pre-treating the starch with alkaline solutions.

Films forming suspensions showed thixotropic behavior. The apparent

viscosity of films forming gels increased exponentially under the increase in

amylose content and the Arrhenius law represented variations with

temperature. The alkaline treatment of high amylose corn starches previous

to gelatinization produced low solubility and opacity and an increase in

crystalline regions reaching an asymptotic value after 60 min. Water

absorption and film opacity increased with the increase of glycerol content

(effect observed at 30% glycerol concentration). Opacity can be attributed to

the formation of crystalline fractions of retrograded amylose molecules.

147

Jansson and Thuvander [158] prepared starch films by using the casting

method with hydroxyl propylated potato starch mixed with 30% glycerol and

water (18% starch based on water). The mixing was carried out for 30 min

at 90oC until dispersion of the starch molecules. Bars of 7X70mm2 were

produced with variable thickness from 0.3 to 2.6 mm. The Tg was taken by

using DSC and DMA and was reported at ~38oC. The films were characterized

in terms of stiffness, strength and failure strain as well as by fracture

toughness which were measured by single edge notch tests. In these

particular experiments, stiffness increased as thickness increased from 0.3 to

1.0 mm. As thickness increased further, the stiffness decreased. This

behavior was attributed to the differences in molecular stretching produced in

response to water evaporation. This explanation is in accordance with the

process by which starch molecules retrograde in presence of water, and form

crystalline regions. As the water evaporates slowly in thicker films, the starch

molecules have lag periods of relaxation and therefore form more crystalline

regions. It was speculated that the higher values in stiffness may be due to

an equal rates in the molecular relaxation versus water evaporation. This

induced artificial stretching may produce a molecular deformation in direction

in the plane of the film. The strain at failure was reported to decrease with

the increase in thickness.

The TS increased with the thickness up to ~1.0 mm (up to 4 MPa) beyond

this point this property decreased (1.5 MPa). The explanation would be

similar to that offered for the behavior observed of the stiffness. The

formation of oriented crystalline fractions (filaments) may induce higher

tensile strengths. Such fractions may be formed by the aligning of the

molecules in the direction of the stress while heating the amorphous material

above its Tg (but below its melting point). The stress would be produced

during the evaporation of the water from the films. Oriented material may be

at least five times stronger than the unoriented material. In general, when

liquids compose of complex molecules or ions (e.g., sucrose or silicates and a

vast number of organic polymers) are cooled rapidly a glass may be formed.

148

Glasses are thus examples of non-crystalline solids. Glasses do not show a

sharply-defined thermal transition peak, but soften over a temperature

interval. The strain at failure (or fracture) was reported to increase with

increasing thickness up to ~1.0 mm, and decreased with the further increase

in thickness. The measured fracture toughness showed similar results.

Fracture toughness showed an increment when the film reached ~1.0mm

thickness.

Pushpadass et al. [159] reported the extrusion of corn starch (at 20%

moisture content based on starch dry basis) in presence of glycerol in a ratio

of 3:1 (~35%) with the further preparation of films. They used stearic acid

sucrose and urea at varying concentrations as secondary plasticizers for the

starch-glycerol mixture. The extrusion was carried out at 110 and 120oC

(barrel temperature). The physical and mechanical properties of the films

were studied by SEM and tensile testing. Tg was determined by DSC. The

interaction between functional groups of starch and plasticizers were

determined by FTIR. Water transmission rate was determined by using the

standard ASTM E96-95. SEM images showed the presence of partially melted

starch granules in the extruded material. The TS was 0.9 to 3.2 MPa, strain

at break 26.9 to 56.2% and Young’s modulus of starch films ranged from 4.5

to 67. 7 MPa. DSC displayed two Tg values in the temperature ranges of 0.1

to 1oC and 9.6 to 12oC (up to know just one peak for Tg had been reported

for glycerol plasticized starches). Multiple endotherms were observed in

thermoplastic extrudates. The gelatinization enthalpies of the extrudates

varied from 0 to 1.7 J/g and it was associated this variation to the extrusion

temperature and plasticizer content. The shift in the FTIR spectral bands

were related to bonding interactions between the starch and plasticizers. The

water transmission rate was reported between 10.9 and 15.7 g mm h-1 m-2

kPa-1, variation which was also associated to the extrusion temperature and

type of plasticizer.

149

Averous et al. [114] reported the production of TPS –wheat starch-glycerol-

water-cellulosic fibers-. These materials were extruded in a single screw

designed machine and further used for injection moulding. Extrusion was

carried out at 150oC and injection at 130oC with an injection pressure of 1000

bar and holding time of 23s. Mechanical properties were measured in an

injection moulded dumbbell parts. Thermomechanical properties were

analyzed by the DMA in the dual cantilever geometry at a frequency of 1Hz

and a heating rate of 1.5oC/min. Plasticized wheat starch had a maximum

strength of 3 MPa which showed an increase in presence of fibers. The strain

at break decreased and the elastic modulus increased as normal. Plots of

storage modulus and tan versus temperature showed the shifting towards

higher modulus and temperatures respectively for reinforces materials.

In general, it can be seen that native starches (~ 30% amylose, 70%

amylopectin) may be extruded between 100 and 150oC, the screw speed

varies with the type of extruder and should be set accordingly to the need of

the extrusion, but some values can be range from 40 to 120 rpm. Injection

molded parts can be prepared according to the standard ASTM D 638, Type I

dog-bone shaped to measure the properties. The temperatures in the barrel

can be kept up to 170oC at 90 bar.

3.12.3. Materials and methods

3.12.3.1. Sample preparation

Native corn starch used in this study was obtained from Casco Inc. Canada.

the results reported are the average of 20 different batches. Production of

modified starches and preparation of films were carried out according to the

methods followed by Jeng et al. [42]. Glycerol was from Aldrich.

HALLOCOTE® 466 hydrasperse clay from HallStarch Co. was from L.V.

LOMAS LTEE.

150

The dry fractions (starches and 30% clay base on starch dry weight) were

premixed followed by the addition of glycerol (40%) and rigorous mixing with

a commercial food blender. The mixture were then stored in plastic buckets

with hermetic lids before extrusion.

3.12.3.2. Extrusion

The process of extrusion was carried out with a laboratory twin-screw

extruder ONYX TEC-25/40, with a screw nominal diameter of 5 mm, twin-

screw centre distance 21.2 mm, L/D 40, ten heating zones, and three

venting ports.

Although, it was variable, the averaged speed in the feeder was

approximately 8 rpm and the screw speed 120 rpm. The energy input was

~4.1 A. The melting averaged temperature varied from 120 to 140 from

native and modified starches respectively. The venting ports were kept open

to allow complete removal of any moisture. The resulting extrudates,

obtained in rods of 3 mm, were immediately cut into pellets, cooling with air

and stored in plastic bags for further processing in the injection molding

machine. The extrusion conditions (speed of the feeder, speed screw,

temperatures, energy input) for the mixtures of native starch-or modified

starch-glycerol clay were similar.

In general, the following heating spectrum was follow for the extrusion of all

materials: screw speed (rpm) 120, temperature (°C)

120/120/130/130/130/140/140/140/150/160°C for modified starches, and

110/110/120/120/120/130/130/135/150/150°C for native starches,

residence time 200-400 (s), motor energy input 4±1 A.

151

3.12.3.3. Injection molding

Pellets after the extrusion were used to produce standardized ASTM D638

Type I dog-bone specimens by using a conventional injection molding

equipment ENGEL, Ludwing ES 80/28. Experimental molding conditions

were: barrel temperature: 140/145/150°C; die temperature: 150 for native

starches 160°C for modified starches; injection pressure, 100 bar; mould

temperature 60°C. The specimens were conditioned at 50% RH, at 2oC 24

hours before testing [117]. Maximum tensile strength, strain rupture, and

elastic modulus were performed on a universal mechanical testing system

Instron series 3360 model 3367 with 30kN (6,750 lbf) capacity.

Films were also produced by casting method and hot press method after

extrusion. These specimens were produced by using a die type D638-5-IMP-

ASTM. A strain rate of 2.5 mm/min was used. Film thickness was measured

with a micrometer at four random positions on each film specimen to an

accuracy of ± 0.02 mm. Tensile tests were carried out in an Instron series

3360, model 3367 with jaws of 2kN capacity.

3.12.4. Results and discussion

A significant improvement of the mechanical properties was observed in

samples loaded with clay even at the high load of this filler. Significant

increase in the tensile strength was also seen in samples of modified starches

prepared in presence of glycerol with respect to native starches. The

elongation at break decreased from native to modified starches in both

composites prepared with glycerol and glycerol/clay, but remained almost

equal from modified-starch/glycerol to modified-starch/clay/glycerol

composites, indicating a small effect in the rheology of these materials. This

effect can be also observed in the tensile strength increased which increased

substantially. By the other hand, the elongation at break increased form

native-starch/glycerol to native-starch/glycerol/clay composites influencing

152

notoriously the flow properties of native starches, effect which can be

attributed to the phase separation of the glycerol in presence of clay as

shown by DMTA analyses (Fig 71).

In native starches/clay/glycerol composites, the enhancement of the

properties of the composites suggests that the stiffness increases with the

silicate content the matrix. This is due to the well known

intercalated/exfoliated of the clay in the matrix of polymers. The increase in

the tensile strength is directly related with smaller particles which may allow

a good dispersion and can migrate throughout the system. The size and the

good dispersion of the clay will not restrict the migration and dispersion of

the glycerol. The reduction of strain in native starches in presence of clay has

been attributed to the agglomeration of the filler and therefore to the

restriction of the migration of the glycerol thus forming a separated phase

[126].

However, in the samples of modified starches, this phenomenon was not

observed. On the contrary an increase in the thermal stability was described

suggesting a good intercalation of clay-modified starch. Moreover, the

elongation at break did not increase after the incorporation of the clay

indicating the chemical reaction of the modified starch with the layers of clay.

A good dispersion of the clay may be due to the more available polar groups

found in modified starches.

It is interesting to note that at 40% glycerol modified starches showed a very

low elongation at break, but above this percentage this value increased

abruptly.

153

0510

15

20

25

30

35

40

MS-G-45%NS-G-45%

NS-G-40%MS-G-40%

NS-G-35%MS-G-35%

MS-G-C-40%NS-G-C-40%

Modulus GPa

Elongationat break (%)

Tensile Strength (MPa)

Figure 71- Tensile strength (MPa) and elongation at break (%) of modified and native starch glycerol and clay composites

3.12.5. Conclusions

An improvement in all material properties of the composites could be

achieved by better dispersion of the clay. Better dispersion can be achieved

by first mixing of starch and filler followed by plasticization. Although an

enhancement of mechanical properties takes place in the clay-filled

composites, and further increase of the properties by the fungal modification,

the water resistance is still too poor to use these composites in packaging

applications, at least for liquids. From a moisture sensitivity point of view,

well-ordered intercalated structures as well as the presence of hydrophobic

pendant groups are also helpful to lower the moisture sensitivity in

comparison to a structure consisting of individual micrometer or even

nanometer dispersion of layered silicate in a starch matrix.

154

4. General conclusions

The analysis of the structural properties of the fungal produced polymers

showed important variation with respect to the substrate. These properties in

overall influenced the functional properties of the polysaccharides.

The glass transition temperature (Tg) of starch films manufactured with

modified starches were investigated by mechanical dynamic analysis (DMA)

and differential scanning calorimeter (DSC). After the starch modification the

tan curve shown by DMA revealed two thermal transitions, phase changes

which were associated to the Tg of different polymers. The Tg by DSC

confirmed the findings of the DMA. The storage modulus of modified starches

(E’) was at least three-fold the strength of native starches. In addition, the

thermogravimetric analyses (TG) showed that complete degradation occurred

at ~630oC in native starches. However, after the modification, almost 20% of

the modified starches supported temperatures up to 1000oC. The second

derivative of the mass loss vs. temperature (2nd Deriv.) showed also two

thermal transitions in modified starches. X-ray diffraction analyses showed

that the crystalline regions of the starch granules remain intact after the

modification. Scanning electron micrographs (SEM) showed the physical

damage of the granules produced by the fungal enzymes as well as the

presence of inclusions blocking the granular’s surface porosity. Analyses by

solid state 13 CP/MAS NMR FTIR, and FT-Raman showed the substitution of

the branching points of the starch.

The various phase transitions (Tg) found in the DMA and DSC, therefore, may

correspond to variations in the molecular weight of the starch chains, as it

was shown by MALDI-TOF analyses. Degradation of the starch may occur

mainly in the amylose fractions. The analyses also showed the potential

presence of exopolysacchrides and protein-like compounds in the mass of

modified starches.

155

5. Future work

The chemical characterization of these materials and/or the description of the

biosynthetic pathway is still of great interest for the future improvement of

the process of production of these materials. In general applied and

fundamental work in the development of these biodegradable polymers for

packaging applications, production by extrusion and reactive extrusion,

injection molding applications, research to improve the processing and

binding properties of starch (starch-PVC, starch polyurethanes, etc., or

fillers). Also, of great importance is the effect of various nano-particles on

the functional properties of these materials.

156

6. References

[1] V. Siracusa, P. Rocculi, S. Romani, M. Rosa: Biodegradable polymers for food

packaging: a review. 2008, 19, 634-643. [2] L. Avérous, P. Halley: Biocomposites based on plasticized starch. Biofuels

Bioprod Bioref, 2009, 3, 329-343. [3] R. Narayan, Biodegradable plastics, in Opportunities For innovation in

Biotechnology, N. i. o. S. T. (NIST), Editor. 1993, National institute of Standards & Technology (NIST).

[4] R. Narayan, C. Pettigrew: ASTM standards help difine and grow a new biodegradable plastic industry. 1999, XII, 36-42.

[5] A. H. Tullo: From refuse to reuse. Chem Eng News 2007, 85, 15-20. [6] R. Septo: The Processing of Starch as a Thermoplastic. 2003, 201, 203-212. [7] A. J. Carvalho, A. A. Curvelo, A. Gandini: Surface modification of thermoplastic

starch: reactions with isocyanates, epoxy functions and steraroyl chloride. Ind Crops Prod, 2005, 21, 331-336.

[8] J. Guan, K. M. Eskridge, M. A. Hanna: Acetylated starch-polylactic acid loose-fill packaging materials. Ind. Crops Prod, 2005, 22, 109-123.

[9] Y. Lu, L. Tighzert, F. Berzin, S. Rondot: Innovative plasticized starch films modified with waterborne polyurethane from renewable resources. 2005, 61, 174-182.

[10] T. Qunyi, Z. Ganwei: Rapid synthesis of a superabsorbent from a saponified starch and acrylonitrile/AMPS graft compolymers. 2005, 62, 74-79.

[11] G. F. Fanta, F. C. Felker, R. L. Shogren: Graft polymerization of acrylonitrile onto spherocrystals formed from jet cooked cornstarch. Carbohydr Polym, 2004, 56, 77-84.

[12] M. Bengtsson, K. Koch, P. Gatenholm: Surface octanoylation of high-amylose potato starch films. Carbohydr Polym, 2003, 54, 1-11.

[13] L. Zhang, J. Gao, R. Tian, J. Yu, W. Wang: J Appl Polym Sci, 2003, 88, 146. [14] G. Chang, L. Kiho: Carbohydr Polym, 2002, 48, 125-130. [15] V. Athawale, V. Lele: Synthesis and characterization of graft copolymers of

maize starch and methacrylonitrile. Carbohydr Polym, 2000, 41, 407-416. [16] D. Rutot, P. Degee, R. Narayan, P. Dubois: 2000, 7, 215-225. [17] D. Peters: Carbohydrates for fermentation. 2006, 1, 806–814. [18] A. Hebeish, M. Beliakova, A. Bayazeed: Improved synthesis of poly(MAA)-

starch graft copolymers. Appl Polym Sci, 1998, 68, 1709-1715. [19] J. Gao, J. Yu, W. Wang, L. Chang, R. Tian: Graft copolymerization of starch-AN

initiated by potassium permanganate. J Appl Polym Sci, 1998, 68, 1965–1972. [20] J. Aburto, S. Thiebaud, I. Alric, E. Borredon, D. Bikiaris, J. Prinos, C.

Panayiotou: Properties of octanoated starch and its blends with polyethylene. Carbohdr Polym, 1997, 34, 101-112.

157

[21] S. Thiebaud, J. Aburto, I. Alric, E. Borredon, D. Bikiaris, J. Prinos, C. Panayiotou: Properties of fatty acid esters of starch and their blends with LDPE. 1997, 65, 705-721.

[22] G. F. Fanta, R. Shogren: Modification of starch-poly(methyl acrylate) graft copolymers by steam jet cooking. J Appl Polym Sci, 1997, 65, 1021-1029.

[23] S. Rath, R. Singh: Flocculation characteristics of grafted and ungrafted starch, amylose, and amylopectin. 1997, 66, 1721-1729.

[24] T. Maharana, B. Singh: 2006, 100, 3229. [25] Y. Sugahara, T. Ohta: 2001, 82, 1437-1343. [26] A. De Graaf, A. Karman, L. Janseen: Material Properties and Glass Transition

Temperatures of Different Thermoplastic Starches After Extrusion Processing. Starch/Stärke, 2003, 55, 80-86.

[27] R. Ellis, M. Cochrane, M. Dale, C. Duffus, A. Lynn, I. Morrison, R. Prentice, J. Swanton, S. Tiller: Starch production and industrial use. J Sci Food Agric 1998, 77, 289-311.

[28] N. Follain, C. Joly, P. Dole, C. Bliard: Mechanical properties of starch-based materials. I. short review and complementary experimental analysis,. J Appl Polym Sci, 2005, 97, 1783-1794.

[29] I. Danjaji, R. Nawang, U. Ishiaku, H. Ismail, Z. Ishak: J Appl Polym Sci, 2001, 79, 29-37.

[30] A. De Graaf, P. Jannssen: Properties and manufacturing of a new starch plastic. Polym Eng Sci, 2001, 41, 584-594.

[31] D. Garlotta, W. M. Doane, R. L. Shogren, J. W. Lawton, J. Willett: Mechanical and thermal properties of starch-filled poly(D-, L-, lactic acid)/poly(hydroxyl ester ether) biodedgradable blends. J Appl Polym Sci, 2003, 88, 1775-1786.

[32] W. Liu, Y. Wang, Z. Sun: 2004, 92, 484-. [33] T. Ohkita, S. Lee: 2005, 97, 1107-1114. [34] W. Ban, J. Song, S. Argyropoulos, L. Lucia: 2006, 100, 2542-2548. [35] D. Bielinkski, L. Pysklo, L. Slusorsky, G. Janowska, G. Levandowicz:

Preliminary studies on the surface layer of starch. Macromol Symp, 2003, 194, 233-239.

[36] J. Willett, V. Finkenstadt: Initiator effects in reactive extrusion of starch-polyacrylamide graft copolymers. J Appl PolymSci, 2006, 99, 52-58.

[37] S. Kalambur, S. Rizvi: Biodegradable and functionality superior starch-polyester nanocomposites from reactive exgtrusion. 2005, 96, 1072-1082.

[38] F. Xie, L. Yu, H. Liu, L. Chen: Starch Modification Using Reactive Extrusion. Starch/Stärke, 2006, 58, 131-139.

[39] J. M. Luengo, B. García, A. Sandoval, G. Naharro, E. R. Olivera: Bioplastics from microorganisms. 2003, 6, 251-260.

[40] K. Sudesh, T. Iwata: Sustainability of Biobased and Biodegradable Plastics. 2008, 36, 433-442.

[41] A. Rajan, T. Abraham: Enzymatic modification of cassava starch by bacterial lipase. 2006, 29, 65-71.

[42] R. Jeng, C. Huang, M. Sain, M. Hubbes, A. Rodriguez, A. Saville: Starch-like exopolysaccharides produced by the filamentous fungi ophiostoma ulmi and O. novo-ulmi. 2007, 37, 80-95.

158

[43] C. Huang, R. Jeng, M. Sain, B. Saville, M. Hubbes: Production, characterization, and mechanical properties of starch modified by Ophiostoma spp. Bioresources, 2006, 1, 234-239.

[44] M. Sain, R. Jeng: Disclosure documents, Biocap-NSERC Report. Laboratory of Pathology, Faculty of Forestry, University of Toronto, Canada. 2006, 1.

[45] M. Sain, R. Jeng, M. Hubbes: US Patent Appl. No. 11/764683. 2007. [46] A. Buléon, P. Colonna, V. Planchot, S. Ball: Starch granules: structure and

biosynthesis. Int J Biol Macromol 1998, 23, 85-112. [47] J. Szymonska, F. Krok: Potato starch granule nanostructure studied by high

resolution non-contact AFM. Int. J. Biol. Macromol 2003, 33, 1–7. [48] T. A. Waigh, K. L. Kato, A. M. Donald, M. J. Gidley, C. J. Clarke, C. Riekel:

Side-Chain Liquid-Crystalline Model for Starch. Starch/Stärke, 2000, 52, 450–460.

[49] T. A. Waigh, A. M. Donald, F. Heidelbach, C. Riekel, M. J. Gidley: Analysis of the Native Structure of Starch Granules with Small Angle X-Ray Microfocus Scattering. Biopolymers 1999, 49, 91–105.

[50] M. J. Ridout, A. P. Gunning, M. L. Parker, R. H. Wilson, V. J. Morris: Using AFM to image the internal structure of starch granules. 2002, 50, 123-132.

[51] M. J. Ridout, M. L. Parker, C. L. Hedley, T. Y. Bogracheva, V. J. Morris: Atomic force microscopy of pea starch granules: granule architecture of wild-type parent, r and rb single mutants, and the rrb double mutant. 2003, 338, 2135 - 2147.

[52] F. Krok, J. Szymonska, P. Tomasik, M. Szymonski: Non-contact AFM investigation of influence of freezing process on the surface structure of potato starch granule. Appl Surf Sci, 2000, 157, 382-386.

[53] T. Oostergetel, J. van Bruggen: The crystalline domains in potato starch granules are arranged in a helical fashion. 1993, 21, 7-12.

[54] K. Gotlieb, A. Capelle: Starch derivatization: Fascinating and unique industrial opportunities. Wageningen Academic Publishers The Netherlands 2005158.

[55] R. F. Tester, J. Karkalas, X. Qi: Sarch-composition, fine structure and architecture. 2004, 39, 151-165.

[56] P. M. Baldwin: Starch Granule-Associated Proteins and Polypeptides: A Review. Starch - Stärke 2001, 53, 475-503.

[57] G. D. Haki, S. K. Rakshit: Developments in industrially important thermostable enzymes: a review. Bioresour Technol, 2003, 89, 17–34.

[58] S. O'Brien, Y. Wang: Susceptibility of annealed starches to hydrolysis by a-amylase and glucoamylase. 2008, 72, 597-607.

[59] H. Li, Z. Chi, X. Duan, L. Wang, J. Sheng, L. Wu: Glucoamylase production by the marine yeast Aureobasidium pullulans N13d and hydrolysis of potato starch granules by the enzyme. Process Biochem, 2007, 42, 462-465.

[60] V. Sharma, K. Rausch, M. Tumbleson, V. Singh: Comparison Between Granular Starch Hydrolyzing Enzyme and Conventional Enzymes for Ethanol Production from Maize Starch with Different Amylose:Amylpectin Ratios. 2007, 59, 549-556.

[61] I. Skwortsov, V. Ignatov: Extracellular polysaccharides and polysaccharide-containing biopolymers from Azospirillum species: properties and the possible role in interaction with plant roots,. 1998, 165, 223-229.

159

[62] V. Moriello, L. Lama, A. Poli, C. Gugliandolo, T. Maugeri, A. Gambacorta, B. Nicolaus: Production of exopolysaccharides from a thermophilic microorganism isolated from a marine hot spring in flegrean areas. 2003, 30, 95-101.

[63] M. M. Girard, C. Schaffer-Lequart: Gelation and resistance to shearing of fermented milk: role of exopolysaccharides, International Dairy Journal. 2007, 17, 666-673.

[64] R. De Philippis, M. Vincenzini: Exocellular polysaccharides from cyanobacteria and their possible applications. FEMS Reviews, 1998, 22, 151-175.

[65] L. Harding, V. Marshall, Y. Hernandez, Y. Gu, M. Maqsood, N. Mclay, A. Laws: Structural characterization of a highly branched exopolysaccharide produced by Lactobacillus delbrueckii subsp. bulgaricus NCFB2074. Carbohydr. Res, 2005, 340, 1170-1111.

[66] M. Girard, C. Schaffer-Lequart: Gelation and resistance to shearing of fermented milk: role of exopolysaccharides. Int Dairy J, 2007, 17, 666-673.

[67] Z. Chi, C. Su, W. Lu: A new exopolysaccharide produced by marine Cyanothece sp. 113. Bioresour Technol, 2007, 96, 1329-1332.

[68] C. Lamelas, M. Benedetti, K. Wilkinson, V. Slaveykova: Characterization of H+ and Cd+ binding properties of the bacterial exopolysaccharides. Chemosphere, 2006, 65, 1362-1370.

[69] A. Kumar, K. Mody, B. Jha: Bacterial exopolysaccharides- a perception,. J Basic Microbiol, 2007, 47, 103-117.

[70] J. Luengo, B. Garcia, A. Sandoval, G. Naharro, E. Olivera: Bioplastics from microorganisms. 2004, 6, 251-260.

[71] J. Guezennec: Deep-sea hydrothermal vents: A new source of innovative bacterial exopolysaccharides of biotechnological interest? J Ind Microbiol, 2002, 29, 204-208.

[72] H. Ruijssenaars, F. Stingele, S. Hartmans: Biodegradability of Food-Associated Extracellular Polysaccharides. 2000, 40, 194-199.

[73] B. Nicolaus, V. Moriello, L. Lama, A. Poli, A. Gambacorta: Polysaccharides from extremophilic microorganims. 2003, 34, 159-169.

[74] A. Laws, Y. Gu, V. Marshall: Biosynthesis, characterization, and design of bacterial exopolysaccharides from lactic acid bacteria. Biotechnol Adv, 2001, 19, 597-625.

[75] F. Rosado, S. Germano, E. Carbonero, S. Da Costa, M. Lacomini, C. Kemmelmeier: Biomass and exopolysaccharides producion in submerged cultures of Pleurotus ostreatoroseus Sing. and Pleurotus ostreatus Florida (Jack.:Fr.) Krumer. 2003, 43, 230-237.

[76] L. Selbmann, F. Stingele, M. Petruccioli: Exopolysaccharide production by filamentous fungi: the example of Botryosphaeria rhodina. 2003, 84, 135-145.

[77] E. Lin, S. Sung: Cultivation conditions influence exopolysaccharide production by the edible Basidiomycete Antrodia cinnamomea in submerged culture. Int J Food Microbiol, 2006, 108, 182-187.

[78] K. Przyby, H. Dahm, A. Ciesielska, K. Molinski: Cellulolytic activity and virulence of Ophiostoma ulmi and O. novo-ulmi isolates. 2006, 36, 58-67.

160

[79] T. Binz, G. Canevascini: Xylanases from the Dutch elm disease pathogens Ophiostoma ulmi and Ophiostoma novo-ulmi. Physiol Mol Plant Pathol, 1996, 49, 159-175.

[80] M. Polakivic, J. Bryjak: Modelling of potato starch saccharification by an Aspergilus niger glucoamylase. 2004, 18, 57-63.

[81] L. Huang, B. Jin, P. Lant, J. Zhou: Simultaneous saccharification and fermentation of potato starch wastewater to lactic acid by Rhizopus oryzae and Rhizopus arrhizus. Biochem Eng J, 2005, 23, 265-276.

[82] S. O'Brien, Y. Wang: Susceptibility of annealed starches to hydrolysis by -amylase and glucoamylase. 2007, 72, 597-607.

[83] K. Kasemwong, K. Plyachomkwan, R. Wansuksri, K. Sriroth: Granule Sizes of Canna (Canna edulis) Starches and their Reactivity Toward Hydration. Starch/Starke, 2008, 60, 624-633.

[84] H. S. Kim, C. H. Kerry: Channels within soft wheat starch A- and B-type granules. 2008, 48, 159-172.

[85] H. M. Wilhelm, M. R. Sierakiwski, G. P. Souza, F. Wypych: Starch films reinforced with mineral clay. Carbohydr Polym, 2003, 52, 101-110.

[86] M. Cerna, A. Barros, A. Nunes, S. Rocha, I. Delgadillo, J. Copikova, M. Coimbra: Use of FT-IR spectroscopy as a tool for the analysis of polysaccharide food additives. Carbohydr Polym, 2003, 51, 383-389.

[87] R. Wilson, M. Kalichevsky, S. Ring, P. Belton: A Fourier-transform infrared study of the gelation and retrogradation of waxy-maize starch. Carbohydr Res, 1987, 166, 162-165.

[88] E. Mills, M. Parker, N. Wellner, G. Toole, K. Feeney, P. Shewry: Chemical imaging: the distribution of ions and molecules in developing and mature wheat grain. 2005, 41, 193-201.

[89] J. Pandey, R. Singh: Green nanocomposites form renewable resources: effect of plasticizer on the structure and material properties of clay-filled starch. 2005, 57, 8-15.

[90] M. Lutfor, S. Sidik, W. Yunus, M. Ab-Rhaman, A. Mansor, H. Jelas: Preparation and swelling of polymeric absorbent constaining hydroxamic acid group form polymer grafted starch. 2001, 45, 96-100.

[91] J. Martin, A. Solla, M. Coimbra, L. Gil: Metabolic distiction of Ulmus minor xylem tissues after inoculation with Ophiostoma novo-ulmi. 2005, 66, 2458-2467.

[92] J. Fang, P. Fowler, J. Tomkinson, C. Hill: The preparation and characterization of a series of chemically modified potato starches. Carbohydr Polym, 2002, 47, 245-252.

[93] L. Thygesen, M. Lokkey, E. Micklander, S. Engelsen: Vibrational microscopy of food. Raman vs. FT-IR. Trends in Food Sci Technol, 2003, 14, 50-57.

[94] M. Xiaofei, Y. Jiugao, F. Jin: Urea and formamide as a mixed plasticizer for thermoplastic starch. Polym Int, 2004, 53, 1780-1785.

[95] M. Huang, J. Yu, X. Ma: Studies on the properties of montomorillonite-reiforced thermoplastic starch composites. Polymer 2004, 45, 7017-7023.

[96] J. Cael, J. Koenig, J. Blackwell: J Carbohydr Res, 1973, 29, 123-127.

161

[97] P. M. Fechner, S. Wartewig, A. Kiesow, A. Heilmann, P. Kleinebudde, R. H. H. Neubert: Influence of Water on Molecular and Morphological Structure of Various Starches and Starch Derivatives. Starch/Stärke, 2005, 57, 605–615.

[98] C. G. Mothe, M. G. Machado, M. I. Tavares: Solid state 13C-NMR and study of the subproducts obtained from corn industry. 2002, 84, 1680-1685.

[99] M. Tavares, A. Bathista, E. Silva, P. Costa, N. Filho, J. Nogueira: 13C-NMR study of Dipteryx Alata Vogel Starch. 2004, 92, 2151-2154.

[100] K. Matsuzaki, U. Toshiyuki, A. Tetsuo: NMR Spectroscopy and Stereoregularity of Polymers. 1996 275.

[101] S. Broberg, K. Koch, R. Andersson, L. Kenne: A comparison between MALDI-TOF mass spectrometry and HPAEC-PAD analysis of debranched starch. Carbohydr Polym, 2000, 43, 285-289.

[102] M. Mohr, K. Bornse, H. Widmer: Matrix assisted laser desorption ionization mass spectrometry: improved matrix for oligosaccharides. 1995, 9, 809-814.

[103] C. Corradini, F. Bianchi, D. Matteuzzi, A. Amoretti, M. Rossi, S. Zanoni: High-performance anion-exchange chromatography coupled with pulsed amperometric detection and capillary zone electrophoresis with indirect

ultra violet detection as powerful tools to evaluate prebiotic properties of fructooligosaccharides and inulin. J Chromatogr A, 2004, 1054, 165–173.

[104] K. Koch, R. Andersson, P. Aman: Quantitative analysis of amylopectin unit chains by means of highperformance anion-exchange chromatography with pulsed amperometric detection. J Chromatogr A, 1998, 800, 199–206.

[105] J. Sanderson, R. Daniels, A. Donald, A. Blennow, S. Engelsen: Exploratory SAXS and HPAEC-PAD studies of starches from diverse plant genotypes. 2004, 64, 433-443.

[106] I. Hanashiro, A. J., S. Hizukuri: A periodic distribution of the chain lenght of amylopectins as reveled by high-performance anion-exchange chromatography. Bioresour Technol, 1996, 283, 151-159.

[107] R. Cataldi, I. Thomas: Carbohydrate analysis by high-performance anion-exchange chromatography with pulsed amperometric detection. J Anal Chem, 2000, 368, 739-758.

[108] J. Sanderson, R. Daniels, A. Donald, A. Blennow, S. Engelsen: Exploratory SAXS and HPAEC-PAD studies of starches form diverse plant genotypes. 2006, 64, 433-443.

[109] J. Jane, Y. Chen, L. Lee, A. McPherson, K. Wong, M. Radosavljevic, T. Kasemsuwan: Effects of Amylopectin Branch Chain Length and Amylose Content on the Gelatinization and Pasting Properties of Starch. Cereal Chem, 1999, 76, 629-637.

[110] K. Wong, J. Jane: Quantitative analysis of debranched amylopectin by HPAEC-PAD wiht a postcolumn enzyme reactor. J Liq Chromatogr Relat Technol, 1997, 20, 297-310.

[111] H. Menard: Dynamic mechanical analysis: a practical introduction. 2008, USA, 202p.

[112] Editor, Atlas of stress strain curves. 2002, Ohio: ASM International. 821. [113] F. Xie, L. Yu, L. Chen, L. Li: A new study of starch gelatinization under stress

using dynamic mechanical analysis. Carbohydr Polym, 2008, 72, 229-234.

162

[114] L. Averous, C. Fringant, L. Moro: Starch-Based Biodegradable Materials Suitable for Thermoforming Packaging. Starch/Starke 2001, 53, 368-371.

[115] R. Septo: The Processing of Starch as a Thermoplastic. 2003, 201, 203-212. [116] S. Gaudin, D. Lourdin, D. Le botlan, L. Ilari, P. Colonna: Plasticization and

mobility in starch-sorbitol films. J Cereal Sci, 1999, 29, 273-284. [117] P. Myllarinen, R. Partanen, J. Seppala, P. Forssell: Effect of glycerol on behavior

of amylose and amylopectin films. 2002, 50, 355-361. [118] X. Tang, S. Alavi, T. Herald: Effects of plasticizers on the structure and properties

of starch-clay nanocomposites films. 2008, 74, 552-558. [119] D. Wood, B. Chiou, M. Glenn, D. Hoffmann, T. Yee, S. Iman, W. Orts: Fine

Structure of Starch-Clay Composites as Biopolymers, Microscopy and Microanalysis. Microscopy Micoanal 2008, 14, 1500-1503.

[120] B. S. Chiou, E. Yee, G. M. Glenn, W. J. Orts: Rheology of starch-clay nanocomposites. Carbohydr Polym, 2005, 59, 467-475.

[121] L. Vertuccio, G. Gorrasi, A. Sorrrentino, V. Vittoria: Nano-clay reinforced PCL/Starch blends obtained by high energy ball milling. Carbohydr Polym, 2009, 75, 172-179.

[122] N. Lilichenko, R. Maksimov, J. Zicans, R. Meri, E. Plume: A biodegradable polymer nanocomposite: mechanical and barrier properties. Mech Compos Mat 2008, 44, 45-56.

[123] Q. Zhang, Z. Yu, X. Xie, K. Naito, Y. Kagawa: Preparation and crystalline morphology of biodegradable starch/clay nanocomposites. Polym Degrad Stab, 2007, 48, 7193-7200.

[124] P. Kampeerapappun, D. Aht-ong, D. Pentrakoon: Preparation of cassava starch/montmorillonite composite film. Carbohdr Polym, 2007, 67, 155-163.

[125] J. Chen, J. Evans: Thermoplastic starch-clay nanocomposites and their characteristics. Carbohydr Polym, 2005, 61, 455-463.

[126] K. Dean, D. L. Yu, Y. Wu: Preparation and characterization of melt-extruded thermoplastic starch/clay nanocomposites. Compos Sci. Technol, 2007, 67, 413-421.

[127] K. Dean, M. Do, E. Petinakis, L. Yu: Key interactions in biodegradable thermoplastic starch/poly(vinyl) alcohol/montmorillonite micro- and nanocomposites. Compos Technol, 2008, 68, 1453-1462.

[128] D. Chaudhary: Understanding amylose crystallinity in starch-clay nanocomposites. J Polym Sci Part B: Polym Phys, 2008, 46, 979-987.

[129] J. Yu, X. Wang, P. Wu: Effect of glycerol on water vapor sorption and mechanical properties of starch/clay composite Films. Starch/Stärke 2008, 60, 257-262.

[130] N. Wang, X. Zhang, N. Han, S. Bai: Effect of citric acid and processing on the performance of thermoplastic starch/montmorillonite nanocomposites. Carbohydr Polym, 2009, 76, 68-73.

[131] E. Rudnik, G. Matuschek, N. Milanov, A. Kettrup: Thermal stability and degradation of starch derivatives. 2006, 85, 267-270.

[132] P. Aggarwal, D. Dollimore: A comparative study of the degradation of different starches using thermal analysis. Talanta 1996, 43, 1527-1530.

163

[133] D. Schlemmer, E. De-Oliveira, M. Sales: Polyestyrene/thermoplastic starch blends with different plasticizers. 2007, 87, 635-638..

[134] J. Bhatty, D. Dollimore, J. Dunn, W. Irwin, M. Ish-Shalom, J. Jordan, M. Reading, J. Sharp, J. Sthal: Treatease on analytical chemistry, Part I, Thermal Methods. 1993, 13, 406.

[135] L. Yu, L. Christie: Measurement of starch thermal transitions using differential scanning calorimetry. Carbohdr Polym. 2001, 46, 179-183.

[136] V. Derycke, G. Vandeputte, R. Vermeylen, W. De Man, B. Goderis, M. Koch, J. Delcour: Starch gelatinization and amylose-lipid interactions during rice parboiling investigated by temperature resolved wide angle X-ray scattering and differential scanning calorimetry. J Cereal Sci, 2005, 42, 334-337.

[137] L. Wasserman, M. Signorelli, A. Schiraldi, V. Yuryev, G. Boggini, S. Bertini: Preparation of wheat resistant starch. J Therm Anal Calorim, 2007, 87, 153-157.

[138] S. Puncha-arnon, W. Pathipanawat, C. Puttanlek, V. Rungsardthong, D. Uttapap: Effects of relative granule size and gelatinization temperature on paste and gel properties of starch blends. 2008, 41, 552-557.

[139] T. Tran, K. Piyachomkwan, K. Sriroth: Gelatinization and Thermal Properties of Modified Cassava Starches. Starch/Stärke, 2007, 59, 46–55.

[140] A. Hagenimana, P. Pu, X. Ding: 2005, 38, 257. [141] S. Berland, P. Relkin, B. Launay: 2003, 71, 311-316. [142] L. Elfstrand, A. Eliasson, M. Jonsson, M. Reslow, M. Wahlgren: From Starch to

Starch Microspheres: Factors Controlling the Microspheres Quality. Starch/Stärke, 2006, 58, 381-386.

[143] K. Poutanen, P. Forssell: Modification of Starch Properties wiht Plasticizers. 1996, 4, 128-132.

[144] A. Sandoval, J. Barreiro: Off-line rheometry of corn starch: Effects of temperature, moisture content and shear rate. 2007, 40, 43-48.

[145] O. Martin, L. Everous, G. Valle: In-line determination of plasticized wheat starch viscoelastic behavior impact of processing. 2003, 53, 169-182.

[146] W. Aichholzer, H. Fritz: Rheological Characterization of Thermoplastic Starch Materials. Starch/Stärke 1998, 50, 77-83.

[147] S. Kalambur, S. Rizvi: Rheological Behavior of Starch-Polycrapolactone (PCL) Nanocomposite Melts Synthesized by Reactive Extrusion. Polym Eng Sci, 2006650-658.

[148] D. Lourdin, H. Bizot, P. Colonna: "Antiplasticization" in starch-glycerol films? J Appl Polym Sci 1998, 63, 1047-1053.

[149] J. W. Lawton: Effect of starch type on the properties of starch containing films. Carbohydr Polym, 1996, 29, 203-208.

[150] A. Dias, M. Suzana, B. Adelaide, M. Grossmann: Effect of glycerol and amylose enrichment on cassava starch film properties. J Food Eng, 2007, 78, 941–946.

[151] L. Godbillot, P. Dole, C. Joly, B. Roge, M. Mathlouthi: Analysis of water binding in starch plasticized films. 2006, 96, 380–386.

[152] N. Laohakunjit, A. Noomhorm: Effect of Plasticizers on Mechanical and Barrier Properties of Rice Starch Film. Starch/Stärke, 2004, 56, 348–356.

[153] O. Lopez, M. Garcia, N. Zaritzky: Film forming capacity of chemically modified corn starches. 2008, 73, 573–581.

164

165

[154] D. Thirathumthavorn, S. Charoenrein: Aging Effects on Sorbitol- and Non-Crystallizing Sorbitol-Plasticized Tapioca Starch Films. 2007, 59, 493- 497.

[155] B. Montaño, P. Torres, B. Ramírez, M. Plascencia, F. Brown: Physical and Mechanical Properties of Durum Wheat (<B><I>Triticum durum</I></B>) Starch Films Prepared with A- and B-type Granules. Starch - Stärke 2008, 60, 559-567.

[156] M. Thunwal, V. Kuthanova, A. Boldizar, M. Rigdahl: Film blowing of thermoplastic starch. 2008, 71, 583– 590.

[157] M. A. Bertuzzi, M. Armada, J. C. Gottifredi: Physicochemical characterization of starch based films. J Food Eng, 2007, 82, 17–25.

[158] A. Jansson, F. Thuvander: Influence of thickness on the mechanical properties for starch films. Carbohydr Polym, 2004, 56, 499–503.

[159] H. Pushpadass, D. Marx, H. Mildford: Effects of Extrusion Temperature and Plasticizers on the Physical and Functional Properties of Starch Films. Starch - Stärke 2008, 60, 527-538.