Use of Z-stack imaging to quantify the phase behaviour of biomaterial composites in relation to theoretical predictions of blending laws from rheological measurements
A thesis submitted in fulfillment of the requirements for the degree of Doctor of Philosophy
Pranita Mhaske
B. Tech. (Food Science and Technology), MIT College of Food Technology, Pune, India
School of Science
Science, Technology, Engineering and Mathematics College
RMIT University
March 2021
Declaration
I certify that except where due acknowledgement has been made, the work is that of the
author alone, and the work has not been submitted previously, in whole or in part, to qualify
for any other academic award. The content of the thesis is the result of work which has been
carried out since the official commencement date of the approved research program at RMIT
University; and any editorial work, paid or unpaid, carried out by a third party has been
acknowledged; and ethics procedures and guidelines have been followed.
Pranita Mhaske
i
23rd March 2021
Copyright
Pranita Mhaske
2020
Copyright Notices
Notice 1
The 1968 Copyright Act, mandates that this thesis be used only under regular conditions of
scholarly fair dealing. In particular, no results or conclusion should be extracted from it, nor
should it be copied or closely paraphrased in whole or in part without the written consent of
the author. Proper written acknowledgement should be made for any assistance obtained from
this thesis.
Notice 2
I certify that I have made all reasonable efforts to secure copyright permissions for third-party
content included in this thesis and have not knowingly added copyright content to my work
ii
without the owner’s permission.
ACKNOWLEDGEMENTS
I owe my sincere gratitude to my senior supervisor, Prof. Stefan Kasapis for his support
and supervision throughout my PhD candidature. His guidance, motivation and scientific
insight has enabled me to develop a thorough understanding of the subject and meet his high
publication standards.
I would also like to express my gratitude to my co-supervisor, A/Prof. Asgar Farahnaky,
for his patient and kind guidance, and critical outlook that helped me be on track during my
candidature and for providing scientific exposure and opportunities that were beyond the scope
of the candidature. Working with him has impressed and inspired me on both the professional
and personal front.
A word of gratitude also goes to Lita Katopo, for patiently guiding, supporting, training
and motivating me during the early days of my candidature, without which I probably wouldn’t
have been able to see this through.
My research would not have reached its successful completion without the support of
the technical staff at RMIT University. I am thankful to Sanaz Salehi, Yan Chen, Spiros
Tsaroumis, Hadi Ranjiburachaloo, Jake Torrisi, Badyn Cook, James Lauer, Chaitali Dekiwadia
and especially Mina Dokouhaki for providing more than just technical assistance throughout
the years.
A big shoutout to my colleagues, associates and friends at the Food Science Laboratory
– Manisha Singh, Billy Lo, Lloyd Condict, Shahla Teimouri, Felicity Whitehead, Courtney
Morrish, Diah Ikasari, Kourosh Abdollahi, Cameron Ince, Carine Semasaka, Jasmeet Singh,
Geethu Kurup, Pham Loc, Arissara Phosanam, Lili Mao, Mayumi Silva, Nelum Pematilleke
and Mithila Jayasundera for their support and company which made the time spent working in
the labs fun and memorable.
Last but certainly not the least, I am grateful to my family for their continual love and
iii
support, to which I owe everything.
TABLE OF CONTENTS
Page i ii iii iv ix xiv xv xvii xx 1 Summary
Title Declaration Copyright Acknowledgements Table of contents List of Figures List of Tables List of abbreviations List of units and symbols Explanatory notes Chapter 1 Background and Literature Review
1.1. Introduction 1.2. Phase behaviour in mixed biopolymer systems 1.3. Phase diagrams of protein-polysaccharide-water systems 1.4. Techniques used to probe phase behaviour
1.4.1. Visual observation and phase volume ratio method 1.4.2. Refractometry 1.4.3. Chemical analysis for concentration estimation 1.4.4. Viscometry 1.4.5. Spectroscopy 1.4.5.1. UV-Vis spectroscopy
1.4.5.2. FTIR spectroscopy
iv
1.4.5.3. Raman spectroscopy 1.4.5.4. NMR spectroscopy 1.4.5.5. Ultrasonic spectroscopy 1.4.5.6. Diffusing wave spectroscopy 1.4.5.7. Electron spin resonance spectroscopy 1.4.6. Diffraction and scattering methods 1.4.6.1. Light scattering 1.4.6.2. X-Ray diffraction 1.4.6.3. SAXS and SANS 1.4.7. Calorimetry 1.4.8. Rheology 1.4.8.1. Mechanical properties of the composite 1.4.8.2. Interfacial tension between phases 1.4.9. Microscopy 1.4.9.1. Optical microscopy 1.4.9.2. Phase contrast microscopy 5 8 10 13 13 18 20 23 26 26 27 30 31 32 34 34 35 36 37 38 40 43 43 47 50 51 52
1.4.9.3. Atomic force microscopy 1.4.9.4. Fluorescent microscopy 1.4.9.5. Electron microscopy
1.5. Significance of the Research 1.6. Research questions 1.7. Research objectives
References
Chapter 2 Materials and Methods
Abstract 2.1. Materials
laser scanning 56 57 58 59 61 61 62 83 84 84 84 85 85 86 86 87 88 88 88 89 89 89 2.1.1. Biomaterials 2.1.1.1. Agarose 2.1.1.2. Gelatin 2.1.1.3. Lipids 2.1.1.4. Microcrystalline cellulose 2.1.2. Chemicals and Dyes 2.1.3. Instruments 2.1.4. Sample preparation
2.1.4.1. Agarose-canola oil composites 2.1.4.2. Agarose-ghee composites 2.1.4.3. Agarose-MCC composites 2.1.4.4. Agarose-gelatin composites 2.1.4.5. Staining for confocal microscopy
2.2. Methods
2.2.2.1. Micro Differential Scanning Calorimetry 2.2.2.2. Modulated Differential Scanning Calorimetry
2.2.1. Fourier Transform Infrared (FTIR) Spectroscopy 2.2.2. Differential Scanning Calorimetry 2.2.3. Rheology 2.2.4. Scanning Electron Microscopy 2.2.5. Confocal Laser Scanning Microscopy 2.2.6. Image Analysis References
90 90 94 97 98 99 107 110 115 116
in comparison to blending
v
123 123 Chapter 3 Quantitative analysis of the phase volume of agarose- canola oil gels law predictions using 3D imaging based on confocal laser scanning microscopy Abstract 31. Introduction
3.2. Materials and methods
3.2.1. Materials 3.2.2. Methods 3.2.2.1. Sample preparation 3.2.2.2. SEM analysis 3.2.2.3. Fourier transform infrared (FTIR) spectroscopy 3.2.2.4. Differential scanning calorimetry 3.2.2.5. Rheological measurements 3.2.2.6. Confocal laser scanning microscopy (CLSM) 3.2.2.6.1. Sample preparation and image acquisition 3.2.2.6.2. Image analysis 3.2.2.7. Statistical analysis
3.3. Results and discussion
the structural 124 124 124 124 124 124 124 124 124 124 125 125 125 125 3.3.1. Experimental observations on
characteristics of agarose-canola oil mixtures
127 3.3.2. Theoretical modelling of the phase behaviour in
agarose-canola oil composites
128 3.3.3. Development of a CLSM protocol and image
analysis for agarose-canola oil composites
3.4. Conclusions References
131 131
Chapter 4 Phase volume quantification of agarose-ghee gels using 3D confocal laser scanning microscopy and blending law analysis: A comparison Abstract 4.1. Introduction 4.2. Materials and methods
4.2.3.1. Scanning electron microscopy
4.2.3.2. Fourier transform infrared spectroscopy (FTIR)
4.2.3.3. Differential scanning calorimetry 4.2.3.4. Rheological measurements 4.2.3.5. Confocal laser scanning microscopy (CLSM) 4.2.3.6. Image analysis 4.2.3.7. Statistical analysis 4.2.1. Materials 4.2.2. Methods 4.2.3. Experimental analysis
4.3. Results and discussion
134 134 135 135 135 135 135 135 135 135 135 135 136 136 136 4.3.1. Investigation of the composite microstructure and its
vi
structural characteristics 4.3.1.1. Composite microstructure 4.3.1.2. Chemical properties 136 136
structure 136 137
138
139
4.3.1.3. Thermal characterization 4.3.1.4. Rheological observations of formation 4.3.2. Theoretical modelling of mechanical functions in support of phase topology 4.3.3. Z-stack CLSM imaging and image analysis of agarose-ghee blends
4.4. Conclusions References
141 141 Chapter 5 3D Confocal Laser Scanning Microscopy
for Quantification of the Phase Behaviour in Agarose- MCC co-gels in Comparison to the Rheological Blending-law Analysis Abstract 5.1. Introduction 5.2. Materials and methods
5.2.2.4. Rheological measurements
144 144 145 145 145 145 145 145 145 146 5.2.1. Materials and Sample Preparation 5.2.2. Experimental analysis 5.2.2.1. SEM Analysis 5.2.2.2. Fourier transform infrared spectroscopy (FTIR) 5.2.2.3. Differential scanning calorimetry
5.2.2.5. Confocal laser scanning microscopy (CLSM) and Image Analysis
5.2.2.6. Statistical analysis
5.3. Results and discussion
146 146 146 5.3.1. Structural Characterisation of the Agarose/MCC
Mixtures
148 5.3.2. Rheological Modelling of the Phase Behaviour in
Agarose/MCC Composite Gels
149 5.3.3. 3D CLSM imaging and image analysis for Phase
Volume Estimations in Agarose/MCC Mixture
5.4. Conclusions References
151 151
scanning microscopy confocal laser
vii
Chapter 6 Comparison of rheological blending-law analysis and for 3D quantification of the phase behaviour in agarose- gelatin co-gels Abstract 153
6.1. Introduction 6.2. Materials and methods
6.2.2.1. Sample preparation 6.2.2.2. Fourier transform infrared spectroscopy (FTIR) 6.2.2.3. Differential scanning calorimetry 6.2.2.4. Rheological measurements
6.2.1. Materials 6.2.2. Methods 153 155 155 155 155 156 156 156 157
6.2.2.5. Confocal laser scanning microscopy (CLSM) and Image Analysis
6.2.2.6.Statistical analysis
6.3. Results and discussion
157 158 158 6.3.1. Structural Characterisation of agarose-gelatin
164
171
mixtures 6.3.2. Theoretical modelling of the Phase Behaviour in Agarose-gelatin Co-Gels using blending law analysis 6.3.3. 3D CLSM imaging and image analysis of agarose- gelatin co-gels
6.4. Conclusions References
175 176 Chapter 7 Conclusions and Future Research
181 7.1. Conceptual basis of the present study
viii
182 185 190 193 7.2. Conclusions 7.3. Future Research References Appendix
List of Figures
Chapter 1
8 Figure 1.1 Scale of measurement of the various techniques used to probe phase
behaviour
11 16 Figure 1.2 Phase diagram in an aqueous biopolymer mixture Figure 1.3 Phase diagram of 3% gelatin/0,9% iota-carrageenan as a function of
sodium chloride concentration and pH
42
Figure 1.4 DSC exotherms for a) 2% agarose, b) 15% gelatin, c) 2% agarose plus 15% gelatin, d) Dalda vanaspati lipid, e) soyabean oil, and f0 2% agarose plus 15% gelatin plus 2.5% Dalda vanaspasti lipid. Scan rate 1°C/min
50
Figure 1.5 A drop of caseinate in alginate rich phase at rest (1), start-up transient of deformation under steady shear (2,3), steady state shape (4), and relaxation after shear is stopped (5,6). Graph represents the deformation parameter as a function of shear rate in comparison with theory
55
Figure 1.6 Kappa carrageenan (A, B) and Iota carrageenan (C, D) in mixture with meat proteins at pH 5.6 and 7.1, respectively. Black arrows indicate carrageenan molecules with attached meat proteins.
Chapter 2
84 85 91 93 94 95 Figure 2.1 Chemical structure of an agarose polymer Figure 2.2 Chemical structure of gelatin Figure 2.3 Schematic Diagram of FTIR Figure 2.4 FTIR- ATR Spectrum 100 at RMIT University Figure 2.5 Schematic diagram of a heat flux type DSC Figure 2.6 Typical material transition as a function of temperature change in a
DSC
97 98 99 102
Figure 2.7 The heat flux micro DSC with a cylinder type measuring system Figure 2.8 Micro DSC Setaram VII at RMIT University Figure 2.9 Modulated DSC Q2000 from TA instruments at RMIT University Figure 2.10 Basic shear rate versus shear stress behaviour of Newtonian, shear thinning, shear-thickening, and Bingham and Herschel-Bulkley materials
104
Figure 2.11 The principle of oscillation viscometry. Applied strain versus time (a) and resultant stress versus time that is measured in an elastic solid (b), Newtonian liquid (c) and viscoelastic liquid (d).
ix
106 108 110 Figure 2.12 Discovery Hybrid Rheometer by TA Instruments at RMIT University Figure 2.13 Schematic diagram of a SEM Figure 2.14 FEI Quanta 200 ESEM from FEI Corporate at RMIT University
112 Figure 2.15
Illustration of the confocal imaging principle (solid lines = in-focus light; dashed lines = out-of-focus light)
Chapter 3
126
Figure 3.1 SEM images of (a) 1% (w/w) agarose at 50x magnification, (b) 1% (w/w) agarose at 400x magnification, (c) 1% (w/w) agarose with 3% (w/w) canola oil at 400x magnification, and (d) 1% (w/w) agarose with 11% (w/w) canola oil at 400x magnification.
126
Figure 3.2 FTIR spectra of 1% (w/w) agarose with 0, 1, 3, 5, 7, 9 and 11% (w/w) canola oil arranged from top to bottom (solid lines), and canola oil (dashed line)
127
Figure 3.3 Cooling profiles of 1% (w/w) agarose with 0, 1, 3, 5, 7, 9 and 11% (w/w) canola oil arranged from top to bottom (solid lines), and canola oil (dotted line) obtained using DSC at a scan rate of 1°C/min.
127
Figure 3.4 Cooling profiles of G' for 1% (w/w) agarose with 0 (○), 1% (■), 3% (x), 5% (●), 7% (Δ), 9% (-) and 11% (♦) (w/w) canola oil from 50°C to 5°C at a scan rate of 2°C/min. The inset corresponds to calibration curve of G' at 5°C as a function of agarose concentration.
128
129 Figure 3.6
Figure 3.5 Modeling the phase topology of 1% (w/w) agarose plus 3% (w/w) canola oil using the isostrain blending law. Predicted storage modulus of agarose (G'ag) and composite (G'C(BL)) are illustrated by dashed and solid lines, respectively, and the experimental composite modulus (G'exp) is shown by a dotted line. Sag is the solvent content of the agarose phase. The inset shows experimental G' values (■) and calculated phase volume of canola oil (▲) for 1% (w/w) agarose with increasing oil concentration of (solid line fit on the inset and the oil phase volume are derived from the isostrain blending law). (a) 2D, (b) Z-stack, and (c) 3D CLSM images of 1% (w/w) agarose plus 5% (w/w) canola oil. The oil droplets are either in purple (a-b) or grey (c), and the agarose is in black (a-c).
130
Figure 3.7 Phase volume of canola oil in mixture with 1% (w/w) agarose determined via a Z-stack method coupled with image analysis software (FIJI and Imaris).
Chapter 4
137
Figure 4.1 SEM micrographs at 50x magnification of (a) 1% (w/w) agarose, and at 400x magnification of 1% (w/w) agarose containing (b) 0% ghee, (c) 3% (w/w) ghee and (d) 15% (w/w) ghee.
137
x
Figure 4.2 FTIR spectra of ghee (dotted line), and 1% (w/w) agarose with 0, 1, 3, 6, 9, 12 and 15% (w/w) ghee arranged from top to bottom (solid lines).
138
Figure 4.3 Exothermic peaks from the cooling profiles of (a) ghee (solid line) and agarose (dotted line) by themselves and (b) 1% (w/w) agarose with 1, 3, 6, 9, 12 and 15% (w/w) ghee arranged from bottom to top.
139
Figure 4.4 Cooling profile of storage modulus for 1% (w/w) agarose with 0 (□), 1% (җ), 3% (-), 6% (x), 9% (●), 12% (Δ) and 15% (♦) (w/w) ghee and ghee by itself (■) during cooling from 50 to 5°C at a scan rate of 2°C/min. The inset denotes the agarose calibration curve of G' at 5°C as a function of its concentration
140
140 Figure 4.6
Figure 4.5 Computerized modeling for 1% (w/w) agarose plus 6% (w/w) ghee, using the Lewis and Nielsen blending law, is illustrated depicting the predicted storage modulus of agarose (G'ag; dashed line), composite (G'C(LN); solid line) and the experimental composite modulus (G'exp; dotted line) as a function of the solvent content of the agarose phase (Sag). The experimental Gʹ values (●) and phase volume estimates of ghee (♦) for 1% (w/w) agarose with increasing ghee concentrations are depicted in the inset (blending law is used to derive the solid line fit of composite modulus and the ghee phase volume in the inset). (a) 2D, (b) Z-stack, and (c) 3D CLSM images of 1% (w/w) agarose plus 6% (w/w) ghee. The green droplets depict the ghee whereas the black background corresponds to agarose.
141 Figure 4.7 Confocal images of 1% (w/w) agarose and 20% (w/w) ghee captured
from (a) the surface and (b) a depth of 80 µm from the surface.
Chapter 5
147
Figure 5.1 SEM images of (a) 1% (w/w) agarose network at 50x and, walls of 1% (w/w) agarose network containing (b) 0, (c) 0.6 and (d) 1.2 % (w/w) MCC under 600x.
148
Figure 5.2 FTIR spectra between 400-4000 cm-1 for agarose with 0, 0.1, 0.3, 0.6, 0.9, 1.2 and 1.5% (w/w) MCC and MCC by itself stacked upwards from the the bottom.
148
Figure 5.3 DSC thermograms of (a) cooling and heating profiles of hydrated MCC paste and (b) cooling profiles of agarose with 0, 0.1, 0.3, 0.6, 0.9, 1.2 and 1.5% (w/w) MCC arranged successively upwards
149
Figure 5.4 Profiles of Gʹ for 1% (w/w) agarose with 0 (●), 0.1% (x), 0.3% (▲), 0.6% (□), 0.9% (♦), 1.2% (-) and 1.5% (○) (w/w) MCC during cooling from 50 to 5°C with a controlled strain of 0.1% and a scan rate of 2°C/min. The agarose calibration curve of Gʹ at 5°C as a function of its concentration is illustrated in the inset.
149
xi
Figure 5.5 Values of composite Gʹ obtained via rheological experiments (x) and theoretical blending laws predictions (solid line) along with the estimated phase volumes of MCC (♦) for samples containing 1% (w/w) agarose with increasing MCC concentrations.
150 Figure 5.6
(a) 2D, (b) Z-stack, and (c) 3D CLSM images of 1% (w/w) agarose containing 0.6% (w/w) MCC. The blue background denotes the stained agarose phase whereas the black cavities correspond to the unstained MCC particles.
150
Figure 5.7 Phase volume estimates of MCC in mixture with 1% (w/w) agarose determined by analysing 3D CLSM images using image analysis software – FIJI and Imaris.
Chapter 6
159
Figure 6.1 FTIR spectra for (a) single systems of 3% agarose and 3% gelatin and (b) 3% gelatin with 0.1, 0.4, 0.8, 1.2, 1.6, 2, 2.5 and 3% agarose arranged upwards.
160
Figure 6.2 Cooling profiles of single systems of 3% gelatin (dotted line), 3% agarose (dashed line), and 3% gelatin with 0.1, 0.4 0.8, 1.2, 1.6, 2, 2.5 and 3% agarose (solid lines) arranged from bottom to top, obtained at a scan rate of 1°C/min using DSC.
162
164
168 Figure 6.5
172 Figure 6.6
Figure 6.3 Variation in storage modulus (a) as a function of temperature and time of observation for single systems of 3% gelatin and 3% agarose, and (b) on heating 3% gelatin plus 0.1, 0.4, 0.8, 1.2, 1.6, 2, 2.5 and 3% agarose, arranged upwards. Scan rate: 2°C/min, frequency: 1 rad/s, strain: 0.1%. Figure 6.4 2D CLSM images of 3% gelatin with (a) 0.1, (b) 0.8, (c) 1.6 and (d) 2.5% agarose, at 10x. Agarose phase is denoted in black and gelatin phase in purple. Scalebar denotes 100 µm. (a) Calibration curves of Gʹ at 5°C as a function of agarose (▲) and gelatin (●) concentrations. Modeling phase topology using – the isostrain (GʹU) and isostress (GʹL) blending laws for 3% gelatin and (b) 0.1, (c) 0.8 and (d) 2.5% agarose, and blending law for bicontinuous binary gels for (e) 3% gelatin plus 1.2,1.6, 2.0, 2.5 and 3% agarose (solid lines) arranged upwards along with the experimental values of the same concentrations (dotted lines) arranged upwards; and (f) predictions of the p factor for all the experimental concentrations based on the isostrain (■), isotress (●) and bicontinuous (♦) blending laws with the inset representing the estimated phase volume of the gelatin phase with increasing agarose concentrations. (a, d) 2D, (b,e) Z-stack and (c, f) 3D CLSM images of 3% gelatin and 0.4 and 2.5% agarose respectively. The gelatin phase is depicted in purple whereas agarose phase is in black.
174
xii
Figure 6.7 Gelatin phase volume estimates plotted against increasing agarose concentrations added to 3% gelatin determined by quantifying 3D CLSM images using image analysis software – FIJI and Imaris, and the rheology based blending laws.
Chapter 7
187 Figure 7.1
188 Figure 7.2
xiii
(a) 2D, (b) Z-stack, and (c) 3D CLSM images of 3% (w/w) agarose plus 2% (w/w) gelatin and 2% (w/w) canola oil. The agarose, gelatin and oil phase is denoted in green, black and red respectively. The composite 3D CLSM image is further split to denote 3D images of the constituent (d) agarose and gelatin, and (e) oil phase separately (Images captured at 10x, scale bar denotes 200μm). (a) 2D, (b) Z-stack, and (c) 3D CLSM images of 10% (w/w) gelatin plus 5% (w/w) WPI and 5% (w/w) canola oil. The gelatin, WPI and oil phase is denoted in red, black, and green, respectively. The composite 3D CLSM image is further split to denote 3D images of the constituent (d) gelatin and WPI, and (e) oil phase separately. (Images captured at 10x, scale bar denotes 200μm)
List of Tables
14
Chapter 1 Table 1.1
28 53 Summary of analytical techniques used to measure biopolymer concentration in separated phases in an aqueous mixture Spectroscopic techniques used to probe phase behaviour Table 1.2 Table 1.3 Microscopic techniques used for structural characterisation of phase
behaviour
87 87 92 Chapter 2 Table 2.1 List of chemicals and dyes used Table 2.2 List of instruments used Table 2.3 Absorption frequencies of various organic functional groups in the mid-
100
130
IR region Table 2.4 Rheological parameters Chapter 3 Table 3.1 Comparison of canola oil phase volume in mixture with agarose calculated using different FIJI thresholding methods (n=27)
141 Chapter 4 Table 4.1 Experimental ghee concentrations in the composite and estimated phase
volumes with FIJI and Imaris image analysis
xiv
189 Chapter 7 Table 7.1 Experimental constituent concentrations in the composites and estimated phase volumes with Imaris image analysis
List of abbreviations
BSE Back scattered electron
CLSM Confocal Laser Scanning Microscopy
Degree of polymerisation DP
Differential Scanning Calorimeter DSC
DMSO Dimethyl sulfoxide
DMA Dynamic Mechanical Analyser
DMTA Dynamic Mechanical Thermal Analyser
Dichlorotriazinyl Aminofluorescein DTAF
Fourier Transform Infrared Spectrometer FTIR
High performance liquid chromatography HPLC
Infrared IR
Linear Viscoelastic Region LVR
Microcrystalline cellulose MCC
MDSC Modulated Differential Scanning Calorimeter
Nuclear Magnetic Resonance NMR
Potential of Hydrogen pH
RMIT Royal Melbourne Institute of Technology
Scanning Electron Microscope SEM
Secondary Electron SE
Transmission Electron Microscope TEM
UV/Vis Ultraviolet/visible
versus vs
xv
X-ray diffraction XRD
2D Two dimensional
3D Three dimensional
ANOVA Analysis of variance
e.g. Example
RMMF RMIT Microscopy and Microanalysis Facility
Milk protein concentrate MPC
Skim milk powder SMP
Diffusing wave spectroscopy DWS
Kilo Dalton kDa
Molar M
milligram mg
Pascal Pa
radian rad
Electronic spin resonance ESR
Small angle neutron scattering SANS
Small angle light scattering SALS
Sodium hydroxide NaOH
nanometer nm
Atomic force microscopy AFM
Weight by weight w/w
Weight by volume w/v
ESEM Environmental scanning electron microscopy
alpha α
xvi
Beta β
List of symbols and units
absorbance AU
Change in enthalpy ΔH
Change in temperature ΔT
Dalton Da
Degree °
Degree Celsius °C
°C/min Degree Celsius per minute
Storage/elastic modulus Gʹ
Loss modulus G′′
Glass transition temperature Tg
g/mol Gram per mole
Heat capacity Cp
Hertz Hz
Kilo volt kV
Litre L
Micrometer µm
Milliampere mA
Milligram mg
mg/kg Milligram per kilogram
Millimeter mm
Millilitre mL
minute min
xvii
molar M
Newton N
Pascal Pa
Per centimetre cm-1
Phase angle δ
Rotation per min rmp
second s
Water activity aw
Weight by weight w/w
Percent %
centimetre cm
nanometer nm
MPa Megapascal
gram g
density ρ
Angular frequency ω
Phase volume ɸ
wavelength λ
Rayleigh ratio Rθ
Solvent avidity parameter p
Complex modulus G*
magnification X
xviii
hour h
Explanatory notes
The following notes briefly delineate the points that were taken into consideration during the
writing of this thesis.
i) Attempts have been made to use British spellings in the text except in the published
journal articles where the journal guidelines have been followed.
ii) Symbols or abbreviations, used in place of a lengthy name or expression, have been
defined or explained in appropriate places as far as practicable.
iii) Wherever possible, SI units have generally been used in expressing results
throughout this thesis.
iv) The current guidelines to authors for Food Hydrocolloids (published by Elsevier)
has been followed in this thesis except in the published articles where journal
xix
guidelines have been followed.
Summary
The importance of estimating phase separation in protein and polysaccharide gels in
order to obtain desired structural properties and textural profile in food product formulations
remains paramount. Due to thermodynamic imbalance, biopolymer mixtures segregate into
distinct phases, with one phase acting as a continuous matrix in which the second phase remains
dispersed as a discontinuous ‘filler’. Studies up till now show that theoretical models relate the
elastic modulus of the biopolymer phases to the topology of their mixtures and can be
successfully employed for an indirect estimation of the solvent partition between the
biopolymers. Theoretical modelling, though robust, is a time consuming and tedious method
of estimating phase volume of phase separated polymers. Acknowledging this, a need for
developing a rapid, efficient, and accurate method of determining solvent partition between the
biopolymers is identified.
Therefore, this PhD study aims at exploiting the accelerated advancement in image
processing technology and developing a novel Confocal Laser Scanning Microscopy (CLSM)-
based approach using Z-stack imaging and image analysis to quantify phase behaviour of
biopolymer composites. This is achieved by employing numerous analytical and
physicochemical techniques such as scanning electron microscopy, Fourier transform infrared
(FTIR) spectroscopy, differential scanning calorimetry, rheology based theoretical modelling
and confocal laser scanning microscopy.
The first experimental chapter focused mainly on developing a protocol for obtaining
3D images of a simple model system of agarose and canola oil and quantifying the lipid phase
volume using two image analysis software, FIJI and Imaris in parallel. For comparison,
rheology based- theoretical blending laws were employed. The phase behaviour of composite
1
gels made of agarose and various concentrations of canola oil were examined using a variety
of techniques including SEM, FTIR, DSC, dynamic oscillation, polymer blending laws and
CLSM-based Z-stack imaging coupled with image analysis. Microscopic, spectroscopic, and
thermomechanical observations recorded continuous agarose networks supporting
discontinuous canola oil inclusions with increasing levels of canola oil reinforcing the rigidity
of the continuous phase. The outcome of the microscopic protocol is in close agreement with
the oil phase volume predictions from the isostrain blending law indicating the suitability of
the developed protocol in quantifying the phase behaviour of composite gels.
The second experimental chapter probed the phase behaviour of a model system
comprising agarose and a varying concentration of a hard lipid, ghee. Results obtained from
SEM, microDSC, FTIR and dynamic oscillation in-shear revealed discontinuous and hard
inclusions of ghee reinforcing the continuous, weaker agarose matrix with increasing
concentrations of the former. Phase behaviour of the system was quantified in parallel with a
novel method combining 3D CLSM imaging and image analysis software - FIJI and Imaris -
in an effort to substantiate the efficacy of the microscopic protocol in quantifying phase
behaviour. Phase volumes recorded with the microscopic protocol were in close agreement to
those modelled with the isostress blending law using small-deformation dynamic oscillation.
However, results indicated that the inner filtering effect or ‘self-shadowing’ observed
commonly in CLSM images due to the diffraction of laser as it passes through components
with different densities and refractive indices may pose a limitation to the application of this
technique, necessitating further development before it can be applied to more complex,
industrially relevant systems.
This merited further consideration of the suitability of the developed microscopic
protocol in estimating solvent partition in an industrially relevant hydrogel. In doing so, the
efficacy of confocal laser scanning microscopy (CLSM) paired with image analysis software
2
– FIJI and Imaris - in quantifying phase volume in a model system of agarose with varying
concentrations of microcrystalline cellulose (MCC) in comparison to the rheological blending
laws was probed. Structural studies performed using SEM, FTIR, differential scanning
calorimetry and dynamic oscillation in-shear unveiled a continuous, weak agarose network
supporting the hard, rod-shaped MCC inclusions where the composite gel strength increased
with higher ‘filler’ concentration. The phase volumes of MCC, estimated with the microscopic
protocol, matched the predictions obtained from computerized modelling using the Lewis-
Nielsen blending laws. Results highlighted the suitability of the microscopic protocol in
estimating the water partition and effective phase volumes in the agarose-MCC composite gel.
These were encouraging outcomes which lead us to estimate phase behaviour in a
mixed gelling system of a polysaccharide (agarose) and a protein (gelatin) in comparison with
the rheology based blending laws. Structural properties of the composites were probed using
FTIR spectroscopy, microDSC and small-deformation dynamic oscillation in shear.
Throughout the experimental range of concentrations, gelatin formed a continuous network
whereas the agarose phase remained dispersed as either a soft or a hard filler at low agarose
concentrations (0.1-0.8%) and formed a continuous network alongside that of gelatin at higher
concentrations (1.2-3%). The phase volumes of gelatin, recorded using Imaris, were a close
match with those obtained from the blending law predictions, whereas FIJI yielded statistically
different estimates. Results suggest that the microscopic protocol using Imaris shows promise
and can potentially be used as an alternative to the theoretical blending laws in estimating phase
behaviour in biomaterial composites.
Work outlined in this thesis thus lays the groundwork for future research, where it can be
used to determine phase behaviour accurately and rapidly in complex binary and tertiary gelling
3
biomaterial composites and perhaps to materials in the nanoscale.
Chapter 1
Background and Literature Review
4
1.1. Introduction
Industrial processing generally involves multiple polymers interacting intricately to
develop products whose physiochemical properties are distinct from those of their constituents.
Mixtures of biopolymers thereby broaden the spectrum of structural and textural variations
more than deemed possible and offer a better control over processability, nutritional value of
the product, kinetics of active compound release and texture and shelf stability of the final
product (Prameela, Mohan, & Ramakrishna, 2018). Hence, it is no surprise that biopolymers
today have a marked predominance in the food, pharmaceutical, cosmetic, and packaging
industry to name but a few.
Biopolymer composites, however, are stable and co-soluble only in very dilute systems
or in cases where the mixing entropy dominates. When the concentration of the biopolymers is
increased beyond the critical concentration, the system becomes unstable and phase
separation/demixing occurs (Walter, Johansson, & Brooks, 1991). Phase separation can either
be segregative, where the biopolymers repel each other leading to an enrichment of each
constituent in a distinct phase, or associative, where due to weak attractive forces and
nonspecific interactions (hydrogen bonding, electrostatic, hydrophobic and Van der Waals
interactions) both the biopolymers concentrate in one phase forming complex coacervates or
insoluble compounds (Tolstoguzov, 2007).
Since the pioneering works of Beijerinck (1896), outlining the formation of tiny
droplets in a mixed solution of starch and gelatin, there is no scarcity of publications
investigating and reviewing phase separation in biopolymer mixtures, the mechanism behind
it, the factors influencing it and ways in which it can be exploited by the processing industries.
Despite the plethora of literature available, phase behaviour of biopolymer mixtures remains
5
of interest and an active research area among researchers and industrialists.
Generally, while investigating mixed biopolymer systems in the aqueous state, a phase
diagram is created of the constituents at desired processing parameters to establish the
boundaries of their compatibility. Depending on the accuracy required, several methods can be
employed to construct a biopolymer phase diagram. Rudimentary methods such as visual
observation (Hemar, Tamehana, Munro, & Singh, 2001) and the phase volume ratio method
(Antonov, Gonçalves, 2018; Schorsch, Clark, Jones, & Norton, 1999), frequently accompanied
by centrifugation (Thaiudom & Goff, 2003) to speed up the process and induce bulk phase
separation, have also been successfully utilised for simple binary mixtures. The construction
of a more detailed and precise phase diagram entails determination of biopolymer
concentration in the separated phases using analytical techniques. Techniques like
refractometry (Semenova & Savilova, 1998), flow injection analysis (Kontogiorgos, Tosh, &
Wood, 2009), phenol-sulphuric method (Masuko et al., 2005), protein analysis method (C
Schorsch, Jones, & Norton, 1999), turbidimetry (Kontogiorgos et al., 2009), viscometry and
quantitative HPLC have been successfully utilized in the past to determine
protein/polysaccharide concentration in the individual phases.
Industrially relevant biopolymer systems, however, often comprise of highly viscous
and gel forming polymers, making it difficult to accurately study their phase behaviour using
the aforementioned techniques. A biopolymer gel, which is essentially a three-dimensional
network of polymer chains, entraps a large volume of solvent within its network and in the
presence of a gel forming polymer; therefore, gelation and phase separation act as competing
processes (Anderson & Jones, 2001). The onset of gelation arrests the polymer demixing in a
state of non-equilibrium, yielding a complex, unique structure for the composite. Although this
opens up endless possibilities of manipulating the properties of the resultant gel by altering the
effects of the two processes, pinning the exact mechanism and extent of phase separation
6
becomes a formidable problem. As such, there is no time-temperature superposition principle
that can sustain a composite gel’s state of thermodynamic equilibrium as the phase diagram
does in solutions (Shrinivas, Kasapis, & Tongdang, 2009).
In composite gels, the concentration of the phases for one, can vary considerably within
the system subject to the degree of phase separation attained when gelation occurred. The
demixed phases may not be pure solutions and can exhibit the presence of multiple inclusions
of each phase within each other. The internal and external conditions of the system may further
lead to changes in the micro and macroscopic structure of the composite gel. With their
complex chemical composition and intricate structural hierarchy, the systems often need to be
‘adjusted’ to overcome interferences from other components. Systems may need to be
physically or chemically treated to make them suitable to be analysed, introducing the risk of
physical contamination and chemical alteration, making the results additionally dubious and
the whole endeavour laborious (Agbenorhevi & Kontogiorgos, 2010; Tolstoguzov, 1999).
Given that biopolymer mixtures encompass a broad range of products with a broader
spectrum of textural and structural properties, accurately determining their phase behaviour
with a single technique is a major challenge as it results in an incomplete, tunnel visioned
understanding of the composite. In the past, different types of gravimetric (Mousia, Farhat,
Blachot, & Mitchell, 2000), calorimetric (Conti, Yoshida, Pezzin, & Coelho), spectroscopic
(Icoz & Kokini, 2007), microscopic (van de Velde et al., 2003), rheological analysis
(Semasaka, Mhaske, Buckow, & Kasapis, 2018), and scattering measurements have been used
to characterize phase behaviour of multiple biopolymer composites from the molecular level
up to their macroscopic behaviour and properties (Fig.1.1). Each of these techniques have their
own advantages and limitations while probing specific parameters of phase-separated
composites under certain experimental conditions and a thorough understanding of the
7
characterisation techniques can facilitate a better, interdisciplinary approach of following phase
behaviour.
Fig.1.1. Scale of measurement of the various techniques used to probe phase behaviour.
This review therefore, aims at summarising the basic principles of the techniques
inherent in the estimation of phase behaviour of mixed biopolymer systems in the aqueous and
gel state, along with their key advantages and limitations with relevant examples without
attempting to provide an exhaustive list of applications.
1.2. Phase behaviour in mixed biopolymer systems
Biopolymers in a mixture seldom behave in the same way as they would in the absence
of the other. In mixture of similar biopolymers and polyelectrolytes, a biopolymer can change
the ionic environment of the solution inducing gelation, complex formation or other associative
interactions of the other (Sarika, Pavithran, & James, 2015), or compete with it for ions from
added salts in the mixture (Donato, Garnier, Novales, & Doublier, 2005). On the other hand, a
mixture of dissimilar biopolymers, proteins and polysaccharides for example, can have one of
three interactions – co-solubility, complexing or segregation due to incompatibility. As
mentioned earlier, a co-soluble biopolymer mixture is a rare possibility, occurring in cases
where the total polymer concentration does not exceed 2-4% for aqueous mixtures of linear
8
polysaccharides or gelatin and 12% for mixtures containing globular proteins, making phase
separation in biopolymer composites a rule rather than an exception (Tolstoguzov, 2003;
Zasypkin, Braudo, & Tolstoguzov, 1997).
Depending on the polymeric nature of the biopolymers, their molecular weights and the
different functional groups present in their macromolecules, the constituent biopolymers in a
mixture are either attracted to or repulsed from each other. This forms the basis for their
complexing or thermodynamic incompatibility (de Kruif & Tuinier, 2001). Electrostatic
attraction between polyelectrolytes such as the anionic polysaccharides and cationic proteins
below their isoelectric points leads to the formation and precipitation of the two polymers as a
complex coacervate. Formation of a coupled network as a result of association between two
gelling biopolymers is also observed in certain, but rare cases (Tanaka, 2002). The most
common outcome of mixing two biopolymers is a segregative interaction or thermodynamic
incompatibility which occurs due to the fact that macromolecules tend to be surrounded by
others of the same type and polarity, resulting in a reduced concentration of one polymer in
close proximity of the other (Grinberg & Tolstoguzov, 1997; Tolstoguzov, 1993). When the
concentration is high enough, the biopolymers separate completely into two phases each rich
in one biopolymer and depleted in another (Frith, 2010).
Gibbs free energy of mixing which considers the entropy and enthalpy of mixing is a
popular method of deducing whether the constituents of a mixture will phase separate or not.
Non-covalent interactions between the solutes including polar and non-polar attractions, ion
pairing, van der Waals forces and hydrogen bonding are exothermic reactions whereas
endothermic reactions include repulsion between molecules or disruption of molecular
bonding. Entropy denotes the energy changes corresponding to the alterations in the molecular
arrangement. A positive entropy corresponds to the increase in the degree of freedom of
9
molecular motions, while a negative value denotes a decrease in it (Schmitt, Sanchez, Desobry-
Banon, & Hardy, 1998). Gibbs free energy of mixing (ΔGmix), is given by the following
equation (Goh, Teo, Sarkar, & Singh, 2020):
ΔGmix = ΔHmix - TΔSmix (1)
Where ΔH is the enthalpy of mixing and ΔS the entropy of mixing at temperature T in Kelvin.
Considering systems in equilibrium, according to the second law of thermodynamics,
co-existing phases must have the lowest ΔGmix at temperature T. In most cases, ΔHmix is
positive, favouring demixing of the solutes. Solutes with a low molecular weight have an
adequately large ΔSmix value that results in the co-solubility of solutes. In biopolymers,
however, ΔSmix is much smaller and ΔHmix dominates, leading to phase separation. Depending
on the charge on the two biopolymers, the system then undergoes either an associative or a
segregative phase separation. In segregative interactions, there is no thermodynamic drive for
heterotypic association and the enthalpic drive for segregation dominates (Tolstoguzov, 1991).
Grinberg & Tolstoguzov (1997), in their work report the incompatibility conditions for over a
hundred different protein-polysaccharide composites. The thermodynamics and kinetics of
phase separation have been discussed in detail elsewhere (Bergfeldt, Piculell, & Linse, 1996;
Flory, 1953; Fredrickson, Liu, & Bates, 1994; Johansson, Karlström, Tjerneld, & Haynes,
1998; Robeson, 2007).
1.3. Phase diagrams of protein-polysaccharide-water systems
A phase diagram depicts the phase behaviour of two thermodynamically distinct
biopolymers (Fig.1.2) (Goh et al., 2020; Tolstoguzov, 1991, 2003, 2006). The concentrations
of biopolymers are generally plotted on the axes in weight percent, the rest of the phase being
the solvent. Each point in the diagram denotes a system composition, with the binodal curve
10
separating the miscible composite concentrations below it from the two-phase system
concentrations above. The closer the binodal is to the axes, the higher the incompatibility
between the biopolymers. Lowering the temperature of the system, increasing the ionic strength
or the molecular weights of the constituents decreases the entropy of mixing and displaces the
binodal towards the origin (Morris, 2009). A composite with a total biopolymer concentration
C, constituting a% of biopolymer 1 and b% of biopolymer 2, phase separates into two bulk
biopolymer-rich phases. The biopolymer 1 enriched phase will have d% biopolymer 1, while
biopolymer 2 enriched phase would have a composition of e%. The biopolymer 1 enriched
phase contains negligible traces of biopolymer 2 and vice-versa. Line d-C-e denotes the tie
line. The ratio of the line segments dC/Ce corresponds to the phase volume ratio of the two
separated phases. Therefore, any composite composition that lies on the tie line would result
in the same effective concentration in the separated phases, but their phase volumes will vary.
A rectilinear diameter can be obtained by joining the midpoints of two or more tie-lines. Any
composition on the rectilinear diameter would phase separate into two phases with the same
phase volume and biopolymer concentration.
11
Fig.1.2. Phase diagram in an aqueous biopolymer mixture.
When the overall composite concentration shifts closer to the origin, the tie lines
become shorter till the two points converge on a single point on the binodal. This point is
known as the critical point, or separation threshold, F in Fig.1.2 which denotes the minimum
critical concentration at which the biopolymer composite will phase separate into two phases.
A lower critical point denotes lower compatibility between the constituent biopolymers. In an
ideal biopolymer composite, when the composite concentration is varied, the corresponding
tie-lines do not cross over. However, this is the case in an ideal biopolymer composite. In reality
most composites exhibit a much more complex behaviour.
An important aspect to consider is the spinodal, the dotted line in Fig.1.2. Lying within
the binodal, the spinodal separates composites based on the nature of their phase separation. If
a composite lies between the binodal and the spinodal, it is likely to phase separate via
nucleation, taking more time than the timescale of the experiment and making it difficult to
study. Composites within the spinodal region undergo spontaneous phase separation owing to
the increasing thermal concentration fluctuations. Gelation of one or more of the biopolymers
of the composite is another factor that needs to be considered as it may arrest phase separation
at different stages depending on various factors or prevent it completely.
As the phase diagram provides critical information of biopolymer compatibility and can
be used to predict the behaviour of two biopolymers in a composite, it is a requisite for the
study of phase behaviour and the fabrication of novel textures and products. Innumerable
publications outline the works of researchers and industrialists studying phase behaviour and /
or constructing phase diagrams of biopolymer composites. Use and development of multiple
techniques is undertaken in an effort to not overlook any of the plenitude of factors that affect
phase behaviour of a composite. The rest of this review will focus on the various techniques
12
used to probe phase behaviour.
1.4. Techniques used to probe phase behaviour
The various analytical techniques used to probe phase behaviour in aqueous mixtures will be
discussed in detail in the following sections and briefly summarised in Table 1.1.
1.4.1. Visual observation and phase volume ratio method
The earliest technique used to study phase behaviour was visual observation. Two
polysaccharides in varying concentrations were mixed to form composites with the same total
polysaccharide concentration. The mixtures would either be left undisturbed over a longer
period of time or be centrifuged to speed up phase separation. The binodal is plotted by
following the different concentration of both biopolymers at which phase separation occurs.
Depending on the degree of incompatibility between the constituents, phase separation within
the composite ranged between microscopic and macroscopic. Michon, Cuvelier, Launay,
Parker, & Takerkart (1995) constructed a phase diagram of acid gelatin and iota carrageenan
at different pH and added salt and further detailed it by classifying the mixtures as ‘clear’ when
it was possible to see through them, ‘cloudy’ when objects appeared blurred when viewed
through the composite, ‘opaque’ when nothing could be seen through them (Fig.1.3).
Thaiudom & Goff (2003) followed the compatibility of skim milk powder with locust bean
gum, guar gum and xanthan gum by centrifuging the mixtures and observing the composite
while Hemar, Tamehana, Munro, & Singh (2001) successfully constructed the phase diagram
of milk protein concentrate (MPC) or skim milk powder (SMP) when mixed with xanthan gum
by visually observing the components phase separate into two distinct layers with a clear phase
boundary.
To qualitatively determine the binodal, Grinberg & Tolstoguzov (1972) added a
13
solution of one constituent biopolymer into the second biopolymer solution till turbidity
Table 1.1: Summary of analytical techniques used to measure biopolymer concentration in separated phases in an aqueous mixture.
Technique
Principle
Advantages
Disadvantages
References
Suitable for a wide
Suitable only for bulk separated
Application (examples) Locust bean
(Antonov & Gonçalves, 2018)
Phase volume ratio
range of composites
phases
gum + gelatin
Doesn’t need sophisticated instruments for analysis
Determination of cloud point concentration results in a coexistence curve which may involve solutions of higher concentrations than initially used
Volumes of two bulk phase separated phases is measured in a calibrated capillary tube; graphical calculation using geometrical analysis is used to calculate weight fraction of the two polymers in individual phases.
Assumes complete depletion of one polymer in the separated phase of the other which isn’t always the case
Refractometry Refractive index of a
Sucrose in agar
Sample needs to have low optical
gels
density
Micellar casein +
Significant differences between the refractive indices of the two biopolymers necessary
(Wang et al., 2014; Yang et al., 2015; Bourriot et al., 1999; Semenova & Savilova, 1998)
More suitable for non-interacting systems that have undergone bulk phase separation
Easy to use Quick Small amounts of sample needed Doesn’t need a sophisticated laboratory set up for estimation
gaur gum 11S globulin of faba beans + fibrinogen/ dextran/ carboxymethylce llulose/ sodium alginate
Phenol sulphuric method
Interferences from protein component significantly impact absorption readings
(Zhang & Foegeding, 2003); (Li et al., 2016); (Perrechil et al., 2009)
β-lactoglobulin + dextran sulfates / kappa- carrageenan soluble rice protein + dextran
Addition of sulphuric acid to proteins produces varying amounts of ammonium sulphate,
Quick Easy Reliable Reagents are easily available and affordable
medium is linearly proportional to its concentration within a range of polymer concentration at a specific temperature. Thus, measuring the refractive index can be an indirect method of estimating the polymer concentration. Phenol and sulphuric acid are added to polysaccharide solution. Dehydrated polysaccharides react with phenol to produce furfural which forms a
14
CO2 and sulphur that alter the native system
As the protein concentration
/ kappa- carrageenan Sodium caseinate + locust bean gum
coloured solution whose absorbance is measured. The intensity of the colour produced is directly proportional to the polysaccharide concentration.
varies, different blank solutions are needed for blank correction Certain systems necessitate the removal of protein from system before estimating the polysaccharide concentration
BSA +
Bradford Assay
Results in varied response between different proteins
Ribonuclease A
(González- González, 2011); (Svensson, 1999)
Assay is linear over a short range
Dextran +
High sensitivity Easy operation Universally accepted
of concentrations
technique
albumin, pluronic acid
Dilution is often necessary
Biuret method
Affected by the association state
Independent of protein composition
(Farouk et al., 2011); (Zhuang et al., 2020)
Meat proteins + carrageenan Myofibrillar
and purity of proteins Isn’t as sensitive as other protein assays
protein + dietary fibre/ modified starch/ konjac glucomannan
Canola protein
Kjeldahl method
(Neirynck et al., 2007); (Klassen et al., 2011)
Universally accepted High precision High reproducibility
isolate + alginate + carrageenan Sodium caseinate
Involves high set up cost Other components containing nitrogen can result in errors in protein estimation
guar gum
Different proteins need different
correction factors Is an indirect method of protein estimation
The anionic part of the Coomassie blue dye, attaches to argenine or lysine residues. The absorbance of the coloured complex is measured. In alkaline conditions, copper sulphate is added to a protein solution. Cupric ions bind with peptide bonds developing a purple colour, the absorbance of which is measured using a colorimeter. Proteins are digested using an acid to release ammonia. The released ammonia/nitrogen is calculated by titration as an indirect estimation or the protein content.
15
occurred and then added the second solution till the composite formed a clear solution. This
was repeated for a series of individual biopolymer solutions with varying concentrations.
Construction of the phase diagram using this cloud point system is convenient when the
constituent biopolymer solutions have a low optical density, are in the aqueous state and the
difference in their refractive coefficient is significant (Tolstoguzov, 2006). However, visual
observation is a qualitative method, where the concentration of biopolymers in the separated
phases and the tie lines of compatibility cannot be determined in the phase diagram.
Fig.1.3. Phase diagram of 3% gelatin/0.9% iota-carrageenan as a function of sodium chloride
concentration and pH (Michon, Cuvelier, Launay, Parker, & Takerkart, 1995).
Polyakov, Grinberg, & Tolstoguzov (1980) developed a more refined technique called
the phase volume ratio method that could be used to deduce the concentration of the polymers
as well as the tie lines in the phase diagram. Similar to visual observation method, mixtures of
two biopolymers A and B with individual weight concentrations, Ai and Bi were mixed together
in varying concentration and allowed to phase separate. This was carried out in graduated
centrifuge tubes or calibrated capillary tubes in order to determine the volumes (V) of the
separated phases. The phase volume ratio (r) of the phases could be calculated using the
following formula (Schorsch et al., 1999):
16
r↑ = V↑/ (V↑+ V↓) (2)
Where ↑ denotes the separated upper phase and ↓ is the lower phase.
The apparent weight fraction of the individual separated phases, y is given by:
(3) y = Ac/Ai
(4) (1-y) = Bc/Bi
Where Ac and Bc are the weight fractions of polymer A and B in the mixture.
The ratio of the volume is then plotted against the apparent weight fraction of the
respective component. Extrapolating the curve of the phase volume ratio as a function of the
apparent weight fraction, with the values of r ranging between 0 and 1, give the values of r on
the boundary of a stable homogeneous mixture and two separated phases. r↑ = 0 denotes one
phase similar to the lower phase while r↑ = 1 corresponds to one phase similar to the upper
phase. This procedure is repeated with solutions of varying polymer concentrations Ai and Bi
with different polymer A/polymer B ratio. The binodal is drawn through the points representing
the composition of the phases after it has been determined. The points denoting composition
of two co-soluble phases are joined to get the tie lines on the phase diagram. The system
composition where r = 0.5 corresponds to the point on the rectilinear diameter (Tolstoguzov,
1999). Antonov & Gonçalves (2018) used the phase volume ratio to construct a phase diagram
of locust bean gum and gelatin. Though this technique overcomes the drawbacks of visual
observation in terms of being able to estimate the polymer concentration in the separated phases
and enables determination on tie-lines, the method only deals with systems that bulk phase
separate and the determination of cloud point concentrations gives a coexistence curve which
isn’t always the same as the initial concentration of phases in the composite left to phase
separate. The method has the added disadvantage that it considers complete depletion of one
polymer from the other leading to anomalies in concentration estimation. To overcome these
17
draw backs, refractometry, turbimetric methods and methods to chemically estimate polymer
concentration in the separated phases are used which have been explained in detail in the
following sections.
1.4.2. Refractometry
The simplest and fastest method of estimating biopolymer concentration in a solution
is refractometry. The speed of light changes in different materials based on the material’s
density. This difference in the speed changes the direction of light as it passes through different
materials. The ratio of the speed of light in vacuum to the speed of light in a specific material
is a constant physical characteristic of the material known as index of refraction (n). The
behaviour of light as it passes between two materials is determined using the Snell’s law of
refraction given by (George, 2001):
(5) 𝑛 = 𝑠𝑖𝑛𝛩𝑖 𝑠𝑖𝑛𝛩𝑟
Where Θi and Θr are angles made by the incident ray and the refracted ray, respectively, at the
intersection of the two materials. The value of n varies according to the wavelength of incident
light and in an aqueous solution is influenced by the temperature of the solution, the solute
added and its concentration. Every solute, or biopolymer, has a specific value of n per
concentration in g/dL, termed as specific refractivity. The refractive index of a biopolymer
solution is linearly proportional to the biopolymer concentration over a wide range of
concentrations at a specific temperature. With an increase in temperature, the value of n
decreases due to the increase in solvent volume. By determining the n of a standard solution at
a specific temperature, it is possible to determine the biopolymer concentration at that
18
temperature.
According to the generally accepted Stockmayer theory (Stockmayer, 1950), in a three
component system, containing biopolymer 1, biopolymer 2 and water, when the concentrations
are expressed in molal units, the interaction between the two biopolymers is measured directly
by the difference between the increments of the refractive index. The refractive indices are
measured under conditions at which the chemical potential of the second biopolymer and its
molality are maintained at the same level as in the solution and reference solvent. The change
in the refractive index, dn is given by (Pittz, Lee, Bablouzian, Townend, & Timasheff, 1973):
𝑇,𝑝,𝑚2
𝑇,𝑝,𝑚1
(6) ) ) 𝑑𝑚2 𝑑𝑛 = ( 𝜕𝑛 𝜕𝑚1 𝑑𝑚1 + ( 𝜕𝑛 𝜕𝑚2
Where n is the refractive index, m1 and m2 is the molal concentration of biopolymer 1 and 2
respectively, p is the pressure (Pa) and T is the temperature (°C).
The value of n can be converted to biopolymer concentration using a refractometer
reticule, or an established calibration curve of the standard. However, the angle of refraction
produced by a biopolymer solution, Θr is a result of the combined concentration of the
constituents also called, “total solids” in a solution. In a biopolymer composite, therefore, if
the constituent biopolymers have similar specific refractivities, the presence of the second
biopolymer will interfere with accurate concentration measurement. To overcome this, Ding et
al. (2002) combined refractometry with UV-Vis spectroscopy to measure the concentration on
dextran and gelatin in separated phases. UV-Vis spectroscopy was first used to determine
concentration of gelatin in both the phases and the increments in refractive index of both the
biopolymers were predetermined. With this information at hand, the concentration of dextran
was calculated using the refractive index of the separated phase. Other examples of applications
of refractometry in concentration estimation include, but aren’t limited to sucrose in agar gels
19
(Wang, Yang, Brenner, Kikuzaki, & Nishinari, 2014; Yang et al., 2015), micellar casein- gaur
gum (Bourriot, Garnier, & Doublier, 1999) and fibrinogen, dextran, carboxymethylcellulose,
sodium alginate and 11S globulin of faba beans (Semenova & Savilova, 1998).
Use of refractometry, however, necessitates that both the biopolymers have a low
optical density with a significant difference in their refractive indices, which is not always the
case with biopolymers. In cases like Ding et al., (2002), it is assumed that the refractive index
of the composite is the sum of the refractive indexes of the individual composites, an
assumption that holds true just for non-interacting systems. The interference of the second
biopolymer makes this inapplicable to systems that haven’t undergone bulk phase separation
and its high temperature sensitivity necessitates that temperature is monitored throughout the
analysis (Loret, Schumm, Pudney, Frith, & Fryer, 2005).
1.4.3. Chemical analysis for concentration estimation
A popular technique of estimating biopolymer concentration in the separated phases is
by chemical analysis. Due to its simplicity and high sensitivity, concentration measurement by
UV absorbance is widely used. However, alternative methods such as optical activity, HPLC,
electrophoresis, and other radiometric and chromatographic methods have also been
successfully used and are discussed in the following sections.
Perhaps the most popular method for estimating the concentration of pure sugar
solution, oligosaccharide of polysaccharides is the phenol sulphuric method (Dubois, Gilles,
Hamilton, Rebers, & Smith, 1956). In this method, trace quantities of phenol are added to an
aliquot of sample along with sulphuric acid. Polysaccharides, when dehydrated due to the
reaction with concentrated sulphuric acid produce furfural derivatives. These furfural
derivatives further react with phenol to form a coloured solution. Absorption, which is
20
proportional to the intensity of the colour developed, is measured at 490nm (Jain, Karibasappa,
Dodamani, & Mali, 2017). This method has been used for quantification of polysaccharides in
a number of studies including the works of Zhang & Foegeding (2003), Li et al. (2016) and
Perrechil, Braga, & Cunha (2009). Masuko et al. (2005) further modified the technique to make
it suitable for microplate analysis. The popularity of the method can be attributed to the fact
that it is quick, easy and reliable involving affordable and easily available reagents. However,
the method is established for pure solutions and the presence of even minor traces of proteins
in the solution have a significant impact on the absorption measurements. Sulphuric acid reacts
with protein to form ammonium sulphate, carbon dioxide and sulphur which alter the overall
chemistry of the reaction. Moreover, multiple blanks are needed for blank correction to
attenuate signals from non-polysaccharide sources (Agbenorhevi & Kontogiorgos, 2010).
A more accurate method of protein estimation is binding it with the anionic dye,
Coomassie blue. This technique is also known as the Bradford assay (Bradford, 1976). The
anionic form of the dye adsorbs to the non-polar amino acid (ideally arginine and lysine)
residues, and the cationic region of proteins. The absorption of the protein-Coomassie blue
complex is measured at 595 nm (González-González, Mayolo-Deloisa, Rito-Palomares, &
Winkler, 2011; Svensson, Berggren, Veide, & Tjerneld, 1999). Since the dye exhibits a
preferential binding to arginine and lysine, it subsequently results in a varied response between
different proteins. The assay is linear typically just within the short range of concentrations
within 0 µg/mL to 2000 µg/mL making dilution necessary before analysis and error in one
dilution is compounded throughout affecting the accuracy of the assay (Ernst & Zor, 2010). In
spite of these complications, the high sensitivity and ease of the assay make it a popular
technique in protein estimation.
Another popular method for the determination of protein concentration in a solution is
the use of biuret reaction, the approaches of which have been around since the works of Gornall
21
et al. in 1949. In this method, copper sulphate is mixed with a protein in a high alkaline solution.
Cupric ions from copper sulphate form a complex with the proteins by binding at the peptide
bonds, producing a violet-purple colour (Farouk, Frost, Krsinic, & Wu, 2011; Zhuang, Wang,
Jiang, Chen, & Zhou, 2020). This reaction, unlike the Bradford assay, is independent of the
protein composition but is affected by association state and purity of proteins. This method,
however, isn’t as sensitive as the other methods of protein estimation and attempts have been
made to modify the technique to increase its sensitivity (Matsushita, Irino, Komoda, &
Sakagishi, 1993). The interference caused by the presence of other components in the solution
such as lactose, fat and turbidity have been successfully eliminated by adding hydrogen
peroxide to the solution (Reichardt & Eckert, 1991). The biuret reaction was amplified by
including a second reaction with Folin phenol reagent in the Lowry method (Waterborg, 2009).
However, in the Lowry method, specific amino acids, namely tyrosine and tryptophan affect
the development of colour resulting in substantial variation in accordance to protein
composition. With the development of simplistic and sensitive protein assays, these methods
however have witnessed a dip in popularity.
The classic Kjeldahl method is an alternative to the colorimetric analysis where the
proteins are ‘digested’ or wet oxidised in the presence of a catalyst in sulphuric acid. The acidic
ammonium sulphate formed is diluted in water and alkalinised. This is followed by titration to
estimate the content of ammonia/nitrogen representing the amount of protein in the sample
(Bradstreet, 1954). Kjeldahl method is popular globally due to its high reproducibility and
precision. However, it involves a high set up cost and is an indirect method of protein
estimation that may result in errors as samples may have other sources nitrogen apart from
protein. Furthermore, different proteins need different correction factors due to the varying
amino acid sequence (Mæhre, Dalheim, Edvinsen, Elvevoll, & Jensen, 2018).
All the chemical assays mentioned here, along with others that aren’t so popular, for
22
example, use of methylene blue, ninhydrin, bicinchoninic acid, Dumas method, BCA assay,
‘blue value’ assay, have been explained in detail elsewhere in literature (Copeland, 1994;
Gilbert & Spragg, 1964; Jiang, Tsao, Li, & Miao, 2014; Marshall, 2005; Sapan, Lundblad, &
Price, 1999; Soedjak, 1994). A common drawback for all chemical assays of concentration
estimation is that it is only feasible for a polymer if there is an established test for it.
1.4.4. Viscometry
Along with its wide application in determining degree of polymerisation, molecular
weight and molecular weight distribution (Masuelli & Illanes, 2014), there have been various
attempts of correlating the viscosity of ternary polymer solutions with its miscibility. When a
biopolymer is added to a solvent, its configuration and hydrodynamic volume is influenced by
the solvent used, affecting the solution’s viscosity. In a biopolymer composite, it is assumed
that beside solvent-biopolymer interactions, mutual interactions of the macromolecules greatly
influence composite viscosity. The attractive or repulsive forces between the unlike polymers
can lead to the contraction or expansion of the polymer strands. According to the classic
Huggins theory, the dimensions of the strands can be correlated to the intrinsic viscosity of a
composite. Viscosity of biopolymers can be measured using viscometers and rheometers that
have been explained in detail elsewhere (Gupta, Wang, & Vanapalli, 2016).
Characterisation of the thermodynamic and hydrodynamic interaction between the
biopolymer strands can help determine the phase behaviour of the composite. According to the
framework established by Cragg-Bigelow (1955) and Krigbaum-Wall (1950), the adaptation
(𝜂𝑠𝑝)
𝑖
of the Huggins equation to a polymer-solvent system is denoted as
𝑐𝑖
23
(7) = [𝜂]𝑖 + 𝑏𝑖𝑖𝑐𝑖
Where (ηsp)i is the specific viscosity and ci is the polymer concentration. [η]i is the intrinsic
viscosity given by [η]i = limCm→0 (ηsp)i/Ci. bii is the viscometric interaction parameter related
i (Vis et al.2015).
to the Huggins coefficient as bii = kH[η]2
(𝜂𝑠𝑝)
𝑚
For a mixture of polymers in a solvent, Eq. 7 can be extended as:
𝑐𝑚
(8) = [𝜂]𝑚 + 𝑏𝑚𝑐𝑚
Where ‘m’ stands for mixture. Since reduced viscosity is an additive property, Eq. 8 can be
(𝜂𝑠𝑝)
(𝜂𝑠𝑝)
𝑚
𝑖
rewritten in the weight-average form as follows:
𝑐𝑚
𝑐𝑖
(9) = ∑ 𝑖 𝑤𝑖
Where wi is the weight fraction of the biopolymer i (i =2,3) given by wi = ci/cm. Combining
(𝜂𝑠𝑝)
𝑚
Eqs. 1 and 3, leads to :
𝑖
𝑖
𝑐𝑚
1/2
+ = ∑ ([𝜂]𝑖 + 𝑏𝑖𝑖𝑐𝑖) 𝑤𝑖 = ∑ [𝜂]𝑖𝑤𝑖 𝑖 + 𝑐𝑚 ∑ 𝑏𝑖𝑖𝑤2 𝑖 = ∑ [𝜂]𝑖𝑤𝑖 𝑖
𝑖
2 𝑤𝑖)
(10) 𝑐𝑚(∑ 𝑏𝑖𝑖
Comparing Eqs. 2 and 4, with i=2,3 it can be deduced that:
𝑖
2
2
(11) [𝜂]𝑚 = ∑ [𝜂]𝑖𝑤𝑖 = [𝜂]2𝑤2 + [𝜂]3𝑤3
1 2 𝑤𝑖)
1 2𝑤2 + 𝑏33
1 2𝑤3)
And 𝑏𝑚 = (∑ 𝑏𝑖𝑖 = (𝑏22
2 + 𝑏33𝑤3
2 + 2𝑏22
1/2𝑏33
1/2𝑤2𝑤3
id between the strands
(12) = 𝑏22𝑤2
The ‘ideal’ value of the universal viscometric interaction parameter, bm
24
of the constituent biopolymers in the mixture is defined using Eq. (6). This is done with an
id, the geometric
assumption that for the ideal value of the specific interaction parameter b23
1/2
mean of those parameters corresponds to interaction between like-stranded polymers as
id = 𝑏22
1/2𝑏33
(13) b23
By parallelism, the experimental value of the same parameter can be defined as:
𝑒𝑥𝑝 = 𝑏22𝑤2
2 + 𝑏33𝑤3
2 + 2𝑏23
𝑒𝑥𝑝𝑤2𝑤3
𝑒𝑥𝑝 is estimated from the slope of Eq. (2), and b22 and b33 are determined using Eq 1
(14) 𝑏𝑚
Where 𝑏𝑚
from the composite formed by polymers 2 and 3, respectively.
𝑒𝑥𝑝
The compatibility between the biopolymers in a mixture is based on the comparison
between the theoretical and experimental b23 values (Krigbaum & Wall, 1950). When 𝑏23
id or Δbm > 0 the biopolymers are compatible or have attractive forces between them and
𝑒𝑥𝑝 < b23
id or Δbm < 0, the polymers are incompatible with repulsive
> b23
𝑒𝑥𝑝 = b23
id or Δbm = 0, the polymers have
conversely when 𝑏23
molecular interactions. In hypothetical cases where 𝑏23
neither attractive nor repulsive forces acting between them.
Viscometry has been used to determine the phase behaviour of a number of systems
involving both organic solvents (Chee, 1990; Danait & Deshpande, 1995) and water (Inamura
& Jinbo, 1991; Parets, Garcia, Soria, & Campos, 1990; Wanchoo & Sharma, 2003) as solvent.
The method is simple and straightforward and does not require sophisticated equipment.
id is based on the geometric
However, there seems to be ambiguity in the theoretical framework that has been a subject of
debate among researchers. As pointed out in Eq. 7, calculation of b23
mean value, while there have been arguments that support the use of arithmetic mean value to
id (Garcı́a, Melad, Gómez, Figueruelo, & Campos, 1999). Garcı́a et al. (1999) also
define b23
25
raise the concern of mathematical errors in the framework that has been explained in their work.
1.4.5. Spectroscopy
Spectroscopy deals with the generation, measurement and interpretation of the spectra
produced during the interaction of electromagnetic radiation with matter. Multiple
spectroscopic techniques are available, differing in terms of the region of the electromagnetic
spectrum used for analysis (e.g. UV, visible light, IR), the type of radiation-matter interaction
being observed (e.g. diffraction, absorption, emission) and the material being analysed (atomic
or molecular spectroscopy). The common techniques used to analyse phase behaviour have
been discussed below and summarised in Table 1.2.
1.4.5.1. UV-Vis spectroscopy
This type of spectroscopy analyses compounds using electromagnetic radiations
between 10 nm to 700 nm and has the most applications in the realms of food biopolymer
science. With relatively low energies, the UV-Vis spectrum can only excite the chromophores,
molecular sub-structures that interact with electromagnetic radiation, to the first couple of
levels of the electronic excited state. Chromophores are able absorb energy from the incident
UV-Vis range of the electromagnetic spectrum, and this absorption gives the analyte its
apparent colour which can be measured to estimate the concentration of samples.
Estimation of sample concentration is done based on the principles of the Beer-Lambert
given as follows:
(15) = 𝜀 × 𝑐 × 𝑑 = 𝐴 lg 𝐼𝑜 𝐼 = lg 1 𝑇
where Io and I are the initial and the drop in the intensity of light as it passes the sample. T, ε,
c, d and A, are transmission, the molar absorption coefficient, molar concentration, path length
26
and the samples absorbance that is displayed on the spectrophotometer.
However, absorbance is not always linearly proportional to the concentration of the
chromophores based on the limitations of the spectrophotometer and at high concentration
chromophores might dimerise resulting in different spectroscopic parameters. In samples
exhibiting substantial turbidity, light may not necessarily be absorbed as much as it is scattered
by the sample. The spectrophotometer in these conditions measures the apparent absorbance
(or attenuance) relating to the optical density of the sample and not its concentration. The most
common application of the UV-Vis spectroscopy while probing the phase behaviour of
biopolymers is estimation of changes in polymer concentration based on the assays discussed
in Section 4.3.
1.4.5.2. FTIR spectroscopy
Fourier transform infrared (FTIR) spectroscopy is type of infrared spectroscopy that
uses the mathematical process termed ‘Fourier transform’ to convert raw data into an actual
spectrum. Infrared (IR) analyses compounds using the infrared spectrum that is further divided
into near IR, mid-IR and far IR. Near IR, having the greatest energy penetrates the sample
much deeper than mid or far IR but is therefore the least sensitive. As the energies involved are
relatively low, they do not excite the electrons. However, when samples are subjected to
infrared radiation, they undergo a change in dipole moment leading to vibrational excitation of
groups and atoms that are covalently bonded (Gaca-Zając et al., 2018). Distinct molecules each
have their own unique vibrational energy spectra characterized by their frequency, band shape
and intensity, that serves as the fingerprint for that particular molecule (Gutiérrez Sama et al.,
2018).
In a composite, where the components are in distinct phases, it can be assumed that the
27
spectral fingerprint of one component isn’t affected by the presence of the other. This would
Table 1.2. Spectroscopic techniques used to probe phase behaviour.
Principle
Advantages
Technique Ultrasonic Measures the intermolecular
Disadvantages Not suitable for dilute
Highly sensitive to
suspensions less than 1% concentration
intermolecular interactions Can investigate a wide range of
References (Chanamai & McClements, 2001, Chappellaz, Alexander, & Corredig, 2010; Corredig, Sharafbafi, & Kristo, 2011; Clark, 1996)
repulsions caused in aqueous samples due to compression in ultrasonic waves
molecular processes Non-destructive and non-
invasive
Can be applied to highly
concentrated, optically opaque systems
Raman
Can be used for optically unclear
Due to the inherent weak
(Normand et al., 2000, Pudney et al., 2003, Haug et al., 2003)
samples In situ analysis possible Sample size as less as 1 µL or
Measures wavelength and intensity of inelastic light scattering from molecules along with changes in the polarizability of molecules
1 mg is sufficient
Enhanced signal-to-noise and
simplification of the resonance Raman spectrum enable time- resolved studies
scattering, not applicable for dilute systems Intense lasers can induce sample lability and photo/thermal decomposition Inherent fluorescence of constituents in sample may obscure the Raman scattering signal
The variety of functional
groups and their microenvironments present result in overlapping broad spectral bands
FTIR
Measures the changes in
Identifies functional groups present in the molecule
(Shamsuri, Abdullah, & Daik, 2012; Rafe & Razavi, 2015)
Sampling chamber has a relatively small size Some materials absorb
intrinsic dipole moment and bond vibration frequencies in molecules
infrared radiations making the results unreliable
Mounted pieces can obstruct
the IR beam
28
Advantages Fast, non-destructive and
Disadvantages Needs additional
Table 1.2. Spectroscopic techniques used to probe phase behaviour (continued). Technique Principle Diffusing Wave
reproducible
Doesn’t necessitate sample
References (Weinbreck, 2004; Blijdenstein, 2003; Gancz, 2005; Liu, 2007)
experimentation to predetermine various parameters
dilution
Quantifies in-situ dynamics of a colloidal turbid system by measuring the intensity fluctuations of scattered occurring due to Brownian motion of colloidal particles
UV-Vis
Measures sample
Necessitates tedious sample
preparation
Quick and accurate Easy to use
Due to high accuracy,
(Jain, Karibasappa, Dodamani, & Mali, 2017; Agbenorhevi & Kontogiorgos, 2010)
concentration based on the absorption of UV-Vis radiation by the sample during electron excitation
readings can be affected by minute contaminants, light or vibration form equipment nearby
Generates a weaker signal
Non-destructive technique Allows examination of materials
Nuclear Magnetic Resonance
compared to other techniques
in the solid state
(Voron’ko, Derkach, Vovk, & Tolstoy, 2016; Dai & Matsukawa, 2013)
Analyses the environment of hydrogen or carbon nuclei in a compound in an externally applied magnetic field
Paramagnetic species have a
Electron spin resonance
(Schorsch, Jones, & Norton, 2000; Williams, Clegg, Langdon, Nishinari, & Piculell, 1993)
Measures the microwave energy absorbed by unpaired electrons in a magnetic field
low chemical lifetime Some species require low temperatures for detection
Employs non-ionizing radiation Preparation steps or extraction not needed prior to analysis Greater sensitivity than NMR Offer higher sensitivity Require low sample volumes Can analyse dilute samples
29
result in a composite spectrum that depicts the peaks of both the components in intensities
corresponding to their concentrations. When composites are partially or completely miscible,
the composite spectrum is likely to show formation of new bands, disappearance of some
component bands or shifts in bands indicating shifts in component specific bonds to bonds
between components (Icoz & Kokini, 2007).
In their work, Shamsuri, Abdullah, & Daik (2012) used FTIR to study the phase
behaviour in agar/biopolymer (cellulose, rice starch or zein protein) blend aerogel fabricated
in ionic liquid and co-solvent mixture. The blends of the composites replicated all the chemical
moieties seen in the component spectral scan, indicating that the components did not undergo
any chemical changes and remained in their distinct phases in the composite. In another study
conducted by Rafe & Razavi (2015) on β-lactoglobulin and basil seed gum composite it was
reported that the components were thermodynamically incompatible as the composite spectra
reproduced the individual peaks of the components with a significant increase in intensity with
relation to concentration.
1.4.5.3. Raman spectroscopy
Raman spectroscopy is a branch of vibrational spectroscopy, complementary to that of
IR spectroscopy. Though both IR absorption and Raman scattering induce transitions within
vibrational levels, they produce distinct spectra and cannot be used as alternative techniques.
While IR absorption depends on a change in the intrinsic dipole moment with molecular
vibration, Raman scattering relies on changes in the polarizability of functional groups due to
vibration of the atoms (Li-Chan, 1996). Therefore, non-polar groups such as C=C, S-S and C-
C have dominant Raman bands compared to polar groups (O-H, N-H and C=O) which exhibit
30
strong IR stretching vibrations.
Water being a polar compound exhibits a strong IR absorption that interferes with the
sample spectra as most food systems are aqueous in nature. Therefore, Raman spectroscopy is
better suited for such systems as it produces less interference. The phase behaviour of
gelatin/maltodextrin composite has been studied using Raman spectroscopy in conjunction
with multivariate curve resolution which enabled determination of individual polymer
concentration in their respective domains (Normand, Pudney, Aymard, & Norton, 2000;
Pudney, Hancewicz, Cunningham, & Gray, 2003). Haug, Williams, Lundin, Smidsrød, &
Draget (2003) probed the phase behaviour of fish gelatin and κ-carrageenan composites using
Raman spectroscopy.
Being a scattering technique, samples of any shape, size or thickness can be analysed,
down to the microscopic range of 10 microns. Samples can be measured in-situ in the solid,
liquid or gaseous state without any sample preparation. The many advantages make Raman
spectroscopy a promising technique in spite of its limited application to materials that
autofluoresce. Sample heating due to radiations can destroy the sample or mask the Raman
spectrum.
1.4.5.4. NMR spectroscopy
Nuclear magnetic resonance (NMR) utilizes nuclear spin states and resonance
spectroscopy for spectroscopic analysis. It is widely used for the structure determination of
biopolymers and their composites. The method utilizes spectral lines of different atomic nuclei
which are excited in a strong magnetic field when a radiofrequency transmitter is applied.
Changes in the 3D structure of molecules can be determined as neighbouring atoms influence
31
the signals from individual nuclei.
It is a non-destructive technique that allows visualization of atoms and molecules in a
solid as well as in various media in solution. NMR spectroscopy uses the lowest irradiation
energy for excitation within the spectroscopic techniques and hence, relaxation and sensitivity
of NMR spectroscopy are significantly different, enabling observation of molecular dynamics.
In the works of Voron’ko, Derkach, Vovk, & Tolstoy (2016), NMR spectroscopy was used to
study the various bonds involved in the formation of κ-carrageenan-gelatin complexes. The
results elucidated that while electrostatic interactions and hydrogen bonds were at play within
the temperature range of the coil to helix transition, the complexes were stabilised by
hydrophobic interactions at higher temperatures. They further expanded their work to study the
kinetics of gel formation using NMR in aqueous mixtures of κ-carrageenan and gelatin and
documented a decrease in the gelation rate as the κ-carrageenan/gelatin ratio increased. NMR
has also been successfully used to study the gelation mechanism of curdlan (Zhang, Nishinari,
Williams, Foster, & Norton, 2002), konjac mannan (Williams et al., 2000) and agarose (Dai &
Matsukawa, 2013) at the molecular level.
1.4.5.5. Ultrasonic spectroscopy
Ultrasonic spectroscopy studies ultrasonic waves resolved into their Fourier frequency
components. As an analytical technique, it is based on precise measurements of structural (size,
shape, orientation, etc) or material (attenuation, velocity, etc) properties. Ultrasound, by
definition, has a frequency greater than 20 kHz and in an ultrasonic wave, the oscillating
pressure (which is stress in general terms) causes oscillation of compressions (mechanical
deformation), thus acting like a rheological wave. Therefore, simplistically ultrasonic
32
parameters deal with elasticity and viscosity. However, ultrasound involves high frequency
deformations as opposed to the low frequency deformations (below 1 kHz) involved in classical
rheology.
Ultrasonic spectroscopy deals with elastic properties of materials in contrast to other
spectroscopic techniques that probe the electric and magnetic properties. Compression in
ultrasonic wave alters the intermolecular distances in a sample causing intermolecular
repulsions as a response. Due to the elasticity of water, sound waves travel up to five times
faster in water than in air making it highly sensitive to intermolecular interactions. With the
high sensitivity of the ultrasonic parameters to intermolecular interaction, ultrasonic analysis
of a wide range of molecular processes in aqueous media are made possible.
The measurable parameters of interest are the speed of sound v and attenuation α which, when
assuming a Newtonian medium with minimal thermo-acoustic losses, are directly related to the
complex longitudinal elastic modulus:
(16) M = Mʹ +iMʹʹ
Where Mʹ = ρv2 and Mʹʹ = (2ραv3)/ω and ω is the angular frequency of the wave and ρ is the
material density. The complex longitudinal elastic modulus is related to the bulk K and shear
moduli G by:
(17) M = K + 4G/3
Ultrasound has been successfully used to study biopolymer gels and sol-gel transitions in the
works of Audebrand, Doublier, Durand, & Emery, (1995) and Toubal, Nongaillard,
Radziszewski, Boulenguer, & Langendorff, (2003). It is also gaining popularity for the
investigation of phase behaviour in mixed biopolymer systems such as gum arabic and
modified starch (Chanamai & McClements, 2001), cellulose and amylose gels (Clark, 1996)
and dairy proteins (Chappellaz, Alexander, & Corredig, 2010; Corredig, Sharafbafi, & Kristo,
33
2011).
1.4.5.6. Diffusing wave spectroscopy
Recently, diffusing wave spectroscopy (DWS) has gained substantial attention as a light
scattering technique used to measure the structural, viscoelastic, and dynamic properties of
turbid colloidal samples. DWS utilises the multiple scattering events that occur when light
passes through a colloidal system to quantify its in-situ dynamics. The diffusion equation can
be used to approximate the propagation of light in such a system. Similar to any light scattering
technique, DWS measures the intensity fluctuations of scattered light occurring due to the
Brownian motion of colloidal particles. By treating the propagation of photons through turbid
samples as a process of diffusion, it makes it possible to extract the dynamics of the scattering
particles from measured correlation functions. DWS is a fast, reproducible, non-destructive
technique that doesn’t necessitate sample dilution unlike the conventional dynamic light
scattering techniques. DWS has been successfully used to probe the phase behaviour of whey
protein and gum arabic (Weinbreck, 2004), β-lactoglobulin and dextran (Blijdenstein, 2003)
and various emulsions (Gancz, 2005; Liu, 2007)
1.4.5.7. Electron spin resonance spectroscopy
Electron spin resonance (ESR) spectroscopy is a technique used for probing materials
with unpaired electrons. Electron spin resonance spectroscopy is similar to nuclear magnetic
resonance with the key difference that in electron spin resonance, the electron spins are excited
due to the absorption of radiation in the microwave radiation instead of spins of atomic nuclei
as in nuclear magnetic resonance. Microwave radiations are absorbed in the presence of an
external magnetic field. Given that electrons have a greater magnetic moment than nuclei, ESR
spectroscopy has a greater sensitivity than NMR spectroscopy. It is also more specific, which
can also be a disadvantage as most biopolymers are not paramagnetic (having unpaired
34
electrons). Sample volumes as low as 300 µL and concentrations as low as 1 µM can be
analysed. Phase behaviour of micellar casein and κ-carrageenan systems and κ-carrageenan
and konjac mannan mixtures have been investigated using ESR spectroscopy in the past
(Schorsch, Jones, & Norton, 2000; Williams, Clegg, Langdon, Nishinari, & Piculell, 1993)
1.4.6. Diffraction and scattering methods
A classic example of the interaction between radiation and matter is the scattering of
light as it strikes particles. This phenomenon initiated a number of theoretical approaches, out
of which the Rayleigh theory, that pursues the scattering of light by particles substantially
smaller than incident wavelength, has been adapted and most widely applied to the realm of
biopolymer science as the Rayleigh-Gans-Debye (RGD) theory. Most biopolymer mixtures
behave as ‘water-in-water’ emulsions and can be modelled as particles in a solvent. This is
made possible as the biopolymer forming the continuous phase acts like the solvent, and the
dispersed phase, which often forms spherical droplets to reduce interfacial energy acts as the
dispersed particles. As both the phases are water rich, with biopolymers being inherently weak
scatterers, they manifest a relatively small difference in the refractive index (Aymard,
Williams, Clark, & Norton, 2000).
A number of scattering techniques such as small-angle neutron scattering (SANS),
small-angle light scattering (SALS), and X-ray diffraction have been used to characterise phase
behaviour in biopolymer mixtures. The underlying principle of all these techniques is very
similar, with the main difference being the wavelengths of the radiation impacting the
investigation of local structure. As a general thumb rule, short wavelengths of X rays (λ =
0.151 nm) and neutrons (λ = 1.0-2.5 nm) are better than the relatively longer wavelengths of
35
visible light (λ = 350-800 nm). A combination of the three scattering techniques can provide
an exhaustive analysis of the morphology, but such comprehensive studies are rare in
biopolymer science.
SANS exhibits a strong angular dependence for the scattered intensity, denoting sharp
interfaces between the continuous and dispersed phase (Tuinier, Ten Grotenhuis, Holt,
Timmins, & De Kruif, 1999). Time-resolved SALS can help distinguish between the spinodal
and nucleation growth mechanisms (Tromp, Rennie, & Jones, 1995) while turbidity
measurements too can help detect the onset of phase separation.
1.4.6.1 Light scattering
According to Debye (1959) and Smoluchowski (1912), in a solution, scattered light
intensity arises from fluctuations in concentration and density. For biopolymer solutions,
fluctuations in the density are much smaller than the fluctuations in concentration even in dilute
solutions and can be approximated by the density fluctuations in pure solvents. Therefore, in
solutions where the radius of gyration, Rg < 1/20 λ, the Rayleigh ratio or Rθ (normalized
scattering intensity) is given by:
(18) Rθ = K (Δc2)
Where Rθ = (iθ /Io) r2 and, Io is the initial beam intensity, iθ is the corresponding intensity of the
scattered beam at angle θ, r is the distance between the detector and the centre of the scattering
volume, and K is the constant of optical contrast. Thus, at low concentrations, the intensity of
the scattered light is directly proportional to concentration c, but decreases eventually due to
the repulsive interactions, which lead to a decrease in osmotic compressibility with an increase
in concentration. This enable differentiation between molecular structures at higher
concentrations. Scattering intensity in such systems is given by:
36
(19) Rθ = KcRT(dc/dπ)
Where RT(dc/dπ) is the osmotic compressibility and π the osmotic pressure (Burchard, 1994).
This principle of light scattering has been used to estimate parameters like interfacial tension
and turbidity, the changes in which can be linked to phase behaviour as discussed in the latter
sections.
Puyvelde et al. (2002) used SALS combined with shear rheometry to measure the
interfacial tension in dextran-gelatin composites. Under shear, the composite yields an
anisotropic SALS pattern, where the degree of anisotropy relates to droplet deformation. The
interfacial tension can be determined from the deformation of the droplet under shear, and from
droplet relaxation after cessation of shear. This technique is also applicable to the theory of
fibril break up set forth by Tomotika (1935), which follows the break-up of fibrils formed under
shear, when shear is halted. Antonov, Puyvelde, & Moldenaers (2004) in their work used SALS
to study the interfacial tension and subsequent phase behaviour of water-sodium caseinate-
sodium alginate systems located at different distances from the binodal. They revealed that
close to the critical point, interfacial tension is much lower than for concentrations farther off
from it.
1.4.6.2 X-ray diffraction
Compared to the huge contribution of X-ray diffraction to conventional structural
chemistry, its application to biopolymers is somewhat confined to molecular biology. The less
direct link between macroscopic behaviour, characteristic of food materials, and detailed
molecular features limit the application of X-rays to biopolymer science.
When a beam of X-ray passes from a material, a small fraction of the radiation is
scattered due to the inhomogeneities in electron density, and the variations in the scattered
37
intensity and the direction of scatter are used to draw conclusions on the material’s microscopic
structure. A simplifying feature, or a limiting factor of the X-ray technique is that fact that
only electrons present scatter the radiation as opposed to both electrons and neutrons for other
radiations. Even out of the small fraction of X-ray diffraction studies that focus on biopolymer
investigations, a smaller fraction deals with gels and solutions, with amorphous, crystalline
powders and fibre being a popular choice.
At more hydrated conditions, in solutions and gels, systems are less ordered in structure
and hence produce a diffused scattering pattern. Miles et al. (1985) in their work successfully
utilized X-ray diffraction and turbidity analysis to describe the gelation of amylose solutions.
The changes in the scattering intensity was studied as gelation commenced. The rate of
crystallinity development was much slower than the time of experimentation for turbidimetric
and rheometric analysis (1000 min vs 100 and 300 min, respectively). It was concluded that
concentration of amylose into dense light scattering domains, leading to network formation
occurred in the early stages accompanied by an increase in shear modulus. On a much slower
time scale the onset of crystallinity was observed, with insignificant contribution to modulus
increase. Other examples of materials investigated by X-ray diffraction include wheat and corn
starch gels (Xu et al., 2013), amylose-amylopectin mixtures (Leloup, Colonna, & Buleon,
1991), agarose (Foord & Atkins, 1989), and carrageenan/ konjac glucomannan compound gel
(Yuan, Xu, Cui, & Wang, 2019).
1.4.6.3. SAXS and SANS
Investigating properties and characteristics of a food product are best done while
maintaining their native surrounding environment. In the given sense, small angle scattering
possesses attractive attributes as a non-invasive technique, enabling analysis in the solution or
38
gel state. Due to the inverse relationship between the scattering angle and scale lengths
investigated, small angle scattering makes it possible to investigate proteins, emulsions,
porosity and phase separation, to name a few. At these larger scales, the technique is sensitive
to the arrangement of assemblies of atoms rather than atomic separation in materials enabling
the determination of the molecular structures and spatial skills. Consequently, small angle
scattering (SAS) techniques, either with X-rays (SAXS) or with neutrons (SANS) have been
widely used in biopolymer systems. The theory behind SAS has been covered in detail in
literature (Hammouda, 2010; King, 1999).
A SAS experiment typically measures the intensity of scattered radiation versus the
scattering vector q, which is given by:
(20) q = (4π/λ) sinθ
where λ is the wavelength of the incident radiation, and θ is the scattering angle.
The scattering length in SAXS is determined by the electronic structure of the target
atom. As the number of electrons in an atom is equal to the number of protons, the intensity of
X-ray scattering increases linearly with atomic number. Naturally then, the heavier elements
in the material dominate the X-ray scattering signal. In contrast, in SANS, neutrons are
scattered by the atomic nucleus, which means that scattering can vary significantly between
isotopes.
In their work, Horkay & Hammouda (2008), studied the effect of temperature, nature
of solvent and charge in solutions of poly-lysine and poly-proline. In semi dilute solutions of
polyaspartic acid, chondroitin sulfate hyaluronic acid and DNA, addition of salts result in
structural changes around the polymer chains. Xu et al. (2018) used SAXS and SANS to
39
characterise the microstructure of the β-lactoglobulin-pectin complex coacervates.
1.4.7. Calorimetry
Even though considerable work was done on the thermodynamic properties of polymer
solutions and blends, the fundamental approach to the thermodynamics of polymeric solutions
appears to have been neglected till the 1990s, with the works of Tseretely and Smirnova (1992)
investigating the melting enthalpy of gelatin being the first of its kind. Flory-Huggins lattice
model (Flory, 1953), along with its generalisation (Doi, Edwards, & Edwards, 1988) had been
the broadly accepted thermodynamic theory used for analysis of experimental data. However,
in case of composite systems, studies seldom split excess Gibbs free energy (parameter χ) to
the constituent enthalpy (χS) and entropy (χH) contributions (Cesàro, Cuppo, Fabri, & Sussich,
1999). This was probably due to the late adoption of calorimeters in the field of biochemistry
making it difficult to accurately determine the enthalpy of biopolymer phase transitions (Wadsö
& Chemistry, 1980). Detailed reviews on the development of calorimeters, their working
principle and their application in biopolymer science has been documented in literature
(Leharne, 2017; Chiu, Prenner, & Sciences, 2011).
Differential scanning calorimetry (DSC) is a physical characterization technique used
to study the purity, thermal behaviour and stability of polymers and polymer blends. First order
transitions like gelation and melting, along with second order transitions like glass transition
are characteristic properties of biopolymers, impacted by the presence of and interaction with
other biopolymers and additives, along with different processing techniques employed
(Gregorova & Microcalorimetry, 2013). Simplistically, a calorimeter is used to measure the
specific heat capacity of thermally induced events in biopolymer systems as a function of
temperature. The apparent specific heat (cp2) of a biopolymer solution is calculated by:
40
(21) cp2 = cp1 + 1/w2(cp-cp1)
where cp is the solution’s specific heat, cp1 is the specific heat of the solvent, and w2 is the
weight fraction of the solute. Principally, in a calorimeter, the reference and the sample are
maintained at the same temperature, while the changes in the energy during a thermal transition
are recorded by the DSC as a rate dQ/dt against temperature. This sudden change in the value
𝑑𝑄
of specific heat Cp corresponds with the transition temperature as:
𝑑𝑇
(22) = 𝑚𝐶𝑝
Where m is the mass of the sample. First order transitions can be described using the following
𝑑𝑄
formula:
𝑑𝑡
𝑑𝑡
(23) = 𝜅𝛥𝑇 = 𝜅Ṫ(𝑡 − 𝑡0) + 𝑑𝑄 |𝑡0
Where κ is the thermal conductance between the sample holder and the sample, Ṫ is the rate of
increase in temperature and t0 the start of transition.
For the analysis of calorimetric results, the most widely accepted model is the van’t Hoff
equation:
(24) 𝛥𝐻° = 𝑅𝑇2 𝑑 ln 𝐾 𝑑𝑇
Where K is the equilibrium constant, ΔH° is the enthalpy change, R is the ideal gas constant
and T is temperature. This equation is applicable to two state processes which do not have
significant intermediate populations during transition, protein denaturation for example.
Change in entropy, ΔS° and change in standard free energy ΔG° are obtained using the
following equations:
(25) 𝛥𝐺° = −𝑅𝑇 ln 𝐾
41
(26) 𝛥𝐺° = 𝛥𝐻° − 𝑇𝛥𝑆°
Calorimetry can provide not only the calorimetric enthalpy ΔH (from the area of the
heat absorption), but also the process’ effective enthalpy using the van’t Hoff equation and the
sharpness of the transition. It can hence be used to observe the impact of a second polymer on
the thermal transitions of the first biopolymer and vice versa both qualitatively and
quantitatively.
Fig.1.4. DSC exotherms for a) 2% agarose, b) 15% gelatin, c) 2% agarose plus 15% gelatin, d)
Dalda vanaspati lipid, e) soyabean oil, and f) 2% agarose plus 15% gelatin plus 2.5% Dalda
vanaspasti lipid. Scan rate 1°C/min.
Shrinivas et al. (2009) in their work used the DSC to study the phase behaviour of
binary mixtures of gelatin and agarose, and the ternary mixtures of agarose, gelatin, and a lipid.
Peaks in Figures 1.4 a and 1.4 b depict the coil to helix transition of single agarose and gelatin
42
solutions, respectively, Fig.1.4 e denotes the crystalline aggregation and lattice formation of
dalda vanaspati, a fat solid at room temperature, as opposed to a featureless baseline heat flow
of soybean oil reflecting that it remains in the liquid state throughout the experimental
temperature range (Fig. 1.4 e). Binary mixtures of agarose and gelatin (Fig. 1.4 c) and ternary
mixtures of agarose/gelatin/lipid (Fig. 1.4 f) maintain the overall peak form of each of the
constituent transition revealing that they remain phase separated without interacting with each
other in a mixture.
1.4.8. Rheology
1.4.8.1. Mechanical properties of the composite
When either one or both of the biopolymers in a biopolymer mixture can form a gel,
the process of phase separation is arrested by network formation. If just one biopolymer can
form a gel, it forms a continuous network in which the second biopolymer remains dispersed.
In a scenario where both the biopolymers are gelling agents, the first component will form a
continuous network, with the solution of the second polymer permeating it. The gelation of the
second polymer may follow in one of three ways - the first component remains as a continuous
network within which the second biopolymer forms a dispersed gel network; in the case the
second biopolymer is a stronger gelling agent, it disrupts the primary network and forms a
continuous network supporting the broken network as a dispersed phase; the third possible
outcome is that the primary network remains intact and the second component forms another
continuous network through the interstices of the primary network leading to a bi-continuous
network (Morris, 2009). Various factors can influence the onset of gelation of one or both of
the polymers and the level of phase separation attained before the structure is trapped by
network formation. This has been discussed in detail in the works of Kasapis (2008) and Norton
& Frith (2001). Biopolymer gelation along with evolution of phase behaviour can thus be used
43
to create a plethora of phase structures, allowing the manipulation of the composite’s final
properties (Firoozmand & Rousseau, 2015). The phase structure of a given composite acutely
determines its bulk mechanical properties.
The rheological approach of determining phase behaviour correlates the mechanical
properties of a simple binary mixture to the corresponding properties of its constituents by a
set of theoretical blending laws (Frith, 2010). The different parameters that account for the
gel’s mechanical properties and the techniques used for characterising them are outlined in the
works of Picout & Ross-Murphy (2003) and Clark & Ross-Murphy (2009).
Adapted from the realm of polymer science, the Takayanagi blending laws state that in
a biopolymer composite where the individual components X and Y have a low miscibility, the
composite modulus (Gc) can be estimated by considering the extreme cases of strain and stress
distribution within the composite as follows (Takayanagi, 1963):
(27) 𝐺𝑐 = ɸ𝑋𝐺𝑋 + ɸ𝑌𝐺𝑌
and
(28) 𝐺𝑐 = (ɸ𝑋 𝐺𝑋⁄ + ɸ𝑌 𝐺𝑌⁄ )−1
Where GX and GY are the storage modulus and ɸX and ɸY (ɸX + ɸY = 1) are the phase volumes
of the polymers X and Y. Eq. 27 refers to the upper bound or isostrain condition where the
stronger gel forms continuous matrix supporting the weaker dispersed phase and the entire
composite has a uniform strain. Eq. 28 on the other hand applies to isostress conditions where
the weaker matrix supports a more rigid filler and the composite is subjected to the same stress.
Davies in his works (Davies, 1971a, 1971b) developed the following equation that
could be applied to phase separated biopolymer composites where the constituents formed a
bi-continuous gel network:
5⁄
5⁄ = ɸ𝑋(𝐺𝑋)1
5⁄ + ɸ𝑌(𝐺𝑌)1
44
(29) (𝐺𝑐)1
Since the majority of the biopolymer composites include a large amount of solvent
(water) with a relatively smaller concentration of the biopolymers, it introduces another
complication in accurately determining the phase behaviour of a three component system
(water-polymer X-polymer Y) - variable phase volume (Morris, 1992). Due to the varying
avidity of the two active gelling biopolymers to attract water, the total amount of water in the
system is split between the two in unequal quantities. This influences the values of ɸx and ɸy
needed in Eqs. 27, 28 and 29. The modulus-concentration relationship is also affected as the
effective biopolymer concentration after phase separation is much higher than the initial
concentration. This was overcome by introducing the ‘solvent avidity parameter’, p as follows
(Clark, Richardson, Ross-Murphy, & Stubbs, 1983):
𝑤𝑋 𝑤𝑌 (30) 𝑝 = ( ⁄ 𝑋⁄ ) ( 𝑌⁄ )
Eq. 30 can be used in systems where the weights of the two biopolymers (X and Y) and
the water content of both (w = wX + wY) is known, in order to determine p, the phase volumes
of the respective phases and subsequently their effective concentration in the separated phases.
Numerous studies have documented the successful application of the blending laws in phase
behaviour estimation in biopolymers (Dekkers, Emin, Boom, & van der Goot, 2018; Hassan
Firoozmand & Rousseau, 2013; Mhaske, Condict, Dokouhaki, Katopo, & Kasapis, 2019).
The original blending laws are applicable for composites where one biopolymer formed
a continuous network and the other remained dispersed as spherical inclusions. This is true in
most scenarios as the dispersed phase tends to form isotropic spheroidal fillers due to the drive
to reduce the interfacial tension between the two phases (Ding et al., 2002). However, when
the spherical inclusions aggregate to form fibre or oriented particles, or when the system
comprises of biopolymer fillers whose inherent structures aren’t spherical, the application of
45
the original blending laws cannot accurately probe the structural properties of these systems.
Kerner (1956a, 1956b), followed by Halpin (1976), were the first to introduce a method that
took into account the shape of the filler as follows:
(31) Gʹ / GʹX = (1 + ABɸY) / (1 - BɸY)
(32) B = [(GʹY/GʹX) – 1]/[(GʹY/GʹX) + A]
Where A is related to Einstein co-efficient kE (A = kE -1), sensitive to phase topology of the
composite and takes into account the shape of the dispersed inclusions and Poisson’s ratio of
the matrix (Kasapis, 2009).
Lewis and Nielsen further refined the blending laws by taking into consideration the
concept of maximum packing fraction of the dispersed phase, ɸm (Lewis & Nielsen, 1970;
Nielsen, 1970):
(33) Gʹc/ GʹX = (1 + ABɸY)/(1-BψɸY)
2]ɸY
(34) Ψ = 1 + [(1 - ɸm)/ɸm
ɸm relates to the shape of the filler particles and its degree of polydispersity (Mhaske, Condict,
Dokouhaki, Farahnaky, & Kasapis, 2020).
When shear is applied to the composite gels, it often permanently orients the filler
particles fibrous filler particles and enhances the network strength. Using the single parameter
exponential equation, the filler orientation distribution can be defined as (Kacir, Narkis, &
Ishai, 1975):
(35) γ = 1 – e-λα
Where λ is the empirical parameter, correlated to material architecture or operational
parameters, α is the orientation angle between 0 and 90 (π/2) degrees and γ is the percent of
46
filler particles aligned between 0 and ± α degrees (Chen & Cheng, 1996). Using a simplified
version of Eq. 35 Koh & Kasapis (2011) further modified Eq. 33 to take into account the
angular orientation of aligned filler particles in the composite as follows:
(36) Gʹc/ GʹX = [(1 + ABɸY)/(1-BψɸY)]eλ
When λ = 0, Eq. (11) will default to the original Lewis and Nielsen model.
The biopolymer structure-rheology relationship in co-relation with phase behaviour has
been studied in detail in multiple other publications (Chronakis & Kasapis, 1995; Clark &
Ross-Murphy, 1987; Stefan Kasapis, 2008, 2009; S. Kasapis & Bannikova, 2017; Kasapis,
Morris, Norton, & Clark, 1993; E. Morris, 2000; Richardson & Kasapis, 1998). Though the
application of the blending laws is well established, it is as such, an indirect approach of
probing phase behaviour that is time consuming, involving tedious experimentation and
substantial modelling expertise. Furthermore, the blending laws are based on certain
assumptions that may not always hold true for the system for example, a) polymer confinement
to individual phases, b) the modulus is proportional to the square of the polymer concentration
and the volume of polymer network is negligible when compared to volume of the solvent and
c) there is massive increase in effective molecular weight at the onset of gelation.
1.4.8.2. Interfacial tension between phases
For phase separation to take place in aqueous polymer mixtures, the driving force of
phase separation must be greater than the increase in interfacial free energy which is the product
of the total interfacial area and interfacial tension between the two phases (Antonov et al.,
2004). This makes interfacial tension between the coexisting phases of a phase separated
47
system a property of interest.
The onset of phase separation is marked by the formation of droplet structures in the
spatial distribution of the two polymers. In the absence of external dynamic effects such as
shear or applied flow field, the phase with the greater volume fraction forms the continuous
phase. Since the microstructure formation is spherical, it is evident that interfacial tension is at
play. It is worth noting however that the interfacial tension is extremely low, typically in the
range of 0.01-10 µN/m (Vis et al., 2015). In addition, the liquid-liquid interface is permeable
to solvent and small ions and the interplay between phase separation of the charged polymers
leads to an electrical potential difference between the two phases which further reduces the
interfacial tension. The effect of biopolymer concentration and molecular weight of the
biopolymers on the interfacial tension has been investigated by a number of researchers (Ding
et al., 2002; Guido, Simeone, & Alfani, 2002; Spyropoulos et al., 2008; Puyvelde et al., 2002).
Bamberger, Seaman, Sharp, & Brooks (1984) reported that interfacial tension could be
correlated either with the length of the tie line (TLL) in the phase diagram:
(37) 𝜎 = 𝐶1 . (𝑇𝐿𝐿)𝐶2
Or with the difference in the polymer concentrations in the separated phases:
𝐶2
(38) 𝜎 = 𝐶1 . (𝛥𝑋𝑝𝑜𝑙𝑦𝑚𝑒𝑟)
Where C1, C2 are experimental constants related to the biopolymer, temperature, concentration,
and molecular weight.
A number of techniques can be used to measure the interfacial tension and are classified
into three broad groups (Ding et al., 2002): measurement of interfacial tension from the balance
between surface forces and gravitational forces; measuring the equilibrium size and shape of
the droplets; dynamic measurements of deformation and breakage of thin threads of one phase
48
(Scholten, Tuinier, Tromp, & Lekkerkerker, 2002), steady-state shape of droplets in a
controlled flow field (Sigillo, Di Santo, Guido, & Grizzuti, 1997) or retraction of deformed
drop after cessation of flow (Guido & Villone, 1999).
Selection of the measurement technique depends on the physical properties of both the
phases, mainly their viscosity and density. Since the density of the phases is quite similar and
at least one or both the phases are highly viscous using the techniques of the first two categories
is either impossible or highly unreliable. Therefore, techniques of dynamic deformation are
most widely used to estimate interfacial tension (Stokes, Wolf, & Frith, 2001). The interfacial
tension can be directly correlated to the complex modulus of the composite using Palierne’s
model as follows (Palierne, 1990, 1991):
∗ (1+3∑𝑖ɸ𝑖𝐻𝑖 1−2∑𝑖ɸ𝑖𝐻𝑖
(39) ) 𝐺 ∗ = 𝐺𝐶
Where Hi = fn(α/Ri, Gc*, Gd*). Here Hi is a function of frequency (ω), given for each particle
size Ri. Gc* and Gd* are the complex modulus of the continuous and the filler phase,
respectively. ɸi is the phase volume for a monodispersed emulsion and α is the interfacial
tension between the two phases.
Alternatively, steady state deformation of droplets as result of applied flow field can be
easily investigated in a mixture at very low shear rates. The relationship between the steady
5(19𝑝+16)
state deformation of the droplet in a shear flow is given by (Cox, 1969):
4(1+𝑝)√(19𝑝)2+ (20/𝐶𝑎)2
(40) 𝐷 =
Where p is the ratio of the viscosity of the internal phase to that of the external phase and Ca
is the capillary number given by (Tucker III & Moldenaers, 2002):
49
(41) 𝐶𝑎 = 𝜂𝑐𝑅𝛾 𝜎
Where ηc is the viscosity of the external phase, Rγ is the radius of the droplet under shear and σ
is the interfacial tension.
Fig.1.5. A drop of caseinate in alginate rich phase at rest (1), start-up transient of deformation under
steady shear (2,3), steady state shape (4), and relaxation after shear is stopped (5,6). Graph represents
the deformation parameter as a function of shear rate in comparison with theory (Guido et al. 2002)
Fig.1.5 shows the results of such an experiment extracted from the works of Guido et
al. (2002) and demonstrates that low shear rates are capable of producing significant
deformations in a caseinate-alginate mixture. This method is combined with microscopy or
small angle light scattering (SALS) to visualise/characterise droplet deformation and
reformation and have been explained in detail elsewhere in the review. The role of interfacial
tension in the phase behaviour of mixed biopolymer systems has been explained in detail in
the works of Frith, (2010).
1.4.9. Microscopy
The different microscopy techniques used to probe behaviour in biopolymer composites
50
have been summarised in Table 1.3 and discussed in detail below:
1.4.9.1. Optical microscopy
The late nineteenth century saw a rapid advancement in the field of microscopy
enabling researchers to probe specimens from a multitude of disciplines in the micron and
submicron level. Optical microscopy being the simplest form of microscopy was the most
accessible and hence most popular. Early microscopists were limited by poor lens design,
blurred images and optical aberration which have been successfully overcome with
technological advancements. In its simplest form, an optical microscope consists of an
objective lens placed at an adjustable distance from the object, known as the focal length. Light
rays reflecting from the object, pass through the objective to form an inverted aerial
intermediate image at a fixed distance within the optical tube. This is further magnified by the
microscope eyepiece that produces an erect image. The total magnification of the microscope
is equal to the magnification of the objective times that of the eyepiece (Mertz, 2019).
While imaging a mixed biopolymer system, light from the microscope lamp reaches
the sample via a condenser. Some of this light passes through the sample undeviated, whilst
some of it gets diffracted due to the differences in density and refractive indices of the
constituent biopolymers within the sample. The diffracted light is rendered 180 degrees or half
a wavelength out of phase with the undeviated light and causes destructive interference with
the direct light at the intermediate image plane. The eye piece lens further magnifies the image
which is finally projected onto the camera film, light-sensitive computer chip or the retina. The
objective projects the undeviated light, spreading it evenly throughout the image plane at the
eyepiece’s diaphragm. The diffracted light is brought to focus at localized places on the same
image plane where it causes a destructive interference reducing intensity resulting in lighter or
darker areas. These lighter or darker patterns are what form the image of the sample (Davidson,
51
Abramowitz, 2002). Optical microscopy is a quick, easy technique that does not require much
sample preparation. However, it is suitable only for dark imaging of samples that have a
significant density difference.
Freitas, Albano, & Telis (2017) in their work used the optical microscope to study the
complexing of soy protein isolate (SPI) and high methoxyl pectin (HM) at varying pH and at
different proportions. Capron, Costeux, & Djabourov (2001) studied the phase separated
alginate and sodium caseinate water in water emulsion using the optical microscope.
1.4.9.2. Phase contrast microscopy
Phase contrast microscopy is based on the advances in optical principles by Frits
Zernike introduced before the World War II, but the technique did not become universally
recognised till the 1950s. Frits Zernike’s study revealed that amplitude and phase differences
between deviated light and zeroth order could be altered to enhance contrast and interference
(Zernike, 1942a, 1942b). As mentioned in the previous section, refracted waves that are ½
wavelength or 180 degrees out of phase exhibit a good contrast in an optical image, but samples
with a lower difference in refractive indices cause a lag of just ¼ wavelength, which isn’t
enough to decrease its intensity. This creates an image without contrast, making the details
obscured. Zernike proposed a method to speed up the undeviated light by ¼ wavelength to
induce a difference of ½ wavelength between the undeviated and deviated light. This would
then subsequently lead to the formation of a dark image against a light background as described
earlier (Pluta & Maksymilian, 1988). Another less popular option is to slow down the direct
light by ¼ wavelength, which results in both the diffracted and direct light reaching the
eyepiece in step causing a constructive interference. This creates a bright image on a dark
52
background (Abramowitz, 1987).
Table 1.3: Microscopic techniques used for structural characterisation of phase behaviour.
Principle
Advantages
Disadvantages
Type of Microscope
Radiation type
Photons
Optical
Simplest of the microscopes to
Have a lower
use
magnification compared to other microscopes
Light reflected/ transmitted through the sample passes through a number of glass lenses that magnify the object
Uses visible light Affordable
Lower resolution compared to other microscopes Internal structures can’t be viewed unless samples are fixed
Cannot operate in
darkness
Electrons
Electron
Has a high magnification and
resolution
Samples need to be under vacuum in most cases Most samples need to be
Quick Capable of generating three-
coated with a non- conducting material
The sample is scanned using a beam of electron. The energy lost by the electrons as it interacts with the specimen is used as a signal to form an image.
dimensional and topographical images
Excess charge can be carried over during imaging and affect the image quality Cannot image live
samples
Only generates black and
white images
53
Photons
Phase contrast
Can observe samples in-situ in
Not suitable for thick
natural state
Differences in refractive indices used to provide contrast in constituents
samples Images usually have a halo effect
No sample preparation required High contrast, high resolution
images obtained
Laser
Confocal
Allows identification of multiple
components as long as the dyes used to stain them have different excitation and emission wavelengths
Depth of view limited to a few hundred microns Smaller field of view compared to other microscopic techniques
High resolution images can be
Staining required
Laser is used to fluoresce only one plane of sample; emitted light is captured through a pinhole to block out-of-focus photons; stepping feature allows to change focal plane along the Z axis
obtained rapidly
Non-invasive 3D sectioning made
possible
Laser
Fluorescence
Capable of generating high
Samples need to be
Instead of transmitted/ reflected light, natural or induced fluorescence is captured
stained in order to be observed
contrast, high resolution images Capable of using and visualising
multiple fluorophores simultaneously
Not best suited for thick samples and cross- sections
Atomic force
N/A (Physical cantilever used)
Being a new technique, it needs to be standardised for widespread application
A cantilever is scanned across the sample surface, changes in the position of the cantilever induced by surface morphology is recorded
Possibility of damaging
High scan speed High resolution Minimal or no sample preparation needed Can be paired with other
soft samples
techniques to give a holistic picture
Can be used at extreme
temperatures
Can be operated in air or in a
liquid
54
To speed up the undeviated zeroth order of light, a phase plate with a ring-shaped phase
shifter is installed to objective’s rear focal plane. The phase plate is optically much thicker than
the ring’s narrow area. Consequently, while traversing the objective, direct light passing
through the phase ring travels a shorter distance than the diffracted light. This causes the
deviated and direct light to be ½ wavelength out of phase when they reach the image plane,
where they interfere destructively and creating a dark image on a light background. To slow
down the speed of direct light, the phase shifter area of the plate is made optically thicker than
the plate. In this scenarios, direct light reaches the image plane with the diffracted light
inducing constructive interference, creating a bright image on a darker background (Delly,
1988).
Fig.1.6: Kappa carrageenan (A, B) and Iota carrageenan (C, D) in mixture with meat proteins
at pH 5.6 and 7.1, respectively. Black arrows indicate carrageenan molecules with attached
meat proteins.
Farouk, Frost, Krsinic & Wu (2011) used phase contrast microscopy to study the phase
behaviour of kappa and iota carrageenan in mixture with meat protein (Fig.1.6). Phase contract
55
microscopy has also been successfully used in the works of Bourriot, Garnier & Doublier
(1999) to study the phase behaviour of κ-carrageenan- micellar casein mixtures using different
ionic forms of carrageenan (Na+, K+).
1.4.9.3. Atomic force microscopy
Also known as scanning-force microscopy, atomic force microscopy (AFM) is a type
of scanning probe microscope technique. Originally designed to characterise surfaces at the
atomic scale, AFM allows imaging of solutions and soft composites at molecular scale with a
resolution of <1 nm.
Disregarding the conventional notions of microscopy, AFM collects data for images by
‘feeling’ the samples, making it a physical method of imaging (Morris, Kirby, & Gunning,
1999). A cantilever probe is run over the sample surface and the changes in the magnitude of
interaction between the sample surface and the probe (mainly van der Waal’s force) is
measured. A laser beam is focused directly over the tip of the cantilever and reflected on to a
mirror. The mirror further reflects the laser onto a position positive photodiode detector.
As the sample is scanned, the cantilever tip moves in accordance with the surface
topography. Consequently, the angle of the laser reflected from the cantilever changes
according to Hooke’s law, resulting in changes in the position of the laser spot falling on the
photodiode. As a result, the intensity in each quadrant of the photodiode changes producing an
electric signal that quantifies the normal motion of the tip. The cantilever deflections are thus
measured to produce a map of the surface topography on the computer.
Kim et al. (2004) successfully employed AFM to probe the morphology of protein-
DNA oligonucleotide complexes. Whereas Ikeda et al. (2004b) studied the concentration
dependence of gel formation in mixed gellan-xyloglucan systems using AFM revealing that at
0.05% gellan and 0.7% xyloglucan, the constituents exhibited a synergistic interaction forming
56
a single-phase gel. Whilst single systems of gellan and xyloglucan did not gel at concentrations
lower than 0.5 and 0.75%, respectively. A detailed account of the applications of AFM in food
and polymer science can be found elsewhere in literature (Nguyen-Tri, Ghassemin, Carriere,
Assadi, & Nguyen, 2020; H. Yang et al., 2007).
Given the high scan speeds and resolution of AFM, it is helping unveil information that
was thus far inaccessible to researchers. Samples probed using AFM can be studied in their
native state without any processing. AFM can also be modified to study aqueous systems in-
situ. Since it is a relatively new technique, it needs to be standardised to find widespread
application. In addition, as the probe comes in direct contact with the samples, there is a
possibility of it damaging the soft samples and gels. Lastly, since AFM is a physical technique,
it doesn’t help in the chemical characterisation of the samples but recent developments in the
technique like infrared-AFM are looking at overcoming these drawbacks (Magonov &
Instrumentation, 2006).
1.4.9.4. Fluorescent microscopy
Epifluorescence or fluorescence microscopy exploits the capability of fluorochromes
to emit light when excited by light of a specific wavelength and is one of the fastest growing
areas of investigation in microscopy (Winter & Shroff, 2014). Some materials auto-fluoresce
or fluorochromes can be externally added to samples. The fluorescent probe molecules spread
over the sample according to its affinity with the different constituents in the sample and local
accessibility and attach to a biopolymer non-covalently or covalently (Patonay, Salon, Sowell,
& Strekowski, 2004). At a particular wavelength of light, fluorochromes, being rich in
chromophores absorb photons and emit them at a longer wavelength. Multiple dyes can be used
to stain different components as long as the emission wavelength of the dyes is different.
Illumination in normal optical microscopy is attained by widefield illumination where
57
the entire sample, and not just the focal plane is simultaneously and uniformly illuminated.
This necessitates that samples are thin and relatively transparent and often causes an out-of-
focus blur that decreases contrast and resolution of the image (Auty, 2013). To overcome these
drawbacks, the confocal microscope was developed. The essential feature of confocal
microscopy is that both illumination and detection systems are focused on a single point in the
sample. This is brought about by focusing the illuminating light or laser on a specific point in
the sample and the emitted light is detected by means of a small pinhole (aperture) in front of
the detector. The pinhole helps in reducing the out of focus blur giving a sharp image. The
detector amplifies the signal and converts the photon electron intensity into pixels (Davidson
et al., 2002). Confocal microscopy has been extensively used to study food microstructure at
the meso scale and the applications can be found summarised in the works of Cardona, Iriart,
& Herrera (2013) and Auty (2013). Confocal microscopy has major advantages over other
forms of microscopy including 3D imaging of bulk samples is possible, minimal sample
preparation required as samples can be imaged in situ, has a higher resolution and more
sensitive detection as compared to conventional microscopy and it is possible to image dynamic
processes. Disadvantages of confocal microscopy are limited to the fact that staining generally
involves solvents that may affect the sample and/or introduce artefacts to the sample. As of
now diffraction limited lateral resolution is around 200nm and prolonged exposure to
fluorescent light may induce photobleaching and loss of fluorescence (Jonkman & Brown,
2015).
1.4.9.5. Electron microscopy
Electron microscopes comprise an electron emitter in a vacuum chamber. Electron
beams have a substantially shorter wavelength compared to light radiation and hence give a
58
much higher resolution. Electromagnets are used to focus the electron beam on the sample and
an image is produced by striking the surfacing of a sample and capturing the emitted electrons
as in scanning electron microscopy (SEM) or by passing the electron beam through a thin
section of sample as in transmission electron microscopy (Ikeda et al. 2004). SEM is more
common in the study of the morphology of biopolymer gels.
For TEM thin sections of the sample need to be prepared either as a metallic replica or
as negatively stained dispersions through which a narrow beam of electrons is passed. Sample
preparation is tedious involving chemical fixation, solvent dehydration and embedding in a
suitable resin. Glass or diamond knives are used to section ultrathin samples of ~90-150 nm
thickness using an ultramicrotome. Sections are further stained to increase contrast and carbon
coated to prevent charging.
On the other hand, SEM provides topographic details with a high depth of field.
Conventional SEM is suitable for studying surface morphology of low moisture samples like
freeze dried gels, dairy powder or starches. Cryo-SEM and environmental SEM (ESEM) are
advancements in SEM which allow a sample to be rapidly frozen and fractured under vacuum
and directly imaged under the electron beam without any sample preparation for successful
visualisation of internal structures, whilst ESEM allows samples to be imaged under low
vacuum mode.
A major challenge to EM techniques is minimising damage to non-conductive
specimens caused by the electron beam. Samples usually need to be coated with a conductive
layer of carbon or heavy metal and may mask minute morphological details.
1.5. Significance of the Research
Phase separation in protein and polysaccharide gels remains one of the basic tools of
obtaining the required structural properties and textural profile in food product formulations.
59
In phase separated biopolymer gels, the overall structural properties and techno-functionality
of the mixture depend on the volume and composition of each phase, hence it is critical to
determine the water (solvent) partition between two hydrocolloid phases. A number of
techniques have been used in the past with varying popularity to probe the driving kinetics,
mechanism and extent of phase separation. These, however, have numerous drawbacks as listed
above and fail to provide a holistic understanding of phase behaviour in biopolymer
composites. There is a need for a simplistic and direct approach to probe phase behaviour in
biopolymer composites. Therefore, the significance of this PhD study is summarised as
follows:
i. It is established that phase behaviour in biopolymer gels affects the overall structural
and textural properties of the composite. Therefore, it is critical to accurately determine
the water partition between the two hydrocolloid phases, along with the volume and
composition of each phase.
ii. Although numerous techniques have attempted to probe phase behaviour in biomaterial
composites, these provide an incomplete, tunnel visioned understanding of phase
behaviour.
iii. Of the many techniques, application of rheology based theoretical blending laws is most
popular, and though robust, these are an indirect method of estimating phase behaviour
which based on numerous assumptions and require tedious experimentation and
modelling expertise.
iv. Confocal scanning laser microscopy is used for qualitative analysis in biomedical
science, biosciences and recently in food science but not for quantitative analysis.
v. This research, therefore, aims at developing a new approach for the estimation of the
60
phase volume in biopolymer mixtures using a Z-stack function in confocal microscopy.
1.6. Research Questions
i. Can rheological data be obtained on the small deformation properties of the model mixed
biopolymer gels?
ii. Can blending law analysis be applied to the rheological data of the model composite
systems?
iii. Can the Z-stack generate three-dimensional images of the biopolymer composites?
iv. Can the three-dimensional images be quantified using the image processing software?
v. Can theoretical predictions of phase behaviour be correlated to tangible evidence from
microscopic observations/analysis?
1.7. Research Objectives
To address the research questions outlined in the previous section, the following research
objectives are developed:
i. Designing model biopolymer composites that are suitable for both rheological as well as
microscopic analysis.
ii. Obtaining rheological data on the small deformation properties of model biopolymer
composites and applying blending law analysis to it.
iii. Developing a method to capture three-dimensional images of the composite gels using
the Z-stack feature of CLSM.
iv. Developing a novel technique of quantifying three-dimensional images to estimate phase
volume using two image analysis software- FIJI and Imaris.
v. Testing the efficacy of the established protocol in probing biopolymer composites of
varying complexity.
vi. Comparing the two image analysis software for accuracy and customizability to ensure
61
wide-spread applicability of the protocol.
References
Abramowitz, M. (1987). Contrast Methods in Microscopy: Transmitted Light: Olympus
Corporation, Precision Instrument Division.
Affdl, J. H., & Kardos, J. (1976). The Halpin‐Tsai equations: a review. Polymer Engineering
& Science, 16(5), 344-352.
Agbenorhevi, J. K., & Kontogiorgos, V. (2010). Polysaccharide determination in
protein/polysaccharide mixtures for phase-diagram construction. Carbohydrate
polymers, 81(4), 849-854.
Anderson, V., & Jones, R. J. P. (2001). The influence of gelation on the mechanism of phase
separation of a biopolymer mixture. 42(23), 9601-9610.
Antonov, Y. A., & Gonçalves, M. (2018). Phase equilibria and the mechanical properties of
gel-like water-gelatin-locust bean gum systems. Journal of Characterization and
Development of Novel Materials, 10(2), 77-87.
Antonov, Y. A., Van Puyvelde, P., & Moldenaers, P. (2004). Interfacial tension of aqueous
biopolymer mixtures close to the critical point. International Journal of Biological
Macromolecules, 34(1), 29-35. doi:https://doi.org/10.1016/j.ijbiomac.2004.01.001
Audebrand, M., Doublier, J., Durand, D., & Emery, J. J. F. h. (1995). Investigation of gelation
phenomena of some polysaccharides by ultrasonic spectroscopy. 9(3), 195-203.
Auty, M. (2013). Confocal microscopy: principles and applications to food microstructures. In
Food microstructures (96-P98): Elsevier.
Aymard, P., Williams, M., Clark, A., & Norton, I. J. L. (2000). A turbidimetric study of phase
separating biopolymer mixtures during thermal ramping. 16(19), 7383-7391.
Bamberger, S., Seaman, G. V., Sharp, K., & Brooks, D. E. (1984). The effects of salts on the
interfacial tension of aqueous dextran poly (ethylene glycol) phase systems. Journal of
62
Colloid and Interface Science, 99(1), 194-200.
Beijerinck, M. J. Z. f. B., Parasiten und Infektionskrankheiten. (1896). Über eine
Eigentümlichkeit der löslichen Stärke. 2, 679-699.
Bergfeldt, K., Piculell, L., & Linse, P. (1996). Segregation and association in mixed polymer
solutions from Flory− Huggins model calculations. The Journal of Physical Chemistry,
100(9), 3680-3687.
Bourriot, S., Garnier, C., & Doublier, J. (1999). Phase separation, rheology and microstructure
of micellar casein–guar gum mixtures. Food Hydrocolloids, 13(1), 43-49.
Bourriot, S., Garnier, C., & Doublier, J. L. (1999). Micellar-casein–κ-carrageenan mixtures. I.
Phase separation and ultrastructure. Carbohydrate polymers, 40(2), 145-157.
Bradford, M. M. (1976). A rapid and sensitive method for the quantitation of microgram
quantities of protein utilizing the principle of protein-dye binding. Analytical
biochemistry, 72(1), 248-254. doi:https://doi.org/10.1016/0003-2697(76)90527-3
Bradstreet, R. B. (1954). Kjeldahl method for organic nitrogen. Analytical chemistry, 26(1),
185-187.
Burchard, W. (1994). Light scattering techniques. In Physical techniques for the study of food
biopolymers (151-213): Springer.
Capron, I., Costeux, S., & Djabourov, M. (2001). Water in water emulsions: phase separation
and rheology of biopolymer solutions. Rheologica Acta, 40(5), 441-456.
Cardona, J. A. R., Iriart, C. H., & Herrera, M. L. (2013). Applications of confocal laser
scanning microscopy (CLSM) in foods. In Confocal Laser Microsc.-Princ. Appl. Med.
Biol. Food Sci. (203-234).
Cesàro, A., Cuppo, F., Fabri, D., & Sussich, F. (1999). Thermodynamic behavior of mixed
biopolymers in solution and in gel phase. Thermochimica Acta, 328(1), 143-153.
Chanamai, R., & McClements, D. J. (2001). Depletion flocculation of beverage emulsions by
63
gum arabic and modified starch. Journal of food science, 66(3), 457-463.
Chappellaz, A., Alexander, M., & Corredig, M. (2010). Phase separation behavior of caseins
in milk containing flaxseed gum and κ-carrageenan: a light-scattering and ultrasonic
spectroscopy study. Food Biophysics, 5(2), 138-147.
Chee, K. K. (1990). Determination of polymer-polymer miscibility by viscometry. European
Polymer Journal, 26(4), 423-426.
Chen, C.-H., & Cheng, C.-H. (1996). Effective elastic moduli of misoriented short-fiber
composites. International Journal of Solids and Structures, 33(17), 2519-2539.
Chiu, M. H., & Prenner, E. J. (2011). Differential scanning calorimetry: An invaluable tool for
a detailed thermodynamic characterization of macromolecules and their
interactions. Journal of Pharmacy and Bioallied Sciences, 3(1), 39.
Chronakis, I. S., & Kasapis, S. (1995). Food applications of biopolymer—theory and practice.
In G. Charalambous (Ed.), Developments in Food Science 37(75-109): Elsevier.
Clark, A., Richardson, R., Ross-Murphy, S., & Stubbs, J. (1983). Structural and mechanical
properties of agar/gelatin co-gels. Small-deformation studies. Macromolecules, 16(8),
1367-1374.
Clark, A. H. (1996). Biopolymer gels. Current Opinion in Colloid & Interface Science, 1(6),
712-717.
Clark, A. H., & Ross-Murphy, S. B. (1987). Structural and mechanical properties of
biopolymer gels. In Biopolymers (57-192): Springer.
Clark, A. H., & Ross-Murphy, S. B. (2009). Biopolymer network assembly: measurement and
theory. In Modern biopolymer science (1-27): Elsevier.
Conti, D. S., Yoshida, M. I., Pezzin, S. H., & Coelho, L. A (n.d.). On the miscibility of
biopolymers and their interactions. Paper presented at the 2nd International
Symposium on Calorimetry and Chemical Thermodynamics P.
64
Copeland, R. A. (1994). Methods for protein analysis: Springer.
Corredig, M., Sharafbafi, N., & Kristo, E. J. F. H. (2011). Polysaccharide–protein interactions
in dairy matrices, control and design of structures. 25(8), 1833-1841.
Cox, R. (1969). The deformation of a drop in a general time-dependent fluid flow. Journal of
fluid mechanics, 37(3), 601-623.
Cragg, L., & Bigelow, C. (1955). The viscosity slope constant k′—ternary systems: Polymer–
polymer–solvent. Journal of Polymer Science, 16(82), 177-191.
Dai, B., & Matsukawa, S. J. C. r. (2013). Elucidation of gelation mechanism and molecular
interactions of agarose in solution by 1H NMR. 365, 38-45.
Danait, A., & Deshpande, D. D. (1995). A novel method for determination of polymer-polymer
miscibility by viscometry. European Polymer Journal, 31(12), 1221-1225.
Davidson, M. W., & Abramowitz, M. (2002). Optical microscopy. Encyclopedia of imaging
science and technology.
Davies, W. (1971a). The elastic constants of a two-phase composite material. Journal of
Physics D: Applied Physics, 4(8), 1176.
Davies, W. (1971b). The theory of composite dielectrics. Journal of Physics D: Applied
Physics, 4(2), 318.
de Kruif, C. G., & Tuinier, R. (2001). Polysaccharide protein interactions. Food Hydrocolloids,
15(4), 555-563.
Debye, P. (1959). Angular dissymmetry of the critical opalescence in liquid mixtures. The
Journal of Chemical Physics, 31(3), 680-687.
Dekkers, B. L., Emin, M. A., Boom, R. M., & van der Goot, A. J. (2018). The phase properties
of soy protein and wheat gluten in a blend for fibrous structure formation. Food
Hydrocolloids, 79, 273-281.
Delly, J. G. (1988). Photography Through The Microscope. 9e. Eastman Kodak Co.,
65
Rochester, New York.
Ding, P., Wolf, B., Frith, W., Clark, A., Norton, I., & Pacek, A. (2002). Interfacial tension in
phase-separated gelatin/dextran aqueous mixtures. Journal of Colloid and Interface
Science, 253(2), 367-376.
Ding, P., Wolf, B., Frith, W. J., Clark, A. H., Norton, I. T., & Pacek, A. W. (2002). Interfacial
Tension in Phase-Separated Gelatin/Dextran Aqueous Mixtures. Journal of Colloid and
Interface Science, 253(2), 367-376.
Doi, M., Edwards, S. F., & Edwards, S. F. (1988). The theory of polymer dynamics (Vol. 73):
Oxford University Press.
Donato, L., Garnier, C., Novales, B., & Doublier, J. L. (2005). Gelation of globular protein in
presence of low methoxyl pectin: effect of Na+ and/or Ca2+ ions on rheology and
microstructure of the systems. Food Hydrocolloids, 19(3), 549-556.
Dubois, M., Gilles, K. A., Hamilton, J. K., Rebers, P. t., & Smith, F. (1956). Colorimetric
method for determination of sugars and related substances. Analytical chemistry, 28(3),
350-356.
Ernst, O., & Zor, T. (2010). Linearization of the Bradford protein assay. Journal of Visualized
Experiments, (38), e1918.
Farouk, M., Frost, D., Krsinic, G., & Wu, G. (2011). Phase behaviour, rheology and
microstructure of mixture of meat proteins and kappa and iota carrageenans. Food
Hydrocolloids, 25(6), 1627-1636.
Farouk, M. M., Frost, D. A., Krsinic, G., & Wu, G. (2011). Phase behaviour, rheology and
microstructure of mixture of meat proteins and kappa and iota carrageenans. Food
Hydrocolloids, 25(6), 1627-1636.
Firoozmand, H., & Rousseau, D. (2013). Microstructure and elastic modulus of phase-
separated gelatin–starch hydrogels containing dispersed oil droplets. Food
66
Hydrocolloids, 30(1), 333-342.
Firoozmand, H., & Rousseau, D. (2015). Controlled phase separation for texture modification.
In Modifying Food Texture (157-181): Elsevier.
Flory, P. J. (1953). Principles of polymer chemistry: Cornell University Press.
Foord, S., & Atkins, E. J. B. O. R. o. B. (1989). New x‐ray diffraction results from agarose:
Extended single helix structures and implications for gelation mechanism. 28(8), 1345-
1365.
Foreman, J., Sauerbrunn, S., & Marcozzi, C. (2013). ‘Thermal Analysis & Rheology. Kyoritsu
Chem. & Co., Ltd, Kisarazu, Japan.
Fredrickson, G. H., Liu, A. J., & Bates, F. S. (1994). Entropic corrections to the Flory-Huggins
theory of polymer blends: Architectural and conformational effects. Macromolecules,
27(9), 2503-2511.
Freitas, M. L. F., Albano, K. M., & Telis, V. R. N. (2017). Characterization of biopolymers
and soy protein isolate-high-methoxyl pectin complex. Polímeros, 27(1), 62-67.
Frith, W. J. (2010). Mixed biopolymer aqueous solutions–phase behaviour and rheology.
Advances in Colloid and Interface Science, 161(1-2), 48-60.
Gaca-Zając, K. Z., Smith, B. R., Nordon, A., Fletcher, A. J., Johnston, K., & Sefcik, J. (2018).
Investigation of IR and Raman spectra of species present in formaldehyde-water-
methanol systems. Vibrational Spectroscopy, 97, 44-54.
Garcı́a, R., Melad, O., Gómez, C. M., Figueruelo, J. E., & Campos, A. (1999). Viscometric
study on the compatibility of polymer–polymer mixtures in solution. European
Polymer Journal, 35(1), 47-55. doi:https://doi.org/10.1016/S0014-3057(98)00106-2
George, J. W. (2001). The usefulness and limitations of hand‐held refractometers in veterinary
laboratory medicine: an historical and technical review. Veterinary Clinical Pathology,
67
30(4), 201-210.
Gilbert, G., & Spragg, S. (1964). Iodine sorption:" blue value. Whistler, RL et al. Methods in
carbohydrate chemistry, 4, 168-169.
Goh, K. K., Teo, A., Sarkar, A., & Singh, H. (2020). Milk protein-polysaccharide interactions.
In Milk proteins (499-535): Elsevier.
González-González, M., Mayolo-Deloisa, K., Rito-Palomares, M., & Winkler, R. (2011).
Colorimetric protein quantification in aqueous two-phase systems. Process
Biochemistry, 46(1), 413-417.
Gornall, A. G., Bardawill, C. J., & David, M. M. (1949). Determination of serum proteins by
means of the biuret reaction. Journal of biological chemistry, 177(2), 751-766.
Gregorova, A. (2013). Application of differential scanning calorimetry to the characterization
of biopolymers. Applications of Calorimetry in a Wide Context-Differential Scanning
Calorimetry, Isothermal Titration Calorimetry and Microcalorimetry, 3-20.
Grinberg, V. Y., & Tolstoguzov, V. (1972). Thermodynamic compatibility of gelatin with some
D-glucans in aqueous media. Carbohydrate Research, 25(2), 313-321.
Grinberg, V. Y., & Tolstoguzov, V. (1997). Thermodynamic incompatibility of proteins and
polysaccharides in solutions. Food Hydrocolloids, 11(2), 145-158.
Guido, S., Simeone, M., & Alfani, A. (2002). Interfacial tension of aqueous mixtures of Na-
caseinate and Na-alginate by drop deformation in shear flow. Carbohydrate polymers,
48(2), 143-152.
Guido, S., & Villone, M. (1999). Measurement of interfacial tension by drop retraction
analysis. Journal of Colloid and Interface Science, 209(1), 247-250.
Gupta, S., Wang, W. S., & Vanapalli, S. A. (2016). Microfluidic viscometers for shear rheology
of complex fluids and biofluids. Biomicrofluidics, 10(4), 043402.
Gutiérrez Sama, S., Farenc, M., Barrère-Mangote, C., Lobinski, R., Afonso, C., Bouyssière,
68
B., & Giusti, P. (2018). Molecular Fingerprints and Speciation of Crude Oils and Heavy
Fractions Revealed by Molecular and Elemental Mass Spectrometry: Keystone
between Petroleomics, Metallopetroleomics, and Petrointeractomics. Energy & Fuels,
32(4), 4593-4605.
Hammouda, B. (2010). SANS from polymers—review of the recent literature. Journal of
Macromolecular Science®, Part C: Polymer Reviews, 50(1), 14-39.
Haug, I., Williams, M. A. K., Lundin, L., Smidsrød, O., & Draget, K. I. (2003). Molecular
interactions in, and rheological properties of, a mixed biopolymer system undergoing
order/disorder transitions. Food Hydrocolloids, 17(4), 439-444.
Hemar, Y., Tamehana, M., Munro, P., & Singh, H. J. F. H. (2001). Viscosity, microstructure
and phase behavior of aqueous mixtures of commercial milk protein products and
xanthan gum. 15(4-6), 565-574.
Hemar, Y., Tamehana, M., Munro, P. A., & Singh, H. (2001). Viscosity, microstructure and
phase behavior of aqueous mixtures of commercial milk protein products and xanthan
gum. Food Hydrocolloids, 15(4), 565-574. doi:https://doi.org/10.1016/S0268-
005X(01)00077-7
Horkay, F., Basser, P. J., Hecht, A.-M., & Geissler, E. J. T. J. o. c. p. (2018). Ionic effects in
semi-dilute biopolymer solutions: A small angle scattering study. 149(16), 163312.
Horkay, F., Hammouda, B. J. C., & Science, P. (2008). Small-angle neutron scattering from
typical synthetic and biopolymer solutions. 286(6-7), 611-620.
Icoz, D. Z., & Kokini, J. L. J. C. p. (2007). Probing the boundaries of miscibility in model
carbohydrates consisting of chemically derivatized dextrans using DSC and FTIR
spectroscopy. 68(1), 68-76.
Ikeda, S., Nitta, Y., Kim, B. S., Temsiripong, T., Pongsawatmanit, R., & Nishinari, K. J. F. H.
69
(2004). Single-phase mixed gels of xyloglucan and gellan. 18(4), 669-675.
Inamura, I., & Jinbo, Y. (1991). Interaction between poly (vinyl alcohol) and poly (ethylene
glycol) in water studied by viscosity and density. Polymer journal, 23(9), 1143-1147.
Jain, V. M., Karibasappa, G. N., Dodamani, A. S., & Mali, G. V. (2017). Estimating the
carbohydrate content of various forms of tobacco by phenol-sulfuric acid method.
Journal of education and health promotion, 6.
Jiang, B., Tsao, R., Li, Y., & Miao, M. (2014). Food Safety: Food Analysis
Technologies/Techniques. In N. K. Van Alfen (Ed.), Encyclopedia of Agriculture and
Food Systems (273-288). Oxford: Academic Press.
Johansson, H.-O., Karlström, G., Tjerneld, F., & Haynes, C. A. (1998). Driving forces for phase
separation and partitioning in aqueous two-phase systems. Journal of Chromatography
B: Biomedical Sciences and Applications, 711(1), 3-17.
Jonkman, J., & Brown, C. M. J. J. o. b. t. J. (2015). Any way you slice it—a comparison of
confocal microscopy techniques. 26(2), 54.
Kacir, L., Narkis, M., & Ishai, O. (1975). Oriented short glass‐fiber composites. I. Preparation
and statistical analysis of aligned fiber mats. Polymer Engineering & Science, 15(7),
525-531.
Kasapis, S. (2008). Phase Separation in Biopolymer Gels: A Low- to High-Solid Exploration
of Structural Morphology and Functionality. Critical reviews in food science and
nutrition, 48(4), 341-359.
Kasapis, S. (2009). CHAPTER 7 - Unified Application of the Materials-Science Approach to
the Structural Properties of Biopolymer Co-Gels throughout the Industrially Relevant
Level of Solids. In S. Kasapis, I. T. Norton, & J. B. Ubbink (Eds.), Modern Biopolymer
70
Science (225-260). San Diego: Academic Press.
Kasapis, S., & Bannikova, A. (2017). Chapter 2 - Rheology and Food Microstructure. In J.
Ahmed, P. Ptaszek, & S. Basu (Eds.), Advances in Food Rheology and Its Applications
(7-46): Woodhead Publishing.
Kasapis, S., Morris, E. R., Norton, I. T., & Clark, A. H. (1993). Phase equilibria and gelation
in gelatin/maltodextrin systems — Part IV: composition-dependence of mixed-gel
moduli. Carbohydrate polymers, 21(4), 269-276. doi:https://doi.org/10.1016/0144-
8617(93)90058-C
Kerner, E. (1956a). The elastic and thermo-elastic properties of composite media. Proceedings
of the physical society. Section B, 69(8), 808.
Kerner, E. (1956b). The electrical conductivity of composite media. Proceedings of the
physical society. Section B, 69(8), 802.
Kim, J. M., Jung, H. S., Park, J. W., Lee, H. Y., & Kawai, T. J. A. c. a. (2004). AFM phase lag
mapping for protein–DNA oligonucleotide complexes. 525(2), 151-157.
King, S. M. (1999). Small-angle neutron scattering. Modern Techniques for Polymer
Characterisation, 12.
Klassen, D. R., Elmer, C. M., & Nickerson, M. T. (2011). Associative phase separation
involving canola protein isolate with both sulphated and carboxylated polysaccharides.
Food Chemistry, 126(3), 1094-1101.
Koh, L. W., & Kasapis, S. (2011). Orientation of short microcrystalline cellulose fibre in a
gelatin matrix. Food Hydrocolloids, 25(5), 1402-1405.
Kontogiorgos, V., Tosh, S. M., & Wood, P. J. J. F. B. (2009). Kinetics of phase separation of
oat β-glucan/whey protein isolate binary mixtures. 4(3), 240.
Krigbaum, W. R., & Wall, F. T. (1950). Viscosities of binary polymeric mixtures. Journal of
71
Polymer Science, 5(4), 505-514.
Leharne, S. J. C. (2017). The physical chemistry of high-sensitivity differential scanning
calorimetry of biopolymers. 3(1), 1.
Leloup, V., Colonna, P., & Buleon, A. J. J. o. C. S. (1991). Influence of amylose-amylopectin
ratio on gel properties. 13(1), 1-13.
Lewis, T. B., & Nielsen, L. E. (1970). Dynamic mechanical properties of particulate‐filled
composites. Journal of applied polymer science, 14(6), 1449-1471.
Li-Chan, E. C. (1996). The applications of Raman spectroscopy in food science. Trends in
Food Science & Technology, 11(7), 361-370.
Li, X., Liu, Y., Li, N., Xie, D., Yu, J., Wang, F., & Wang, J. (2016). Studies of phase separation
in soluble rice protein/different polysaccharides mixed systems. LWT-Food Science
and Technology, 65, 676-682.
Loret, C., Schumm, S., Pudney, P. D., Frith, W. J., & Fryer, P. J. (2005). Phase separation and
molecular weight fractionation behaviour of maltodextrin/agarose mixtures. Food
Hydrocolloids, 19(3), 557-565.
Mæhre, H. K., Dalheim, L., Edvinsen, G. K., Elvevoll, E. O., & Jensen, I.-J. (2018). Protein
determination—method matters. Foods, 7(1), 5.
Magonov, S. N. (2006). Atomic force microscopy in analysis of polymers. Encyclopedia of
Analytical Chemistry: Applications, Theory and Instrumentation.
Marshall, R. J. (2005). FOOD AND NUTRITIONAL ANALYSIS | Dairy Products. In P.
Worsfold, A. Townshend, & C. Poole (Eds.), Encyclopedia of Analytical Science
(Second Edition) (312-319). Oxford: Elsevier.
Masuelli, M. A., & Illanes, C. O. (2014). Review of the characterization of sodium alginate by
intrinsic viscosity measurements: Comparative analysis between conventional and
72
single point methods.
Masuko, T., Minami, A., Iwasaki, N., Majima, T., Nishimura, S.-I., & Lee, Y. C. (2005).
Carbohydrate analysis by a phenol–sulfuric acid method in microplate format.
Analytical biochemistry, 339(1), 69-72.
Matsushita, M., Irino, T., Komoda, T., & Sakagishi, Y. (1993). Determination of proteins by a
reverse biuret method combined with the copper-bathocuproine chelate reaction.
Clinica chimica acta, 216(1-2), 103-111.
Melish, K. A Literature Review on Viscometers that Measure the Viscosity at Extreme
Conditions.
Mertz, J. (2019). Introduction to optical microscopy: Cambridge University Press.
Mhaske, P., Condict, L., Dokouhaki, M., Farahnaky, A., & Kasapis, S. (2020). Phase volume
quantification of agarose-ghee gels using 3D confocal laser scanning microscopy and
blending law analysis: A comparison. LWT, 109567.
Mhaske, P., Condict, L., Dokouhaki, M., Katopo, L., & Kasapis, S. (2019). Quantitative
analysis of the phase volume of agarose-canola oil gels in comparison to blending law
predictions using 3D imaging based on confocal laser scanning microscopy. Food
Research International, 125, 108529.
Michon, C., Cuvelier, G., Launay, B., Parker, A., & Takerkart, G. (1995). Study of the
compatibility/incompatibility of gelatin/iota-carrageenan/water mixtures.
Carbohydrate polymers, 28(4), 333-336.
Miles, M. J., Morris, V. J., Orford, P. D., & Ring, S. G. J. C. r. (1985). The roles of amylose
and amylopectin in the gelation and retrogradation of starch. 135(2), 271-281.
Morris, E. (2000). Rheology of biopolymer co-gels. In Hydrocolloids (135-146): Elsevier.
Morris, E. R. (1992). The effect of solvent partition on the mechanical properties of biphasic
biopolymer gels: an approximate theoretical treatment. Carbohydrate polymers, 17(1),
73
65-70.
Morris, E. R. (2009). CHAPTER 5 - Functional Interactions in Gelling Biopolymer Mixtures.
In S. Kasapis, I. T. Norton, & J. B. Ubbink (Eds.), Modern Biopolymer Science (167-
198). San Diego: Academic Press.
Morris, V. J., Kirby, A. R., & Gunning, A. P. (1999). Atomic force microscopy for biologists
(Vol. 57): World Scientific.
Mousia, Z., Farhat, I., Blachot, J., & Mitchell, J. J. P. (2000). Effect of water partitioning on
the glass-transition behaviour of phase separated amylopectin–gelatin mixtures. 41(5),
1841-1848.
Neirynck, N., Dewettinck, K., & Van der Meeren, P. (2007). Influence of pH and biopolymer
ratio on sodium caseinate—guar gum interactions in aqueous solutions and in O/W
emulsions. Food Hydrocolloids, 21(5-6), 862-869.
Nguyen-Tri, P., Ghassemin, P., Carriere, P., Assadi, A. A., & Nguyen, D. D. (2020). Recent
Progressive Use of Advanced Atomic Force Microscopy in Polymer Science: A
Review.
Nielsen, L. E. (1970). Generalized equation for the elastic moduli of composite materials.
Journal of Applied Physics, 41(11), 4626-4627.
Normand, V., Pudney, P. D., Aymard, P., & Norton, I. T. J. J. o. A. P. S. (2000). Weighted‐
average isostrain and isostress model to describe the kinetic evolution of the mechanical
properties of a composite gel: Application to the system gelatin: Maltodextrin. 77(7),
1465-1477.
Norton, I. T., & Frith, W. J. (2001). Microstructure design in mixed biopolymer composites.
Food Hydrocolloids, 15(4), 543-553.
Palierne, J. (1990). Linear rheology of viscoelastic emulsions with interfacial tension.
Rheologica acta, 29(3), 204-214.
74
Palierne, J. (1991). Rheologica Acta Rheol Acta 30: 497 (1991). Rheologica acta, 30(497).
Parets, M. J., Garcia, R., Soria, V., & Campos, A. (1990). Solution properties of
polyelectrolytes—V. Viscometric study of mixed polyanions in pure water. European
Polymer Journal, 26(7), 767-773. doi:https://doi.org/10.1016/0014-3057(90)90127-P
Patonay, G., Salon, J., Sowell, J., & Strekowski, L. J. M. (2004). Noncovalent labeling of
biomolecules with red and near-infrared dyes. 9(3), 40-49.
Perrechil, F., Braga, A., & Cunha, R. (2009). Interactions between sodium caseinate and LBG
in acidified systems: Rheology and phase behavior. Food Hydrocolloids, 23(8), 2085-
2093.
Picout, D. R., & Ross-Murphy, S. B. (2003). Rheology of biopolymer solutions and gels.
TheScientificWorldJOURNAL, 3.
Pittz, E. P., Lee, J. C., Bablouzian, B., Townend, R., & Timasheff, S. N. (1973). [10] Light
scattering and differential refractometry. In Methods in enzymology 27(209-256):
Elsevier.
Pluta, M., & Maksymilian, P. (1988). Advanced light microscopy (Vol. 1): Elsevier
Amsterdam.
Polyakov, V. I., Grinberg, V. Y., & Tolstoguzov, V. B. (1980). Application of phase-volume-
ratio method for determining the phase diagram of water-casein-soybean globulins
system. Polymer Bulletin, 2(11), 757-760. doi:10.1007/BF00255893
Prameela, K., Mohan, C. M., & Ramakrishna, C. (2018). Biopolymers for food design:
Consumer-friendly natural ingredients. In Biopolymers for food design (pp. 1-32):
Elsevier.
Pudney, P. D. A., Hancewicz, T. M., Cunningham, D. G., & Gray, C. (2003). A novel method
for measuring concentrations of phase separated biopolymers: the use of confocal
Raman spectroscopy with self-modelling curve resolution. Food Hydrocolloids, 17(3),
75
345-353.
Rafe, A., & Razavi, S. M. (2015). Effect of thermal treatment on chemical structure of β-
lactoglobulin and basil seed gum mixture at different states by ATR-FTIR
spectroscopy. International Journal of Food Properties, 18(12), 2652-2664.
Rahman, M. S., Al-Marhubi, I. M., & Al-Mahrouqi, A. (2007). Measurement of glass transition
temperature by mechanical (DMTA), thermal (DSC and MDSC), water diffusion and
density methods: A comparison study. Chemical Physics Letters, 440(4), 372-377.
Reichardt, W., & Eckert, W. (1991). The determination of protein content of milk, cheese, and
meat with the use of the biuret reaction. Die Nahrung, 35(7), 731.
Richardson, R. K., & Kasapis, S. (1998). Rheological methods in the characterisation of food
biopolymers. In Developments in Food Science, 39(1-48): Elsevier.
Robeson, L. M. (2007). Polymer blends. A Comprehensive Review.
Sapan, C. V., Lundblad, R. L., & Price, N. C. (1999). Colorimetric protein assay techniques.
Biotechnology and applied Biochemistry, 29(2), 99-108.
Sarika, P. R., Pavithran, A., & James, N. R. (2015). Cationized gelatin/gum arabic
polyelectrolyte complex: Study of electrostatic interactions. Food Hydrocolloids, 49,
176-182.
Schmitt, C., Sanchez, C., Desobry-Banon, S., & Hardy, J. (1998). Structure and
technofunctional properties of protein-polysaccharide complexes: a review. Critical
reviews in food science and nutrition, 38(8), 689-753.
Scholten, E., Tuinier, R., Tromp, R. H., & Lekkerkerker, H. (2002). Interfacial tension of a
decomposed biopolymer mixture. Langmuir, 18(6), 2234-2238.
Schorsch, C., Clark, A., Jones, M., & Norton, I. (1999). Behaviour of milk
protein/polysaccharide systems in high sucrose. Colloids and surfaces B: Biointerfaces,
76
12(3-6), 317-329.
Schorsch, C., Jones, M., & Norton, I. J. F. H. (1999). Thermodynamic incompatibility and
microstructure of milk protein/locust bean gum/sucrose systems. 13(2), 89-99.
Schorsch, C., Jones, M., & Norton, I. J. F. H. (2000). Phase behaviour of pure micellar casein/κ-
carrageenan systems in milk salt ultrafiltrate. 14(4), 347-358.
Semasaka, C., Mhaske, P., Buckow, R., & Kasapis, S. J. F. c. (2018). Modelling water partition
in composite gels of BSA with gelatin following high pressure treatment. 265, 32-38.
Semenova, M., & Savilova, L. (1998). The role of biopolymer structure in interactions between
unlike biopolymers in aqueous medium. Food Hydrocolloids, 12(1), 65-75.
Shamsuri, A., Abdullah, D. K., Daik, R. J. C. C., & Technology. (2012). FABRICATION OF
AGAR/BIOPOLYMER BLEND AEROGELS IN IONIC LIQUID AND CO-
SOLVENT MIXTURE. 46, 45-52.
Shrinivas, P., Kasapis, S., & Tongdang, T. J. L. (2009). Morphology and mechanical properties
of bicontinuous gels of agarose and gelatin and the effect of added lipid phase. 25(15),
8763-8773.
Sigillo, I., Di Santo, L., Guido, S., & Grizzuti, N. (1997). Comparative measurements of
interfacial tension in a model polymer blend. Polymer Engineering & Science, 37(9),
1540-1549.
Smoluchowski, M. J. T. L., Edinburgh,, Magazine, D. P., & Science, J. o. (1912). XIII. On
opalescence of gases in the critical state. 23(133), 165-173.
Soedjak, H. S. (1994). Colorimetric determination of carrageenans and other anionic
hydrocolloids with methylene blue. Analytical chemistry, 66(24), 4514-4518.
Spyropoulos, F., Ding, P., Frith, W., Norton, I., Wolf, B., & Pacek, A. (2008). Interfacial
tension in aqueous biopolymer–surfactant mixtures. Journal of Colloid and Interface
77
Science, 317(2), 604-610.
Stockmayer, W. (1950). Light scattering in multi‐component systems. The Journal of Chemical
Physics, 18(1), 58-61.
Stokes, J., Wolf, B., & Frith, W. (2001). Phase-separated biopolymer mixture rheology:
Prediction using a viscoelastic emulsion model. Journal of Rheology, 45(5), 1173-1191.
Svensson, M., Berggren, K., Veide, A., & Tjerneld, F. (1999). Aqueous two-phase systems
containing self-associating block copolymers: Partitioning of hydrophilic and
hydrophobic biomolecules. Journal of Chromatography A, 839(1), 71-83.
Takayanagi, M. (1963). Viscoelastic behavior of polymer blends and its comparison with
model experiments. J. Soc. Mater. Sci. Jpn., 12, 389-394.
Tanaka, F. (2002). Theoretical study of molecular association and thermoreversible gelation in
polymers. Polymer journal, 34(7), 479-509.
Thaiudom, S., & Goff, H. D. (2003). Effect of κ-carrageenan on milk protein polysaccharide
mixtures. International Dairy Journal, 13(9), 763-771.
Thaiudom, S., & Goff, H. D. (2003). Effect of κ-carrageenan on milk protein polysaccharide
mixtures. International Dairy Journal, 13(9), 763-771.
Tolstoguzov, V. (1991). Functional properties of food proteins and role of protein-
polysaccharide interaction. Food Hydrocolloids, 4(6), 429-468.
Tolstoguzov, V. (1999). Compositions and phase diagrams for aqueous systems based on
proteins and polysaccharides. In International review of cytology 192(3-31): Elsevier.
Tolstoguzov, V. (2003). Some thermodynamic considerations in food formulation. Food
Hydrocolloids, 17(1), 1-23. doi:https://doi.org/10.1016/S0268-005X(01)00111-4
Tolstoguzov, V. (2006). 17 Phase Behavior in Mixed Polysaccharide Systems. Food
polysaccharides and their applications, 589.
Tolstoguzov, V. B. (1993). Thermodynamic Incompatibility of Food. Food colloids and
78
polymers: Stability and mechanical properties, 113, 94.
Tolstoguzov, V. B. (2007). 7 - Ingredient interactions in complex foods: aggregation and phase
separation. In D. J. McClements (Ed.), Understanding and Controlling the
Microstructure of Complex Foods (185-206): Woodhead Publishing.
Tomotika, S. J. P. o. t. R. S. o. L. S. A. (1935). On the Instability of a Cylindrical Thread of a
Toubal, M., Nongaillard, B., Radziszewski, E., Boulenguer, P., & Langendorff, V. (2003). Ultrasonic
monitoring of sol–gel transition of natural hydrocolloids. Journal of Food Engineering, 58(1),
1-4.
Viscous Liquid Surrounded by Another Viscous Fluid. 150, 322.
Tromp, R., Rennie, A., & Jones, R. J. M. (1995). Kinetics of the simultaneous phase separation
and gelation in solutions of dextran and gelatin. 28(12), 4129-4138.
Tseretely, G. I., & Smirnova, O. I. (1992). DSC study of melting and glass transition in
gelatins. Journal of Thermal Analysis and Calorimetry, 38(5), 1189-1201.
Tucker III, C. L., & Moldenaers, P. (2002). Microstructural evolution in polymer blends.
Annual Review of Fluid Mechanics, 34(1), 177-210.
Tuinier, R., Ten Grotenhuis, E., Holt, C., Timmins, P. A., & De Kruif, C. G. (1999). Depletion
interaction of casein micelles and an exocellular polysaccharide. Physical Review
E, 60(1), 848.
van de Velde, F., Weinbreck, F., Edelman, M. W., van der Linden, E., & Tromp, R. H. (2003).
Visualisation of biopolymer mixtures using confocal scanning laser microscopy
(CSLM) and covalent labelling techniques. Colloids and Surfaces B:
Biointerfaces, 31(1-4), 159-168.
Van Puyvelde, P., Antonov, Y. A., & Moldenaers, P. (2002). Rheo-optical measurement of the
interfacial tension of aqueous biopolymer mixtures. Food Hydrocolloids, 16(5), 395-
79
402.
Vis, M., Peters, V. F., Blokhuis, E. M., Lekkerkerker, H. N., Erné, B. H., & Tromp, R. H.
(2015). Decreased interfacial tension of demixed aqueous polymer solutions due to
charge. Physical review letters, 115(7), 078303.
Voron’ko, N. G., Derkach, S. R., Vovk, M. A., & Tolstoy, P. M. (2017). Complexation of κ-
carrageenan with gelatin in the aqueous phase analysed by 1H NMR kinetics and
relaxation. Carbohydrate polymers, 169, 117-126.
Voron’ko, N. G., Derkach, S. R., Vovk, M. A., & Tolstoy, P. M. (2016). Formation of κ-
carrageenan–gelatin polyelectrolyte complexes studied by 1H NMR, UV spectroscopy
and kinematic viscosity measurements. Carbohydrate polymers, 151, 1152-1161.
Wadsö, I. (1980). Calorimetric instrumentation for studies of biopolymer model
compounds. Pure and Applied Chemistry, 52(2), 465-477.
Walter, H., Johansson, G., & Brooks, D. E. (1991). Partitioning in aqueous two-phase systems:
recent results. Analytical biochemistry, 197(1), 1-18.
Wanchoo, R. K., & Sharma, P. K. (2003). Viscometric study on the compatibility of some
water-soluble polymer–polymer mixtures. European Polymer Journal, 39(7), 1481-
1490.
Wang, Z., Yang, K., Brenner, T., Kikuzaki, H., & Nishinari, K. (2014). The influence of agar
gel texture on sucrose release. Food Hydrocolloids, 36, 196-203.
Waterborg, J. H. (2009). The Lowry method for protein quantitation. In The protein protocols
handbook (7-10): Springer.
Williams, M. A., Foster, T. J., Martin, D. R., Norton, I. T., Yoshimura, M., & Nishinari, K. J.
B. (2000). A molecular description of the gelation mechanism of konjac mannan. 1(3),
440-450.
Williams, P., Clegg, S., Langdon, M., Nishinari, K., & Piculell, L. J. M. (1993). Investigation
80
of the gelation mechanism in. kappa.-carrageenan/konjac mannan mixtures using
differential scanning calorimetry and electron spin resonance spectroscopy. 26(20),
5441-5446.
Winter, P. W., & Shroff, H. (2014). Faster fluorescence microscopy: advances in high speed
biological imaging. Current opinion in chemical biology, 20, 46-53.
Xu, A. Y., Melton, L. D., Ryan, T. M., Mata, J. P., Rekas, A., Williams, M. A. K., &
McGillivray, D. J. (2018). Effects of polysaccharide charge pattern on the
microstructures of β-lactoglobulin-pectin complex coacervates, studied by SAXS and
SANS. Food Hydrocolloids, 77, 952-963.
Xu, J., Fan, X., Ning, Y., Wang, P., Jin, Z., Lv, H., ... & Xu, X. (2013). Effect of spring dextrin
on retrogradation of wheat and corn starch gels. Food Hydrocolloids, 33(2), 361-367.
Yang, H., Wang, Y., Lai, S., An, H., Li, Y., & Chen, F. (2007). Application of atomic force
microscopy as a nanotechnology tool in food science. Journal of food science, 72(4),
65-75.
Yang, K., Wang, Z., Brenner, T., Kikuzaki, H., Fang, Y., & Nishinari, K. (2015). Sucrose
release from agar gels: Effects of dissolution order and the network inhomogeneity.
Food Hydrocolloids, 43, 100-106.
Yuan, C., Xu, D., Cui, B., & Wang, Y. (2019). Gelation of κ-carrageenan/Konjac
glucommanan compound gel: Effect of cyclodextrins. Food Hydrocolloids 87, 158-
164.
Zasypkin, D. V., Braudo, E. E., & Tolstoguzov, V. B. (1997). Multicomponent biopolymer
gels. Food Hydrocolloids, 11(2), 159-170.
Zernike, F. J. P. (1942a). Phase contrast, a new method for the microscopic observation of
transparent objects. 9(7), 686-698.
Zernike, F. J. P. (1942b). Phase contrast, a new method for the microscopic observation of
81
transparent objects part II. 9(10), 974-986.
Zhang, G., & Foegeding, E. A. (2003). Heat-induced phase behavior of β-
lactoglobulin/polysaccharide mixtures. Food Hydrocolloids, 17(6), 785-792.
Zhang, H., Nishinari, K., Williams, M. A., Foster, T. J., & Norton, I. T. J. I. J. o. B. M. (2002).
A molecular description of the gelation mechanism of curdlan. 30(1), 7-16.
Zhuang, X., Wang, L., Jiang, X., Chen, Y., & Zhou, G. (2020). The effects of three
polysaccharides on the gelation properties of myofibrillar protein: Phase behaviour and
82
moisture stability. Meat Science, 108228.
Chapter 2 - Materials and Methods
Abstract
This chapter summarises the specifications of the materials used in this work along with
the equipment, instrumentation and the methods employed in this research. Specifications of
the biopolymers, chemicals and reagents used in this research are provided along with the
experimental methods of analysis such as scanning electron microscopy (SEM), Fourier
transform infrared spectroscopy (FTIR), differential scanning calorimetry (DSC), rheology,
and confocal laser scanning microscopy. The working principle and reasoning behind the
analytical instruments are provided in detail. Further explanation specific to each experimental
83
chapter can be found in Chapters 3 to 6.
2.1. Materials
2.1.1. Biomaterials
2.1.1.1. Agarose
Agarose is a natural polysaccharide and the gelling component of Agar, obtained from the
Rhodophyceae family of red seaweeds. Agarose has negligible traces of sulphur content as
opposed to the charged fraction of agar, agaropectin and exhibits an unbranched linear structure
(Figure 2.1) . In its solid states, it is in the powder form, existing as a threefold, left-handed
double helix. The hydroxyl groups line the central cavity along the helix axis and are
responsible for the hydrogen bonding with water. When dissolved in hot water (~70°C),
agarose remains dispersed as fluctuating disordered coils. Upon cooling, these disordered coils
form an ordered double helix, forming a gel network. In this study, agarose is used as the main
gelling biopolymer within which other biomaterials are entrapped. Type-1 agarose used for the
experiments was supplied by Sigma-Aldrich Co., Sydney, Australia. The powder has an off-
white colour, with the moisture content being less than 10 g H2O/g dry matter and less than
0.15% sulphate according to the supplier. Agarose forms a strong gel with a storage modulus
of 7208 Pa at 1% w/w (at 5°C, angular frequency of 1 rad/sec and 0.01% strain).
84
Figure 2.1 Chemical structure of an agarose polymer (Lee & Bahaman, 2012)
2.1.1.2. Gelatin
As a common animal-derived protein, gelatin is extracted via partial acid or alkaline hydrolysis
and proteolysis of collagen present in the skin, cartilage, bone and connective tissue of fish,
cows or pigs. The physical and chemical properties of gelatin, such as the gel strength,
isoelectric point and molecular weight, vary with the source of gelatin, as well as the type and
extent of extraction. Figure 2.2 illustrates the chemical structure of gelatin. Type-B bovine
gelatin generally results in higher viscosity, obtained via chemically intensive alkaline
hydrolysis for extended time and was used in this work to form a bicontinuous network along
with agarose. It was purchased from Sigma-Aldrich Co., Sydney, Australia. The powder was
light yellow in colour with a bloom strength of 225 with a molecular weight of 173kDa. Gelatin
formed a soft gel matrix with a storage modulus of 81 Pa at 1% w/w (at 5°C, angular frequency
of 1 rad/sec and 0.01% strain).
Figure 2.2. Chemical structure of gelatin (Deshmukh et al. 2017)
2.1.1.3. Lipids
Canola oil is produced from the variety of rapeseeds that have a low erucic acid content. It
mainly constitutes of a mixture of α-linolenic acid and oleic acid. In the present study it was
used as a discontinuous filler that remained dispersed in a continuous gel network exhibiting
85
distinct phase separation. Food grade canola oil was purchased from Manildra Group, NSW,
Australia. 0.06% free fatty acid as oleic acid was present in the oil with a peroxide value of 0.1.
It remains in the liquid state at room temperature with a negligible storage modulus.
Pure cow ghee, a form of clarified butter manufactured by Amul (Gujrat, India) was
used in the study. The density of ghee was measured to be 0.911g/ml. Ghee is a complex
mixture of triacylglycerols (TAGs), which constitutes 98% of its composition. In this research,
ghee is seen to crystalize between 10-15°C.
2.1.1.4. Microcrystalline cellulose
Microcrystalline cellulose (MCC) is widely used in the food and pharmaceutical industries as
a texture modifier/fat replacer and tableting excipient, respectively. MCC is obtained from the
acid hydrolysis of naturally occurring cellulose, such as wood pulp. The extracted colloidal
MCC when dispersed in water, can form a three-dimensional network as the hydrocolloid layer
that surrounds the cellulose particles absorbs water and swells. The high-water holding capacity
of the MCC particles make it ideal as a fat replacer. As it cannot be digested, MCC is also used
in food products to increase their fibre content. MCC was obtained from Sigma Aldrich Co.,
Sydney, Australia. The product was in the form of white fibres with a bulk density of 0.6g/cm3.
The cellulose crystals, being hygroscopic, absorbed water about 4.1 times their weight from
the system without disintegrating.
2.1.2 Chemicals and Dyes
Details of all the chemicals including dyes used for staining in this study are stated in Table
86
2.1. Throughout the study, Milli-Q water is used for preparing chemical solutions.
Table 2.1 List of chemicals and dyes used.
Chemical CAS Number Supplier
Agarose 9004-34-6 Sigma Aldrich
DMSO 67-68-5 Sigma Aldrich
Nile Red 7385-67-3 Sigma Aldrich
DTAF 51306-35-5 AAT Bioquest
FITC 27072-45-3 Sigma Aldrich
Rhodamine B 81-88-9 Sigma Aldrich
Microcrystalline Cellulose 9004-34-6 Sigma Aldrich
Gelatin 9000-70-8 Sigma Aldrich
2.1.3 Instruments
The instruments used in this study along with their manufacturers are listed in Table 2.2
Table 2.2 List of Instruments used
Instrument Type Manufacturer
MDSC Q2000 Calorimeter TA Instruments
(New Castle, DE, USA)
Quanta™ 200 Microscope FEI
87
(Hillsboro, OR, USA)
FriZone Triade freeze dryer Labconco (Kansas City,
Missouri, USA)
Spectrum Two FTIR spectrometer Perkin Elmer (Norway, CT,
USA)
Setaram VII Calorimeter Setaram (Seturau, Caluire,
France)
Discovery Hybrid Rheometer TA Instruments (New Castle,
DE, USA)
Eclipse Ti-E Microscope Nikon (Minato, Tokyo,
Japan)
2.1.4 Sample preparation
2.1.4.1. Agarose-Canola oil composites
The concentration of agarose was kept constant at 1% (w/w) and the concentration of canola
oil was varied (1, 3, 5, 7, 9 and 11%, w/w). Agarose powder was dissolved in milliQ water at
80°C and stirred for 5 min before cooling to 45°C to prepare agarose solution. Respective
quantity of canola oil was then added to the agarose solution and homogenised at 3000 rpm for
30 s using an IKA T25 Ultra Turrax homogeniser (Wilmington, North Carolina, USA).
2.1.4.2. Agarose-ghee composites
Binary composites were prepared by maintaining the concentration of agarose at 1% (w/w) and
varying the ghee concentrations (0, 1, 3, 6, 9, 12 and 15%) with the remainder being milliQ
water to make the total volume of sample, 100 mL. I g of agarose powder was dispersed in
88
calculated amount of milliQ water and heated to 75°C. It was stirred for 10 minutes and cooled
to 45°C before the addition of ghee. The mixture was then homogenised using the IKA
homogeniser-T25 Turrax at 5000 rpm for 30s.
2.1.4.3. Agarose-MCC composites
99 ml of MCC solutions were prepared by adding varying concentrations of MCC (0.1, 0.3,
0.6, 0.9, 1.2 and 1.5 g) to calculated quantities of milliQ water. These solutions were stood
overnight at room temperature to ensure complete hydration of the MCC particles. These
suspensions were then heated to 80°C with constant magnetic stirring before adding 1g agarose
to the mixture. The agarose/MCC composites were stirred for 15 min for the complete
dissolution of agarose and cooled to 50°C for further analysis.
2.1.4.4. Agarose-gelatin composites
The concentration of gelatin in the composite was maintained at 3% and the concentration of
agarose was varied to yield 0.1, 0.4, 0.8, 1.2, 1.6, 2.0, 2.5 and 3% of agarose in the final
composite. Required amounts of agarose were dissolved in milliQ water at 75°C and allowed
to dissolve completely before cooling the mixture down to 60°C and adding gelatin. Samples
were stirred for 60 minutes for homogenous dissolution prior to analysis.
2.1.4.5. Staining for confocal laser scanning microscopy
The agarose/lipid composites in Chapters 3 and 4, were stained using 5mg/L Nile red, which
is extensively used to stain lipids. The stained sample solution was carefully pipetted into
precooled Coverwell imaging chamber for quick and homogenous gel formation.
In Chapter 5, the procedure outlined in the works of Russ et al. (2013) was followed to
covalently label agarose using 5-DTAF (5-(4, 6-Dichlorotriazinyl) Aminofluorescein). 1g of
agarose powder was dispersed in 50 ml milliQ water at heated to 80°C with constant stirring.
89
After cooling the solution to 60°C, 8mg 5-DTAF was added followed by the slow addition of
15 ml of 1%(w/v) Na2SO4. After the addition of 3-4 drops of 10% (w/v) NaOH, the mixture
was stirred for 2 hours before adding 120 ml absolute ethanol to get a yellow precipitate. The
solution was refrigerated for 20 min at -20°C before filtering it and washing the residue with
ethanol. The yellow residue was dried to obtain a fine agarose powder.
In chapter 6, the protein in the agarose-gelatin mixture was stained using 0.001% w/w FITC.
The dye was added to the composite and stirred for 60 min before allowing gel formation for
imaging.
2.2.2 Methods
2.2.2.1 Fourier Transform Infrared (FTIR) Spectroscopy
Spectroscopy focuses on, measures and interprets the spectra obtained due to the interaction of
electromagnetic radiation with materials (Penner, 2010). Spectroscopy can be further classified
into different types based on the region of electromagnetic spectrum (UV, visible light or IR),
type of interaction (absorption, transmission or diffraction) and type of matter (atomic or
molecular spectroscopy). FTIR specifically focuses on the infrared region of the
electromagnetic spectrum which can be further classified into near – IR (14000-4000 cm-1),
mid IR (4000-400 cm-1) and far-IR (400-10 cm-1) based on their wavenumbers. The mid-IR
spectra is the most commonly used range for FTIR analysis (Türker-Kaya & Huck, 2017).
When the material interacts with infrared radiation, it excites the molecules changing the
charge distribution of the molecule and subsequently, its electric dipole moment. The most
essential vibration modes include stretching and bending (rocking, twisting, torsioning,
scissoring and wagging) (Gaca-Zając et al., 2018). Distinct molecules have a unique vibrational
energy spectra characterized by its frequency, band shape and intensity, that serves as the
90
fingerprint for that particular molecule (Gutiérrez Sama et al., 2018).
Figure 2.3 Schematic Diagram of FTIR (Dole, Patel, Sawant, & Shedpure, 2011)
As depicted in Figure 2.3, a typical FTIR comprises of an infrared light source, a detector, a
fixed and a sliding mirror and a beam splitter. Infrared radiation from the source hits the beam
splitter and is ‘split’ into two and each half is diverted to the sliding and fixed mirror,
respectively. The beams are reflected back to the splitter at a point, where they combine to
induce either a destructive or constructive interference. The combined beam passes through the
sample to a detector which compiles an interferogram. The interferogram is transferred into the
source spectra with and without the absorption spectra of the sample. The proportion of the
source spectra to the absorption spectra is considered to be the sample’s FTIR transmission
91
spectrum (Komal Kumar & Devi Prasad, 2012).
Table 2.3 Absorption frequencies of various organic functional groups in the Mid- IR region
adapted from the works of Stuart, 2005
Groups Frequency (cm-1)
1470-1420 and 1380-1340 -CH2 and -CH3 bending Alkanes -CH stretching and bending 3000-2800
Olefinic –CH stretching 3100-3000 Alkenes
Acetylenic –CH stretching 3300 Alkynes
-CH stretching 3100-3000
-C=C- stretching 1600 Aromatics -OH bending 1500-1300
C-O stretching 1220-1000
C-O asymmetric stretching 1220-1000 Ethers
-NH stretching 3500-3300 Amines
-C=O stretching 1670-1640 Amides
-NH stretching 3500-3100
-NH bending 1640-1550
-C=O stretching 1735-1700 Aldehydes and Ketones
-C=O stretching 1740-1720 Carboxylic Acids
The IR radiation of a molecule can be determined by measuring its frequency as follows:
E = hv (1)
Where E represents the system’s energy, h denotes Planck’s constant (6.63 x 10-27 J.s) and
the frequency (in hertz) is given by v. Frequency is generally denoted in wavenumbers, ṽ
(cm-1) as follows:
92
ṽ = 1/ λ (2)
where λ is light’s wavelength in cm.
In the food industry, FTIR analysis is widely used as a technique in quality control. FTIR can
rapidly produce an IR spectra of the sample with great accuracy and with extreme convenience.
This can be co-related with the reference spectra to detect any anomalies that can affect the
quality of the product. The interpretation of the spectra and the peaks needs a fundamental
grasp of the composition and properties of food in terms of the micro and macro components
present and their molecular structure. Table 2.3 gives a broad classification of the absorption
frequencies of organic functional groups in the mid – IR region.
A Spectrum 100 with MIRacleTM ZnSe single reflection ATRplate from Perkin Elmer
(Norwalk, Connecticut) given in Figure 2.4 was employed for analysis.
93
Figure 2.4 FTIR- ATR Spectrum 100 at RMIT University
2.2.2.2 Differential Scanning Calorimetry
Calorimetric approaches have been widely used in food research for the quantitative and
qualitative analysis of the thermodynamic properties of the food systems. It studies and
measures the temperatures, the heat flow and enthalpy during the kinetic reactions and thermal
transitions of a material like heating, cooling, melting, crystallisation, denaturation, etc. All of
the reactions are classified into either endothermic (absorption of heat by the system) or
exothermic (release of heat) reactions.
Figure 2.5: Schematic diagram of a heat flux type DSC (Barron et al., 2012)
There are two types of Differential Scanning Calorimetry (Markus B. Dürrenberger, Stephan
Handschin, Béatrice Conde-Petit, & Felix Escher) – heat flux DSC and power compensation
DSC. Both types of DSC essentially study the heatflow in a system secluded from
environmental influences (Figure 2.5). In power compensation DSC, the sample is kept in a
hermetically sealed pan and an empty reference pan are places in two different chambers. Both
94
the pans are designed to be heated or cooled at the same rate (dT/dt). However, since one pan
is empty and the other contains the sample, the rate of supply of the thermal energy or the heat
flow (dH/dt) is different. In the heat flow DSC, the hermetically sealed sample and the empty
reference pan are kept in one sample chamber and heated. The heat flow (dH/dt) is the same
for both the pans but the rate of heating (dT/dt) achieved by both of them is different. A
computer is set up with the instrument to monitor and record, the initial temperature, the rate
of heating and the heat flow of the system (Guigo & Sbirrazzuoli, 2018).
When the sample is heated, the molecules or atoms gain energy to break free from the ordered
rigid state in solids, to a more unordered state in liquids. Figure 2.6 shows a typical example
of the effect of change in temperature on the material. Both exothermic (crystallisation) and
endothermic (melting) reactions or transitions can occur in a sample as a result of heating.
Figure 2.6: Typical material transition as a function of temperature change in a DSC adapted
from Verdonck, Schaap, & Thomas, 1999
A transition that is result of energy absorbed by the system, is termed as endothermic transition,
95
eg. melting. The energy absorbed by the system per unit mass to bring about the transition, is
known as latent heat (Emadi, Whiting, DJURDJEVIC, Kierkus, & Sokolowski, 2018). Glass
transition in amorphous materials has no latent heat associated with it and is referred to as a
second order transition.
At a fixed heating rate (dT/dt), the instrument can measure the system’s heat flow (dH/dt). The
heat capacity (Cp) of the material can be further calculated by dividing the heat flow (dH/dt)
by the rate of heating (dT/dt).
⁄ ] (3) 𝐶𝑝 = [(𝑑𝐻 𝑑𝑡⁄ ) (𝑑𝑇 𝑑𝑡⁄ )
When the mass of the sample (m), is known, the specific heat capacity (Cp/m) of the sample
𝑑𝐻
can be obtained as a function of temperature as follows (Thomas, 2005):
𝑑𝑡
(4) + 𝑓(𝑇, 𝑡) = 𝐶𝑝 𝑑𝑇 𝑑𝑡
Where dH/dt refers to the rate of total heat flow (mW or mJ/s)
Cp refers to the heat capacity of the sample in (J/°C)
dT/dt refers to the rate of heating (°C/min)
f(T,t) refers to heat flow as a function of temperature and time (mW)
A user defined heating/cooling rate and regime is used to measure the gain or loss of energy of
the system during it. The result is displayed as a thermogram as displayed in Figure 2.6. The
area of a peak in the thermogram relates to the energy lost or gained during the thermal
transition (Liu, Yao, Fan, Wang, & Hu, 2018). Two different types of DSC – micro and
96
modulated DSC, were employed in this study.
2.2.2.2.1 Micro Differential Scanning Calorimetry
The micro DSC is fabricated to work on bulk samples and dilute solutions with great versatility,
precision, and sensitivity. The characteristic feature of the micro DSC is its three-dimensional
sensor facilitated to achieve greater contact area with the sample yielding a higher degree of
sensitivity and accuracy. Sensitive Peltier elements are used to ensure contact with the
calorimetric block as shown in Figure 2.7 (Kodre, Attarde, Yendhe, Patil, & Barge, 2014).
Thermal conductors are used to ensure that the temperature of the vessel is same as that of the
calorimetric block. Thermistors, thermocouples, thermophiles, and resistance wires are used as
detectors to detect the heat flux of the sample in the vessel. Transducers are fit on both the
sample and reference vessel to eliminate changes common in both.
Figure 2.7: The heat flux micro DSC with a cylinder type measuring system (Kodre, Attarde,
Yendhe, Patil, & Barge, 2014).
97
Setaram VII from Setaram was employed in this research (Figure 2.8).
Figure 2.8: Micro DSC Setaram VII at RMIT University
2.2.2.2.2 Modulated Differential Scanning Calorimetry
Most food systems are complex composites made up of two or more components. Hence,
during a thermal regime, the heat flux of a given sample may comprise of multiple weak, broad
or overlapping transition signals arising from the different components, material impurities or
the presence of water. The modulated DSC overcomes this issue by delineating complex
transitions into individual contributing transitions. This is achieved by heating the samples with
two heating rates simultaneously, resulting in complementary information that facilitates the
deconvolution of the total heat flow into reversing and non reversing heat flow (Jana & Das,
98
2018). In addition, a sinusoidal modulation is applied in the modulated DSC on the linear
heating ramp, supplying uneven heatflow to the sample in a periodic fashion (Verdonck et al.,
1999). This causes to average sample temperature to continuously rise but in a non linear
fashion. The rate of the heating ramp, the period of modulation and the temperature amplitude
of modulation are critical parameters that need to be considered in the modulated DSC.
Figure 2.9: Modulated DSC Q2000 from TA instruments at RMIT University
2.2.2.3 Rheology
Rheology is the study of the flow and deformation of matter. First developed in synthetic
biopolymers owing to their extensive application in daily utilities, rheology has found
widespread application in various industries like food, pharmaceutical, ceramics, metals, to
name but a few. Food rheology has a direct impact on the final product and its acceptability as
it influences the product’s texture, flavour and appearance (Cheremisinoff, 2018). For example
the ease with which ketchup can be poured from its bottle, the disintegration of food in the
99
mouth affecting the release of flavour compounds, etc. Hence, rheology is one of the most
important components in product development and quality control. The main objectives of
performing rheological studies are:
i) Obtaining quantitative data of the materials’ mechanical properties
ii) Evaluating food texture and correlating it with sensory analysis
iii) Acquiring information pertaining to the material’s molecular composition and
arrangement
iv) Measuring and correlating quality of the medial and final products
v) Evaluating shelf life of products
In rheological measurements, the most important parameter used for quantification is viscosity,
which is defined as a material’s resistance to flow. The knowledge of a product’s viscosity
helps determine the processing and packaging parameters of a product. Viscosity
measurements are generally involve the use of a Brookefield viscometer and/or a rheometer.
Besides viscosity, other key rheological parameters used in rheological analysis are given in
Table 2.4.
Table 2.4 Rheological parameters
Parameter Definition Symbol Unit (SI)
Stress Force per unit area σ Pa
Strain Relative deformation in shear γ -
Shear rate Change of strain per unit time γ̇ s-1
Viscosity Resistance to flow η Pa.s
Storage Modulus Measure of solid like properties of Gʹ Pa
material
Loss Modulus Measure of fluid like properties of a Gʹʹ Pa
100
material
Phase angle Degree of viscoelasticity Tan δ -
Complex viscosity Resistance to flow of the sample in a η* Pa.s
structured state
Dynamic viscosity Internal friction of a liquid ηʹ Pa.s
Rheological behaviour of fluids and semi-solids can be broadly categorized as Newtonian or
Non-Newtonian behaviour. In Newtonian behaviour shear rate and shear stress exhibit a linear
relationship, with a plot beginning at the origin. Water, clarified juices, milk, etc show a
Newtonian behaviour.
Non-Newtonian materials like yoghurt, salad dressings ketchup, etc show a rheological
behaviour depended on time as a result of structural changes in the material. Non-Newtonian
behaviour includes shear thinning, shear thickening and time dependent shear shinning or shear
thickening. The shear stress-shear strain plot of Non-Newtonian fluids is either non-linear
and/or doesn’t begin at the origin. Shear thinning fluids, also known as pseudoplastic materials
exhibit an increase in shear stress with an increase in shear rate at a rate that is less than
proportional. Salad dressings, concentrated fruit juices are examples of shear thinning
materials. Materials like gelatinised starch dispersions show a shear thickening behaviour
where an increase in shear rate gives a more than proportional increase in the shear stress.
Many materials do not commence flow till the yield stress σ0 (the threshold of stress) is
exceeded, once they start flowing they may further exhibit shear thinning (thixotropic) or shear
thickening (antithixotropic) behaviour (Rao, 2010). The shear rate vs shear stress behaviour of
101
Newtonian and Non-Newtonian fluids is given in Figure 2.10.
Figure 2.10: Basic shear rate versus shear stress behaviour of Newtonian, shear thinning,
shear-thickening and Bingham and Herschel-Bulkley materials
Rheological parameters can be measured in a broad spectrum of materials ranging from solids
to liquids. Data available to the present day researchers has been based on the rheological
study of two extremes – Hookean solids and Newtonian liquids. The Hookean solid and the
Newtonian liquids are idealised examples of material behaviour as both establish a linear
relationship between stress and strain (or shear rate) based on direct proportionality,
irrespective of the applied force and amount of stress (Douglas, 2018). Modulus (Gʹ) is the
proportionality constant for solids,while viscosity (η) is the proportionality constant of liquids,
where modulus = σ / γ and viscosity = σ/(dγ/dt).
Majority of the food materials however, exhibit the ideal conditions of a Hookean solid or
Newtonian liquid only within a limited range of time and stress applied. More commonly, both
viscous (liquid like) and elastic (solid like) behaviour, termed as viscoelastic behaviour is
observed. Figure 2.11 shows the difference in behaviour between an ideal elastic solid, ideal
102
viscous liquid and a viscoelastic material (Shaw & MacKnight, 2018). When the time of
experimentation is short, all substances show a solid like behaviour. Whereas in an experiment
conducted over a longer period of time, the material is more likely to exhibit a viscous
behaviour. Dr. Marcus Reiner states that when given enough time, everything will flow
(Reiner, 1964). He proposed the following equation to describe this phenomenon
(5) 𝐷 = 𝜏 𝑇
Where D is the Deborah number, τ is the characteristic time of the material and T is the time
over which the experiment is run. When D < 1, liquid like behaviour is observed whilst, at D>1
materials exhibit a solid like behaviour (Barborik & Zatloukal, 2018).
Viscoelasticity of a material can be measured by using a rheometer to run small scale
deformation experiments. A bi-directional (oscillating) shear is used for viscoelastic
measurements as contrasted with the unidirectional oscillations used in viscosity
measurements. A sinusoidal stress or strain is applied on the material within what is known as
its ‘linear viscoelastic region (LVR)’. The pre-edtablished LVR is the domain where the
magnitudes of stress and strain are linearly related and the behaviour of the material is
independent of the stress or the strain applied. This implies that both the applied stress (or
strain) and the consequent response follow a sine wave and ‘excite’ the sample but does not
destroy its elastic structure. This is achieved by using a deformation that is relatively low or
slow, so as to maintain the sample’s structure close to its equilibrium. In large scale
deformation measurements used for measuring viscosity, permanent deformation of the sample
103
structure occurs.
Figure 2.11: The principle of oscillation viscometry. Applied strain versus time (a) and
resultant stress versus time that is measured in an elastic solid (b), Newtonian liquid (c) and
viscoelastic liquid (d).
Interpreting the data from rheological studies necessitates the knowledge of the terms: Elastic
modulus (Gʹ), loss modulus (Gʹʹ), tan delta (δ) and complex viscosity (η*). The elastic
properties of a sample, the strength of its network and its subsequent ability to withstand
deformation under applied stress or strain are all denoted by Gʹ, the storage modulus of the
sample. It takes into consideration the number and strength of the molecular interactions within
a sample. Whilst, Gʹʹ or the viscous modulus denotes the viscous properties of a sample and
deals with just the number of molecular interactions within the sample. If Gʹ > Gʹʹ, then the
sample is more like a solid than like a liquid. Conversely, when Gʹ < Gʹʹ the sample exhibits
104
liquid like properties(Dealy, Read, & Larson, 2018).
The complex modulus G*, which combines the storage and the loss moduli with equation 6
indicates the sample’s overall resistance to oscillatory shear.
G* = (Gʹ2 + Gʹʹ2)1/2 (6)
Another method for assessing the comparative importance of Gʹ and Gʹʹ is calculating the tan
δ or phase angle of the substance that is given by Equation 7. Tan δ is regarded as a measure
of viscosity.
Tan δ = Gʹʹ/Gʹ (7)
Materials with tan δ above 1 are extremely liquid-like whilst the value of tan δ is less than 1
for highly elastic substances.
In this study, TA Instruments’ Discovery Hybrid Rheometer (New Castle, DE, USA) was used
to perform small scale deformation oscillation experiments to obtain insight on the rheology of
materials (Figure 2.10). Keeping all variables constant while varying the following (one at a
time): time, temperature, frequency and strain provides important data about the materials
being examined (Puig-Rigall et al., 2018).
Strain Sweep
This is generally the first experiment conducted in rheological studies. It helps in determining
the LVR of the sample. The amplitude of the stress/strain is varied over a range and the
response of the sample is measured. The LVR is the region where the stress/strain and the Gʹ
are linearly related. The point where Gʹ shows a 10% deviation from the constant value, is
105
taken to be the end of the LVR region.
Figure 2.12: Discovery Hybrid Rheometer by TA Instruments at RMIT University
Time sweep
During time sweep, the sample is observed over a specific period while the other variables
(frequency, temperature, strain) are held constant. It enables the user to determine if there are
any changes in the properties of the sample over time. It indicates clearly the time taken by a
sample to reach an equilibrium state and also provides information regarding the storage
stability of the sample.
Temperature sweep
During the temperature sweep test, the sample is subjected to a predefine temperature range
while the other parameters are held constant. This allows the user to observe changes in the
sample’s properties as a function of temperature eg. gelation, gelatinisation, denaturation. It is
106
important to note that the scan rate employed for the test has a significant impact on the results.
Frequency sweep
Frequency, expressed as the inverse of time is the time needed to complete one oscillation. A
high frequency corresponds to a shorter time span and vice versa. The resultant spectrum of
the frequency sweep helps classify the sample that’s being examined into a weak gel, a strong
gel, a dilute solution or an entangled solution (Richardson & Kasapis, 1998).
2.2.2.4 Scanning Electron Microscopy
Scanning Electron Microscopy(SEM) was used to study the microstructure and phase topology
of the individual components in the biopolymer composites. It was employed to provide
tangible evidence of the phase separation and the presence of an architecture suitable for
rheological modeling.
SEM is capable of revealing surface information of a wide range of materials at the microspace
and nanospace which are inaccessible by light microscopy (magnification range 10 – 100,000x)
(Dudkiewicz et al., 2011). Conventional optical microscopy, on the other hand is limited to
examining particles 0.2 µm or greater in size (Cardott & Curtis, 2018). Another advantage of
SEM is its large imaging chamber which enables analysis of large samples. Furthermore, the
capacity to rotate, tilt and move the sample in three dimensions in the imaging chamber
increases its efficiency in probing the structure of the sample. SEM however, can’t capture
coloured images and has a lower resolution compared to transmission electron microscopy (Z.
Zhang, Law, & Lian, 2010). Typically, SEM consists of an electron gun that emits electrons, a
lens system that focuses the electrons, an electron detector that collects electrons from the
beam-specimen interaction, cathode ray tubes (CRT) that enable visual and photo recording
107
and miscellaneous electronics as shown in Figure 2.13.
Figure 2.13 Schematic diagram of a SEM (AMMRF, 2010)
The cathodic electron gun most commonly made of tungsten, emits electrons which are
accelerated to the anode through a voltage difference ranging between 0.1 to 50 keV. Pairs of
electromagnetic coils act as the focusing (condenser) lenses, the power of which depends on
the current passing through the lenses. The lenses focus the electron beam on to the specimen,
adjusting the spot size as much as necessary to improve the sharpness of the image. The
deflection coils in the objective lens are used to direct the beam across the sample in a raster
fashion scan. When the electron beam hits the specimen surface, and excites the atoms on the
surface creating signals. The signal created is specific to the structural features, chemical
proportion and crystallography characteristics of the specimen. The most common signals
generated by the SEM are secondary electrons (SE) and back scattered electrons (BSE) (Wang
& Petrova, 2012). Low energy secondary electrons (<50eV) provide the best resolution and
are best suited for generating the information about the sample topology. Whereas the
backscattered electrons have high energy (>50 eV), and reduce the resolution of the images.
108
BSE is most commonly used to probe into the sample composition (Fowler et al., 2018) .
Depending on the conditions of imaging, SEM is further classified into four main categories as
follows (Wandrol, Novák, Wu, Vesseur, & Vystavěl, 2018 ; Carter & Shieh, 2015):
Conventional (High Vacuum) SEM
This is the most common type of SEM where the sample is coated with a conducting
metal by a method called sputtering. A beam of electrons is used to excite the conducting
surface and create secondary electron emission. Imaging is done by recording either the SE or
BSE.
Low Vacuum SEM
When imaging non conducting samples like biological, archaeological or geological samples
among others which cannot be coated, SEM in a low vacuum mode is used. This works along
similar principle as the High vacuum SEM, other than the fact that the pressure in the chamber
can be adjusted (lowered) to prevent charging of the non conducting samples. Imaging is done
using BSE and the image quality is generally lower than that achieved in the high vacuum
mode due unpredictable electron discharge causing jumps or black spots in the image.
Cryo SEM
Cryo- SEM is specially used for imaging hydrated samples in their frozen state without altering
the sample properties. A cryo set up is added to the conventional high vacuum SEM that allows
the flash frozen samples to be viewed at sub-zero temperatures. The frozen samples can be
fractured or cracked to reveal internal structures. This is most suited for samples like hydrogels,
biofilms, foams, etc.
Environmental SEM (ESEM)
Imaging dynamic processes like melting, crystallisation or growth of biological specimens can
109
be made possible with use of an ESEM. The equipment set up allows the sample to be imaged
in its natural state without the need of any pre-processing steps. Sample temperature and the
relative humidity of the specimen chamber can be controlled allowing the user to wet, dry, heat
and cool the sample.
Figure 2.14: FEI Quanta 200 ESEM from FEI Corporate (Oregon, USA) at RMIT University
2.2.2.5 Confocal Laser Scanning Microscopy
The Confocal Microscope is based on the principle underlying Paul Nipkow’s Nipkow disk
which was a rotating disk with spirally arranged holes, which made it possible to scan and
110
transmit images via a telegraphic wire. The first application of confocal microscopy in food
science was published in 1987 (Heertje, 1987) and has witnessed an exponential rise in
applications since then.
Conventional optical microscopy, has been widely employed to get an insight of the food
structure at the meso scale and relate it to the texture and sensory attributes, along with the
physico-chemical properties of the final product since decades (Auty, 2013). However, light
microscopy requires pre-processing of the sample, such as freeze drying, sectioning, fixation,
or embedding, which may alter its microstructure (Plucknett et al., 2001). Confocal microscopy
is an adapted form of optical microscopy that has much in common with the epiflourescence
microscope. The development of confocal microscopy has helped overcome the problems in
light microscopy and has supplemented its advantages with the ability to capture images in
situ, image bulk specimens without sectioning them and obtain 3 dimensional images of the
specimen (N. Zhang et al., 2018).
In confocal microscopy, a diffraction limited laser beam from an optical fibre (generally from
an Argon source) is focused on a small subsurface spot on the sample via an objective lens as
shown in Figure 2.15. This enables imaging of a small section of the specimen as opposed to
the widefield illumination achieved by a typical optical microscope that illuminates the entire
sample. The incident laser excites the naturally present or labelled flourophores in the samples
at the given spot. The flourophores absorb the incident light and emit it at a longer wavelength.
The fluorescence emission is deflected via a dichoric mirror onto the condenser lens that
focuses it onto a photon detector via a pinhole. The photon detector, such as a photomultiplier
tube or CCD measures the intensity of the emitted fluorescence for this one spot. The signal is
then amplified and converted to picture element or pixel that makes up a digital image. The
focused spot is scanned across the sample, in the given focal plane in a raster fashion,
generating a 2D image pixel by pixel (Bayguinov et al., 2018). The pinhole is the distinguishing
111
feature of a confocal microscope as it controls the amount of fluorescence that reaches the
detector. The fluorescence from the out-of-focus planes is blocked by the pinhole stopping it
from reaching the photon detector. This makes it possible for the confocal microscope to focus
on any single focal plane along the z axis allowing emitted fluorescence from that specific
plane to reach the photon detector barring emissions from the out of focus planes. By adjusting
the pinhole, the z-depth or thickness can be adjusted. When the pinhole is completely open, the
confocal microscope is equivalent to a conventional scanning microscope (Blouin et al., 2018).
This 3D imaging technique, also known as Z-stack imaging has widely been used for
qualitative analysis, but the its use in quantitative analysis is scanty (Markus B. Dürrenberger
et al., 2001).
Figure 2.15: Illustration of the confocal imaging principle (solid lines = in-focus light; dashed
lines = out-of-focus light) (Franck, Hong, Maskarinec, Tirrell, & Ravichandran, 2007)
The laser is scanned across the sample in a raster fashion to form an image of various sizes,
112
most common being 512 x 512 or 1024 x 1024 pixels for a 2D image. A stepping motor fitted
to the microscope can be used to change the focal plane and a number of 2D images (x-y planes)
can be captured at consecutive focal planes through the z-axis. These multiple 2D images form
a Z – Stack that can be rendered into a 3D image. The presence of the pinhole facilitates the
CLSM to not only capture images with excellent resolution within the focal plane (≥ 0.25µm
in the x and y direction) but also obtains a good resolution between consecutive focal planes (≥
0.25µm in the z direction).
The resolution of the confocal microscope has the same limitations as those that affect light
microscopy. Theoretically, the diffraction limitation is given by the following equation (Zuo,
Liu, Cheng, Chen, & Tian, 2018):
∆x, ∆y = λ/2 n sinα (8)
Where α, n and λ are the aperture of the objective lens, refractive index of the sample and the
wavelength of the incident light respectively.
Along the optical axis, z-resolution, ∆z = λ/2 n sinα a minimum lateral resolution of 190nm
and an axial resolution of about 500nm is theoretically possible (Balashov et al., 2018). Under
the ideal conditions, in the absence of out of focus blur, close resolutions are attainable,
however, in practice, a z resolution of 750nm and an xy resolution of 250 nm is realistic (Auty,
2013).
Most commonly, optical sections are imaged in the z direction, in the xy planes perpendicular
to the laser beam. However, using the stepping principle, it is possible to capture images in the
xz or the yz plane. Further, scanning one image in time yields an x-t image. This makes it
possible to study any change in the fluorescent intensity in the given time frame. This feature
of the CLSM has been largely used in the biosciences to study chemical reactions in cells,
changes in ion concentrations, changes in pH and release of caged fluorescing probes due to
113
photo-activation. This technique is being explored in food studies and has been successful in
capturing in-situ images of gelation, aggregation and break up of β – lactoglobulin (Ko &
Gunasekaran, 2009; Olsson, Langton, & Hermansson, 2002) and baking of bread (Markus B
Dürrenberger, Stephan Handschin, Béatrice Conde-Petit, & Felix Escher, 2001) among others.
The advantages of CLSM over conventional light microscopy include (Pawley, 2006):
1) Ability to image bulk samples
2) Minimal sample preparation enables samples to be visualised in their native form
3) Optical sectioning makes it possible to capture a 3D Z Stack and subsequent 3D
reconstruction
4) Better resolution than conventional optical microscopy
5) In situ imaging enables capturing images of dynamic processes under controlled
environmental conditions using appropriate sample stages.
The disadvantages of CLSM include:
1) A flat surface is needed for imaging to ensure that the field of view has an even
illumination and emission signal
2) Flourescent labelling of the sample with a fluoroschrome may change intrinsic
properties and/or introduce non-native particles in the sample.
3) As the lateral resolution is limited to 250 nm by diffraction, colloidal food systems such
as casein micelles or interfacial films are difficult to properly resolve.
4) Common lasers have a limited number of excitation wavelengths, which occur over
narrow bands and risk overlapping and colocalisation
5) Since image is formed pixel by pixel as opposed to a snapshot in light microscopy,
imaging with CLSM is a much slower process. Imaging can be made faster by reducing
114
the sensitivity, resolution, etc.
2.2.2.6 Image Analysis
The Z-stack images captured by the CLSM were further analysed using two image analysis
software – FIJI with Java 6 and Imaris 9.0. Both the software employ distinct methods of
analysis and were hence used in the study to determine and compare their suitability in
accurately measuring the phase volume.
FIJI (FIJI Is Just ImageJ) is a free image processing software provided by ImageJ. It regards a
3D Z-stack as multiple 2D images and processes each image on the xy plane separately. The
result for each of the 2D images is then added and multiplied by the distance between two
consecutive xy planes to get volumetric measurements (Schneider, Rasband, & Eliceiri, 2012)
(Schindelin et al., 2012).
This is contrasted by Bitplane’s core scientific software module, Imaris. Imaris is made
exclusively for visualising and analysing 3D and 4D microscopy datasets. The Z stack images
captured using CLSM are rendered into 3D images by default and all measurements are done
on the single rendered 3D image (Li, Zhou, Voss, Quick, & LaBarbera, 2016).
The image analysis procedures followed and the results obtained are discussed and compared
115
in the experimental chapters.
References
AMMRF (2010). Electron Microscopy. Booklet.
Auty, M. (2013). Confocal microscopy: principles and applications to food microstructures
Food microstructures (pp. 96-P98): Elsevier.
Balashov, V., Efimov, A., Agapova, O., Pogorelov, A., Agapov, I., & Agladze, K. (2018). High
resolution 3D microscopy study of cardiomyocytes on polymer scaffold nanofibers
reveals formation of unusual sheathed structure. Acta Biomaterialia, 68, 214-222.
Barborik, T., & Zatloukal, M. (2018). Effect of die exit stress state, Deborah number, uniaxial
and planar extensional rheology on the neck-in phenomenon in polymeric flat film
production. Journal of Non-Newtonian Fluid Mechanics, 255, 39-56.
Barron, A. R., Allen, J., Bott, S., Jebb, M., Jiang, C., & Kienast, W. (2012). Physical methods
in chemistry and nano science. CONNEXIONS Rice University: Houston, TX, USA.
Bayguinov, P. O., Oakley, D. M., Shih, C. C., Geanon, D. J., Joens, M. S., & Fitzpatrick, J. A.
(2018). Modern Laser Scanning Confocal Microscopy. Current protocols in cytometry,
e39.
Blouin, S., Roschger, A., Varga, F., Misof, B., Spitzer, S., Roschger, P., & Klaushofer, K.
(2018). Confocal laser scanning microscopy—a powerful tool in bone research. Wiener
Medizinische Wochenschrift, 1-8.
Cardott, B. J., & Curtis, M. E. (2018). Identification and nanoporosity of macerals in coal by
scanning electron microscopy. International Journal of Coal Geology, 190, 205-217.
Carter, M., & Shieh, J. (2015). Chapter 5 – Microscopy. In M. Carter & J. Shieh (Eds.), Guide
to Research Techniques in Neuroscience (Second Edition) (pp. 117-144). San Diego:
Academic Press.
116
Cheremisinoff, N. P. (2018). Introduction to Polymer Rheology and Processing: 0: Crc Press.
Dealy, J. M., Read, D. J., & Larson, R. G. (2018). Structure and rheology of molten polymers:
from structure to flow behavior and back again: Carl Hanser Verlag GmbH Co KG.
Dole, M. N., Patel, P. A., Sawant, S. D., & Shedpure, P. S. (2011). Advance applications of
Fourier transform infrared spectroscopy. Int. J. Pharm. Sci. Rev. Res, 7(2), 159-166.
Douglas, J. (2018). Weak and Strong Gels and the Emergence of the Amorphous Solid State.
Gels, 4(1), 19.
Dudkiewicz, A., Tiede, K., Loeschner, K., Jensen, L. H. S., Jensen, E., Wierzbicki, R., . . .
Molhave, K. (2011). Characterization of nanomaterials in food by electron microscopy.
TrAC Trends in Analytical Chemistry, 30(1), 28-43.
Dürrenberger, M. B., Handschin, S., Conde-Petit, B., & Escher, F. (2001). Visualization of
Food Structure by Confocal Laser Scanning Microscopy (CLSM). LWT – Food Science
and Technology, 34(1), 11-17.
Dürrenberger, M. B., Handschin, S., Conde-Petit, B., & Escher, F. (2001). Visualization of
food structure by confocal laser scanning microscopy (CLSM). LWT-Food Science and
Technology, 34(1), 11-17.
Emadi, D., Whiting, L. V., DJURDJEVIC, M., Kierkus, W. T., & Sokolowski, J. (2018).
Comparison of newtonian and fourier thermal analysis techniques for calculation of
latent heat and solid fraction of aluminum alloys. Metallurgical and Materials
Engineering.
Fowler, C., Lynch, R., Shingler, D., Walsh, D., Carson, C., Neale, A., . . . Brown, A. (2018).
A Novel Electron-Microscopic Method for Measurement of Mineral Content in Enamel
Lesions. Archives of oral biology.
Franck, C., Hong, S., Maskarinec, S., Tirrell, D., & Ravichandran, G. (2007). Three-
117
dimensional full-field measurements of large deformations in soft materials using
confocal microscopy and digital volume correlation. Experimental Mechanics, 47(3),
427-438.
Gaca-Zając, K. Z., Smith, B. R., Nordon, A., Fletcher, A. J., Johnston, K., & Sefcik, J. (2018).
Investigation of IR and Raman spectra of species present in formaldehyde-water-
methanol systems. Vibrational Spectroscopy, 97, 44-54.
Guigo, N., & Sbirrazzuoli, N. (2018). Thermal Analysis of Biobased Polymers and
Composites. Handbook of Thermal Analysis and Calorimetry, 6, 399-429.
Gutiérrez Sama, S., Farenc, M., Barrère-Mangote, C., Lobinski, R., Afonso, C., Bouyssière,
B., & Giusti, P. (2018). Molecular Fingerprints and Speciation of Crude Oils and Heavy
Fractions Revealed by Molecular and Elemental Mass Spectrometry: Keystone
between Petroleomics, Metallopetroleomics, and Petrointeractomics. Energy & Fuels,
32(4), 4593-4605.
Jana, P. P., & Das, J. (2018). Precise estimation of glass transition and crystallization
temperatures of Zr55Cu30Ni5Al10 metallic glass using step-scan modulated
temperature differential scanning calorimeter. Thermochimica Acta, 660, 18-22.
Ko, S., & Gunasekaran, S. (2009). In situ microstructure evaluation during gelation of β-
lactoglobulin. Journal of food engineering, 90(2), 161-170.
Kodre, K., Attarde, S., Yendhe, P., Patil, R., & Barge, V. (2014). Differential scanning
calorimetry: A review. Research and Reviews: J. Pharmaceutical Analysis (RRJPA),
3(3), 11-22.
Komal Kumar, J., & Devi Prasad, A. (2012). Fourier transform infrared spectroscopy an
advanced technique for identification of biomolecules. Drug Invention Today, 4(12,
Co), 616-618.
Li, L., Zhou, Q., Voss, T. C., Quick, K. L., & LaBarbera, D. V. (2016). High-throughput
118
imaging: Focusing in on drug discovery in 3D. Methods, 96, 97-102.
Liu, Y., Yao, X., Fan, C., Wang, B., & Hu, J. (2018). Effect of a smectic liquid crystal polymer
as new β-nucleating agent on crystallization structure, melting, and rheological
behavior of isotactic polypropylene. Polymer Bulletin, 1-15.
Olsson, C., Langton, M., & Hermansson, A.-M. (2002). Dynamic measurements of β-
lactoglobulin structures during aggregation, gel formation and gel break-up in mixed
biopolymer systems. Food Hydrocolloids, 16(5), 477-488.
Pawley, J. B. (2006). Fundamental limits in confocal microscopy Handbook of biological
confocal microscopy (pp. 20-42): Springer.
Penner, M. H. (2010). Ultraviolet, visible, and fluorescence spectroscopy Food analysis (pp.
387-405): Springer.
Plucknett, K., Pomfret, S., Normand, V., Ferdinando, D., Veerman, C., Frith, W., & Norton, I.
(2001). Dynamic experimentation on the confocal laser scanning microscope:
application to soft‐solid, composite food materials. Journal of microscopy, 201(2), 279-
290.
Puig-Rigall, J., Obregon-Gomez, I., Monreal-Pérez, P., Radulescu, A., Blanco-Prieto, M. J.,
Dreiss, C. A., & González-Gaitano, G. (2018). Phase behaviour, micellar structure and
linear rheology of tetrablock copolymer Tetronic 908. Journal of colloid and interface
science, 524, 42-51.
Rao, M. A. (2010). Rheology of fluid and semisolid foods: principles and applications:
Springer Science & Business Media.
Reiner, M. (1964). The deborah number. Physics today, 17(1), 62.
Richardson, R. K., & Kasapis, S. (1998). Rheological methods in the characterisation of food
119
biopolymers Developments in Food Science (Vol. 39, pp. 1-48): Elsevier.
Schindelin, J., Arganda-Carreras, I., Frise, E., Kaynig, V., Longair, M., Pietzsch, T., . . .
Schmid, B. (2012). Fiji: an open-source platform for biological-image analysis. Nature
methods, 9(7), 676.
Schneider, C. A., Rasband, W. S., & Eliceiri, K. W. (2012). NIH Image to ImageJ: 25 years of
image analysis. Nature methods, 9(7), 671.
Shaw, M. T., & MacKnight, W. J. (2018). Introduction to polymer viscoelasticity: John Wiley
& Sons.
Stuart, B. (2005). Infrared spectroscopy. Kirk‐Othmer Encyclopedia of Chemical Technology.
Thomas, L. C. (2005). MDSC Paper# 2 Modulated DSC Basics; Calculation and Calibration
of MDSC Signals: New Castle: TA Instruments.
Türker-Kaya, S., & Huck, C. (2017). A review of mid-infrared and near-infrared imaging:
principles, concepts and applications in plant tissue analysis. Molecules, 22(1), 168.
Verdonck, E., Schaap, K., & Thomas, L. C. (1999). A discussion of the principles and
applications of Modulated Temperature DSC (MTDSC). International journal of
pharmaceutics, 192(1), 3-20.
Wandrol, P., Novák, L., Wu, M., Vesseur, E. J., & Vystavěl, T. (2018). In-situ Heating in SEM
and FIB/SEM Systems. Microscopy and Microanalysis, 24(S1), 336-337.
Doi:10.1017/S1431927618002179
Wang, Y., & Petrova, V. (2012). Scanning electron microscopy. Nanotechnology Research
Methods for Foods and Bioproducts, 103-126.
Zhang, N., Thompson, C. E., Townend, I. H., Rankin, K. E., Paterson, D. M., & Manning, A.
J. (2018). Non-destructive 3D imaging and quantification of hydrated biofilm-sediment
aggregates using X-ray micro-computed tomography. Environmental Science &
120
Technology.
Zhang, Z., Law, D., & Lian, G. (2010). Characterization methods of encapsulates
Encapsulation Technologies for Active Food Ingredients and Food Processing (pp.
101-125): Springer.
Zuo, R., Liu, W., Cheng, H., Chen, S., & Tian, J. (2018). Breaking the Diffraction Limit with
Radially Polarized Light Based on Dielectric Metalenses. Advanced Optical Materials,
121
1800795.
Chapter 3
Quantitative analysis of the phase volume of agarose- canola oil gels in comparison to blending law predictions using 3D imaging based on confocal laser scanning microscopy
Mhaske, P., Condict, L., Dokouhaki, M., Katopo, L., & Kasapis, S. (2019).
Quantitative analysis of the phase volume of agarose-canola oil gels in comparison to
blending law predictions using 3D imaging based on confocal laser scanning microscopy. Food
122
Research International, 125, 108529.
Food Research International 125 (2019) 108529
Contents lists available at ScienceDirect
Food Research International
journal homepage: www.elsevier.com/locate/foodres
T
Quantitative analysis of the phase volume of agarose-canola oil gels in comparison to blending law predictions using 3D imaging based on confocal laser scanning microscopy
⁎
Pranita Mhaske, Lloyd Condict, Mina Dokouhaki, Lita Katopo, Stefan Kasapis
School of Science, RMIT University, Bundoora West Campus, Plenty Road, Bundoora, VIC 3083, Australia
A R T I C L E I N F O
A B S T R A C T
Keywords: Blending law analysis Phase volume Confocal laser scanning microscopy Z-stack 3D imaging
Studying the phase behaviour of composite gels facilitates understanding of their structural and textural prop- erties at low and intermediate levels of solids. In this work, the phase behaviour of a model system of agarose including various concentrations of canola oil was studied. This was pursued using a variety of techniques including SEM, FTIR, microDSC and dynamic oscillation in-shear. The structural studies recorded strong, con- tinuous agarose networks supporting soft, discontinuous canola oil inclusions, with increasing levels of canola oil strengthening the composite system. A novel confocal laser scanning microscopy (CLSM) method for quantita- tive in situ examination of the oil phase volume was developed using three-dimensional (3D) imaging and image analysis software - FIJI and Imaris. Microscopic observations were assessed in relation to theoretical predictions from rheology-based blending-law analysis. Quantitative outcomes from the combined 3D imaging and image analysis are in close agreement with the volume predictions for the oil phase obtained from the isostrain blending law indicating the suitability of this approach in quantifying the phase behaviour of composite ma- terials. The results of this work indicate that the developed microscopic method shows promise and could be used in the determination of phase volume in more complex and industrially relevant systems.
1. Introduction
2018) and the tertiary system of agarose-gelatin-lipid (Shrinivas, Kasapis, & Tongdang, 2009).
synthetic polymer
Accurate determination of the phase volume for each component in the aqueous mixture is critical, since it determines the structural properties of the commercial formulation. Blending laws aim to achieve this but they reflect an indirect approach that relies on certain as- sumptions (e.g. the massive increase in effective molecular weight of the biopolymer at the onset of gelation) and a time-consuming method that requires considerable rheological work. On the other hand, con- focal laser scanning microscopy (CLSM) may offer a direct and rapid alternative to phase volume determination. One of CLSM's advantages is minimal sample manipulation, hence its capability to analyse the original state of a material without the pre-processing steps of freeze drying, sectioning or fixation, which are typically employed in other techniques (e.g. electron or optical microscopy) and may alter micro- structure (Auty, 2013).
CLSM can capture a sequence of two-dimensional (2D) optical sec- tions through a soft matrix to produce a three-dimensional (3D) image. Multiple 2D images are obtained by moving a focal plane, i.e. an x-y plane, on which the laser is focused, along the z-axis in steps of
By and large, industrial processing results in complex systems con- taining two or more components with well-defined physicochemical properties, with end products often possessing structural characteristics that are distinct from the individual constituents (Sikorski, Pokorny, & Damodaran, 2008). Examples include processed food, ceramics and plastic composites to name but a few. In the absence of a thermo- dynamic drive to heterotypic association, components tend to segregate leading to phase separation, which increases the “effective” con- centration of each constituent in its phase (Morris, 2009). Drawing from the “sophisticated” research (Nielsen, 1974; Takayanagi, Harima, & Iwata, 1963), workers in the area of bioma- terials adapted the blending law analysis to predict the mechanical properties of composites from the shear modulus of the individual components in relation to their effective concentration and phase vo- lume. Successful applications have been documented in various systems including agarose-whey protein (Katopo, Kasapis, & Hemar, 2012), bovine serum albumin-gelatin (Semasaka, Katopo, Buckow, & Kasapis, 2018), soy protein-wheat gluten (Dekkers, Emin, Boom, & van der Goot,
⁎
Corresponding author. E-mail address: stefan.kasapis@rmit.edu.au (S. Kasapis).
https://doi.org/10.1016/j.foodres.2019.108529 Received 18 November 2018; Received in revised form 18 May 2019; Accepted 26 June 2019
Available online 28 June 2019 0963-9969/ © 2019 Published by Elsevier Ltd.
123
P. Mhaske, et al.
Food Research International 125 (2019) 108529
specified thickness through the depth of the sample. This technique, also known as Z-stack imaging, offers unique insights into the material's true 3D structure. Numerous studies have demonstrated the applic- ability of CLSM in phase-separated mixtures including wheat protein- starch (Peighambardoust, Dadpour, & Dokouhaki, 2010), gelatin- starch-olive oil (Firoozmand & Rousseau, 2014) and whey protein iso- late-locust bean gum (Ur-Rehman et al., 2016), but its use has been focused on qualitative observations.
concentrations of canola oil were freeze dried using Labconco FriZone Triade freeze dryer (Kansas City, Missouri, USA) at −40 °C for 3 days. Then, they were placed on 12.6 mm carbon coated aluminium pins and gold coated. Surface micrographs were obtained using the FEI Quanta 200 ESEM (Hillsboro, Oregon, USA). An overview of the gel network was captured in the low vacuum mode with an accelerated electron beam of 25 kV and spot size of 5, with the working distance being about 10 mm at 50× magnification. Details of the agarose-canola oil topology were captured at a higher magnification of 400× in the high vacuum mode also using an accelerated electron beam of 25 kV and a spot size of 5.
2.2.3. Fourier transform infrared spectroscopy (FTIR)
A Perkin Elmer Spectrum Two FTIR spectrometer (Norway, CT, USA) fitted with a GladiATR from Pike Technologies (Wisconsin, United States) was used to obtain absorbance spectra for the samples of this −1 investigation. Measurements were carried out from 400 to 4000 cm −1 and averaged using 64 scans. The final spectra at a resolution of 4 cm of agarose-canola oil mixtures were collected by subtracting the water spectrum, and measurements were done in triplicate.
2.2.4. Differential scanning calorimetry (DSC)
Some examples of CLSM work, coupled with image analysis soft- ware, have been reported for the visualisation of water droplets in table spreads (Van Dalen, 2002), starch granules in guar gum-maize starch mixtures (Nagano, Tamaki, & Funami, 2008) and the fractal dimension of proteins in wheat dough (Li, Liu, Wu, & Zhang, 2017). Earlier at- tempts to develop CLSM-based techniques that can directly determine phase volume in mixtures were restricted to viscous aqueous solutions (Blonk, Van Eendenburg, Koning, Weisenborn, & Winkel, 1995) or single 2D images (Lorén, Langton, & Hermansson, 1999), hence they have not had widespread appeal that prevented further development. In this communication, a novel CLSM-based approach is developed using 3D imaging and quantitative image analysis. For the first appli- cation of this approach, we have chosen a mixture of agarose hydrogel and canola oil that remains in the liquid state at the experimental temperatures of this investigation (5 to 35 °C). The aqueous agarose network is thermodynamically incompatible with canola oil ensuring conditions of phase separation in the binary mixture (McClements, 2015). The aim is to estimate the phase volume of canola oil globules dispersed homogeneously within the supporting matrix of an agarose gel.
Thermal analysis of agarose-oil mixtures was performed using a microDSC Setaram VII (Seturau, Caluire, France). Samples of about 700 mg were transferred into aluminium cells and sealed, with the re- ference cell containing water at an equal weight. Both cells were then deposited into the instrument chamber and equilibrated at 25 °C for 60 min prior to ramping the temperature to 80 °C, cooling to 5 °C and then heating again to 80 °C. A constant scan rate of 1 °C/min was used throughout and all experiments were carried out in duplicate returning effectively overlapping thermograms.
Results from the microscopic experimentation will be compared with estimates from the corresponding rheological tests thus affording an evaluation of the predictions of rheological modelling on the phase behaviour of this model system. Successful application of this protocol to the agarose-canola oil gel will allow expansion to other systems of commercial interest, including mixtures with a solid lipid phase, sus- pending refined wood pulp (for example, microcrystalline cellulose), and matrices with two phase separated biopolymer networks where it is important but difficult to record directly the volumes and composition of the two phases; these, however, determine the textural character- istics of formulations (for example, blends of agarose and gelatin).
Additional measurements on canola oil were carried out using a modulated DSC by TA Instruments, Q2000 (New Castle, DE, USA). The instrument is fitted with a refrigerated cooling unit to reach tempera- tures well below 0 °C and a nitrogen purge cell with a flow rate of 20 ml/min. About 11 mg of the oil was hermetically sealed in an alu- minium pan with an empty pan being used as the reference. It was equilibrated at 25 °C for 60 min followed by successive heating and cooling routines between −80 and 80 °C at a scan rate of 1 °C/min. A modulated amplitude of ± 0.53 °C for 40 s was used throughout. Work was performed in duplicate producing consistent results.
2. Materials and methods
2.1. Materials
2.2.5. Rheological measurements
Agarose (Type-1) was purchased from Sigma-Aldrich (Sydney, Australia). The powder has an off-white colour, with its moisture and sulphate content being < 10 and 0.15%, respectively, according to the specification. Canola oil was purchased from Manildra supplier (Gladesville, Australia), and has a density of 0.918 g/ml.
2.2. Methods
2.2.1. Sample preparation
Small-deformation dynamic oscillation in-shear was used to mea- sure the elastic modulus (G′) and viscous modulus (G″) of both agarose and canola oil as well as their binary mixtures. Measurements were conducted with TA Instruments' Discovery Hybrid Rheometer (New Castle, DE, USA). A 40 mm diameter parallel plate geometry with 1 mm gap was employed. Samples were loaded on the Peltier plate at 45 °C and coated with polydimethylsiloxane to minimize evaporation. They were then cooled from 45 to 5 °C, held at 5 °C for 30 min followed by a frequency sweep at 5 °C from 0.1 to 100 rad/s and a heating run to 80 °C. A constant rate of 2 °C/min and a strain of 0.01%, which was within the linear viscoelastic region (LVR), were used throughout. Triplicate tests were carried out yielding congruent outcomes.
Binary mixtures of agarose and canola oil were prepared by adding a fixed amount of the polysaccharide power that would have produced 1% (w/w) agarose in the aqueous preparation, to the remainder of water once various concentrations of the oil (0, 1, 3, 5, 7, 9 and 11%, w/w) were added to a total of a hundred. In doing so, the agarose so- lution was made first by dissolving the powder in milliQ water at 80 °C with agitation for 10 min followed by cooling to 45 °C. Canola oil was then added to the agarose solution at this temperature followed by homogenisation using an IKA T25 Ultra Turrax homogeniser (Wilmington, North Carolina, USA) at 3000 rpm for 30 s.
2.2.2. SEM analysis
2.2.6. Confocal laser scanning microscopy (CLSM) 2.2.6.1. Sample preparation and image acquisition. Agarose-canola oil composites containing various concentrations of the oil were stained with Nile red (0.005 g/l), which is a sensitive stain for the detection of oil droplets, and then were placed into the well of a precooled Coverwell Press-Seal imaging chamber with 20 mm diameter and 2.8 mm depth (Sigma-Aldrich, NSW, Australia). Precooling of the chamber at 4 °C allowed a rapid and homogenous gel formation in situ following deposition of the warm solution. A Nikon A1R laser on an ECLIPSE Ti-E inverted microscope (Minato, Tokyo, Japan) was used to
Gel cubes of about 10 mm3 in size of agarose with varying
124
2
P. Mhaske, et al.
Food Research International 125 (2019) 108529
3,6-anhydro-galactose
−1), CeC bending (1396 cm (1068 cm bridges
generate confocal images. An excitation wavelength of 488 nm was used to induce fluorescence and an emission filter of 500–530 nm was applied to observe the stained oil droplets and collect images (Auty, 2013).
−1), CeH a series of chemical moieties, i.e. OeH stretching (2980 cm −1), CeOeC vibration vibration (2900 cm −1), of esters sulfate −1), and CeH angular deformation of anomeric carbon (1246 cm −1) (Almrhag et al., 2013; Samiey & Ashoori, 2012; Wang, Guo, (877 cm & Yang, 2011; Zamora-Mora, Velasco, Hernández, Mijangos, & Kumacheva, 2014).
Each 3D image was composed of 15 consecutive z-series of 2D images collected using a 10× objective and a z-axis step of 5.72 μm. The 3D image was generated with the xyz dimensions of 1279 μm × 1279 μm × 80 μm, respectively. Stained oil droplets were observed at up to 80 μm from the surface due to the significant loss of fluorescent signal at greater depths. A Z-stack of 2D images from nine sections across each sample was captured to avoid the effect of any inhomogeneous distribution of oil droplets.
Canola oil produces characteristic peaks representing symmetrical and asymmetrical CeH stretching vibration from CH2 groups of organic −1, respectively), C]O stretching of fatty acids (2856 and 2920 cm −1), CH3 aliphatic group bending vibration aliphatic esters (1746 cm −1), CH2 bending vibration or CeO stretching vibration of (1452 cm −1), and out of plane bending vibration of cis- aliphatic esters (1155 cm −1) (Rohman & Man, 2010; Zhang, Ma, Miao, HC]CHe (721 cm Tuchiya, & Chen, 2015). Mixtures of agarose and canola oil reproduce the characteristics of the individual components, with fingerprints from agarose dominating the interferograms. Since the mixtures show marked similarity with the individual components both in band wa- venumber and intensity, the earlier assumption in relation to the ab- sence of specific molecular interactions between the two components stands.
2.2.6.2. Image analysis. Obtained images were analysed using two image analysis software: FIJI with Java 6 software (Madison, Wisconsin, USA) and Imaris 9.1.0 (Bitplane, Belfast, Northern Ireland) in parallel. The former utilised the iterative ‘Default’ automatic global thresholding method (Calvard & Ridler, 1978), with the ‘analyse particle’ function then being used to estimate pixels denoting the lipid phase prior to multiplication by the Z-depth to obtain the volume fraction. The latter processed the obtained images using the ‘surface’ feature, the 3D images were then thresholded using the default method and the phase volume was estimated for the thresholded images.
2.2.7. Statistical analysis
Volume measurements of the oil phase were carried out in duplicate and data are expressed in mean ± standard deviation. Results were evaluated by one-way analysis of variance (ANOVA) using Minitab 18 (Minitab Inc., Pennsylvania, USA). Significance differences were de- fined as p < .05 with the Tukey test.
3. Results and discussion
Complementary characterization of the phase behaviour in aqueous preparations of agarose-canola oil mixtures is performed using micro differential scanning calorimetry. Fig. 3 depicts thermograms of the polysaccharide and mixtures that were cooled from 80 to 5 °C at 1 °C/ min. The former exhibits a relatively sharp exothermic peak with a mid- point transition temperature of 26 °C signifying the coil-to-helix tran- sition traversing the solution-to-gel consistency. Lipid inclusion also reveals the presence of a single peak, which remains unaltered in terms of temperature band and cooperativity. Increasing levels of canola oil generate thermal events of comparable enthalpy change arguing that the gelation process of the polysaccharide remains unaffected. Results further support the statement of microscopic phase separation between agarose and canola oil in the blend.
3.1. Experimental observations on the structural characteristics of agarose- canola oil mixtures
Scanning electron microscopy has been used in the past to unveil complex composite morphology (Shrinivas et al., 2009). In the present study it is employed to provide tangible evidence of phase separation in agarose-canola oil mixtures. Fig. 1a exhibits the SEM image of freeze- dried agarose gels at 50× magnification, where it is evident that there is a considerable degree of porosity in a honeycomb pattern. At the same magnification, we observed a similar mesh-type assembly in the presence of oil arguing for a continuous agarose phase in the composite but the oil droplets were not visible (images not shown). This suggests that the size of oil droplets is much smaller than the agarose pores and as such it requires a higher level of magnification for observation.
The absence of a noticeable calorimetric event attributed to the lipid constituent of the mixture prompted us to examine it separately from the polysaccharide. Thermal analysis of canola oil using microDSC fails to record a visible peak during cooling from 80 to 5 °C. A plausible explanation to this outcome lies in the composition of triacylglycerols (TAGs) in canola oil, which contains a high proportion of unsaturated TAGs crystallizing at very low temperature (Marikkar, Ghazali, Che Man, & Lai, 2002; Tang & Marangoni, 2007). To examine this, we employed modulated differential scanning calorimetry that was able to look for potential thermodynamic transitions at subzero temperatures. We observed an intense exothermic peak, culminating at about −57 °C, during a cooling routine between 80 and - 80 °C, as recorded in Fig. 3. This should correspond to the aggregation of unsaturated TAGs into clusters and the subsequent formation of a crystalline network. The subzero crystallisation temperature of canola oil shown in this work strongly supports our statement that canola oil exists as liquid inclu- sions in the agarose gel.
The agarose structure at higher magnification (400×) reveals a smooth featureless surface in Fig. 1b. In contrast, oil addition generates spheroidal clumps that correspond to the oil phase being dispersed within the agarose network (Fig. 1c,d). There is a noticeable alteration in the polysaccharide architecture showing a decreased level of smoothness with increasing oil concentration from 3 to 11% (w/w). It is worth noting that the size of the lipid domains remains fairly consistent (i.e. < 100 (80 ± 20) μm in diameter) with varying concentration arguing for limited lipid coalescence. Overall, the agarose-canola oil blends demonstrate a continuous polysaccharide structure supporting homogeneously dispersed oil globules. This topology provides the basis for quantitative analysis of the phase behaviour of the agarose-canola oil gel via theoretical modelling discussed in the next section.
Supramolecular evidence on the structure of the materials of this investigation is afforded by mechanical measurements, which en- compassed the individual components as a prelude to work on their mixtures. The general pattern of elastic modulus development for the single agarose preparation during controlled cooling at 2 °C/min from 50 to 5 °C is shown in Fig. 4. There is a sharp increase in the values of G′ at temperature below 30 °C indicating a gelation process in which the long and flexible coils of the polysaccharide transform into rigid double helices via hydrogen bonds. Subsequent aggregation of the double he- lices yields a three-dimensional network (Katopo et al., 2012). In con- trast, canola oil shows a Newtonian behaviour of a very low viscous modulus (G″ ~ 0.02 Pa), with the complex dynamic viscosity (η*) re- maining constant at about 0.16 Pa.s at 5 °C.
Following the measurements on single components, a series of binary mixtures was prepared by keeping the agarose at 1% (w/w) and
Results obtained from SEM indicate phase separation between polysaccharide and lipid. To confirm the absence of molecular inter- actions between the two constituents, we carried out Fourier transform infrared spectroscopy, and the interferograms for agarose, canola oil and their blends are illustrated in Fig. 2. The agarose spectrum exhibits
125
3
P. Mhaske, et al.
Food Research International 125 (2019) 108529
Fig. 1. SEM images of (a) 1% (w/w) agarose at 50× magnification, (b) 1% (w/w) agarose at 400× magnification, (c) 1% (w/w) agarose with 3% (w/w) canola oil at 400× magnification, and (d) 1% (w/w) agarose with 11% (w/w) canola oil at 400× magnification.
1.8
1068
2980
2900
1396
1246
1.6
877
1.4
1.2
1.0
0.8
e c n a b r o s b A
0.6
1746
2920
2856
1155
0.4
721
1452
illustrates the cooling profiles of blends recorded under the same ex- perimental routine as for the single systems. These follow similar pat- terns with the agarose network, an outcome that supports phase se- paration behaviour with the continuous agarose network suspending discontinuous inclusions of canola oil. Results from various techniques in Figs. 1 to 4 are consistent, and it is evident that the storage modulus of the composite gels increases with higher oil addition; G′ records values from 103.80 to 104.03 Pa at 1 to 11% (w/w) oil, respectively. Such enhancement in the overall rigidity of the mixture should be attributed to the rising agarose concentration in the polymer phase (effective concentration), which is the result of volume exclusion between aqu- eous and lipid phase.
0.2
0.0
3500
3000
2500
1500
1000
500
2000 Wavenumber (cm-1)
Fig. 2. FTIR spectra of 1% (w/w) agarose with 0, 1, 3, 5, 7, 9 and 11% (w/w) canola oil arranged from top to bottom (solid lines), and canola oil (dashed line).
varying the canola oil from 1 to 11% (w/w) in concentration steps of increases the effective concentration of the 2% which, of course, polysaccharide in the gradually diminishing aqueous phase. Fig. 4 also
This outcome is distinctly different to the results reported by Shrinivas et al. (2009) who have shown that in tertiary systems of 1% agarose, 10% gelatin and variable concentrations of soybean oil at 25 °C, when gelatin dominates the phase topology, the strength of the network gradually decreases on increasing concentrations of oil. These differences should be attributed to the much steeper calibration curve of agarose when compared to gelatin; for example, the log G' value of gelatin at a concentration of 30% (w/w) is ~ 4.5 (Pa), i.e. considerably lower than that of agarose at concentrations of < 5% (w/w) with a log G' value of ~5 (Pa). This seems to indicate that the initial strength of the continuous matrix has a substantial effect on how the addition of oil affects the mechanics of the aqueous/lipid composite system.
126
4
P. Mhaske, et al.
Food Research International 125 (2019) 108529
2.25
900
850
1.75
)
800
)
W m
(
750
1.25
W m
700
s e r u t x i m
0.75
650
600
0.25
550
500
l i o / e s o r a g a d n a
-0.25
( l i o a l o n a c f o w o l f t a e H
450
e s o r a g a
-0.75
400
350
f o w o l f
-1.25
300
t a e H
-1.75
250
-80
-60
-40
-20
0
20
40
60
Temperature (°C)
Fig. 3. Cooling profiles of 1% (w/w) agarose with 0, 1, 3, 5, 7, 9 and 11% (w/w) canola oil arranged from top to bottom (solid lines), and canola oil (dotted line) obtained using DSC at a scan rate of 1 °C/min.
+
=
1
ϕ x
ϕ y
(2)
3.2. Theoretical modelling of the phase behaviour in agarose-canola oil composites
where, G'c, G'x and G'y are the shear elastic moduli of the composite, X- phase constituent and Y-phase constituent, respectively, with ϕx and ϕy being the phase volumes of the two constituents. Eq. (1) is applicable to a format of a strong matrix supporting a soft filler (isostrain relationship in Morris, 2009), which is the topology of the current system; agarose network suspending liquid inclusions of canola oil.
To put the experimental observations of the preceding section on a quantitative footing, we constructed a calibration curve of elastic modulus for agarose networks at 5 °C. As shown in the inset of Fig. 4, increasing polysaccharide concentration from 1.0 to 2.5% (w/w) yields a double logarithmic plot with a good linear fit (r2 = 0.995). This in- formation is vital for modelling work that relates the overall mechan- ical property of the composite gel to the modulus of the two component phases via the so-called “blending laws”. Their mathematical expres- sion, being appropriate for this work, is as follows (Semasaka et al., 2018):
G
′
′ = C
′ + x
y
ϕ G x
ϕ G y
(1)
Fig. 5 displays an example of the application of the isostrain blending law (designed to model hard matrices supporting soft filler particles), for 1% (w/w) agarose with 3% (w/w) canola oil, but results are consistent for all combinations in the mixture. It shows a compu- terised calculation of shear modulus of the continuous agarose phase (G'ag) for all possible solvent (water) distributions in the agarose phase (Sag), i.e. from 0 to 1, in steps of 0.01. The values of the shear modulus
4.2
3.8
3.4
5.5
) a P ( '
3.0
5.0
) a P ( '
G g o L
4.5
2.6
y = 3.805x + 3.8521 R² = 0.995
G g o L
4.0
3.5
2.2
0
0.1
0.2
0.3
0.4
0.5
Log agarose concentration (%)
1.8
5
10
15
25
30
35
20 Temperature (°C)
Fig. 4. Cooling profiles of G' for 1% (w/w) agarose with 0 (○), 1% (■), 3% (x), 5% (●), 7% (Δ), 9% (-) and 11% (♦) (w/w) canola oil from 50 °C to 5 °C at a scan rate of 2 °C/min. The inset corresponds to a calibration curve of G' at 5 °C as a function of agarose concentration.
127
5
P. Mhaske, et al.
Food Research International 125 (2019) 108529
Fig. 5. Modelling the phase topology of 1% (w/w) agarose plus 3% (w/w) canola oil using the isostrain blending law. Predicted storage modulus of agarose (G'ag) and composite (G'C(BL)) are illustrated by dashed and solid lines, respec- tively, and the experimental composite modulus (G'exp) is shown by a dotted line. Sag is the solvent content of the agarose phase. The inset shows experimental G' values (■) and calculated phase volumes of canola oil (▲) for 1% (w/w) agarose with increasing oil concentration (solid line fit on the inset and the oil phase volume are derived from the isostrain blending law).
analysis of phase behaviour in agarose-canola oil composites, which involves selective fluorescent staining of oil droplets followed by con- focal imaging using Z-stack function and processing by image analysis software. A typical example of a single optical section (2D image) in the xy-plane captured from a section of 1% agarose with 5% canola oil gel is depicted in Fig. 6a. The oil droplets appear as purple spheres whereas the black regions correspond to the unstained agarose network. Fur- thermore, the droplets that are in the focal plane display bright purple colour whilst those that are outside the focal plane show a reduction in the colour intensity. Consecutive 2D images of the sample were ac- quired at a constant z-axis step of 5.72 μm (Fig. 6b) and the Z-stack image was rendered into a 3D image shown in Fig. 6c. To obtain esti- mations of the oil phase volume, the stack of 2D images (Fig. 6b) was processed using FIJI, whilst Imaris analysed the 3D image (Fig. 6c).
of the composite (G'C(BL)), which are calculated via the isostrain model of Eq. (1), and the experimental composite modulus (G'exp) are also depicted as a function of Sag. Gratifyingly, calculated and experimental values of the composite shear modulus intersect at the Sag value of 1.0, an outcome which indicates that the entirety of the solvent is dis- tributed within the agarose phase, hence demonstrating the capacity of the blending law to predict physically realistically results. This is also shown for the whole experimental range of increasing lipid con- centrations in the inset of Fig. 5, which illustrates a perfect fit between experimental moduli, obtained following frequency sweeps of the composites at 5 °C, and predicted moduli from blending-law analysis. Knowledge of solvent partition is crucial in estimating the phase volume of the individual components in composite systems, since it also predicts the effective (final) polymer concentration in its phase; generic relationships that predict solvent distribution for phase separated sys- tems are given in the following framework (Kasapis, Morris, Norton, & Clark, 1993; Morris, 2009):
tw
w
x = +
S (
×
)
(3)
x
x
tw
S
w
y = +
((1
−
)
×
)
y
x
(4)
=
ϕ x
tw
tw
tw x +
x
y
(5)
FIJI is an open source analysis software that processes individual 2D images, which make up a Z-stack (Schindelin et al., 2012). The 8-bit 2D image provided by CLSM comprises numerous pixels with the value of each pixel ranging from 0 to 255 (Luo et al., 2018). Pixels carry the intensity value of images, hence being vital to their quantification, an unstained sample results in a completely black image with the pixel intensity of all the image pixels being 0. This also acts as the negative control for imaging and the baseline calibration for image analysis. To correctly determine the number of pixels constituting an object (i.e. the oil droplet presently), an accurate description of the object edge is necessary. Thresholding was selected as the method for estimating the boundary of an object (Unnikrishnan, Pantofaru, & Hebert, 2007), since it can define a precise pixel intensity as the cut-off point between oil droplet and agarose background. Choosing a threshold that is too low leads to false positives, i.e. overestimating the phase volume, whereas too high a threshold causes false negatives that overlook oil droplets emitting a low-intensity signal. The threshold can be defined either manually or by applying automated algorithms. The former relies on the operator's practical skills for visually detecting the threshold and can be arbitrary, with inter-operator subjectivity impacting the re- producibility of results (Bergouignan et al., 2009). In contrast, auto- matic thresholding, which uses an algorithm to set the threshold, eliminates subjectivity and as such it was employed in this work.
In Eqs. (3) and (4), w, x, y and Sx are the total weight of water in the system, weight of polymer X, weight of polymer Y and the solvent fraction in X-phase, respectively. In Eq. (5), twx and twy refer to the total weights of X- and Y-phase, respectively. The phase volumes (ϕx and ϕy), in Eqs. (5) and (2), are defined by the total weight of the phases. The phase volume of canola oil in the mixtures, calculated from the isostrain blending law based on the Sag values of 1.0, is also plotted against the experimental oil concentration in the inset of Fig. 5. It was found that the predicted phase volume of the oil using Eqs. (3) to (5) tracks well its actual concentration in the mixture across the entire experimental range, hence its phase volume depends solely on its weight. Outcomes of theoretical modelling from rheological algorithms will be compared with those from 3D imaging and image analysis in the following sec- tion.
3.3. Development of a CLSM protocol and image analysis for agarose- canola oil composites
There are two categories of automatic thresholding that could be employed, i.e. global and local, with the former utilising a single threshold value for the entire image while the latter employs a different threshold value for each pixel depending on the intensity of adjacent pixels. Global thresholding works well for high quality images where
We developed a CLSM microscopic protocol
for quantitative
128
6
P. Mhaske, et al.
Food Research International 125 (2019) 108529
Fig. 6. (a) 2D, (b) Z-stack, and (c) 3D CLSM images of 1% (w/w) agarose plus 5% (w/w) canola oil. The oil droplets are either in purple (a-b) or grey (c), and the agarose is in black (a-c). (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)
estimated oil phase volume is closely matched to its experimental concentration in the composites, and Table 1 demonstrates that the best estimation is provided by the Default algorithm.
there is a good contrast between object and background pixels, while local thresholding is better suited for poorly illuminated images with either low contrast or multiple colour spectra (Sakila & Vijayarani, 2017). The CLSM images obtained in this study (Figs. 6a,b) reveal that the fluorescently labelled oil droplets can be clearly distinguished from the surrounding agarose background that has negligible fluorescence, indicating high contrast images, hence automatic global thresholding was utilised.
The ‘Default’ thresholding method employs an iterative algorithm, which randomly selects pixels from the image (oil droplets and agarose) and averages them to form an initial threshold value (Calvard & Ridler, 1978). This is used to achieve a separation, with the average of the pixels at or above the threshold value (i.e. oil droplets) and the average of the pixels below the threshold value (i.e. agarose) being computed. The average of the two means is then used as the optimal threshold limit for re-thresholding of the image. The process is repeated until the optimal threshold is greater than the composite average, as follows:
μ
μ TH (
)
TH (
)
+
bg
obj
THopt
2
(6)
= ⎛ ⎜ ⎝
⎞ ⎟ ⎠
where, THopt is the optimal threshold, and μbg and μobj are the average of the background and object pixel intensities separated by the initial threshold (TH).
FIJI includes sixteen distinct methods of global thresholding, each employing a different algorithm (Landini, 2017). To ascertain which algorithm works best for oil phase volume calculations, individual 3D images, from one section of each agarose-canola oil mixture, were thresholded using the sixteen algorithms. The oil volume calculation was carried out by extracting the total area of all droplets from the thresholded 2D images, regardless of the droplet diameter or circu- larity, and then multiplying the total area by the z-step between two optical cross sections in the xy-plane (i.e. 5.72 μm in this case). Table 1 summarises estimations of the oil phase volume derived from all the thresholding algorithms for a stack of 2D images in our composites. Blending law modelling in the preceding section argues that the
In our samples, an optimal threshold value of 55 was obtained from
129
7
P. Mhaske, et al.
Food Research International 125 (2019) 108529
Table 1 Comparison of canola oil phase volume in mixture with agarose calculated using different FIJI thresholding methods (n = 27).
Thresholding method
Phase volume of canola oil (%)
1% agarose plus 1% oil
1% agarose plus 3% oil
1% agarose plus 5% oil
1% agarose plus 9% oil
1% agarose plus 7% oil
1% agarose plus 11% oil
Default
Huang
Intermodes
IsoData
IJ IsoData
Li
MaxEntropy
Mean
MinError
9.39 ± 0.03 17.80 ± 0.59 12.99 ± 0.04 11.74 ± 0.52 11.53 ± 0.28 16.94 ± 0.21 12.28 ± 1.22 24.23 ± 0.62 84.04 ± 0.82
Minimum
Moments
Otsu
Percentile
Reniyentropy
3.02 ± 0.01 8.09 ± 1.12 3.36 ± 0.06 3.59 ± 0.03 3.47 ± 0.02 6.55 ± 0.08 3.89 ± 0.01 13.94 ± 0.11 46.02 ± 0.57 2.20 ± 0.11 3.69 ± 0.09 3.59 ± 0.08 42.34 ± 0.75 3.85 ± 0.24 0.01 ± 0.00
5.32 ± 0.03 18.90 ± 0.22 2.59 ± 0.38 7.36 ± 0.33 6.78 ± 0.33 14.08 ± 0.72 6.16 ± 0.47 21.61 ± 0.48 48.50 ± 0.49 0.01 ± 0.00 6.89 ± 0.15 7.36 ± 0.07 45.44 ± 1.32 6.36 ± 0.54 0.0013 ± 0.02
Shanbagh
1.16 ± 0.03 36.11 ± 0.14 1.18 ± 0.04 1.21 ± 0.09 1.18 ± 0.03 2.25 ± 0.09 1.66 ± 0.17 17.58 ± 0.58 32.31 ± 0.23 0.26 ± 0.05 1.23 ± 0.05 1.24 ± 0.04 37.15 ± 0.17 1.66 ± 0.13 0.0006 ± 0.00
Triangle
Yen
10.99 ± 0.19 4.88 ± 0.23
41.98 ± 0.13 6.46 ± 0.50
7.33 ± 0.16 30.11 ± 1.27 3.55 ± 0.27 12.51 ± 0.43 11.28 ± 0.21 23.49 ± 0.28 5.10 ± 0.15 26.99 ± 0.23 56.19 ± 0.86 0.05 ± 0.01 7.65 ± 0.47 8.84 ± 0.30 46.90 ± 0.76 7.65 ± 0.20 0.004 ± 0.00 50.29 ± 0.02 9.60 ± 0.26
8.23 ± 0.49 11.95 ± 0.54 11.74 ± 0.25 46.96 ± 0.88 12.39 ± 0.09 1.57 ± 0.11 22.13 ± 1.70 13.67 ± 0.38
10.72 ± 0.15 15.32 ± 0.29 9.71 ± 0.18 9.38 ± 0.21 9.24 ± 0.08 14.09 ± 0.30 11.92 ± 0.11 17.77 ± 0.30 65.92 ± 0.91 4.36 ± 0.44 10.06 ± 1.05 10.60 ± 0.29 43.44 ± 0.40 11.60 ± 1.56 0.06 ± 0.04 21.51 ± 1.12 12.06 ± 1.08
29.59 ± 0.40 1.61 ± 0.08
12
Fiji
Imaris
11
)
10
%
9
8
7
6
5
4
3
2
( l i o a l o n a c f o e m u l o v e s a h P
1
0
1%
3%
5%
7%
9%
11%
Concentration of canola oil in composite with agarose (%)
Fig. 7. Phase volume of canola oil in mixture with 1% (w/w) agarose determined via a Z-stack method coupled with image analysis software (FIJI and Imaris).
software was found to be utilising the ‘Default’ automatic global thresholding method and the ‘analyse particle’ function, prior to mul- tiplication by the Z-depth to obtain the volume fraction of the lipid phase.
Microscopic work was complemented with another image proces- sing tool (Imaris) that employs an alternative way of phase volume estimation (Calvard & Ridler, 1978; Fishman et al., 2006; Fogarty, Hammond, Kanjhan, Bellingham, & Noakes, 2013). We used a feature of this licensed software called maximum intensity projection (MIP) to transform a stack of 2D confocal images of the agarose-canola oil gel
the Default algorithm using Eq. (6). In addition, we found that this is the only algorithm that can consistently compute the phase volumes as a function of increasing oil levels in the various composites (Table 1). Therefore, the Default thresholding method was applied to determine the phase volume of oil in all 3D images, captured from nine sections across each sample, and outcomes are given in Fig. 7. Accordingly, FIJI's image-analysis predictions of oil phase volumes are very close to the experimental concentrations in the composites, and these are also congruent with the phase volume estimates from blending law model- ling (inset of Fig. 5). Therefore, the best suited methodology for the FIJI
130
8
P. Mhaske, et al.
Food Research International 125 (2019) 108529
approach lays the groundwork for future research, which could utilise more complex biomaterial composites, e.g. two hydrocolloid aqueous phase supporting lipid inclusions, with the goal of accurately and ra- pidly determining phase behaviour in industrially relevant matrices.
into a 3D image. Then, we activated another feature in Imaris called “Surfaces” for identification of the object of interest (i.e. oil droplets) and subsequent image quantification. In doing so, a Gaussian filter was applied on a 3D image of the present composite to smoothen it followed by subtraction of the background intensity from the image to imple- ment a thresholding process.
Acknowledgement
The authors acknowledge the facilities, and the scientific and the RMIT University's Microscopy &
assistance of
technical Microanalysis Facility.
Funding
This research did not receive any specific grant from funding
agencies in the public, commercial, or not-for-profit sectors.
Declarations of interest
None.
References
Once the above were accomplished, measurements could be carried out with a marching cube algorithm (Lorensen & Cline, 1987). The algorithm works by creating a polygonal mesh around the oil droplets and those contained within the mesh were analysed volumetrically. To ensure that all oil droplets are included in the polygonal mesh selection, we used the diameter of the largest oil droplet as the limiting factor. Measurement of the droplet diameter is performed manually in Imaris, using a feature called “Slice”, and it was found that the largest diameter of oil droplets in the images of all mixtures varied from 80 to 100 μm. It should be noted that in some cases boundaries are in contact with an- other and they may be considered as one larger droplet in Imaris. Therefore, measurements were carried out taking into consideration the possible presence of fused oil droplets yielding an imaging limiting factor of a large diameter. As such the methodology determined to be best suited for this application in the Imaris software was found to be processing of the obtained images using the ‘surface’ feature, followed by thresholding using the default method followed by phase volume estimation.
Almrhag, O., George, P., Bannikova, A., Katopo, L., Chaudhary, D., & Kasapis, S. (2013). Investigation on the phase behaviour of gelatin/agarose mixture in an environment of reduced solvent quality. Food Chemistry, 136, 835–842.
Auty, M. A. E. (2013). 4 - Confocal microscopy: Principles and applications to food mi- crostructures. In V. J. Morris, & K. Groves (Eds.). Food microstructures (pp. 96–98). Woodhead Publishing.
Bergouignan, L., Chupin, M., Czechowska, Y., Kinkingnéhun, S., Lemogne, C., Le Bastard,
G., & Fossati, P. (2009). Can voxel based morphometry, manual segmentation and automated segmentation equally detect hippocampal volume differences in acute depression? NeuroImage, 45, 29–37.
Blonk, J., Van Eendenburg, J., Koning, M., Weisenborn, P., & Winkel, C. (1995). A new CSLM-based method for determination of the phase behaviour of aqueous mixtures of biopolymers. Carbohydrate Polymers, 28, 287–295.
Calvard, T., & Ridler, S. (1978). Picture thresholding using an iterative selection method.
IEEE Transactions on Systems, Man, and Cybernetics, 8, 630–632.
Dekkers, B. L., Emin, M. A., Boom, R. M., & van der Goot, A. J. (2018). The phase
properties of soy protein and wheat gluten in a blend for fibrous structure formation. Food Hydrocolloids, 79, 273–281.
Firoozmand, H., & Rousseau, D. (2014). Tailoring the morphology and rheology of phase-
separated biopolymer gels using microbial cells as structure modifiers. Food Hydrocolloids, 42, 204–214.
Fishman, E. K., Ney, D. R., Heath, D. G., Corl, F. M., Horton, K. M., & Johnson, P. T.
The above protocol allowed volume determination of oil droplets within the specified mesh. It was applied on 3D images from nine sections of each composite gel, and results are also reproduced in Fig. 7. They show that the Imaris-based phase volumes are closely matched to the experimental oil concentration in the mixtures and as such are in good agreement with those predicted using the rheology-based blending law. Statistical analysis conducted on the phase volumes ob- tained by Imaris and FIJI for each composite preparation show no significant difference (p > .05) between the two imaging tools. It ap- pears that although two distinct processing techniques have been de- veloped, i.e. Imaris processes 3D images whereas FIJI does so on the 2D counterparts, they can both determine accurately the phase volume of canola oil in agarose gels. These inclusions appear to belong to the same statistical population, giving us confidence in the application of the developed protocols for comparison with the theoretical modelling from rheology.
(2006). Volume rendering versus maximum intensity projection in CT angiography: What works best, when, and why. Radiographics, 26(3), 905–922.
4. Conclusions
Fogarty, M., Hammond, L., Kanjhan, R., Bellingham, M., & Noakes, P. (2013). A method for the three-dimensional reconstruction of Neurobiotin™-filled neurons and the lo- cation of their synaptic inputs. Frontiers in Neural Circuits, 7, 153.
Kasapis, S., Morris, E. R., Norton, I. T., & Clark, A. H. (1993). Phase equilibria and ge-
lation in gelatin/maltodextrin systems - Part IV: Composition-dependence of mixed- gel moduli. Carbohydrate Polymers, 21, 269–276.
Katopo, L., Kasapis, S., & Hemar, Y. (2012). Segregative phase separation in agarose/ whey protein systems induced by sequence-dependent trapping and change in pH. Carbohydrate Polymers, 87, 2100–2108.
Landini, G. (2017, 29 September). Auto threshold. Retrieved from http://imagej.net/
Auto_Threshold.
Li, Q., Liu, R., Wu, T., & Zhang, M. (2017). Interactions between soluble dietary fibers and wheat gluten in dough studied by confocal laser scanning microscopy. Food Research International, 95, 19–27.
Lorén, N., Langton, M., & Hermansson, A.-M. (1999). Confocal laser scanning microscopy and image analysis of kinetically trapped phase-separated gelatin/maltodextrin gels. Food Hydrocolloids, 13, 185–198.
Lorensen, W. E., & Cline, H. E. (1987). Marching cubes: A high resolution 3D surface construction algorithm. Paper presented at the ACM siggraph computer graphics.
Luo, T. L., Eisenberg, M. C., Hayashi, M. A. L., Gonzalez-Cabezas, C., Foxman, B., Marrs, C. F., & Rickard, A. H. (2018). A sensitive thresholding method for confocal laser scanning microscope image stacks of microbial biofilms. Scientific Reports, 8(1), 13013.
Through the years, rheology-based blending law analysis has been extensively utilised to predict phase equilibria and gelation properties in binary gels. Ultimately, blending laws provide estimates of the component phase volume that determines structural performance in composites. Recent advances in imaging technology (mainly for clinical analysis and medical intervention) have allowed us to develop a novel microscopic method using Z-stack CLSM, in combination with image processing tools, for application in a soft biocomposites. In particular, the model system of this study comprised a binary composite of a continuous agarose network supporting discontinuous inclusions of canola oil that show a distinct phase separation. The developed mi- croscopic method was successful in determining the phase volume of spheroidal oil inclusions in a range of mixed gel formulations. Results agree remarkably well with phase volume predictions from the isostrain model of blending laws also obtained presently for agarose-canola oil systems.
Marikkar, J. M. N., Ghazali, H. M., Che Man, Y. B., & Lai, O. M. (2002). The use of cooling and heating thermograms for monitoring of tallow, lard and chicken fat adulterations in canola oil. Food Research International, 35, 1007–1014.
McClements, D. J. (2015). Emulsion formation. Food emulsions: principles, practices, and
techniques(3rd ed.). Boca Raton: CRC Press.
Morris, E. R. (2009). Funtional interactions in gelling biopolymer mixtures. In S. Kasapis, I. T. Norton, & J. B. Ubbink (Eds.). Modern biopolymer science – Bridging the divide between fundamental treatise and industrial application. New York: Elsevier. Nagano, T., Tamaki, E., & Funami, T. (2008). Influence of guar gum on granule
morphologies and rheological properties of maize starch. Carbohydrate Polymers, 72,
Furthermore, this is the first time that the validity of rheological predictions in determining phase volume in an aqueous/lipid phase has been tested, as opposed to the widely used application of rheological modelling in aqueous mixtures. Known concentrations of canola oil were used in this study to ensure known phase volumes, such that the validity of both the theoretical/rheological and microscopic meth- odologies for determining phase volume could be assessed. The working
131
9
P. Mhaske, et al.
Food Research International 125 (2019) 108529
95–101.
of components in food system. In S. Damodaran, K. L. Parkin, & O. R. Fennema (Eds.). Fennema's food chemistry(4th ed.). Boca Raton: CRC Press.
Nielsen, L. E. (1974). Morphology and the the elastic modulus of block polymers and
polyblends. Rheologica Acta, 13, 86–92.
Peighambardoust, S., Dadpour, M., & Dokouhaki, M. (2010). Application of epi-
Takayanagi, M., Harima, H., & Iwata, Y. (1963). Viscoelastic behavior of polymer blends and its comparison with model experiments. 23, Memoirs-Faculty of Engineering Kyushu University1–13.
fluorescence light microscopy (EFLM) to study the microstructure of wheat dough: A comparison with confocal scanning laser microscopy (CSLM) technique. Journal of Cereal Science, 51, 21–27.
Rohman, A., & Man, Y. B. C. (2010). Fourier transform infrared (FTIR) spectroscopy for
Tang, D., & Marangoni, A. G. (2007). Modeling the rheological properties and structure of colloidal fat crystal networks. Trends in Food Science & Technology, 18, 474–483. Unnikrishnan, R., Pantofaru, C., & Hebert, M. (2007). Toward objective evaluation of image segmentation algorithms. IEEE Transactions on Pattern Analysis and Machine Intelligence, 29, 929–944.
analysis of extra virgin olive oil adulterated with palm oil. Food Research International, 43, 886–892.
Ur-Rehman, A., Khan, N. M., Ali, F., Khan, H., Khan, Z. U., Jan, A. K., & Ahmad, S. (2016). Kinetics study of biopolymers mixture with the help of confocal laser scanning mi- croscopy. Journal of Food Process Engineering, 39, 533–541.
Sakila, A., & Vijayarani, S. (2017). A hybrid approach for document image binarization. Paper presented at the 2017 International Conference on Inventive Computing and Informatics (ICICI).
Samiey, B., & Ashoori, F. (2012). Adsorptive removal of methylene blue by agar: Effects of
NaCl and ethanol. Chemistry Central Journal, 6, 14.
Van Dalen, G. (2002). Determination of the water droplet size distribution of fat spreads using confocal scanning laser microscopy. Journal of Microscopy, 208, 116–133. Wang, W., Guo, X., & Yang, Y. (2011). Lithium iodide effect on the electrochemical be-
Schindelin, J., Arganda-Carreras, I., Frise, E., Kaynig, V., Longair, M., Pietzsch, T., & Cardona, A. (2012). Fiji: An open-source platform for biological-image analysis. Nature Methods, 9, 676.
havior of agarose based polymer electrolyte for dye-sensitized solar cell. Electrochimica Acta, 56, 7347–7351.
Semasaka, C., Katopo, L., Buckow, R., & Kasapis, S. (2018). Modeling water partition in composite gels of BSA with gelatin following thermal treatment. Food Hydrocolloids, 76, 141–149.
Zamora-Mora, V., Velasco, D., Hernández, R., Mijangos, C., & Kumacheva, E. (2014). Chitosan/agarose hydrogels: Cooperative properties and microfluidic preparation. Carbohydrate Polymers, 111, 348–355.
Shrinivas, P., Kasapis, S., & Tongdang, T. (2009). Morphology and mechanical properties of bicontinuous gels of agarose and gelatin and the effect of added lipid phase. Langmuir, 25, 8763–8773.
Zhang, H., Ma, J., Miao, Y., Tuchiya, T., & Chen, J. Y. (2015). Analysis of carbonyl value of frying oil by fourier transform infrared spectroscopy. Journal of Oleo Science, 64, 375–380.
Sikorski, Z. E., Pokorny, J., & Damodaran, S. (2008). Physical and chemical interactions
132
10
Chapter 4
Phase volume quantification of agarose-ghee gels using
3D confocal laser scanning microscopy and blending law
analysis: A comparison
Mhaske, P., Condict, L., Dokouhaki, M., Farahnaky, A., & Kasapis, S. (2020). Phase volume
quantification of agarose-ghee gels using 3D confocal laser scanning microscopy and
133
blending law analysis: A comparison. LWT, 129, 109567.
LWT - Food Science and Technology 129 (2020) 109567
Contents lists available at ScienceDirect
LWT - Food Science and Technology
journal homepage: www.elsevier.com/locate/lwt
T
Phase volume quantification of agarose-ghee gels using 3D confocal laser scanning microscopy and blending law analysis: A comparison
∗
Pranita Mhaske, Lloyd Condict, Mina Dokouhaki, Asgar Farahnaky, Stefan Kasapis
School of Science, RMIT University, Bundoora West Campus, Plenty Road, Melbourne, VIC, 3083, Australia
A R T I C L E I N F O
A B S T R A C T
Keywords: Lewis-Nielsen blending law Phase behaviour Confocal laser scanning microscopy 3D imaging Image analysis
A thorough understanding of the phase behaviour of biomaterial composites is imperative for manipulating the structural and textural properties in novel food products. This study probed the phase behaviour of a model system comprising agarose and a varying concentration of ghee. Results obtained from scanning electron mi- croscopy (SEM), micro differential scanning calorimetry (DSC), Fourier transform infrared spectroscopy (FTIR) and dynamic oscillation in-shear revealed discontinuous and hard inclusions of ghee reinforcing the continuous, weaker agarose matrix with increasing concentrations of the former. Phase behaviour of the system was quantified in parallel with a novel method combining 3D confocal laser scanning microscopy (CLSM) imaging and image analysis software - FIJI and Imaris - in an effort to substantiate the efficacy of the microscopic protocol in quantifying phase behaviour. Phase volumes recorded with the microscopic protocol were in close agreement to those modelled with the Lewis-Nielsen blending law using small-deformation dynamic oscillation. However, results indicated that the inner filtering effect or ‘self-shadowing’ observed commonly in CLSM images may pose a limitation to the application of this technique, necessitating further development before it can be applied to more complex, industrially relevant systems.
1. Introduction
adapting the classic works of Nielsen (1974) and Takayanagi et al. (1963) in biopolymer research (Semasaka, Buckow, & Kasapis, 2018). This approach has been successfully applied to a plethora of studies in relation to the development specific textures, fat replacement, etc. (Dekkers, Emin, Boom, & van der Goot, 2018; Kasapis & Tay, 2009). However, the application of blending laws is a time-consuming exercise necessitating a significant amount of rheological experimentation and modeling expertise. They are, as such an indirect application, and until recently were confined to aqueous biopolymer mixtures (Mhaske, Condict, Dokouhaki, Katopo, & Kasapis, 2019).
Over the years, attempts have been made to develop a direct method using confocal laser scanning microscopy (CLSM) to study and quantify phase behaviour and the effective phase volume of components in binary mixtures. Such studies aimed mainly to qualitatively probe various images of biopolymer mixtures, e.g. interpenetrating networks of soy protein and sugar beet pectin (Hou et al., 2015). They also at- tempted to quantitatively analyse aqueous systems and 2D images (Blonk, van Eendenburg, Koning, Weisenborn, & Winkel, 1995; Lorén, Langton, & Hermansson, 1999), but were not met with widespread appeal leading to further development.
In order to keep up with the increasing competition in the market, the food industry is continuously challenged to improve products in terms of value addition, cost efficiency and contribution to maintaining a healthy lifestyle without compromising on their quality, taste and texture (Kasapis, 2008). As a consequence, the industry has seen an increase in the use of biopolymers such as proteins and dietary fibre not only to reinforce the nutritional value but also to act as thickening and stabilizing agents, creating formulations with intriguing and distinct textural properties (Çakır & Foegeding, 2011). Mixtures of biopolymers widen the sensory window of products by enabling the manipulation of microstructure and textural properties in semi-solid foods. This is made possible due to the thermodynamic imbalance that segregates the binary mixture into distinct phases, depending on the morphology of the microstructure, hence altering the product's textural properties. Segregative phase separation commonly leads to the enrichment of the two components in distinct phases with one phase acting as a con- tinuous matrix supporting the second discontinuous “filler” (Gaaloul, Turgeon, & Corredig, 2010).
Recently, a novel technique of quantifying phase volume in agarose- canola oil composites using 3D imaging coupled with image analysis
By considering these phase-separated components as distinct in- dividual systems, their mechanical properties can be predicted by
∗
Corresponding author. E-mail address: stefan.kasapis@rmit.edu.au (S. Kasapis).
https://doi.org/10.1016/j.lwt.2020.109567 Received 19 September 2019; Received in revised form 5 May 2020; Accepted 6 May 2020
Available online 16 May 2020 0023-6438/ © 2020 Published by Elsevier Ltd.
134
P. Mhaske, et al.
LWT - Food Science and Technology 129 (2020) 109567
2.3.2. Fourier transform infrared spectroscopy (FTIR)
Single and binary mixtures of agarose and ghee were analysed using a PerkinElmer Spectrum 2 FTIR spectrometer (Norwalk, Connecticut, United States) with a GladiATR (Pike Technologies, Maddison, Wisconsin, United States) diamond crystal attenuated total reflectance (ATR) device. The absorbance interferograms were obtained over 400- −1. The −1 and averaged over 32 scans at a resolution of 4 cm 4000 cm spectra of agarose and agarose-ghee composites were finalised by subtracting the spectrum of water in triplicate measurements.
was developed by Mhaske et al. (2019). Results obtained were in ex- cellent agreement with estimates from traditional theoretical modeling, supporting the efficacy and accuracy of this work. CLSM is a quick, direct method that enables the user to study phase behaviour while keeping the composite microstructure intact by making tedious sample preparation like sectioning, coating and freeze drying redundant. This, coupled with the accuracy of image analysis software in quantifying phase volume shows potential for application over a wide range of materials saving time and effort. The primary objective of this paper, therefore, is to test the efficacy of the protocol in quantifying the phase volume of a more complex, hard lipid-biopolymer composite.
2.3.3. Differential scanning calorimetry (DSC)
Setaram Micro DSC VII (Setu-rau, Caluire, France) was used to carry out thermal analysis of the agarose-ghee mixtures. Samples (700 mg) were weighed into cylindrical vessels and sealed. A vessel containing equal weight of milliQ water (Millipore Corporation, Burlington, United States) served as a reference. Samples were equilibrated at 40 °C for 30 min before heating to 70 °C, cooling to 5 °C and heating to 70 °C again. Triplicate runs of averaged exotherms obtained at a scan rate of 1 °C/min are reported in this work.
The research also deals with the application of blending laws in quantifying the phase volumes in a hard lipid-biopolymer mixture in contrast to their widespread application to aqueous gels. Since lipids are an integral part of most food products (e.g. dressings, spreads), this finding holds potential in broadening the spectrum of scrutinizable products using blending laws. Attempts will be made to quantify the phase volume of a model system comprising of known quantities of hard lipid (i.e. ghee) in an agarose-ghee mixture using the blending laws and theoretical estimates will be compared with phase volumes obtained by the microscopic protocol. Ghee (anhydrous milk fat made by clarifying butter) remains in the solid form at ambient temperature (and within the temperature range of the present experimentation) and is expected to phase separate due to thermodynamic incompatibility with the aqueous network of agarose (Mortensen, 2016).
2. Materials and methods
2.1. Materials
Ghee by itself was analysed using a modulated DSC-Q2000, by TA Instruments (New Castle, DE, USA), fitted with a cooling unit, capable of reaching temperatures of about −50 °C. The accumulation of con- densate was prevented by purging nitrogen at a flow rate of 20 ml/min. About 10 mg of ghee was weighed and hermatically sealed in an alu- minium pan with an empty pan used as a reference. The sample was held at 40 °C for 30 min, heated to 70 °C and cooled down to −35 °C before heating it to 70 °C at a scan rate of 1 °C/min. A modulated amplitude of ± 0.53 °C for 40 s was used throughout. Averaged results of triplicates yielding effectively overlapping exotherms are reported in this study.
2.3.4. Rheological measurements
Agarose: Agarose (Type-1) was obtained from Sigma-Aldrich (Sydney, Australia). According to the supplier, the moisture, ash and sulphate content was less than 10, 0.25 and 0.15 g/100 g, respectively. Ghee: Pure cow ghee was purchased from Amul (Gujrat, India), the
density of which was measured to be 0.911 g/ml.
2.2. Methods
Density measurement: The density of ghee was volumetrically mea- sured using a 10 ml measuring cylinder. The cylinder was filled up to the 10 ml mark with ghee to measure its weight, which was divided by the weight of equal volume of water.
Discovery Hybrid Rheometer by TA Instruments (New Castle, DE, USA) was used to carry out small-deformation dynamic oscillation in- shear tests to measure the storage modulus of single and binary systems of agarose and ghee. A 20 mm geometry was used for ghee by itself while a 40 mm geometry was used for agarose and the composites with a 1 mm gap. Samples were loaded on a Peltier plate pre-heated at 45 °C and covered with polydimethylsiloxane to minimize evaporation. These were then cooled to 5 °C (at a rate of 2 °C/min) with a controlled strain of 0.01% and a constant oscillatory frequency of 1 rad/s prior to a 30 min time sweep. At the conclusion of the time sweep, a frequency sweep was performed in the range of 0.01–100 rad/s, followed by a heating ramp to 80 °C. Tests were carried out in triplicate for each sample.
2.3.5. Confocal laser scanning microscopy (CLSM)
Sample preparation: Binary mixtures were prepared by maintaining agarose concentration at 1 g/100 g and varying concentrations of ghee (0, 1, 3, 6, 9, 12 and 15 g/100 g) while adjusting the amount of water added to make the total volume of sample, 100 ml. In doing so, 1 g agarose was dissolved in calculated amount of milliQ water at 75 °C, stirred for 10 min and cooled to 45 °C prior to the addition of ghee. This mixture was then homogenized at 5000 rpm for 30 s using an IKA homogeniser-T25 Turrax (Wilmington, North Carolina, USA).
2.3. Experimental analysis
2.3.1. Scanning electron microscopy (SEM)
Ghee was stained with a lipophilic stain, Nile red (0.005 g/L) and then used to prepare agarose-ghee composites containing 1 g/100 g agarose and a varying concentration of ghee as outlined in Section 2.2. The homogenized sample solutions were pipetted into the well of a CoverWell imaging chamber, having a diameter of 20 mm and a depth of 2.8 mm (Sigma-Aldrich, NSW, Australia). Pre-cooling the chamber to 4 °C was essential for rapid in-situ gelation where the ghee is kinetically trapped within the agarose network and not bulk phase separated.
Confocal images were captured using a Nikon ECLIPSE Ti-E inverted microscope (Minato, Tokyo, Japan) fitted with an A1R laser. A 488 nm laser was used to excite Nile red and the emissions were captured be- tween 500 and 530 nm. A stack of 2D images was captured using a 10x objective. Fifteen consecutive images along the z-axis with a z-step of 5.72 μm were captured, imaging a depth of about 86 μm from the surface. Nine images were captured across each sample to avoid the effect of any inhomogeneous distribution of the ghee droplets.
2.3.6. Image analysis
The Z-stack obtained from CLSM imaging was quantified using two
Agarose gel cubes (10 mm3) with varying concentrations of ghee were freeze dried for 72 h at −40 °C in a Labconco FriZone Triade freeze dryer (Kansas City, Missouri, USA). They were then glued on to 12.6 mm carbon coated aluminium pins prior to being subjected to gold coating. An overview of the surface microstructure was obtained using FEI Quanta 200 ESEM (Oregon, USA) in low vacuum mode at 50x magnification. The working distance was 12 mm with an accelerating voltage of 25 kV and a spot size of 5. A 400x magnification was used to zoom into the pores of the agarose network to capture details of the agarose-ghee topology using a spot size of 5 and an electron beam of 25 kV.
135
2
P. Mhaske, et al.
LWT - Food Science and Technology 129 (2020) 109567
3.1.2. Chemical properties
image analysis software – FIJI with Java 6 (Madison, Wisconsin, USA) and Imaris 9.1.0 software (Bitplane, Belfast, Northern Ireland) in par- allel to estimate the lipid phase volume using the protocol outlined by Mhaske et al. (2019) as follows:
bridges),
(sulphate
FIJI: The Z-stack was thresholded using the “Default” algorithm, after which, the total area of object pixels was calculated irrespective of the size and circularity of the droplet. The generated area was finally multiplied by the z-step to determine the volume of the fat droplets present in that Z-stack. Averaged results of the nine images captured across each sample are reported.
1246 −1 (C–H vibration) and 2980 cm
−1 (C–C bending), 2900 cm
The molecular nature of agarose, ghee and their composites was examined using Fourier transform infrared (FTIR) spectroscopy. This was done to identify the presence of potential molecular interactions between the two components using as reference their molecular fin- gerprints. The interferograms obtained are illustrated in Fig. 2. Spectral −1 (anomeric scans of agarose exhibited absorption bands at 877 cm −1 (C–O–C vibration of carbon's angular C–H deformation), 1068 cm −1 3,6-anhydro-galactose cm esters), −1 1396 cm (O–H stretching). These bands are characteristic of the polysaccharide as reported earlier (Awadhiya, Tyeb, Rathore, & Verma, 2017; Wang, Guo, & Yang, 2011; Zamora-Mora, Velasco, Hernández, Mijangos, & Kumacheva, 2014).
Imaris: The 3D image was opened in the “Surfaces” feature and a Gaussian filter was applied to smoothen it, prior to subtracting the background intensity to reduce noise. The volume of ghee is measured in voxels by setting the diameter of the largest droplet as the limiting factor. This was repeated for the nine images and the averaged results are reported.
−1 (C–H rocking of CH2 and cis double bond), 1155 cm
2.3.7. Statistical analysis
Volume measurement of the lipid phase was done in triplicate and the data is expressed in mean ± standard deviation. Minitab 18 (Minitab Inc., Pennsylvania, USA) was used to carry out one-way ana- lysis of variance (ANOVA). Significance differences were defined as p < 0.05 with the Tukey test.
3. Results and discussion
The interferogram of ghee produced typical peaks, which reflect distinct characteristics of edible oils (Mansor, Man, & Rohman, 2011; Upadhyay, Jaiswal, & Jha, 2016). Absorption bands were observed at −1 721 cm −1 (bending of CH3 (stretching vibration of C–O in esters), 1452 cm −1 (C]O stretching of aliphatic esters) and at aliphatic group), 1746 cm −1 (symmetrical and asymmetrical C–H stretching of 2856 and 2920 cm CH2 groups in the organic fatty acids) (Antony, Sharma, Mehta, Ratnam, & Aparnathi, 2017; Upadhyay, Jaiswal, & Jha, 2018). The spectral scans of the blends replicate the characteristics of the two constituents with the spectra of the polymer dominating. Since its peaks deviated neither in band wavenumber nor intensity in the composites, and no new peaks were formed, it can be inferred that there is no molecular interaction between ghee and agarose.
3.1. Investigation of the composite microstructure and its structural characteristics
3.1.3. Thermal characterisation
3.1.1. Composite microstructure
In order to be able to apply the blending law analysis, it is vital to have a phase separated system with distinct, non-interacting compo- nent phases. To attest that the agarose-ghee composites meet this morphological requirement, and avoid bulk phase separation, freeze dried samples were analysed using SEM. The efficacy of SEM in probing into the complex microstructures of composite gels has been well documented in the past (Pang, Deeth, Sharma, & Bansal, 2015).
Complementary characterisation to the preceding sections for the phase behaviour of our composite is afforded by calorimetric tests (Du, Brenner, Xie, & Matsukawa, 2016). The typical exotherms of agarose and ghee during controlled cooling are shown in Fig. 3a. Changes ac- companying the exothermic profile of agarose exhibited a mid-point transition temperature of about 26 °C, an outcome of the coil-to-helix transition of the polysaccharide leading to gelation at 1 g/100 g solids. The crystallization and lattice formation of the fatty acids, having a medium melting fraction (MMF; mainly stearin), lead to the formation of one minor peak at 15 °C, with the low melting fraction (LMF; mainly olein) resulting in a major peak at 9.5 °C. The bimodal behaviour re- flected the thermodynamically distinct molecular domains present in ghee (Dimick, Reddy, & Ziegler, 1996; Verret et al., 2009).
An overview of the freeze-dried agarose gel (1 g/100 g) captured using 50x magnification is depicted in Fig. 1a. The image reveals a highly porous structure with the polysaccharide network forming a honeycomb pattern. Addition of ghee did not have any perceivable effect on the continuous network of agarose, as all samples displayed a similar morphology (results not shown). Since the fat droplets weren't visible at this magnification, a higher exposure (400x) was used to zoom into the walls of the agarose network. This indicates that the size of the ghee inclusions is much smaller than the pores of the agarose matrix.
At 400x, in the absence of ghee, the walls of agarose network ex- hibited a smooth, uninterrupted surface as shown in Fig. 1b. This was contrasted in the micrographs of the binary mixture seen in Fig. 1c and d, which show the spherical inclusions of ghee embedded within the continuous polysaccharide mesh. The micrographs reveal that when the concentration of ghee was increased from 3 to 15 g/100 g, the number of lipid inclusions increased, while their size remained fairly constant (60 ± 15 μm).
Ghee is a complex mixture of triacylglycerols (TAGs), which con- stitutes 98% of its composition and comprise at least 60 different fatty acids, with their characteristic chemical and thermal properties (Christelle Lopez, Bourgaux, Lesieur, Riaublanc, & Ollivon, 2006). The onset of crystallization and peak intensity depend largely on the rate of cooling (C. Lopez, Lesieur, Bourgaux, & Ollivon, 2005). Based on the above, Fig. 3b examines the cooling exotherms of the blends of 1 g/ 100 g agarose with increasing concentrations of ghee. Dominant peaks denoting the crystallization of fat are visible between −10 and 15 °C, with the peaks containing higher concentrations of ghee depicting two independent but partially overlapping phase transitions of the con- stituent melting fractions. The absence of the native agarose peaks in the blend can be attributed to the much larger intensity of the peaks of ghee masking the low intensity coil-to-helix events of the former.
Single ghee preparations or in a composite with agarose generated comparable thermal events, varying only in intensity as a function of concentration. This argues that the crystallization of fat is unaffected by the presence of the polysaccharide in their blends reinforcing the con- cept of phase separation in the mixture. Furthermore, agarose gels at temperatures higher than ghee crystallization and the isothermal test used in the following rheological experiments (5 °C for 30 min), hence the composite morphology should be that of a continuous biopolymer matrix supporting discontinuous fat crystals.
From the micrographs, it is clear that the blending of agarose and ghee lead to the formation of isotropic structures where the dis- continuous inclusions of ghee was supported by the continuous poly- meric matrix. The spheroidal lipid inclusions will serve as the basis for modeling the mechanical properties of the binary composite. Similar results were reported by Mhaske et al. (2019), who studied the mi- crostructure of agarose-canola oil composite and unveiled a continuous polysaccharide network supporting the oil droplets acting as a dis- continuous “filler”.
136
3
P. Mhaske, et al.
LWT - Food Science and Technology 129 (2020) 109567
Fig. 1. SEM micrographs at 50x magnification of (a) 1% (w/w) agarose, and at 400x magnification of 1 g/100 g agarose containing (b) 0 g/100 g ghee, (c) 3 g/100 g ghee and (d) 15 g/100 g ghee.
Fig. 2. FTIR spectra of ghee (dotted line), and 1 g/100 g agarose with 0, 1, 3, 6, 9, 12 and 15 g/100 g ghee arranged from top to bottom (solid lines).
3.1.4. Rheological observations of structure formation
Finally, oscillatory tests were performed to investigate the me- chanical behaviour of the agarose network and crystalline ghee as a prerequisite to advancing a study on the effect of increasing the con- centration of the latter in the composite. The profile of storage modulus
(Gʹ) for 1 g/100 g agarose during controlled cooling at 2 °C/min is depicted in Fig. 4. On hydration and subsequent solubilisation, the disordered coils of agarose undergo a conformational transition at about 30 °C to form a strong gel at the end of the cooling routine (103.79 Pa at 5 °C). On the other hand, formation of the crystalline ghee
137
4
P. Mhaske, et al.
LWT - Food Science and Technology 129 (2020) 109567
Fig. 3. Exothermic peaks from the cooling profiles of (a) ghee (solid line) and agarose (dotted line) by themselves and (b) 1 g/100 g agarose with 1, 3, 6, 9, 12 and 15 g/100 g ghee arranged from bottom to top.
network lead to a sharp increase in the values of Gʹ at around 22 °C and a reinforced structure of 105.66 Pa at 5 °C.
aqueous phase and the hardening of ghee. Rheological observations suggest that ghee acts as the strong discontinuous filler (its contribution is noticeable in the mixture) supported by a weaker agarose phase hence providing the basis for theoretical modeling using blending law.
3.2. Theoretical modeling of mechanical functions in support of phase topology
Turning our attention to the composites, these were made by adding 1 g of agarose to, for example, 9 g of ghee in a beaker. Subsequent amount of water was added to make up the total sample volume, re- sulting in an increase in the effective concentration of agarose in the diminished aqueous phase as the concentration of fat increased in the composite (from 1 to 15 g/100 g). Addition of ghee to the agarose gels produced a systematic variation in viscoelasticity of the cooling pro- files, as depicted in Fig. 4. Blends showed a similar trend of gelation as that of the single agarose system up till 15 °C, with higher ghee con- centrations increasing the composite Gʹ value, as a result of volume exclusion between the aqueous and lipid phase, leading to effective polysaccharide concentrations higher than the initial 1 g/100 g.
As shown in Fig. 4, a second wave of structure formation was wit- nessed at about 15 °C that reflected the onset of fat crystallization. Storage modulus values correspond to the concentration of added ghee reaching 103.84 and 104.32 Pa for 1 and 15 g/100 g ghee, respectively. Thus, the overall composite strength at the end of the cooling run (5 °C) is the outcome of higher effective concentrations of the polymer in the
Having qualitatively established in the preceding sections that agarose and ghee remain in mixture in non-interacting phase-separated domains, we turned to an in-depth quantitative analysis using rheolo- gical blending law. In doing so, a double-logarithmic calibration curve was obtained (inset of Fig. 4 with a correlation coefficient, R2 of 0.995) to provide information on the concentration dependence of storage modulus for agarose ranging from 1.0 to 2.5 g/100 g. This was indis- pensable for the interpolation between measured values of Gʹ at dif- ferent concentrations during theoretical modeling. The topology of a soft continuous polymer phase supporting a hard discontinuous filler can follow progress in viscoelasticity of our results using the mathe- matical expression of Lewis and Nielsen (Shrinivas, Kasapis, &
138
5
P. Mhaske, et al.
LWT - Food Science and Technology 129 (2020) 109567
Fig. 4. Cooling profile of storage modulus for 1% (w/w) agarose with 0 (□), 1 (җ), 3 (−), 6 (x), 9 (●), 12 (Δ) and 15 g/100 g (♦) (w/w) ghee and ghee by itself (■) during cooling from 50 to 5 °C at a scan rate of 2 °C/min. The inset denotes the agarose calibration curve of G′ at 5 °C as a function of its concentration.
Tongdang, 2009):
1
+
ABΦ y
=
1
−
BψΦ y
(1)
′ G ′ G x
1
−
B
=
fraction in the polymer-X phase and the total weights of the X and Y phases, respectively. Further, the relative weight of the two phases needed to be adjusted for density differences. This was carried out by dividing Wx and Wy by their respective densities (water = 1 g/ml; ghee = 0.911 g/ml) to obtain the values of Vx and Vy, i.e. the volumes of phase X and Y, respectively. Finally, the relative phase volume in the composite is given for Φx, for example, as follows:
A
+
(2)
′ G y ′ G x ′ G y ′ G x
=
Φ x
V x +
V x
V y
(6)
1
m
⎜
⎟
ψ
Φ y
Φ − 2 Φ m
(3)
⎞ ⎠
= + ⎛ 1 ⎝
1,
=Φ y
+Φx
where, Gʹ, Gʹx, and Gʹy are the storage modulus of the composite, soft continuous polymer phase and rigid dispersed phase, respectively and ɸy is the relative phase volume of the filler. The approach takes into consideration the shape of the filler via the Einstein coefficient (A), which is 1.5 for dispersed spheroids in an elastomeric matrix, and the maximum packing fraction (ɸm) of the filler, which is 0.48 for loose packing of spheroidal aggregates reflecting the random arrangement of extremely hard ghee particles within the agarose matrix.
and these relative phase volumes are used in where, equations (1) and (3). Accordingly, the phase volumes of ghee were calculated for Sag = 1 and equations (4)–(6), which are also reproduced in the inset of Fig. 5 against the experimental lipid concentrations. Estimated phase volumes of ghee correlate well with the experimental concentrations in the composite, upholding that the phase volume de- pended entirely on its concentration. Phase volume predictions by theoretical modeling will be compared with those of 3D imaging cou- pled with image analysis to assess its efficacy in estimating phase vo- lumes in demixed systems.
3.3. Z-stack CLSM imaging and image analysis of the agarose-ghee blends
Fig. 5 illustrates a typical outcome of calculations using equations (1)–(3) for 1 g/100 g agarose with 6 g/100 g ghee. A detailed section of the computerized output reproduced the storage modulus of the con- tinuous agarose phase (Gʹag) for all possible water distributions from 0.5 to 1.0, in steps of 0.01, in the agarose phase (Sag). The calculated hy- pothetical values of the composite (GʹC(LN)) and the experimental composite values (Gʹexp) as a function of Sag are also depicted. Calcu- lated and experimental values intersected at Sag of 1.0, positing that the total amount of water remained in the polymer phase. Modeling of the entire working range is depicted in the inset of Fig. 5 for Sag = 1, re- vealing a close agreement between the experimental and modelled composite moduli as a function of lipid concentration.
The above understanding has now been utilised to estimate the phase volume of the individual phases. This can be obtained by the generic equations developed by Kasapis, Morris, Norton, and Clark (1993):
w
x = +
S (
×
)
(4)
W x
x
S
w
= +
((1
−
)
×
)
W y y
x
(5)
3D images of all the agarose-ghee blends were captured and pro- cessed using image analysis software, FIJI and Imaris, following the protocol developed by Mhaske et al. (2019). In doing so, fluorescent probes were employed to selectively stain the lipid. A single, 2D image captured from a section of a sample containing 1 g/100 g agarose and 6 g/100 g ghee in the xy plane is represented in Fig. 6a, as an example. The stained ghee droplets appeared as green spheres whilst the un- stained agarose formed the black background. The bright green spheres correspond to the droplets in the focal plane, which are contrasted to those outside the focal plane showing a reduced colour intensity. Each 2D depiction is an 8 bit image with 512 × 512 pixel in size (1 pixel = 2.49 μm on the X and Y-axis). Keeping a fixed z-step of 5.72 μm, 15 consecutive 2D images along the Z-axis were captured as shown in Fig. 6b, collecting data up to a depth of 85 μm from the surface. This stack of 2D images was processed with FIJI, an open source analysis software for phase volume estimation, by considering separately each 2D plane. On the other hand, Imaris rendered a 3D image using the Z- stack (Fig. 6c) and performed volumetric measurements.
where, w, x, y, Sx, Wx and Wy denote the total weight of water in the composite, weight of polymer X, weight of polymer Y, the solvent
139
6
P. Mhaske, et al.
LWT - Food Science and Technology 129 (2020) 109567
Fig. 5. Computerized modeling for 1 g/100 g agarose plus 6 g/100 g ghee, using the Lewis and Nielsen blending law, is illustrated depicting the predicted storage modulus of agarose (G'ag; dashed line), composite (G'C(LN); solid line) and the experi- mental composite modulus (G'exp; dotted line) as a function of the solvent content of the agarose phase (Sag). The experimental Gʹ values (●) and phase vo- lume estimates of ghee (♦) for 1 g/100 g agarose with increasing ghee concentrations are depicted in the inset (blending law is used to derive the solid line fit of composite modulus and the ghee phase volume in the inset).
Each pixel in the 2D images denotes a specific intensity value ran- ging from 0 to 255 (Xu, Li, Li, & Hua, 2016). A sample containing no florescent probe produces a completely black image with all the pixels having an intensity value of 0. This is exploited as the negative control that defines the baseline calibration in image analysis. While analysing the images, thresholding was the most crucial step to accurately de- lineate the boundary of fat droplets so that the object (fat) pixels are distinguished from the background (agarose) pixels. The phase volume of ghee estimated using FIJI is reported in Table 1 along with the weight-per-weight experimental concentration of the lipid added to the mixture and its conversion to the corresponding weight-per-volume concentration considering that its density is below 1.0. Thus, the phase volumes obtained using FIJI are very close to the experimental values (entries in the first two columns), and the theoretical estimates in the inset of Fig. 5.
Turning our attention to the second image analysis software – Imaris, we once again followed the procedure outlined by Mhaske et al. (2019), with the phase volume estimates being also reported in Table 1. It is evident that the Imaris estimates are in good agreement with those obtained from FIJI, the experimental values, and the theoretical pre- dictions from rheological blending-law analysis seen in the inset of Fig. 5. Statistically, there is no significant difference (p ˃ 0.05) between the measurements of phase volume obtained using FIJI and Imaris, indicating that both types of software systems are capable of generating reliable results.
A closer look at Table 1, revealed a discernible trend of decreasing proximity between the weight-per-volume ghee concentrations and the phase volumes estimated using the microscopic protocol. To scrutinize this further, a binary gel comprising 1 g/100 g agarose and 20 g/100 g ghee was prepared and analysed. Taking into account density con- siderations, this created a weight-per-volume ghee concentration of 21.55%, with FIJI and Imaris returning values of 20.12 ± 0.05 and 20.09 ± 0.11% (w/v), respectively. Such outcomes are rationalised in Fig. 7a that depicts the first optical section captured along the Z-axis and Fig. 7b with the 15th optical section at a depth of approximately 85 μm from the surface. As before, green spheres denote the stained ghee droplets whereas the agarose is present as the black background. The clear spherical inclusions of ghee in Fig. 7a are contrasted with those in Fig. 7b, which appear to be eclipsed and akin to crescents.
Fig. 6. (a) 2D, (b) Z-stack, and (c) 3D CLSM images of 1 g/100 g agarose plus 6 g/100 g ghee. The green droplets depict the ghee whereas the black back- ground corresponds to agarose. Scale bar denotes 100 μm. (For interpretation of the references to colour in this figure legend, the reader is referred to the Web version of this article.)
This outcome is attributed to the limitation of CLSM in imaging deep sections, with the illuminating beam undergoing partial reflection
140
7
P. Mhaske, et al.
LWT - Food Science and Technology 129 (2020) 109567
Table 1 Experimental ghee concentrations in the composite and estimated phase volumes with FIJI and Imaris image analysis.
Weight per weight (g/100 g) ghee concentrations in the composite
Weight per volume (g/100 cm3) ghee concentrations in the composite
Lipid phase volumes estimated using FIJI (g/100 cm3)
Lipid phase volumes estimated using Imaris (g/100 cm3)
1 3 6 9 12 15
1.10 3.30 6.59 9.89 13.10 16.40
1.14 ± 0.02 3.42 ± 0.16 6.48 ± 0.11 9.68 ± 0.18 12.78 ± 0.13 15.60 ± 0.22
1.21 ± 0.07 3.41 ± 0.09 6.59 ± 0.14 9.54 ± 0.04 12.63 ± 0.21 15.79 ± 0.19
CRediT authorship contribution statement
Pranita Mhaske: Methodology, Formal analysis, Writing - original draft. Lloyd Condict: Writing - review & editing. Mina Dokouhaki: Writing - review & editing. Asgar Farahnaky: Writing - review & editing. Stefan Kasapis: Funding acquisition, Conceptualization, Supervision, Writing - review & editing.
Acknowledgement
The authors acknowledge the facilities, and the scientific and the RMIT University's Microscopy and
technical assistance of Microanalysis Facility.
References
Fig. 7. Confocal images of 1 g/100 g agarose and 20 g/100 g ghee captured from (a) the surface and (b) a depth of 80 μm from the surface. Scale bar de- notes 100 μm.
Antony, B., Sharma, S., Mehta, B. M., Ratnam, K., & Aparnathi, K. (2017). Study on FT- MIR spectra of ghee (anhydrous milk fat). British Food Journal, 119(1), 181–189. Awadhiya, A., Tyeb, S., Rathore, K., & Verma, V. (2017). Agarose bioplastic‐based drug delivery system for surgical and wound dressings. Engineering in Life Sciences, 17(2), 204–214.
Blonk, J. C. G., van Eendenburg, J., Koning, M. M. G., Weisenborn, P. C. M., & Winkel, C. (1995). A new CSLM-based method for determination of the phase behaviour of aqueous mixtures of biopolymers. Carbohydrate Polymers, 28(4), 287–295. Çakır, E., & Foegeding, E. A. (2011). Combining protein micro-phase separation and
protein–polysaccharide segregative phase separation to produce gel structures. Food Hydrocolloids, 25(6), 1538–1546.
Dekkers, B. L., Emin, M. A., Boom, R. M., & van der Goot, A. J. (2018). The phase
properties of soy protein and wheat gluten in a blend for fibrous structure formation. Food Hydrocolloids, 79, 273–281.
Dimick, P., Reddy, S. Y., & Ziegler, G. R. (1996). Chemical and thermal characteristics of milk‐fat fractions isolated by a melt crystallization. Journal of the American Oil Chemists Society, 73(12), 1647–1652.
Du, L., Brenner, T., Xie, J., & Matsukawa, S. (2016). A study on phase separation behavior
and scattering that reduces the photon intensity striking the specimen being sampled (“inner filter” effect). Loss of excitation reduces the fluorescence yield from the focal plane, which is further attenuated on its return to the detector by the same factors (Pioch, Sorg, Stadler, Hagge, & Dörfer, 2004; Laurent et al., 1994). In lipid-biopolymer composites, component phases have significantly different refractive indices, and when the laser beam passes from a phase with a specific refractive index to another some of it is scattered back towards the source. Reduction of the laser intensity reaching deeper sections causes a “shadowing” effect observed in Fig. 7b. It can be minimized by the use of an objective with a higher numerical aperture, increasing illumina- tion intensity, using the microscope in the transmission mode, etc. but cannot be completely overcome within the bounds of current tech- nology (Sandison, Williams, Wells, Strickler, & Webb, 1995, pp. 39–53).
in kappa/iota carrageenan mixtures by micro DSC, rheological measurements and simulating water and cations migration between phases. Food Hydrocolloids, 55, 81–88.
Gaaloul, S., Turgeon, S. L., & Corredig, M. (2010). Phase behavior of whey protein ag- gregates/κ-carrageenan mixtures: Experiment and theory. Food Biophysics, 5(2), 103–113.
4. Conclusions
Hou, J.-J., Guo, J., Wang, J.-M., He, X.-T., Yuan, Y., Yin, S.-W., et al. (2015). Edible
double-network gels based on soy protein and sugar beet pectin with hierarchical microstructure. Food Hydrocolloids, 50, 94–101.
Kasapis, S. (2008). Phase separation in biopolymer gels: A low-to high-solid exploration of structural morphology and functionality. Critical Reviews in Food Science and Nutrition, 48(4), 341–359.
Kasapis, S., Morris, E. R., Norton, I. T., & Clark, A. H. (1993). Phase equilibria and ge- lation in gelatin/maltodextrin systems — Part IV: Composition-dependence of mixed- gel moduli. Carbohydrate polymers. 21(4), 269–276.
Kasapis, S., & Tay, S. L. (2009). Morphology of molecular soy protein fractions in binary
composite gels. Langmuir, 25(15), 8538–8547.
Laurent, M., Johannin, G., Gilbert, N., Lucas, L., Cassio, D., Petit, P. X., et al. (1994). Power and limits of laser scanning confocal microscopy. Biology of the Cell, 80(2), 229–240.
Lopez, C., Bourgaux, C., Lesieur, P., Riaublanc, A., & Ollivon, M. (2006). Milk fat and
primary fractions obtained by dry fractionation: 1. Chemical composition and crys- tallisation properties. Chemistry and Physics of Lipids, 144(1), 17–33.
Lopez, C., Lesieur, P., Bourgaux, C., & Ollivon, M. (2005). Thermal and structural be- havior of anhydrous milk fat. 3. Influence of cooling rate. Journal of Dairy Science, 88(2), 511–526.
Lorén, N., Langton, M., & Hermansson, A. M. (1999). Confocal laser scanning microscopy and image analysis of kinetically trapped phase-separated gelatin/maltodextrin gels. Food Hydrocolloids, 13(2), 185–198.
Rheology-based blending laws have been widely applied over the years to estimate the phase volumes and water partitioning between segregative biopolymers in the composite gel. These, however, are te- dious, demanding a substantial amount of experimental work and specialised expertise for successful modeling of phase behaviour. Advances in imaging technology and high precision image analysis software have made it possible to probe specimens non-invasively and capture in-situ 3D images that can be further quantified to estimate the phase volume in binary composites. Our microscopic protocol was tested on a model system of agarose-ghee that showed distinct phase separation in order to estimate the phase volume of the discontinuous filler. Estimates compared well with those obtained from the rheolo- gical modeling, making the microscopic protocol a potential substitute for accurately determining phase behaviour in industrially relevant composite gels. Perceivable effects of self-shadowing during imaging merit further research in order to achieve wide-spread applicability for this protocol.
Mansor, T. S. T., Man, Y. B. C., & Rohman, A. (2011). Application of fast gas chroma- tography and Fourier transform infrared spectroscopy for analysis of lard adultera- tion in virgin coconut oil. Food Analytical Methods, 4(3), 365–372.
141
8
P. Mhaske, et al.
LWT - Food Science and Technology 129 (2020) 109567
Mhaske, P., Condict, L., Dokouhaki, M., Katopo, L., & Kasapis, S. (2019). Quantitative
of bicontinuous gels of agarose and gelatin and the effect of added lipid phase. Langmuir, 25(15), 8763–8773.
analysis of the phase volume of agarose-canola oil gels in comparison to blending law predictions using 3D imaging based on confocal laser scanning microscopy. Food Research International, 125, 108529.
Mortensen, B. K. (2016). Anhydrous milk fat/butter oil and ghee reference Module in food
science. Elsevier.
Takayanagi, M. (1963). Viscoelastic behavior of polymer blends and its comparison with model experiments. Journal of the Society of Materials Science Japan, 12, 389–394. Upadhyay, N., Jaiswal, P., & Jha, S. N. (2016). Detection of goat body fat adulteration in pure ghee using ATR-FTIR spectroscopy coupled with chemometric strategy. Journal of Food Science and Technology, 53(10), 3752–3760.
Nielsen, L. E. (1974). Morphology and the elastic modulus of block polymers and poly-
blends. Rheologica Acta, 13(1), 86–92.
Pang, Z., Deeth, H., Sharma, R., & Bansal, N. (2015). Effect of addition of gelatin on the
Upadhyay, N., Jaiswal, P., & Jha, S. N. (2018). Application of attenuated total reflectance Fourier Transform Infrared spectroscopy (ATR–FTIR) in MIR range coupled with chemometrics for detection of pig body fat in pure ghee (heat clarified milk fat). Journal of Molecular Structure, 1153, 275–281.
rheological and microstructural properties of acid milk protein gels. Food Hydrocolloids, 43, 340–351.
Verret, C., El Moueffak, A., Largeteau, A., Frimigacci, M., Demazeau, G., Leal-Calderon, F., et al. (2009). Effects of high pressure on anhydrous milk fat crystallization in emulsion. High Pressure Research, 29(1), 57–60.
Wang, W., Guo, X., & Yang, Y. (2011). Lithium iodide effect on the electrochemical be-
Pioch, T., Sorg, T., Stadler, R., Hagge, M., & Dörfer, C. E. (2004). Resin penetration through submicrometer hiatus structures: A SEM and CLSM study. Journal of Biomedical Materials Research Part B: Applied Biomaterials: An Official Journal of The Society for Biomaterials, The Japanese Society for Biomaterials, and The Australian Society for Biomaterials and the Korean Society for Biomaterials, 71(2), 238–243. Sandison, D. R., Williams, R. M., Wells, K. S., Strickler, J., & Webb, W. W. (1995).
havior of agarose based polymer electrolyte for dye-sensitized solar cell. Electrochimica Acta, 56(21), 7347–7351.
Xu, L., Li, Z., Li, J., & Hua, W. (2016). A novel bit-level image encryption algorithm based
Quantitative fluorescence confocal laser scanning microscopy (CLSM) Handbook of bio- logical confocal microscopy. Springer.
on chaotic maps. Optics and Lasers in Engineering, 78, 17–25.
Semasaka, C., Buckow, R., & Kasapis, S. (2018). Modeling counterion partition in com- posite gels of BSA with gelatin following thermal treatment. Food Hydrocolloids, 74, 97–103.
Zamora-Mora, V., Velasco, D., Hernández, R., Mijangos, C., & Kumacheva, E. (2014). Chitosan/agarose hydrogels: Cooperative properties and microfluidic preparation. Carbohydrate Polymers, 111, 348–355.
Shrinivas, P., Kasapis, S., & Tongdang, T. (2009). Morphology and mechanical properties
142
9
Chapter 5
3D Confocal laser scanning microscopy for quantification
of the phase behaviour in agarose-MCC co-gels in
comparison to the rheological blending-law analysis
Mhaske, P., Farahnaky, A., & Kasapis, S. (2020). 3D Confocal Laser Scanning Microscopy
for Quantification of the Phase Behaviour in Agarose-MCC co-gels in Comparison to the
143
Rheological Blending-law Analysis. Food Biophysics, 1-8.
Food Biophysics https://doi.org/10.1007/s11483-020-09656-6
ORIGINAL ARTICLE
3D Confocal Laser Scanning Microscopy for Quantification of the Phase Behaviour in Agarose-MCC co-gels in Comparison to the Rheological Blending-law Analysis
Pranita Mhaske 1 & Asgar Farahnaky 1 & Stefan Kasapis 1
Received: 3 September 2020 / Accepted: 5 October 2020 # Springer Science+Business Media, LLC, part of Springer Nature 2020
Abstract The need for a rapid and direct alternative to the rheology-based blending laws in quantifying phase behaviour in biopolymer composite gels is explored in this study. In doing so, the efficacy of confocal laser scanning microscopy (CLSM) paired with image analysis software – FIJI and Imaris - in quantifying phase volume was studied. That was carried out in a model system of agarose with varying concentrations of microcrystalline cellulose (MCC) in comparison to the rheological blending laws. Structural studies performed using SEM, FTIR, differential scanning calorimetry and dynamic oscillation in-shear unveiled a continuous, weak agarose network supporting the hard, rod-shaped MCC inclusions where the composite gel strength increased with higher ‘filler’ concentration. The phase volumes of MCC, estimated with the microscopic protocol, matched the predictions obtained from computerized modelling using the Lewis-Nielsen blending laws. Results highlight the suitability of the micro- scopic protocol in estimating the water partition and effective phase volumes in the agarose-MCC composite gel.
Keywords Phase behaviour . Blending law analysis . 3D imaging . Confocal laser scanning microscopy . Image analysis
Introduction
using the classical methods of centrifugal separation or deter- mination of osmotic pressure from composite solutions [3, 4]. Other attempts to probe phase behaviour of composite sys- tems employed turbidimetric [5], spectrometric [6] and rheo- logical protocols. The latter, in particular, paired with blend- ing law analysis estimated accurately the phase behaviour of synthetic and natural polymer mixtures [7, 8]. In the latter, blending laws correlate the overall mechanical properties of the composite to those of the individual constituents taking into consideration the solvent partition between the two poly- meric phases. Though accurate in estimating phase volumes, their application demands considerable experimentation and modelling expertise. Clearly, there is a need for a rapid and direct approach capable of predicting the phase behaviour in biopolymer mixtures.
The recent years have seen a drastic increase in the use of proteins and polysaccharides in food product formulations to improve/control processability, shelf life, texture/mouthfeel and the kinetics of bioactive compound release [1]. For the most part, such techno-functionality is governed by the sol- vent partitioning amongst the macromolecules in the phase- separated mixture [2]. Fundamental understanding of solvent distribution and molecular interactions between constituents depends on their concentration, water addition, ionic strength, pH and processing, for example, thermal treatment or applied shear. Despite the range and depth of research in the literature, a thorough understanding of phase behaviour and interactions in complex biomaterial systems remains of great interest. These systems generally comprise highly viscous and gelled phases, whose phase behaviour cannot be accurately predicted
* Stefan Kasapis
stefan.kasapis@rmit.edu.au
1
Over the years, confocal laser scanning microscopy (CLSM) has been used sporadically in this area and proved popular for qualitative analysis [9–11]. A few studies coupled CLSM imaging with image analysis but they were limited to the investigation of single 2D images [12] and viscous aque- ous solutions [13]. Exploiting advances in image processing technology, Mhaske et al. (2019) developed a technique of quantifying phase volume in a lipid-biopolymer system using 3D CLSM imaging paired with image analysis [14]. The
School of Science, RMIT University, Bundoora West Campus, Plenty Road, Melbourne, VIC 3083, Australia
144
Food Biophysics
protocol yielded values for the volume of the lipid phase that depend entirely on its concentration (w/w) added to the sys- tem. These were in close agreement with the estimates obtain- ed from the rheology based theoretical modelling, highlight- ing the potential of the combined approach.
days in a Labconco FriZOne Triade freeze drier (Kansas City, Missouri, USA). These were then placed on 12.6 mm carbon coated aluminium stubs and gold coated. FEI Quanta 200 ESEM (Hillsboro, Oregon, USA) was used to capture the surface micrographs of the freeze dried samples. Images were captured in the high vacuum mode, with a spot size of 5, working distance of 12 mm and an accelerating voltage of 30 kV. A 50x magnifica- tion was used to obtain an overview of the gel network, whilst details of the composite topology were captured at 600 x.
Fourier Transform Infrared Spectroscopy (FTIR) A Spectrum 2 FTIR spectrometer by Perkin Elmer (Norwalk, Connecticut, USA) with a diamond crystal attenuated total reflectance de- vice from GladiATR (Pike Technologies, Maddison, Wisconsin, USA) was used to analyse single and binary mix- tures of agarose and MCC. The absorbance spectra were ob- tained at a resolution of 4 cm− 1 between 400 and 4000 cm− 1 and averaged over 64 scans in triplicate measurements by subtracting the spectrum of water.
The present work aims to quantify the phase behaviour of a complex, industrially relevant hydrogel comprising a structur- ing polysaccharide (agarose) and a rigid dispersion in the form of microcrystalline cellulose (MCC). The latter is also capable of absorbing and retaining water, hence being used in the cosmetics, plastic, pharmaceutical and cement industries. Applications in processed food are based on its functionality as a dietary fibre and bulking agent to manipulate the nutri- tional profile and consistency of preparations [15]. Thus, MCC’s interactions with other biopolymers is a topic of cur- rent fundamental and applied interest. In the present system, the retention of water by agarose and MCC is unknown and yet it will go a long way in determining the structural proper- ties of this mixture. Phase volume estimations will be attempted with the combined protocol of microscopic exami- nation and blending law modelling in order to confirm the suitability of this approach in identifying the solvent partition between two hydrophilic polymeric phases.
Materials and Methods
Materials and Sample Preparation
Differential Scanning Calorimetry (DSC) Thermal analysis of agarose and agarose-MCC mixtures was carried out using the Setaram Micro DSC VII (Setu-rau, Caluire, France). About 700 mg sample was transferred into a cylindrical vessel and sealed, with an equal weight of milliQ water being added to the reference vessel. Both vessels were then deposited into the instrument chamber and held at 40 °C for 30 min before ramping the temperature to 70 °C, cooling to 5 °C and once again heating to 70 °C. Averaged exotherms from triplicate runs, obtained at a constant scan rate of 1 °C/min, are reported. MCC paste, obtained by hydrating MCC overnight in excess water and centrifuging the suspension, was also analysed using a modulated DSC Q2000 (TA Instruments, New Castle, DE, USA). The instrument was interfaced to a refrigeration unit to achieve temperatures well below 0 °C and a nitro- gen purge at a flow rate of 25 mL/min was used. About 12 mg of sample was weighed in an aluminium pan and hermetically sealed, with an empty pan being the reference. Samples were subjected to successive heating and cooling routines between 10 and 350 °C at a constant scan rate of 1 °C/min. Triplicate runs produced consistent results.
Agarose (Type-1) and microcrystalline cellulose (Avicel, PH- 101) were supplied by Sigma-Aldrich (Sydney, Australia). The sulphate, ash and moisture content of agarose was less than 0.15, 0.25 and 10%, respectively, according to the sup- plier. The particle size of cellulose crystals was ~ 50 µm, with a relative density of 0.6 g/cm3. 5-DTAF [5-(4, 6– Dichlorotriazinyl)aminofluorescein], a bright yellow coloured powder was purchased from AAT Bioquest (California, USA) and used to covalently label agarose. Varying quantities of MCC (0.1, 0.3, 0.6, 0.9, 1.2, 1.5 g) were added to specific amounts of milliQ water to make the total volume of the aque- ous solutions 99 ml. This was stood overnight at ambient temperature to ensure complete hydration of the cellulose crystals. MCC suspensions were then heated to 80 °C in a water bath with constant magnetic stirring before dispersing 1 g agarose powder. Mixtures (1% (w/w) agarose and 0.1, 0.3, 0.6, 0.9, 1.2 and 1.5% (w/w) MCC) were stirred continuously for 15 min to dissolve the polysaccharide and cooled to 50 °C for subsequent analysis.
Experimental
Rheological Measurements Small-deformation dynamic oscil- lation in-shear was carried out on a Discovery Hybrid Rheometer (TA Instruments, New Castle, DE, USA) to mea- sure the elastic modulus (Gʹ) of the hydrated MCC paste, aga- rose gel and agarose-MCC composites. Samples were loaded at 50 °C onto the preheated Peltier plate with a 40 mm parallel plate geometry and 1 mm gap. The exposed edges of the com- posite were coated with polydimethylsiloxane (PDMS) to
SEM Analysis Gel cubes of agarose (10 mm3) with varying concentrations of MCC were freeze dried at -40 °C for 3
145
Food Biophysics
Statistical Analysis Phase volume results obtained were eval- uated using Minitab 18 (Minitab Inc., Pennsylvania, USA) by one-way analysis of variance (ANOVA). Volume measure- ment is done in duplicate and expressed in mean ± standard deviation, with significant differences being p < 0.05 (Tukey test).
minimise moisture loss. Analysis involved a temperature sweep from 50 to 5 °C at 2 °C/min, a 30 min isothermal step at 5 °C with a frequency of 1 rad/s, followed by a frequency sweep from 0.1 to 100 rad/s prior to ramping the temperature to 70 °C with a controlled strain of 0.01% throughout, which is within the linear viscoelastic region of this matrix (LVR).
Results and Discussion
Structural Characterisation of the Agarose/MCC Mixtures
Confocal Laser Scanning Microscopy (CLSM) and Image Analysis Agarose was covalently labelled with 5-DTAF (5-(4, 6-Dichlorotriazinyl) Aminofluorescein) using the procedure outlined by Russ et al. (2013) [16]. In doing so, 1 g of agarose was dissolved in 50 ml milliQ water at 80 °C with constant stirring. The solution was cooled to 60 °C and 8 mg 5-DTAF was added before slowly adding 15 ml of 1% (w/v) Na2SO4. Then, 3–4 drops of 10% (w/v) NaOH were added and stirred for 2 hours prior to adding 120 ml absolute ethanol to obtain a yellow precipitate. The mixture was refrigerated at -20 °C for 15 min before filter- ing it and washing the residue with absolute ethanol. The residue was then dried to obtain a fine yellow agarose powder.
Scanning electron microscopy is a technique capable of pro- viding tangible evidence on the morphology of biomaterials [17], and it is used in the present work to visualise the structure of agarose/MCC composites. As shown in Fig. 1a, agarose exhibits a highly porous, honeycomb like assembly under a low magnification of 50x. This assembly remains unchanged upon addition of MCC (images not shown), indicating that the gelation of the polysaccharide is independent of the presence of the cellulose crystal. The presence of MCC particles could not be seen at 50x magnification, which prompted us to in- crease it to 600x. As depicted in Fig. 1b, the walls of the agarose network appear uninterrupted and smooth at the higher magnification. Addition of MCC results in the devel- opment of rough surfaces, with rod-like MCC particles being embedded within the featureless background (Fig. 1c, d). As the concentration of MCC is raised from 0.6 to 1.2% (w/w), the surface roughness corresponding to the volume of MCC particles entrapped in the polysaccharide mesh also increases. SEM micrographs suggest that the topology of the composite gel consists of agarose strands forming a continuous matrix that supports the dispersed rod-shaped particles of the MCC filler.
The 5-DTAF-labelled agarose was used to make samples as outlined earlier. These were transferred into the well of a CoverWell imaging chamber (diameter – 20 mm; depth – 2.8 mm) and precooled to 4 °C to ensure rapid gelation of agarose before the settling of cellulose crystals. Nikon A1R laser fitted to an inverted microscope, ECLIPSE Ti-E (Minato, Tokyo, Japan) was used to obtain confocal images. An argon laser at 488 nm was used to excite the 5-DTAF-labelled aga- rose and the emissions were recorded between 500 and 530 nm. Fifteen 8-bit images, 1279 µm x 1279 µm in size were captured along the Z-axis, at an interval of 5.72 µm. The distribution of MCC from 9 different regions across the samples was studied.
Two image analysis software, Imaris 9.1.0 (Bitplane, Belfast, Northern Ireland) and FIJI with Java 6 (Madison, Wisconsin, USA) were used to quantify the Z-stacks obtained in parallel. The MCC phase volume was determined using the steps outlined by Mhaske et al. (2019) [14]. In Imaris, the Z- stack was opened in the “Surfaces” feature. A Gaussian filter was applied before subtracting the background intensity to smoothen the image and reduce noise. The volume of hollow spaces within the agarose network, attributed to the presence of unstained MCC particles, was measured. In FIJI, the “Default” algorithm was used to threshold the images, and the images were inverted to make the unstained MCC particles the object and the agarose phase the background. The area of the object pixels was calculated for the Z-stack and then mul- tiplied by the Z-step to get the volume of the MCC particles. Results obtained for the nine images captured across each sample were averaged before reporting the phase volume of the MCC particles.
Infrared spectroscopy was employed next to record the chemical fingerprints of agarose, MCC and their composites (Fig. 2). The former featured characteristic absorbance bands at 3650 cm− 1 (O-H stretching), 2980 and 2900 cm− 1 (C-H vibration), 1396 cm− 1 (C-C bending), 1246 cm− 1 (sulphate esters), 1068 cm − 1 (C-O-C vibration of 3,6-anhydro-galac- tose bridges) and 877 cm− 1 (anomeric carbon’s angular C-H deformation) [18, 19]. MCC interferograms showed the gen- eral attributes of cellulose, reproducing typical absorbance bands at 1059 cm− 1 (ring vibration and C-OH bending), 1160 cm− 1 (C-O and C-O-C stretching), 1640 cm− 1 (-O- ten- sile vibration neighbouring H atoms) and a broad band be- tween 3225 and 3350 cm− 1 (stretching vibrations of O-H groups) [20–22]. Composites reproduced the absorbance bands of both individual components, with no variation in their wavenumber, hence suggesting that there is no chemical interaction between the constituents of this mixture. Furthermore, as the concentration of MCC increased in the
146
Food Biophysics
Fig. 1 SEM images of (a) 1% (w/w) agarose network at 50x and, walls of 1% (w/w) agarose network containing (b) 0, (c) 0.6 and (d) 1.2% (w/w) MCC under 600x
exhibit a single exothermic event that reproduces the charac- teristics of the agarose thermogram in terms of shape and temperature band. Once more, it is confirmed that the pres- ence of MCC does not affect the structural properties of the agarose gel.
composite, the absorbance bands of both agarose and MCC displayed a slight amplification in intensity. This is an expect- ed outcome, congruent with the increase in polymer concen- trations in their respective domains as explained in the follow- ing sections.
In addition to infrared spectroscopy, differential scanning calorimetric tests were carried out to obtain information on the micromolecular aspects of the structural behaviour of our mix- ture. Featureless thermograms are recorded from the cooling cycles of the MCC paste in Fig. 3a indicating the absence of a first-order thermodynamic transition. The major features in the heating run are two endothermic peaks. The first event with a peak at about 108 °C indicates the desorption of water from the cellulose molecule, whereas the second event at 320 °C traverses the thermal decomposition of the MCC crys- tal [22]. Figure 3b depicts the exothermic peak of agarose upon cooling with a midpoint at 23 °C, reflecting the coil-to- helix transition of the polysaccharide leading to subsequent gel formation. Corresponding profiles of the blends also
Finally, we studied the mechanical behaviour of single and binary mixtures of agarose and MCC. Figure 4 reproduces typical patterns of elastic modulus variation as a function of temperature when the polymer is cooled from 40 to 5 °C at a rate of 2 °C/min. A significant increase in Gʹ values is ob- served at around 35 °C (onset of gelation), approaching as- ymptotically an equilibrium at the end of the cooling cycle (G’ is about 103.8 Pa at 5 °C). In contrast, the storage modulus of the MCC paste produced a relatively flat pattern that remained unaffected by temperature, yielding a storage modulus value of about 105.4 Pa (results not shown). Qualitatively, the me- chanical spectra of the composite exhibit a similar trend of storage modulus development upon cooling from vestigial gelation to eventual levelling off. As the MCC concentration
147
Food Biophysics
1160
hydrated MCC paste
1640
1059
Fig. 2 FTIR spectra between 400–4000 cm-1 for agarose with 0, 0.1, 0.3, 0.6, 0.9, 1.2 and 1.5% (w/w) MCC and MCC by itself stacked upwards from the bottom
3225-3350
1% agarose + 1.5% MCC
1% agarose + 1.2% MCC
1% agarose + 0.9% MCC 1% agarose + 0.6% MCC
e c n a b r o s b A
1% agarose + 0.3% MCC
1% agarose + 0.1% MCC
1068
2980
2900
1396 1246
3650
1% agarose
877
4500
4000
3500
3000
1500
1000
500
0
2500
2000
Wavenumber (cm-1)
agarose phase and increasing its effective concentration, an outcome that leads to the formation of a stronger hydrogel.
Rheological Modelling of the Phase Behaviour in Agarose/MCC Composite Gels
increases, however, so does the mechanical strength of the composite, reaching 104.1 Pa with 1.5% (w/w) MCC addition. This outcome is partially attributed to the reinforcing effect of the hard MCC particles on the softer agarose matrix. It should also be noted that the cellulose particles absorb water [23], thereby decreasing the amount of solvent available to the
1000
Cooling
500
)
0
Heating
W m
(
-500
w o l f t a e H
-1000
-1500
Analysis thus far, established qualitatively the noninteractive na- ture of MCC crystal and agarose molecules leading to a filler composite with a relatively weak matrix supporting hard inclu- sions. We are now interested to relate quantitatively the mechan- ical strength of the constituent phases to the overall behaviour of the composite gel. That was initiated by obtaining a double- logarithmic calibration curve of storage modulus for the agarose network at concentrations between 1.0 to 2.5% (w/w); reported in the inset of Fig. 4. Mechanical properties were then followed by a set of equations adapted from the ‘sophisticated synthetic polymer research’ proposed by Lewis and Nielsen [24]:
-2000
0
50
100
300
350
400
150
250
ð1Þ
200 Temperature (°C)
G0 G
¼ 1 þ ABΦy 1 − BψΦy
0 x
0
G
1.5
− 1
1% (w/w) agarose + 1.5 % MCC
ð2Þ
B ¼
1
1% (w/w) agarose + 1.2 % MCC
y 0 G x 0 G
1% (w/w) agarose + 0.9 % MCC
)
G
0.5
y 0 x
1% (w/w) agarose + 0.6 % MCC
!
W m
(
1% (w/w) agarose + 0.3 % MCC
0
ð3Þ
Φy
1% (w/w) agarose + 0.1 % MCC
þ A ψ ¼ 1 þ 1−Φm Φ2 m
w o l f t a e H
-0.5
1% (w/w) agarose
-1
a
-1.5
0
100
150
-50
50 Temperature (°C)
Fig. 3 DSC thermograms of (a) cooling and heating profiles of hydrated MCC paste and (b) agarose cooling profiles with 0, 0.1, 0.3, 0.6, 0.9, 1.2 and 1.5% (w/w) MCC arranged successively upwards scanned at a rate of 1 °C/min
where, Gʹ, Gʹx, Gʹy are the elastic modulus of the composite, the continuous agarose network and the discontinuous MCC phase, respectively, and ɸy is the phase volume of the filler. The framework considers the concept of maximum packing fraction (ɸm) of the filler particles, which corresponds to 0.52 for random MCC rods. The shape of the filler is taken into account via the Einstein co-efficient, rendering A to be 2.08 for dispersed cylindrical, rod-shaped inclusions [25].
148
b
Food Biophysics
4.2
3.7
5.5
5
y = 3.7321x + 3.8717 R² = 0.9945
) a P (
′
3.2
Fig. 4 Profiles of Gʹ for 1% (w/w) agarose with 0 (●), 0.1% (x), 0.3% (▲), 0.6% (□), 0.9% (♦), 1.2% (-) and 1.5% (○) (w/w) MCC during cooling from 50 to 5 °C at a scan rate of 2 °C/min; the agarose calibration curve of Gʹ at 5 °C as a function of its concentration is illustrated in the inset
4.5
) a P (
′
G g o L
4
G g o L
2.7
3.5
0
0.1
0.3
0.4
0.2 Log agarose concentration (%)
2.2
5
10
15
25
30
35
20 Temperature (°C)
3D CLSM Imaging and Image Analysis for Phase Volume Estimations in the Agarose/MCC Mixture
In an effort to determine the contribution of each com- ponent to the elastic modulus of the composite, prepara- tions were centrifuged at 40 °C and 10,000 rpm for 15 min. Supernatants of clear agarose solutions were decanted and their weights were recorded in order to determine the ef- fective agarose concentration, which ranged from 1.01 to 1.08% (w/w). Respective values of Gʹx for these agarose concentrations were calculated using the calibration curve from the inset of Fig. 4. Centrifuged MCC pellets were also weighed to work out the amount of absorbed water, which was ~ 4.1 times the weight of the cellulose powder in each composite. Weights of the decanted agarose supernatant and MCC pellet obtained after centrifugation were used to calculate the values of ɸx and ɸy (phase volumes of agarose and MCC, respectively) for each composite, with ɸx + ɸy = 1.
In order to capture and then process 3D images of our com- posites, agarose was covalently labelled with DTAF and the procedure outlined by Mhaske et al. (2019) [14] was followed. Figure 6a depicts a typical 2D image of the surface of the gel containing 1% (w/w) agarose and 0.6% (w/w) MCC. Stained agarose forms a continuous blue background in which the unstained MCC particles are dispersed as black inclusions. This is an 8 bit 2D image with a dimension of 512 × 512 pixels, corresponding to an area of 1279 × 1279 µm. Fifteen such images were obtained along the Z-axis at a fixed interval or Z-step of 5.72 µm, reproduced in Fig. 6b, which allowed imaging the sample up to a depth of ~ 85 µm. The Z-stack was then utilized to generate a 3D image, as seen in Fig. 6c. FIJI, which is an open source analysis software, analyses the Z- stack for phase volume estimation by quantifying each 2D
7
4.1
6
4
)
%
,
5
(
3.9
) a P
C C M
4
, x (
′
3.8
3
G g o L
3.7
2
f o e m u l o v e s a h P
3.6
1
0
3.5
0
0.6
0.3
1.2
1.5
1.8
0.9 MCC concentration (%, w/w) Fig. 5 Values of composite Gʹ obtained via rheological experiments (x) and blending law predictions for G’ (solid line) and phase volumes of MCC (♦) for samples containing 1% (w/w) agarose with increasing MCC concentrations
The storage modulus value of the MCC paste (about 251 kPa), obtained from rheological measurements in the preceding section, was used as a constant Gʹy for all com- posites, with the assumption that equally hydrated MCC particles in the various composites (0.1 to 1.5% w/w MCC) would yield the same elastic response. Blending law modelling was able to produce a Gʹ trace that exhib- ited a close fit with the corresponding experimental values of the composites depicted in Fig. 5, an outcome that showcases its ability to predict phase behaviour in binary composites. Corresponding volumes of the MCC phase in all composites, as predicted by blending law, are also depicted in Fig. 5 being congruent with the amount of experimentally added polysaccharide. Phase volume pre- dictions of MCC will be compared with those obtained from microscopy, paired with image analysis, in the fol- lowing section in order to assess the suitability of the microscopic protocol in quantifying phase behaviour in binary co-gels.
149
Food Biophysics
Fig. 6 (a) 2D, (b) Z-stack, and (c) 3D CLSM images of 1% (w/w) agarose containing 0.6% (w/w) MCC. The blue background denotes the stained agarose phase whereas the black cavities correspond to the unstained MCC particles
image separately. In contrast, Imaris renders a 3D image using the stack of 2D images before quantifying the 3D image volumetrically.
et al. (2019) [14] was followed to estimate the MCC phase volumes and results are also shown in Fig. 7. Similarly, the Imaris predictions are a close fit to those estimated using FIJI image analysis and rheological blending laws. Statistically, the difference between the phase volume values obtained from the two types of imaging software is not significant (p > 0.05) indicating that both can equally measure water absorption and its retention in the polymeric phases of the binary mixture. Nevertheless, the two types of software have distinct ap- proaches of image quantification and their “plug-ins” vary in
7
6
FIJI
Imaris
)
%
(
5
C C M
4
3
2
f o e m u l o v e s a h P
1
0
0.1
0.3
0.6
0.9
1.2
1.5
Concentration of MCC (%)
In an 8 bit image, pixels denote a specific intensity value of the wavelength emitted by the sample within the range of 0 and 255, where a pixel with the intensity of 0 appears black and refers to an unstained sample. This serves as the negative control and is considered to be the baseline calibration in image analysis [26]. Thresholding is the most critical step in image analysis in order to accurately segregate the object pixels (MCC particles) from the background pixels (agarose network) [27]. The phase volume of MCC particles in the composite was measured using the image analysis software, FIJI. Figure 7 illustrates the final phase volume of MCC in the composite in relation to the original amount of added cellulose powder. The phase volumes estimated by FIJI are comparable with those obtained from the rheological blending-law analy- sis in Fig. 5. Increase in the MCC phase volume in the com- posite also corresponds to ~ 4.1 times the initial added amounts documented in Fig. 7 microscopically and in Fig. 5 rheologically from the MCC concentration (%, w/w) vs.. phase volume of MCC (%) relationship. Phase volume esti- mations, of course, allow quantification of the amount of sol- vent held in a given polymeric phase.
Fig. 7 Phase volume estimates plotted against MCC concentrations in mixture with 1% (w/w) agarose determined by analysing 3D CLSM images using image analysis software – FIJI and Imaris
Next, we turned our attention to the second image analysis software, Imaris. Once more the protocol used by Mhaske
150
Food Biophysics
Rheological Properties of Materials· Rheo-Optics· Biorheology. (Springer, Berlin, 1975), p. 594–600
customisability, which may affect their capacity in estimating phase volume in more complex systems, for example, in ter- tiary preparations of protein, polysaccharide and a lipid phase.
8. M. Takayanagi, H. Harima, Y. Iwata, Viscoelastic behaviour of polymer blends and its comparison with model experiments. Mem. Fac. Eng. Kyusha Univ. 23, 1–13 (1963)
Conclusions
9. X. Li et al., Studies of phase separation in soluble rice protein/ different polysaccharides mixed systems. LWT Food Sci. Technol. 65, 676–682 (2016)
10. T. Moschakis et al., in Microrheology and microstructure of water- in-water emulsions containing sodium caseinate and locust bean gum. Food Funct. 9(5), 2840–2852 (2018)
13.
11. H. Pan et al., Effect of the extent and morphology of phase separa- tion on the thermal behavior of co-blending systems based on soy protein isolate/alginate. Food Hydrocoll. 52, 393–402 (2016) 12. N. Lorén, M. Langton, A.M. Hermansson, Confocal laser scanning microscopy and image analysis of kinetically trapped phase- separated gelatin/maltodextrin gels. Food Hydrocoll. 13(2), 185– 198 (1999) J.C.G. Blonk et al., A new CSLM-based method for determination of the phase behaviour of aqueous mixtures of biopolymers. Carbohyd. Polym. 28(4), 287–295 (1995)
15.
14. P. Mhaske et al., Quantitative analysis of the phase volume of agarose-canola oil gels in comparison to blending law predictions using 3D imaging based on confocal laser scanning microscopy. Food Res. Int. 125, 108529 (2019) J. Nsor-Atindana et al., Functionality and nutritional aspects of microcrystalline cellulose in food. Carbohyd. Polym. 172, 159– 174 (2017)
Rheology-based blending laws have been used extensively in the literature to predict the mechanical strength of synthetic polymer matrices, and this is augmented in biomaterial com- posite gels by the estimation of solvent partition between polymeric phases, since these are largely aqueous prepara- tions. Nevertheless, such attempts are built around painstaking rheological experimentation and they also require consider- able modelling expertise, hence making it difficult for labora- tories to implement or reproduce. Given this, the suitability of CLSM based 3D imaging, coupled with image analysis, in quantifying phase volumes in a binary mixture was investigat- ed as a rapid alternative approach. The microscopic protocol could successfully render phase volumes of hydrated MCC particles suspended in an agarose network, an outcome that agreed remarkably well with those obtained from rheological modelling. The protocol could cope with increasing concentrations/density of the filler phase in phase volume es- timation and merits further consideration in more complex preparations.
16. N. Russ et al., Influence of nongelling hydrocolloids on the gelation of agarose. Biomacromolecules 14(11), 4116–4124 (2013) 17. C. Kyomugasho et al., Evaluation of cation-facilitated pectin-gel properties: Cryo-SEM visualisation and rheological properties. Food Hydrocoll. 61, 172–182 (2016)
18. X. Qi et al., Facile formation of salecan/agarose hydrogels with tunable structural properties for cell culture. Carbohyd. Polym. 224, 115208 (2019)
Acknowledgements The authors acknowledge the facilities and technical assistance of the RMIT University’s Microscopy and Microanalysis Facility.
19. V. Zamora-Mora et al., Chitosan/agarose hydrogels: Cooperative properties and microfluidic preparation. Carbohydr. Polym. 111, 348–355 (2014)
References
20. C.P. Azubuike, A.O. Okhamafe, Physicochemical, spectroscopic and thermal properties of microcrystalline cellulose derived from corn cobs. Int. J. Recycl. Org. Waste Agric. 1(1), 9 (2012) 21. H. Toğrul, N. Arslan, Flow properties of sugar beet pulp cellulose and intrinsic viscosity–molecular weight relationship. Carbohyd. Polym. 54(1), 63–71 (2003)
1. K. Prameela, C.Murali Mohan, C. Ramakrishna, Chap. 1 - Biopolymers for Food Design: Consumer-Friendly Natural Ingredients, in Biopolymers for Food Design, ed. by A.M. Grumezescu, A.M. Holban (Academic Press, Cambridge, 2018), p. 1–32
22. M. Murigi et al., Comparison of physicochemical characteristics of microcrystalline cellulose from four abundant Kenyan biomasses. IOSR J. Polym. Text. Eng. 1(2), 53–63 (2014)
2. C. Stenger et al., Formation of concentrated biopolymer particles composed of oppositely charged WPI and pectin for food applica- tions. J. Dispersion Sci. Technol. 38(9), 1258–1265 (2017) 3. V. Tolstoguzov, Ionic strength and poly ion charges. Int. Food
23. A.J. Gravelle, S. Barbut, A.G. Marangoni, Food-grade filler parti- cles as an alternative method to modify the texture and stability of myofibrillar gels. Sci. Rep. 7(1), 11544 (2017)
Ingred. 2, 8–9 (1990)
24. C. Semasaka et al., Modeling water partition in composite gels of BSA with gelatin following high pressure treatment. Food Chem. 265, 32–38 (2018)
25. L.W. Koh, S. Kasapis, Orientation of short microcrystalline cellu- lose fibers in a gelatin matrix. Food Hydrocoll. 25(5), 1402–1405 (2011)
4. M.M.G. Koning, J. van Eendenburg, D.W. de Bruijne, Mixed Biopolymers in Food Systems: Determination of Osmotic Pressure, in Food Colloids and Polymers, ed. by E. Dickinson, P. Walstra (Woodhead Publishing. Cambridge, 2005), p. 103–112 5. T. Giancone et al., Protein–polysaccharide interactions: Phase be- haviour of pectin–soy flour mixture. Food Hydrocoll. 23(5), 1263– 1269 (2009)
26. A.A. Karamitrou, G.N. Tsokas, M. Petrou, A pixel-based semi‐ stochastic algorithm for the registration of geophysical images. Archaeol. Prospect. 24(4), 413–424 (2017)
27. T.Y. Goh et al., Performance analysis of image thresholding: Otsu
technique. Measurement 114, 298–307 (2018)
6. P.D. Pudney, T.M. Hancewicz, D.G. Cunningham, The use of con- focal Raman spectroscopy to characterise the microstructure of complex biomaterials: foods. J. Spectrosc. 16(3–4), 217–225 (2002)
Publisher’s Note Springer Nature remains neutral with regard to jurisdic- tional claims in published maps and institutional affiliations.
7. L.E. Nielsen, in Morphology and the elastic modulus of block poly- mers and polyblends, in Rheological Theories· Measuring Techniques in Rheology Test Methods in Rheology· Fractures
151
Chapter 6
Comparison of rheological blending-law analysis and 3D
confocal laser scanning microscopy for quantification of the
phase behaviour in agarose-gelatin co-gels
Abstract
The efficacy of image analysis software – Imaris and FIJI – in quantifying 3D images
captured using confocal laser scanning microscopy (CLSM) to estimate phase behaviour in a
mixed gelling system of a polysaccharide (agarose) and a protein (gelatin) in comparison with
the rheology based blending laws is explored. Structural properties of the composites were
probed using FTIR spectroscopy, micro DSC and small-deformation dynamic oscillation in
shear. Throughout the experimental range of concentrations, gelatin formed a continuous
network whereas the agarose phase remained dispersed as either a soft or a hard filler at low
agarose concentrations (0.1-0.8%) and formed a continuous network alongside that of gelatin
at higher concentrations (1.2-3%). The phase volumes of gelatin, recorded using Imaris, were
a close match with those obtained from the blending law predictions, whereas FIJI yielded
statistically different estimates. Results suggest that the microscopic protocol using Imaris
shows promise and can potentially be used as an alternative to the theoretical blending laws in
estimating phase behaviour in biomaterial composites.
Key words: Blending law analysis; phase behaviour; 3D imaging; image analysis; confocal
152
laser scanning microscopy;
6.1. Introduction
A food system typically consists of several components, from simple molecules to macro–
molecular aggregates, held together by intermolecular forces and arranged in several structural
levels, for instance, colloidal phase. Geometric features and interactions among food
components are responsible for the ‘food structure’ which has a direct impact on the texture of
the product. A thorough understanding of such interactions is therefore important for
developing food products with intriguing texture, different from that of the original components
(Sikorski, Pokorny, & Damodaran, 2008).
Texture of the food product can be manipulated by altering the processing conditions (e.g.
ionic strength, pH, temperature), polymer concentration used, or by using a combination of
polymers at varying concentrations. When used in combination, biopolymer can exhibit
attractive or repulsive interactions resulting in complexation or thermodynamic
incompatibility, respectively based on their nature and the solvent conditions (Çakır &
Foegeding, 2011). Out of these, thermodynamic incompatibility leading to phase separation in
a protein and polysaccharide mixture is widely observed due to a low mixing entropy (Grinberg
& Tolstoguzov, 1997). However, when one or more components of a composite are gel
forming, gelation and phase separation act as competing processes, with their kinetics affecting
the morphology of the final microstructure.
The interplay between gel formation and phase separation has been successfully studied
using techniques like light scattering (Norton & Frith, 2001), spectroscopy (Anvari, Pan, Yoon,
& Chung, 2015), microscopy (Geonzon, Bacabac, & Matsukawa, 2019), calorimetry (Wang,
Dumas, & Gharsallaoui, 2019) and rheology (Wurm, Pham, & Bechtold, 2019). Rheology
based blending law analysis in particular, have been extensively used to estimate phase
153
behaviour of synthetic and natural polymer composites. These laws estimate the mechanical
properties of the mixtures from the shear modulus of the constituents in relation to their phase
volume and effective concentration (Nielsen, 1975; Takayanagi, 1963). The blending laws
however are an indirect approach, based on assumptions (e.g. the onset of gelation brings about
a substantial increase in the molecular weight of the biopolymer), and are a time-consuming
approach necessitating considerable rheological work and modelling expertise.
In the recent past, confocal laser scanning microscopy has seen increasing popularity for
qualitatively probing phase behaviour in biopolymer mixtures (Lan, Ohm, Chen, & Rao, 2020;
Zhang et al., 2017). In their work, Mhaske et al. (2019) developed a method of estimating phase
volume in a binary model system (lipid-polysaccharide) using CLSM based 3D imaging
coupled with image analysis. The technique could also accurately quantify solvent partition
and phase volume in a composite where both components were hydrophilic (agarose and
microcrystalline cellulose) (Mhaske, Farahnaky, & Kasapis, 2020). Given that the technique is
quick, easy, requiring minimum sample preparation and capable of producing phase volume
estimates that are highly comparable to those obtained from theoretical blending laws, it is
worth probing its efficacy in measuring phase volume in a complex, industrially relevant
system comprising of two gel forming polymers.
Agarose and gelatin are both extensively used as texturisers, stabilisers, emulsifiers and
gelling agents in the food, pharma, and cosmetic industry. There is a plethora of research
probing their gelation mechanism, effect induced by changes in physical and chemical
environment and their subsequent applications. Therefore, in the present study, the phase
behaviour and topology of agarose-gelatin mixtures is revisited and followed using the classical
blending laws at increasing polysaccharide concentration. Due to their inherent thermodynamic
incompatibility, the two polymers are expected to phase separate, yielding varying phase
154
topologies depending on their concentrations. The phase volume estimates obtained using the
blending laws will be compared to those from the microscopic protocol coupled with image
analysis software.
Successful implementation of this protocol will allow application to other composites of
commercial interest, making it easier for laboratories to probe the phase behaviour of
composites without painstaking rheological experimentation and modelling expertise.
6.2. Materials and methods
6.2.1. Materials
Type -1 Agarose and Type- B bovine gelatin was procured from Sigma-Aldrich (Sydney,
Australia). Agarose had less than 0.15% sulphate, 0.25% ash and 10% moisture content, whilst
gelatin had a molecular weight of about 173kDa and a Bloom value of 225 according to the
supplier. Fluorescein isothiocyanate (FITC) was also purchased from Sigma-Aldrich (Sydney,
Australia) and used to stain the protein.
6.2.2. Methods
6.2.2.1. Sample preparation
Single solutions of agarose and gelatin were prepared by dispersing powders in milliQ
water at 75 and 60°C, respectively to produce 3% aqueous solutions of each. Binary solutions
had a constant gelatin concentration of 3% and varied agarose concentrations to yield 0.1, 0.4,
0.8, 1.2, 1.6, 2.0, 2.5 and 3% of agarose in the final composite. Solutions were prepared by
dissolving required quantities of agarose in milliQ water at 75°C and cooling the mixture down
to 60°C before adding gelatin. Samples were stirred at 60°C using a magnetic stirrer for 60
155
minutes for homogeneous dissolution prior to analysis.
6.2.2.2. Fourier Transform Infrared Spectroscopy (FTIR)
Single and binary solutions of agarose and gelatin were analysed using a Perkin Elmer
(Norwalk, Connecticut, USA) spectrometer, Spectrum 2 FTIR with an attenuated total
reflectance device by GladiATR (Pike Technologies, Maddison, Wisconsin, USA). The
absorbance spectra were recorded between 400 and 4000 cm-1 with a resolution of 4 cm-1 and
averaged over 32 scans before subtracting the spectrum of water.
6.2.2.3. Differential Scanning Calorimetry (DSC)
A Setaram Micro DSC VII (Setu-rau, Caluire, France) was employed for the thermal
analysis of agarose, gelatin, and their composites. About 700 mg sample was added to
aluminium vessels and sealed, while equal weight of water was used as reference. The vessels
were transferred into an insulated instrument chamber and equilibrated at 25°C for 30 min. The
temperature was then raised to 80°C, lowered to 5°C and ramped to 80°C again. Thermographs
reported are the averaged triplicate scans obtained at a scan rate of 1°C/min.
6.2.2.4. Rheological analysis
The elastic moduli of agarose, gelatin and their composites were measured using small-
deformation dynamic oscillation in-shear on a Discovery Hybrid Rheometer by TA instruments
(New Castle, DE, USA). Samples were loaded on a preheated Peltier plate at 45°C with a 40
mm parallel plate geometry and a fixed gap of 1mm. Exposed edges of the sample were coated
with silicon oil to prevent moisture loss. Samples were cooled to 5°C, followed by a 30 min
156
time sweep and a frequency sweep from 0.1 to 100 rad/s at 5°C with a final heating run to
80°C. For all the tests, a scan rate of 2°C/min and a strain of 0.1%, which is within the linear
viscoelastic region (LVR), were used.
6.2.2.5. Confocal Laser Scanning Microscopy (CLSM) and Image Analysis
0.001% w/w FITC was added to the agarose-gelatin composites prepared in section
2.2.1 and stirred for 60 min to stain the protein. Samples were then transferred into wells of
CoverWell imaging chambers (20mm wide and 2.8 mm deep) and allowed to gel. An inverted
microscope, ECLIPSE Ti-E (Minato, Tokyo, Japan) fitted with a Nikon A1R laser was used to
capture confocal images at 10x. A 488nm argon laser was used to excite the FITC-labelled
gelatin and the emissions were captured between 500 and 530 nm. Thirty 8-bit images, 1279
µm x 1279 µm in size were captured along the Z-axis at an interval of 10 µm. A Z-stack of 2D
images from ten sections across each sample was captured to study the distribution of agarose
and gelatin phases.
The images captured were analysed in parallel using two image analysis software: FIJI
with Java 6 software(Madison, Wisconsin, USA) and Imaris 9.1.0 (Bitplane, Belfast, Nothern
Ireland). The phase volume of gelatin was estimated using the procedure outlined by Mhaske
et al. (2019). In FIJI, the images were thresholding using the “Default” algorithm. The area of
the object (gelatin) pixels was calculated for each 2D image of the Z-stack. This, multiplied by
the Z-step generated the volume of the gelatin phase. Whereas in Imaris, the Z-stack was
opened using the ‘Surfaces’ feature and the background intensity was subtracted to reduce
noise and smoothen the image. The total volume of the gelatin phase was estimated using the
generated 3D image.
6.2.2.6. Statistical Analysis
The phase volume estimates were evaluated using one-way analysis of variance
157
(ANOVA) using Minitab 18 (Minitab Inc., Pennsylvania, USA). Volume measurement was
carried out in duplicate and expressed as mean ± standard deviation, with the significant
differences being p<0.05 (Tukey test).
6.3. Results and discussion
6.3.1. Structural characterisation of agarose-gelatin mixtures
Fourier transform infrared spectroscopy was used as a preliminary test to confirm the
absence of molecular interaction between agarose and gelatin by comparing the chemical
fingerprints of the constituents with those of the composites (Fig. 6.1). The interferogram of
agarose featured a series of characteristic chemical moieties, i.e. C-H angular deformation of
anomeric carbon (877 cm-1), sulphate esters (1246 cm-1), C-O-C vibration of 3,6-anhydro-
galactose bridges (1078 cm-1), C-C bending (1396 cm-1), C-H vibration (2900 cm-1) and O-H
stretching (2980 cm-1) (Awadhiya, Tyeb, Rathore, & Verma, 2017; Qi et al., 2019). Spectral
scan of gelatin yielded major peaks in four different amide regions – Amide A denoting the N-
H bending vibrations (3325-3157), Amide I arising due to C=O stretching vibration, NH
bending and C-N stretching (1700-1600 cm-1), Amide II absorption band containing
contributions from C-N stretching and N-H bending vibrations (1565-1520 cm-1) and Amide
III representing N-H in-bending vibrations paired with C-N stretching with contributions from
C=O in-plane bending and C-C stretching (1240-670 cm-1) (Fig. 6.1a) (Cebi, Durak, Toker,
Sagdic, & Arici, 2016; Mirzapour‐Kouhdasht, Sabzipour, Taghizadeh, & Moosavi‐Nasab,
2019).
The spectral scans of the composites reproduced the absorbance bands of both the
constituents without a variance in their wavenumber, thereby suggesting that there is no
chemical interaction between the individual components of the composite (Fig. 6.1b).
158
However, a slight increase in the intensity of all the absorbance bands is observed on increasing
the agarose concentration. This is an expected outcome attributed to the increase in the effective
polymer concentration in the respective phases with the increase in agarose concentration in
the composite.
Figure 6.1.: FTIR spectra for (a) single systems of 3% agarose and 3% gelatin and (b) 3%
159
gelatin with 0.1, 0.4, 0.8, 1.2, 1.6, 2, 2.5 and 3% agarose arranged upwards.
Calorimetric tests were carried out for the complementary characterisation of phase
behaviour in aqueous mixtures of agarose and gelatin. Exotherms of the protein,
polysaccharide, and their mixtures during a cooling run from 80 to 5°C are depicted in Fig. 6.2.
A single system of 3% agarose denoted by the dashed line exhibits an exothermic peak with a
midpoint at 25°C, reflecting the formation of a gel network due to the coil to helix transition
of the agarose strands. The exotherm of gelatin (dotted line) yielded a peak culminating at a
much lower temperature of 12°C attributed to its gelation.
Figure 6.2: Cooling profiles of single systems of 3% gelatin (dotted line), 3% agarose (dashed
line), and 3% gelatin with 0.1, 0.4 0.8, 1.2, 1.6, 2, 2.5 and 3% agarose (solid lines) arranged
from bottom to top, obtained at a scan rate of 1°C/min using DSC.
When mixed, agarose and gelatin maintained the temperature sequence and overall
peak form of each transition in the composite (Fig. 6.2). A dominant peak denoting the gelation
of gelatin was observed between 0 and 20°C in all composites. Whereas the peak attributed to
the gelation of agarose is virtually absent in samples with the lowest concentrations of agarose
160
and increases in intensity as the polysaccharide concentration is increased in the composite. An
obvious explanation of the observed behaviour is that gelatin forms a continuous matrix in all
the samples, whereas contributions from the formation of agarose network are only evident in
samples containing increased agarose concentrations. As the composites neither exhibit
distortion of the individual peaks of the constituents nor generate new thermal events it can be
argued that agarose and gelatin remain in two distinct, phase separated domains in the
composite without any heterotypic association. It is worth noting that as the agarose
concentration in the composites is increased, the temperature band of agarose in the composites
shifts to higher temperatures signifying an increased thermal stability of the concentrated
agarose phase in the composite. Similar results supporting phase separation are reported in the
works of Shrinivas, Kasapis, & Tongdang (2009), who investigated the phase behaviour of
binary and ternary mixtures of agarose, gelatin, and a lipid.
Finally, mechanical measurements were carried out to obtain supramolecular evidence
on structure formation of the constituents and their mixtures. The asymmetric variations in the
values of storage modulus of single agarose and gelatin systems during a cooling and a heating
run separated by a 30 min isothermal run at 5°C exhibit a similar sequence of structure
formation as for the calorimetric thermograms in Fig. 6.2, and are depicted in Fig. 6.3a. When
the disordered and fluctuating coils of 3% agarose in an aqueous solution are cooled to 5°C,
they form an ordered three-dimensional network resulting in a significant increase in the
storage modulus, commencing at about 33.5°C. Substantial thermal hysteresis is observed
during the heating run, with the network melting at 80°C. This is a result of the highly
aggregated intermolecular associations forming the structural knots of the enthalpic network
(Fernández et al., 2008). Gelatin on the other hand, exhibits a gelation and a melting
temperature of 20.7°C and 26.2°C, respectively. A constant thermal hysteresis of six to seven
degrees throughout the experimental range for gelatin argues for a gelation brought about by
161
helices as opposed to aggregation (Parker & Povey, 2012).
Figure 6.3: Variation in storage modulus (a) as a function of temperature and time of
observation for single systems of 3% gelatin and 3% agarose, and (b) on heating 3% gelatin
plus 0.1, 0.4, 0.8, 1.2, 1.6, 2, 2.5 and 3% agarose, arranged upwards. Scan rate: 2°C/min,
162
frequency: 1 rad/s, strain: 0.1%.
Figure 6.3b shows the melting profile of the composites containing 3% gelatin with
increasing concentrations of agarose during controlled heating at 2°C/min. At the lower
experimental concentrations of agarose (0.1, 0.4 and 0.8% in Fig. 6.3b), the composite network
melts completely at the temperature range of gelatin on heating. This disintegration of the
composite with gelatin melting suggests that at low concentrations of agarose, gelatin forms a
continuous network within which agarose is dispersed as a discontinuous filler. In contrast,
when the concentration of agarose is increased to 1.2% or more, upon heating, composites
show a decrease in composite moduli on the melting of the gelatin network, but the gel remains
intact within the experimental range of temperature. This postulates that at increased
concentrations of agarose, the polysaccharide too forms a continuous network that holds the
composite gel together after gelatin melts. Thus, the melting of gelatin weakens the composite
gel but does not destroy it. This is congruent with the calorimetry results which yield a single
peak of gelatin gelation at low agarose concentrations along with the gradual development of
the second peak illustrating the formation of agarose network in composites containing a higher
concentration of agarose.
Concluding evidence on the phase behaviour of agarose-gelatin co-gels is afforded by
confocal microscopy in Fig. 6.4. The gelatin phase which is stained using FITC appears purple
in the images while the unstained agarose appears black. Figs. 6.4a and 6.4b denote samples
containing 3% gelatin with 0.1 and 0.8% agarose, respectively. It is evident from the purple
background of the images that gelatin forms a continuous matrix in which agarose is dispersed
as a discontinuous filler. The filler particles increase in number as the concentration of agarose
163
increases (Fig. 6.4b vs Fig. 6.4a).
Figure 6.4: 2D CLSM images of 3% gelatin with (a) 0.1, (b) 0.8, (c) 1.6 and (d) 2.5% agarose,
at 10x. Agarose phase is denoted in black and gelatin phase in purple. Scalebar denotes 100
µm.
Fig. 4c representing the composite containing 3% gelatin and 1.6% agarose shows
larger aggregates of agarose branching out to form a continuous network. In Fig. 4d, sample
containing 3% gelatin and 2.5% agarose depicts a clear bicontinuous network meshed together.
This confirms the hypothesis made in the preceding sections that at lower concentrations of
agarose (0.1, 0.4 and 0.8%), the composite gel has a continuous gelatin phase with agarose as
the discontinuous filler, and at higher agarose concentrations (1.2, 1.6, 2.0, 2.5 and 3%),
agarose too forms a continuous matrix along with gelatin.
6.3.2. Theoretical modelling of phase behaviour in agarose-gelatin co-gels using blending
law analysis
In order to advance a quantitative footing to the qualitative observations on phase behaviour
164
made thus far, we obtained double logarithmic calibration curves of elastic moduli within the
experimental range of concentrations for agarose (between 1 and 3.25%) and gelatin (between
1 and 10%) at 5°C; reported in Fig. 6.5a. The calibration curves with a correlation coefficient,
R2 of 0.995, are indispensable for interpolating between the experimentally obtained Gʹ values
at varying concentrations during theoretical modelling. The isostrain and isostress blending
laws given by Eqs. 1 and 2, respectively are used to follow progress in viscoelasticity of the
composites (Morris, 2009):
−1
(1) 𝐺ʹ𝑐 = ɸ𝑥𝐺ʹ𝑥 + ɸ𝑦𝐺ʹ𝑦
ɸ𝑦 𝐺ʹ𝑦
(2) + ) 𝐺ʹ𝑐 = (ɸ𝑥 𝐺ʹ𝑥
Where Gʹc, Gʹx, and Gʹy are the elastic moduli of the composite and the constituent phases X
and Y, respectively, with ɸx and ɸy being the respective phase volumes of the constituents and
ɸx + ɸy = 1. According to this framework, the isostrain conditions or upper bound behaviour is
applicable when a strong continuous matrix supports a weaker discontinuous filler, and in the
converse situation where a stronger filler is dispersed in a weak supporting matrix, the isostress
conditions or lower bound predictions are suitable.
The Takayanagi (isostrain and isostress) blending laws predict the solvent distribution
between the constituent biopolymer phases by estimating the elastic modulus on shear for all
possible solvent (water) distributions in the gelatin phase (SGEL), via the calibration curves in
Fig. 6.5a, and finding the one estimated value that matches the experimental storage modulus
of the composite (Gʹexp). Examples of the application of blending law analysis to the agarose-
gelatin co-gels in the present experimental framework are represented in Figs. 6.5b, 6.5c and
6.5d. On plotting the experimental modulus for the composite containing 3% gelatin and 0.1%
agarose in Fig. 6.5b, it intersects both agarose continuous (GʹU(Ag)) and gelatin continuous
165
(GʹU(gel)) topologies of the upper bound. Results in the preceding sections and specifically the
microscopy image presented in Fig. 6.4a, eliminate the possibility of an agarose continuous
topology at low agarose concentrations. On the other hand, a strong gelatin continuous phase
supporting agarose as a weak filler with gelatin claiming about 3/4th of the water (SGEL =0.76)
is plausible.
In contrast, when the agarose concentration in the composite was increased to 0.8%,
the experimental Gʹ value intersects just the gelatin continuous topology of the lower bound
(GʹL(gel)) of the computerized modelling at SGEL = 0.52 as illustrated in Fig. 6.5c. Results for the
composite containing 0.4% agarose were consistent to these (with SGEL = 0.69) and are hence,
not represented to avoid repetition. However, application of the blending laws to agarose-
gelatin composites containing greater than 0.8% agarose, yielded inconclusive results as shown
in Fig. 6.5d, which represents the composite containing 2.5% agarose. The experimental
modulus intersects the agarose and gelatin continuous topologies of both the upper and the
lower bound at their convergence. Evidence presented in Figs. 6.2, 6.3b and 6.4 suggests that
at higher agarose concentrations, agarose too forms a continuous network. Therefore, the
possibility of the system undergoing phase inversion at higher agarose concentrations and
forming an agarose continuous matrix with gelatin as a filler, cannot be entirely dismissed. In
addition, the arrangement of either an agarose or gelatin continuous phase, with agarose
trapping a disproportionately large amount of water in its phase (SGEL < 0.07) is physically
improbable as it would render the gelatin phase effectively dehydrated.
Results from the theoretical modelling along with the supporting evidence in the
preceding sections help conclude that in the composites containing 3% gelatin with 0.1, 0.4
and 0.8% agarose, the system is gelatin continuous with the dispersed agarose phase at 0.1%
concentration being weaker than the gelatin matrix, and stronger than it at 0.4 and 0.8%
166
concentrations. The isostrain and isostress blending law analysis however, fails to rationalise
the co-relation between the composite moduli and the constituents at higher concentrations of
agarose.
Since the microscopy images (Figs. 6.4c and 6.d) provide a direct visualisation of the
distribution of agarose and gelatin in the composite and suggest that agarose forms a continuous
network alongside that of gelatin, we pursued this molecular architecture next. This was done
by using the theoretical framework set forth by Davies that relates the composite modulus to
phase volumes of the individual networks and their respective moduli in a bicontinuous gel
1 5⁄
network (Davies, 1971):
1 5⁄ = ɸ𝑥𝐺ʹ𝑥
1 5⁄ + ɸ𝑦𝐺ʹ𝑦
(3) 𝐺ʹ𝑐
Fig. 6.5e summarises the results of this analysis for gelatin gels in the presence of the higher
concentrations of agarose studied in the present work (1.2-3%). Gratifyingly, values of
experimental composite moduli when plotted on the theoretically estimated bounds yielded
physically realistic results and the intersected predictions shift to lower volumes of water in the
gelatin phase on increasing the agarose content in the mixture. This once again confirms the
hypothesis that at agarose concentrations of 1.2% or more, the composite comprises a
bicontinuous network of agarose and gelatin.
Since agarose and gelatin undergo a segragative phase separation in the composite, the
solvent fraction in their respective phases, the final phase volumes along with the concentration
of each biopolymer in its phase and its corresponding mechanical contribution all depend on
the avidity of the two biopolymers for the aqueous medium. Thus, we utilized the ‘solvent-
avidity’ factor, p to allow for a meaningful comparison of water partition in the composites in
167
the given study.
168
169
Figure 6.5: (a) Calibration curves of Gʹ at 5°C as a function of agarose (▲) and gelatin (●)
concentrations. Modeling phase topology using – the isostrain (GʹU) and isostress (GʹL)
blending laws for 3% gelatin and (b) 0.1, (c) 0.8 and (d) 2.5% agarose, and blending law for
bicontinuous binary gels for © 3% gelatin plus 1.2,1.6, 2.0, 2.5 and 3% agarose (solid lines)
arranged upwards along with the experimental values of the same concentrations (dotted lines)
arranged upwards; and (f) predictions of the p factor for all the experimental concentrations
based on the isostrain (■), isotress (●) and bicontinuous (♦) blending laws with the inset
representing the estimated phase volume of the gelatin phase with increasing agarose
concentrations.
The p factor is the ratio of the solvent (S) to the initial biopolymer concentration © of
constituent X divided by the corresponding ratio of constituent Y, given by (Clark, Richardson,
Ross-Murphy, & Stubbs, 1983; A. J. F. s. Clark, Blanshard, & Lillford, 1987):
(4) 𝑝 = 𝑆𝑋 𝑐𝑋⁄ 𝑆𝑌 𝑐𝑌⁄
With X and Y corresponding to gelatin and agarose, respectively, and SX + SY = S, the total
amount of water (solvent) in the sample. Intersection data in all composites (Figs. 6.5b, 6.5c
and 6.5e) were replotted as a function of the p factor in Fig. 5f. It was found that as phase
behaviour of the composites varied between isostrain, isostress and biocontinuous conditions
at different agarose concentrations, according to Eqs. 1, 2 and 3, respectively, it resulted in
three distinct clusters of p values. As seen in Fig. 6.5f, there is an increase in the values of p as
the composites move from an isostrain to isotress to biocontinuous makeup, with gelatin
displaying a greater relative affinity for water with the changes in phase behaviour. The p
values within a specific family however (isostress or bicontinuous in Fig. 6.5f), exhibit a
systematic decrease, indicating a reduced proportion of water associated with the gelatin as the
170
concentration of agarose is increased. The asystematic variation of p with polymer
concentration as the system moves from a gelatin continuous to a bicontinuous makeup
suggests fundamental differences in the interplay between gelation kinetics and microscopic
phase separation at different architectural setup.
Finally, we estimated the phase volumes of the gelatin phase for the SGEL values of
intersection across all experimental composites (represented in the inset of Fig. 6.5f in percent)
using the framework established by Kasapis, Morris, Norton, & Clark (1993):
(5) Wx = x + (Sx x w)
(6) Wy = y +((1 – Sx) x w)
𝑊𝑥+ 𝑊𝑦
(7) ɸ𝑥 = 𝑊𝑥
Where, x, y, w, Sx, Wx and Wy denote the weight of polymer X, weight of polymer Y, weight of
total water in the system, solvent fraction in polymer-X phase, and the total weights of
constituent X and Y phases, respectively. These phase volume predictions will be compared
with those of the microscopic protocol to assess the efficacy of the latter in quantifying phase
volumes in phase separated systems.
6.3.3. 3D CLSM imaging and image analysis of agarose-gelatin co-gels
A Z-stack of all the experimental composites was captured and analysed using two image
analysis software – FIJI and Imaris, in parallel, according to the procedure outlined by Mhaske
et al. (2019). A typical 2D image captured from the surface of the composite containing 3%
gelatin with 0.4 and 2.5% agarose is depicted in Fig. 6.6a and Fig. 6.6d, respectively. As in
Fig. 6.4, FITC stained gelatin appears purple, whereas the unstained agarose appears black. In-
171
line with the discussion thus far, agarose phase remains as dispersed inclusions in a continuous
gelatin matrix at a low concentration of 0.4% in Fig. 6.6a, and forms a continuous matrix
alongside
Figure 6.6: (a, d) 2D, (b,e) Z-stack and (c, f) 3D CLSM images of 3% gelatin and 0.4 and 2.5%
172
agarose respectively. The gelatin phase is depicted in purple whereas agarose phase is in black.
gelatin when the concentration increases to 2.5% in the composite (Fig. 6.6d). The 2D images
are 8-bit images, 512 x 512 pixels in dimension corresponding to an area of 1279 x 1279 µm.
30 such images were captured along the Z axis at a fixed distance or Z-step of 10 µm, imaging
a ~300 µm deep section from the surface, as seen in Figs. 6.6b and 6.6e (depicting samples
containing 0.4 and 2.5% agarose, respectively). These Z-stacks (Figs. 6.6b and 6.6e) are then
used to render the 3D images seen in Figs. 6.6c and 6.6f, respectively. The Z-stacks are
analysed using an open source analysis software, FIJI, which quantifies area fraction in each
2D image and multiples it with the Z-step to give the volume fraction of a phase in the imaged
cross-section. In contrast, Imaris generates a 3D image using the Z-stack and then performs
volumetric quantification.
In the 8-bit images captured, pixels denote the intensity of the wavelength emitted by the
sample between the range of 0 and 255. An unstained sample appears black with a pixel
intensity of 0, serving as a negative control and baseline calibration in image analysis (Lu,
Chen, Wang, & Chang, 2016). The phase volume of the gelatin phase measured using the two
image processing software – Imaris and FIJI are illustrated in Fig. 6.7 along with the phase
volume estimates obtained from theoretical modelling for comparison. As seen in the figure,
the volume estimates obtained using Imaris are comparable with the theoretical predictions
indicating its suitability as an alternative for the theoretical blending laws. Phase volumes
estimated using FIJI however vary between 40 and 60% throughout the experimental range and
173
are statistically, significantly different from the theoretical outcomes.
Figure 6.7: Gelatin phase volume estimates plotted against increasing agarose concentrations
added to 3% gelatin determined by quantifying 3D CLSM images using image analysis
software – FIJI and Imaris, and the rheology based blending laws.
Given that FIJI has sixteen distinct methods of thresholding (separating the object
pixels from the background pixels), with each method employing a different algorithm, it is
possible that the method used for thresholding in the present work isn’t the most appropriate
for the system investigated. During the initial exploration of the applicability of 3D imaging
paired with image analysis in estimating phase behaviour in biopolymer composites, Mhaske
et al. (2019) found that the ‘Default’ thresholding method that utilized the iterative algorithm
yielded values closest to the theoretical predictions. This iterative algorithm by Ridler &
Calvard (1978), has also been successfully used to estimate the phase volumes in systems
comprising agarose-lipid (Mhaske, Condict, Dokouhaki, Farahnaky, & Kasapis, 2020) and
agarose-microcrystalline cellulose (Mhaske et al., 2020) previously. It is probable that while
the algorithm could accurately quantify components with a distinct phase interface, it is not
best suited for a system comprising constituents that are co-soluble. Further exploration would
174
perhaps help find the suitable algorithm in FIJI. However, as the aim of this study was to
evaluate the efficacy of the microscopic protocols developed previously in conjunction with
image analysis software in estimating phase behaviour, this wasn’t pursued. A nonadaptive
thresholding, as is the case with FIJI, would also necessitate that theoretical modeling is carried
out to verify the suitability of the thresholding method for every system, negating the purpose
of its development as an alternative to theoretical modeling.
6.3.4. Conclusions
Exploiting the advances made in imaging technology, recent attempts have been made at
developing a microscopic protocol paired with image analysis for quantifying phase volumes
in simple biopolymer composites as an alternative to the theoretical blending laws. Successful
implementation of the quick, direct microscopic protocol could potentially render the tedious
experimental work and modelling expertise that the theoretical blending laws entail,
dispensable. Therefore, the suitability of 3D CLSM imaging, paired with image analysis using
two software – Imaris and FIJI in parallel, in estimating the phase volume in a gelling
biopolymer composite of agarose and gelatin was investigated. Imaris, estimated the phase
volume of the continuous gelatin phase in close comparison with the rheological modelling as
the composite phase topology shifted from the upper bound and lower bound to a bicontinuous
network with increasing agarose concentration. Whereas, the ‘Default’ thresholding method of
FIJI which had been successfully used in the past for simpler composites, failed to cope up and
yielded inconsistent results.
The microscopic protocol thus lays the groundwork for future research, where it can be
used to determine phase behaviour accurately and rapidly in complex binary and tertiary gelling
175
biomaterial composites.
References
Anvari, M., Pan, C. H., Yoon, W. B., & Chung, D. (2015). Characterization of fish gelatin–
gum arabic complex coacervates as influenced by phase separation
temperature. International Journal of Biological Macromolecules, 79, 894-902.
Awadhiya, A., Tyeb, S., Rathore, K., & Verma, V. (2017). Agarose bioplastic‐based drug
delivery system for surgical and wound dressings. Engineering in Life Sciences, 17(2),
204-214.
Çakır, E., & Foegeding, E. A. (2011). Combining protein micro-phase separation and protein–
polysaccharide segregative phase separation to produce gel structures. Food
Hydrocolloids, 25(6), 1538-1546.
Cebi, N., Durak, M. Z., Toker, O. S., Sagdic, O., & Arici, M. (2016). An evaluation of Fourier
transforms infrared spectroscopy method for the classification and discrimination of
bovine, porcine and fish gelatins. Food Chemistry, 190, 1109-1115.
Clark, A. H., Richardson, R. K., Ross-Murphy, S. B., & Stubbs, J. M. (1983). Structural and
mechanical properties of agar/gelatin co-gels. Small-deformation
studies. Macromolecules, 16(8), 1367-1374.
Clark, A. H. (1987). Application of network theory to food systems. Food structure and
behaviour;edited by JMV Blanshard and P. Lillford. 13-34.
Davies, W. E. A. (1971). The theory of elastic composite materials. Journal of Physics D:
Applied Physics, 4(9), 1325.
Fernández, E., López, D., Mijangos, C., Duskova‐Smrckova, M., Ilavsky, M., & Dusek, K.
(2008). Rheological and thermal properties of agarose aqueous solutions and
176
hydrogels. Journal of Polymer Science Part B: Polymer Physics, 46(3), 322-328.
Geonzon, L. C., Bacabac, R. G., & Matsukawa, S. (2019). Microscopic characterization of
phase separation in mixed carrageenan gels using particle tracking. Journal of The
Electrochemical Society, 166(9), B3228.
Grinberg, V. Y., & Tolstoguzov, V. B. (1997). Thermodynamic incompatibility of proteins and
polysaccharides in solutions. Food Hydrocolloids, 11(2), 145-158.
Kasapis, S. (2008). Phase separation in biopolymer gels: A low to high-solid exploration of
structural morphology and functionality. Critical Reviews in Food Science and
Nutrition, 48, 341-359.
Kasapis, S., Morris, E. R., Norton, I. T., & Clark, A. H. (1993). Phase equilibria and gelation
in gelatin/maltodextrin systems—Part IV: Composition-dependence of mixed-gel
moduli. Carbohydrate Polymers, 21(4), 269-276.
Lan, Y., Ohm, J. B., Chen, B., & Rao, J. (2020). Phase behavior, thermodynamic and
microstructure of concentrated pea protein isolate-pectin mixture: Effect of pH,
biopolymer ratio and pectin charge density. Food Hydrocolloids, 101, 105556.
Lu, C. T., Chen, Y. Y., Wang, L. L., & Chang, C. F. (2016). Removal of salt-and-pepper noise
in corrupted image using three-values-weighted approach with variable-size
window. Pattern Recognition Letters, 80, 188-199
Mhaske, P., Condict, L., Dokouhaki, M., Farahnaky, A., & Kasapis, S. (2020). Phase volume
quantification of agarose-ghee gels using 3D confocal laser scanning microscopy and
blending law analysis: A comparison. LWT, 109567.
Mhaske, P., Condict, L., Dokouhaki, M., Katopo, L., & Kasapis, S. (2019). Quantitative
analysis of the phase volume of agarose-canola oil gels in comparison to blending law
predictions using 3D imaging based on confocal laser scanning microscopy. Food
177
Research International, 125, 108529.
Mhaske, P., Farahnaky, A., & Kasapis, S. (2020). 3D Confocal Laser Scanning Microscopy for
Quantification of the Phase Behaviour in Agarose-MCC co-gels in Comparison to the
Rheological Blending-law Analysis. Food Biophysics, 25, 1-8.
Mirzapour‐Kouhdasht, A., Sabzipour, F., Taghizadeh, M. S., & Moosavi‐Nasab, M. (2019).
Physicochemical, rheological, and molecular characterization of colloidal gelatin
produced from Common carp by‐products using microwave and ultrasound‐assisted
extraction. Journal of texture studies, 50(5), 416-425.
Morris, E. R. (2009). Functional interactions in gelling biopolymer mixtures. In Modern
biopolymer science (pp. 167-198): Elsevier.
Nielsen, L. E. (1975). Morphology and the elastic modulus of block polymers and polyblends.
In Rheological Theories· (pp. 594-600): Springer.
Norton, I. T., & Frith, W. J. (2001). Microstructure design in mixed biopolymer
composites. Food Hydrocolloids, 15(4-6), 543-553.
Parker, N. G., & Povey, M. J. W. (2012). Ultrasonic study of the gelation of gelatin: Phase
diagram, hysteresis and kinetics. Food Hydrocolloids, 26(1), 99-107.
Qi, X., Su, T., Tong, X., Xiong, W., Zeng, Q., Qian, Y., ... & He, X. (2019). Facile formation
of salecan/agarose hydrogels with tunable structural properties for cell
culture. Carbohydrate polymers, 224, 115208.
Ridler, T. W., & Calvard, S. (1978). Picture thresholding using an iterative selection
method. IEEE trans syst Man Cybern, 8(8), 630-632.
Shrinivas, P., Kasapis, S., & Tongdang, T. (2009). Morphology and mechanical properties of
bicontinuous gels of agarose and gelatin and the effect of added lipid
phase. Langmuir, 25(15), 8763-8773.
Sikorski, Z. E., Pokorny, J., & Damodaran, S. (2008). 14Physical and Chemical Interactions of
178
Components in Food Systems. Fennema’s food chemistry, 849-884.
Takayanagi, M. (1963). Viscoelastic behavior of polymer blends and its comparison with
model experiments. Journal of the Society of Material Science Japan, 12, 389-394.
Wang, J., Dumas, E., & Gharsallaoui, A. (2019). Low Methoxyl pectin/sodium caseinate
complexing behavior studied by isothermal titration calorimetry. Food
Hydrocolloids, 88, 163-169.
Waterborg, J. H. (2009). The Lowry method for protein quantitation. In The protein protocols
handbook (pp. 7-10): Springer.
Wurm, F., Pham, T., & Bechtold, T. (2019). Modelling of phase separation of alginate-
carrageenan gels based on rheology. Food Hydrocolloids, 89, 765-772.
Zhang, T., Xu, X., Ji, L., Li, Z., Wang, Y., Xue, Y., & Xue, C. (2017). Phase behaviors involved
in surimi gel system: Effects of phase separation on gelation of myofibrillar protein and
179
kappa-carrageenan. Food Research International, 100, 361-368.
Chapter 7
Conclusions and Future Research
180
7.1 Conceptual basis of the present study
Mixed biopolymer systems are increasingly used in biomedical, pharmaceutical and
food industries due to their functional and structure forming properties. While broadening the
spectrum of textural variations, they offer a better control over the nutritional value of the
product, processibility, kinetics of active compound release and shelf stability of the final
product (Prameela, Mohan, & Ramakrishna, 2018). With the growing appreciation of
biopolymers in altering techno-functionality in a wide range of products, a thorough
understanding of the structure-function relationship of biopolymer composites is crucial for the
fabrication of materials with desired functional properties (Sikorski, Pokorny & Damodaran,
2008).
This has been successfully obtained via theoretical modelling using rheology based
blending laws which assume that constituents of a biopolymer composite undergo extensive
segragative phase separation. Characterisation of solvent partition between the constituents is
done with respect to the relative affinity for water of each component. Theories borrowed from
the realm of polymer science are utilized to follow the mechanical properties of the mixtures
as outlined by Takayangi, Lewis and Nielsen (Nielsen, 1974; Takayanagi et al., 1963):
(𝐺𝑐)𝑛 = ɸ𝑋(𝐺𝑋)𝑛 + ɸ𝑌(𝐺𝑌)𝑛 (1)
(2) Gʹc/ GʹX = (1 + ABɸY)/(1-BψɸY)
2]ɸY
(3) Ψ = 1 + [(1 - ɸm)/ɸm
(4) B = [(GʹY/GʹX) – 1]/[(GʹY/GʹX) + A]
Where n equals to 1, -1 and 0.2 for isostrain, isostress and bicontinuous conditions,
respectively. GʹC, GʹX and GʹY are the storage modulus of the composite, continuous and
dispersed phases, respectively while ɸY is the phase volume of the filler. A is the Einstein
coefficient which corresponds to the shape of the filler, and ɸm is the maximum packing factor
of the filler.
Though the blending laws are robust, they are an indirect method of estimating phase
181
volume and require considerable experimentation and modeling expertise. On the other hand,
confocal laser scanning microscopy (CLSM) which is widely used in the biosciences and
biomedical sciences for qualitative analysis shows potential in being utilised for direct
quantitative estimation of phase volume in biomaterial composites when paired with the
accelerating advancement in image processing. CLSM allows the sample to be analysed in its
native state without needing pre-processing steps that other microscopic techniques entail like
fixation, sectioning and/or freeze drying. Multiple two-dimensional (2D) images can be
captured along the Z-axis by moving the focal plane in steps of specified thickness through the
sample depth. This Z-stack feature of the confocal offers unique insight into a gel’s true 3D
microstructure (Auty, 2013).
Numerous studies have demonstrated the qualitative application of CLSM in phase-
separated biopolymer mixtures including wheat protein/starch (Peighambardoust, Dadpour &
Dokouhaki, 2010), gelatin/starch/olive oil (Firoozmand & Rousseau, 2014), and whey protein
isolate/locust bean gum (Ur-Rehman et al., 2016). Examples of application of Z-stack CLSM
imaging, paired with image analysis software (e.g. ImageJ, FIJI or Image-Pro) for quantitative
analysis do exist but are limited for example, measuring size of starch granules (Nagano,
Tamaki & Funami, 2008), analysis of water droplet size distribution in spreads (Van Dalen,
2002), and protein’s fractal dimension and area in wheat dough (Li, Liu, Wu & Zhang, 2017).
7.2. Conclusions
To address the gaps outlined above, this PhD study began by developing a novel
microscopic method using Z-stack CLSM in combination with image processing software for
application in soft biomaterials. Work started off with a simple agarose-canola oil system,
where agarose upon cooling formed a continuous network within which the discontinuous
fillers of canola oil remained dispersed exhibiting distinct phase separation. As there was no
182
water partitioning between the two components, water remained in the agarose phase in its
entirety and the volume of the lipid phase depended solely on its concentration added to the
sample. Using the lipid phase volume estimates obtained via rheology based theoretical
modelling as reference, different algorithms in image analysis, using two software, FIJI and
Imaris in parallel were tested for suitability by comparing the accuracy of the phase volume
estimates obtained. Gratifyingly, the microscopic protocol developed employing the iterative
algorithm for thresholding (separating the object pixels from the background pixels) (Ridler &
Calvard, 1978), successfully determined the phase volume of the oil droplets across the
experimental range in congruence with the isostrain blending laws. The successful
implementation of the microscopic protocol merited further consideration of the protocol’s
application to more complex mixtures.
In addition, this was the first time that the rheological blending laws were applied to an
aqueous/lipid mixture, as opposed to their popular application in aqueous composites. Using
known quantities of lipid allowed for the validation of both the rheology based theoretical
modeling and the microscopic procedures. The successful application of the blending laws
suggested that they could potentially be used to analyse samples containing lipids, thus
broadening the range of biomaterials whose phase behaviour could be probed.
Next, a binary system consisting of agarose and a hard lipid, ghee that showed distinct
phase separation was probed using the developed microscopic protocol in order to quantify the
phase volume of the discontinuous filler. Ghee, being a lipid, primarily helped verify the
applicability of the blending laws to non-aqueous composites. Exhibiting a greater storage
modulus than the agarose network, it changed the composite topology from isostrain to
isostress, verifying the applicability of the isostress blending law in probing non aqueous
mixtures, while validating the novel microscopic protocol. Values of the estimated phase
volumes compared well with the theoretical estimates making the microscopic procedure a
183
potential alternative to rheological modeling in soft biopolymer hydrogels. However, at higher
lipid concentrations, perceivable inner-filtering was observed during imaging due to diffraction
of the illuminating laser as it passed through constituents with different densities and refractive
indices (Pioch, Sorg, Stadler, Hagge, & Dörfer, 2004). This in turn impacted the obtained phase
volume estimates and highlighted the need for further research in order to achieve wide-spread
applicability of the microscopic procedure.
Thus far, the efficacy of the microscopic protocol had just been tested in estimating
phase volume in a lipid-polymer system where the lipid didn’t absorb any water/solvent and
there was no water partitioning between the two components. Therefore, in the third
experimental chapter, attempts were made at estimating the water absorbed by, and the
resultant phase volume of microcrystalline cellulose (MCC) suspended in an agarose network.
The MCC particles, being hygroscopic, were expected to absorb water from the system but not
disintegrate (Nsor-Atindana et al., 2017). The microscopic protocol rendered phase volume
estimates of hydrated cellulose crystals in close comparison with those obtained from
rheological modelling and could cope with the increasing density of the MCC fillers. This
proved that the developed protocol can be used to accurately estimate not just the phase volume
of a component in a composite but also the amount of water it absorbs from the system.
With such encouraging results, work was expanded to a mixed biopolymer system
comprising two gelling biopolymers – agarose and gelatin. Upon cooling, both biopolymers
were expected to form a gel network entrapping a fraction of water within its domain. The two
image processing software were used to estimate the phase volume of the gelatin phase with
increasing agarose concentration in the composite. One of the image processing software,
Imaris, estimated the phase volume of the continuous gelatin phase in close comparison with
the rheological modelling as the composite phase topology shifted from the isostrain to
184
isostress to a bicontinuous network with increasing agarose concentration. Whereas the
‘Default’ thresholding method of FIJI which had been successfully used in the past for simpler
composites, failed to cope and yielded inconsistent results.
7.3. Future research
The microscopic protocol thus lays the groundwork for future research, where it can be
used to determine phase behaviour accurately and rapidly in complex, industrially relevant
gelling biomaterial composites. To further advance research in this area, a crucial step would
be overcoming or minimising the effect of self-shadowing or inner filtering. Within the bounds
of current research, the effect of self-shadowing has been successfully minimized by a number
of measures including - increasing the illumination intensity, using an objective with a higher
numerical aperture (NA), using the microscope in the transmission mode, etc (Sandison,
Williams, Wells, Strickler, & Webb, 1995). However, it cannot be completely overcome. Non-
diffractive bessel beams and airy beams could probably be used for illumination instead of
lasers to reduce shadowing (Di Domenico, G., Ruocco, G., Colosi, C. et al., 2018).
It would be interesting to test the efficacy of the developed protocol in quantifying
phase volume of composites in the nanoscale. CLSM has been successfully used to image nano
particles in the field of biomedical sciences, with its application limited to qualitative analysis
(Zhang, L., & Monteiro-Riviere, N., 2012; Lee et al., 2014). As nanoparticles are increasingly
used in food, packaging and pharmaceutical industries to alter microstructure, rheology and
functional properties at the nanoscale (Ameta et al., 2020), the successful application of this
technique could help researchers probe the structure-function relationship of nanoparticles with
ease.
Another interesting endeavour would be to modify the developed protocol to include
185
the additional dimension of time. With the recent advancements in imaging technology, there
have been successful applications of continuous CLSM imaging of live cells to follow progress
of biological events over time (Grossi et al., 2016). If this technique is adapted in the realm of
polymer science, it would be possible to visualise and quantify the dynamic changes in phase
behaviour in composites during gelation and/or storage.
Finally, though the technique was developed as an alternative to the blending laws, it
has the potential to surpass them in terms of application. Exploiting the capability of CLSM to
excite multiple fluorophores simultaneously, it is possible to accurately differentiate between
multiple constituents of a composite, provided they are adequately labelled (Dürrenberger et
al., 2001). This implies that the microscopic protocol can be used to capture and further
quantify images of complex tertiary or quaternary composites, whilst the application of
blending laws mostly remains restricted to binary composites.
An example of this is illustrated in Figures 7.1 and 7.2. Figure 7.1a is a 2D image of
Sample 1, containing 3% (w/w) agarose plus 2% (w/w) gelatin and 2% (w/w) canola oil. The
stained agarose and lipid phases appear green and red, respectively and the unstained gelatin is
denoted in black. The agarose phase forms the continuous phase within which the gelatin and
lipid phase remains dispersed. Figure 7.2a which is a 2D image of another tertiary composite,
Sample 2, comprising 10% (w/w) gelatin plus 5% (w/w) WPI and 5% (w/w) canola oil denotes
the stained gelatin and oil phases in red and green respectively while the whey protein isolate
appears black. Here, the discontinuous WPI and oil fillers are dispersed in the continuous
gelatin matrix. For both samples, multiple 2D images were captured along the Z-axis to obtain
a Z-stack as shown in Figures 7.1b and 7.2b, respectively. These Z-stacks are then used to
generate 3D images using the image processing software, Imaris, as shown in Figures 7.1c and
186
7.2c, respectively.
Figure 7.1: (a) 2D, (b) Z-stack, and (c) 3D CLSM images of 3% (w/w) agarose plus 2% (w/w)
gelatin and 2% (w/w) canola oil. The agarose, gelatin and oil phase is denoted in green, black
and red respectively. The composite 3D CLSM image is further split to denote 3D images of
the constituent (d) agarose and gelatin, and (e) oil phase separately (Images captured at 10x,
scale bar denotes 200μm).
Image analysis allows the user to split the 3D images based on the wavelength used for
illumination, making it possible to view and capture images of the distribution of a specific
component in a composite. Figure 7.1c is split into Figures 7.1d and 7.1e which are 3D images
187
of the agarose and gelatin, and oil phase, respectively. Figure 7.2c is split into Figures 7.2d and
7.2e which denote the 3D images of the distribution of gelatin and WPI, and the oil phase
within the composite.
(a) 2D, (b) Z-stack, and (c) 3D CLSM images of 10% (w/w) gelatin plus 5% (w/w) WPI and
5% (w/w) canola oil. The gelatin, WPI and oil phase is denoted in red, black and green
respectively. The composite 3D CLSM image is further split to denote 3D images of the
constituent (d) gelatin and WPI, and (e) oil phase separately. (Images captured at 10x, scale
188
bar denotes 200μm).
These images can then be quantified using the developed microscopic protocol using Imaris as
described in previous chapters. Phase volume estimates obtained by processing Figures 7.1d,
7.1e, 7.2d and 7.2e are summarised in Table 7.1.
Table 7.1: Experimental constituent concentrations in the composites and estimated phase
volumes with Imaris image analysis
Phase volume estimated using Imaris (%) Concentration added in the composite (%, w/w) Sample 1
3 2 2 74.5 1.9 23.6 Agarose Oil Gelatin Sample 2
10 5 5 60.6 5.1 34.3 Gelatin Oil WPI
The phase volume of canola oil in the composite depends entirely on its concentration
added to the composite, as there is no water partitioning between the aqueous and the lipid
phases. Table 7.1 shows close agreement in the oil phase volume estimates generated by Imaris
and the actual concentration of canola oil added to both the samples. Though this gains
confidence regarding the accuracy and credibility of the results obtained, further
189
experimentation using complementary techniques is needed for confirmation.
References
Ameta S.K., Rai A.K., Hiran D., Ameta R., Ameta S.C. (2020) Use of Nanomaterials in Food
Science. In: Ghorbanpour M., Bhargava P., Varma A., Choudhary D. (eds) Biogenic
Nano-Particles and their Use in Agro-ecosystems. Springer, Singapore.
https://doi.org/10.1007/978-981-15-2985-6_24
Athimethphat, M. (2011). A review on global binarization algorithms for degraded document
images. AU JT, 14(3), 188-195.
Auty, M. A. E. (2013). 4 - Confocal microscopy: principles and applications to food
microstructures. In V. J. Morris & K. Groves (Eds.), Food Microstructures (pp. 96-
P98): Woodhead Publishing.
Calvard, T. & Ridler, S. (1978). Picture Thresholding Using an Iterative Selection Method.
IEEE Transactions on Systems, Man, and Cybernetics, 8, 630-632
Di Domenico, G., Ruocco, G., Colosi, C. et al. (2018).Cancellation of Bessel beam side lobes
for high-contrast light sheet microscopy. Sci Rep 8, 17178
Dürrenberger, M. B., Handschin, S., Conde-Petit, B., & Escher, F. (2001). Visualization of
food structure by confocal laser scanning microscopy (CLSM). LWT-Food Science and
Technology, 34(1), 11-17.
Firoozmand, H., & Rousseau, D. (2014). Tailoring the morphology and rheology of phase-
separated biopolymer gels using microbial cells as structure modifiers. Food
Hydrocolloids, 42, 204-214.
Grossi, M., Morgunova, M., Cheung, S., Scholz, D., Conroy, E., Terrile, M., ... & O’Shea, D.
F. (2016). Lysosome triggered near-infrared fluorescence imaging of cellular
trafficking processes in real time. Nature communications, 7(1), 1-13.
Lee, P. L., Chen, B. C., Gollavelli, G., Shen, S. Y., Yin, Y. S., Lei, S. L., ... & Ling, Y. C.
190
(2014). Development and validation of TOF-SIMS and CLSM imaging method for
cytotoxicity study of ZnO nanoparticles in HaCaT cells. Journal of hazardous
materials, 277, 3-12.
Li, Q., Liu, R., Wu, T. & Zhang, M. (2017). Interactions between soluble dietary fibers and
wheat gluten in dough studied by confocal laser scanning microscopy. Food Research
International, 95, 19-27.
Nagano, T., Tamaki, E., & Funami, T. (2008). Influence of guar gum on granule morphologies
and rheological properties of maize starch. Carbohydrate Polymers, 72, 95–101
Nielsen, L. E. (1974). Morphology and the the elastic modulus of block polymers and
polyblends. Rheologica Acta, 13, 86-92.
Nsor-Atindana, J., Chen, M., Goff, H. D., Zhong, F., Sharif, H. R., & Li, Y. (2017).
Functionality and nutritional aspects of microcrystalline cellulose in
food. Carbohydrate Polymers, 172, 159-174.
Peighambardoust, S., Dadpour, M., & Dokouhaki, M. (2010). Application of epifluorescence
light microscopy (EFLM) to study the microstructure of wheat dough: a comparison
with confocal scanning laser microscopy (CSLM) technique. Journal of Cereal
Science, 51, 21-27.
Pioch, T., Sorg, T., Stadler, R., Hagge, M., & Dörfer, C. E. (2004). Resin penetration through
submicrometer hiatus structures: a SEM and CLSM study. Journal of Biomedical
Materials Research Part B: Applied Biomaterials: An Official Journal of The Society
for Biomaterials, The Japanese Society for Biomaterials, and The Australian Society
for Biomaterials and the Korean Society for Biomaterials, 71(2), 238-243.
Prameela, K., Mohan, C. M., & Ramakrishna, C. (2018). Biopolymers for food design:
Consumer-friendly natural ingredients. In Biopolymers for food design (pp. 1-32):
191
Elsevier.
Sikorski, Z.E., Pokorny, J. & Damodaran, S. (2008). Physical and chemical interactions of
components in food system. In S. Damodaran, K.L. Parkin and O.R. Fennema (Eds),
Fennema’s Food Chemistry (4th ed). Boca Raton: CRC Press.
Takayanagi, M., Harima, H., & Iwata, Y. (1963). Viscoelastic behavior of polymer blends and
its comparison with model experiments. Memoirs-Faculty of Engineering Kyushu
University, 23, 1-13.
Ur‐Rehman, A., Khan, N. M., Ali, F., Khan, H., Khan, Z. U., Jan, A. K., Ahmad, S. (2016).
Kinetics Study of Biopolymers Mixture with the Help of Confocal Laser Scanning
Microscopy. Journal of food process engineering, 39, 533-541.
Van Dalen, G. (2002). Determination of the water droplet size distribution of fat spreads using
confocal scanning laser microscopy. Journal of Microscopy, 208, 116-133.
Zhang, L. W., & Monteiro-Riviere, N. A. (2012). Use of confocal microscopy for nanoparticle
192
drug delivery through skin. Journal of biomedical optics, 18(6), 061214.
APPENDIX
PUBLICATIONS AND PRESENTATIONS
I. Journal articles:
Mhaske, P., Condict, L., Dokouhaki, M., Katopo, L., & Kasapis, S. (2019). Quantitative
analysis of the phase volume of agarose-canola oil gels in comparison to blending law
predictions using 3D imaging based on confocal laser scanning microscopy. Food
Research International, 125, 108529.
(This published paper is presented as Chapter 3 in this Thesis)
Mhaske, P., Condict, L., Dokouhaki, M., Farahnaky, A., & Kasapis, S. (2020). Phase volume
quantification of agarose-ghee gels using 3D confocal laser scanning microscopy and
blending law analysis: A comparison. LWT, 109567.
(This published paper is presented as Chapter 4 in this Thesis)
Mhaske, P., Farahnaky, A., & Kasapis, S. (2020). 3D Confocal Laser Scanning Microscopy
for Quantification of the Phase Behaviour in Agarose-MCC co-gels in Comparison to
the Rheological Blending-law Analysis. Food Biophysics, 25, 1-8.
(This published paper is presented as Chapter 5 in this Thesis)
II. Poster presentation
Mhaske, P., Condict, L., Dokouhaki, M., Katopo, L. and Kasapis, S.: 3D imaging using
confocal laser scanning microscopy and image analysis to quantify phase volume in
composite mixtures – a novel, direct approach. In 15th International Hydrocolloids
conference (2-5th March), Melbourne, Australia.
III. Oral presentation
Mhaske, P., Condict, L., Dokouhaki, M., Farahnaky, A. and Kasapis, S.: Quantitative phase
volume estimation of agarose-ghee gels using 3D confocal laser scanning microscopy
and blending law analysis: A comparison. In 15th International Hydrocolloids
193
conference (2-5th March), Melbourne, Australia.