Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Processing and Development of Polysaccharide-Based Biopolymers for Packaging Applications
Processing and Development of Polysaccharide-Based Biopolymers for Packaging Applications
Processing and Development of Polysaccharide-Based Biopolymers for Packaging Applications
Ebook653 pages16 hours

Processing and Development of Polysaccharide-Based Biopolymers for Packaging Applications

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Processing and Development of Polysaccharide-Based Biopolymers for Packaging Applications presents the latest cutting-edge research into the processing and utilization of bio-based polymers for packaging applications, covering materials derived from polysaccharides, polylactic acid (PLA), polyhydroxyalkanoates (PHAs), polybutylene and bio-polyethylene. The book also covers the principles of biopolymer plasticization, experimental and modeling techniques, the use of nanotechnology, and key advances relating to biopolymer-based packaging, including anti-microbials, anti-oxidative agents, and modified atmosphere packaging (MAP).

  • Introduces the principles of biopolymer plasticization and summarizes experimental and modeling techniques
  • Covers a range of important bio-based polymer resources, explaining resources, availability, characterization methods, and extraction and refining techniques
  • Supports the processing and development of bio-based polymers with enhanced functionality for advanced packaging applications
LanguageEnglish
Release dateJun 2, 2020
ISBN9780128187968
Processing and Development of Polysaccharide-Based Biopolymers for Packaging Applications

Related to Processing and Development of Polysaccharide-Based Biopolymers for Packaging Applications

Related ebooks

Materials Science For You

View More

Related articles

Reviews for Processing and Development of Polysaccharide-Based Biopolymers for Packaging Applications

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Processing and Development of Polysaccharide-Based Biopolymers for Packaging Applications - Yachuan Zhang

    Processing and Development of Polysaccharide-Based Biopolymers for Packaging Applications

    Edited by

    Yachuan Zhang, PhD

    Lethbridge Research and Development Centre, Agriculture and Agri-Food Canada, Lethbridge, Alberta, Canada

    Table of Contents

    Cover image

    Title page

    Copyright

    Contributors

    Chapter 1. Principle of biopolymer plasticization

    1. Introduction

    2. Plasticization and antiplasticization

    3. Glass transition

    4. Retrogradation of polysaccharide-based biopolymer

    5. Summary

    Chapter 2. Experimental and modeling techniques used in determining properties of biopolymers

    1. Introduction

    2. Mechanical properties

    3. Water vapor and oxygen barrier properties

    4. Moisture sorption isotherm

    5. Retrogradation and crystallinity

    6. Surface properties

    7. Summary

    Chapter 3. Characteristics of biopolymers from natural resources

    1. Introduction

    2. Starch-based films

    3. Cellulose and derivatives

    4. Kefiran

    5. Pullulan

    6. Soluble soy polysaccharide

    7. Pectin

    8. Alginate

    9. Carrageenan

    10. Agar

    Chapter 4. Biobased polymer composite from poly(lactic acid): processing, fabrication, and characterization for food packaging

    1. Introduction

    2. Overview of biopolymers

    3. Poly(lactic acid)

    4. PLA nanocomposites in packaging industry

    5. Conclusions

    Chapter 5. Processes and characterization for biobased polymers from polyhydroxyalkanoates

    1. Introduction: petroleum-based plastics/conventional plastics

    2. PHA thermoplastics

    3. Conclusion

    Chapter 6. Processes and characterization for biobased polymers from polybutylene succinate

    1. Introduction

    2. Synthesis of PBS

    3. Processing and properties of PBS

    4. Applications of PBS

    5. Synthesis of its copolymer

    6. Blends of PBS and other biodegradable polymers

    7. Compatibilization of PBS/biodegradable polymer blend

    Chapter 7. Processes and characterization of bioethanol and biopolyethylene

    1. Introduction

    2. Processes and characteristics of biopolyethylene

    3. Characterization of biopolyethylene

    4. Applications of biopolyethylene

    Chapter 8. Nano-technologies and reinforcements

    1. Introduction

    2. Polymer-based composite

    3. Biopolymer-based packaging

    4. Nanocomposites

    5. Conclusion

    Chapter 9. Biodegradable packaging antimicrobial activity

    1. Introduction

    2. Foodborne pathogens

    3. Methods to assess edible and coating film antimicrobial property

    4. Natural and biological antimicrobial compounds: characteristics and action mechanism

    5. Active packaging with antimicrobial activity

    6. Case study: Antimicrobial activity of gelatin/starch blend films incorporated with turmeric extract and cross-linked with citric acid

    7. Conclusion

    Chapter 10. Antioxidant incorporated biopolymer composites for active packaging

    1. Introduction

    2. Conclusion

    Chapter 11. Modified atmosphere packaging development

    1. Introduction

    2. Development of MAP

    3. Principles of MAP

    4. Strategies for MAP

    5. Effects of MAP on food products

    6. Polysaccharides used in MAP

    7. Other polysaccharides

    8. Mathematical modeling of MAP

    Chapter 12. Latest development of biopolymers based on polysaccharides

    1. Introduction

    2. Polysaccharide-based packaging

    3. Future prospective

    Index

    Copyright

    Elsevier

    Radarweg 29, PO Box 211, 1000 AE Amsterdam, Netherlands

    The Boulevard, Langford Lane, Kidlington, Oxford OX5 1GB, United Kingdom

    50 Hampshire Street, 5th Floor, Cambridge, MA 02139, United States

    Copyright © 2020 Elsevier Inc. All rights reserved.

    No part of this publication may be reproduced or transmitted in any form or by any means, electronic or mechanical, including photocopying, recording, or any information storage and retrieval system, without permission in writing from the publisher. Details on how to seek permission, further information about the Publisher’s permissions policies and our arrangements with organizations such as the Copyright Clearance Center and the Copyright Licensing Agency, can be found at our website: www.elsevier.com/permissions.

    This book and the individual contributions contained in it are protected under copyright by the Publisher (other than as may be noted herein).

    Notices

    Knowledge and best practice in this field are constantly changing. As new research and experience broaden our understanding, changes in research methods, professional practices, or medical treatment may become necessary.

    Practitioners and researchers must always rely on their own experience and knowledge in evaluating and using any information, methods, compounds, or experiments described herein. In using such information or methods they should be mindful of their own safety and the safety of others, including parties for whom they have a professional responsibility.

    To the fullest extent of the law, neither the Publisher nor the authors, contributors, or editors, assume any liability for any injury and/or damage to persons or property as a matter of products liability, negligence or otherwise, or from any use or operation of any methods, products, instructions, or ideas contained in the material herein.

    Library of Congress Cataloging-in-Publication Data

    A catalog record for this book is available from the Library of Congress

    British Library Cataloguing-in-Publication Data

    A catalogue record for this book is available from the British Library

    ISBN: 978-0-12-818795-1

    For information on all Elsevier publications visit our website at https://www.elsevier.com/books-and-journals

    Publisher: Matthew Deans

    Acquisitions Editor: Edward Payne

    Editorial Project Manager: Mariana L. Kuhl

    Production Project Manager: Maria Bernard

    Cover Designer: Alan Studholme

    Typeset by TNQ Technologies

    Contributors

    Guilherme José Aguilar, Msc ,     Departamento de Química, Faculdade de Filosofia, Ciências e Letras de Ribeirão Preto, Universidade de São Paulo, Ribeirão Preto, São Paulo, Brazil

    Nur Fazreen Alias, Msc ,     School of Materials & Mineral Resources Engineering, Engineering Campus, Universiti Sains Malaysia, Seri Ampangan, Nibong tebal, Seberang Perai Selatan, Pulau Pinang, Malaysia

    Hadi Almasi,     Department of Food Science and Technology, Faculty of Agriculture, Urmia University, Urmia, Iran

    Mohammad Israil Ansari,     Department of Botany, University of Lucknow, Lucknow, Uttar Pradesh, India

    Mochamad Asrofi, PhD ,     Department of Mechanical Engineering, University of Jember, Jember, Jawa Timur, Indonesia

    Amin Babaei-Ghazvini

    Faculty, Sustainable Design Engineering, University of Prince Edward Island, Charlottetown, PE, Canada

    Department of Chemistry, University of Prince Edward Island, Charlottetown, PE, Canada

    Anuj Kumar Chandel,     Department of Biotechnology, Engineering School of Lorena (EEL), University of São Paulo, Lorena, Brazil

    Marcia Eliana da Silva Ferreira, PhD ,     Faculdade de Ciências Farmacêuticas de Ribeirão Preto, Universidade de São Paulo, Ribeirão Preto, São Paulo, Brazil

    Farha Deeba, PhD ,     International Centre for Genetic Engineering and Biotechnology, New Delhi, India

    Xin Geng, PhD ,     College of Food Science and Engineering, Qingdao Agricultural University, Qingdao, Shandong, China

    Xiue Han, PhD ,     College of Food Science, Northeast Agricultural University, Harbin, Heilongjiang, China

    Jissy Jacob

    Department of Chemistry, St. Thomas College, Pala, Kerala, India

    Mahatma Gandhi University, Kottayam, Kerala, India

    Bambang Kuswandi, PhD ,     Chemo and Biosensors Group, Faculty of Pharmacy, University of Jember, Jember, Jawa Timur, Indonesia

    Usman Lawal,     Electrochemical Power Source Division, CSIR-Central Electro Chemical Research Institute (CECRI), Karaikudi, Tamil Nadu, India

    Denis Jansen Lemos Costa, Msc ,     Departamento de Química, Faculdade de Filosofia, Ciências e Letras de Ribeirão Preto, Universidade de São Paulo, Ribeirão Preto, São Paulo, Brazil

    Xiaodong Li, PhD ,     College of Food Science, Northeast Agricultural University, Harbin, Heilongjiang, China

    Sravanthi Loganathan, PhD ,     Process Engineering Division, CSIR-Central Electrochemical Research Institute (CECRI), Karaikudi, Tamil Nadu, India

    Jyoti Mala,     D. P. C. S. S. (Miller) Senior Secondary School, Patna, Bihar, India

    K.I Ku Marsilla, PhD ,     School of Materials & Mineral Resources Engineering, Engineering Campus, Universiti Sains Malaysia, Seri Ampangan, Nibong tebal, Seberang Perai Selatan, Pulau Pinang, Malaysia

    Yuvraj Singh Negi, PhD ,     Department of Polymer and Process Engineering, Indian Institute of Technology Roorkee, Saharanpur Campus, Saharanpur, Uttar Pradesh, India

    Brijesh Pandey,     Department of Biotechnology, School of Life Sciences, Mahatma Gandhi Central University, Motihari, Bihar, India

    Swarnaprabha Pany, MSc ,     Laboratory of Biomaterials and Regenerative Medicine for Advanced Therapies; ICMR-Regional Medical Research Centre, Department of Health Research, Bhubaneswar, Odisha, India

    Sanghamitra Pati, MBBS, MD, MPH ,     Laboratory of Biomaterials and Regenerative Medicine for Advanced Therapies; ICMR-Regional Medical Research Centre, Department of Health Research, Bhubaneswar, Odisha, India

    Paras Porwal,     Amity Institute of Biotechnology, Amity University Uttar Pradesh, Lucknow Campus, Lucknow, Uttar Pradesh, India

    Ruchir Priyadarshi, PhD ,     Department of Food and Nutrition, BioNanocomposite Research Centre, Kyung Hee University, Seoul, Republic of Korea

    Saurabh Singh Rathore,     Department of Biotechnology, School of Life Sciences, Mahatma Gandhi Central University, Motihari, Bihar, India

    Rajashree Sahoo, MSc ,     Laboratory of Biomaterials and Regenerative Medicine for Advanced Therapies; ICMR-Regional Medical Research Centre, Department of Health Research, Bhubaneswar, Odisha, India

    Sangram Keshari Samal, PhD ,     Laboratory of Biomaterials and Regenerative Medicine for Advanced Therapies; ICMR-Regional Medical Research Centre, Department of Health Research, Bhubaneswar, Odisha, India

    A. Swaroop Sanket, MSc, MTech ,     Laboratory of Biomaterials and Regenerative Medicine for Advanced Therapies; ICMR-Regional Medical Research Centre, Department of Health Research, Bhubaneswar, Odisha, India

    Sauraj, PhD ,     Department of Polymer and Process Engineering, Indian Institute of Technology Roorkee, Saharanpur Campus, Saharanpur, Uttar Pradesh, India

    Iman Shahabi-Ghahfarrokhi,     Department of Food Science and Technology, Faculty of Agriculture, University of Zanjan, Zanjan, Iran

    Akhilesh Kumar Singh,     Department of Biotechnology, School of Life Sciences, Mahatma Gandhi Central University, Motihari, Bihar, India

    Satarudra Prakash Singh,     Department of Biotechnology, School of Life Sciences, Mahatma Gandhi Central University, Motihari, Bihar, India

    Janmejai Kumar Srivastava,     Amity Institute of Biotechnology, Amity University Uttar Pradesh, Lucknow Campus, Lucknow, Uttar Pradesh, India

    Delia Rita Tapia-Blácido, PhD ,     Departamento de Química, Faculdade de Filosofia, Ciências e Letras de Ribeirão Preto, Universidade de São Paulo, Ribeirão Preto, São Paulo, Brazil

    Sabu Thomas, PhD, FRSC

    School of Chemical Sciences, Mahatma Gandhi University, Kottayam, Kerala, India

    International and Inter University Centre for Nanoscience and Nanotechnology, Mahatma Gandhi University, Kottayam, Kerala, India

    Ravi Babu Valapa, BTech, MTech, PhD ,     Electrochemical Power Source Division, CSIR-Central Electro Chemical Research Institute (CECRI), Karaikudi, Tamil Nadu, India

    Yachuan Zhang, PhD ,     Lethbridge Research and Development Centre, Agriculture and Agri-Food Canada, Lethbridge, Alberta, Canada

    Rob Zhang, E.I.T ,     Tower Engineering, Winnipeg, Manitoba, Canada

    Chapter 1

    Principle of biopolymer plasticization

    Yachuan Zhang, PhD ¹ , and Xin Geng, PhD ²       ¹ Lethbridge Research and Development Centre, Agriculture and Agri-Food Canada, Lethbridge, Alberta, Canada      ² College of Food Science and Engineering, Qingdao Agricultural University, Qingdao, Shandong, China

    Abstract

    To overcome the brittleness of native polysaccharide, plasticization is required to soften polysaccharide structure and improve flexibility. Three plasticization mechanisms, that is, lubricity theory, gel theory, and free volume theory, were summarized in this chapter. The efficiency of two major plasticizers, that is, polyols and amide-containing molecules, was discussed and compared. Mechanisms of glass transition and retrogradation were explicated. Plasticizer depresses the glass transition temperature (T g ) of the polysaccharide-based polymer from very high, such as 243°C for native starch, to normal temperature range. The plasticizer can either prevent retrogradation from occurring or induce it to happen. Therefore, a critical concentration of plasticizer is required to achieve successful plasticization and prevent retrogradation. Moreover, 10% urea or 15% glycerol is considered the critical concentration to successfully plasticize the thermoplastic starch. Factors impacting the retrogradation rate of thermoplastic starch were reviewed, including the botanical source of the starch, storage temperature, storage time, storage relative humidity, and plasticizer content. Waxy corn starch is considered to be the least prone to retrogradation. Amylopectin containing long glucan chains is relatively prone to retrogradation. Low temperature is considered to be in favor of the retrogradation. However, when stored at a temperature below T g , the starch molecule is in a stable glassy state, and retrogradation does not occur or is extremely slow.

    Keywords

    Antiplasticization; Biopolymer; Plasticization; Plasticizer; Retrogradation

    1. Introduction

    Similar to conventional synthetic polymers, native polysaccharides, such as starch, cellulose, and chitin, contain strong intermolecular and intramolecular hydrogen bonding, which make them very brittle and difficult to handle in producing bioplastics. Consequently, the plasticizer is an essential component in the formation of these polysaccharide-based bioplastics with decreased brittleness and improved flexibility and elasticity. Plasticization mechanisms have been proposed. Some plasticizers have been reported in the literature. Polysaccharide-based biopolymers experience multiple physical phase transitions by thermal treatment, including glass transition, gelatinization, crystallization, and melting. It is very necessary to investigate the thermal properties of polysaccharide-based biopolymers as they behave differently under different end-use conditions ranging from freezing to cooking. Among these thermal properties, the glass transition is the most important parameter in determining the mechanical properties and in controlling the kinetics of crystallization of amorphous polymers (Shi et al., 2007). During storage, retrogradation, also known as aging, occurs in the amorphous phase, which impacts the properties of the polysaccharide-based biopolymers. This chapter discusses plasticization principles and provides an overview of the research regarding plasticization, glass transition, and retrogradation of the polysaccharide-based biopolymers.

    2. Plasticization and antiplasticization

    Native polysaccharides do not exhibit thermoplasticity. They are brittle and susceptible to cracking due to the strong cohesive energy density of polymer molecules (Lim et al., 2002). These drawbacks can be resolved by the addition of plasticizers (Sanyang et al., 2015). In the presence of plasticizers, at high temperatures (normally 90–180°C) and under shear, polymers readily melts and flows, allowing for its use as an extruding, injection molding, or blowing material (Ma et al., 2007). The ultimate role of plasticizers is to break the inter/intramolecular hydrogen bonds and reduce the strong inter/intramolecular interactions. This process of overcoming the brittleness in polymers by softening the structure and by increasing the mobility of the macromolecular chains (Pushpadass et al., 2008), resulting in a lowering of processing temperature, is termed as plasticization (Zhang et al., 2014). Three theories have been established to explain how plasticizers work to accomplish plasticization effects. The most popular one is the lubricity theory, in which plasticizer acts as a lubricant to facilitate movements of the macromolecules over each other. The gel theory considers the disruption of polymer–polymer interactions (hydrogen bonds and van der Waals or ionic forces). The free volume theory states that a plasticizer increases free volume between the polymers and lowers its glass transition temperature (T g ). Regardless of which theory is the most appropriate, the role of a plasticizer molecule is to interpose itself between the polymer chains and reduce the forces that hold the chains together (Gioia and Guilbert, 1999). As a result, the mobility of polymeric chains increases, which improves the flexibility, extensibility, and ductility of plasticized polymers (Sanyang et al., 2015).

    Requirements for a successful plasticizer have been summarized by Zhang et al. (2014), which include small molecule, polarity, hydrophilicity, compatibility with polymer, and high boiling point. The working parts or active sites in a plasticizer are the hydrophilic parts, such as hydroxyl groups (Zhang and Han, 2006a). For a starch-based biopolymer, successful plasticizers are categorized into two groups. One is a polyol, including water, glycerol, xylitol, sorbitol, sucrose, fructose, and glucose, while another is amide-containing molecules, including urea, amides, and amino acids (Wang et al., 2014). Theoretically, water is the most effective plasticizer due to its small size. However, due to the hydrophilic nature of the most polysaccharide-based biopolymers, the water content in the polymers is not stable. It varies as per the environmental relative humidity (RH). When the environmental RH increases, the polysaccharide-based biopolymer picks up water from the environment. Otherwise, it loses moisture to the environment. For this reason, glycerol attracts more attention and becomes the most preferred plasticizer in polysaccharide-based biopolymers. Zhang and Han (2006a,b) investigated the effect of glycerol in pea starch films. They found glycerol-plasticized films contained prominently high moisture content and announced glycerol might not act as a plasticizer. Instead, it acted as a water-holding agent to utilize the plasticizing activity of water molecules. Water held by the glycerol molecules positioned between the starch polymers and glycerol molecules, leading to increased spatial distance between starch polymers (Zhang and Han, 2006a). They further assumed the high water holding capacity of the glycerol is probably due to the high polarity of the glycerol molecule, which has a dielectric constant of 42.5 at 25°C. Meanwhile, the thermomechanical analysis demonstrated that the glycerol lowered the glass transition temperature (T g ) of the starch film down to −70°C, indicating that the plasticization effect of glycerol is super good. The thermomechanical analysis also demonstrated the glycerol-plasticized starch film needed very little apparent activation energy (ΔH α ) to go through the glass transition, which is also termed as α-relaxation, exhibiting the glycerol-plasticized starch film absorbed a small amount of energy from the environment to overcome the cohesive forces between the starch polymers when the film went from the glassy state through the glass transition zone to the rubbery state (Zhang and Han, 2006b). Pushpadass et al. (2008) announced that 20% glycerol is required to successfully plasticize the starch-based film, while 25% glycerol is the optimal concentration to reach the maximum tensile strength (TS) and optimum modulus of elasticity (EM). Sanyang et al. (2015) investigated the effect of glycerol content on the physical and thermomechanical properties of sugar palm starch. They concluded that 30% (w/w) glycerol exhibited better mechanical properties.

    Small amide-containing molecules have also proved to be good plasticizers (Ma and Yu, 2004; Huang et al., 2006; Yang et al., 2006a, 2006b; Wang et al., 2014). Wang et al. (2014) used urea as a plasticizer on starch film. They revealed that only when the urea content was between 10% and 30% (w/w, urea rate relative to the total dry basis), did the plasticization occur. When below 10%, an antiplasticization phenomenon occurred. When above 30%, a phase separation appeared. Based on their experimental results, Wang et al. (2014) proposed a schematic representation for urea states in starch films as shown in Fig. 1.1. According to their explanations, when the urea content was 10% or less, urea was caught by starch polymer through H-bonding, resulting in the obstruction of the mobility of the starch segments, leading to a typical antiplasticization occurred, such as enhanced TS and reduced elongation at break (EB). When urea was between 10% and 30%, the urea content became saturated. Free urea began to attract moisture. Both urea and water molecules facilitated the mobility of the starch segments. As a result, TS decreased and EB increased. When the urea was above 30%, urea became oversaturated. Phase separation appeared.

    Figure 1.1 Schematic representation of urea states in urea-plasticized thermoplastic starch with different urea concentrations. 

    From Wang, J.L., Cheng, F., Zhu, P.X., 2014. Structure and properties of urea-plasticized starch films with different urea contents. Carbohydr. Polym. 10, 1109–1115.

    Compared to the polyol, urea, formamide, and acetamide were considered to be more effective to plasticize the thermoplastic starch. For example, Zullo and Iannace (2009) announced that the urea/formamide mixture worked more effectively than glycerol in producing homogenous and robust films. They found the presence of urea/formamide resulted in higher EB and lower EM than glycerol-plasticized starch films. Scanning electron microscope further verified the starch film containing urea/formamide exhibited a smoother surface than the glycerol-plasticized starch films, as shown in Fig. 1.2. Ma and Yu (2004) explained why urea, formamide, and acetamide plasticized thermoplastic starch more efficiently. They calculated the hydrogen bond energy of urea-starch, formamide-starch, acetamide-starch, and glycerol-starch and established a strength order of the hydrogen bonding, which is urea   >   formamide   >   acetamide   >   polyols. However, even though urea, formamide, and acetamide work well as plasticizers in thermoplastic starch, they are not allowed to be used in food packaging and edible films for food safety because of the high toxicity of these reactants (Yu et al., 2010).

    To achieve successful plasticization, a critical content of plasticizer is required. If the plasticizer content is below the critical content, the plasticizer shows an antiplasticization effect, resulting in increased TS and EM and decreased EB of the starch film. The critical content of glycerol is not constant. It varies depending on factors, such as starch resources and production conditions. As mentioned earlier, Wang et al. (2014) indicated 10% urea was required to plasticize starch film, while Pushpadass et al. (2008) revealed that 20% glycerol was required to plasticize starch film. Lourdin et al. (1997), Godbillot et al. (2006), and Zhang and Han (2010) suggested that the glycerol critical content was 15% when plasticizing the starch film. Moreover, 25% of glycerol was further considered to be the optimal concentration because the starch film would exhibit maximum TS and optimum EM (Pushpadass et al., 2008).

    Figure 1.2 SEM micrographs of glycerol-plasticized starch film (left) and urea/formamide-plasticized starch film (right). 

    Adapted from Zullo, R., Iannace, S., 2009. The effects of different starch sources and plasticizers on film blowing of thermoplastic starch: correlation among process, elongational properties and macromolecular structure. Carbohydr. Polym. 77, 376–383.

    3. Glass transition

    All polymers will experience several thermal transitions when temperature increases or decreases. Glass transition, which is also denoted as α-relaxation, is one of them. When the environmental temperature is below a specific temperature range, due to the inter- or intramolecular interactions, polymer molecules in the amorphous zones, such as amylose and amylopectin in gelatinized starch, tightly held together. As a result, the long-scale segmental motion of the polymer molecules halts. Then, the biopolymer appears hard and brittle. When the temperature rises, the polymer molecules gradually obtain energy from the environment. Once the environmental temperature rises to above the specific temperatur1e range, the polymer molecules undergo large-scale segmental motion. As a result, the polymer, in turn, becomes soft and pliable. During this process, the polymer is called to go through from a glassy state to a rubbery state. This process is defined as the glass transition. This temperature range is defined as the glass transition temperature, which is noted T g . T g is not a specific temperature point. Instead, it is a temperature range. When a polymer experiences glass transition, its properties, such as free volume, heat capacity (C p ), thermal expansion coefficient (α), dielectric coefficient (ε), and viscoelastic properties, change (Bhandari and Howes, 1999). Accelerated changes in these properties as a function of temperature are the basis for detecting glass transition. In practice, two technologies are most commonly used to test glass transition. They are respectively differential scanning calorimetry (DSC) and dynamic mechanical analysis (DMA). DSC detects changes in heat capacity (C p ) (Ribeiro et al., 2003; Kumar et al., 2011). Polymer acquires a higher heat capacity when above the T g than it does when below T g . This is a working principle behind DSC. Change in heat capacity does not suddenly occur but takes place over a temperature range—the T g . DMA measures viscoelastic moduli, such as storage modulus (E′), loss modulus (E″), and tanδ (E″/E′) of the polymer as they are deformed under a period (sinusoidal) deformation (stress or strain) with temperature scan (Zhang and Han, 2006a; Zhang et al., 2014). E′ is a measurement of the energy stored and recovered in a cyclic deformation, and tanδ is the ratio of the energy lost to the energy stored in a cyclic deformation (Zhang et al., 2012). Over the glass transition zone, the E′ decreases with the temperature increasing, meaning polymer becomes less elastic or more permanently deformed. In general, DMA is considered to be more sensitive to glass transition than DSC.

    Figs. 1.3 and 1.4, respectively, indicate the examples of DSC thermograms and DMA curves of E′ and tanδ as a function of temperature ramping low temperature to high temperature, with thermoplastic starch being a representative (Zhang et al., 2013; Zhang and Han, 2006b). In Fig. 1.3, T g is defined as the middle point of the heat flow step, which are respectively 17.56°C, 14.13°C, and 10.64°C. Meanwhile, in Fig. 1.4, T g is defined as the midpoint between the onsets of the fall in E′ which is about −30°C, or as the peak of tanδ curve which is about −10°C. The tanδ is normally found at a temperature higher than the onset or the midpoint temperature of the E′ drop. So, it is very necessary to indicate how the discreet T g is defined as it is reported. Fig. 1.4 shows that the E′ decreased from −60 to 20°C, which is the glass transition zone or the α-relaxation, suggesting that the thermoplastic starch became less elastic.

    Some studies of tanδ displayed two relaxations in the temperature range where the E′ dropped (Garcia et al., 2011; Montero et al., 2017), representing a lack of miscibility or microheterogeneity of the polymer matrix because each phase exhibits its glass-rubber relaxation. For example, Montero et al. (2017) carried out studies on glycerol-plasticized starch and found two relaxations. One was at around −60   Cº, while another at 10–50°C. They claimed the relaxation that occurred at the lower temperature is ascribed to the glass transition of glycerol-rich domains, while the relaxation that happened in the higher temperature is associated with the movement of amylopectin chains in the amorphous regions of the starch-rich domains.

    Figure 1.3 DSC thermograms of glass transition of thermoplastic starch/MMT samples, with MMT content respectively being 0%, 4%, and 8%. 

    From Zhang, Y., Liu, Q., Hrymak, A., Han, J.H., 2013. Characterization of extended thermoplastic starch reinforced by montmorillonite nanoclay. J. Polym. Environ. 21, 122–131.

    Figure 1.4 DMA storage modulus E′ and tanδ curves of the starch films, at frequencies of 1, 5, and 10   Hz, respectively. 

    From Zhang, Y., Han, J.H., 2006b. Mechanical and thermal characteristics of pea starch films plasticized with monosaccharides and polyols. J. Food Sci. 71, E109–E118.

    To go through the glass transition zone from low temperature to high temperature, polymer acquires energy from the environment. This energy is termed as apparent activation energies (ΔH a ), which can be determined by the Arrhenius relationship (Kalichevsky et al., 1993):

    (1.1)

    where f is the frequency of DMA mechanical oscillation, ΔH a is the apparent activation energy, R is the universal gas constant, and T is the temperature. E′ and tanδ peaks shift to a higher temperature as the frequency of DMA mechanical oscillation increases, as shown in Fig. 1.4. It is expected for any thermally activated relaxation process (Lazaridou and Biliaderis, 2002). Fig. 1.5 shows an example of the variation of T g as a function of the frequency of DMA mechanical oscillation (1, 5, and 10   Hz) for sorbitol-plasticized starch films at different sorbitol content levels (Zhang and Han, 2006b).

    Figure 1.5 Arrhenius plots of ln(frequency) versus reciprocal temperature (K −¹) for sorbitol-plasticized starch films at different sorbitol content levels. 

    From Zhang, Y., Han, J.H., 2006b. Mechanical and thermal characteristics of pea starch films plasticized with monosaccharides and polyols. J. Food Sci. 71,E109–E118.

    Zhang and Han (2006b) conducted experiments to study the ΔH a of starch-based biodegradable films plasticized by glycerol, sorbitol, glucose, fructose, and mannose, respectively. According to Eq. (1.1), glycerol-plasticized starch films acquired ΔH a of 102.2–133.0   kJ   mol −¹ to go through the glass transition zone, while the sorbitol-, glucose-, fructose-, and mannose-plasticized starch films acquired about 200–300   kJ   mol −¹. Lazaridou and Biliaderis (2002) reported chitosan and starch composite films needed ΔH a of 259.9–400.3   kJ   mol −¹. Corn starch and pullulan blends respectively plasticized by sorbitol and xylose needed ΔH a of 226 and 296   kJ   mol −¹ (Biliaderis et al., 1999). Lazaridou and Biliaderis (2002) announced the relaxation that is associated to the movement of polymer chains in the amorphous regions of the polysaccharide-rich domains always acquires more ΔH a than the relaxation that is ascribed to the glass transition of plasticizer rich domains, indicating larger degree of changes in segments mobility and cooperativity of the polysaccharide.

    T g of the polysaccharide is greatly impacted by the plasticizer. For example, native starch is predicted to have T g of 243°C, which will never be achieved as the starch decomposes before 243°C (Bhandari and Howes, 1999). However, the addition of plasticizer depresses T g to the normal temperature range. Zhang et al. (2011) reported that starch film containing 25%–27% glycerol achieved T g of 22.4–23.4°C. Zhang et al. (2012) further confirmed that the addition of 30% glycerol in starch film achieved a T g of 22.2°C. Change et al. (2006) reported that tapioca starch film containing 20% glycerol and 11.23% moisture possessed T g of 37.6°C. With increasing plasticizer amount, T g decreases. Several T g values of polysaccharide-based polymers are listed in Table 1.1.

    Table 1.1

    From Zhang, Y., Rempel, C., Liu Q., 2014. Thermoplastic starch processing and characteristics – a review. Crit. Rev. Food Sci. Nutr. 54, 1353–1370.

    4. Retrogradation of polysaccharide-based biopolymer

    Retrogradation happens with time in polysaccharide-based polymers. It is the reassociation or recrystallization process of polysaccharide chains in amorphous regions. For example, starch chains in amorphous regions are metastable and, as a result, amylose and amylopectin chains vibrate, move, and finally retrogradate over time during storage. In other words, amylose and amylopectin chains realigned themselves into an ordered structure—the crystalline structure, if they are left alone for a long enough period. Goesaert et al. (2005) schematically presented the retrogradation process of starch molecules as shown in Fig. 1.6, in which the retrogradation was described in three continuous steps. In the first step, native starch exists in the granule. In the second step, the starch granules swell and disrupt with water and under heating, resulting in amylose and amylopectin molecules leaching from granules. In the third step, amylose and amylopectin associate, realign, and form a network under cooling conditions. Finally, the amylose and amylopectin recrystallize to form an ordered structure. Retrogradation has been the subject of research throughout the last 20   years. Liu and Han (2005), respectively, studied the crystalline structure of amylose and amylopectin film under an inverted phase-contrast microscope and found dendrites with 90-degree branching in amylose film and a networking of interlinked clusters in amylopectin film. Tako and Hizukuri (2002) proposed a retrogradation mechanism at a molecular level shown in Fig. 1.7. When retrogradation occurs, O-6 of D-glucosyl residues of amylose molecules and OH-2 of D-glucosyl residues of short side-chains of amylopectin molecules interact through a hydrogen bonding. Meanwhile, OH-2 of D-glucosyl residues of amylose molecules and O-6 of D-glucosyl residues of short side-chains of amylopectin molecules also interact through a hydrogen bonding (Fig. 1.7A). In addition, hydrogen bonding between O-3 and OH-3 of D-glucosyl residues on different amylopectin molecules also occurs (Fig. 1.7B). Zhang and Han (2010) studied the retrogradation of the pea starch film and announced that the crystallite size in the starch film was in the range of 0.7–7   nm and the crystal model was B-type nanoscale crystal. They also found that the glycerol-plasticized film contained more crystallinity than the pure starch film, which were, respectively, 10%–20% and 5.7%. One explanation they proposed is that the plasticizer enhanced the starch molecules' mobility that led to the sliding of starch polymers to parallel position and developed the stable crystalline structure during the storage period. Without plasticizer, starch molecules interacted with each other through the strong hydrogen bonding which decreased the mobility of the starch molecules. Throughout the realignment process, the reduction in the intermolecular distance between glucan chains leads to the removal of water or plasticizer, a phenomenon known as syneresis (Vamadevan and Bertoft, 2018). The retrogradation could result in changes in the properties of starch-based biopolymers, such as mechanical properties, water vapor permeability, and moisture sorption isotherms. Zhang and Han (2010) reported that when the crystallinity increased from 1% to 20%, the moisture content, oxygen permeability, water vapor permeability, and EB decreased, while EM increased. This phenomenon that the plasticizer causes the retrogradation and leads to the decrease of the polymer quality is termed as antiplasticization. Fig. 1.8 indicates plasticization and antiplasticization caused by glycerol and mannose. Fig. 1.8 shows the curve of 25% glycerol starch film had a lower slope, lower TS, and extended strain at break, than the pure starch film, exhibiting the plasticization effect. However, at 10% of glycerol, the film exhibits antiplasticization because the EB became lower than the pure starch film. Fig. 1.8B shows the obvious antiplasticization effect of mannose at both 15% and 25% levels (Zhang and Han, 2010). The retrogradation rate of thermoplastic starch depends on a number of variables, including storage temperature, botanical source of the starch, ratio of amylose to amylopectin, water, or plasticizer content, and structure of amylose and amylopectin (Zhang and Rempel, 2012). Some of these factors are summarized as follows:

    Figure 1.6 Schematic representation of changes that occur in a starch–water mixture during heating, cooling, and storage. (I) Native starch granules; (II) gelatinization, associated with swelling [a] and amylose leaching and partial granule disruption [b], resulting in the formation of a starch paste; (III) retrogradation: formation of an amylose network (gelation/amylose retrogradation) during cooling of the starch paste [a] and formation of ordered or crystalline amylopectin molecules (amylopectin retrogradation) during storage [b]. 

    From Goesaert, H., Brijs, K., Veraverbeke, W.S., Courtin, C.M., Gebruers, K., Delcour, J.A., 2005. Wheat flour constituents: how they impact bread quality, and how to impact their functionality. Trends Food Sci. Technol. 16, 12–30.

    Figure 1.7 (A) Hydrogen bonding between amylose and amylopectin molecules. (B) Retrogradation mechanism

    Enjoying the preview?
    Page 1 of 1