Edible Insects Processing for Food and Feed: From Startups to Mass Production 0367746948, 9780367746940

Consumers around the world are becoming increasingly aware of the significant impacts of food consumption on the environ

583 130 78MB

English Pages 369 [370] Year 2023

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Edible Insects Processing for Food and Feed: From Startups to Mass Production
 0367746948, 9780367746940

Table of contents :
Cover
Half Title
Title Page
Copyright Page
Table of Contents
Preface
About the Editors
List of Contributors
Chapter 1 The Role of Insects for Nature and Humans
1.1 Introduction
1.1.1 Insects in Plant Reproduction
1.1.2 Insects in Waste Biodegradation
1.1.3 Insects in Controlling Harmful Pest Species
1.1.4 Insect Food in China: History and Current Status
1.1.4.1 Long history of insect food in China
1.1.4.2 Current insect food in China
1.2 Conclusion
1.3 Prospective and Challenges
References
Chapter 2 Environmental Impact of Edible Insect Processing
2.1 Introduction
2.2 Insect Feed Processing
2.3 Killing, Decontamination, Separation
2.4 Fractionation – Allocation or Substitution
2.5 Complex Processing Technologies
2.6 Food or Feed Applications
2.7 Circular Economy Relevance (Side Streams, Special Products)
2.8 Conclusions and Outlook
References
Chapter 3 Legislation
3.1 Disclaimer
3.2 Regulatory Aspects
3.2.1 Introduction
3.2.2 Traditional Entomophagy – Insects as Food
3.2.3 Traditional Entomophagy – Insects for Feed
3.2.4 Non-Traditional Entomophagy – Insects for Food
3.2.5 Non-Traditional Entomophagy – Insects for Feed
3.3 Practical Application of Legislation
3.3.1 Initial Steps
3.3.2 Traditional Production Systems
3.3.2 Non-traditional Production Systems
3.4 Possible Future Developments and Needs for Regulation
Notes
References
Chapter 4 Nutrient Content and Functionalities of Edible Insects
4.1 Introduction
4.2 Nutritional Composition of Edible Insects
4.2.1 Protein and Amino Acids
4.2.2 Energy Value
4.2.3 Lipids and Fatty Acids
4.2.4 Minerals
4.3 Digestibility of Edible Insects
4.4 Health Benefits
4.4.1 Gastrointestinal Health
4.5 Insect Protein Hydrolysates
4.6 Protein Functionalities
4.6.1 Protein Solubility
4.6.2 Emulsifying Properties
4.6.3 Coagulation
4.6.4 Surface Charge
4.6.5 Surface Hydrophobicity
4.6.6 Water Holding Capacity
4.6.7 Oil Holding Capacity
4.6.8 Colour
4.6.9 Foaming Properties
4.7 Bioactivities
4.7.1 Antioxidant Properties
4.7.2 Angiotensin Converting Enzyme (ACE) Inhibitory Activity
4.7.3 Antimicrobial Properties
4.7.4 Antidiabetic Properties
4.8 Conclusion
References
Chapter 5 Consumer Attitudes towards Insects as Food
5.1 Introduction
5.2 Traditions and Culture
5.3 Psychological Barriers and the Yuck Factor
5.4 Sensory and Palatability Aspects
5.5 Current Trends in the Use of Insects as Food
5.6 Conclusions
References
Chapter 6 Edible Insect Farming
6.1 Introduction
6.1.1 Entomophagy: Wild Harvesting to Insect Farming
6.2 Mini-Livestock: An Advantageous Farming Choice
6.2.1 Environmental Aspects
6.2.1.1 Life cycle assessment
6.2.1.2 Footprints of land and water
6.2.1.3 Greenhouse gas emissions
6.2.1.4 Minimize pesticides
6.2.1.5 Biodegradation of waste material
6.2.1.6 Resource inventory for insect farming
6.2.1.7 Feeding requirements
6.2.1.8 Energy consumption
6.2.1.9 Edible insects services for the agriculture system
6.2.1.10 Feed conversion ratio
6.2.1.11 Economical aspects
6.2.1.12 Transport
6.3 Types of Insect Farming
6.3.1 Traditional Insect Farming
6.3.2 Indoor Insect Farming
6.3.2.1 Mulberry silkworm
6.3.2.2 House cricket
6.3.2.3 Yellow mealworm
6.3.2.4 Black soldier fly (BSF)
6.3.2.5 Housefly (HF)
6.3.3 Outdoor Insect Farming
6.3.3.1 Grasshoppers
6.3.3.2 Palm weevil or Sago larvae
6.3.3.3 Bamboo caterpillar
6.3.3.4 Weaver ants
6.4 Cost of Cultivation
6.5 Challenges and Way Forward
Acknowledgment
References
Chapter 7 Startups
7.1 Introduction
7.2 Case Studies
7.2.1 Food-Based Startups
7.2.1.1 Mighty Cricket
7.2.1.2 Illegal Oats
7.2.1.3 Jiminy’s
7.2.2 Technology-Based Startups
7.2.2.1 Aspire Food Group
7.2.2.2 BeoBia (The Bug Factory)
7.2.2.3 FarmInsect
7.3 Insect-Focused Foodtech Startups in Europe
7.3.1 Ÿnsect
7.3.2 Ÿnsect Human Nutrition & Health (the Dutch Food Branch of Ÿnsect)
7.3.3 Innovafeed
7.3.4 Protix Biosystems
7.3.5 nextProtein
7.3.6 Nextalim
7.3.7 Nasekomo
7.3.8 Hexafly
7.3.9 Entocycle
7.3.10 Hargol FoodTech
7.3.11 BetaHatch
7.3.12 Grubbly Farm
7.3.13 Plento
7.3.14 Insectta
7.3.15 Protenga
7.3.16 Bugsolutely
7.3.17 Magalarva
7.3.18 Entobel
Further Reading
Chapter 8 Mass Production Technologies
8.1 Introduction
8.2 Characteristics of Insects for Automated Rearing
8.2.1 Distinct Features of Insect Farming
8.2.2 Different Species of Insects for Food and Feed
8.2.3 Other Applications
8.3 General Methodology for Mass Production
8.3.1 Feed Principles for the Mass Production of Insects
8.3.1.1 Solid-feed
8.3.1.2 Semisolid feed
8.3.1.3 Liquid feed
8.4 Feed and Nutritional Requirements for Insect Rearing
8.4.1 Macronutrients
8.4.2 Micronutrients
8.4.3 Plant Material
8.4.4 Laboratory Diet
8.5 Equipment and Mechanization for Insect Mass Rearing
8.5.1 Production and Operation Management
8.5.2 Rearing Area
8.5.3 Feeding and Watering
8.5.4 Separation and Sorting
8.5.5 Cleaning Room
8.5.6 Dung Area
8.6 Production and Processing Technologies by Species
8.6.1 Black Soldier Fly
8.6.1.1 Adult colony and its management
8.6.1.2 Mating and oviposition
8.6.1.3 Production of larvae and its maintenance
8.6.1.4 The feed used for rearing
8.6.1.5 Costs and quality maintenance
8.6.1.6 Impact of different factors on the growth of BSFL
8.6.2 Crickets
8.6.2.1 Production of crickets worldwide
8.6.2.2 Rearing units
8.6.2.3 Diets and feeds
8.6.2.4 Environmental conditions
8.6.2.5 Reproduction
8.6.3 Mealworm
8.6.3.1 Rearing
8.6.3.2 Feed
8.6.4 Housefly
8.6.4.1 The feed and its maintenance in rearing
8.6.4.2 Process of production
8.6.5 Waxworm
8.6.5.1 Development of larvae and their diet maintenance
8.6.5.2 Rearing and reproduction
8.7 Environmental Control and Conditions
8.7.1 Physical Factors
8.7.1.1 Light, temperature and location
8.7.2 Mechanical factors
8.7.2.1 Filtration system
8.7.2.2 Panels and pads for evaporation
8.7.2.3 Humidifiers
8.8 Basic Needs for the Supply Chain System
8.8.1 Feed
8.8.2 Farms and Farmed Species
8.8.3 Transportation, Storage, and Distribution
8.8.4 Processing and Manufacturing Infrastructure
8.9 Challenges
Acknowledgement
References
Chapter 9 Insect Farming for Feed: Case Study
9.1 Introduction
9.2 Strategy of the Company
9.2.1 Modular Approach
9.2.2 Local Approach
9.2.3 Energetic Approach
9.2.4 Contract Approach
9.3 The Black Soldier Fly
9.3.1 Biology
9.3.2 Why Black Soldier Fly?
9.4 Products and Services of BEF Biosystems
9.5 Insect Farming
9.5.1 Side Characteristics
9.5.2 Feeding System
9.6 Insect Mass Production Technologies
9.6.1 Cages for Reproduction
9.6.2 Nursery
9.6.3 Fattening System
9.7 Environmental Impact of Our Plant
9.8 Strategies of Industry Marketing
9.9 Investments in the Insect Sector
9.10 Future Trends in Insect Products
9.11 Future Prospects for BEF Biosystems
9.12 Conclusion
References
Chapter 10 Insect Farming for Food: Case Study Company – Horizon Insects
10.1 Company Overview
10.1.1 Site and Facilities
10.1.1.1 Farming shed
10.1.1.2 Equipment
10.2 Primary Production
10.2.1 The Tenebrio molitor Lifecycle
10.2.2 The Production Cycle
10.2.3 Feed and Substrate
10.2.4 Pest Control
10.2.5 Good Practices and Regulation
10.2.6 Initial and Ongoing Costs
10.3 Additional Revenue Streams
10.3.1 Insect Frass
10.3.2 Outreach Events
10.3.3 Farm Tours and Cookery Classes
10.3.4 “Grow Your Own Mealworms” Kits
10.4 How Much Can Edible Insect Farming Yield?
10.5 Challenges and Future Plans
Chapter 11 Food Safety and Allergies
11.1 Introduction
11.2 Safety Considerations
11.2.1 Microbial Safety
11.2.2 Mould and Mycotoxins
11.2.3 Parasites
11.2.4 Allergic Reactions to Edible Insects
11.2.5 Toxicity
11.2.6 Contamination with Heavy Metals and Organic Pollutions
11.2.7 Allergens
11.2.8 Pesticide Residues
11.2.9 Mycotoxins
11.2.10 Hazard Analysis Critical Control Point for the Edible Insect Industry
11.2.11 Environment
11.2.12 Waste Management
11.2.13 Storage and Transportation
11.2.14 Recording
11.2.15 Food Safety Characterization of Insects
11.2.16 Bacteria
11.2.17 Endospore-Forming Bacteria
11.2.18 Viruses
11.2.19 Fungi
11.2.20 Yeasts and Moulds
11.2.21 Chemical Hazards
11.2.22 Toxic Metals
11.2.23 Other Chemical Contaminants from Production and Processing
11.2.24 Allergenic Potential
11.2.25 Anti-Nutritional Factors in Insects
11.3 Toxicological Hazards
11.4 Labelling and Health Claims
11.5 Concluding remarks
References
Chapter 12 Subsequent Processing of Insects
12.1 Introduction
12.2 Processing Steps
12.2.1 Starvation
12.2.2 Killing
12.2.3 Drying Methods
12.2.4 Milling and Fractionation
12.2.5 Dry Processing Technologies
12.2.6 Wet Processing Technologies
12.2.7 Protein Solubilization and Recovery
12.3 Post-Processing Handling of Food
12.4 Whole Insects as Culinary Ingredients
12.5 Insect Powder
12.6 Special Ingredients
12.7 Outlook
References
Chapter 13 Storage and Packaging of Edible Insects
13.1 Introduction
13.1.1 Basic Principles of Storage
13.1.2 Basic Principles of Packaging
13.1.3 Different Types of Packaging
13.2 Storage of Whole Insects
13.2.1 Fresh (Frozen and Chilled) Insects
13.2.2 Dried Insects
13.3 Storage of Paste and Powder Derived from Insects
13.3.1 Insect Paste
13.3.2 Insect Powder
13.4 Storage of Insect Fractions
13.4.1 Protein Powder
13.4.2 Insect Oil
13.5 Packaging and Storage of Insect-Based Food Products
13.5.1 Materials Used for Packaging
13.5.2 Examples of Packed Insect-Based Food Products
13.6 Further Challenges and Perspectives
References
Chapter 14 Market Potential and Statistics on Current Insect Consumption as Food
14.1 Background
14.1.1 Ecological Issues Involved In Rearing Edible Insects
14.2 Importance of Insect Consumption
14.2.1 Ento-Technology
14.2.2 Consumer Acceptance
14.2.3 Food Welfare and Security
14.2.4 Importance of Insects as Nutrimental Source
14.2.4.1 Role of insects in improving human gut microbiota
14.3 Market Potential of Edible Insects
14.3.1 Production of Edible Insects
14.3.2 Processing of Edible Insects
14.3.3 Estimation and Forecast Parameters of the Market
14.3.3.1 Cash income
14.3.3.2 Enterprise development
14.4 Marketing Strategies
14.5 Statistical Analysis of Insect Consumption
14.5.1 Asia
14.5.2 Europe
14.5.3 United States
14.5.4 Africa
14.6 Conclusion
Acknowledgement
References
Chapter 15 Example of Business Plan for Producers
15.1 Introduction
15.2 Your Business Plan’s Seven Building Blocks
15.2.1 Building Block 1: The Executive Summary
15.2.2 Building Block 2 – The Business idea
15.2.2.1 Description of the product or service
15.2.2.2 Customer advantages or benefits
15.2.2.3 Status of development of the products and services
15.2.3 Building Block 3 – Market and Competitor Analysis
15.2.3.1 Sector
15.2.3.2 Customer segment
15.2.3.3 Competitors
15.2.4 Building Block 4 – Marketing and Distribution
15.2.4.1 Product
15.2.4.2 Promotion
15.2.4.3 Distribution concept (Place)
15.2.4.4 Price
15.2.5 Building Block 5 – Business System
15.2.5.1 Preconditions for the business’ establishment/production’s requirements
15.2.6 Building Block 6 – Enterprise Management
15.2.6.1 Commercial know-how
15.2.6.2 Organization of the enterprise
15.2.6.3 The enterprise’s legal form
15.2.7 Building Block 7 – Financial Planning
15.2.7.1 Creating a Milestone Plan
15.2.7.2 Sources of financing (Capital requirement)
Notes
References
Chapter 16 Future Challenges for a Sustainable Edible Insect Industry
16.1 Introduction
16.2 A Case Study: Tebrito AB Edible Insect Company in Sweden
16.3 Future Challenges for the Edible Insect Industry
16.3.1 Sustainability, Environmental and Biodiversity Issues
16.3.1.1 Sustainability claims
16.3.1.2 Climate change and edible insect biodiversity
16.3.1.3 “Monoculture”: Mono-rearing of a few edible insect species and biodiversity
16.3.1.4 Diseases/nutrition programs similar to plant breed and animal breeding (e.g., bees)
16.3.2 Legislation, Safety and Nutritional Issues
16.3.2.1 International protocols and accepted limits for insect food and feed-analysis
16.3.2.2 Population nutritional studies in various countries
16.3.2.3 HACCP: safety control
16.3.2.4 Legislation: different approaches
16.3.3 Industrial Production Issues
16.3.3.1 Scaling up edible insect industrial production
16.3.3.2 Systematized artificial diets
16.3.3.3 Use of by-products as feeding substrate
16.3.4 Market Issues
16.3.4.1 Consumer acceptance
16.3.5 Bioethics and Insect Biotechnology
16.3.5.1 Molecular biology/insect biotechnology/genetically modified insects
16.3.5.2 Animal ethics, bioethics, biosecurity
16.4 Conclusions and Proposals
References
Index

Citation preview

Edible Insects Processing for Food and Feed Consumers around the world are becoming increasingly aware of the significant impacts of food consumption on the environment, and demand for more sustainable foods is expanding rapidly. Edible Insects Processing for Food and Feed: From Startups to Mass Production focuses on the growing topic of insects as food and feed, covering not only production elements, but also case studies and several other areas of interest, such as environmental aspects, nutritional value, consumers, food safety and market statistics. Key Features: • Includes several case studies and latest advancements in the area • Contains multidisciplinary approach, covering farm-to-fork aspects • Contains full account of contemporary developments in mass production of edible insects Written by passionate leading academics and industry partners around the globe, this book aims to bring together the latest advancements in edible insect production in a dynamic, modern and multidisciplinary approach. It is a one-stop shop that gives readers a flavour of where the fascinating topic of edible insect production is now, but more importantly of where it might be heading to in the future, showcasing several related challenges and opportunities.

Edible Insects Processing for Food and Feed From Startups to Mass Production

Edited by

Simona Grasso and Matteo Bordiga

First edition published 2023 by CRC Press 6000 Broken Sound Parkway NW, Suite 300, Boca Raton, FL 33487-2742 and by CRC Press 4 Park Square, Milton Park, Abingdon, Oxon, OX14 4RN CRC Press is an imprint of Taylor & Francis Group, LLC © 2023 selection and editorial matter, Simona Grasso and Matteo Bordiga; individual chapters, the contributors Reasonable efforts have been made to publish reliable data and information, but the author and publisher cannot assume responsibility for the validity of all materials or the consequences of their use. The authors and publishers have attempted to trace the copyright holders of all material reproduced in this publication and apologize to copyright holders if permission to publish in this form has not been obtained. If any copyright material has not been acknowledged please write and let us know so we may rectify in any future reprint. Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, reproduced, transmitted, or utilized in any form by any electronic, mechanical, or other means, now known or hereafter invented, including photocopying, microfilming, and recording, or in any information storage or retrieval system, without written permission from the publishers. For permission to photocopy or use material electronically from this work, access www​.copyright​.com or contact the Copyright Clearance Center, Inc. (CCC), 222 Rosewood Drive, Danvers, MA 01923, 978-750-8400. For works that are not available on CCC please contact mpkbookspermissions​@tandf​.co​​.uk Trademark notice: Product or corporate names may be trademarks or registered trademarks and are used only for identification and explanation without intent to infringe.

Library of Congress Cataloging‑in‑Publication Data Names: Grasso, Simona, editor. | Bordiga, Matteo, editor. Title: Edible insects processing for food and feed : from startups to mass production / edited by Simona Grasso and Matteo Bordiga. Description: First edition. | Boca Raton : CRC Press, 2023. | Includes bibliographical references and index. Identifiers: LCCN 2022050237 (print) | LCCN 2022050238 (ebook) | ISBN 9780367746940 (hbk) | ISBN 9780367761516 (pbk) | ISBN 9781003165729 (ebk) Subjects: LCSH: Edible insects. | Entomophagy. | Animal feeding. | Insect rearing. | Feed industry. Classification: LCC TX388.I5 E35 2023 (print) | LCC TX388.I5 (ebook) | DDC 641.3/9--dc23/eng/20221122 LC record available at https://lccn.loc.gov/2022050237 LC ebook record available at https://lccn.loc.gov/2022050238 ISBN: 978-0-367-74694-0 (hbk) ISBN: 978-0-367-76151-6 (pbk) ISBN: 978-1-003-16572-9 (ebk) DOI: 10.1201/9781003165729 Typeset in Times by Deanta Global Publishing Services, Chennai, India

Contents Preface vii About the Editors ix List of Contributors xi 1 The Role of Insects for Nature and Humans Liqiong Niu, Xinquan Yang, Ronghai He, and Yuan Liu

1

2 Environmental Impact of Edible Insect Processing Sergiy Smetana, Dusan Ristic, and Volker Heinz

21

3 Legislation Nils Th. Grabowski

33

4 Nutrient Content and Functionalities of Edible Insects Ruchita Rao Kavle, Ellie Pritchard, Alaa El-Din Ahmed Bekhit, Alan Carne, and Dominic Agyei

53

5 Consumer Attitudes towards Insects as Food Marina Carcea, Valentina Narducci, and Valeria Turfani

85

6 Edible Insect Farming Rimsha Naseem, Waqar Majeed, Mian Muhammad Awais, Muhammad Noman Naseem, Naureen Rana, and Uzma Ramzan

107

7 Startups Simona Grasso and Matteo Bordiga

133

8 Mass Production Technologies Waqar Majeed, Rimsha Naseem, Naureen Rana, Hammad Ahmad Khan, Uzma Ramzan, Sobia Kanwal, Elmo Borges de Azevedo Koch, Nazia Ehsan, Muhammad Sarfraz Ahmed, and Muhammad Naveed

155

9 Insect Farming for Feed: Case Study Marco Meneguz and Sihem Dabbou

193

10 Insect Farming for Food: Case Study Company – Horizon Insects Laurence Mohan

211

11 Food Safety and Allergies Dele Raheem, Fernando Ramos, António Raposo, Ariana Saraiva, Oluwatoyin Bolanle Oluwole, and Conrado Carrascosa

225

12 Subsequent Processing of Insects Dusan Ristic, Sergiy Smetana, and Volker Heinz

247 v

vi Contents 13 Storage and Packaging of Edible Insects Giacomo Rossi, Shikha Ojha, Namrata Pathak, Pramod Mahajan, and Oliver K. Schlüter

261

14 Market Potential and Statistics on Current Insect Consumption as Food Umm E Ummara, Aqsa Riaz, Waqar Majeed, Sobia Kanwal, Ayesha Parveen, Tehrim Liaqat, Kaynaat Akbar, Iffa Maryam, Aqsa Shareef, and Uzma Ramzan

277

15 Example of Business Plan for Producers Ileana Maricruz Bermúdez-Serrano, Rodrigo Llauradó, and Utz Dornberger

305

16 Future Challenges for a Sustainable Edible Insect Industry Vassileios Varelas

329

Index 347

Preface The vision that inspired the idea of this book was to create a source of information on the timely topic of innovations and advancements in the area of insect processing. Over the years, several businesses and startups have been trying to improve the production of insects for both food and feed to create large-scale production systems that would be of interest to the Western world. Several aspects that require improvement include insect quality, the safety of processing and transportation and storage issues, as well as marketing. Consumers worldwide are becoming increasingly aware of the significant impacts of food consumption on the environment. In this context, the demand for more sustainable foods is expanding rapidly. Insect-producing businesses are innovating advanced processing and scaling up technologies to meet this emerging consumer demand. Many consumers in Western countries are opening up to the concept of insects as food, considering their nutritional quality and sustainable nature. Numerous research efforts have been dedicated to exploring the use of insects as food in the Western world, identifying the best strategies to overcome psychological barriers and the “yuck” factor associated with insect consumption. Several regulatory challenges also remain in Western countries. This book is a complete account of contemporary developments in the area of the mass production of edible insects, and is aimed at interested readers as a valuable resource on this contemporary topic. This book is valuable reading material for students (particularly at post-graduate/graduate level or higher), research scholars, teachers, scientists, industry professionals, entrepreneurs and all others interested in the area of edible insects.

vii

About the Editors Dr. Simona Grasso is an assistant professor at University College Dublin, Ireland. Before taking on this role she was a senior research fellow at the University of Reading, School of Agriculture, Policy and Development, UK. Simona received her PhD in Food Science from University College Dublin, Ireland and obtained a first class honours BSc and MSc in Food Science from the University of Catania, Italy. She has experience working for the food industry covering a variety of roles both in the UK and in the Republic of Ireland. Her main research activity concerns new product development, functional foods and healthier meat products. She is also interested in consumer and sensory science, as well as the valorization of food byproducts and the development of upcycled foods. She has more than 10 years of experience in food science, and in her academic career she has published more than 30 research papers and literature reviews in peer-reviewed international journals. Dr. Matteo Bordiga is an assistant professor of Food Chemistry at the Department of Pharmaceutical Sciences, Università del Piemonte Orientale (UPO), Novara, Italy. He received his PhD in Food Science and his MS in Chemistry and Pharmaceutical Technologies from the same university. The main research activity of Dr. Bordiga concerned food chemistry, investigating the different classes of polyphenols from analytical, technological and nutritional points of view. More recently, his research interests have included wine chemistry, focusing on the entire production process – from vine to glass. He has published more than 50 research papers in peer-reviewed international journals. He is the editor of LWT – Food Science and Technology (Elsevier), associate editor of Food Science & Nutrition (Wiley), and editorial board member of International Journal of Food Science & Technology; Future Foods. He is also the editor of the books Valorization of Wine Making By-Products – CRC Press | Taylor & Francis Group (2016); Post-Fermentation and Distillation Technology: Stabilization, Aging, and Spoilage – CRC Press | Taylor & Francis Group (2018); Food Aroma Evolution During Food Processing, Cooking, and Aging – CRC Press | Taylor & Francis Group (2019); and Concise Encyclopedia of Science and Technology of Wine – CRC Press | Taylor & Francis Group (2021).

ix

List of Contributors Dominic Agyei Department of Food Science University of Otago Dunedin, New Zealand

Alan Carne Department of Biochemistry University of Otago Dunedin, New Zealand

Muhammad Sarfraz Ahmed Department of Zoology, Wildlife and Fisheries University of Agriculture Faisalabad, Pakistan

Conrado Carrascosa Department of Animal Pathology and Production, Bromatology and Food Technology Universidad de Las Palmas de Gran Canaria Arucas, Spain

Kaynaat Akbar Department of Zoology, Wildlife and Fisheries University of Agriculture Faisalabad, Pakistan Mian Muhammad Awais Department of Pathobiology Bahauddin Zakariya University Multan, Pakistan Alaa El-Din Ahmed Bekhit Department of Food Science University of Otago Dunedin, New Zealand Ileana Maricruz Bermúdez-Serrano SEPT Competence Center Leipzig University Leipzig, Germany Matteo Bordiga Department of Pharmaceutical Sciences Università degli Studi del Piemonte Orientale Novara, Italy Marina Carcea Research Centre for Food and Nutrition Council for Agricultural Research and Economics (CREA) Rome, Italy

Sihem Dabbou Center Agriculture Food Environment (C3A) University of Trento San Michele All’adige (TN), Italy Utz Dornberger SEPT Competence Center Leipzig University Leipzig, Germany Nazia Ehsan Department of Zoology, Wildlife and Fisheries University of Agriculture Faisalabad, Pakistan Nils Th. Grabowski Institut für Lebensmittelqualität und -sicherheit (LMQS) Stiftung Tierärztliche Hochschule Hannover Hannover, Germany Simona Grasso School of Agriculture and Food Science University College Dublin Dublin, Ireland Ronghai He School of Food and Biological Engineering Jiangsu University Zhenjiang, China

xi

xii List of Contributors Volker Heinz German Institute of Food Technologies Quakenbrueck, Germany

Marco Meneguz BEF Biosystems Company Torino, Italy

Sobia Kanwal Department of Zoology, Wildlife and Fisheries University of Agriculture Faisalabad, Pakistan

Laurence Mohan Horizon Insects Ltd London, England

Ruchita Rao Kavle Department of Food Science University of Otago Dunedin, New Zealand

Valentina Narducci Research Centre for Food and Nutrition Council for Agricultural Research and Economics (CREA) Rome, Italy

Hammad Ahmad Khan Department of Zoology, Wildlife and Fisheries University of Agriculture Faisalabad, Pakistan

Rimsha Naseem Department of Zoology, Wildlife and Fisheries University of Agriculture Faisalabad, Pakistan

Elmo Borges de Azevedo Koch Department of Ecology Federal University of Rio Grande do Norte Natal-RN, Brazil

Muhammad Noman Naseem Center of Animal Sciences QAAFI University of Queensland St Lucia, Australia

Tehrim Liaqat Department of Zoology, Wildlife and Fisheries University of Agriculture Faisalabad, Pakistan Yuan Liu Department of Food Science & Technology School of Agriculture and Biology Shanghai Jiao Tong University Shanghai, China

Muhammad Naveed Department of Zoology, Wildlife and Fisheries University of Agriculture Faisalabad, Pakistan Liqiong Niu School of Life Sciences Guangzhou University Guangzhou, China

Rodrigo Llauradó Chepulines: Cocina Ento Buenos Aires, Argentina

Shikha Ojha Department of Systems Process Engineering Leibniz Institute for Agricultural Engineering and Bioeconomy (ATB) Potsdam, Germany

Pramod Mahajan Department of Systems Process Engineering Leibniz Institute for Agricultural Engineering and Bioeconomy (ATB) Potsdam, Germany

Oluwatoyin Bolanle Oluwole Department of Food Technology Federal Institute of Industrial Research, Oshodi Lagos, Nigeria

Waqar Majeed Department of Zoology, Wildlife and Fisheries University of Agriculture Faisalabad, Pakistan Iffa Maryam Department of Zoology, Wildlife and Fisheries University of Agriculture Faisalabad, Pakistan

Ayesha Parveen Department of Zoology, Wildlife and Fisheries University of Agriculture Faisalabad, Pakistan Namrata Pathak Department of Systems Process Engineering Leibniz Institute for Agricultural Engineering and Bioeconomy (ATB) Potsdam, Germany

List of Contributors  xiii Ellie Pritchard Department of Food Science University of Otago Dunedin, New Zealand Dele Raheem Arctic Centre University of Lapland Rovaniemi, Finland Fernando Ramos Faculty of Pharmacy University of Coimbra Coimbra, Portugal Uzma Ramzan Institute of Zoology University of the Punjab Lahore, Pakistan Naureen Rana Department of Zoology, Wildlife and Fisheries University of Agriculture Faisalabad, Pakistan António Raposo Research Center for Biosciences and Health Technologies Universidade Lusófona de Humanidades e Tecnologias Lisboa, Portugal Aqsa Riaz Department of Zoology, Wildlife and Fisheries University of Agriculture Faisalabad, Pakistan Dusan Ristic German Institute of Food Technologies Quakenbrueck, Germany and Institute of Food Technology University of Natural Resources and Life Sciences (BOKU) Vienna, Austria Giacomo Rossi Department of Systems Process Engineering Leibniz Institute for Agricultural Engineering and Bioeconomy (ATB) Potsdam, Germany

Ariana Saraiva Department of Animal Pathology and Production Universidad de Las Palmas de Gran Canaria Arucas, Spain Oliver K. Schlüter Department of Systems Process Engineering Leibniz Institute for Agricultural Engineering and Bioeconomy (ATB) Potsdam, Germany Aqsa Shareef Department of Zoology, Wildlife and Fisheries University of Agriculture Faisalabad, Pakistan Sergiy Smetana German Institute of Food Technologies Quakenbrueck, Germany Valeria Turfani Research Centre for Food and Nutrition Council for Agricultural Research and Economics (CREA) Rome, Italy Umm E Ummara Department of Zoology, Wildlife and Fisheries University of Agriculture Faisalabad, Pakistan Vassileios Varelas Laboratory of Industrial Chemistry, Department of Chemistry University of Athens Athens, Greece Xinquan Yang School of Life Sciences Guangzhou University Guangzhou, China

The Role of Insects for Nature and Humans

1

Liqiong Niu, Xinquan Yang, Ronghai He, and Yuan Liu Contents 1.1 Introduction 1 1.1.1 Insects in Plant Reproduction 4 1.1.2 Insects in Waste Biodegradation 5 1.1.3 Insects in Controlling Harmful Pest Species 6 1.1.4 Insect Food in China: History and Current Status 8 1.1.4.1 Long history of insect food in China 9 1.1.4.2 Current insect food in China 10 1.2 Conclusion 15 1.3 Prospective and Challenges 15 References 16

1.1 INTRODUCTION The search for new alternative proteins that can be sustainably developed has become the focus of future research, as continuous population growth, the increasing demand on the food supply and aggravated global climate change have become increasingly prominent. Insect protein has received widespread attention. Edible insects are a promising industry, and consumption of insects is a solution that could solve the world’s food security problems and feed the growing population (Schiemer et al. 2018; Patel et al. 2019; Kim et al. 2019). As a new type of protein resource, it has the advantages of being high in protein, being highly nutritious, having organic elements and vitamins, being environmentally friendly, and allowing for sustainable development. Edible insects should be closely linked with the food industry, agriculture and pharmaceutical industries (Paoletti and Dreon 2005; Chen et al. 2009; Jideani and Netshiheni 2017). Studies have shown that the crude protein content of edible insects is higher than that of common meat, and they are also rich in vitamins and essential amino acids (Tang et al. 2019). It is estimated that at least 2 billion people in more than 113 countries worldwide, including Asia, Africa and South America, regularly consume insects (van Huis et al. 2013; Tao and Li 2018). Insects can be considered as one of the most ideal high-nutrient food materials in future foods. As the largest group of species on the planet, insects are almost ubiquitous. Insects have the characteristics of strong reproductive ability, a short life cycle, diverse feeding habits, disease resistance and strong adaptability. They have become the most prosperous species group in nature, and traces of insects may be found in any living environment (Feng and Chen 2016a; Kok and van Huis 2021). In addition, DOI: 10.1201/9781003165729-1

1

2  Edible Insects Processing for Food and Feed insects play a very important role in the ecosystem. According to their feeding characteristics, they can be divided into herbivorous, carnivorous, caustic insects and omnivorous insects. Insects can act as prey, predators and decomposers in the food chain, and they perform very complex functions and roles in maintaining ecological balance (Feng and Chen 2016a). In general, rearing insects has many advantages for the environment, such as high feed conversion efficiency – the ability to convert feed into the increased weight of the fed animal. Insects can reduce environmental pollution by the transformation and biodegradation of harmful waste, thereby increasing the value of bio-waste. Insects produce fewer greenhouse gases and less ammonia. Less water and agricultural land are required compared with traditional cattle rearing. Insects can be reared at a high density, using less resources in exchange for more food, and at the same time reducing the burden on the environment. Insect farming does not pose animal welfare issues. From a health aspect, insects could prevent or reduce the risk of spreading zoonotic diseases (van Huis et al. 2013; Doi and Mulia 2021; Kok and van Huis 2021; van Huis et al. 2021). In addition, insects as food also perform many functions that are beneficial to people’s livelihoods, such as improving the livelihoods of poor areas, and providing opportunities for entrepreneurship (FAO 2013; Doi and Mulia 2021). Insects have the potential to be a source of food and to feed the population on a global scale, and at the same time have a small negative effect on the environment (Govorushko 2019). Functional foods are playing an increasingly leading role in the 21st century, and the development of insect proteins has become a hot topic. In recent years, the number of published studies has shown an increasing tendency (Figure 1.1). According to surveys conducted by Wageningen University in the Netherlands, there are currently more than 2,000 insects being eaten by humans in the world, but most of them are collected in the wild (Jongema 2017; Tao and Li 2018; van Huis 2020). The common insect groups include beetles, caterpillars, bees, wasps, ants, grasshoppers, locusts, crickets, cicadas, and so on. In Asian countries with long history of traditional food culture, such as China and Thailand, residents often consume insect food including beetles (31%), bees and ants (14%), crickets (13%) etc. (van Huis et al. 2013). As a traditional food, edible insects contain high values of essential fatty acids, protein and mineral content (Rumpold and Schlüter 2013; Tao and Li 2018; de Carvalho et al. 2020). Edible insects have been 600

Research Article Agriculture Food Science and Technology

400

200

0 2020

2019

2018

2017

2016

2015 Year

2014

2013

2012

2011

2010

FIGURE 1.1  Number of published research articles from the web of science using “edible insects,” further refined by the “agriculture’ or “food science and technology” research fields.

1  •  The Role of Insects for Nature and Humans  3 shown to have high nutritional value and good sensory qualities and to be safe for human consumption (Kok and van Huis 2021). Insects also have the characteristics of fast reproduction, wide distribution and easy absorption. They can also be used as potential substitutes and feed sources for fish meal in feed (Barroso et al. 2014). Edible insects, as food materials that could be provided to humans and animals, have good potential for providing nutrition and allowing for sustainable and efficient development (Halloran et al. 2016; de Carvalho et al. 2020). Chinese residents have an extremely long history and relatively high degree of acceptance of insect consumption. Many people have been obsessed with insect food since ancient times (Luo et al. 2005; Feng et al. 2018). Since most people would not eat insects directly, various types of cooking methods, such as frying and baking, are used to process the raw material into cooked food; this produces a distinctive delicious dish and unique flavor, and edible insects have formed part of a special food culture (Chen et al. 2009). According to historical records, Chinese people began consuming insects more than 3,000 years ago (Feng et al. 2020). People often describe the sensation of eating fried grasshoppers as “crispy texture, chicken taste”, and many people like to eat this type of insect food, which has a charred smell and sweet taste (Tang et al. 2019). This dietary culture has been continuously developed and enriched by people at different times. Now China has identified more than 178 species of commonly consumed edible insects, and restaurants in Yunnan province and other places serve insect dishes all year round (Chen et al. 2009). Also, Doudan and Jinchan cuisine are very famous and popular in Jiangsu Province. Insect foods have been shown to have a unique texture and flavor, consumer acceptance of insect consumption is relatively high, and many people like to try and eat various insect foods prepared using different methods. Moreover, the culture of edible insects is heavily influenced by different regions, showing a vigorous development. Driven by the economic and environmental value of edible insects, many companies have already begun to focus on the insect industry, and various type of products, such as insect dishes and insect snacks, have been developed (Figure 1.2). In the future, the inheritance and protection of insect food culture will also

FIGURE 1.2  Edible insect dishes and snack products sold in China. (Photo based on Internet websites.)

4  Edible Insects Processing for Food and Feed be particularly important (van Huis et al. 2013; Feng et al. 2018). In this chapter, we will discuss this topic in detail.

1.1.1 Insects in Plant Reproduction Insects are an ecological service provider that are fundamental to human survival and play an important role in plant reproduction. There are many insects in nature that can pollinate plants; a study has found that 98% of the pollinators for 100,000 plants are insects, and 78.5% of flowering plants in temperate zones require insect pollinators to reproduce successfully (Che 2018; Ingram et al. 1996; Mehrwar and Uniyal 2021). Plants rely less on non-biological pollination, such as wind and water; most pollination is performed by wild insects, including bees, flies, butterflies and moths. Bees also rely on flower resources such as pollen and nectar. There are already more than 20,000 insect species worldwide, of which about 5,000 species appear in the neotropical regions and are the main pollinators of wild and cultivated plants (Feng and Chen 2016a; Pires and Maués 2020). Insects also play an important role in the growth and reproduction of economic crops. With the rapid increase in the number of people in recent years, the area of farmland has been rapidly reduced, and the quality and quantity of edible food are at greater risk (Yi et al. 2010). If there were no insects to carry out pollination, food would be scarce and famine might even occur (Che 2018). Among 100 food crops in the world, more than 75% of them rely on insects for pollination. Domesticated bees can pollinate 15% of the plants in these species, and more than 80% of plants are pollinated by wild bees and other wild insects (Ingram et al. 1996). Insect pollination could bring huge economic benefits to the production of food, fruit trees and vegetables globally (Yi et al. 2010). In modern agriculture, scientists use artificial rearing drones to pollinate vegetables in protected areas, which could improve the quality and yield of crops and save a lot of labor. In addition, it also solves the environmental pollution caused by pollination with chemicals and the product safety problems caused by pollination (Che 2018). As bees provide ecological services for global food production, a decline in their number would pose a huge threat to global food production (McConnell and Burger 2020). Insects could visit and pollinate plants, strengthen the gene exchange and interaction and exchange of self-pollinated plants, adapt to the changing environment, improve the quality and quantity of self-pollinated plant seeds, and increase the yield of crops, fruits and vegetables, which enable humans to obtain more food. Insects, an ecosystem service provider, are very important to agriculture and nature (Che 2018; Feng and Chen 2016a; van Huis et al. 2013). Most insects are directly or indirectly beneficial to humans. Insects can produce natural products for humans, such as honey (bees), silk (silk moths), dyes and shellac (scaled insects) and tannic acid and ink (insect galls) (Weinzierl and Henn 2020). In particular, bees can pollinate plants, and they can also provide large amounts of honey, royal jelly, propolis, beeswax and pollen. As we know, honey is the raw material for making cakes, beverages, medicines, functional foods and other products, and bee venom can also be used to treat diseases such as rheumatoid arthritis (Che 2018). The aspects of insects that are directly beneficial to humans include insect predators and pest parasites (Weinzierl and Henn 2020). By collecting insects, crops and plants could be reproductive (van Huis et al. 2013). Many insects have the ability to reorient the development of plants to form galls, which could provide insects with unique and enhance plants and protect them from enemies and other substances (Schultz et al. 2019). These actions could better promote the growth and reproduction of plants. In addition, fly larvae can use a variety of ways to reduce the content of important elements (N, P, K, and C) to 40%–60%. These products could be used as compost, fertilizer, or soil remediation materials to promote plant growth and reproduction (Petkova 2019). Ordinary forest managers have little knowledge about sustainable management and insect capture. There is also little experience in regulating forest vegetation or practices for maximizing and maintaining insect population. On the contrary, some insects cause huge damage and even death to commercial trees, and as a result, many forest managers regard all insects as potentially destructive pests (van Huis et al. 2013).

1  •  The Role of Insects for Nature and Humans  5 For a long time, forest managers have regarded caterpillars as pests, as they feed on fresh leaves and are therefore considered to be harmful to trees. In fact, trees can grow more leaves and edible insect species could be protected through a forest management system (van Huis et al. 2013; van Huis and Oonincx 2017). Edible insects such as grasshoppers, weaver ants and giant water bugs are usually caught in the wild (Hanboonsong et al. 2013). By capturing pests such as grasshoppers and bamboo caterpillars, crops could be protected and pests prevented even without using pesticides (Lamsal et al. 2019). It should be noted that many edible insects are phytophagous pests, which may be parasitic on plants. Within a certain range and time, and only if collected from the wild, the density of insects can be reduced, which is conducive to the growth and reproduction of agricultural and forestry plants. If such insects are cultivated artificially, appropriate management measures should be taken and fully considered, and methods such as rearing insects in an artificial closed environment and artificial feed rearing could reduce the harmfulness of insects to plants (Feng 2016a).

1.1.2 Insects in Waste Biodegradation Every year, about one-third of the food produced in the world will be discarded. Only a small part of this waste is used for biofuel and compost production, and most of the waste will be landfilled, causing further damage and pollution to the environment. With the rapid growth of the global population and the emergence of a large demand for animal protein, large-scale production of edible insects for human consumption and livestock feed seems to be a sustainable solution (Varelas 2019). Reducing food waste and loss is in line with the UN’s sustainable development goal of reducing food waste, achieving the goal of eliminating hunger, improving nutrition and promoting sustainable agricultural development (Desa 2016). Insects play an important role in the cycle of the ecosystem. Caustic insects are important components of the decomposers in the ecosystem. Caustic insects can further be divided according to their living environment: those that feed on dead organic matter, plant roots, rotting plants, etc.; those that feed on animal feces (which contains large amounts of organic compounds); and those that feed on dead animals or trees in the forest. Insects clean up the environment and promote nutrient circulation in the ecological cycle. In the forest ecosystem, insects help decompose animal and plant organisms, decompose dead wood, and accelerate circulation and energy flow in the forest; in the treatment of urban garbage, insects can clean up corrupt organic compounds, decompose and clean the environment for human living; and insect activity can change the soil structure, spread microorganisms and disperse animal feces, which is conducive to the survival and reproduction of other organisms (Feng and Chen 2016a). Some insects perform the function of removing garbage and manure, cleaning our environment. For example, at first, there were no cattle in Australia. Once Europeans introduced cattle to Australia, the cattle quickly reproduced everywhere, which eventually caused a large accumulation of cow manure and caused a serious environmental pollution problem for society. In the 1980s, Australians introduced the dung beetle from China. Dung beetles were able to eat the manure and bury them it the soil, thus successfully solving this environmental problem (Che 2018). Rearing insects has been shown to have many advantages for the environment and to have high feed conversion efficiency (the ability of an animal to convert feed to its increased weight, expressed as the feed per kg required to increase 1 kg of body weight). For example, when feeding crickets, an average of only 2 kg of feed is needed to gain 1 kg of body weight (van Huis et al. 2013). In addition, insects can reduce environmental pollution by biodegrading harmful waste, thereby increasing the value of bio-waste. Fewer greenhouse gases and relatively little ammonia are emitted by insects. Insects require less water and agricultural land compared with traditional cattle rearing (van Huis et al. 2013). Fly larvae could reduce the water content and important elements (N, P, K, and C) in waste, reducing the weight of organic waste by 60% in 10 days as a result (Petkova 2019). Insects can be sustainably reared on organic waste (for example, manure, pig slurry and compost). It has been reported that the black soldier fly (Hermetica illucens), housefly (Musca domestica), and yellow mealworm (Tenebrio molitor) can effectively bio-transform organic wastes. In addition, it is reported that

6  Edible Insects Processing for Food and Feed some coleopteran insect larvae show the ability to rapidly biodegrade plastics, and the larvae of mealworms in particular degrade plastics quickly. Mealworm larvae could synergize with their intestinal flora to digest and biodegrade polystyrene (PS), polyethylene (PE), polypropylene (PP) and polyvinyl chloride (PVC) in the intestines in the body (Wu and Criddle 2021). The above insects could bio-transform 1.3 billion tons of organic waste every year (van Zanten et al. 2015; Veldkamp et al. 2012; Baiano 2020). The black soldier fly has also been found to reduce the amount of manure accumulation and to convert and decompose biological waste (Tingle et al. 1975; Sheppard et al. 1994). Interestingly, publications before the 1960s indicated that the black soldier fly was considered purely as a pest (Tomberlin and van Huis 2020). Subsequent studies have shown that black soldier fly larvae can rapidly degrade municipal solid waste into biofuels and related by-products at a low cost (Kumar et al. 2018). In addition, it has been reported that some larvae can biodegrade plastics, and the larvae of yellow mealworm in particular degrade plastics very fast (Wu and Criddle 2021). The study found that the use of insect flies to biodegrade livestock and poultry manure does not produce any secondary waste or secondary environmental pollution, making it a low-cost, energy-saving and efficient method. It could not only use as resources in livestock manure and reduce manure pollution, but also in rearing maggots. Improved the livestock and poultry production environment and reduces environmental pollution could be obtained environmentally friendly. Finally, maggots could also be used as animal protein feed and organic fertilizer to promote the healthy development of ecological recycling in animal husbandry (Huang et al. 2013). It is estimated that in the next 30–40 years, many waste products will inevitably be generated due to large-scale livestock breeding. However, insects could consume these waste products as food and convert them into high value-added materials such as high levels of protein. Biodegradation of organic waste is also an important reason for using insect protein in animal feed (Petkova 2019). It should be noted that although there have been studies that have proved the feasibility of insects using food waste for sustainable food and feed production, this function is currently subject to some legal restrictions and mainly depends on consumer acceptance. The main strategy to solve consumer-related problems is to disguise insects as products that consumers are familiar with and make the products taste good (van Huis 2020). In the future, it may be necessary to develop strategies to promote consumer acceptance of edible insects as raw materials for food and feed (Rumpold and Langen 2020). These wastes provide suitable untapped resources for protein production. Insects not only have the potential to extract protein from waste, but they also help to reduce the amount of waste significantly, including for use as compost, fertilizer, soil remediation materials and raw materials for biogas production. Taking into account the recent changes in global fertilizer prices, the use of waste after insect farming as fertilizer is particularly attractive. For example, in the past 10 years, the price of some brands of fertilizer has fluctuated by more than 400%. In order to ensure the yield of the land, the demand for highquality organic fertilizers will continue to increase in the future (Petkova 2019). It should be noted that when organic waste streams are used, food safety is not a major issue, but it requires careful monitoring (van Huis 2020). Therefore, insects could not only convert biological waste into high-quality protein that could be used in animal feed, but they also have the potential to obtain protein from waste, and provide solutions for meeting the growing demand for animal protein and protein production in the future. Also, insects provide suitable untapped resources. It is also possible to use food waste as a new substrate or dietary ingredient in the large-scale insect breeding process for human food and animal feed to improve the sustainable development of the food industry. In addition, insects could be very helpful in reducing the impact of climate change and food and feed production on the environment, and in significantly decomposing and reducing related waste (Van Huis et al. 2013; Varelas 2019; Petkova 2019).

1.1.3 Insects in Controlling Harmful Pest Species As mentioned here, many insects play many important roles in interactions between humans and their environment. Insects are decomposers, pollinators, predators and prey of organic matter in the ecosystem.

1  •  The Role of Insects for Nature and Humans  7 Insects can also be used as food for other animals, but they are also important natural enemies and pests in agriculture and for forest plants (Feng et al. 2018). Pests are an important limiting factor for the sustainable development of agriculture. About half of insect species feed on plants. The offspring of insects are characterized by fast reproduction, large populations and short reproduction cycles. Insects could feed on crops; many insects also have a strong ability to fly, which poses a huge threat to crops. Some insects also carry pathogenic microorganisms and nematodes, which may cause serious crop diseases and adversely affect crops. Agricultural pests feed on crops such as grains, vegetables, fruits, etc., reducing crop yields until there is no harvest. This poses a huge and potential threat to the agricultural industry, which mankind depends on and which affects the survival of mankind (Feng and Chen 2016a). When the insect population erupts, it may destroy forests. Forest is the most important part of the earth ecosystem and an important aspect of human survival and reproduction. The outbreak of forest pests may cause huge damage to forest plants and even partially destroy the forest ecosystem. If insects are collected only from the wild and within a certain range and time, the density of insects would be reduced, which is beneficial to agricultural and forest plants (Feng and Chen 2016a). For example, the longhorn beetle is a resident pest that harms forest plants and can destroy forests in severe cases. However, the longhorn beetle is an edible insect with high nutritional value. In ethnic minority areas such as Yunnan Province, China, the custom of eating longhorn beetle larvae still endures. The larvae of longhorn beetle are generally fried or charred before being eaten, and it is common to eat raw larvae (Feng and Chen 2016b). Studies have found that the effect of chemical control of the longhorn beetle is not ideal. If they can be used as resource, manual catching or trapping can be used to effectively control pests, and they can also be made into commodities for sale, which will generate greater economic value (Feng and Chen 2016b). On the other hand, if such insects are cultivated artificially, their safety must be fully considered, and methods such as artificial closed-environment rearing and appropriate management measures and methods must be adopted to reduce the harmfulness of insects to plants (Feng and Chen 2016a). With the continuous development of global agricultural intensification, the control of natural pests has been identified as an ecological service (Wilby and Thomas 2002). Also, the prevention and control of natural pests or diseases play a vital role in maintaining the stability of the global agricultural system, national food security, the national income and rural household income (Naylor and Ehrlich 1997). Many edible insects are significant pests in the growth of crops, and spraying chemical insecticides is often used to control pests. By catching and eating insects, the use of chemical pesticides and other substances could be reduced, resulting in less residues in plant-derived foods, less chemical environmental pollution, and prevention of insect resistance (Imathiu 2020). Entomophagy is an ideal alternative to increasing food production and reducing the impact of crop pests. It has also been proven to be a successful method to control crop pests (van Huis 2018). The locust (Patanga succincta) is a kind of pest that seriously harms crops. There are reports of locust damage all over the world due to locust infestation. Due to the large population of locusts and their strong migratory ability, they pose a great threat to agricultural production (Feng and Chen 2016b). According to the Food and Agriculture Organization of the United Nations (FAO), the locust infestation that erupted in 2020 may have been caused by factors such as climate change, hurricanes, and heavy rain. Then this locust infestation affected 70,000 hectares of land in Somalia and Ethiopia (Peng et al. 2020). Chemical pesticides are usually used to kill locusts so as to achieve control over their population. Although this method could result in good control in the short term, chemical pesticide residues pollute food and the environment and harm human health (Feng and Zhao 2016; Peng et al. 2020). Locusts could be the kind of biological resource that could be used by humans. If artificial trapping methods can be used to collect locusts, it can not only achieve the effect of controlling them, but means they can also be used as a protein resource to benefit mankind. In China, locusts have brought great harm to Chinese agricultural production throughout history, and locusts are a favorite food for the population. Farmers often use wheat piles and straws to collect locusts in the fields. Fresh and frozen locusts are mostly sold

8  Edible Insects Processing for Food and Feed in markets. In recent years, locusts as food have become popular around the world. Locusts, as a characteristically high-protein food resource, have the characteristics of high added value and a good taste (Feng and Zhao 2016). In 1978 in Thailand, large numbers of locusts broke out in maize-growing farmland, and the government launched a campaign to promote the consumption of locusts. The government promoted the use of locusts as raw material for deep-fried snacks, the addition of ground locusts to biscuits, or the use of ground locusts in cooking sauces, and this promotion was a success (van Huis 2018). After that they were not considered as pests of crops, locusts became very popular. As the demand for insects as food continued to increase, some farmers began specializing in growing crops to breed them, resulting in a shift from wild collection to large-scale rearing. In addition, rearing insects has brought additional economic income to local farmers (Roffey 1979; Hanboonsong 2010). Moreover, harvesting and consuming insects would bring additional economic benefits (Cerritos 2009) and provide organic plant products. Some farms that require mechanical harvesting often occur stink bugs. In this case, on the one hand manually collecting stink bugs can protect crops, and on the other hand, the harvested stink bugs can be sold to create income. This manual collection method has become more and more common in controlling agricultural pests while obtaining precious nutrients and income (van Huis et al. 2013). Collecting insects could protect crops and promote plant reproduction, which can increase the production of plant food (Van Huis et al. 2013; Yen 2015). Natural enemy insects could eliminate pests and ensure a good harvest of crops, and they are the guardians of the land. For example, ladybugs eat aphids, predatory mites prey on harmful mites such as red spiders, and Trichogramma parasitize and eliminate pests such as pine caterpillars (Che 2018). Capturing pests such as grasshoppers and bamboo caterpillars could prevent and control pests even without using pesticides (Lamsal et al. 2019). Studies reported that edible insects considered to be pests could be controlled by mechanical methods, which could reduce the harmful effects of pesticides on crops and the natural environment. Capture of insects is a practical pest control method that could be widely used in crop systems all over the world (Cerritos and Cano-Santana 2008; Imathiu 2020). In addition, with the increase of cultivated species of edible insects, the safety risks for agricultural and forestry plants may increase. It is necessary to evaluate the safety of the development and utilization of insect capture on the basis of insect biology and ecology, and then perform a large-scale artificial rearing (Feng and Chen 2016a).

1.1.4 Insect Food in China: History and Current Status Insects are one of the oldest biological species in the world, in existence as early as 400 million years ago. During this long history, Chinese people have continuously developed and utilized insect resources in social practice, which promoted the development of their own material production and the improvement and enrichment of human spiritual culture (Yi et al. 2010; Liu 2021). An important aspect of the development and utilization of insect resources is the consumption of insects. Chinese people have eaten edible insects since ancient times (Yi et al. 2010). Until now, 178 species of edible insects have been scientifically identified and investigated. However, the number of edible insects is much greater than the number recorded, which means that there are still many insects that could be used as human food resources (Chen et al. 2009). The preparation methods for edible insects include frying, braising, stewing, stewing after frying, boiling and roasting. They are eaten at stages from eggs and larvae to pupae to adult insects (Chen et al. 2009; Gou et al. 2020b). Chinese edible insect culture has a long history with extensive and profound content. Factors such as the economy, politics, the natural environment and witchcraft ideology have all played an important role in its development. Ancestors’ behavior of eating insects was repeated by later generations, and this habit was passed on from generation to generation, gradually evolving into a unique food culture (Yi et al. 2010). This kind of food culture has been continuously developed and enriched by people in different times and has become a valuable part of the history of Chinese culture.

1  •  The Role of Insects for Nature and Humans  9

1.1.4.1 Long history of insect food in China Chinese insect resources are widely distributed, and China was one of the earliest countries in the world to eat insects and one of the earliest countries in the world to develop insect resources. Chinese residents have a lot of experience in the use of insects (Feng 2016a) (Zhang et al. 2008; Yi et al. 2010). Chinese people have used insect resources for more than 5,000 years. According to historical records, the Chinese started silkworm breeding more than 5,200 years ago (Yanhuang period; Yi et al. 2010). As China is a big silkworm country, eating silkworm pupae should have been an early custom in food culture. However, it was not until the Jin Dynasty that there was a written record of eating silkworm pupae (Liu and Zhai 2017). Chinese residents started to eat insects more than 3,200 years ago and to use insects for entertainment more than 2,000 years ago. The early Chinese books Zhou Ji and Shi Jing recorded the history of insects as medicines (Yi et al. 2010). Judging from the existing literature, the history of Chinese edible insect culture can be traced back to the pre-Qin Dynasty period and the Western Zhou Dynasty period (Liu and Zhai 2017). A sauce made with white ant eggs was not only a precious delicacy liked by the royal family and aristocrats in the Zhou Dynasty, but was also use as an offering in temples. At present, there is no recorded way of consumption. Considering that insect food was consumed the royal family, it must have tasted good (Liu and Zhai 2017; Liu et al. 2021). During the Han, Wei, Jin and Northern and Southern Dynasties, Chinese edible insect culture developed further. Edible insects such as ants and cicadas were still popular. In the Northern Wei Dynasty, there were records of eating insects such as silkworm chrysalises, locusts, cyanosis, grubs, etc., that were not recorded in previous documents (Liu and Zhai 2017). Since ancient times, locusts have been one of the most common pests in farmland that endanger agricultural production. Many folk cultures evidence people’s aversion to locusts and other pests and the problems they caused. When locust infestation occurred, farmland crops often failed to yield a harvest, causing people everywhere to starve. Chinese ancients eliminated the locust, not only by using physical methods such as catching and trapping, but also by using “sympathetic witchcraft” to “curse” the locusts across time and space, hoping to eliminate the harm caused by the infestation and improve their lives. It could be speculated that eating locusts during the Three Kingdoms period became a food culture. Also, when people didn’t have enough food to eat, they cooked locusts as a staple food to satisfy their hunger. Gradually, people not only came to rely on eating locusts to reduce the impact of disaster during locust plagues, but also began to catch large numbers of locusts to use as a everyday food, and to process and store excess locusts to remediate food shortages. Edible locusts were very popular in folk food culture during the Tang and Song Dynasties, and there are many records of edible locusts (Liu and Zhai 2017; Liu et al. 2021). During the Tang and Song Dynasties, with the prosperity of the feudal economy, the culture of edible insects was further developed. Edible ant-egg paste spread in the food culture of the Southern Region of China. There are many peculiarities in the edible insect culture during the Tang and Song Dynasties that are different from the previous generations, such as the use of edible bugs, prawns and other insects. For example, bed bugs are insects that parasitize humans and animals. They are deeply disgusting to some people, but other people like to eat them (Liu and Zhai 2017). The Yuan, Ming and Qing Dynasties were the period when Chinese traditional culture was integrated. Adapting to this, the food culture also developed into a prosperous period. The idea of the “culinary” gradually emerged, and the menus of these cuisines contain various insect dishes. For example, the fried bean worm in Shandong cuisine is the larva of the bean hawk moth, commonly known as Doudan. Its protein content is extremely high, and it is loved by the people of northern Shandong (Liu and Zhai 2017). The books that intensively reflect the cultural achievements of Chinese edible insects are the Ben Cao Gang Mu (Compendium of Materia Medica) by Li Shizhen in the Ming Dynasty and the Supplement to Compendium of Materia Medica by Zhao Xuemin in the Qing Dynasty. The Ben Cao Gang Mu records more than 100 kinds of edible insects such as cicadas, bees, silkworms and crickets, and the Supplement

10  Edible Insects Processing for Food and Feed to Compendium of Materia Medica, the sequel to Ben Cao Gang Mu, records a large number of edible insects that did not appear in the Ben Cao Gang Mu, such as the honey tiger, dragon lice, foreign insects, brown insects and so on. The authors of the two books not only explain the eating methods and medicinal principles of these insects, but also organize and compile their eating history and customs, which are a concentrated reflection of China’s extensive culture of eating insects (Liu and Zhai 2017).

1.1.4.2 Current insect food in China In the past 40 years, research on edible insects has gradually increased. Especially in recent years, as human awareness and acceptance of wild edible insects has increased, and researchers around the world have actively advocated for edible insects, edible insects have been the subject of more widespread research. In the past 30 years, Chinese research on edible insects and the publication of papers and books have increased rapidly. The content of this research includes edible insect customs, edible species, identification of edible insects, nutrient analysis and evaluation, health value and artificial rearing, insect food production or safety and many other aspects, and edible insects in China are forming a new industry (Feng 2016a; Feng 2016c; Feng et al. 2018). There are currently 324 species in 11 orders of edible or related insects that have been recorded in China, including edible species, some less commonly used species and some medicinal insects; some of these have been collected and investigated. The edible insects have not been specifically studied and identified and are not included. The nutritional value of 178 insect species in China has been studied, analyzed and reported, including edible insects and species used for feed and medicinal purposes. Studies have found that the nutritional value of edible insects varies from species to species, but that their protein, fat, vitamin and mineral content can meet human nutritional requirements (Chen et al. 2009; Feng 2016c; Feng et al. 2018). For a long time, edible insects have been eaten by residents of different ethnic groups in many regions of China. Due to the vast territory and large ethnic groups and populations involved, a rich and colorful dietary culture has developed. The acceptance of eating insects as food varies greatly, and varies from region to region (Feng 2016b). Over time, Chinese culture has been closely related to insects, and a unique Chinese insect culture has been formed. Until now, China has seen the emergence of more than 100 traditional festivals related to insects, such as “Silkworm Day” in Nanchong City, Sichuan Province; the “Longcan Festival” in Tongxiang County, Zhejiang Province; the “Insect Delivering Festival” in Nantong City, Jiangsu Province; and the “Double Butterfly Festival” in Yixing City, Jiangsu Province (Yi et al. 2010). Chinese residents eat insects directly or eat products made from insects ( Figures 1.3, 1.4, and 1.5). In Chinese cities, as edible insects are delicious and come from nature without pesticides, food additives and other artificial products, residents like to eat edible insects. Among Chinese minority groups, people consider edible insects as part of their custom (Chen et al. 2009). However, it was reported that there are less than 100 species that are commonly eaten, and the frequently eaten species are 10–20 edible insects including bees and wasps, silkworms, crickets, bamboo caterpillars, dragonflies and beetles (Feng 2016b; Feng et al. 2018). In order to better utilize and promote insects as food resources in their daily diet, humans can use insects directly or indirectly (Feng 2016b; Feng et al. 2018). Direct entomophagy involves cooking and processing insects directly for human consumption. In some areas where edible insects are often eaten, cicada pupae, oriental migratory locusts, meyen, wasp pupae, bamboo worms, crickets, wasps and other edible insects will be eaten directly (Feng 2016b; Feng et al. 2018). As mentioned above, edible insects also have significant regional and ethnic characteristics (Figures 1.6 and 1.7). Generally speaking, the southern part of the country uses more varieties than the northern part, and the processing methods include deep-frying, boiling, stir-frying and steaming; they are also eaten raw. The main species used in northern China are silkworm pupae and yellow mealworms. Cooking methods include boiling, frying and deep-frying. People in northeast China like to eat fried silkworm pupae, and Shandong people like to eat bean hawk moth larvae (Feng 2016b), common silkworms, bamboo worms, crickets, wasps, etc. (Feng et al. 2018).

1  •  The Role of Insects for Nature and Humans  11

FIGURE 1.3  Industrialized cultivation of Tenebrio molitor.

FIGURE 1.4  Cultured Tenebrio molitor.

In Zhejiang Province of south China, located in the south wing of the Yangtze River Delta on the southeast coast of China, deep-fried cicadas are very popular. It has been estimated that residents could eat five tons a day as a delicious dish for banquets (Zhao et al. 2015; Feng 2016b). In Southwest China’s Yunnan Province, there are many kinds of edible insects. Vespa pupae, bamboo worms, grasshoppers, termites, etc. are insects that the locals love to eat. Some restaurants prepare distinctive insect dishes with

12  Edible Insects Processing for Food and Feed

FIGURE 1.5  Artificial breeding of grasshoppers.

a meaningful dish name, which has the benefit of attracting diners and popularizing knowledge about eating insects. As the acceptance of using insects increases, it is believed that more insect food with rich nutrition, delicious flavor and appealing color will be developed (He et al. 2019; Feng 2016b). Cooking plays an important role in the direct utilization of insects, and it could promote the popularity of insects as food (Feng et al. 2018). Cooking is a process in which edible raw materials are processed into food by appropriate methods in order to meet consumers’ physical and psychological needs. After cooking, insects can provide consumers with a nutritious, healthy and safe diet to ensure dietary hygiene and meet human material dietary needs. They also provide a diet with good color, shape and taste to meet human psychological needs (Feng 2016b). Various factors, such as the insect species, developmental stage and processing methods, result in different flavors of insect dishes. The taste of insect dishes can be compared with known dishes. Researchers have reported that 48 kinds of volatile aroma components have been detected from roasted yellow mealworm larvae and barley larvae; 8 of these are aroma pyrazines formed by the Maillard reaction during food thermal processing, 11 kinds are substances related to the aroma compounds of traditionally baked dishes (roasted meat, bread, potatoes, etc.) such as grilled meat and oil. Yellow mealworm larvae produce

1  •  The Role of Insects for Nature and Humans  13

FIGURE 1.6  Cicadas sold in Shandong province farmers’ market.

FIGURE 1.7  Hermetia illucens after microwave drying.

14  Edible Insects Processing for Food and Feed a pleasant aroma and bready aroma after roasting at 180°C, while pyrazine and carbonyl compounds have a decisive effect on the aroma of insect products (Żołnierczyk and Szumny 2021). Edible insects can be processed by various methods, such as deep-frying, frying, boiling and steaming and in cold salad dressings. Fried insects are a common snack, and insects are cooked on barbecue stands. Some restaurants have produced distinctive insect dishes, and have given a meaningful dish name, which has a good effect on attracting diners and popularizing knowledge about eating insects. As the acceptance of consuming insects grows, it is believed that more insect foods with rich nutrition and flavor will appear in Chinese markets (Feng 2016b). Insect dishes are part of traditional Chinese cooking, so they are also reflected in the Chinese cultural consciousness. Chinese residents are familiar with the concept of eating insects (Verneau et al. 2021). However, insect food has not yet become the main food staple. The reasons why people cannot accept insects as food are psychological aversion and the lack of knowledge about them. The lack of insect processing and cooking methods is also one of the reasons (Feng 2016b). It is very important to explore consumers’ perceptive and appetites for insect foods, and to develop dishes that not only fully retain the rich nutrients of insects, but also conform to the local food culture, so that ultimately most consumers can accept insect food (Aarts 2020; Feng 2016b). Further taste education, sharing information on edible insects and tasting edible insects are necessary to improve consumers’ familiarity with the taste and texture of edible insects (Mishyna et al. 2020). Before insect food products are developed, attention should be paid to the types of insect food and their cooking methods, which may be influenced by consumers’ personal culture and expectations. The emphasis on healthy environmental benefits and sustainability is not enough to encourage the promotion of insect consumption (Chen et al. 2009). It is difficult for some Chinese people to accept insects as food and eat them directly. Although insects have been proven to have high nutritional value, some people think that insects are dirty and disgusting as food material, and they often refuse to eat insects. A Chinese researcher has investigated 614 Chinese consumers from Beijing City and Nanjing City on the main factors affecting the purchase of insect food. The results showed that the main factors include insect phobia, aversion, knowledge level and socio-demographic factors, such as age, housing size, family income and whether they are from the North or South China region. In addition, the results show that the positive factors affecting the frequency of eating insects are the preferences of the children in the family, as well as their age and knowledge level. On the contrary, negative factors include concerns about food safety and the shape of insects. The result shows that improving consumers’ awareness and education about eating insects can increase their frequency of consuming insect food. On the other hand, high-end restaurants often prepare edible insects as raw materials to cook unique dishes for high-income groups. High-income consumer groups may be target consumer groups for insect food (Liu et al. 2020). The main difference between insect eaters and non-eaters in China lies in the aversion to insects. Disgusters are better able to predict the intentions of non-eaters, while positive attitudes are more predictive of the interest of insect-eaters (Verneau et al. 2021). Therefore, in addition to eating insects directly in restaurants, the industrialization of insect food needs to be changed to overcome prejudice. It is necessary to conduct more in-depth scientific tests on insect functional foods to confirm their health value and potential functional activity, and to apply modern science and technology to further process insect foods. For example, edible insect products can be processed into insect powder, insect alcohol and nutrient insect liquid (Chen et al. 2009). Insects are rich in nutrients, and high-quality insect protein can also be used as an additive or nutrient for food production. Processing edible insects into granular and paste forms is more suitable for consumers who are not used to eating whole insects (Govorushko 2019). Studies have reported that yellow mealworm pupa has the function of replenishing protein. Adding yellow mealworm pupa protein can increase the protein content of bread, improve the bread’s nutritional value and flavor and enhance its sensory quality (Peng et al. 2013). Bread produced by adding 15% cricket protein pulp has excellent color and flavor, high protein and unsaturated fatty acid content and strong nutritional value, and it meets food safety standards (Gou et al. 2020a). A new type of edible insect fish tofu product could be obtained by

1  •  The Role of Insects for Nature and Humans  15 adding 10% Clanis bilineata tsingtauic powder. The protein uniformity in the gel matrix is good, and its gel structure is also the most compact (Cao et al. 2019). Chinese researchers have conducted research on insect rearing, product development, and rearingrelated technology for the artificial cultivation and utilization of edible insects. At present, China has improved research on the development of insect protein powder, oil, processed chitin products and health foods. The Chinese edible insect industry is forming a new industrial sector. In addition, Chinese residents also eat insects indirectly by eating insect-fed livestock in order to better utilize and promote insects as food resources in everyday diets, which may be a more acceptable way to use insects in human diets (Feng 2016b; Feng et al. 2018). Indirect entomophagy is the feeding of insects. Compared with direct entomophagy, it has the obvious characteristics of being easy to accept and very safe; it also has multiple functions, can be standardized and cultivated and is beneficial for environmental protection (Feng 2016b; La Barbera et al. 2019). In China, insect-raised chickens and eggs are recognized as green nutritious food. At present, the use of insects as feed protein requires more in-depth animal feeding research, and the most important thing is to develop an efficient and low-cost feed insect breeding model, so that insect feed has a price advantage (Feng 2016b). Recently, interest in direct and indirect entomophagy has seen a huge increase around the world. Using insects as raw materials for food and feed is a potential way to meet the growing global food demand and to improve sustainable animal diets (Verneau et al. 2021). Insects will be used as a generally accepted protein food resource in the future, and more in-depth research needs to be carried out in many of its aspects. As a new protein resource, insects can be developed and utilized for food through a variety of ways such as the direct use of insect protein, health food and protein feed to meet the needs of various applications of this new protein resource (Feng 2016b).

1.2 CONCLUSION This chapter summarized the role of insects on ecosystems and humans, and also traced the history of the use of edible insects in China. Insects will be used as a generally accepted protein food resource in the future, and more in-depth research needs to be carried out in many of its aspects. Using insects as raw materials for food and feed is a potential way to meet the growing global food demand and improve sustainable development. As a new protein resource, insect food could be developed and utilized in various ways, such as the direct use of insect protein, insect dishes, health food, personalized insect snacks and protein feed, etc., to meet all applications of this new protein resource.

1.3  PROSPECTIVE AND CHALLENGES Although the edible insect industry shows good development prospects, it also faces many challenges. Firstly, the feeding cost of edible insects remains high, relying on human breeding because of the lack of specialized equipment for breeding and processing. Secondly, the cost of raw materials for breeding is high. For example, the cost of wheat bran, the main raw material for breeding Tenebrio molitor, has been rising in recent years. At present, some Chinese enterprises use food waste and other organic wastes as raw materials, which reduces the cost of raw materials. Thirdly, the lack of insect product hygiene standards and quality standards means that the quality of insect products on the market varies greatly.

16  Edible Insects Processing for Food and Feed Therefore, the complex issues in this emerging field must be resolved by transcending traditional boundaries and requiring cooperation between all stakeholders, especially between the public sector, academia and private companies. It requires a combination of multiple disciplines and methods to promote the development of new agricultural fields and feeds that feed on insects. The development of edible insect technology requires multidisciplinary cooperation between disciplines, and it is necessary to strengthen exchanges and cooperation among social scientists, entomologists and non-academic stakeholders. In addition, as an acceptable food source for sustainable development, insects also require joint research in psychology and gastronomy.

REFERENCES Aarts, K. W. P. 2020. How to develop insect-based ingredients for feed and food? A company’s perspective. Journal of Insects as Food and Feed 6(1):67–68. Baiano, A. 2020. Edible insects: An overview on nutritional characteristics, safety, farming, production technologies, regulatory framework, and socio-economic and ethical implications. Trends in Food Science & Technology 100:35–50. Barroso, F. G., de Haro, C., Sánchez-Muros, M. J., et al. 2014. The potential of various insect species for use as food for fish. Aquaculture 422:193–201. Cao, J., Yue, C., and Guo, Y. 2019. Effect of the addition of Clanis bilineata tsingtauica on qualitative properties of fish tofu. Food Research and Development 40(8):133–138. Cerritos, R. 2009. Insects as food: An ecological, social and economical approach. CAB Reviews: Perspectives in Agriculture, Veterinary Science, Nutrition and Natural Resources 4(27):1–10. Cerritos, R., and Cano-Santana, Z. 2008. Harvesting grasshoppers Sphenarium purpurascens in Mexico for human consumption: A comparison with insecticidal control for managing pest outbreaks. Crop Protection 27(3):473–480. Che, J. 2018. Introduction to Insects. Pictorial book on 100 kinds of edible and medicinal insects. Chemical Industry Press. Chen, X., Feng, Y., and Chen, Z. 2009. Common edible insects and their utilization in China. Entomological Research 39(5):299–303. de Carvalho, N. M., Madureira, A. R., and Pintado, M. E. 2020. The potential of insects as food sources–a review. Critical Reviews in Food Science and Nutrition 60(21):3642–3652. Desa, U.N. 2016. Transforming our world: The 2030 agenda for sustainable development. Goal 12. Ensure sustainable consumption and production patterns. https://stg​-wedocs​.unep​.org​/ bitstream​/ handle​/20​.500​.11822​/11125​/ unepswiosm1inf7sdg​.pdf​?sequence=1 Doi, H., and Mulia, R. N. 2021. Future land ese for insect meat production among countries: A global classification. Frontiers in Nutrition 8:661056. FAO. 2013. The Contribution of Insects to Food Security, Livelihoods and the Environment. Food and Agriculture Organization. http://www​.fao​.org​/3​/i3264e​/i3264e00​.pdf Feng, Y. 2016a. Introduction. In Edible Insects of China, eds. Feng, Y., Chen, X., and Zhao, M., 3–14. Science Press. Feng, Y. 2016b. Utilization and prospect of edible insects. In Edible Insects of China, eds. Feng, Y., Chen, X., and Zhao, M., 100–106. Science Press. Feng, Y. 2016c. Edible insects of China. In Edible Insects of China, eds. Feng, Y., Chen, X., and Zhao, M., 107. Science Press. Feng, Y., and Chen, X. 2016a. The potential influence of edible insect utilization on humans and the environment. In Edible Insects of China, eds. Feng, Y., Chen, X., and Zhao, M., 48–59. Science Press. Feng, Y., and Chen, X. 2016b. Coleoptera (beetles). In Edible Insects of China, eds. Feng, Y., Chen, X., and Zhao, M., 153–190. Science Press. Feng, Y., Chen, X. M., and Zhao, M., et al. 2018. Edible insects in China: Utilization and prospects. Insect Science 25(2):184–198. Feng, Y., and Zhao, M. 2016. Orthoptera (grasshoppers, locusts, katydids, crickets). In Edible Insects of China, eds. Feng, Y., Chen, X., and Zhao, M., 127–139. Science Press. Feng, Y., Zhao, M., Ding, W., et al. 2020. Overview of edible insect resources and common species utilisation in China. Journal of Insects as Food and Feed 6(1):13–25.

1  •  The Role of Insects for Nature and Humans  17 Gou, M., Li, J., Zhao, Y., et al. 2020a. Application and research progress of insect protein in food. Meat Industry 06:50–54. Gou, M., Li, J., Guo, X., et al. 2020b. Research on the preparation of cricket protein bread. Food and Fermentation Sciences & Technology 56(05):30–35. Govorushko, S. 2019. Global status of insects as food and feed source: A review. Trends in Food Science & Technology 91:436–445. Halloran, A., Roos, N., Eilenberg, J., et al. 2016. Life cycle assessment of edible insects for food protein: A review. Agronomy for Sustainable Development 36(4):1–13. Hanboonsong, Y. 2010. Edible insects and associated food habits in Thailand. In Forest Insects as Food: Humans Bite Back, Proceedings of a Workshop on Asia-Pacific Resources and Their Potential for Development, eds. Durst, P.B., Johnson, D. V., Leslie, R.N., et al., 173–182. FAO Regional Office for Asia and the Pacific Press. Hanboonsong, Y., Jamjanya, T., and Durst, P. B. 2013. Six-legged livestock: Edible insect farming, collection and marketing in Thailand. RAP Publication 3:1–57. He, Z., Zhao, M., Wang, C. Y., et al. 2019. Purine and uric acid contents of common edible insects in Southwest China. Journal of Insects as Food and Feed 5(4):293–299. Huang, G., Long, K., Zhu, F., et al. 2013. The research progress and utilization prospect of Maggot in low carbon husbandry’s development. Hubei Agricultural Sciences 14:3233–3237. Imathiu, S. 2020. Benefits and food safety concerns associated with consumption of edible insects. NFS Journal 18:1–11. Ingram, M., Nabhan, G., and Buchmann, S. L. 1996. Our forgotten pollinators: Protecting the birds and bees. Global Pesticide Campaigner 6(4):1–12. Jideani, A. I., and Netshiheni, R. K. 2017. Selected edible insects and their products in traditional medicine, food, and pharmaceutical industries in Africa: Utilization and prospects. In Future Foods, ed. Mikkola, H., 55–69. Intech Open Press. Jongema, Y. 2017. List of edible insect species of the world. Laboratory of Entomology, Wageningen University. https://www​.wur​.nl ​/en ​/ Research​-Results​/Chair​-groups​/ Plant​- Sciences​/ Laboratory​- of​-Entomology​/ Edible​ -insects​/ Worldwide​-species​-list​.htm. Kim, T. K., Yong, H. I., Kim, Y. B., et al. 2019. Edible insects as a protein source: A review of public perception, processing technology, and research trends. Food Science of Animal Resources 39(4):521–540. Kok, R., and A. van Huis. 2021. Insect food in space. Journal of Insects as Food and Feed 7(1):1–4. Kumar, S., Negi, S., and Mandpe, A., et al. 2018. Rapid composting techniques in Indian context and utilization of black soldier fly for enhanced decomposition of biodegradable wastes-A comprehensive review. Journal of Environmental Management 227:189–199. La Barbera, F., Verneau, F., and Coppola, A. 2019. Entomophagy: A contribution to the understanding of consumer intention. Calitatea 20(S2):329–334. Lamsal, B., Wang, H., Pinsirodom, P., et al. 2019. Applications of insect-derived protein ingredients in food and feed industry. Journal of the American Oil Chemists’ Society 96(2):105–123. Liu, A. J., Li, J., and Gómez, M. I. 2020. Factors influencing consumption of edible insects for Chinese consumers. Insects 11(1):10. Liu, M. 2021. Brief introduction to Chinese insect-associated Folklore culture. Agricultural Archaeology 1:237–244. Liu, M., and Zhai, R. 2017. The history and industrial development of edible insect culture in China. Agricultural Archaeology 3:223–232. Luo, Z. 2005. Insects as traditional food in China. In Ecological Implications of Minilivestock: Potential of Insects, Rodents, Frogs and Snails, ed. Paoletti, M.G., 475–480. Science Publishers Press. McConnell, M. D., and Burger Jr, L. W. 2020. Precision conservation to enhance wildlife benefits in agricultural landscapes. In Precision Conservation: Goespatial Techniques for Agricultural and Natural Resources Conservation, eds. Delgado,J. A., Sassenrath, G.F., and Mueller, T., 1–28. Acsess Press. Mehrwar, V., and Uniyal, V. P. 2021. Insect pest diversity of standing crops and traditional pest management in agricultural areas of Mandakini Valley, Garhwal Himalaya, Uttarakhand, India. International journal of Horticulture, Agriculture and Food Science 5(4):1–9. Mishyna, M., Chen, J., and Benjamin, O. 2020. Sensory attributes of edible insects and insect-based foods-Future outlooks for enhancing consumer appeal. Trends in Food Science & Technology 95:141–148. Naylor, R., and Ehrlich, P. R. 1997. Natural pest control services and agriculture. In Nature’s Services: Societal Dependence on Natural Ecosystems, ed. Daily, G. C., 151–174. Island Press. Paoletti, M. G., and Dreon, A. L. 2005. Minilivestock, environment, sustainability, and local knowledge disappearance. In Ecological implications of minilivestock: Potential of insects, rodents, frogs, and snails, ed. Paoletti, M. G., 1–18. CRC Press.

18  Edible Insects Processing for Food and Feed Patel, S., Suleria, H. A. R., and Rauf, A. 2019. Edible insects as innovative foods: Nutritional and functional assessments. Trends in Food Science & Technology 86:352–359. Peng, W., Ma, N. L., Zhang, D., et al. 2020. A review of historical and recent locust outbreaks: Links to global warming, food security and mitigation strategies. Environmental Research 191:110046. Peng, Y., Lin, H., Yue, X., et al. 2013. Processing formula of Tenebrio molitor pupae protein bread. Journal of Anhui Agricultural University 40(5):790–794. Petkova, M. 2019. Edible insects and related products. In The Role of Alternative and Innovative Food Ingredients and Products in Consumer Wellness, ed. C. M. Galanakis, 139–164. Academic Press. Pires, C. S. S., and Maués, M. M. 2020. Insect pollinators, major threats and mitigation measures. Neotropical Entomology 49:469–471. Roffey, J. 1979. Locusts and grasshoppers of economic importance in Thailand. Locusts and Grasshoppers of Economic Importance in Thailand 14:139–150. Rumpold, B. A., and Langen, N. 2020. Consumer acceptance of edible insects in an organic waste-based bio-economy. Current Opinion in Green and Sustainable Chemistry 23:80–84. Rumpold, B. A., and Schlüter, O. K. 2013. Nutritional composition and safety aspects of edible insects. Molecular Nutrition & Food Research 57(5):802–823. Sheppard, D. C., Newton, G. L., Thompson, S. A., et al. 1994. A value added manure management system using the black soldier fly. Bioresource Technology 50(3):275–279. Schiemer, C., Halloran, A., and Jespersen, K. 2018. Marketing insects: superfood or solution-food? In Edible Insects in Sustainable Food Systems, eds. Halloran, A., Flore, R., Vantomme, P., et al., 213–236. Springer. Cham Press. Schultz, J. C., Edger, P. P., Body, M. J., et al. 2019. A galling insect activates plant reproductive programs during gall development. Scientific Reports 9(1):1–17. Tao, J., and Li, Y. O. 2018. Edible insects as a means to address global malnutrition and food insecurity issues. Food Quality and Safety 2(1):17–26. Tang, C., Yang, D., Liao, H., et al. 2019. Edible insects as a food source: A review. Food Production, Processing and Nutrition 1(1):1–13. Tingle, F. C., Mitchell, E. R., and Copeland, W. W. 1975. The soldier fly, Hermetia illucens in poultry houses in north central Florida [Insect pests]. Journal-Georgia Entomological Society 10:179–183. Tomberlin, J. K., and van Huis, A. 2020. Black soldier fly from pest to ‘crown jewel’ of the insects as feed industry: an historical perspective. Journal of Insects as Food and Feed 6(1):1–4. van Huis, A. 2018. Insects as human food. In Ethnozoology, eds. R. R. N. Alves and U. P. Albuquerque, 195–213. Academic Press. van Huis, A. 2020. Nutrition and health of edible insects. Current Opinion in Clinical Nutrition & Metabolic Care 23(3):228–231. van Huis, A. 2021. Prospects of insects as food and feed. Organic Agriculture 11(2):301–308. van Huis, A., and Oonincx, D. G. 2017. The environmental sustainability of insects as food and feed. A review. Agronomy for Sustainable Development 37(5):1–14. van Huis, A., Van Itterbeeck, J., Klunder, H., et al. 2013. Edible Insects: Future Prospects for Food and Feed Security (No. 171). Food and Agriculture Organization of the United Nations. http://www​.fao​.org​/3​/i3253e​/i3253e​.pdf Varelas, V. 2019. Food wastes as a potential new source for edible insect mass production for food and feed: A review. Fermentation 5(3):81. Van Zanten, H. H., Mollenhorst, H., Oonincx, D. G., et al. 2015. From environmental nuisance to environmental opportunity: Housefly larvae convert waste to livestock feed. Journal of Cleaner Production 102:362–369. Veldkamp, T., Van Duinkerken, G., van Huis, A., et al. 2012. Insects as a sustainable feed ingredient in pig and poultry diets: A feasibility study No. 638. Wageningen ur Livestock Research. https://library​.wur​.nl​/ WebQuery​ /wurpubs​/fulltext ​/234247. Verneau, F., Zhou, Y., Amato, M., et al. 2021. Cross-validation of the entomophagy attitude questionnaire (EAQ): A study in China on eaters and non-eaters. Food Quality and Preference 87:104029. Weinzierl, R., and Henn, T. 2020. Beneficial insects and mites. In Handbook of Integrated Pest Management for Turf and Ornamentals, ed. Leslie, A. R., 443–452. Lewis Publishers Press. Wilby, A., and Thomas, M. B. 2002. Natural enemy diversity and pest control: Patterns of pest emergence with agricultural intensification. Ecology Letters 5(3):353–360. Wu, W. M., and Criddle, C. S. 2021. Characterization of biodegradation of plastics in insect larvae. Methods in Enzymology 648:95–120. Yen, A. L. 2015. Insects as food and feed in the Asia Pacific region: Current perspectives and future directions. Journal of Insects as Food and Feed 1(1):33–55.

1  •  The Role of Insects for Nature and Humans  19 Yi, C., He, Q., Wang, L., et al. 2010. The utilization of insect-resources in Chinese rural area. Journal of Agricultural Science 2(3):146–154. Zhao, R., Wang, G., and Wang, A. Y. 2015. Ingredients in Chinese traditional cooking. In A History of Food Culture in China, ed. Zhao, R., 8–9. World Scientific Press. Zhang, C., Tang, X., and Cheng, J. 2008. The utilization and industrialization of insect resources in China. Entomological Research 38:S38–S47. Zołnierczyk, A. K. and Szumny A. 2021. Sensory and chemical characteristic of two insect species: Tenebrio molitor and Zophobas morio larvae affected by roasting processes. Molecules, 26(9), 2697.

Environmental Impact of Edible Insect Processing

2

Sergiy Smetana, Dusan Ristic, and Volker Heinz Contents 2.1 Introduction 21 2.2 Insect Feed Processing 23 2.3 Killing, Decontamination, Separation 24 2.4 Fractionation – Allocation or Substitution 24 2.5 Complex Processing Technologies 25 2.6 Food or Feed Applications 26 2.7 Circular Economy Relevance (Side Streams, Special Products) 26 2.8 Conclusions and Outlook 27 References 28

2.1 INTRODUCTION Traditional life cycle assessment (LCA) applications include determining a product’s environmental impact across its full life cycle, identifying environmental hotspots (stages with the largest contribution), performing product comparison, and defining trade-offs. Comparison of alternative products and trade-off minimization for the selection of fewer environmentally damaging solutions (Cucurachi et al., 2019) has become a “golden” standard in environmental product and service assessment. Furthermore, LCA serves as a foundation for Environmental Product Declaration (del Borghi, 2013; Schau & Fet, 2008) and Product Environmental Footprint (Bach et al., 2018), all of which are included in the European Commission’s guidelines for environmental impact assessment and declaration (Allio, 2007). LCA is a complex method aiming to assess environmental impacts. LCA is divided into four stages, according to ISO standards: (1) goal and scope definition, (2) life cycle inventory (LCI), (3) life cycle impact assessment (LCIA), and (4) interpretation (ISO 14040, 2006; ISO 14044, 2006). Any LCA should include details on the functional unit (FU), system boundaries, impact assessment techniques, and a timeline, as well as assumptions, constraints, data quality and needs, reference flows, and so on (Thabrew et al., 2009). LCIA approaches assign an impact factor to each basic inventory flow, linking the quantity of resources utilized and emissions to the potential environmental impact (Zampori et al., 2016). In addition, DOI: 10.1201/9781003165729-2

21

22  Edible Insects Processing for Food and Feed there are two basic techniques of LCA: attributional (information on the environmental burden connected with a specific product life cycle) and consequential (information on the environmental burden arising because of decision-making with market changes). The type of LCA is determined by the goal and scope of the study. Some research projects focus on an attributional LCA, while others focus on both attributional and consequential LCAs. The current chapter includes clear differentiation on the application of the approaches to insect processing chains. While it will highlight a few processing technologies applied in insect production chains, it will not determine the processing characteristics, which is the scope of Chapter 12. A great diversity of insect species and processing techniques applied to insect production chains create a challenge in terms of environmental impact assessment (Smetana, Spykman et al., 2021). The modular LCA approach is suggested as one of the efficient options to deal with the diversity. The modular LCA approach is based on breaking down a production system or a product’s life cycle into modules that may then be recombined to build complete value chains (Steubing et al., 2016). These modules, defined by practitioners, cover life cycle stages or unit processes (e.g., feed production, storage, harvesting). Aside from basic flows, the modules only contain input and output flows that connect them to other modules in the production system under investigation. This is accomplished by expanding the foreground process(es) of a module to include any necessary background processes (e.g., utilities, waste treatment, infrastructure). This process is repeated until the complete production chain is represented in modular life cycle inventories (LCI). Life cycle impact assessment (LCIA) is performed for each of the modular LCIs, yielding separate LCIA results for each module. The LCIA result for a value chain is calculated by adding the LCIA values of all the modules involved (Rebitzer, 2005), following certain conditions and rules. LCA modularity is seen as a viable technique to deal with various versions of insect production chains (Smetana, Spykman et al., 2021; Spykman et al., 2021). In classic LCA, each alternative life cycle must be simulated separately, even if only one of the stages has changed. Within a production system, the modular LCA method allows for the assessment of many alternative insect production chains by assessing separate blocks (modules). A product’s life cycle is modelled in multiple options in LCA to create some pre-conditions for the comparability of studies in the same scope (depending on the purpose and scope of the study: from cradle to grave, cradle to gate, gate to gate, etc.). Alternative value chains emerge when several modules within a production system produce interchangeable products. A module–product matrix shows how the modules can be linked to construct different value chains, taking into consideration scaling factors and interdependencies (Steubing et al., 2016). In comparison to traditional LCA, the modelling work can be significantly reduced because it scales with the number of modules rather than the number of alternative value chains. However, the modular approach requires an upfront time investment in modularization, which entails the appropriate specification of modules to reflect important production system choices. Through optimization models, which allow for the detection of missing data points using optimization methods, modular LCA can simplify (aggregate) scenario analysis (for example, using Pareto optimization). It can thus not only enable value chain optimization by using module data as inputs to optimization models (Steubing et al., 2016) but also provide answers to data limitations in emerging technology assessment (Thomas et al., 2020). Among the important modules of insect production, there are a few which are recognized as environmental hotspots. Feed production, insect farming, and processing are recognized as being among the most impacting life cycle stages (Smetana, Spykman et al., 2021; Spykman et al., 2021). Processing technologies can act as catalyzers of environmental impact, either increasing or decreasing it. All those stages include processing technologies to be applied either to the handling of insect feed, or to insect farming (e.g., feeding, harvesting), or to insect biomass processing. Therefore, this chapter will further concentrate on the environmental impact of processing aspects of feed, insect killing and decontamination, fractionation, and complex methods of insect processing as well as variations in environmental impact due to the system changes (food, feed, circularity). Specific attention will be devoted to the identification of trends with changes in environmental impact initial feed or insect biomass.

2  •  Environmental Impact of Edible Insect Processing  23

2.2  INSECT FEED PROCESSING Insect production chains rely to a great degree on existing solutions for feed production and processing. The type of feed used (vegetable residuals, complex feed, food waste, etc.) and its properties (nutrient content, moisture content, etc.) determine the performance and environmental impact of the insect production system to a significant extent (Bosch et al., 2019; Ites et al., 2020; Oonincx & de Boer, 2012; Smetana et al., 2016; Smetana, Schmitt et al., 2019). In many circumstances, higherquality insect feed leads to higher environmental impacts, but also shorter growing cycles. While lower nutritional quality insect feed (which could have a reduced impact on productivity) results in smaller insects, longer growth cycles, and higher conversion ratios, greater nutritional quality insect feed results in larger insects, longer growing cycles, and higher conversion ratios (Bosch et al., 2019; Smetana et al., 2016). Producers and processors of insects must to take this feed intercorrelation into consideration. Furthermore, the system is complicated by the ability of many insect species to grow on different waste types. Biogenic waste treatment may have environmental benefits, especially if the feeding substrate is of high nutritional quality ratios (Bosch et al., 2019; Ites et al., 2020; Salomone et al., 2017; Smetana et al., 2016). Depending on the impact of avoided waste treatment processes, the environmental impact of waste handling using insects could be beneficial (Roffeis et al., 2017, 2020; Smetana, 2020; Smetana et al., 2016). Due to the limitation on data availability, a wide range of viable alternatives, application scales, and feed and food processing phases are now regarded as extremely difficult to model for LCA assessment (Ites et al., 2020; Smetana et al., 2016; Smetana, Schmitt et al., 2019). Insects are typically fed with dried or moisturized feeds. Mealworms, crickets, and grasshoppers are fed dry feeds, whereas fly larvae require moisturized feeds. Dried feeds derived from grains require only minor processing, such as dehulling, mixing, cutting, and grinding. Additional sources of moisture (e.g., vegetables) are sometimes provided in addition to the dried meals (Halloran et al., 2017; Oonincx & de Boer, 2012). Moisturized feeds can be produced by mixing water with dry ingredients or by using feeds that already have a high moisture content. Such high-moisture feeds originate from foodprocessing industries (e.g., wet milling, breweries, dairy companies), animal farming (manure), and anaerobic digestion facilities (digestate). The influence of the feed processing stage is considered to be minor compared to the other production stages. Being similar in impact to the transportation, it is usually disregarded or assessed in an aggregate form. At the same time, if the wet feeds are considered, their transportation and pre-treatment might result in increased impacts. Most current LCA studies do not include insect feed pre-treatment alternatives (Isibika et al., 2019; Ravindran & Jaiswal, 2019). In extreme cases, wet biomass can be dried out to obtain dried feeds for insects, which results in extreme impacts due to the drying process and concentration of impacts (Smetana, Schmitt et al., 2019; Smetana, Spykman et al., 2021). Feed processing can have a considerable role to play in the impact of certain categories. Thus, feed processing is responsible for the highest impact share of non-renewable energy consumption and blue water use, with little variations if the crops are processed for food or feed (Detzel et al., 2018). Similar feed processing gas been shown to be responsible for high energy use impacts when feed is produced for animal farming (Bandekar et al., 2016; Mungkung et al., 2013), especially if it is characterised as high for land-based systems (Jerbi et al., 2012). In terms of greenhouse gas emissions (GHGe), the processing of compound (commercial) feed is accounted for in the range of 13.1 kg CO2eq per ton of feed (Baek et al., 2014), with more complex processing involving brans and meals separation and mixing reaching as high as 148.5 kg CO2eq per ton of feed (Chen et al., 2021). Despite the high variations in the impacts, in general, feed processing is expected to be responsible for 1%–5% of overall environmental impact of the insect production chain.

24  Edible Insects Processing for Food and Feed

2.3  KILLING, DECONTAMINATION, SEPARATION Insects are treated in two phases after they attain the necessary size, age, and composition requirements. Primary processing involves operations such as sieving, separation, blanching, decontamination, or freezing to clean insect biomass and remove microbial load, whereas secondary processing (discussed further) involves milling, fractionation via centrifugation, drying, and fat separation to improve the properties of insect derived components. Entire live insects, whole fresh/frozen insects, fresh insect puree, dried whole larvae, defatted meal, insect oil, and intermediary items incorporating insect components are examples of end products offered to businesses or other customers. Because of differences in processing depths, most LCA studies may be divided into three categories: (1) cradle-to-farm gate, (2) cradle-to-processing gate, and (3) cradle-to-plate. For the first system boundary (1), the end products are live insects that are usually used as feed (Bosch et al., 2019; Komakech et al., 2015; Oonincx & de Boer, 2012; Suckling et al., 2020). The use of living insects for direct food intake is questionable since no firms presently produce live larvae for food. On the other hand, whole larvae and adult insects, decontaminated and sometimes dried, are recommended for use as food and feed (Bava et al., 2019; Halloran et al., 2017; Ites et al., 2020; Mertenat et al., 2019; Roffeis et al., 2015; Salomone et al., 2017). Whole live insects, sold to feed pets and other animals (e.g., poultry), require minimal processing, associated with insect harvesting (sieving, separation, sometimes storage at low temperatures) and transportation. Such fresh insect biomass, sometimes milled in the form of a puree, results in products with low environmental impacts in the range of 1.0–3.1 kg CO2eq per 1 kg of fresh weight for global warming potential, 1.6–3.6 m2 annually for land use occupation, 24.3–33.7 MJ of non-renewable energy use, and 0.38–0.42 m3 of water resource demand (Smetana, Spykman et al., 2021).

2.4  FRACTIONATION – ALLOCATION OR SUBSTITUTION The two basic methods of LCA are attributional and consequential. The first is widely used to determine the environmental impact of a product or a technology, as well as to compare them to alternatives. From this perspective, any by-products generated during the fractionation of insect biomass can uphold a certain environmental burden. Depending on the weight and economic value of a by-product, they may serve to increase or decrease the environmental impact of the main product. Indirect positive and negative impacts connected with possible market effects owing to the avoidance or greater use of other resources are included in consequential modelling (Parajuli et al., 2017). This is particularly relevant in the case of waste valorization, which is frequently intended to prevent waste treatment (Bartocci et al., 2020; Pérez-Camacho et al., 2018), reuse or remanufacturing of potentially avoidable waste into useable resources (Bartocci et al., 2020), and market replacement of energy or other commodities (Bartocci et al., 2020; Hijazi et al., 2020; Pérez-Camacho et al., 2018). According to the LCA, effective circulation of nutrients from organic waste may be characterized as a waste management system that reduces the environmental impact of waste treatment and replaces high-impacting goods or services on the market with bioconversion products. A good example of this is the substitution of food waste for typical anaerobic digestion (AD) feedstock (avoiding the need for raw materials and waste treatment) (Pérez-Camacho et al., 2018). Fractionation (separation of insect biomass into a few fractions: water, lipid, and protein) is the second stage of processing. The quality of fractionation is determined by the technology characteristics. Insect drying, which uses a variety of methods (heat, sun, and freeze-drying), has been identified as one of the most prevalent processing processes in a few studies (Bava et al., 2019; Ites et al., 2020; Mertenat et al., 2019; Roffeis et al., 2015, 2020; Salomone et al., 2017; Smetana, Schmitt et al., 2019). It has a high energy

2  •  Environmental Impact of Edible Insect Processing  25 requirement and might have significant environmental consequences. Heat for insect production might be obtained as a by-product from other businesses in some situations, resulting in little environmental effect (Joensuu & Silvenius, 2017). When a high grade of insect biomass is required for food applications, lyophilization methods (freeze-drying of entire larvae) (Bava et al., 2019; Lenaerts et al., 2018) are used, which have significant energy costs and high environmental implications. Centrifugation, cold or hot pressing (Alles et al., 2020; Smetana, Leonhardt et al., 2020; Smetana, Schmitt et al., 2019), and, in certain cases, supercritical liquid extraction are used to separate water, lipid, and protein fractions from fresh and dry biomass (Purschke et al., 2017). Insect biomass fractionation is also benefiting from emerging food processing technology such as pulsed electric fields (PEF) (Alles et al., 2020; Shorstkii et al., 2020; Smetana, Mhemdi et al., 2020). As a result of these improvements, both energy use and environmental effects are reduced. When it comes to agri-food waste valorization and LCA, the most promising, but also the most difficult, issue is how to deal with residues at the end of the chain and utilize them as a feedstock for biotechnologies (Hijazi et al., 2020). Food waste at the consumer level aggregates the influence of the whole value chain, and its treatment or reuse is critical for improving the food system’s sustainability. The food system, on the other hand, is not self-contained; it draws on resources from other industries and generates its own. The idea of the ‘water–energy–food nexus’ (Udugama et al., 2020) is the most well-known interindustry interaction system. From a holistic circularity and trade-off perspective, such a nexus emphasizes the need to focus research and development on (1) reducing or preventing produced food that may be wasted (Slorach et al., 2019) and (2) ensuring the (calculated economic and environmental impact) management of unpreventable food waste (Castell-Perez et al., 2017; Kibler et al., 2018). The water–energy–food nexus needs more LCA development and adaption. For example, (Bartocci et al., 2020) discovered optimal food waste management scenarios by integrating economic analysis and consequential LCA. The waste from the food industry is highlighted as an environmentally advantageous bioenergy resource. However, such findings may only be reached if just one effect category of the carbon footprint is considered (Bartocci et al., 2020). Some authors propose upgrading the classical LCA with additional coefficients of feedstock intensity and circular-process feedstock intensity, the waste factor and circular-process waste factor, product renewability, process material circularity, and energy intensity and circular-process energy intensity to deal with the complex problems of assessing the circularity of resources in a waste-based bioeconomy, which can provide additional insight into the potential benefits of bio-residues.

2.5  COMPLEX PROCESSING TECHNOLOGIES The development of efficient, sustainable, and low-cost processing methods, as well as waste minimization, recovery, and inclusion of by-products/co-products, are the key difficulties posed by insect processing (Ojha et al., 2021). Although the requisite know-how and high prices remain obstacles, novel food processing technologies could increase the processing efficiency, safety, quality, and sustainability of insect-based products. Economies of scale are required for several of the technologies outlined, which necessitates higher production quantities than most current systems (Sindermann et al., 2021). Even if the insect business is soon likely to come out with a few new product improvements, the primary dry and wet processing technologies most probably will remain and get into joined processing lines. Some equipment may require small changes during upscaling. Animal welfare is a major concern in the insect food and feed industry (van Huis, 2021). In this discussion, the key topic is whether insects are ‘sentient entities.’ This may be the case when considering (1) their brains’ highly efficient functional organization and (2) their ability to learn socially and associatively, as well as to communicate in a variety of ways. When insects are farmed and killed, the precautionary principle is used, assuming that they can feel pain. (van Huis, 2021) discussed the ramifications of welfare concerns for the edible insect industry using Brambell’s framework of five freedoms: freedom

26  Edible Insects Processing for Food and Feed from hunger and thirst; freedom from discomfort; freedom from pain, injury, or sickness; freedom from fear and anguish; and freedom to exhibit natural behaviour. For the production of meat substitutes, extrusion cooking and pelletizing or high-moisture extrusion is used (González et al., 2019; Roncolini et al., 2020; Smetana, Pernutz et al., 2019; Ulmer et al., 2020). While both technologies have the potential to result in the same product (moisturized minced meat), they can be different in terms of the environmental impact. Thus, a study by (Saerens et al., 2021) indicated that high-moisture extrusion is a more productive and efficient technology comparing to extrusion cooking. However, it varies depending on the type of raw materials being processed. The inclusion of insect biomass in the extruded products has the potential to reduce the environmental impact and be competitive with animal and even plant-based products (Smetana, Profeta et al., 2021).

2.6  FOOD OR FEED APPLICATIONS Several proven and commercialized products are associated with the various application possibilities: pelleted feeds, bars, pasta, spreads, and so on. When fresh or dried insect biomass is combined with plant material, hybrid goods are created, which may have a lesser environmental effect. However, overly thorough processing (in the case of isolates) could worsen the effects of the finished good. The utilization of concentrated protein fractions (insect flour, insect meal, or defatted protein concentrate) and insect lipids (fats and oils) as part of a more complex matrix is related to further product development. Both protein and lipid fractions have uses in feed and pet food (Gasco et al., 2020; Surendra et al., 2016; Zorrilla & Robin, 2019), as well as food (Smetana, Ashtari Larki et al., 2018b; Smetana et al., 2015, 2016; Smetana, Leonhardt et al., 2020; Tzompa-Sosa et al., 2019; Zorrilla & Robin, 2019). Extrusion is one of the example technologies (discussed previously which can rely on the concentrated protein fractions. Lipid fractions are utilized as an animal feed additive (Gasco et al., 2019), in baking (Delicato et al., 2020; Tzompa-Sosa et al., 2019), and in complex fat products like spreads and margarines (Smetana, Leonhardt et al., 2020). Even though cooking of the product at the consumer stage may pose significant environmental impacts associated with long preservation, excessive wasting, and high energy use at inefficient cooking practices, there are only a few studies that performed LCA research in the scope of cradle-to-product application boundaries (Smetana, Spykman et al., 2021). Insect products are frequently rated as having a similar or lower environmental effect as traditional food and feed products, despite many problems associated with poor manufacturing chains.

2.7  CIRCULAR ECONOMY RELEVANCE (SIDE STREAMS, SPECIAL PRODUCTS) Even on a prepared diet, insect production can be more efficient than raising some traditional animals (van Huis & Oonincx, 2017). However, insect production can result in a considerable amount of low-value organic waste, including as frass, exuviae, and uneaten feed, which is frequently disposed of by spreading on agricultural fields (Jucker et al., 2020). The notion of applying insect frass to the ground allows for the restoration of the insect-rearing side stream back into the food production chain, which is in line with circular economy concepts. It has been suggested that it is a superior alternative to linear models that terminate in energy recovery or disposal through incineration and landfilling (IPIFF, 2019). The chemical and physical qualities of insect frass are compatible with those of other commercial fertilizers, and it has shown tremendous promise as a fertilizer (e.g., soil improver, organic fertilizer, or compost material) (Salomone et al., 2017). For example, some researchers suggest that the frass generated by

2  •  Environmental Impact of Edible Insect Processing  27 TABLE 2.1  Matrix of agri-food waste application potential for the production of food substitutes from biomass in relevance to technology readiness levels (Smetana, Aganovic et al., 2018) SOURCE OF BIOMASS FOR SUBSTITUTES

Plant Milk Insect Cultured Single cell protein Microalgae Fungi Y&B

SUBSTITUTED ANIMAL DERIVED PRODUCTS MEAT

DAIRY

EGGS

FAT

IX/9 IX/0 VI-IX/7 VI/4

IX/6 n/a I-VII/8 VI/4

VIII/6 VI/0 VI-VII/8 VI/4

IX/9 IX/O IX/9 VII/6

VII/5 IX/6 VII/6

VIII/5 IX/5 VI/6

VIII/5 VIII/5 VIII/6

IX/7 IX/9 VI/6

Note I–IX—Technology Readiness Level (TRL) after (Mankins, 1995) for the application of source biomass to substitute animal derived products; 0–9—potential of agri-food waste application for the source biomass production (with 0—no potential identified to 9 confirmed industrial application possible); Y&B—yeast and bacteria

Tenebrio molitor may be employed as a biofertilizer in organic farming because of its nutritional value and associated microbiota, which may aid in nutrient absorption (Poveda et al., 2019). In addition, a novel potential for incorporating insect frass into biogas production has demonstrated promising results in a cost-effective and long-term way (Bulak et al., 2020). Some manure, plant wastes, and sewage sludges have great biomethane potential, according to the authors. Insect fat has been successfully utilized to produce biodiesel with similar properties to oilseed-derived fuel, taking advantage of the high fat content of some insect species (Wang et al., 2017; Zheng et al., 2013). Table 2.1 depicts an ideal insect-for-food and feed production system. It shows a simple idea for mass insect generation utilizing food waste from the food production cycle. The image depicts insect rearing for bioconversion with many phases for creating important goods and side streams in a holistic manner. The C-LCA performed on waste stream utilization for insect production (Smetana, Schmitt et al., 2019) indicates that a change in the feed supply chain to insect production relates to demand and availability of high-protein side streams from the food sector in the base scenario, where side streams that may be utilized for other animals are fed to Hermetia illucens (milling, alcohol production, and breweries). Currently, these by-products are used to make feed components. Increased demand for this feed might lead to increased demand for other high-protein feeds (such as soybean meal) to fill the protein feed gap for other animals. Furthermore, the C-LCA method classifies food manufacturing side streams as single goods, allowing for market replacement of milling and brewery products. As a result of this strategy, the environmental effect of black soldier fly (BSF) goods has improved. Therefore, the scenarios based on low-value agri-food items with strong nutritional profiles (e.g., dried distillers grains with solubles [DDGS]) or waste use with significant environmental effect are the most promising for feed production, as the impact of increasing resource usage should not outweigh the advantages of manure utilization.

2.8  CONCLUSIONS AND OUTLOOK The processing of edible insects, like the processing of food, can change the environmental impact of initial biomass. Insect feed is known to be responsible for the highest influence on insect-derived products. The application of processing technologies, like drying or extraction, fractionating, and concentrating nutrients results in better quality feed with higher environmental impact. At the same time, there are other technologies with relatively low environmental impact (fermentation, thermal treatment) which can be

28  Edible Insects Processing for Food and Feed applied to increase the digestibility of feed for insects. In such cases, the environmental impact might be even improved. Therefore, it is first necessary to differentiate the technologies of feed processing according to their ability to change the environmental impact and to deal with those which result in improved environmental impact and improved quality of feed. Further processing technologies dealing with insect biomass result either in an increase in the impact of the final product or its decrease. Processing technologies aiming for the purification of ingredients (milling, separation, concentration, isolation, filtration, etc.) generate new products with higher environmental impact. Especially impactful are those technologies that separate and evaporate water. Different drying technologies result in a dried product with a much higher impact than initial insect biomass. On the other hand, technologies associated with an increase in moisture content result in a “dissolution” of environmental impact and the generation of insect-based products with lower impact. High-moisture extrusion, emulsification and homogenization, or hydration are some examples technologies that can be used for such an approach. Insect-producing companies should therefore rely on the use of fresh or frozen insects (not fractionated), defining the potential to process insect-based food with the inclusion of water in the process. Despite such conclusions, further research is needed to define the impact of insect-processing technologies, especially if the development of new forms of products is expected.

REFERENCES Alles, M. C., Smetana, S., Parniakov, O., Shorstkii, I., Toepfl, S., Aganovic, K., & Heinz, V. (2020). Bio-refinery of insects with pulsed electric field pre-treatment. Innovative Food Science & Emerging Technologies, 64, 102403. https://doi​.org​/10​.1016​/j​.ifset​.2020​.102403 Allio, L. (2007). Better regulation and impact assessment in the European Commission. In Colin Kirkpatrick and David Parker Regulatory Impact Assessment: Towards Better Regulation (pp. 72–105). Edward Elgar Publishing Limited, Glos, UK. https://doi.org/10.4337/9781847208774.00011 Bach, V., Lehmann, A., Görmer, M., & Finkbeiner, M. (2018). Product environmental footprint (PEF) pilot phase— Comparability over flexibility? Sustainability, 10(8), 2898. https://doi​.org​/10​.3390​/su10082898 Baek, C.-Y., Lee, K.-M., & Park, K.-H. (2014). Quantification and control of the greenhouse gas emissions from a dairy cow system. Journal of Cleaner Production, 70, 50–60. https://doi​.org​/10​.1016​/j​.jclepro​.2014​.02​.010 Bandekar, P., Matlock, M., Ulrich, R., & Thoma, G. (2016). Comparative Life Cycle Assessment of Field-to-Fork Pork Supply Chains in the United States and the European Union. Final Report. University of Arkansas. Bartocci, P., Zampilli, M., Liberti, F., Pistolesi, V., Massoli, S., Bidini, G., & Fantozzi, F. (2020). LCA analysis of food waste co-digestion. Science of the Total Environment, 709, 136187. https://doi​.org​/10​.1016​/j​.scitotenv​.2019​ .136187 Bava, L., Jucker, C., Gislon, G., Lupi, D., Savoldelli, S., Zucali, M., & Colombini, S. (2019). Rearing of Hermetia illucens on different organic by-products: Influence on growth, waste reduction, and environmental impact. Animals, 9(6), 289. https://doi​.org​/10​.3390​/ani9060289 Bosch, G., van Zanten, H. H. E., Zamprogna, A., Veenenbos, M., Meijer, N. P., van der Fels-Klerx, H. J., & van Loon, J. J. A. (2019). Conversion of organic resources by black soldier fly larvae: Legislation, efficiency and environmental impact. Journal of Cleaner Production, 222, 355–363. https://doi​.org​/10​.1016​/j​.jclepro​.2019​ .02​.270 Bulak, P., Proc, K., Pawłowska, M., Kasprzycka, A., Berus, W., & Bieganowski, A. (2020). Biogas generation from insects breeding post production wastes. Journal of Cleaner Production, 244, 118777. https://doi​.org​/10​.1016​ /j​.jclepro​.2019​.118777 Castell-Perez, E., Gomes, C., Tahtouh, J., Moreira, R., McLamore, E. S., & Knowles, H. S. (2017). Food processing and waste within the nexus framework. Current Sustainable/Renewable Energy Reports, 4(3), 99–108. https:// doi​.org​/10​.1007​/s40518​- 017​- 0079-z Chen, X., Chen, Y., Liu, X., Li, Y., & Wang, X. (2021). Investigating historical dynamics and mitigation scenarios of anthropogenic greenhouse gas emissions from pig production system in China. Journal of Cleaner Production, 296, 126572. https://doi​.org​/10​.1016​/j​.jclepro​.2021​.126572 Cucurachi, S., Scherer, L., Guinée, J., & Tukker, A. (2019). Life cycle assessment of food systems. One Earth, 1(3), 292–297. https://doi​.org​/10​.1016​/j​.oneear​.2019​.10​.014

2  •  Environmental Impact of Edible Insect Processing  29 del Borghi, A. (2013). LCA and communication: Environmental product declaration. The International Journal of Life Cycle Assessment, 18(2), 293–295. https://doi​.org​/10​.1007​/s11367​- 012​- 0513-9 Delicato, C., Schouteten, J. J., Dewettinck, K., Gellynck, X., & Tzompa-Sosa, D. A. (2020). Consumers’ perception of bakery products with insect fat as partial butter replacement. Food Quality and Preference, 79, 103755. https://doi​.org​/10​.1016​/j​.foodqual​.2019​.103755 Detzel, A., Krüger, M., Busch, M., Drescher, A., Wriessnegger, C. L., & Köppen, S. (2018). Deliverable 5.3 – Part I: Report on the Life Cycle Assessment Results. Protein2Food Project Report. Institut für Energie- und Umweltforschung Heidelberg (ifeu). Gasco, L., Biancarosa, I., & Liland, N. S. (2020). From waste to feed: A review of recent knowledge on insects as producers of protein and fat for animal feeds. Current Opinion in Green and Sustainable Chemistry. https:// doi​.org​/10​.1016​/j​.cogsc​.2020​.03​.003 Gasco, L., Dabbou, S., Trocino, A., Xiccato, G., Capucchio, M. T., Biasato, I., Dezzutto, D., Birolo, M., Meneguz, M., Schiavone, A., & Gai, F. (2019). Effect of dietary supplementation with insect fats on growth performance, digestive efficiency and health of rabbits. Journal of Animal Science and Biotechnology, 10(1), 4. https://doi​ .org​/10​.1186​/s40104​- 018​- 0309-2 González, C. M., Garzón, R., & Rosell, C. M. (2019). Insects as ingredients for bakery goods. A comparison study of H. illucens, A. domestica and T. molitor flours. Innovative Food Science & Emerging Technologies, 51, 205–210. https://doi​.org​/10​.1016​/j​.ifset​.2018​.03​.021 Halloran, A., Hanboonsong, Y., Roos, N., & Bruun, S. (2017). Life cycle assessment of cricket farming in northeastern Thailand. Journal of Cleaner Production, 156, 83–94. https://doi​.org​/10​.1016​/j​.jclepro​.2017​.04​.017 Hijazi, O., Abdelsalam, E., Samer, M., Attia, Y. A., Amer, B. M. A., Amer, M. A., Badr, M., & Bernhardt, H. (2020). Life cycle assessment of the use of nanomaterials in biogas production from anaerobic digestion of manure. Renewable Energy, 148, 417–424. https://doi​.org​/10​.1016​/j​.renene​.2019​.10​.048 IPIFF. (2019). Contribution Paper on the Application of Insect Frass as Fertilising Product in Agriculture. IPIFF Brussels, Belgium. Isibika, A., Vinnerås, B., Kibazohi, O., Zurbrügg, C., & Lalander, C. (2019). Pre-treatment of banana peel to improve composting by black soldier fly (Hermetia illucens (L.), Diptera: Stratiomyidae) larvae. Waste Management, 100, 151–160. https://doi​.org​/10​.1016​/j​.wasman​.2019​.09​.017 Ites, S., Smetana, S., Toepfl, S., & Heinz, V. (2020). Modularity of insect production and processing as a path to efficient and sustainable food waste treatment. Journal of Cleaner Production, 248, 119248. https://doi​.org​/10​ .1016​/j​.jclepro​.2019​.119248 Jerbi, M. A., Aubin, J., Garnaoui, K., Achour, L., & Kacem, A. (2012). Life cycle assessment (LCA) of two rearing techniques of sea bass (Dicentrarchus labrax). Aquacultural Engineering, 46, 1–9. https://doi​.org​/10​.1016​/j​ .aquaeng​.2011​.10​.001 Joensuu, K., & Silvenius, F. (2017). Production of mealworms for human consumption in Finland: A preliminary life cycle assessment. Journal of Insects as Food and Feed, 3(3), 211–216. https://doi​.org​/10​.3920​/JIFF2016​.0029 Jucker, C., Lupi, D., Moore, C. D., Leonardi, M. G., & Savoldelli, S. (2020). Nutrient Recapture from insect farm waste: Bioconversion with Hermetia illucens (L.) (Diptera: Stratiomyidae). Sustainability, 12(1), 362. https:// doi​.org​/10​.3390​/su12010362 Kibler, K. M., Reinhart, D., Hawkins, C., Motlagh, A. M., & Wright, J. (2018). Food waste and the food-energy-water nexus: A review of food waste management alternatives. Waste Management, 74, 52–62. https://doi​.org​/10​ .1016​/j​.wasman​.2018​.01​.014 Komakech, A. J., Sundberg, C., Jönsson, H., & Vinnerås, B. (2015). Life cycle assessment of biodegradable waste treatment systems for sub-Saharan African cities. Resources, Conservation and Recycling, 99, 100–110. https://doi​.org​/10​.1016​/j​.resconrec​.2015​.03​.006 Lenaerts, S., van der Borght, M., Callens, A., & van Campenhout, L. (2018). Suitability of microwave drying for mealworms (Tenebrio molitor) as alternative to freeze drying: Impact on nutritional quality and colour. Food Chemistry, 254, 129–136. https://doi​.org​/10​.1016​/j​.foodchem​.2018​.02​.006 Mankins, J. C. (1995). Technology readiness levels. White Paper, April, 6. Mertenat, A., Diener, S., & Zurbrügg, C. (2019). Black soldier fly biowaste treatment – Assessment of global warming potential. Waste Management, 84, 173–181. https://doi​.org​/10​.1016​/j​.wasman​.2018​.11​.040 Mungkung, R., Aubin, J., Prihadi, T. H., Slembrouck, J., van der Werf, H. M. G., & Legendre, M. (2013). Life cycle assessment for environmentally sustainable aquaculture management: A case study of combined aquaculture systems for carp and tilapia. Journal of Cleaner Production, 57, 249–256. https://doi​.org​/10​.1016​/j​.jclepro​.2013​ .05​.029 Ojha, S., Bußler, S., Psarianos, M., Rossi, G., & Schlüter, O. K. (2021). Edible insect processing pathways and implementation of emerging technologies. Journal of Insects as Food and Feed, 7(5), 877–900. https://doi​.org​/10​ .3920​/JIFF2020​.0121

30  Edible Insects Processing for Food and Feed Oonincx, D. G. A. B., & de Boer, I. J. M. (2012). Environmental impact of the production of mealworms as a protein source for humans—A life cycle assessment. PLOS ONE, 7(12), e51145. https://doi​.org​/10​.1371​/journal​.pone​ .0051145 Parajuli, R., Knudsen, M. T., Birkved, M., Djomo, S. N., Corona, A., & Dalgaard, T. (2017). Environmental impacts of producing bioethanol and biobased lactic acid from standalone and integrated biorefineries using a consequential and an attributional life cycle assessment approach. Science of the Total Environment, 598, 497–512. https://doi​.org​/10​.1016​/j​.scitotenv​.2017​.04​.087 Pérez-Camacho, M. N., Curry, R., & Cromie, T. (2018). Life cycle environmental impacts of substituting food wastes for traditional anaerobic digestion feedstocks. Waste Management, 73, 140–155. https://doi​.org​/10​.1016​/j​.wasman​.2017​.12​.023 Poveda, J., Jiménez-Gómez, A., Saati-Santamaría, Z., Usategui-Martín, R., Rivas, R., & García-Fraile, P. (2019). Mealworm frass as a potential biofertilizer and abiotic stress tolerance-inductor in plants. Applied Soil Ecology, 142, 110–122. https://doi​.org​/10​.1016​/j​.apsoil​.2019​.04​.016 Purschke, B., Stegmann, T., Schreiner, M., & Jäger, H. (2017). Pilot-scale supercritical CO2 extraction of edible insect oil from Tenebrio molitor L. larvae - Influence of extraction conditions on kinetics, defatting performance and compositional properties. European Journal of Lipid Science and Technology, 119(2), 1600134. https://doi​.org​/10​.1002​/ejlt​.201600134 Ravindran, R., & Jaiswal, A. K. (2019). Wholesomeness and safety aspects of irradiated foods. Food Chemistry, 285, 363–368. https://doi​.org​/10​.1016​/j​.foodchem​.2019​.02​.002 Rebitzer, G. (2005). Enhancing the Application Efficiency of Life Cycle Assessment for Industrial Uses. https://doi​ .org​/10​.5075​/epfl​-thesis​-3307 Roffeis, M., Almeida, J., Wakefield, M., Valada, T., Devic, E., Koné, N., Kenis, M., Nacambo, S., Fitches, E., Koko, G., Mathijs, E., Achten, W., & Muys, B. (2017). Life cycle inventory analysis of prospective insect based feed production in West Africa. Sustainability, 9(10), 1697. https://doi​.org​/10​.3390​/su9101697 Roffeis, M., Fitches, E. C., Wakefield, M. E., Almeida, J., Alves Valada, T. R., Devic, E., Koné, N., Kenis, M., Nacambo, S., Koko, G. K. D., Mathijs, E., Achten, W. M. J., & Muys, B. (2020). Ex-ante life cycle impact assessment of insect based feed production in West Africa. Agricultural Systems, 178, 102710. https://doi​.org​ /10​.1016​/j​.agsy​.2019​.102710 Roffeis, M., Muys, B., Almeida, J., Mathijs, E., Achten, W. M. J., Pastor, B., Velásquez, Y., Martinez-Sanchez, A. I., & Rojo, S. (2015). Pig manure treatment with housefly (Musca domestica) rearing – An environmental life cycle assessment. Journal of Insects as Food and Feed, 1(3), 195–214. https://doi​.org​/10​.3920​/JIFF2014​.0021 Roncolini, A., Milanović, V., Aquilanti, L., Cardinali, F., Garofalo, C., Sabbatini, R., Clementi, F., Belleggia, L., Pasquini, M., Mozzon, M., Foligni, R., Federica Trombetta, M., Haouet, M. N., Serena Altissimi, M., di Bella, S., Piersanti, A., Griffoni, F., Reale, A., Niro, S., & Osimani, A. (2020). Lesser mealworm (Alphitobius diaperinus) powder as a novel baking ingredient for manufacturing high-protein, mineral-dense snacks. Food Research International, 131, 109031. https://doi​.org​/10​.1016​/j​.foodres​.2020​.109031 Saerens, W., Smetana, S., van Campenhout, L., Lammers, V., & Heinz, V. (2021). Life cycle assessment of burger patties produced with extruded meat substitutes. Journal of Cleaner Production, 306. https://doi​.org​/10​.1016​ /j​.jclepro​.2021​.127177 Salomone, R., Saija, G., Mondello, G., Giannetto, A., Fasulo, S., & Savastano, D. (2017). Environmental impact of food waste bioconversion by insects: Application of life cycle assessment to process using Hermetia illucens. Journal of Cleaner Production, 140, 890–905. https://doi​.org​/10​.1016​/j​.jclepro​.2016​.06​.154 Schau, E. M., & Fet, A. M. (2008). LCA studies of food products as background for environmental product declarations. The International Journal of Life Cycle Assessment, 13(3), 255–264. https://doi​.org​/10​.1065​/ lca2007​.12​ .372 Shorstkii, I., Comiotto Alles, M., Parniakov, O., Smetana, S., Aganovic, K., Sosnin, M., Toepfl, S., & Heinz, V. (2020). Optimization of pulsed electric field assisted drying process of black soldier fly (Hermetia illucens) larvae. Drying Technology, 1–9. https://doi​.org​/10​.1080​/07373937​.2020​.1819825 Sindermann, D., Heidhues, J., Kirchner, S., Stadermann, N., & Kühl, A. (2021). Industrial processing technologies for insect larvae. Journal of Insects as Food and Feed, 7(5), 857–875. https://doi​.org​/10​.3920​/JIFF2020​.0103 Slorach, P. C., Jeswani, H. K., Cuéllar-Franca, R., & Azapagic, A. (2019). Environmental and economic implications of recovering resources from food waste in a circular economy. Science of the Total Environment, 693, 133516. https://doi​.org​/10​.1016​/j​.scitotenv​.2019​.07​.322 Smetana, S. (2020). Life cycle assessment of specific organic waste-based bioeconomy approaches. Current Opinion in Green and Sustainable Chemistry. https://doi​.org​/10​.1016​/j​.cogsc​.2020​.02​.009 Smetana, S., Aganovic, K., Irmscher, S., & Heinz, V. (2018). Agri-food waste streams utilization for development of more sustainable food substitutes. In Designing Sustainable Technologies, Products and Policies (pp. 145– 155). Cham: Springer.

2  •  Environmental Impact of Edible Insect Processing  31 Smetana, S., Ashtari Larki, N., Pernutz, C., Franke, K., Bindrich, U., Toepfl, S., & Heinz, V. (2018). Structure design of insect-based meat analogs with high-moisture extrusion. Journal of Food Engineering, 229, 83–85. https:// doi​.org​/10​.1016​/j​.jfoodeng​.2017​.06​.035 Smetana, S., Leonhardt, L., Kauppi, S.-M., Pajic, A., & Heinz, V. (2020). Insect margarine: Processing, sustainability and design. Journal of Cleaner Production, 264, 121670. https://doi​.org​/10​.1016​/j​.jclepro​.2020​.121670 Smetana, S., Mathys, A., Knoch, A., & Heinz, V. (2015). Meat alternatives: Life cycle assessment of most known meat substitutes. The International Journal of Life Cycle Assessment, 20(9), 1254–1267. https://doi​.org​/10​ .1007​/s11367​- 015​- 0931-6 Smetana, S., Mhemdi, H., Mezdour, S., & Heinz, V. (2020). Pulsed electric field–treated insects and algae as future food ingredients. In Pulsed Electric Fields to Obtain Healthier and Sustainable Food for Tomorrow (pp. 247–266). Elsevier. https://doi​.org​/10​.1016​/ B978​- 0​-12​-816402​- 0​.00011-2 Smetana, S., Palanisamy, M., Mathys, A., & Heinz, V. (2016). Sustainability of insect use for feed and food: Life cycle assessment perspective. Journal of Cleaner Production, 137, 741–751. https://doi​.org​/10​.1016​/j​.jclepro​ .2016​.07​.148 Smetana, S., Pernutz, C., Toepfl, S., Heinz, V., & van Campenhout, L. (2019). High-moisture extrusion with insect and soy protein concentrates: Cutting properties of meat analogues under insect content and barrel temperature variations. Journal of Insects as Food and Feed, 5(1), 29–34. https://doi​.org​/10​.3920​/JIFF2017​.0066 Smetana, S., Profeta, A., Voigt, R., Kircher, C., & Heinz, V. (2021). Meat substitution in burgers: Nutritional scoring, sensorial testing, and life cycle assessment. Future Foods, 100042. https://doi​.org​/10​.1016​/j​.fufo​.2021​.100042 Smetana, S., Schmitt, E., & Mathys, A. (2019). Sustainable use of Hermetia illucens insect biomass for feed and food: Attributional and consequential life cycle assessment. Resources, Conservation and Recycling, 144, 285–296. https://doi​.org​/10​.1016​/j​.resconrec​.2019​.01​.042 Smetana, S., Spykman, R., & Heinz, V. (2021). Environmental aspects of insect mass production. Journal of Insects as Food and Feed, 7(5), 553–571. https://doi​.org​/10​.3920​/JIFF2020​.0116 Spykman, R., Hossaini, S. M., Peguero, D. A., Green, A., Heinz, V., & Smetana, S. (2021). A modular environmental and economic assessment applied to the production of Hermetia illucens larvae as a protein source for food and feed. The International Journal of Life Cycle Assessment, 26(10), 1959–1976. https://doi​.org​/10​.1007​/ s11367​- 021​- 01986-y Steubing, B., Mutel, C., Suter, F., & Hellweg, S. (2016). Streamlining scenario analysis and optimization of key choices in value chains using a modular LCA approach. The International Journal of Life Cycle Assessment, 21(4), 510–522. https://doi​.org​/10​.1007​/s11367​- 015​-1015-3 Suckling, J., Druckman, A., Moore, C. D., & Driscoll, D. (2020). The environmental impact of rearing crickets for live pet food in the UK, and implications of a transition to a hybrid business model combining production for live pet food with production for human consumption. The International Journal of Life Cycle Assessment, 25(9), 1693–1709. https://doi​.org​/10​.1007​/s11367​- 020​- 01778-w Surendra, K. C., Olivier, R., Tomberlin, J. K., Jha, R., & Khanal, S. K. (2016). Bioconversion of organic wastes into biodiesel and animal feed via insect farming. Renewable Energy, 98, 197–202. https://doi​.org​/10​.1016​/j​.renene​ .2016​.03​.022 Thabrew, L., Wiek, A., & Ries, R. (2009). Environmental decision making in multi-stakeholder contexts: Applicability of life cycle thinking in development planning and implementation. Journal of Cleaner Production, 17(1), 67–76. https://doi​.org​/10​.1016​/j​.jclepro​.2008​.03​.008 Thomas, C., Grémy-Gros, C., Perrin, A., Symoneaux, R., & Maître, I. (2020). Implementing LCA early in food innovation processes: Study on spirulina-based food products. Journal of Cleaner Production, 268, 121793. https://doi​.org​/10​.1016​/j​.jclepro​.2020​.121793 Tzompa-Sosa, D. A., Yi, L., van Valenberg, H. J. F., & Lakemond, C. M. M. (2019). Four insect oils as food ingredient: Physical and chemical characterisation of insect oils obtained by an aqueous oil extraction. Journal of Insects as Food and Feed, 1–14. https://doi​.org​/10​.3920​/JIFF2018​.0020 Udugama, I. A., Petersen, L. A. H., Falco, F. C., Junicke, H., Mitic, A., Alsina, X. F., Mansouri, S. S., & Gernaey, K. v. (2020). Resource recovery from waste streams in a water-energy-food nexus perspective: Toward more sustainable food processing. Food and Bioproducts Processing, 119, 133–147. https://doi​.org​/10​.1016​/j​.fbp​.2019​ .10​.014 Ulmer, M., Smetana, S., & Heinz, V. (2020). Utilizing honeybee drone brood as a protein source for food products: Life cycle assessment of apiculture in Germany. Resources, Conservation and Recycling, 154, 104576. https:// doi​.org​/10​.1016​/j​.resconrec​.2019​.104576 van Huis, A. (2021). Prospects of insects as food and feed. Organic Agriculture, 11(2), 301–308. https://doi​.org​/10​ .1007​/s13165​- 020​- 00290-7 van Huis, A., & Oonincx, D. G. A. B. (2017). The environmental sustainability of insects as food and feed. A review. Agronomy for Sustainable Development, 37(5), 43.

32  Edible Insects Processing for Food and Feed Wang, H., Rehman, K. ur, Liu, X., Yang, Q., Zheng, L., Li, W., Cai, M., Li, Q., Zhang, J., & Yu, Z. (2017). Insect biorefinery: A green approach for conversion of crop residues into biodiesel and protein. Biotechnology for Biofuels, 10(1), 304. https://doi​.org​/10​.1186​/s13068​- 017​- 0986-7 Zampori, L., Saouter, E., Schau, E., Cristobal Garcia, J., Castellani, V., & Sala, S. (2016). Guide for Interpreting Life Cycle Assessment Result. Luxembourg: Publications Office of the European Union. Zheng, L., Hou, Y., Li, W., Yang, S., Li, Q., & Yu, Z. (2013). Exploring the potential of grease from yellow mealworm beetle (Tenebrio molitor) as a novel biodiesel feedstock. Applied Energy, 101, 618–621. https://doi​.org​/10​.1016​ /j​.apenergy​.2012​.06​.067 Zorrilla, M., & Robin, N. (2019). Nutrition technologies: Offering price competitive black soldier fly protein and oil to the animal feed and pet food sectors. Industrial Biotechnology, 15(6), 328–329. https://doi​.org​/10​.1089​/ind​ .2019​.29195​.mzo

3

Legislation Nils Th. Grabowski Contents

3.1 Disclaimer 33 3.2 Regulatory Aspects 34 3.2.1 Introduction 34 3.2.2 Traditional Entomophagy – Insects as Food 35 3.2.3 Traditional Entomophagy – Insects for Feed 36 3.2.4 Non-Traditional Entomophagy – Insects for Food 36 3.2.5 Non-Traditional Entomophagy – Insects for Feed 39 3.3 Practical Application of Legislation 40 3.3.1 Initial Steps 40 3.3.2 Traditional Production Systems 41 3.3.2 Non-traditional Production Systems 41 3.4 Possible Future Developments and Needs for Regulation 41 Notes 51 References 52

3.1 DISCLAIMER First, any contribution to a legal subject is condemned to temporality. In this way, information presented here may already be outdated when consulting this contribution, and the reader is strongly advised to scrutinize the content before relying on it. Then, legislation varies with each country, even each region, and not all relevant texts are available via the internet. Besides, the complete description on what every country resp. region regulates would exceed the margin given here. In this way, a method to approach the issue of productive insect legislation is provided rather than a comprehensive description of all laws pertaining to productive insects. Finally, only approbated legal practitioners are entitled to provide legal counselling. As the author is not a lawyer, he recommends consulting such a practitioner in any case where legal aspects of insect use are addressed.

DOI: 10.1201/9781003165729-3

33

34  Edible Insects Processing for Food and Feed

3.2  REGULATORY ASPECTS 3.2.1 Introduction Although it can be expected that regulations were applied in human societies from their origins on, the oldest preserved legal documents were found in Mesopotamia and dated from approx. 2100 BCE–2050 BCE. Since then, sets of general codes of conduct regulating human societies have been established and accepted by them. Among the several branches of public law, health law is focused on the different aspects of ensuring the health of a nation’s population within societal limits and norms. Public health law is part of this law to protect people from disease and promote their health. Correspondingly, and within public health law, food law specifically deals with providing a framework to protect the population from harmful foodstuffs (that may imperil health) and consumer deception, setting a basis for the evaluation of the quality of the foodstuffs that are on the market (Gostin et al., 2016). Food law is considered a cross-section legislation as it includes consumer protection, hazard control, and trade law.1 It ideally covers all stages of the production chain, from the primary production and associated areas (e.g. feeding, housing, animal health and welfare, etc.) to processing, storage, logistics, marketing, and public health. From the moment of emergence as a taxonomic group more than 400 million years ago, insects have been on the menu. Taxa emerging later, like primates (some 55 million years ago), have evolved to eat insects, and hominids including the modern Homo sapiens are no exception to the rule. The amount of modern food regulations varies markedly, from countries with a food hygiene code and some derived regulations to those with a highly complex set of regulations that could have become as specific as the so-called “cucumber regulation” (REG [EC] 1766/882, abolished in 2009), which set a standards as to how curved a cucumber can be. Surprisingly, and despite their consumption over millennia (cucumbers have been cultivated “only” for over 3,000 years), edible insects have virtually been almost non-existent in food law, even in countries in which insects are relatively common foodstuffs. Why is that so, and why are regulations pertaining to edible insects emerging in the last decade increasing – at least in comparison to the previous millennia – at an unprecedented speed? Ideally, legislation always seeks to create a common set of rules in case the need arises. This is the key to understanding the situation that current food insect legislation is in. Seen globally, there are two types of entomophagy that have an impact on food insect legislation worldwide, i.e. traditional and non-traditional one. • Traditional entomophagy has developed over the millennia, and typically implies catching wild-ranging insects from crops, in nearby areas or the further afield; transporting them home; processing them; and either eating them at home or selling them on local markets. • Non-traditional entomophagy, which has developed in recent decades (although the idea and basic studies for it have been performed from last century on), relies typically on farming, and insects are harvested and processed in enterprises of various sizes from small to large. This type also includes producing insects for other purposes, e.g. feed. Although there seems to be a clear difference between the two, the reality is different. Processing red palm weevil larvae (Rhynchophorus ferruguineus) so that they can be exported from Thailand to other countries is one example of how traditional entomophagy is profiting from advances in food hygiene as developed by science, and farming silkworms (Bombyx mori) is a traditional yet modernized agricultural activity. Thus the traditional and non-traditional may be blending into each other, depending on the species and the production system they are subjected to. Gathering the weaver ant Oecophylla smaragdina for personal consumption in Papua New Guinea (Meyer-Rochow, 2005) and producing tons of black

3 • Legislation  35 soldier flies (Hermetia illucens) weekly3 are probably extreme examples of how diverse insect production can be nowadays. Parallel to the diversity in production systems is the great diversity of edible insects themselves. Out of the approximately one million species of insect, some 2,000–3,000 are known to be edible and consumed, at least traditionally. To put this in context, currently there are some 5,000 mammal, 18,000 avian, 11,000 reptile, and 32,500 fish species, and of those vertebrates, some 1,000 species are consumed regularly. In addition, the traditional use of these insects is highly local and specific in order to suit specific needs, risks, and social and cultural frame conditions. Some species are consumed after heating, and some raw (or even alive); some consumption traditions involve removing the wings before eating, some even the antennae; in some cases, only certain insect species or taxa may be consumed, while others are banned for cultural reasons (Grabowski, 2017). All in all, edible insect consumption is a complex area, and this complexity may be one reason that generalizing about “insects” may be so difficult. Within other animal-based foodstuffs, only fishery products may pose a similar complexity, while meat, dairy, and honey products are also diversified (particularly dairy products), but more “general rules” in terms of food safety apply to them, which makes handling amd regulating them easier.

3.2.2 Traditional Entomophagy – Insects as Food Before the “entomophagy movement” started, eating insects had been restricted almost exclusively to the traditionalists, i.e., societies beyond the borders of the Muslim and Western civilizations. In the Mid-East, which is the cradle of the three large monotheistic religions, orthopterans (crickets, grasshoppers, locusts etc.) seem to have been a popular foodstuff, just as in other parts of the world. The Jewish food law (kashrut) bans all insects except four types of orthopterans as “unclean”, limiting entomophagy to a set of what is believed to be crickets, grasshoppers, and locusts. Islam also includes food guidelines, and entomophagy is restricted exclusively to locusts as stated in the Sunna (Grabowski, 2017). Christianity does not have religious food taboos, but the role some insects play in the bible seems to have affected the relations between humans and insects. Three of the seven plagues of Egypt were related to insects, and one of the demons associated with the devil is known as the “lord of the flies” (“Beelzebub”; Bromiley, 2002). Europe practised entomophagyat one time, e.g. in the ancient world, but abandoned this habit largely after Christianization, although not completely (Payne et al., 2019). Insects have been associated with the so-called 3 Ds – “dirt, disease, and death” (Dagevos, 2021). Once Muslims and Christians started exploring the planet and conquered a large part of it, the civilization (as they conceived it in those times) they exported also affected the culinary customs of the countries they occupied, and food habits they rejected out of their own traditions were discouraged actively or passively. Where entomophagous societies remained independent, entomophagy as practised by the majority of its population continued. In countries that were subject to the European conquistas and imperialism, this food habit either disappeared completely or was marginalised to those sectors that in terms of power were not relevant in colonial or postcolonial times, e.g. indigenous tribes in the Americas (Ramos Elorduy and Pino Moreno, 1989, von Paczensky and Dünnebier, 1994). In this way, eating insects was considered retrograde, and this could be a reason why edible insects have not found their way into most of the world’s food legislation. If that were the case, then they would appear explicitly in the laws of the countries relatively untouched by foreign ambitions to rule them. Afghanistan, Bhutan, China, Ethiopia, Iran, Saudi Arabia, Nepal, Thailand,4 and, in a broader sense, Japan are considered the modern states that were not subjected to foreign colonization. Of those, all but Afghanistan and Bhutan are mentioned in the list of Jongema (2017),5 i.e. the consumption of at least one species of edible insect has been recorded so far. Surveying FAOLEX,6 which is a platform for agriculture-related regulations from the Food and Agriculture Organization (FAO), it was seen that only China makes reference to edible insects in the subsection “food and nutrition”, specifically B. mori, which is one of the 224 species listed by Jongema (2017) for that country. This is

36  Edible Insects Processing for Food and Feed particularly interesting in the case of Thailand, a hotspot in terms of entomophagy with not only a large amount of insects (284) listed as edible (Jongema, 2017), but also an extensive set of food regulations (116). It is more common to include insects as contaminants, e.g. in Bhutan, China, and Nepal. Explaining this becomes speculation. On one hand, the food law of a given country may not be as detailed and regulations simply apply to all foodstuffs including insects. This may be valid in most countries that work with a definition of “foodstuff”. In a survey pertaining to African countries (Grabowski et al., 2020), no definition of the term could be encountered in 20 out of 54 countries. When searching food law, care must be taken when working with definitions. Although most foodstuff definitions contain “animals” as one possible source of food, “animals” by themselves may be defined differently from what would be expected. To give an example, the definition of “food” in the Ugandan food and drug act does not include birds nor fish. On the other hand, edible insects may not be included in nations practising traditional entomophagy simply because there was no need to regulate the market as it is either too small resp. too local, or because no major food-borne disease outbreaks have been associated with edible insects so far in that country. Finally, indifference and/or negligence are other possible explanations.

3.2.3 Traditional Entomophagy – Insects for Feed Not much can be said about traditional methods for feeding animals with insects. Some pet owners may collect insects for their insectivore pets. In extensive production systems, particularly poultry, pigs, and animals farmed in aquaculture, eventually catch and eat insects in a natural way. Thus, actively supplying insects to livestock seems an uncommon tradition (Menzel and D’Alusio, 1998), because collecting insects from the wild can be tedious and time consuming, and on a global scale it may be questionable that large amounts of farm animals could be fed with wild-caught insects on a wider and more consistent basis. An exception to this rule are swarming locusts, which can be gathered with less effort (if the livestock animals do not catch them themselves). The only insects with a constant supply are bees (Apis mellifera) and the silkworm (B. mori). One of the few examples is feeding bee larvae to chicken or fish. This has become more popular as eliminating drone brood is one way to control varroosis, an economically important bee disease. While many beekeepers throw away the honeycombs with drone larvae or leave them in the open so that wild birds eat them (possibly spreading American foulbrood), cooking and using them as feed for (mostly) the beekeeper’s own animals (even as fish bait) is a sustainable option which has been practised in private and on a small scale (Naber, 2016). No information on the regulatory aspect could be encountered. If not used as a foodstuff, silkworm pupae can also be used as feedstuff. It has a high nutrient value, but unprocessed pupae spoil quickly (Wei et al., 2009), and it may be suggested that this was one of the main reasons (along with a possibly low palatability) that silkworm pupae have not been used on a broader scale.

3.2.4 Non-Traditional Entomophagy – Insects for Food Until recently, countries with little or no insect consumption have not had the need to establish rules for edible insects. It may be speculated that the only way insect consumption takes place in these nations is a) via those introducing them from abroad, either because they are migrants living in that country that want to have a “taste of home” or returning tourists that got acquainted with entomophagy abroad and brought back a souvenir; b) aficionados that e.g. purchase feed insects from pet shops or rear them at home, and consume them with their nearest and dearest; and c) those few consuming the last remnants of traditional insect dishes, e.g. European cheeses in which piophilid cheese flies contribute to their ripening. From the legal point of view, it must be stressed that buying feed insects at pet shops is a rededication; eating them is, from the regulatory point of view, the same as eating canned cat food. From the safety point of view, it may be stated that buying living insects means buying them as fresh as possible, but the quality and safety

3 • Legislation  37 must be at least doubted. Many of the typical feed insects such as crickets, locusts, and mealworms reflect the quality of the feeds they were offered, and depending on the national feedstuff regulation, the feedstuffs may be controlled or not. In a worst-case scenario, feed offered to the feed insects is of low quality and/ or contaminated with microbes or chemical substances that affect human health. Consuming fly-ripened cheeses from European countries along the Mediterranean was considered rural (in contrast to emerging urban lifestyle) and therefore retrograde, and was even banned in Italy when diary product legislation did not consider the idiosyncrasies of e.g. the casu marzu and banned it from the market (Payne et al., 2019). According to Payne et al. (2019), non-traditional entomophagy started in the universities. Although eating insects was inconceivable for Western civilization, and at best a somewhat strange tradition of peoples from far away, anthropologists still recorded it, assessing what tribe consumed which species under which circumstances. They would surely have tried them during their fieldwork, discovering their benefits, i.e. in terms of taste and nutrition.7 Scientific evidence of these benefits emerged step by step, and opinions thematizing the also currently non-entomophagous societies should start eating insects have been published from the late 20th century on. Putting this into practice, entomologists interested in entomophagy started to connect and even organize tasting events. All this culminated in the early 2010s when the FAO recognized insects’ potential to address the challenges facing the human population; the first congress on food and feed insects, held in The Netherlands in 2014, was the “big bang” for many stakeholders in the non-traditional sector, but also raised marked interest in those dealing with traditional entomophagy. Continent-wide lobby organizations, e.g. the International Platform for Insects as Food and Feed (IPIFF) in Europe or the North American Coalition for Insect Agriculture (NACIA), emerged, and farmers interested in rearing productive insects (i.e. those raised for a specific purpose like food, feed, industry, waste management etc.) started to put pressure on national governments. As an example, in Europe this process started in 2014 in Belgium. The Scientific Committee of the Federal Agency for the Food Chain and Hoge Gezondheitsraad published the “Common Advice SciCom 14-2014 and SHC Nr. 9160”, a scientific opinion about the food safety aspects of 12 insect species in which recommendations for the Belgian insect producers were laid down. Similar papers were also published for e.g. Austria (2017), Denmark (2017), Germany (2016/2019), Finland (2018), France (2015), Norway (2016), The Netherlands (2014), Portugal (2017), Spain (2018), and Switzerland (2017). Technically speaking, these papers were recommendations rather than official regulations, although most of them were issued by government departments. These recommendations usually start with a reference to the EU novel food regulation in its currently valid version and a disclaimer that these recommendations are applicable until a corresponding EU document will have been published. Then it contains a list of the insect species to which the recommendation can be applied along with a statement that all general regulations used in food and primary production also apply to insects. The final sections refer to specifications for edible insects, typically regarding farming, feeding, heating, storage, labelling, and trade (import/export; Grabowski et al., 2019). In these first years, some changes in these documents were made, e.g. the reduction of allowed species to three in some countries, i.e. the house cricket (Acheta domesticus), the migratory locust (Locusta migratoria), and the yellow mealworm (Tenebrio molitor; Jansen and Grabowski, 2019). In time it became apparent that with productive insects, a new agricultural branch was evolving, and that the lack of a reliable legal framework hampered this development (de Magistris et al., 2015). The European Food Safety Authority (EFSA), the foremost institution for public health issues regarding the EU, reacted in 2015 with a scientific opinion on edible insects (Table 3.1). The basic conclusion the EFSA reached was that with the scientific evidence available at the time of writing, edible insects do not pose a larger risk than other, more conventional foodstuffs of animal origin. It points out that more research is necessary to evaluate the specific risks, particularly the chemical ones. For readers not familiar with EU food legislation, EU food laws have a double origin. On one hand, there is a set of regulations issued by the EU which are directly valid for all member states. The so-called “hygiene package” is the basis for this growing apparatus and comprises the principles for foods and feeds in general, i.e. regulation (REG) (EC) 178/20028; the one for foods in general (REG 852/20049); the one for animal-derived foods (REG 853/200410); the one for official controls (REG 2017/62511); and the one for microbial criteria of foodstuffs (REG 2073/200512), each of them in a constant process of amending. The

38  Edible Insects Processing for Food and Feed TABLE 3.1  Summary of risks evaluated by the EFSA scientific opinion on edible insects (2015) SECTION Microbiology

RISKS Prions

Species barrier, vector?

Viruses

Species barrier?

Bacteria

Entomopathogens, ordinary flora s. ordinary livestock Relevant for farmed insects? Substrate-depending ibid. Substrate, treatment Substrate-depending Cross-reactions Spoilage Similar to ordinary livestock

Fungi Parasites Chemistry

Allergens Processing Environment

PROFILE

Heavy metals Toxins Veterinary drugs others Tropomyosin Bacterial counts Farming

RISK EVALUATION

UNCERTAINTIES

≤, when not fed with risk material ≤, when fed with permitted foodstuffs and efficient heat treatments are applied ibid.

Unsufficient data Viral metabolites?

ibid. ibid.

ibid. ibid.

? ? ? ? +; labelling! +; no raw cosumption ≤; manure control

ibid. ibid. ibid. ibid. ibid. ibid. ibid.

Insufficient data

same is true for associated regulations. For practical matters, the last three regulations are the most important ones. However, this framework is sometimes general and deliberately leaves “gaps” in food legislation, gaps that are filled with national laws, which, naturally, are valid only in the member state they were issued by. A prominent example of how these laws interact is the consumption of raw milk. REG (EU) 853/2004 contains the general rules pertaining to dairy products and claims that each member state may (or may not) permit the consumption of raw milk. For Germany, the national hygiene regulation for animal-derived foodstuffs establishes that consuming raw milk is not allowed unless offered in two specified ways.13 With this hierarchy, it becomes clear why the first national papers regarding edible insects could be nothing more than interim recommendations, and although the scientific opinion of the EFSA in 2015 was a sign of the recognition of the growing importance of non-traditional entomophagy, this did not provide any legal base to work with. The way edible insects went in the EU was that of so-called novel foods, i.e. foodstuffs that have not been consumed in a given country to a large extent. The first version of the EU novel food regulation (REG 258/9714) created confusion in relation to insects as it was not clear whether all insect products would fit the definitions of novel food. In fact, insect extracts like protein or fats were novel foods in the sense of that regulation, while whole animals were not. The new version of the regulation (REG 2015/228315) ended this confusion by including all kinds of insect-based products as novel foods. With that, the most important point so far in EU food insect law was made. REG 2015/2283 establishes that a novel food can be certified as such and placed legally on the market if a stakeholder applies for it and provides sufficient proof of its safety. There are two administrative ways, but eventually, the commission has to decide on this application. While this process is ongoing, similar products already placed on the market may still be traded, but if the decision of the commission is rejection, they will have to be taken off the market. It should be stressed that the application is always a combination of a given species in a certain product type, and that the applicants present tailored evidence that this species/product combination is safe, meaning that for each new species/product combination, a new application must be handed in. Although the regulation foresees a schedule by which the application has to be decided, delays may emerge when questions and doubts turn up during the certifying process. In those situations, the official schedule is stopped and continues once the applicant has provided this missing information (Schiel et al., 2020).

3 • Legislation  39  This is why the first final decisions on insect-related novel food applications were issued in 2021, i.e. regarding T. molitor (yellow mealworm) and L. migratoria (migratory locust). Accordingly, the corresponding regulations (2021/882 resp. 2021/197516) include specifications as to • • • • • • • • •

The species/product combination. The specified food category the specification relates to. Maximum levels of the insect component in the complete foodstuff. Labelling requirements. Data protection issues. Chemical composition. Maximum permissible levels of contaminants. Microbiological criteria. Labelling (incl. allergens).

So far, both applicants opted for data protection, which basically concedes 5 years of patent protection. After that, data protection will expire and the species–product combination may be used by any entrepreneur.

3.2.5 Non-Traditional Entomophagy – Insects for Feed Using insects as feedstuff has become a common practice worldwide. Farming crickets, mealworms, and flies is the decisive difference between traditional and non-traditional approaches, as a constant supply is ensured, logistics make the transport of insects possible, and industrial methods are able to preserve them (e.g. dried or canned), so that shelf life is even increased. By itself, the pet food insect sector is well established and works efficiently.17 However, the main focus has been on insectivorous pets and zoo animals rather than on livestock. In fact, insects have been banned from livestock nutrition, most recently because of the BSE crisis at the turn of the millennium. The BSE prion originally infected sheep but skipped species frontiers when sheep that succumbed to this disease (maedi/visna) were not heated high enough to inactivate the prion; it was then used in bovine nutrition from where it skipped to humans, who also developed a (fatal) disease. In the EU, this event was a true stress test, and the current food legislation already described in the previous section originated because of this (Wissenschaftlicher Beirat der Bundesärztekammer, 2003). It is also no wonder that precaution became the leading element of this food law, and the corresponding EU legislation (REG 999/200118) established, in Article 7, “The feeding to ruminants of protein derived from animals shall be prohibited”. The abbreviation “PAP” (processed animal protein) emerged, and with this, insects were banned from the livestock nutrition sector, as the regulation referred to all livestock species. In fact, the leading opinion during the 2014 Wageningen congress was that it would be easier to allow insects onto the food market than onto the feed market, as the author witnessed in person. It should be stressed that it is unlikely that policy makers had insects in mind when writing this regulation, which was almost a decade before productive insects began to be promoted. As mentioned before, pigs, fish, and poultry consume insects when given the opportunity. Another obstacle to feeding insects to larger animal populations is biosecurity, if feeding with live insects is intended. There is no doubt that using living insects ensures a maximum of freshness, and being able to express predatory behaviour is beneficial (Calber and Albright, 2021), but biosecurity is the decisive factor of influence as feed insect species are also capable of transmitting production-related diseases and zoonoses, e.g. salmonellosis (e.g. Jensen et al., 2020). Another dimension may be added when the feed insect is an invasive species.19 While species like the black soldier fly (H. illucens) are considered largely inoffensive, escaping termites could cause serious problems. However, stakeholders promoted the use of insects as livestock feedstuffs further. Nutritional benefits have been described in many publications, e.g. DiGiacomo and Leury (2019). To cite the EU example, in a

40  Edible Insects Processing for Food and Feed first step, REG 2017/89320 allowed the use of seven species (A. domesticus, the banded cricket [Gryllodes sigillatus], the Jamaican field cricket [Gryllus assimilis], T. molitor, the buffalo worm [Alphitobius diaperinus], H. illucens, and the housefly [Musca domestica]). With this, these species become “farmed animals” as defined in the REG 1069/2009, destined for the production of insect-based PAP. Being so, all other rules for farmed animal feeding and nutrition apply. This means that these insects may not be fed with “ruminant proteins, catering waste, meat-and-bone meal, manure, nor faeces (REG 2017/893, REG 1069/2009,21 REG 767/200922). From the legal point of view and in view of the precautions taken to prevent a new outbreak of transmissible spongiform encephalites (TSE), this measure is comprehensible. As stated in the EFSA opinion, insects are not expected to develop TSE. The risk associated with these feedstuffs is rather that it cannot be guaranteed that while separating harvested insects from the feed substrate, the latter is eliminated completely. This why REG 2017/893 foresees using these insects as a source for PAP in aquaculture, while whole animals and other ways of processing them remains limited to the non-livestock sector. 2021 was an important year for insect production in Europe, not only because of the implementation of two insects as novel food, but also for REG 2021/1372.23 This is the second step in legalizing productive insects as feedstuff. Following their natural feeding behaviour, this regulation amends REG 2017/893 by permitting the use of insects in poultry and pig farming. Livestock farmers have welcomed the initiative.24

3.3  PRACTICAL APPLICATION OF LEGISLATION Although for the non-traditional entomophagy in the previous section, most examples were taken from EU legislation, the goal of this chapter is to enable every reader to explore the regulatory possibilities when intending to initiate an insect-rearing facility, either as a completely new business or as part of an already existing agriculture unit. This section will mainly consist of tables with questions that may be useful in orienting oneself in the regulatory framework each insect farmer will be subjected to. Apart from the purely legal issues, the tables also contain some questions pertaining to other important areas that may not have a legal base but should be evaluated using other information sources. In addition, other fields like economy, building, marketing, etc. round up the picture, but are not discussed in this contribution, although they are nevertheless important. In order to provide a quick reference, some questions may be repeated if they are relevant to different production systems. The base for the tables is the FAOLEX website,25 which seems to be the most comprehensive and updated list of regulatory acts in terms of food and feed. The different regulations are subsumed under certain headings like “Policies”, “Food and nutritions”, etc. For the tables, these headings were used as a way to orient the reader when she/he consults this website in search of a given issue. Of course, many governments also have corresponding websites, but experience has shown that a) not all governments or minsteries have them, b) the access to the different documents may be limited, c) laws are not published, or d) the sites are not updated periodically. As the FAOLEX may be affected by similar problems, caution is still advised, and the author emphasizes again that this chapter only provides a general orientation; the actual valid regulatory frame can only be guaranteed by an approved legal practitioner (see section 3.1).

3.3.1 Initial Steps Before starting, some basic questions must be asked, e.g.: • Will there be a market for what is being produced? • If not, can it be created? • If so, how?

3 • Legislation  41 • Are there funds available to subsist for the time while the product is getting established? • Does the food regulation allow the placement of food and/or feed insects on the market? If these questions can be answered affirmatively, then the process of exploring the specific regulatory framework begins. Then it is necessary to determine the basic production system that is intended to be used – traditional or non-traditional.

3.3.2 Traditional Production Systems Table 3.2 contains questions regarding the traditional use of food insects, while Table 3.3 refers to feed insects. As can be seen, a series of issues must be addressed if the traditional use of edible insects is to be practiced responsibly. Although using feed insects traditionally is typically restricted to feeding surplus reared insects to small local populations of livestock that will be consumed within the domestic environment and may go unnoticed by the authorities, a true commercial activity would be a different thing, and more so if the animals to be fed were livestock and not non-livestock (Table 3.3.). A thorough approach is advised here.

3.3.2 Non-traditional Production Systems Questions pertaining to non-traditional are summarized in Table 3.4​. Since there are more possibilities in terms of the species to be chosen (naturally occurring, foreign but already existing in the country, foreign and new in the country), the corresponding field contains more questions. The basic setting is to farm the insects and to place them on the market. Since non-traditional production systems are expected to have a larger economic impact and produce higher yields, and may reach more consumers than traditional ones, very close contact with the public health authorities is recommended, as they will be in charge of certifying farms and granting permits. Another excellent source for regulatory framework that is prone to be updated regularly are the websites of insect farming stakeholders, e.g. the IPIFF (International Platform for Insects as Food and Feed)29 in Europe. Farming feed insects will be a major branch within this new agricultural sector, because, among other reasons, non-traditional societies are more likely to consume ordinary livestock animals that have been fed with insects rather than eating the insects themselves. So far, producing insects for non-livestock has already developed, but rearing them for livestock species will add another dimension to it, and moreover a greater responsibility will be laid on the farmers. Depending on the country, requirements for feed production for livestock can be very strict, and the future will show if the processing technology will be able to cope with these standards.

3.4  POSSIBLE FUTURE DEVELOPMENTS AND NEEDS FOR REGULATION A look into the future can only be hypothetical, as the current pandemic has shown clearly how quickly frame conditions believed to be somewhat “eternal” can change within weeks. If the development of the insect sector continues, feed and food law will have to reply to the facts the sector is already producing. Although many start-up enterprises have already left the sector (not only because of less consumer acceptance than expected, but also because of regulatory uncertainties), many

42  Edible Insects Processing for Food and Feed TABLE 3.2  Key to exploring the regulatory framework in relation to the traditional use of food insects; terms in “Legal Information” were taken from FAOLEX homepage FIELD Species

Policies

Traditional use

Dealing with known problems

Land use

QUESTION

LEGAL INFORMATION

What edible insect species occur in your area naturally?

.

What other species are there, that could be productive in terms of feed, industry, etc.? Of those species of your interest, are there some that are protected, so that using them is not legal? Is/Are the contemplated species a pest species?

.

If so, are there any regulation prohibiting the use of the species? Does your enterprise reflect or oppose policy goals (see next question)? Does your enterprise promote rural development, sustainability, climate change effect mitigation, gender equality etc.?

Of those that occur, has there been a tradition of consuming them? What do you know about the traditional way of harvesting, processing, storing, and consuming them? What is known of flaws, problems, and obstacles that have occurred? If such exist, are there practical ways to overcome them? Are there any restrictions of making use of the land from which the animals will be taken (natural reserves, biospheres, agroforestry etc.)? Is it necessary to apply for a permit to collect the species? Does the natural population become affected when extracting a given percentage of specimens per time unit?

OTHER SOURCES Literature, information from the local population ibid.

Policies, wild species and ecosystems, international agreements (CITES)

.

Cultivated plants, disaster risk management, environment, forestry ibid.

Literature, information from the local population .

Policies, international agreements

.

Policies, agricultural and rural development, climate change, environment, food and nutrition, forestry, gender, social protection, water, wild species and ecosystems, international agreements .

Literature

.

Literature, information from the local population ibid.

.

ibid.

.

ibid.

Agricultural and rural development, environment, forestry, land and soil, wild species and ecosystems

.

Forestry, wild species and ecosystems ibid.

. Literature, information from the local population (Continued)

3 • Legislation  43 TABLE 3.2 (CONTINUED)  Key to exploring the regulatory framework in relation to the traditional use of food insects; terms in “Legal Information” were taken from FAOLEX homepage FIELD

Harvest

Post-primary production

QUESTION

LEGAL INFORMATION

OTHER SOURCES

Is the use of the selected species limited to a specific sector of the population, e.g. ethnic? Would the insect species fall under the definition of “game”? If so, what implications would that have for the intended use, particularly regarding the catching resp. harvesting method? Are there any legal requirements pertaining to killing the animals? If not, can you ensure killing according to animal welfare principles (i.e. freezing)? Do you have an effective approach to ensuring operators’ health (preventing zoonotic pathogens, dusts, allergies, physical injuries etc.)? Are there any legal requirements for transporting the living/dead animals? Are there any legal requirements for processing them (trimming, heating etc.)? Are there any legal requirements for storing them? Is there a need for veterinary inspection in terms of food safety (microbiological criteria, chemical composition, document soundness etc.)? What must be considered in terms of labelling (species, nutritional composition, allergens etc.)?

Forestry, social protection, wild species and ecosystems

.

ibid.

.

ibid.

.

Food and nutrition, forestry, livestock, wild species and ecosystems

.

Agricultural and rural development, food and nutrition

ibid.

Literature, information from the local population, public health authorities .

ibid.

.

ibid.

.

Food and nutrition

.

Food and nutrition

.

companies have evolved, and being organised in larger organisations, they are able to create, maintain, and even develop a market in which insects play an important role as feed and/or food. It can therefore be assumed that in the food sector, a more solid basis for placing different kinds of insect species (and products made thereof) on the market will be created. To cite Europe again, it may be expected that new food applications will be definitely decided next time. For those with inherent data protection, more farmers may engage in insect production once the 5-year period of data protection is over, and possibly the amount of new applications with no data protection may increase in years to come. If more insect-based foodstuffs are produced, prices may be expected to lower, and the total amount of insects consumed may rise. This translates into a need to establish a legal framework comparable to ordinary animal-based foodstuffs as already existing for meat, dairy, and fishery products.

44  Edible Insects Processing for Food and Feed TABLE 3.3  Key to exploring the regulatory framework in relation to the traditional use of feed insects; terms in “Legal Information” were taken from FAOLEX homepage FIELD Species

Policies

Traditional use Dealing with known problems

Land use

QUESTION

LEGAL INFORMATION

OTHER SOURCES

Is there any record that a wild-ranging species has been used as a feedstuff?

.

If it has, has it been used to feed livestock or non-livestock? Have traditionally reared insects been used as a feedstuff (honeybees, stingless bees, silkworms, cochineal,26 etc.)? If it was livestock, are there any regulations pertaining the use of feed insects for these livestock species? What other species are there that could be productive in terms of feed, industry, etc.? Of those species of your interest, are there some that are protected such that using them is not legal? Is/Are the contemplated species a pest species?

.

Literature, information from the local population ibid.

.

ibid.

Agricultural and rural development, food and nutrition .

.

If so, are there any regulations prohibiting the use of the species? Does your enterprise reflect or oppose policy goals (see next question)? Does your enterprise promote rural development, sustainability, climate change effect mitigation, gender equality etc.?

What do you know about the traditional way of harvesting, processing, storing, and consuming them? What is known of flaws, problems, and obstacles that have occurred? If such exist, are there practical ways to overcome them? Are there any restrictions concerning making use of the land from which the animals will be taken (natural reserves, biospheres, agroforestry etc.)? Is it necessary to apply for a permit to collect the species? Does the natural population become affected when extracting a given percentage of specimens per time unit?

Policies, wild species and ecosystems, international agreements (CITES) Cultivated plants, disaster risk management, environment, forestry ibid. Policies, international agreements Policies, agricultural and rural development, climate change, environment, food and nutrition, forestry, gender, social protection, water, wild species and ecosystems, international agreements .

Literature, information from the local population .

Literature, information from the local population . . Literature

.

Literature, information from the local population ibid.

.

ibid.

Agricultural and rural development, environment, forestry, land and soil, wild species and ecosystems Forestry, wild species and ecosystems ibid.

.

. Literature, information from the local population

(Continued)

3 • Legislation  45 TABLE 3.3 (CONTINUED)  Key to exploring the regulatory framework in relation to the traditional use of feed insects; terms in “Legal Information” were taken from FAOLEX homepage FIELD

Traditional insect farming

QUESTION

LEGAL INFORMATION

Is the use of the selected species limited to a specific sector of the population, e.g. ethnic? Do you need a permit to perform traditional insect rearing?

Forestry, social protection, wild species and ecosystems Agricultural and rural development, environment, forestry, wild species and ecosystems Agricultural and rural development, food and nutrition Agricultural and rural development, food and nutrition

Do you feed the insects in a way that complies with feedstuff regulations? Do you have an effective approach to ensuring operators’ health (preventing zoonotic pathogens, dusts, allergies, physical injuries etc.)? Harvest

Post-primary production

Are there any legal requirements pertaining catching resp. harvesting methods? Are there any legal requirements pertaining to killing the animals? If not, can you ensure killing according to animal welfare principles (i.e. freezing)? Are there any legal requirements for transporting the living/dead animals? Are there any legal requirements for processing them (trimming, heating etc.)? Are there any legal requirements for storing them? What must be considered in terms of labelling (species, nutritional composition, allergens etc.)? Do you require a permit to place feed insects on the market? Are there any specifications as to what insect species may be fed to which livestock species in which way and in what amounts? Does livestock fed with feed insects have to be labelled in a certain way?

Agricultural and rural development, food and nutrition Food and nutrition, forestry, livestock, wild species and ecosystems

OTHER SOURCES .

.

.

Literature, information from the local population, public health authorities .

Public health authorities

ibid.

.

ibid.

.

ibid.

.

Agricultural and rural development, food and nutrition Agricultural and rural development ibid.

.

. .

Food and nutrition

Countries with less elaborate food law may want to detail it further in the future, and insects could be included, in particular if amounts traded reach significant levels or may be relevant as an export good. For the feed sector, a similar tendency is possible. With the permission to use them also in the feed of some livestock species, this may grow even faster than the food insect sector. Section 3.3 has shown that there are many questions that should be addressed when engaging in using insects as food or feed. The author is aware that many questions may not be answered satisfyingly, simply because there is no regulatory basis on which to do so. The extent of these gaps varies with the issue and the country. For feed and food, the precaution principle is fundamental to ensure human health. In order to create a sound and fair regulatory framework for the insect sector, policy makers need more information. This

46  Edible Insects Processing for Food and Feed TABLE 3.4  Key to exploring the regulatory framework in relation to the non-traditional use of food insects; terms in “Legal Information” were taken from FAOLEX homepage FIELD Species

Policies

Bases from traditional uses

Land use

QUESTION

LEGAL INFORMATION

If you want to work with native species, which edible species occur in your area naturally?

.

Of those species of your interest, are there some that are protected such that using them is not legal? If you want to work with foreign species that already exist in the country (e.g. T. molitor, A. domesticus, G. sigillatus etc.), is it legal to farm them? If you want to work with foreign species that have not been introduced into your country so far, is it allowed to import and farm them? Is/Are the contemplated species a pest species?

Policies, wild species and ecosystems, international agreements (CITES) Agricultural and rural development, environment, food and nutrition, forestry, wild species and ecosystems Disaster risk management, environment, forestry, wild species and ecosystems, International Agreements Cultivated plants, disaster risk management, environment, forestry

If so, are there any regulations prohibiting the use of the species? If so, can you elaborate a biosecurity plan that effectively prevents the neozoa or pest species from escaping? Does your enterprise reflect or oppose policy goals (see next question)? Does your enterprise promote rural development, sustainability, climate change effect mitigation, gender equality, etc.?

ibid.

Of those that occur, has there been a tradition of consuming them?

What do you know about the traditional way of harvesting, processing, storing, and consuming them? What is known of flaws, problems, and obstacles that have occurred in traditional uses? If such exist, are there practical ways to overcome them? Are there any restrictions of making use of farmland?

Is it necessary to apply for a permit to farm the species?

OTHER SOURCES Literature, information from the local population .

.

Trade legislation

Literature, information from the local population .

Disaster risk management, environment

Literature

Policies, international agreements Policies, agricultural and rural development, climate change, environment, food and nutrition, forestry, gender, social protection, water, wild species and ecosystems, international agreements .

. Literature

.

Literature, information from the local population ibid.

.

ibid.

.

ibid.

Agricultural and rural development, environment, forestry, Land and soil, wild species and ecosystems Forestry, wild species and ecosystems

.

Public health authority

(Continued)

3 • Legislation  47 TABLE 3.4 (CONTINUED)  Key to exploring the regulatory framework in relation to the non-traditional use of food insects; terms in “Legal Information” were taken from FAOLEX homepage FIELD

Farming

Harvest

Post primary production

QUESTION

LEGAL INFORMATION

If you are working with native species, does the locality support periodic catching of some individuals to refresh to bloodlines? Do you meet the requirements to build and operate a farm? Do you know the biological cycle of the species to be farmed and do you have the means and ways to reproduce it under farming conditions? Have you developed a biosecurity plan to prevent both the escape of farmed insects and the entry of unwanted fauna (wild flies, predators, etc.)? Are there any official recommendations resp. regulations pertaining to insect farming? Do you need to register the farm unit? If you are both killing and processing the animals on your farm, do you need another permit27? Do you exclusively use permitted feedstuff to feed the insects? Do you have an effective approach to ensuring operators’ health (preventing zoonotic pathogens, dusts, allergies, physical injuries etc.)?

.

Are there any legal requirements pertaining to harvesting methods?

Agricultural and rural development, food and nutrition Food and nutrition, forestry, livestock, wild species and ecosystems

Are there any legal requirements pertaining to killing the animals? If not, can you ensure killing according to animal welfare principles (i.e. freezing)? Are there any legal requirements for transporting the living/dead animals? Are there any legal requirements for processing them (trimming, heating etc.)? Are there any legal requirements for storing them? Is there a need for veterinary inspection in terms of food safety (microbiological criteria, chemical composition, document soundness etc.)? What needs to be considered when selling the animals to an end consumer (via farm shop, e-commerce, export, etc.)? What must be considered in terms of labelling (species, nutritional composition, allergens etc.)?

Agricultural and rural development, environment .

Agricultural and rural development, environment, forestry, wild species and ecosystems Agricultural and rural development

OTHER SOURCES Literature, information from the local population Public health authority Literature, information from the local population ibid.

Public health authority

ibid. ibid.

ibid. ibid.

ibid.

ibid.

Agricultural and rural development, food and nutrition

Literature, information from the local population, public health authorities .

Public health authorities

ibid.

.

ibid.

.

ibid.

.

Food and nutrition

.

Food and nutrition

Trade legislation

Food and nutrition

.

48  Edible Insects Processing for Food and Feed TABLE 3.5  Key to exploring the regulatory framework in relation to non-traditional use of feed insects; terms in “Legal Information” were taken from FAOLEX homepage FIELD Species

Policies

Bases from traditional uses

Land use

QUESTION

LEGAL INFORMATION

If you want to work with native species, which feed species occur in your area naturally?

.

Of those species of your interest, are there some that are protected such that using them is not legal? If you want to work with foreign species that already exist in the country (e.g. T. molitor, A. domesticus, G. sigillatus etc.), is it legal to farm them? If you want to work with foreign species that have not been introduced into your country so far, is it allowed to import and farm them? Is/Are the contemplated species a pest species?

Policies, wild species and ecosystems, international agreements (CITES) Agricultural and rural development, environment, food and nutrition, forestry, wild species and ecosystems Disaster risk management, environment, forestry, wild species and ecosystems, international agreements Cultivated plants, Disaster risk management, environment, forestry

If so, are there any regulations prohibiting the use of the species? If so, can you elaborate a biosecurity plan that effectively prevents the neozoa or pest species from escaping? Does your enterprise reflect or oppose policy goals (see next question)? Does your enterprise promote rural development, sustainability, climate change effect mitigation, gender equality etc.?

ibid.

Of those that occur, has there been a tradition of using them as feedstuff?

What do you know about the traditional way of harvesting, processing, storing, and feeding them? What is known of flaws, problems, and obstacles that have occurred in traditional uses? If such exist, are there practical ways to overcome them? Are there any restrictions on making use of farmland?

Is it necessary to apply for a permit to farm the species?

OTHER SOURCES Literature, information from the local population .

.

Trade legislation

Literature, information from the local population .

Disaster risk management, environment

Literature

Policies, international agreements Policies, agricultural and rural development, climate change, environment, food and nutrition, forestry, gender, social protection, water, wild species and ecosystems, international agreements .

. Literature

.

Literature, information from the local population ibid.

.

ibid.

.

ibid.

Agricultural and rural development, environment, forestry, land and soil, wild species and ecosystems Forestry, wild species and ecosystems

.

Public health authority

(Continued)

3 • Legislation  49 TABLE 3.5 (CONTINUED)  Key to exploring the regulatory framework in relation to non-traditional use of feed insects; terms in “Legal Information” were taken from FAOLEX homepage FIELD

Farming

Harvest

Post primary production

QUESTION

LEGAL INFORMATION

If you are working with native species, does the locality support periodic catching of some individuals to refresh to bloodlines? Do you meet the requirements to build and operate a farm? Do you know the biological cycle of the species to be farmed and do you have the means and ways to reproduce it under farming conditions? Have you developed a biosecurity plan to prevent both the escape of farmed insects and the entry of unwanted fauna (wild flies, predators etc.)? Are there any official recommendations resp. regulations pertaining to insect farming? Do you need to register the farm unit? If you are both killing and processing the animals on your farm, do you need another permit?28 Do you exclusively use permitted feedstuff to feed the insects? Do you have an effective approach to ensuring operators’ health (preventing zoonotic pathogens, dusts, allergies, physical injuries etc.)?

.

Are there any legal requirements pertaining to harvesting methods?

Agricultural and rural development, food and nutrition Food and nutrition, forestry, livestock, wild species and ecosystems

Are there any legal requirements pertaining to killing the animals? If not, can you ensure killing according to animal welfare principles (i.e. freezing)? Are there any legal requirements for transporting the living/dead animals? Are there any legal requirements for processing them (trimming, heating etc.)? Are there any legal requirements for storing them? Is there a need for veterinary inspection in terms of food safety (microbiological criteria, chemical composition, document soundness etc.)? What must be considered in terms of labelling (species, nutritional composition, allergens etc.)? Is it allowed to sell the feed insects for use in livestock and non-livestock farming alike?

Agricultural and rural development, environment .

Agricultural and rural development, environment, forestry, wild species and ecosystems Agricultural and rural development

OTHER SOURCES Literature, information from the local population Public health authority Literature, information from the local population ibid.

Public health authority

ibid. ibid.

ibid. ibid.

ibid.

ibid.

Agricultural and rural development, food and nutrition

Literature, information from the local population, public health authorities .

Public health authorities

ibid.

ibid.

ibid.

ibid.

ibid.

ibid.

Food and nutrition

ibid.

ibid.

ibid.

Agricultural and rural development, food and nutrition

ibid.

(Continued)

50  Edible Insects Processing for Food and Feed TABLE 3.5 (CONTINUED)  Key to exploring the regulatory framework in relation to non-traditional use of feed insects; terms in “Legal Information” were taken from FAOLEX homepage FIELD

QUESTION Do you require a permit to place feed insects on the market? Are there any specifications as to what insect species may be fed to which livestock species in which way and in what amounts? Does livestock fed with feed insects have to be labelled in a certain way? What needs to be considered when selling the animals to an end consumer (pet owner, livestock farmer, via e-commerce, export etc.)?

LEGAL INFORMATION

OTHER SOURCES

Agricultural and rural development ibid.

ibid.

Agricultural and rural development, food and nutrition ibid.

ibid.

ibid.

Trade legislation

information is ideally supplied – at least in part – by the farmers that have first-hand experience of possible risks and pitfalls. This situation runs the risk of becoming a vicious circle; authorities are reluctant to certify insect farms because they do not have the tools to evaluate them and the possible risks associated with them, and entrepreneurs are reluctant to run a farm because they cannot judge if the authorities will issue corresponding permits and can therefore not create the information the authorities would need for appropriate certification. Here, science can act as a buffer between the two parties, creating practicerelated knowledge and interacting with them to foster communication and the development of regulations. For insects, the goal is to reach harmonization with other, more established feeds and foods. To do so, research is necessary; food and feed legislation for ordinary livestock has developed over the centuries, following scientific advancements and adapting to the changing socio-economic conditions, even creating new branches like animal welfare concerns in livestock production. Insects have been used for much longer than ordinary livestock species, but have remained largely unregulated because of a combination of low economic impact and low risks associated with them when using them traditionally. This situation has changed in the last decade, and modern productive insect stakeholders developing non-traditional approaches need a reliable framework in which to operate safely, as consumers have the right to consume safe foodstuffs that contain insects either directly or indirectly. Likewise, conditions regarding traditional entomophagy have also changed in the globalized world (in terms of transport, storage, chemical contaminants etc.), and the safety of consumption cannot be guaranteed as it may have been before. There are thousands of edible insects, and the number of potential feed insects has not been determined yet. As each species tells “its own story” and has an individual set of benefits and challenges, the research field is wide open. Some of the areas relevant to insect legislation that should be elucidated are: • Productive insect feeding and nutrition, particularly when using sustainable feeds like agricultural side-streams, and the risks associated with this practice • Productive insect physiology and pathology in order to prevent production diseases and endemics that endanger insect colonies and animal batches on farms • The impact of farming conditions (including building materials and surfaces in contact with the animals) on insect health • Animal welfare in productive insects • The ecological impact of insect farming • Safe processing methods • Operator and consumer health

3 • Legislation  51

NOTES 1. Munzinger Online/Brockhaus – Enzyklopädie in 30 Bänden. 21st ed. https://www​.munzinger​.de/, accessed on April 24th, 2012. 2. https://eur​-lex​.europa​.eu​/ legal​-content​/ EN​/ TXT​/ HTML/​?uri​= CELEX​:31988R1677​&from​=DE 3. https://www​.newframe​.com ​/makhanda​-entrepreneur​-aims​-to​-spread​-his​-wings/ However, the production is not for feed but rather for food. 4. https://wor​ldpo​pula​tion​review​.com ​/country​-rankings​/countries​-never​-colonized 5. https://www.wur.nl/en/Research-Results/Chair-groups/Plant-Sciences/Laboratory-of-Entomology/Edibleinsects/Worldwide-species-list.htm 6. https://www​.fao​.org​/faolex​/country​-profiles​/en/ 7. Some insects are also eaten as medicine (e.g. Tchibozo et al., 2016), but this ethnomedicinal aspect of entomophagy will not be addressed in this paper. 8. https://eur​ - lex ​ . europa​ . eu​ / legal​ - content​ / EN​ / TXT/​ ? uri​ = CELEX​ %3A02002R0178​ -20210526​ & qid​ =1638277259477 9. https://eur​ - lex​ . europa​ . eu​ / legal​ - content​ / EN​ / TXT/​ ? uri​ = CELEX​ %3A02004R0852​ -20210324​ & qid​ =1638277302593 10. https://eur​ - lex​ . europa​ . eu​ / legal​ - content​ / EN​ / TXT/​ ? uri​ = CELEX​ %3A02004R0853​ -20211028​ & qid​ =1638277347106 11. https://eur​-lex​.europa​.eu​/ legal​-content​/ EN​/ TXT​/ HTML/​?uri​= CELEX​:32017R0625​&from​=EN 12. https://eur​-lex​.europa​.eu​/ legal​-content​/ EN​/ TXT/​?uri​= CELEX​%3A02005R2073​-20200308 13. https://www​.gesetze​-im​-internet​.de​/tier​-lmhv/ 14. https://eur​ - lex ​ . europa ​ . eu ​ / legal ​ - content ​ / EN ​ / TXT/ ​ ? uri ​ = CELEX​ %3A01997R0258 ​ -20090807​ & qid​ =1638278197295 15. https://eur​ - lex ​ . europa ​ . eu ​ / legal ​ - content ​ / EN ​ / TXT/ ​ ? uri ​ = CELEX​ %3A02015R2283 ​ -20210327​ & qid​ =1638278281492 16. https://eur​-lex​.europa​.eu​/ legal​-content​/ EN​/ TXT/​?uri​= CELEX​%3A32021R0882​&qid​=1638279238858 and https://eur​-lex​.europa​.eu​/ legal​-content​/ EN​/ TXT/​?uri​= CELEX​%3A32021R1975​&qid​=1638279365057; by the time of writing the manuscript, REG 2021/1975 was scheduled to become valid by December 5th, 2021. 17. In fact, the Swiss recommendation for edible insects opens the oportunity for pet food farmers to switch to the food sector if they they heed the recommendations over four generations of insects, i.e. after four generation, feed insects can be food insects. 18. https://eur​-lex​.europa​.eu​/ legal​-content​/ EN​/ TXT/​?uri​= CELEX​%3A02001R0999​-20210907 19. In fact, there are different degrees of impacts invasive species can have on a “new” environment, from beneficial to neutral to detrimental. Some countries evaluated the impact of productive insects. To give an example, species like Alphitobius diaperinus (buffalo worm), T. molitor, A. domesticus, Gryllus assimilis (Jamaican field cricket), H. illucens, and Musca domestica (housefly) are not estimated to be detrimental, and so escaping into the wild does not endanger the ecosystem (https://www​.knaq​-sh​.de​/aktuell​/news​/2017​/insekten​-futtermittel​.html). 20. https://eur​-lex​.europa​.eu​/ legal​-content​/ EN​/ TXT/​?uri​= CELEX​%3A32017R0893​&qid​=1638518882435 21. https://eur​ - lex ​ . europa ​ . eu ​ / legal ​ - content ​ / EN ​ / TXT/ ​ ? uri ​ = CELEX​ %3A02009R1069 ​ -20191214 ​ & qid​ =1638520509889 22. https://eur​ - lex ​ . europa​ . eu​ / legal​ - content​ / EN​ / TXT/​ ? uri​ = CELEX​ %3A02009R0767​ -20181226​ & qid​ =1638520548178 23. https://eur​-lex​.europa​.eu​/ legal​-content​/ EN​/ TXT/​?uri​= CELEX​%3A32021R1372​&qid​=1638521668822 24. https://www​.nweurope​.eu ​/projects​/project​-search ​/valusect ​/news​/why​-should​-we​-feed​-pigs​-and​-poultry​-with​ -insects​-protein/ 25. https://www​.fao​.org​/faolex ​/en/ 26. Dactylopius coccus, a carmine dye-producing scale insect. 27. The background of this question refers to the definition of “primary production”, which in some countries includes killing, freezing, cooking, and storing, and in others does not. In the latter case, a seperate permit may be necessary. 28. The background of this question refers to the definition of “primary production”, which in some countries includes killing, freezing, cooking, storing, and in others does not. In the latter case, a seperate permit may be necessary. 29. https://ipiff​.org/

52  Edible Insects Processing for Food and Feed

REFERENCES Bromiley, G.W., Ed. (2002) [1988]. “Baal-Zebub”. The International Standard Bible Encyclopedia. 1 (Revised (381) ed.). Grand Rapids, MI: Eerdmans Publishing Company, 47. Calber, C., Albright, J. (2021). Chicken behaviour. In Backyard Poultry Medicine and Surgery: A Guide for Veterinary Practitioners, Second Edition, Eds. Greenacre, C.B., Morishita, T.Y., Hoboken, NJ: Wiley Blackwell, 343–354. Dagevos, H. (2021). A literature review of consumer research on edible insects: Recent evidence and new vistas from 2019 studies. Journal of Insects as Food and Feed 7, 249–259. De Magistris, T., Pascucci, S., Mitsopoulos, D. (2015). Paying to see a bug on my food: How regulations and information can hamper radical innovations in Europe. British Food Journal 117, 1777–1792. DiGiacomo, K., Leury, B.J. (2019). Review: Insect meal: A future source of protein feed for pigs? Animal 13, 3022–3030. Gostin, L.O., Wiley, L.F. (2016). Public Health Law: Power, Duty, Restraint. Third Edition. Oakland, CA: University of California Press. Grabowski, N.T. (2017). Speiseinsekten. Hamburg: Behr’s Verlag. Grabowski, N.T., Ahlfeld, B., Lis, K.A., Jansen, W., Kehrenberg, C. (2019). The current legal status of edible insects in Europe. Berliner Und Münchener Tierärztliche Wochenschrift 132(5/6), 295–311. https://doi​.org​/10​.2376​ /0005​-9366​-18087 Grabowski, N.T., Tchibozo, S., Abdulmawjood, A., Acheuk, F., M’Saad Guerfali, M., Sayed, W.A.A., Plötz, M. (2020). Edible insects in Africa in terms of food, wildlife resource, and pest management legislation. Foods 9:502. https://​doi​.org​/10​.3390​/foods9040502 Jansen, W., Grabowski, N.T. (2019). Insects as novel food. The Belgian pioneer experience with insects for human consumption placed on the domestic market. Berliner Und Münchener Tierärztliche Wochenschrift 132(5/6), 312–316. Jensen, A.N., Hansen, S.H., Baggesen, D.L. (2020). Salmonella typhimurium level in mealworms (Tenebrio molitor) after exposure to contaminated substrate. Frontiers in Microbiology 11, 1613. https://doi​.org​/10​.3389​/fmicb​ .2020​.01613 Menzel, P., D’Alusio, F. (1998). Man Eating Bugs – The Art and Science of Eating Insects. New York: A Material World Book. Meyer-Rochow, V.B. (2005). Traditional food insects and spiders in several ethnic groups of Notheast India, Papua New Guinea, Australia, and New Zealand. In Ecological Implications of Minilivestock, Ed. M.G. Paoletti. Enfield: Science Publishers, 389–413. Naber, F. (2016). Drohenbrut als Futter. Deutsches Bienen-Journal 5/2016, 72. Payne, C., Caparros Megido, R., Dobermann, D., Frédéric, F., Shokley, M., Sogari, G. (2019). Insects as food in the global north – the evolution of the entomophagy movement. In Edible Insects in the Food Sector: Methods, Current Applications and Perspectives, Eds. G. Sogari, C. Mora, D. Menozzi. Cham: Springer, 11–26. Ramos Elorduy, J., Pino Moreno, J.M. (1989). Los insectos comestibles en el México antiguo. México, DF: AGT Editor, S.A. Tchibozo, S., Malaisse, F., Mergen, P. (2016). Insectes consommés par l’homme en Afrique occidentale francophone/ Edible insects by Human in Western French Africa. Geo-Eco-Trop 40, 105–14. Von Paczensky, G., Dünnebier, A. (1999). Kulturgeschichte des Essens und Trinkens. Munich: Orbis. Wei, Z.J., Lia, A.M., Zhang, H.Z., Liu, J., Jiang, S.T. (2009). Optimization of supercritical carbon dioxide extraction of silkworm pupal oil applying the response surface methodology. Bioresource Technology 100, 4214–4219. Wissenschaftlicher Beirat der Bundesärztekammer, Ed. (2003). BSE und die Variante der Creutzfeldt-JakobKrankheit (vCJK) Merkblatt für Ärztinnen und Ärzte. Deutsches Ärzteblatt 100, A578–A582.

Nutrient Content and Functionalities of Edible Insects

4

Ruchita Rao Kavle, Ellie Pritchard, Alaa El-Din Ahmed Bekhit, Alan Carne, and Dominic Agyei Contents 4.1 Introduction 54 4.2 Nutritional Composition of Edible Insects 56 4.2.1 Protein and Amino Acids 58 4.2.2 Energy Value 60 4.2.3 Lipids and Fatty Acids 60 4.2.4 Minerals 63 4.3 Digestibility of Edible Insects 64 4.4 Health Benefits 66 4.4.1 Gastrointestinal Health 66 4.5 Insect Protein Hydrolysates 67 4.6 Protein Functionalities 68 4.6.1 Protein Solubility 69 4.6.2 Emulsifying Properties 69 4.6.3 Coagulation 70 4.6.4 Surface Charge 71 4.6.5 Surface Hydrophobicity 71 4.6.6 Water Holding Capacity 72 4.6.7 Oil Holding Capacity 72 4.6.8 Colour 72 4.6.9 Foaming Properties 73 4.7 Bioactivities 74 4.7.1 Antioxidant Properties 74 4.7.2 Angiotensin Converting Enzyme (ACE) Inhibitory Activity 75 4.7.3 Antimicrobial Properties 75 4.7.4 Antidiabetic Properties 76 4.8 Conclusion 76 References 76

DOI: 10.1201/9781003165729-4

53

54  Edible Insects Processing for Food and Feed

4.1 INTRODUCTION Food production and distribution around the world is a massive system that encompasses a wide range of social, economic and environmental aspects. Due to increasing concerns about food production security and sustainability, there is increasing science-based investigation of alternative sources of dietary protein that can be partially or completely obtained by more sustainable production (Nadathur, Wanasundara, & Scanlin, 2016). A report by the Food and Agriculture Organization (FAO) in 2006 highlighted that livestock production is particularly unsustainable (FAO et al., 2006). An updated report in 2009 suggested that by the year 2050, there will be a 70% increase in demand for meat, which may not be able to be met by increased production, resulting in meat becoming a scarce and expensive commodity (FAO, 2009). As the feed for livestock in many parts of the world mainly consists of soy and legumes, increased livestock production will lead to competition for soy and legume staple foods, contributing to global food insecurity (Post, 2018). Additionally, it has been estimated that 15%–29% of world greenhouse gas production is emitted by livestock used for meat production. The promotion and production of protein from alternative sources beside meat could also help alleviate increasing consumer concerns about animal welfare (Dawkins, 2006; DeBacker & Hudders, 2015). Therefore, an alternative to meat as a source of protein, such as edible insects, as well as the possibilities of algae, seaweed, and cultured meat materials, are considered to be crucial for future food security and sustainability (Hamm, 2018). Edible insect species are being prospected as a resource with a considerable potential to contribute to improving the sustainability of the food supply, biodiversity, and the ecosystem worldwide (EUFIC, 2017). Several insect species are being considered as suitable candidates for the sustainable production of protein-containing food (Belluco et al., 2013). Entomophagy, the consumption of insect material, could be one of several ways to solve issues of food insecurity, malnutrition, hunger, climate change, environmental impact, and food distribution issues (Raheem et al., 2019; Zielińska, Baraniak, & Karaś, 2017). Consumption of insects has been practised in several countries for some time (Figure 4.1) as part of a

FIGURE 4.1  World map summarizing the numbers of recorded edible insects in different countries. (From Jongema, Y. [2017]. Recorded edible insect species, by country. Wageningen University and Research. https:// www​.wur​.nl ​/en ​/ Research ​- Results ​/Chair​- groups ​/ Plant​-Sciences ​/ Laboratory​- of​- Entomology​/ Edible ​- insects ​/ Worldwide​-species​-list​.htm [accessed August 23, 2021].)

4  •  Nutrient Content and Functionalities of Edible Insects  55

FIGURE 4.2  Nutritional and health aspects associated with edible insect material consumption.

FIGURE 4.3  Diagrammatic representation of the publications in the Scopus database for the top ten edible insects. (Data from Kavle et al. [2022].) The node size is proportional to the number of studies published.

56  Edible Insects Processing for Food and Feed

FIGURE 4.4  Illustration of the flow of dietary protein and amino acids in a biological system. (Reprinted with permission from Nadathur et al., Sustainable Protein Sources (Elsevier, 2017), 1–19.)

regular diet (DeFoliart, 1997). Consumption of insect material can have numerous nutritional, health, and technological benefits (Laureati, Proserpio, Jucker, & Savoldelli, 2016) as summarized in Figure 4.2. A challenging task is to encourage people in many cultures to make changes in relation to food choices and preference habits and attitudes. Regional food authorities such as the European Food Safety Agency have recognized insect materials as a “Novel Food” since 2018 (EFSA, 2021).​​ In 1975, Meyer-Rochow published a paper proposing that edible insects could solve the problem of global food shortages (Meyer-Rochow, 1975). Edible insects are reported to be high in protein, dietary fibre, fatty acids, vitamins, minerals, and energy content, and to have high levels of macronutrients such as copper, iron, magnesium, manganese, phosphorus, and zinc (Nowak, Persijn, Rittenschober, & Charrondiere, 2016) that can contribute to meeting human and animal nutritional requirements. For example, 100 g of dried caterpillars is reported to provide 76% of the recommended daily intake of protein (Agbidye, Ofuya, & Akindele, 2009). Mitsuhashi (2010) reported that silkworm pupae contain about 30% lipids and 50% protein, and three pupae are as nutritious as a chicken egg. Hence, there is an increasing interest not only in developing countries but also in developed countries for the use of edible insect materials to increase food security and sustainability. However, studies have shown that the composition and the nutritional value of insects are affected by their diet and life cycle (Hamm, 2018; Zielińska, Baraniak, Karaś, Rybczyńska, & Jakubczyk, 2015).

4.2  NUTRITIONAL COMPOSITION OF EDIBLE INSECTS A compilation of the top ten most consumed insects as reported in the literature is shown in Table 4.1, along with the common name, order, origin, and diet of these insects, as well as the development stage at which they are usually consumed. The top ten insects were obtained from a Scopus search for “edible insects” or “insects as food and feed” and analyzed by VOS viewer (v.1.6.16 Centre for Science and Technology Studies [CWTS], Leiden University, Leiden, The Netherlands), a program which is able to create a network visualization map of the studied insects (Figure 4.3). Of the ten insect species, 40% belong to the order Orthoptera, 20% are in each of the Coleoptera and Lepidoptera orders, and 10% are in each of the Diptera and Hymenoptera orders.​ Some of the insects, such as Tenebrio molitor, Rhynchophorus phoenicis, Apis mellifera, and Imbrasia oyemensis, are typically consumed in their larval form. At this developmental stage, the high fat content of these insect larva can provide unique flavours. The production of Tenebrio molitor has been industrialized due to its high growth rate, even when fed on dry low-nutrient waste (Moreno &

Mealworm

Two-spotted cricket House crickets

Black solider fly

Migratory locust

Silkworm Desert locust African palm weevil Western honey bee Caterpillars

Tenebrio molitor

Gryllus bimaculatus

Hermetia illucens

Locusta migratoria

Bombyx mori Schistocerca gregaria Rhynchophorus phoenicis Apis mellifera

Imbrasia oyemensis

Acheta domesticus

COMMON NAME

SCIENTIFIC NAME

Flowers, leaves

Nectar

Hymenoptera Lepidoptera

Mulberry leaves Any vegetation Tree barks

Any vegetation

Grains, organic material fresh or decay Plants and other insects Plants and other insects Organic waste

NATURAL DIET OF THE INSECT

Lepidoptera Orthoptera Coleoptera

Orthoptera

Diptera

Orthoptera

Orthoptera

Coleoptera

ORDER

Larvae

Larvae and pupae

Pupae and larvae Adult Larvae, pupae, and adult

Adult

Adult, nymph, and larvae Larvae, adult, and pupae

Adult

Larvae, adult, and pupa

DEVELOPMENTAL STAGE CONSUMED

Ivory Coast

Canada, Mexico, USA

tropical/subtropical Western Hemisphere, and Australia Africa, Asia, Australia, and New Zealand USA, Mexico, India Mexico Nigeria

Africa, Mediterranean, parts of Asia USA, Mexico

USA, Mexico

GEOGRAPHICAL ORIGIN

TABLE 4.1  The top ten edible insects reported in literature, in terms of scientific name, common name, order, natural diet of the insect, insect developmental stage consumed, and geographical origin

4  •  Nutrient Content and Functionalities of Edible Insects  57

58  Edible Insects Processing for Food and Feed Elorduy, 2002). Crickets and locusts are mainly eaten in the adult form. They are easy to harvest in the wild, especially when they are travelling in swarms. Although silkworm (Bombyx mori) have been domesticated in several parts of Asia for silk production, silkworm pupae are traditionally consumed as food in several countries (Zhang, Tang, & Cheng, 2008). Edible insects play an important role in the food systems of many countries where their consumption has been accepted and practiced for a long time. In fact, they are a vital source of essential nutrients. The nutritional composition of the ten most common edible insect species is summarized in Table 4.2. While considerable variation in nutrient composition exists within the same insect species obtained from the wild, it is expected that commercially farmed and produced edible insects will have a more consistent nutritional composition, as the farming conditions such as feed type/composition and environment can be controlled (Churchward-Venne, Pinckaers, van Loon, & van Loon, 2017).

4.2.1 Protein and Amino Acids Insects have been promoted as one of the most sustainable alternative sources of protein (Figure 4.4). Edible insects have a relatively higher protein content on a mass basis compared to other animal and plant sources (Teffo, Toms, & Eloff, 2007). Dietary protein is central to biological activities in living organisms and makes up important components such as haemoglobin, hormones, and enzymes as well as structural protein components, and as a supply source of essential amino acids. Protein is an important component of antibodies, which bolsters the immune function of the body. Protein can be hydrolyzed and the amino acids utilized for energy production by the body, although it is the least preferred source compared to carbohydrates and lipids. The nutritional value of a food is largely dependent on the quality of the protein (i.e. based on its digestibility and amino acid composition). On a dry weight basis, proteins are the most abundant nutrient in insects (Table 4.2). For example, the species Acheta domesticus, Locusta migratoria, Bombyx mori, and Schistocerca gregaria contain protein contents of 75.8%, 65.9%, 69.8%, and 69.05% respectively. Compared with the protein contents of insects, raw beef (21.2%), whole eggs (12.6%), and soy milk (3.55%) contain lower amounts of protein. Most of the protein contents reported for edible insects have been determined by the Kjeldahl method using a nitrogen conversion factor (Kp) of 6.25. However, it has recently been pointed out that this factor may not be suitable for insects and leads to protein overestimation due to the presence of insect exoskeleton, which contains a large amount of chitin fibre, rich in non-protein nitrogen (Da Silva Lucas, De Oliveira, Da Rocha, and Prentice, 2020). Hence, a nitrogen conversion factor for insects should be based on the sum of amino acid residues (Ka) (Jantzen, de Oliveira, da Rocha, & Prentice, 2020). Therefore, both nitrogen-to-protein conversion factors, Kp and Ka, need to be determined for edible insects to determine the protein/total nitrogen and protein/protein nitrogen for each insect species (Boulos, Tännler, & Nyström, 2020). Table 4.2 shows a large variation in the nutrient content of various insects. Protein contents can vary by more than 50% within the same insects. For example, Imbrasia oyemensis has been found to contain 12.78%–61.59% protein, Acheta domesticus 15.4%–75.8%, and Rhynchophorus phoenicis 25.7%–41.7% protein. This suggests that the large variation may arise from extrinsic factors such as feed, ecology, and the geographical location where insects were reared or harvested. Other parameters such as the developmental stage of the insect (Kulma et al., 2020) and the manner in which they are processed (involving, for example, the use of mechanical and/or thermal treatments) may also affect the nutritional composition of insect materials. Edible insect protein generally has good amino acid profiles, with essential amino acids such as isoleucine, leucine, lysine, methionine, phenylalanine, threonine, tryptophan, and valine being present at nutritionally adequate levels (FAO/WHO, 2007). However, some variations exist across different insect species. Some edible insects have amino acid contents that exceed the requirements recommended by the Food and Agriculture Organisation (FAO) and the World Health Organization (WHO), for various age groups. For example, Imbrasia oyemensis contains, per 100 g of protein, 12 g histidine, 8.9 g lysine, 2.2 g

15.8 15.4–75.8

42–56.1

65.9 52.6–69.8

41.7–56.8 52.3–69.05 25.7–41.7

12.78– 61.59

Gryllus bimaculatus Acheta domesticus

Hermetia illucens

Locusta migratoria Bombyx mori

Apis mellifera Schistocerca gregaria Rhynchophorus phoenicis

Imbrasia oyemensis

*Carbohydrate, chitin, vitamin

3.55 12.4 21.2

45.6–52.8

Tenebrio molitor

Soymilk, plain Egg, whole, raw Beef, raw

PROTEIN

SPECIES

2.12 9.51 3.68

6.59–25.36

7.5–18.82 12–15.04 19.50– 59.43

23.81 8.09–19.21

12.80–33

5.5–5.8 4.4–30.7

34.5–37.1

FAT

0.64 0.85 1.02

0.72–2.78

3–4 10–3.97 3.27

3.09 4.8–6.36

8.8–10.2

2.8 4.54

3.5–4.1

ASH

2.3 1.91 0.23

10.64

11.1 12.02–19 2.82–5.70

25.56 23.4

36.4

10.1 13.2

8.5–16

OTHERS*

38 143 119

137.27

475 163.2– 429.50 478.60

490.13 172.27–389

247

120 137.5

239–653

ENERGY Costa et al., 2020; EFSA et al., 2021; Jones, Cooper, & Harding, 1972; Kim, Weaver, & Choi, 2017 Orkusz, 2021 Bawa, Songsermpong, Kaewtapee, & Chanput, 2020; Orkusz, 2021 Bosch et al., 2014; Diener et al., 2009; Spranghers et al., 2017 Purschke et al., 2018; Liu et al., 2017 Finke, 2002, 2007; Kim et al., 2017; Rumpold & Schlüter, 2013 Rumpold & Schlüter, 2013 Bosch et al., 2014; Mariod, 2020 Elemo, Elemo, Makinde, & Erukainure, 2011; Mlcek, Rop, Borkovcova, & Bednarova, 2014; Onyeike, Ayalogu, & Okaraonye, 2005; Rumpold & Schlüter, 2013 Yisa Njowe & Sop, 2017 Akpossan et al., 2014; Akpossan, Dué, Kouadio, & Kouamé, 2009; Loh, Njowe, & Sop, 2017a; Rumpold & Schlüter, 2013; Yisa Njowe & Sop, 2017 USDA, 2015 USDA, 2015 USDA, 2015

REFERENCES

TABLE 4.2  Nutritional composition (%) and energy content (kcal/100 g) of edible insects, and conventional foods based on reported on dry weight basis

4  •  Nutrient Content and Functionalities of Edible Insects  59

60  Edible Insects Processing for Food and Feed phenylalanine, and 7.9 g threonine, while Apis mellifera contains, per 100 g protein, 5.9 g isoleucine, 8.7 g leucine, and 3.2 g methionine. Essential amino acids (gram per 100 gram protein) are highest in Imbrasia oyemensis (58.7 g), followed by Tenebrio molitor (45.5 g), which is higher than some other conventional food sources, such as whole egg (5.6 g), beef (8.4 g), and soymilk (1.3 g). The presence of lysine in edible insects and the inclusion of edible insects in the diet could have considerable benefits in complementing lysine-poor staple cereals. Several studies have demonstrated a similar trend whereby edible insects are found to possess levels of essential amino acids comparable to that of soybean and cow casein (Rumpold & Schlüter, 2013; Yi et al., 2013), and animal feeding trials demonstrated the high biological value of edible insect proteins. For example, the quality of Zonocerus variegatus (giant grasshopper) protein was evaluated in rats, and compared to that of soybean and crayfish in a 28-day feeding trial (Soloman, Ladeji, & Umoru, 2008). The giant grasshopper proteins resulted in an 80%–100% higher weight gain in rats, as well as higher food and protein efficacy ratios compared to control chow, and diets containing soybean and crayfish treatments. In addition, the protein efficiency ratio of cricket (Gryllus assimilis), African palm weevil (Rhynchophorus phoenicis), coconut rhinoceros beetle (Oryctes rhinoceros), and termite (Macrotermes bellicoccus) have been reported to be higher or comparable to those of casein, a standard high-quality protein (Ekpo, 2011; Oibiokpa, 2017; Oibiokpa, Akanya, Jigam, Saidu, & Egwim, 2018). Furthermore, the biological value of cricket (Gryllus assimilis), moth (Cirina forda), grasshopper (Melanoplus foedus), and termite (Macrotermes nigeriensis) was higher than that of casein (Oibiokpa et al., 2018). However, in some cases, the inclusion of insect protein in animal feed has been found to have the opposite effect, such as, for example, when Tenebrio molitor larvae was included at 5% and 10% in the diet of piglets under isocaloric and iso-nitrogenous conditions.​

4.2.2 Energy Value Based on their nutrient profile, calorie values of between 120 and 653 kcal/100 g can be obtained from edible insects (Table 4.2). Edible insects are a great source of caloric energy. Generally, adult insects contain higher amounts of chitin, which is indigestible and contributes little caloric input, while larvae and pupae provide high amounts of protein and fats and therefore have relatively high caloric inputs (Tang et al., 2019). Tenebrio molitor (larvae) can provide 653 kcal/100 g of energy, while Gryllus bimaculatus (adult) can provide 120 kcal/100 g, Locusta migratoria (larvae) 490.13 kcal/100 g, Apis mellifera (larvae) 475 kcal/100 g, and Rhynchophorus phoenicis (larvae) 478.60 kcal/100 g. Therefore, products made from different developmental stages of edible insects may provide flexibility in designing meals for consumers with different calorie requirements (Table 4.3).

4.2.3 Lipids and Fatty Acids Lipids represent the second largest nutrient fraction in edible insects. The lipid content ranges from 4.4% to 59.4% on a dry weight basis (Table 4.4). These values are comparable to or higher than those in beef (3.68%) and whole egg (9.51%). It is worth noting that the highest fat content is found in the larval stages, ranging from Rhynchophorus phoenicis (59.43%) down to Acheta domesticus adult stage (4.4%) (Imathiu, 2020). Fatty acids (FAs) are organic acids with at least one carboxyl group and a long carbon chain that may contain one or more double bonds, as in unsaturated fatty acids, or only single bonds, as in saturated fatty acids. Fatty acids are generally classified as saturated fatty acid (SFAs), monosaturated fatty acid (MUFA), and polysaturated fatty acid (PUFA) (Moghadasian & Shahidi, 2017), and, depending the degree of saturation, their physicochemical, nutritional, stability, and sensory attributes can vary. Lipids in edible insects that are liquid at room temperature (25°C) are called “insect oils”. These oils mainly contain high levels of unsaturated fatty acids (for example, linoleic acid, alpha-linolenic acid, and ω-3). The liquid nature of insect oils makes them ideal for use in frying and in mayonnaise and other food products.

4.7

7.8

6.1

1.4

4.1

3.5

7.8

6.6

45.5

Finke, 2002; Rumpold & Schlüter, 2013

Isoleucine

Leucine

Lysine

Methionine

Phenylalanine

Threonine

Tryptophan

Valine

EAA

Reference

Udomsil et al., 2019

19.2

3.5

0.3

1.7

2.2

0.9

2.9

3.9

2.4

1.6

GB (ADULT)

Igual et al., 2020; Udomsil et al., 2019

21.6

4.5

0.4

1.7

2.4

1.0

3.2

3.8

2.9

1.7

AD (ADULT)

de Marco et al., 2015

13.4

2.2

-

1.5

1.4

0.9

2.2

2.4

1.7

1.1

HI (LARVAE)

Purschke et al., 2018

45.2

7.9

0.7

3.6

9.3

2.1

5.4

8.7

4.8

2.8

LM (ADULT)

Finke, 2007

27.6

3.9

0.7

2.8

2.8

1.3

5.0

4.9

3.3

2.9

BM (LARVAE)

Mishyna, Martinez, Chen, & Benjamin, 2019

43.6

7.2

-

3.0

4.4

3.2

8.5

8.7

5.9

2.9

AM (LARVAE)

Mishyna, Martinez, Chen, & Benjamin, 2019

39.6

7.8

-

2.7

3.9

1.9

6.6

8.3

5.1

3.0

SG (ADULT)

Onyeike et al., 2005; Rumpold & Schlüter, 2013

31

3.5

-

3.1

4.8

1.9

4.5

5.4

3.9

3.9

RP (LARVAE)

Gérard et al

58.7

6.1

2.1

7.9

8.8

0.4

8.9

6.9

5.6

12.0

IO (LARVAE)

USDA 2015

8.4

0.9

0.2

0.8

0.8

0.7

1.7

1.6

0.9

0.8

BEEF (RAW)

USDA 2015

1.3

0.1

0.05

0.1

0.2

0.05

0.2

0.3

0.2

0.1

SOY MILK

USDA 2015

5.6

0.9

0.2

0.6

0.7

0.4

0.9

1.1

0.7

0.3

EGG (WHOLE)

EAA: Essential amino acid; TM – Tenebrio molitor; GB – Gryllus bimaculatus; AD – Acheta domesticus; HI – Hermetia illucens; LM – Locusta migratoria; BM – Bombyx mori; AM – Apis mellifera; SG – Schistocerca gregaria; RP – Rhynchophorus phoenicis; IO – Imbrasia oyemensis

3.5

Histidine

TM (LARVAE)

TABLE 4.3  Amino acid content (g/100 g of protein) of edible insects: Comparison with some values for some conventional foods (reported on dry weight basis)

4  •  Nutrient Content and Functionalities of Edible Insects  61

0.40

0.29

0.14

22.58

0.11

3.47

0.17

ND

ND

27.3

2.15

2.09

29.74

ND

ND

33.98

36.98

0.09

ND

ND

1.67

ND

ND

38.74

1.67

36.98

Rumpold & Schlüter, 2013

C12:0

C14:0

C15:0

C16:0

C17:0

C18:0

C20:0

C22:0

C24:0

SFA

C14:1

C16:1

C18:1n-9 C

C20:1

C22:1n-9

MUFA

C18:2n-6

C20:2

C22:6n-3 (DHA)

C18:3n-6

C18:3n-3

C20:3n-6

C20:3n-3

PUFA

Omega-3

Omega-6

Reference

Udomsil, Imsoonthornruksa, Gosalawit, & Ketudat-Cairns, 2019

1.39

0.01

1.72

ND

0.08

0.01

0.06

ND

0.17

1.39

0.45

0.02

0.08

0.04

0.30

0.01

12.8

0.04

0.05

0.21

2.73

0.10

9.26

0.04

0.27

0.05

0.01

GB

Udomsil et al., 2019

1.17

0.01

1.40

0.01

0.01

0.01

0.01

ND

0.19

1.17

0.24

0.01

0.02

0.03

0.15

0.02

8.2

0.02

0.06

0.13

1.83

0.08

5.87

0.01

0.11

0.03

0.01

AD

Lawal, Kavle, Akanbi, Mirosa, & Agyei, 2021

4.82

0.49

5.31

ND

ND

0.49

ND

ND

ND

4.82

3.01

0.58

0.14

0.16

2.13

ND

86.8

ND

ND

0.54

1.49

0.14

9.41

0.13

7.53

62.05

5.46

HI

Udomsil et al., 2019

5.24

11.69

17.59

0.27

ND

11.69

0.11

0.12

0.16

5.24

39.23

ND

0.10

38.00

1.13

ND

40.3

0.09

0.10

0.46

7.33

0.55

29.52

0.11

1.90

0.11

0.12

LM

Paul & Dey, 2014

0.23

0.24

1.69

0.32

ND

0.24

0.42

0.23

0.26

0.23

0.32

ND

0.04

0.09

ND

0.20

1.6

ND

0.16

0.14

0.40

0.06

0.42

ND

0.18

0.24

ND

BM

Ghosh, Jung, & MeyerRochow, 2016

7.80

ND

7.80

ND

ND

ND

ND

ND

ND

7.80

67.00

ND

19.20

45.20

2.60

ND

25.2

ND

ND

ND

9.30

0.40

14.40

ND

0.60

0.30

0.20

AM

Zielińska et al., 2015

14.04

11.35

25.57

ND

ND

11.35

ND

ND

0.18

14.04

38.16

ND

0.14

36.22

1.80

ND

35.3

ND

0.07

0.40

9.27

0.24

23.26

0.09

1.68

0.23

0.07

SG

Paul & Dey, 2014

3.10

0.90

4.00

ND

ND

ND

ND

ND

ND

3.10

66.0

ND

ND

30.0

36.0

ND

44.5

ND

ND

0.70

5.10

ND

36.0

0.10

2.50

0.10

ND

RP

Akpossan et al., 2009; Rumpold & Schlüter, 2013

11.2

ND

11.2

ND

ND

ND

ND

ND

ND

11.2

34.6

ND

ND

34.6

ND

ND

53.7

ND

ND

ND

7.21

ND

45.9

ND

0.48

ND

ND

IO

USDA 2015

0.103

1.02

1.53

0.004

0.003

0.061

0.002

0.002

0.002

0.103

1.182

0.06

ND

1.02

0.09

0.012

1.22

0.001

0.002

0.007

0.772

0.06

0.728

ND

0.079

0.001

0.001

BEEF

USDA 2015

0.99

0.17

1.32

ND

ND

0.17

0.17

ND

ND

0.99

0.42

ND

0.01

0.41

ND

ND

0.3

ND

0.01

0.01

0.07

ND

0.22

ND

ND

ND

ND

SOYMILK

USDA 2015

1.46

0.01

1.52

0.02

0.02

0.01

0.01

ND

ND

1.46

3.59

ND

ND

3.35

0.24

ND

3.2

ND

ND

ND

0.81

ND

2.32

ND

0.03

ND

ND

EGG

SFA – saturated fatty acid, MUFA- monounsaturated fatty acids, PUFA- polyunsaturated fatty acids, ND- no data (Data not reported in mentioned references); TM – Tenebrio molitor; GB – Gryllus bimaculatus; AD – Acheta domesticus; HI – Hermetia illucens; LM – Locusta migratoria; BM – Bombyx mori; AM – Apis mellifera; SG – Schistocerca gregaria; RP – Rhynchophorus phoenicis; IO – Imbrasia oyemensis

0.14

C10:0

TM

TABLE 4.4  Fatty acid composition of various insect species, with comparison to some conventional foods (beef, soymilk, egg) [% fatty acid], dry weight basis

62  Edible Insects Processing for Food and Feed

4  •  Nutrient Content and Functionalities of Edible Insects  63 Insect fats, on the other hand, are solid at room temperature. These fats contain saturated fatty acids, especially suitable for making products such as pasta, margarine, and confectionery (Imathiu, 2020). Fatty acids play critical roles in health, disease, and human metabolism. The fatty acid profile of conventional food sources is slightly comparable to that of insects (Jayanegara, Gustanti, Ridwan, & Widyastuti, 2020). Within and across insect species, larvae are considered a better source of lipids compared to the later developmental stages such as adult or pupae. Similar to the essential amino acid content, the essential fatty acids and the cholesterol contents of edible insects are dependent on environmental factors and the developmental stage, feed, and gender (Nowakowski, Miller, Miller, Xiao, & Wu, 2020; Rumpold & Schlüter, 2015). The saturated fatty acids (SFAs) in insects at different developmental stages are predominantly stearic and palmitic acids. However, monosaturated fatty acids (MUFAs) are usually healthier for human diets. Oleic acid, a widely available fatty acid in edible insects, helps to reduce blood pressure and has great potential in the reduction of cardiovascular, immune, and inflammatory diseases (Sales-Campos, de Souza, Peghini, da Silva, & Cardoso, 2013). Compared with beef, egg, and soymilk, many mature edible insects contain a higher amount of polyunsaturated fatty acids (PUFAs) (Nowakowski et al., 2020), as summarized in Table 4.4. Linolenic acid (C18:2 n6) is a major PUFA in insects. It has been proven to have anti-inflammatory, skin-lightening, and acne reduction potential (Tang et al., 2019). The highest source of linolenic acid is found in insects in the orders Orthoptera and Lepidoptera. Such insects are especially rich in alpha-linolenic acids (Tang et al., 2019). The essential fatty acids 18:2 n-6 (omega 6) and 18:3 n-3 (omega 3) are essential in that they are not synthesized by mammalian cells and therefore need to be consumed in the diet (Jantzen et al., 2020). PUFAs are considered healthy fats usually found in plant oils, seeds, nuts, and fish. Edible insects do not contain docosahexaenoic acid (DHA) and eicosapentaenoic acid (EPA), two essential PUFAs that are usually found in fish (Dobermann, Swift, & Field, 2017; Kinyuru, Mogendi, Riwa, & Ndung'u, 2015; Nowakowski et al., 2020). The high content of PUFA in edible insects indicates that edible insects are likely to have a relatively low cholesterol content (Kinyuru et al., 2015). As shown in Table 4.4, it is not uncommon to see wide variations in the fatty acid profile among insect species due to reasons such as differences in developmental stage, location of rearing, feed, and gender. Additionally, it can be seen from Table 4.4 that the amounts (expressed in % of total fatty acids) of linoleic (18:2 n-6) and alpha-linolenic (18:3 n-3) acids in edible insects are higher than those in some conventional food sources. Hence, edible insects are potentially a good source of fatty acids.

4.2.4 Minerals Insects are a good source of many minerals. The micro and macro mineral compositions of the ten most studied insects (on a dry weight basis) are summarized in Table 4.5. Edible insects are a good source of calcium, phosphorus, magnesium, zinc, copper, and manganese. It has been reported that edible insects contain more calcium, zinc, and iron than beef, pork, and chicken, and therefore entomophagy can be considered as being good way to obtain minerals to beat “hidden hunger” in developing countries (Gibson, 2015). Most insects have an ~3-fold higher level of iron (Fe) than beef (Payne, Scarborough, Rayner, & Nonaka, 2016a; Voelker, 2019). Edible insects have a higher iron content than certain plants and animal food sources, but there are no studies that have identified the type of iron present in insects (Dobermann et al., 2017). House crickets contain approximately 8.83 mg iron per 100 g of dry matter, and it is estimated that crickets have a 180% higher iron content than beef, making crickets a potential source of iron comparable to traditional food sources (Jantzen et al., 2020; Mwangi et al., 2018; Payne, Scarborough, Rayner, & Nonaka, 2016b). As insects typically contain low sodium (Na) levels, they can be included in the diet of individuals sensitive to high sodium levels (Imathiu, 2020). Potassium (K) is an important body mineral, essential to both electrical and cellular functions (Haas & Forbush, 2000). Humans require a regular intake of ionic potassium, as it is one of the main blood minerals that serve as electrolytes (Silverthorn, 2015).

64  Edible Insects Processing for Food and Feed Obesity, a major risk for diabetes, can potentially be associated with low potassium levels (Chatterjee, Yeh, Edelman, & Brancati, 2011). Conventional food sources such as meat, chicken, and soy are known to be a good source of potassium, and as seen in Table 4.5, insects such as Bombyx mori are one of the richest sources of potassium (1826.6 mg/100 g). The magnesium (Mg) level in edible insects such as Acheta domesticus is 136.5 mg/100 g dry weight (DW), Bombyx mori contains 287.9 mg/100 g DW, and Imbrasia oyemensis 49.4 mg/100 g DW, while beef and soymilk contains comparable magnesium levels of 21 and 21.5 mg/100 g DW respectively. Magnesium acts as a co-factor of various enzymes that are involved in protein and nucleic acid metabolism. Mg is also involved in the regulation of potassium and reduces excretion of calcium from the body, facilitating the maintenance of bone density (FAO/WHO, 2007). Several diets and ready-to-eat meals contain Mg levels which are insufficient for the delivery of adequate Mg to a person with inborn genetic defects. Individuals with low Mg levels in the body can be susceptible to cardiovascular diseases (King, Mainous, Geesey, & Woolson, 2005). Manganese (Mn) is an essential trace metal found in tissues. This mineral acts as a precursor for the metabolism of amino acids, lipids, proteins, and carbohydrate. Manganese is an essential metal for body functions, such as wound healing and development of bone and cartilage, but the ingestion of high amounts can be toxic (Crossgrove & Zheng, 2004). Conventional dietary sources of manganese include green vegetables, blueberries, ginger, and egg yolk. As indicated in Table 4.5, Apis mellifera contains 21.9 mg/100 g dry weight (DW) and can be a potential source of manganese. Zinc (Zn) content in edible insects ranges from 1.9 mg to 24.9 mg/100 g dry weight (Table 4.5). Zn is an integral component of more than 300 enzymes participating in the synthesis and metabolism of micro and macro nutrients (Higdon, 2003). This mineral helps maintain cell and organ integrity and plays a role in building immunity and gene expression (Ghosh, Jung, & Meyer-Rochow, 2017). Edible insects such as Schistocerca gregaria and Tenebrio molitor contain 24.9 and 21.8 mg/100 g DW of Zn, respectively (Table 4.5). As mentioned above, edible insects also contain substantial levels of iron (Fe). For example, the level of Fe in cricket (12.9 mg/ 100 g dry weight) is similar to that in beef sirloin (Latunde-Dada, Yang, & Vera Aviles, 2016) and is higher than that in grasshopper, mealworm, and buffalo worm (Latunde-Dada et al., 2016). It has been suggested (Hurrell & Egli, 2010; Lonnerdal, 2000; Mwangi et al., 2018) that histidine and methionine improve the absorption of Fe and Zn, but protein from most edible insects has a lower content of these amino acids compared to animal muscle. Termites (Macrotermis nigierensis) and moths (Cirina forda) particularly have low levels of methionine (Oibiokpa et al., 2018) (Table 4.6).

4.3  DIGESTIBILITY OF EDIBLE INSECTS Edible insects have been subjected to various processing steps (blanching, drying, frying, and roasting) for various reasons, including reduction of microbial load, removal of the exoskeleton to enhance drying, transformation of the insect appearance to a potentially more acceptable form as a result of cooking, or conversion into a dry powder. All the nutrients present in the food will not necessarily be automatically available for absorption following consumption. Therefore, it is important to gain some insights as to the digestibility of edible insects. The digestibility of insect materials depends on several factors such as insect type and the processing conditions that the insect has been exposed to. In one study, the digestibility of grasshopper was reported to be decreased by toasting and drying (Kinyuru, Kenji, Njoroge, & Ayieko, 2010), whereas in a different study, the digestibility of mealworm was improved by boiling and oven cooking (Megido et al., 2018). Results of published studies support the contention that the effects of processing by conventional cooking methods depend somewhat on the insect species. For example, Madibela, Seitiso, Thema, & Letso (2007) reported that Mopane worm, when subjected to cooking, resulted in lower crude protein, ash, and zinc content and poorer dry-matter digestibility. Conversely, Lautenschläger, Neinhuis, Kikongo, Henle, and Förster., (2017) found no effect of thermal processing on the nutritional properties of caterpillars. The protein digestibility in vivo of dried (at 60–70°C) silkworm (Samia ricinii) pupae in

435.1 1126.6 132.1 957.8 109.4 6.3 21.8 2.0 3.7 Rumpold & Schlüter, 2013

88.8 321.7 105.1 702.0 72.9 7.2 14.4 3.9 3.4 Udomsil et al., 2019

GB

101.4 389.9 149.7 899.3 136.5 8.83 19.6 4.86 4.40 Udomsil et al., 2019

AD 118 820 207.3 105.3 30.7 2.4 1.9 0.16 2.5 Chia et al., 2020

HI 221.9 796.0 129.8 697.2 86.0 6.6 12.7 1.2 0.4 Fombong et al., 2021

LM 274.6 1826.6 102.3 1369.9 287.9 9.5 17.8 2.1 2.5 Barker, Fitzpatrick, & Dierenfeld, 1998; Rumpold & Schlüter, 2013

BM 75.6 1585.4 222.9 860.1 201.7 37.7 14 4.6 21.9 Ghosh et al., 2016

AM 285.1 1309.5 80.5 968.7 128.3 7.3 24.9 4.9 1.3 Fombong et al., 2021

SG 52 1025 54.1 685.0 131.8 30.8 15.8 1.6 3.5 Loh, Njowe, & Sop, 2017b; Rumpold & Schlüter, 2013

RP 0.4 167.0 27.9 69.9 49.4 3.4 2.6 0.3 1.7 Loh et al., 2017b

IO 55 340 6 182 21 1.95 4.28 0.074 0.008 USDA 2015

BEEF

34 158 101 69 21.5 0.54 0.31 0.108 0.208 USDA 2015

SOYMILK

142 138 56 198 12 1.75 1.29 0.072 0.028 USDA 2015

EGG

Na – sodium, K – potassium, Ca – calcium, P – phosphorous, Mg – magnesium, Fe – iron, Zu – zinc, Cu – copper, Mn – manganese TM – Tenebrio molitor; GB – Gryllus bimaculatus; AD – Acheta domesticus; HI – Hermetia illucens; LM – Locusta migratoria; BM – Bombyx mori; AM – Apis mellifera; SG – Schistocerca gregaria; RP – Rhynchophorus phoenicis; IO – Imbrasia oyemensis

Na K Ca P Mg Fe Zn Cu Mn Reference

TM

TABLE 4.5  Mineral profile of various insect species compared to that of beef, soymilk, and egg is expressed as mg/100 g dry weight basis

4  •  Nutrient Content and Functionalities of Edible Insects  65

66  Edible Insects Processing for Food and Feed TABLE 4.6  Summary of the hydrolysis of insect, larvae, and worm protein reported in the recent literature INSECT (COMMON NAME AND ORDER)

ENZYMES

DEGREE OF HYDROLYSIS (%)

BIOACTIVITY OF PEPTIDES/ HYDROLYSATES

Crickets (Orthoptera:Grylliade)

Alcalase and gastrointestinal proteases

15–85

Antioxidant activity, angiotensin converting enzyme (ACE), and dipeptidyl peptidase-4 (DPP-IV)- inhibition

Cotton leafworm, (Lepidoptera:Noctuidae)

Alcalase, Gastrointestinal thermolysin, Gastrointestinal enzymes

101.92 101.92 93.73 -

ACE inhibitory

Tropical house cricket Order: Orthoptera

Gastrointestinal enzymes

32

Silkworm (Lepidoptera: Bombycidae)

Gastrointestinal enzymes

-

Mealworm larvae (Coleoptera: Tenebrionidae), Locusts (Orthoptera: Acrididae)

DPPH radical scavenging activity assay, ABTS radical scavenging activity assay, Fe2+ chelating activity ferric-reducing power, antiinflammatory activity In vitro cytotoxicity assays DPPH radical scavenging, ferrous ions (Fe2+) chelation, reducing power activity assay

REFERENCES Hall et al., 2018; Hall, Jones, O’Haire, & Liceaga, 2017; Zielińska et al., 2017 Vercruysse et al., 2005 Zielińska et al., 2017

Zielińska et al., 2017; Zielińska et al., 2015 Wu, Jia, Tan, Xu, & Gui, 2011

a rat model was about 87% that of untreated insect. A water-soluble protein fraction of Tenebrio molitor was found to be easier to digest compared to the water-insoluble fraction using a gastric and duodenal digestion model (Yi, Van Boekel, Boeren, & Lakemond, 2016). Traditional preparation methods, such as boiling, roasting, or a combination of boiling and roasting, of beetle (Eulepida mashona) and cricket (Henicus whellani) decreased the protein digestibility (Manditsera, Luning, Fogliano, and Lakemond., 2019). Together, these studies suggest that processing conditions can either positively or negatively affect the bio-accessibility and also bioactivity of nutrients in insect materials. Similar observations have also been reported by Zielińska et al. (2015) and Manditsera et al. (2019).

4.4  HEALTH BENEFITS This section provides a summary of the health benefits of edible insects and their potential implications as foodstuff or dietary ingredients.

4.4.1 Gastrointestinal Health Edible insects contain several components of potential benefit to human health, including protein, fatty acids, essential amino acids, glycosaminoglycans, and chitin in the adult form. According to Stull et al.

4  •  Nutrient Content and Functionalities of Edible Insects  67 (2018), supplementing the diets of healthy adults with 25 g/day of roasted cricket dried powder for 14 days decreased the presentation of plasma tumour necrosis and increased levels of the probiotic bacterium Bifidobacterium animais in the gut. This suggests that the consumption of cricket powder may exert a protective effect on inflammatory processes in the human body. Similarly, when black solider fly was fed to rainbow trout, the microbial diversity in the trout gut was found to be enhanced (Bruni, Pastorelli, Viti, Gasco, & Parisi, 2018). Feeding broiler chickens dried mealworm and super mealworm was found to reduce the incidence of infections such as E.coli and Salmonella, and was likely due to the probiotic effects of the edible insects (Islam & Yang, 2017). The chitin component of insects has been reported to be associated with antimicrobial activity as well as gut microbe benefits (Nowakowski et al., 2020). Unsaturated fatty acids derived from mealworm and cricket are reported to help reduce the risk of cardiovascular disease (Makkar, Tran, Heuzé, & Ankers, 2014). Additionally, the substantial vitamin B12 content in cricket could aid in the reduction of cognitive decline, anaemia, and bone fractures and to reducing the risk of cardiovascular diseases (Makkar et al., 2014; Spranghers et al., 2018). When provided in diets, edible insect materials can contribute to improving bone density and decreasing the incidence of bone fractures in the elderly. Other mineral components such as iron and zinc can have potential benefits in improving immune and gastrointestinal functions in humans (Imathiu, 2020). Chitin and its by-products such as chitosan have been shown to have antioxidant, anti-inflammatory, anticancer, and antimicrobial activities (Liaqat & Eltem, 2018). Insect powders can also be used to provide dietary protein supplements in order to build muscle mass by enhancing muscle protein synthesis. And in addition, edible insects are also a potential source of carbohydrates (Churchward-Venne, Pinckaers, van Loon, & van Loon, 2017).

4.5  INSECT PROTEIN HYDROLYSATES Peptides derived from natural sources have attracted huge interest because of the health benefits associated with high bioactivity, and the general lack of deleterious side effects arising from peptides derived from natural sources. Hydrolysis of proteins using proteases is an effective approach for the production of bioactive peptides, without affecting the nutritional value of the protein material. The most studied insect for the production of bioactive peptides has been the silkworm, B. mori, and high production yields of bioactive peptides, particularly those exhibiting antioxidant and angiotensin converting enzyme (ACE) inhibiting activities by use of proteases, in comparison to the use of ultrasound and micrionisation treatments (Zhou, Ren, Yu, Jia, & Gui, 2017). Generally, hydrolysis with proteases has been the main method used for generation of protein hydrolysates of a range of insect species, and various hydrolysis protocols have been reported in the literature. Microbial proteases appear to be the most frequently used for the generation of hydrolysates of insect proteins to produce bioactive peptides. In a study conducted by Vercruysse, Smagghe, Herregods, & van Camp (2005), protein from insects such as B. mori, Bombus terrestris, S. gregaria, and Spodeptera littoralis were sequentially hydrolyzed with pepsin, trypsin, and chymotrypsin, which resulted in the production of peptides exhibiting ACE inhibitory activity. Protein in farmed insects such as Amphiacusta annulipes, S. gregaria, and L. migratoria, B. dubia, Gromphadorhina portentosa, along with T. molitor and Z. morio larvae, have also been subjected to hydrolysis by simulated gastrointestinal digestion (Zielińska, Baraniak, et al., 2017; Zielińska et al., 2015). Hydrolysis of B. mori chrysalises with papain has also been studied (Yang et al., 2018). Some studies have determined the degree of hydrolysis to analyze the protease mediated digestion of insect protein, and the degree of hydrolysis has been found to vary from 3% to 100% of the total protein content, depending on the insect species and hydrolysis protocol. For example, Zielińska, Baraniak et al. (2017) applied the same hydrolysis protocol to both L. migratoria and A. annulipes, and the degree of hydrolysis was found to range from 15.8% to 36.3%. Hydrolysis of insect protein with proteases has

68  Edible Insects Processing for Food and Feed not only been carried out with the aim to generate bioactive peptides, but also to improve the technofunctionality of the material, including the solubility, foaming, and thermal stability of the insect material. An exploration of the nutritional and functional properties of insect protein hydrolysates will advance the interest of the food industry in the use of peptides as functional food ingredients. Nonetheless, there has been very little research carried out on the functionality of insect proteins. The ability to fulfil one or more functionality requirements such as solubility, emulsion/foam stabilization, or gel formation can be a very successful use of proteins as ingredients in food products (Jantzen et al., 2020). Further, enzymatic modification of protein is a useful mechanism to improve these functionalities (Hall, Johnson, & Liceaga, 2018). Akpossan et al. (2014) studied the functionality of defatted and non-defatted protein fractions in flours made from caterpillar larvae. A low solubility was observed in both flours, and this was a characteristic of the isoelectric pH of the proteins. Both flours exhibited low to no foaming properties; however, they showed good emulsion and water absorption characteristics. According to Nongonierma and FitzGerald (2017), the solubility of insect (B. mori) hydrolysates was not affected by a pH between 2 and 10. However, a higher solubility was observed after enzymatic hydrolysis by Alcalase of the insect proteins (G.sigillatus) at pH 3, 7, 8, and 10. The proteins of African cricket exhibit a low solubility at pH 3 and 4, suggesting that the isoelectric pH for African cricket is near or at pH 3. The extent of enzymatic hydrolysis of insect proteins can be determined by the degree of hydrolysis of the proteins, and has been reported in literature. The degree of hydrolysis reported among published studies range from 3% to 100%. This large variation in degree of hydrolysis is due to differences in insect species, nature of the harvest, and hydrolysis procedure used (protease, hydrolysis time, temperature, and pH) (Nongonierma & FitzGerald, 2017). For example, when the same hydrolysis protocol was applied to A. annulipes and L. migratoria, the degree of hydrolysis was 15.8% and 36.3% respectively (Zielińska, Karaś, & Jakubczyk, 2017).

4.6  PROTEIN FUNCTIONALITIES The functional properties of foods are the result of physiochemical interactions between food composition, structure, and molecular conformation (Siddiq, Nasir, Ravi, Dolan, & Butt, 2009). The functionality profile directly influences the form, texture, and appearance of food (Singh, Vanga, Orsat, & Raghavan, 2018). Therefore, it is imperative to investigate the functional properties of novel protein from sources such as insect materials to assess the feasibility of using them to substitute for traditional protein in relation to final product quality and ease of processing (van Huis, Dicke, & van Loon, 2015). The functional properties assessed for insect materials include protein solubility, emulsifying properties, surface coagulation, surface charge, surface hydrophobicity, water holding capacity, oil holding capacity, surface charge, colour, foaming capacity, and foam stability (Villaseñor, Enriquez-Vara, Urías-Silva, & Mojica, 2021). As these functionalities have a structural dependency, they are influenced by amino acid composition and physical structure (Gravel & Doyen, 2020). Protein functionalities are influenced by pH, temperature, protein concentration, protein fraction, prior treatment, ionic strength, and the dielectric constant of the medium. Other conditions that affect the functional properties of proteins in a food system include interactions with other macromolecules, and the type and extent of modification (physical, chemical, or enzymatic) that the food has been subjected to (Zayas, 2012). The functional properties of insect proteins have received increasing interest over recent years. This is due to the increased interest in sustainable protein sources, which includes utilizing protein sources which are perceived as more environmentally friendly (Gravel & Doyen, 2020). The insect species most extensively studied for their protein functionalities include mealworm (T. Molitor), black soldier fly (Hermetia illucens), and cricket (G. sigillatus) (Bußler, Rumpold, Jander, Rawel, & Schlüter, 2016; Gravel & Doyen, 2020; Zielińska, Karaś, & Baraniak, 2018).

4  •  Nutrient Content and Functionalities of Edible Insects  69

4.6.1 Protein Solubility Protein solubility refers to the amount of protein that is soluble under specified and controlled conditions (Wang, Wang, Feng, Wang, & Wang, 2021) and is expressed as the percentage of protein that is not precipitated when subjected to centrifugation. A protein is soluble in a solvent, such as an aqueous solution, when the attractive forces between the solvent and the protein and the forces of repulsion between proteins are maximized, preventing precipitation (Gravel & Doyen, 2020). In a food system, protein solubility is one of the fundamental protein functionalities, as protein solubility influences texture, colour, emulsification, foaming, and sensory properties (Haque, Timilsena, & Adhikari, 2016). Low protein solubility is undesirable as this restricts potential applications in liquified food systems containing conventional food solvents. Poor solubility of protein is one of the main disadvantages arising from novel protein sources (Wang et al., 2021). High protein solubility is desirable for the formation of colloidal systems with many potential applications. Further, there is a positive correlation between protein solubility and emulsifying and foaming functionalities (He, Wang, Feng, Chen, & Wang, 2020). The solubility of insect protein has been widely studied and the lowest protein solubility is typically found between pH 3 and 5, and highest solubility typically between pH 8 and 9 (Zielińska et al., 2018). Zielińska et al. (2018) found that G. sigillatus, Schistocerca gregaria, and T. molitor protein isolates exhibited the highest protein solubility, at 97%, 96%, and 90% respectively, at pH 11. Albeit with slight differences in methods and sample dilutions, these results of protein solubility compare with sources of plant protein such as Ginko biloba seed albumin fraction (Mundi & Aluko, 2012) and fenugreek seed (El Nasri & El Tinay, 2007).

4.6.2 Emulsifying Properties Emulsification is a surface-active property of protein in which a colloidal system is composed of a stable dispersion of two immiscible phases (McClements, 2004b). Protein emulsifiers function to stabilize emulsion droplets of oil in water emulsion systems by forming a charged layer around fat globules, resulting in repulsion at the oil/water interface (Culbertson, 2005). Emulsions are formed following high shearing conditions. From this, colloidal fat particles become dispersed in the continuous phase and there is a simultaneous adsorption of surface-active proteins at the oil/water interface. This stabilizes the interfacial surface layer by lowering the surface tension around the oil droplet, thereby aiding further dispersion (Singh & Ye, 2013). A conformational change termed interfacial denaturation follows, whereby the adsorbed surfaceactive proteins assume the most energetically favourable state (Singh & Ye, 2013). In this environment, hydrophobic functional groups are orientated toward the oil droplet while the hydrophilic functional groups face the water environment. The composition and structure of the protein influences this reorientation, whereby proteins which are more amphiphilic and have a relatively flexible and open confirmation will assume interfacial denaturation more readily and be more effective emulsifiers (Singh & Ye, 2013). Emulsifiers play a key role in stabilizing food systems by preventing coalescence, imparting texture, and binding flavour-active and bioactive substances (Culbertson, 2005). To prevent water loss, protein flours, concentrates, and isolates are typically used as emulsifiers in foods such as sausages, soups, and cakes (Villaseñor et al., 2021). Common emulsifying agents used in food include milk, egg, and soy protein (Gravel & Doyen, 2020; Lam & Nickerson, 2013). Emulsifying capacity, emulsion stability, and creaming all play a key role in the appropriateness of applications of protein emulsifying agents in the food industry. The differences in these functionalities are attributable to the protein amphiphilicity, solubility, and other components (Zielińska et al., 2018). The most common applications include milk, ice cream, cream, mayonnaise, and gravies (Culbertson, 2005). The emulsifying capacity measures the amount of oil that protein can emulsify and is quantified by the height of the emulsified layer in relation to the total volume of the system following emulsification (Lam & Nickerson, 2013; Pearce & Kinsella, 1978).

70  Edible Insects Processing for Food and Feed Emulsion capacity has been widely studied among insects, with Tenebrio molitor and Gryllodes sigillatus concentrates being found to display emulsifying capacities of 66.6% and 72.62% respectively (Zielińska et al., 2018). However, this is dependent on pH, and some species such as Mexican fruit fly larvae (A. ludens) display no emulsifying capacity, which may be explained by differences in amino acid sequence and protein structure (Del Valle, Mena, & Bourges, 1982; Villaseñor et al., 2021). Emulsion stability refers to the capacity of an emulsion to remain stable over time. It is well documented that, of the many commonly studied insects, Tenebrio molitor protein extract displays an emulsion stability of 50%–51% within 30 minutes after preparation of the extract (Kim, Setyabrata, Lee, Jones, & Kim, 2016; Zielińska et al., 2018). In comparison, fenugreek seed protein isolates display an emulsion stability of 61% (El Nasri & El Tinay, 2007). Typically, a higher emulsion stability is associated with a high protein content that contains a high proportion of hydrophobic amino acids (Zielińska et al., 2018). A study conducted by Kim et al. (2019) reported the emulsifying capacity and stability of 1% (w/v) Tenebrio molitor, Allomyrina dichotoma, and Protaetia brevitarsis seulensis protein extract solutions. Samples were homogenized at 18,000 rpm for 2 min and the emulsion capacity measured after holding for 10 min at 20°C. Emulsion stability was measured at 500 nm absorbance of the emulsion at different time points (10, 20, 30, 60, 90, and 120 minutes). From this, the salt soluble P. brevitarsis seulensis protein extract exhibited a significantly higher emulsion capacity of 100% compared to other samples, owing to a higher proportion of hydrophobic amino acids present, aiding emulsion formation by reducing interfacial surface tension (Mishyna et al., 2019). Correspondingly, the emulsion stability was higher for water soluble T. molitor and salt soluble A. dichotoma protein extracts due to the presence of a higher proportion of low molecular weight proteins (Kim et al., 2019).

4.6.3 Coagulation Surface coagulation refers to the ability of proteins to form aggregates at the liquid-air surface of a solution, usually following mechanical agitation (Heller & Peters, 1970; Liu, Ru, & Ding, 2012). In a colloidal system stabilized by hydrophobic interactions, the coagulation value decreases with a decrease in the zeta potential (the electrostatic potential at the electrical double layer surrounding a nanoparticle in solution) (Clogston & Patri, 2011). As typically determined by turbidity, heat induced coagulation is a measure of the percentage of total soluble protein that aggregates following heat treatment at a set temperature, such as 100°C (Mishyna et al., 2019). The application of heat results in protein aggregation by inducing protein denaturation and consequently exposing sulfhydryl groups, which become oxidized and form intermolecular disulphide bonds (Mishyna et al., 2019). Hydrophobic domains are also exposed by heat induced denaturation, which contribute to aggregation through hydrophobic interactions (Mishyna et al., 2019). In food applications, it is often desirable to have a higher proportion of coagulated proteins as they contribute structural functionality to food, and non-coagulated proteins may be lost to leeching (Nicolai, Britten, & Schmitt, 2011). However, the rheology of some proteins does not facilitate coagulation but can contribute to increasing viscosity, which is also another beneficial protein additive function (Nicolai et al., 2011). Compared with conventional sources such as animal- and plant-based sources, insect protein coagulation capabilities are not usually reported (Gravel & Doyen, 2020; Zielińska et al., 2018). This may be because the mechanism for the heat induced coagulation of insect proteins is not fully understood. However, pH and temperature are the main parameters documented that influence heat induced protein coagulation (Mishyna et al., 2019). If investigated, the methods used involve consideration of pH, the data are expressed as a function of pH, and the pH at which proteins exhibit the lowest solubility correlates with the highest coagulation (Mishyna et al., 2019). For example, at 55°C, honey bee brood proteins (Apis melliferais) exhibit significantly higher coagulation at the isoelectric pH of 5 (39.0 ± 2.0%), compared to that of other pH values (9.1–9.4%)

4  •  Nutrient Content and Functionalities of Edible Insects  71

4.6.4 Surface Charge Surface charge can be quantified by the zeta potential, the electric potential of colloidal dispersions of electrophoretically charged particles at the plane of shear under an electric field (Kaszuba, Corbett, Watson, & Jones, 2010). There is a positive relationship between zeta potential and the degree of particle flocculation, whereby higher electrostatic repulsion yields higher particle dispersion and therefore an absolute zeta potential (Thaiphanit, Schleining, & Anprung, 2016). In a food system, protein surface charge is a key determinant of electrostatic interactions in an aqueous environment as influenced by the presence of salts and counterions (Karaca, Low, & Nickerson, 2011). There is a positive correlation with negative surface charge and protein solubility owing to water binding by acidic amino acids (Kramer, Shende, Motl, Pace, & Scholtz, 2012). Surface charge is not usually investigated when establishing the protein functionality profile of edible insects (Kim, Yong, Chun, et al., 2020). However, as Tenebrio molitor is the most extensively studied edible insect larvae globally, the surface hydrophobicity of the comprising protein has been determined as a function of pH and ionic strength. The results obtained show that increased aggregation due to a lack of repulsive forces results in a lower zeta potential (Azagoh et al., 2016). There is a negative relationship between the pH of proteins and the zeta potential, whereby a decrease in pH correlates with a decrease in zeta potential and therefore protein solubility. Additionally, as the ionic strength increases, the zeta potential decreases due to salts binding water, thus hindering protein–water interactions and decreasing solubility (Azagoh et al., 2016). Mishyna et al. (2019) found that near the pI of heated honey bee brood protein (pH 4.5–5), there was a significant decrease in surface charge, which could be explained by decreased electrostatic repulsion. Interestingly, there was no significant difference in protein surface charge between pH 3 and 9 on unheated honey bee brood protein samples (Mishyna et al., 2019).

4.6.5 Surface Hydrophobicity Surface hydrophobicity is a structure-dependent protein functionality that can be determined by fluorescence spectroscopy, which refers to the relative proportion of hydrophobic functional groups at the protein surface. Surface hydrophobicity therefore illustrates the relationship between the exterior surface of protein and the surrounding aqueous environment (Wang, Li, Jiang, Qi, & Zhou, 2014). Surface hydrophobicity plays a key role in determining many other protein functionalities that rely on protein–water interactions, including foaming, water holding capacity, oil holding capacity, solubility, and emulsifying capability (Hua, Cui, Wang, Mine, & Poysa, 2005). In the native state of a protein, hydrophobic functional groups are buried. Therefore, following protein unfolding, there is a positive relationship between the proportion of hydrophobic functional groups located at the protein surface and the surface hydrophobicity (Wang et al., 2014). Surface hydrophobicity is important in determining the suitability of insect proteins in food systems, in particular with regard to interactions with other components. However, surface hydrophobicity is often not determined when analyzing the functionality profile of a protein system, as there is no standardized method (Konieczny & Uchman, 2002). In a study conducted by Okagu, Verma, McClements, and Udenigwe (2020), the surface hydrophobicity of Tenebrio molitor protein was measured using an 8-anilino-1-naphthalenesulfonic acid (ANS) assay. Results revealed that the surface hydrophobicity of Tenebrio molitor was comparable to that of traditional protein sources such as soy concentrate, whey concentrates, and egg white protein preparations. To investigate the surface hydrophobicity of honey bee brood protein, Mishyna et al. (2019) measured the fluorescence intensity at 365 nm (excitation) and 465 nm (emission) of heated (85°C, 30 min) and unheated protein samples. They found that at pH 7 and 9, the surface hydrophobicity of unheated protein was significantly lower – with values of 55.5 ± 3.2 and 31.3 ± 3.7 respectively – than at pH 3 and 5. At the lower pH values, it was suggested that partial denaturation resulted in increased exposure of hydrophobic domains with higher surface hydrophobicity.

72  Edible Insects Processing for Food and Feed

4.6.6 Water Holding Capacity Water holding capacity (WHC) describes the amount of free or bound water retained by a protein matrix and physiochemically against gravity per gram of sample following emulsion (Bußler et al., 2016). Because water is polar, water holding capacity is directly correlated to the relative amount of hydrophilic proteins present in a system (Bußler et al., 2016). The WHC of proteins is influenced by protein concentration, pH, ionic strength, temperature, and storage conditions (Kinsella, 1979; Verhoeckx et al., 2014). There is a positive relationship between WHC and protein concentration owing to an increased number of functional groups and surface area available for interaction with water. As with protein solubility, the WHC of protein decreases near the isoelectric pH (pI) as protein–protein interactions are maximized due to the lack of charge-induced repulsion (Gravel & Doyen, 2020). WHC plays a key role in food formulation by influencing mouthfeel, water activity, and the processing conditions used (Aryee, Agyei, & Udenigwe, 2018). In a food system, the interactions between different components need to be carefully formulated, as a high WHC of protein can potentially dehydrate other components (Kinsella, 1979; Kotoki & Deka, 2010). Hence, the WHC of proteins limits their application in food processing. For example, water retention is one of the first processing steps when manufacturing baked goods, and if the WHC of the protein is poor, the protein will not be suitable for the application (Kinsella, 1979; Kotoki & Deka, 2010). A high WHC of whole dried insect protein powder would be preferable to a protein isolate fraction in food applications, because crude protein isolates are more cost effective to produce compared to protein extract concentrates. Crude protein isolates also offer additional nutritional value contributed by vitamins, minerals, fatty acids, and chitin (Zielińska et al., 2018). The potential for insect flours to replace traditionally used flours is encouraging; however, the WHC of each species protein needs to be individually assessed, owing to differences in the protein profile and thus the WHC (Villaseñor et al., 2021). Zielińska et al. (2018) centrifuged 5% dispersions of Tenebrio molitor and Rhynchophorus phoenicis protein concentrate samples at 8000 g for 15 min and reported WHC values of 3.95 mL/g and 2.82 mL/g respectively. The WHC of Tenebrio molitor and Rhynchophorus phoenicis concentrates were comparable to that of milk and soy concentrates, which have been reported to display WHC values of 3.32 mL/g and 2.27 mL/g (Gravel & Doyen, 2020).

4.6.7 Oil Holding Capacity Aided by non-polar and hydrophobic proteins, oil holding capacity (OHC) describes the ability of a protein matrix to absorb oil and retain lipids in an emulsion (Kinsella, 1979; Villaseñor et al., 2021). Oil can be absorbed by the protein matrix by capillary action resulting from interactions between hydrophobic residues and hydrocarbon chains which comprise fat (Zielińska et al., 2018). OHC is closely related to the emulsifying properties where small, low density, hydrophobic proteins bind more fat and thus exhibit a higher OHC (Aryee et al., 2018). OHC is important for imparting desirable mouthfeel, flavour, and texture in foods such as baked goods and imitative meats (El Nasri & El Tinay, 2007). Insect protein displays high oil holding that is attributable to differences in amino acid composition, and therefore has the potential to replace traditional protein (Villaseñor et al., 2021; Zielińska et al., 2018). For example, following a 15 minute centrifugation at 8000 g, a 5% dispersion of Tenebrio molitor concentrate in vegetable oil was found to have an OHC of 2.74 g/g. This is higher than that of both animal and plant sourced material, such as egg white and soy protein concentrate, which have respective values of 1.35 g/g and 1.33 g/g (Gravel & Doyen, 2020; Zielińska et al., 2018).

4.6.8 Colour The colour of samples can be assessed using a colourimeter, such as a Miniscan XE Plus (HunterLab, Reston Virginia, USA), which generates L*, a* and b* values. L* denotes lightness, which has a ranges of

4  •  Nutrient Content and Functionalities of Edible Insects  73 0–100 (black to white). The parameters a∗ and b* are the two chromatic components denoting green to red and blue to yellow, respectively (León, Mery, Pedreschi, & León, 2006). Colour is an important parameter to investigate because colour is one of the main quality criteria assessed by consumers that dictates acceptance (León et al., 2006). The colour of insect proteins is highly dependent on the species and the innate molecular composition (Yi et al., 2013). Processing parameters such as alkalinity levels can also influence the colour of insect powder. This is demonstrated by edible grasshopper (Apis. Mellifera) and honey bee (Schistocera. Gregaria), for which it was shown that a decrease in L* and b* values and an increase in a* occurred following alkaline extraction of the proteins (Mishyna et al., 2019). This phenomenon can be explained by oxidation of phenolic compounds in insect cuticles, resulting in a lighter colour (Atkinson, Brown, & Gilby, 1973; Mishyna, Martinez, Chen, & Benjamin, 2019). The colour of edible insect materials is also sometimes investigated as a secondary parameter to quantify other parameters such as pigment content and coagulation of protein in supernatants extracted from edible insect materials (Kim, Yong, Jang, Kim, & Choi, 2020). When insect material was incorporated into food products such as energy bars, consumers had a higher product acceptance when the colour of the food matches the expected characterizing flavour. Hence, foods containing insect flour would have a higher acceptance to the consumer if the colour, texture, and flavour attributes match, for example as in chocolate (Clarkson, Mirosa, & Birch, 2018).

4.6.9 Foaming Properties Food foams are essentially air bubbles entrapped at the water/air interface boundary (Damodaran, 2017; Kinsella, 1981). Protein foams have important applications in food systems for foods such as in meringues, soufflés, whipped toppings, and leavened bakery products, where a light aerated texture is needed (McClements, 2004a). Foam formation is highly dependent on surface-active protein surfactants which reduce the surface tension at the interface (Damodaran, 2017). There are three steps involved in the formation of a protein foaming system. Firstly, soluble globular proteins diffuse to the air–water interface, concentrate, and reduce surface tension. While at the interface, some polypeptides are unfolded and are then reoriented to form the most energetically favourable arrangement. A continuous film entrapping air bubbles (foam) is then formed (Kinsella, 1981). Foams contribute to the mouthfeel, rheological structure, and flavour release in a food system (Damodaran, 2017; Kinsella, 1979). Foaming capacity is measured as the height of the foam in a system following timed agitation in a buffer solution (Mishyna et al., 2019). However, foam capacity is not a meaningful measurement until related to stability, which refers to the height retention of the foam following periodic time intervals (Mishyna et al., 2019). The most common proteins utilized in foods for their foaming properties include egg white and milk proteins (Kinsella, 1981; Wierenga, van Norél, & Basheva, 2009). The foaming properties of insect materials have been investigated as a potential replacement for traditional materials. Results show that milk protein concentrate has a foam capacity and stability of 88.7% and 55.9% respectively, which is lower than egg white flour, exhibiting 159.1% and 145% respectively. The foaming capacity and stability of Tenebrio molitor protein is significantly lower than milk and egg sources, with the insect protein concentrates exhibiting a foaming capacity of 32.7% and a stability value of 30.3% (Gravel & Doyen, 2020). The results obtained for the foaming properties of Tenebrio molitor protein agree with those reported by Zielińska et al. (2018) and Yi et al. (2013), who reported poor foaming properties for protein hydrolysates of five different species of edible insects. These findings potentially limit the applications of insect protein as a foaming agent. Similar findings were reported by Zielińska et al. (2018), who measured the foaming capacity and stability of Gryllodes sigillatus, Schistocerca gregaria, and Tenebrio molitor protein, where 1% suspensions of protein were homogenized at 16,000 rpm for 2 min before being transferred to measuring cylinders of 20 mm diameter. The highest foaming capacity was reported as 99% for G. sigillatus. There was significant variability in the foam stability for the three species studied, ranging between 6.17% and 99.0%. Enzymatic hydrolysis of insect proteins has been carried out to generate bioactive peptides or to improve the techno-functional properties of the protein. The bio- and techno-functionalities of edible insect protein are discussed in the following sections.

74  Edible Insects Processing for Food and Feed

4.7 BIOACTIVITIES Proteins offer the starting material for the generation of peptides, and these peptides constitute two groups (Aluko, 2018). Firstly, from a nutritional aspect, some provide small peptides and free amino acids that can be absorbed into the body for growth and maintenance. Secondly, some of the peptides exhibit bioactive properties, resulting in various health benefits that can positively influence physiological functions such as inhibition of cancer cell growth, reduction of blood pressure, oxidative stress relief, reduction of hypertension, and amelioration of insulin resistance. This group of peptides are known as bioactive peptides (Chakrabarti, Guha, & Majumder, 2018). Bioactive peptides are generated from various protein sources and are in high demand globally. The trend is for an increased diversification of sources of high-quality food-grade protein, including protein from non-traditional (to much of the Western world) sources such as insects. There is also an emphasis on the range of biological properties that can be obtained from peptides derived from these alternative protein sources (see Figure 4.5) (Nongonierma & FitzGerald, 2017).

4.7.1 Antioxidant Properties Reactive oxygen species (ROS) play a crucial role in stimulating mitochondrial cytochrome c release and initiating apoptosis (Azad & Iyer, 2014). Antioxidant peptides are cell-permeable and are highly effective at reducing intracellular ROS and inhibiting cell death caused by the oxidants. ROS are important in health, but they become a problem (requiring the intervention of antioxidants) when their levels become elevated. These antioxidants then help in the reduction of oxidative damage and decrease the risk of developing chronic illnesses such as cancer or cardiovascular diseases (Agyei, Danquah, Sarethy, & Pan, 2015). According to Liu et al. (2017), there is increasing interest in the search for antioxidant peptides derived from edible insects. The antioxidant activity of peptides derived from Spodotera littoralis were reported by Lieselot Vercruysse, Smagghe, Beckers, and van Camp (2009). The protein hydrolysis was performed using thermolysin, Alcalase, and a simulated gastrointestinal digestion involving several peptidases. The antioxidant assays conducted were FRAP (ferric reducing activity) and DPPH (2,2-diphenylpicrylhydrazyl

FIGURE 4.5  Overview of the process of the production of bioactive peptides from insect protein and the potential properties of the bioactive peptides.

4  •  Nutrient Content and Functionalities of Edible Insects  75 radical scavenging activity). The protein hydrolysate with the highest radical scavenging activity was that produced by simulated gastrointestinal digestion, exhibiting 14% antioxidant activity measured by FRAP and 24% by DPPH. In another study, the antioxidant properties of peptides obtained by in vitro gastrointestinal digestion of edible insects were reported by Zielińska, Karaś, and Jakubczyk (2017). The antioxidant activities of hydrolysates from locust, super worm, and cricket were evaluated based on free radical scavenging activity metrics, such as ion chelating activity and reducing power assay. The enzymes used for the hydrolysis were alpha-amylase, pepsin, a simulator of saliva, and intestinal juice of the human digestive system. Amphhiacusta annulipes hydrolysates presented a high antioxidant activity measured by antiradical activity against DPPH (IC50 = 19.11 g/ml), Fe2+ chelation ability (58.82%) and reducing power (0.652). The highest ability to chelate Cu2+ was exhibited by Locusta migratoria peptides (86.05%). It was observed that there was a high concentration of peptides after digestion (Zielińska, Karaś, et al., 2017), indicating that edible insect protein hydrolysates could be a good source of peptides exhibiting antioxidant activity. However, a more physiologically relevant in vivo analysis of the antioxidant properties of insect protein hydrolysates needs to be conducted to confirm the in vitro findings.

4.7.2 Angiotensin Converting Enzyme (ACE) Inhibitory Activity Hypertension, or high blood pressure, is a common health presentation worldwide, and is also a major risk factor for cardiovascular disease (Nongonierma & FitzGerald, 2017). One of the mechanisms for controlling hypertension is the inhibition of enzymes such as angiotensin-I converting enzyme (ACE), which plays a key role in the control of blood pressure. Antihypertensive peptides inhibit the action of ACE and reduce arterial blood pressure. Hall et al. (2018) studied cricket protein hydrolysates and showed that the ACE inhibition capability of these hydrolysates increased with the degree of hydrolysis. Cricket protein hydrolysates with a degree of hydrolysis values of more than 40% inhibited 80% of ACE activity, and those with a degree of hydrolysis ranging from 60% to 85% inhibited over 90% of ACE activity. In another study, Vercruysse et al. (2005) observed the presence of inhibitory activity against ACE in protein hydrolysates obtained from various insects such as Bombyx mori, Spodoptera littoralis, Schistocera gregaria, and Bombus terrestris. Simulated gastrointestinal digestion was performed using pepsin, trypsin, and chymotrypsin. This study verified that insect protein hydrolysates can generate inhibitory activity against ACE. ACE inhibitor peptides have been identified from food sources other than edible insects. Nonetheless, the activity of ACE inhibitory peptides from edible insects was comparable to that of other dietary protein hydrolysate peptides. For example, the IC50 values of ACE inhibiting properties for animal and plant protein hydrolysates range from 0.028 to 2.8 mg/ml and from 0.01 to 8.2 mg/ml respectively (Daskaya-Dikmen, Yucetepe, Karbancioglu-Guler, Daskaya, & Ozcelik, 2017; Vercruysse et al., 2005), while B. mori hydrolysates had ACE inhibiting IC50 values ranging from 0.7 to 0.9 mg/ml (Staljanssens et al., 2011). In addition, a novel ACE inhibitor peptide (Try-Ala-Asn, IC50: 46 uM) was identified in a hydrolysate of Tenebrio molitor larvae protein when hydrolyzed by Alcalase (Dai, Ma, Luo, & Yin, 2013). Several edible insects are yet to be studied for the potential of protein hydrolysates to generate ACE inhibiting peptides.

4.7.3 Antimicrobial Properties Naturally occurring antimicrobial peptides are ribosomal synthesized polypeptides consisting of about 30–60 amino acids. They are substantially cationic, stable to heat (100°C, 15 min), and have isoelectric pHs between 8.9 and 10.7 (Li, Xiang, Zhang, Huang, & Su, 2012). Edible insects are known to be good sources of antimicrobial peptides such as the defensins, lebocins, attacins, and other proline-rich peptides (Yi, Chowdhury, Huang, & Yu, 2014). Rahnamaeian et al. (2015) studied two antimicrobial peptides, abaecin (a proline-rich peptide) and hyme-noptacin (a glycine-rich peptide) that were isolated from Bombus

76  Edible Insects Processing for Food and Feed pascuorum and Bombus terrestris. When abaecin (IC50 = 20uM) and hyme-noptacin (IC50 = 0.5uM) were combined, the antibacterial effects were enhanced by 18%. Hence, edible insects are potentially a good source of antimicrobial peptides, but there is a need for further investigation (Jantzen et al., 2020).

4.7.4 Antidiabetic Properties Diabetes is a metabolic disorder that is associated with increasing formation of free radicals, which creates oxidative stress. According to the World Health Organisation (WHO), about 3% of the world’s population have diabetes, and this figure is expected to increase to 5.4% by the year 2025 (King, Aubert & Herman, 1998). The antidiabetic properties of peptides are not well studied compared with other bioactivities such as antioxidant and antihypertensive activities. Nevertheless, a recent study assessed the alpha-glucosidase inhibitory activity of peptides derived from hydrolysates of Bombyx mori, protein (Du et al., 2016), which identified two bioactive peptides exhibiting in vitro alpha-glucosidase inhibition. The peptides were SerGln-Ser-Pro-Ala and Gln-Pro-Gly-Arg, having IC50 values of 20 and 65.8 uM, respectively. Interestingly, the Ser-Gln-Ser-Pro-Ala had a similar activity to that of acarbose, the positive control used in the assay. The fact that there are only a few in vitro and in vivo studies of insect protein hydrolysates in relation to antidiabetic activity indicates a need for future studies.

4.8 CONCLUSION While insects have been consumed by many cultures for a considerable time, an increasing interest globally is emerging as to the prospects for edible insects as a food source. Edible insect material is high in protein, fat, minerals, and energy content. The amino acid and fatty acid compositions are reported to meet the nutritional requirements for both adults and children. The fatty acid compliment is high in PUFAs (polyunsaturated fatty acids), which are known to be very beneficial to human health, along with several micronutrients such as phosphorus, calcium, zinc, and potassium; additionally, insect material is low in sodium. When compared with other protein sources such as dairy and meat, the protein from edible insect material, although at a slightly lower level in some species, is still present at substantial levels. A number of studies of the techno-functionality of insect derived protein have been reported which indicate the potential of insect protein to contribute to the functionality of formulated foods; more studies are required. Studies conducted with insect derived protein hydrolysates indicate that a range of peptide bioactivities are exhibited, indicating the health promoting potential of insect materials. The studies conducted so far indicate a considerable potential for development of edible insect material as an alternative food product to support the increasing demand for future human and animal nutrition. There is also an emphasis on the range of biological properties that can be obtained from peptides derived from these alternative protein sources. However, various bioactive properties and techno-functionalities have not been fully elucidated. Future studies must therefore address these research gaps.

REFERENCES Agbidye, F., Ofuya, T., & Akindele, S. (2009). Some edible insect species consumed by the people of Benue State, Nigeria. Pakistan Journal of Nutrition, 8(7), 946–950. Agyei, D., Danquah, M. K., Sarethy, I. P., & Pan, S. (2015). Antioxidative peptides derived from food proteins. In Free radicals in human health and disease (pp. 417–430). Springer.

4  •  Nutrient Content and Functionalities of Edible Insects  77 Akpossan, R., Digbeu, Y., Parfait, J., Eugène, N., Dabonné, S., Dué, E., & Kouamé, L. (2014). Nutritional characteristics of the caterpillars (Imbrasia oyemensis) from Côte d’Ivoire. International Journal of Recent Biotechnology, 2, 1–5. Akpossan, R. A., Dué, E. A., Kouadio, J. P. E., & Kouamé, L. P. (2009). Nutritional value and physico-chemical characterization of the fat of the caterpillar (Imbrasia oyemensis) dried and sold at the Adjamé market in Abidjan, Côte d'Ivoire. Journal of Animal and Plant Sciences (JAPS), 3(3), 243–250. Aluko, R. (2018). Food protein-derived peptides: Production, isolation, and purification. In Proteins in food processing (pp. 389–412). Elsevier. Aryee, A., Agyei, D., & Udenigwe, C. (2018). Impact of processing on the chemistry and functionality of food proteins. In Proteins in food processing (pp. 27–45). Elsevier. Atkinson, P., Brown, W., & Gilby, A. (1973). Phenolic compounds from insect cuticle: Identification of some lipid antioxidants. Insect Biochemistry, 3(11), 309–315. Azad, N., & Iyer, A. (2014). Reactive oxygen species and apoptosis. In System biology of free radical and antioxidants (pp. 113–127). Springer-Verlag Berlin Heidelberg. Azagoh, C., Ducept, F., Garcia, R., Rakotozafy, L., Cuvelier, M. E., Keller, S., & Mezdour, S. (2016). Extraction and physicochemical characterization of Tenebrio molitor proteins. Food Research International, 88, 24–31. https://doi​.org​/10​.1016​/j​.foodres​.2016​.06​.010 Barker, D., Fitzpatrick, M. P., & Dierenfeld, E. S. (1998). Nutrient composition of selected whole invertebrates. Zoo Biology, 17(2), 123–134. Bawa, M., Songsermpong, S., Kaewtapee, C., & Chanput, W. (2020). Effects of microwave and hot air oven drying on the nutritional, microbiological load, and color parameters of the house crickets (Acheta domesticus). Journal of Food Processing and Preservation, 44(5), e14407. https://doi​.org​/10​.1111​/jfpp​.14407 Belluco, S., Losasso, C., Maggioletti, M., Alonzi, C. C., Paoletti, M. G., & Ricci, A. (2013). Edible insects in a food safety and nutritional perspective: A critical review. Comprehensive Reviews in Food Science and Food Safety, 12(3), 296–313. https://doi​.org​/10​.1111​/1541​- 4337​.12014 Bi, F. G. F., Meite, A., Dally, T., Ouattara, H., Kouame, K. G., & Kati-Coulibaly, S. (2015). Biochemical and nutritional study of dried powder from caterpillars Imbrasia oyemensis eaten in west of Côte d’Ivoire. Journal of Applied Biosciences, 96, 9039–9048. Bosch, G., Zhang, S., Oonincx, D. G., & Hendriks, W. H. (2014). Protein quality of insects as potential ingredients for dog and cat foods. Journal of Nutritional Science, 3, e29. Boulos, S., Tännler, A., & Nyström, L. (2020). Nitrogen-to-protein conversion factors for edible insects on the Swiss market: T. molitor, A. domesticus, and L. migratoria. Frontiers in Nutrition, 7(89). https://doi​.org​/10​.3389​/fnut​ .2020​.00089 Bruni, L., Pastorelli, R., Viti, C., Gasco, L., & Parisi, G. (2018). Characterisation of the intestinal microbial communities of rainbow trout (Oncorhynchus mykiss) fed with Hermetia illucens (black soldier fly) partially defatted larva meal as partial dietary protein source. Aquaculture, 487, 56–63. Bußler, S., Rumpold, B. A., Jander, E., Rawel, H. M., & Schlüter, O. K. (2016). Recovery and techno-functionality of flours and proteins from two edible insect species: Meal worm (Tenebrio molitor) and black soldier fly (Hermetia illucens) larvae. Heliyon, 2(12), e00218. https://doi​.org​/10​.1016​/j​.heliyon​.2016​.e00218 Chakrabarti, S., Guha, S., & Majumder, K. (2018). Food-derived bioactive peptides in human health: Challenges and opportunities. Nutrients, 10(11), 1738. Chatterjee, R., Yeh, H.-C., Edelman, D., & Brancati, F. (2011). Potassium and risk of Type 2 diabetes. Expert Review of Endocrinology & Metabolism, 6(5), 665–672. Chia, S. Y., Tanga, C. M., Osuga, I. M., Cheseto, X., Ekesi, S., Dicke, M., & van Loon, J. J. (2020). Nutritional composition of black soldier fly larvae feeding on agro‐industrial by‐products. Entomologia Experimentalis et Applicata, 168(6–7), 472–481. Churchward-Venne, T. A., Pinckaers, P. J., van Loon, J. J., & van Loon, L. J. (2017). Consideration of insects as a source of dietary protein for human consumption. Nutrition Reviews, 75(12), 1035–1045. Clarkson, C., Mirosa, M., & Birch, J. (2018). Consumer acceptance of insects and ideal product attributes. British Food Journal, 120(12), 2898–2911. Clogston, J. D., & Patri, A. K. (2011). Zeta potential measurement. In Characterization of nanoparticles intended for drug delivery (pp. 63–70). Springer. Costa, S., Pedro, S., Lourenço, H., Batista, I., Teixeira, B., Bandarra, N. M., & Pires, C. (2020). Evaluation of Tenebrio molitor larvae as an alternative food source. NFS Journal, 21, 57–64. https://doi​.org​/10​.1016​/j​.nfs​ .2020​.10​.001 Crossgrove, J., & Zheng, W. (2004). Manganese toxicity upon overexposure. NMR in Biomedicine, 17(8), 544–553. Culbertson, J. D. (2005). Food protein functionality. In Handbook of food science, technology, and engineering, 4.

78  Edible Insects Processing for Food and Feed Da Silva Lucas, A. J., De Oliveira, L. M., Da Rocha, M., & Prentice, C. (2020). Edible insects: An alternative of nutritional, functional and bioactive compounds. Food Chemistry, 311, 126022. Dai, C., Ma, H., Luo, L., & Yin, X. (2013). Angiotensin I-converting enzyme (ACE) inhibitory peptide derived from Tenebrio molitor (L.) larva protein hydrolysate. European Food Research and Technology, 236(4), 681–689. Damodaran, S. (2017). Protein-stabilized foams and emulsions. In Food proteins and their applications (pp. 57–110). CRC Press. Daskaya-Dikmen, C., Yucetepe, A., Karbancioglu-Guler, F., Daskaya, H., & Ozcelik, B. (2017). Angiotensin-Iconverting enzyme (ACE)-inhibitory peptides from plants. Nutrients, 9(4), 316. Dawkins, M. S. (2006). A user’s guide to animal welfare science. Trends in Ecology & Evolution, 21(2), 77–82. https://doi​.org​/10​.1016​/j​.tree​.2005​.10​.017 De Backer, C. J. S., & Hudders, L. (2015). Meat morals: Relationship between meat consumption consumer attitudes towards human and animal welfare and moral behavior. Meat Science, 99, 68–74. https://doi​.org​/10​.1016​/j​ .meatsci​.2014​.08​.011 De Marco, M., Martínez, S., Hernandez, F., Madrid, J., Gai, F., Rotolo, L., & Schiavone, A. (2015). Nutritional value of two insect larval meals (Tenebrio molitor and Hermetia illucens) for broiler chickens: Apparent nutrient digestibility, apparent ileal amino acid digestibility and apparent metabolizable energy. Animal Feed Science and Technology, 209, 211–218. https://doi​.org​/10​.1016​/j​.anifeedsci​.2015​.08​.006 DeFoliart, G. R. (1997). An overview of the role of edible insects in preserving biodiversity. Ecology of Food and Nutrition, 36(2–4), 109–132. Del Valle, F., Mena, M., & Bourges, H. (1982). An investigation into insect protein. Journal of Food Processing and Preservation, 6(2), 99–110. Diener, S., Zurbrügg, C., & Tockner, K. (2009). Conversion of organic material by black soldier fly larvae: Establishing optimal feeding rates. Waste Management & Research, 27(6), 603–610. Dobermann, D., Swift, J. A., & Field, L. M. (2017). Opportunities and hurdles of edible insects for food and feed. Nutrition Bulletin, 42(4), 293–308. https://doi​.org​/10​.1111​/nbu​.12291 Du, S., Zhang, J., Zhou, W. T., Li, Q. X., Greene, G. W., Zhu, H. J., & Wang, X. G. (2016). Interactions between fibroin and sericin proteins from Antheraea pernyi and Bombyx mori silk fibers. Journal of Colloid and Interface Science, 478, 316–323. EFSA Panel on Nutrition, N. F., Allergens, F., Turck, D., Castenmiller, J., De Henauw, S., Hirsch-Ernst, K. I., & Knutsen, H. K. (2021). Safety of dried yellow mealworm (Tenebrio molitor larva) as a novel food pursuant to Regulation (EU) 2015/2283. EFSA Journal, 19(1), e06343. https://doi​.org​/10​.2903​/j​.efsa​.2021​.6343 Ekpo, K. (2011). Effect of processing on the protein quality of four popular insects consumed in Southern Nigeria. Archives of Applied Science Research, 3(6), 307–326. El Nasri, N. A., & El Tinay, A. H. (2007). Functional properties of fenugreek (Trigonella foenum graecum) protein concentrate. Food Chemistry, 103(2), 582–589. https://doi​.org​/10​.1016​/j​.foodchem​.2006​.09​.003 Elemo, B. O., Elemo, G. N., Makinde, M., & Erukainure, O. L. (2011). Chemical evaluation of African palm weevil, Rhychophorus phoenicis, larvae as a food source. Journal of Insect Science, 11(1), 146. EUFIC. (2017). Sustainable protein: Meeting future needs. https://www​.eufic​.org​/en​/food​-production​/article​/sustainable​-protein​-meeting​-futureneeds FAO. (2009). How to feed the world in 2050. http://www​.fao​.org​/fileadmin​/templates​/wsfs​/docs​/expert​_paper​/ How​ _to​_ Feed​_the​_World​_in​_2050​.pdf (accessed July 12, 2021). FAO, S. H., Gerber, P., Wassenaar, T., Castel, V., Rosales, M., & De Haan, C. (2006). Livestock’s long shadow: Environmental issues and options. FAO. FAO/WHO. (2007). Food labelling. http://www​.fao​.org​/documents​/card ​/en ​/c​/5eb2d33b​-d2ad​-505e​-898a​ -6aeba60ec896/ (accessed on July 23, 2021). Finke, M. D. (2002). Complete nutrient composition of commercially raised invertebrates used as food for insectivores. Zoo Biology, 21(3), 269–285. Finke, M. D. (2007). Estimate of chitin in raw whole insects. Zoo Biology, 26(2), 105–115. Fombong, F. T., Kinyuru, J., Ng’ang’a, J., Ayieko, M., Tanga, C. M., Vanden Broeck, J., & Van Der Borght, M. (2021). Affordable processing of edible orthopterans provides a highly nutritive source of food ingredients. Foods, 10(1), 144. Food and Agriculture Organization, & World Health Organization. (2007). Protein and amino acid requirements in human nutrition: Report of a joint FAO/WHO/UNU expert consultation. World Health Organization. Gérard, F. b. f., Alassane, M., Théodor, D., Howélé, O., Koffi Gustave, K., & Séraphin, K. C. Biochemical and nutritional study of powder from caterpillars Imbrasia oyemensis eaten in west of Côte d’Ivoire. Ghosh, S., Jung, C., & Meyer-Rochow, V. B. (2016). Nutritional value and chemical composition of larvae, pupae, and adults of worker honey bee, Apis mellifera ligustica as a sustainable food source. Journal of Asia-Pacific Entomology, 19(2), 487–495. https://doi​.org​/10​.1016​/j​.aspen​.2016​.03​.008

4  •  Nutrient Content and Functionalities of Edible Insects  79 Ghosh, S., Jung, C., & Meyer-Rochow, V. B. (2017). Snail as mini-livestock: Nutritional potential of farmed Pomacea canaliculata (Ampullariidae). Agriculture and Natural Resources, 51(6), 504–511. https://doi​.org​/10​.1016​/j​ .anres​.2017​.12​.007 Gibson, R. S. (2015). Dietary-induced zinc deficiency in low income countries: Challenges and solutions the avanelle kirksey lecture at Purdue University. Nutrition Today, 50(1), 49–55. Gravel, A., & Doyen, A. (2020). The use of edible insect proteins in food: Challenges and issues related to their functional properties. Innovative Food Science & Emerging Technologies, 59, 102272. https://doi​.org​/10​.1016​ /j​.ifset​.2019​.102272 Haas, M., & Forbush, B. (2000). The Na-K-Cl cotransporter of secretory epithelia. Annual Review of Physiology, 62(1), 515–534. Hall, F., Johnson, P. E., & Liceaga, A. (2018). Effect of enzymatic hydrolysis on bioactive properties and allergenicity of cricket (Gryllodes sigillatus) protein. Food Chemistry, 262, 39–47. Hall, F. G., Jones, O. G., O'Haire, M. E., & Liceaga, A. M. (2017). Functional properties of tropical banded cricket (Gryllodes sigillatus) protein hydrolysates. Food Chemistry, 224, 414–422. Hamm, M. W. (2018). Sustainable protein provisioning. Nature Sustainability, 1(12), 733–734. Haque, M. A., Timilsena, Y. P., & Adhikari, B. (2016). Food proteins, structure, and function. In Reference Module in Food Science. Elsevier. He, J., Wang, R., Feng, W., Chen, Z., & Wang, T. (2020). Design of novel edible hydrocolloids by structural interplays between wheat gluten proteins and soy protein isolates. Food Hydrocolloids, 100, 105395. https://doi​.org​/10​ .1016​/j​.foodhyd​.2019​.105395 Heller, W., & Peters, J. (1970). Mechanical and surface coagulation: I. Surface coagulation of α-FeOOH sols. Journal of Colloid and Interface Science, 32(4), 592–605. https://doi​.org​/10​.1016​/0021​-9797(70)90152-9 Higdon, J. (2003). An evidence-based approach to vitamins and minerals health benefits and intake recommendations. Thieme Medical Publishers, Inc. Hua, Y., Cui, S. W., Wang, Q., Mine, Y., & Poysa, V. (2005). Heat induced gelling properties of soy protein isolates prepared from different defatted soybean flours. Food Research International, 38(4), 377–385. Hurrell, R., & Egli, I. (2010). Iron bioavailability and dietary reference values. The American Journal of Clinical Nutrition, 91(5), 1461S–1467S. Igual, M., García-Segovia, P., & Martínez-Monzó, J. (2020). Effect of Acheta domesticus (house cricket) addition on protein content, colour, texture, and extrusion parameters of extruded products. Journal of Food Engineering, 282, 110032. Imathiu, S. (2020). Benefits and food safety concerns associated with consumption of edible insects. NFS Journal, 18, 1–11. Islam, M. M., & Yang, C.-J. (2017). Efficacy of mealworm and super mealworm larvae probiotics as an alternative to antibiotics challenged orally with Salmonella and E. coli infection in broiler chicks. Poultry Science, 96(1), 27–34. Jayanegara, A., Gustanti, R., Ridwan, R., & Widyastuti, Y. (2020). Fatty acid profiles of some insect oils and their effects on in vitro bovine rumen fermentation and methanogenesis. Italian Journal of Animal Science, 19(1), 1310–1317. Jones, L., Cooper, R., & Harding, R. (1972). Composition of mealworm Tenebrio molitor larvae. The Journal of Zoo Animal Medicine, 3(4), 34–41. Jongema, Y. (2017). List of edible insect species of the world. Laboratory of Entomology. Kavle, R. R., Pritchard, E. T. M., Bekhit, A. E. D. A., Carne, A., & Agyei, D. (2022). Edible insects: A bibliometric analysis and current trends of published studies (1953–2021). International Journal of Tropical Insect Science, 42(5), 3335–3355. Karaca, A. C., Low, N., & Nickerson, M. (2011). Emulsifying properties of chickpea, faba bean, lentil and pea proteins produced by isoelectric precipitation and salt extraction. Food Research International, 44(9), 2742–2750. Kaszuba, M., Corbett, J., Watson, F. M., & Jones, A. (2010). High-concentration zeta potential measurements using light-scattering techniques. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 368(1927), 4439–4451. Kim, H.-W., Setyabrata, D., Lee, Y. J., Jones, O. G., & Kim, Y. H. B. (2016). Pre-treated mealworm larvae and silkworm pupae as a novel protein ingredient in emulsion sausages. Innovative Food Science & Emerging Technologies, 38, 116–123. https://doi​.org​/10​.1016​/j​.ifset​.2016​.09​.023 Kim, S.-K., Weaver, C. M., & Choi, M.-K. (2017). Proximate composition and mineral content of five edible insects consumed in Korea. CyTA-Journal of Food, 15(1), 143–146. Kim, T.-K., Yong, H. I., Chun, H. H., Lee, M.-A., Kim, Y.-B., & Choi, Y.-S. (2020). Changes of amino acid composition and protein technical functionality of edible insects by extracting steps. Journal of Asia-Pacific Entomology, 23(2), 298–305.

80  Edible Insects Processing for Food and Feed Kim, T.-K., Yong, H. I., Jang, H. W., Kim, Y.-B., & Choi, Y.-S. (2020). Functional properties of extracted protein from edible insect larvae and their interaction with transglutaminase. Foods, 9(5). https://doi​.org​/10​.3390​/ foods9050591 Kim, T.-K., Yong, H. I., Jeong, C. H., Han, S. G., Kim, Y.-B., Paik, H.-D., & Choi, Y.-S. (2019). Technical functional properties of water- and salt-soluble proteins extracted from edible insects. Food Science of Animal Resources, 39(4), 643–654. https://doi​.org​/10​.5851​/ kosfa​.2019​.e56 King, H., Aubert, R. E., & Herman, W. H. (1998). Global burden of diabetes, 1995–2025: Prevalence, numerical estimates, and projections. Diabetes Care, 21(9), 1414–1431. King, D. E., Mainous, A. G., Geesey, M. E., & Woolson, R. F. (2005). Dietary magnesium and C-reactive protein levels. Journal of the American College of Nutrition, 24(3), 166–171. Kinsella, J. (1979). Functional properties of soy proteins. Journal of the American Oil Chemists' Society, 56(3Part1), 242–258. https://doi​.org​/10​.1007​/ BF02671468 Kinsella, J. (1981). Functional properties of proteins: Possible relationships between structure and function in foams. Food Chemistry, 7(4), 273–288. https://doi​.org​/10​.1016​/0308​-8146(81)90033-9 Kinyuru, J. N., Kenji, G. M., Njoroge, S. M., & Ayieko, M. (2010). Effect of processing methods on the in vitro protein digestibility and vitamin content of edible winged termite (Macrotermes subhylanus) and grasshopper (Ruspolia differens). Food and Bioprocess Technology, 3(5), 778–782. Kinyuru, J. N., Mogendi, J. B., Riwa, C. A., & Ndung'u, N. W. (2015). Edible insects—a novel source of essential nutrients for human diet: Learning from traditional knowledge. Animal Frontiers, 5(2), 14–19. Konieczny, P., & Uchman, W. (2002). Comparative characterization of surface hydrophobicity and other physicochemical properties of selected protein prepartions. Electronic Journal of Polish Agricultural Universities. Series Food Science and Technology, 5(2), #10. Kotoki, D., & Deka, S. C. (2010). Baking loss of bread with special emphasis on increasing water holding capacity. Journal of Food Science and Technology, 47(1), 128–131. https://doi​.org​/10​.1007​/s13197​- 010​- 0008-2 Kramer, R. M., Shende, V. R., Motl, N., Pace, C. N., & Scholtz, J. M. (2012). Toward a molecular understanding of protein solubility: Increased negative surface charge correlates with increased solubility. Biophysical Journal, 102(8), 1907–1915. https://doi​.org​/10​.1016​/j​.bpj​.2012​.01​.060 Kulma, M., Kouřimská, L., Homolková, D., Božik, M., Plachý, V., & Vrabec, V. (2020). Effect of developmental stage on the nutritional value of edible insects. A case study with Blaberus craniifer and Zophobas morio. Journal of Food Composition and Analysis, 92, 103570. https://doi​.org​/10​.1016​/j​.jfca​.2020​.103570 Lam, R. S., & Nickerson, M. T. (2013). Food proteins: A review on their emulsifying properties using a structure– function approach. Food Chemistry, 141(2), 975–984. Latunde-Dada, G. O., Yang, W., & Vera Aviles, M. (2016). In vitro iron availability from insects and sirloin beef. Journal of Agricultural and Food Chemistry, 64(44), 8420–8424. https://doi​.org​/10​.1021​/acs​.jafc​.6b03286 Laureati, M., Proserpio, C., Jucker, C., & Savoldelli, S. (2016). New sustainable protein sources: Consumers’ willingness to adopt insects as feed and food. Italian Journal of Food Science, 28(4), 652–668. Lautenschläger, T., Neinhuis, C., Kikongo, E., Henle, T., & Förster, A. (2017). Impact of different preparations on the nutritional value of the edible caterpillar Imbrasia epimethea from northern Angola. European Food Research and Technology, 243(5), 769–778. Lawal, K. G., Kavle, R. R., Akanbi, T. O., Mirosa, M., & Agyei, D. (2021). Enrichment in specific fatty acids profile of Tenebrio molitor and Hermetia illucens larvae through feeding. Future Foods, 3, 100016. https://doi​.org​/10​ .1016​/j​.fufo​.2021​.100016 León, K., Mery, D., Pedreschi, F., & León, J. (2006). Color measurement in L∗a∗b∗ units from RGB digital images. Food Research International, 39(10), 1084–1091. https://doi​.org​/10​.1016​/j​.foodres​.2006​.03​.006 Lewis, J. C., Snell, N. S., Hirschmann, D. J., & Fraenkel-Conrat, H. (1950). Amino acid composition of egg proteins. Journal of Biological Chemistry, 186(1), 23–35. https://doi​.org​/10​.1016​/S0021​-9258(18)56283-9 Li, Y., Xiang, Q., Zhang, Q., Huang, Y., & Su, Z. (2012). Overview on the recent study of antimicrobial peptides: Origins, functions, relative mechanisms and application. Peptides, 37(2), 207–215. Liaqat, F., & Eltem, R. (2018). Chitooligosaccharides and their biological activities: A comprehensive review. Carbohydrate Polymers, 184, 243–259. Liu, J., Ru, Q., & Ding, Y. (2012). Glycation a promising method for food protein modification: Physicochemical properties and structure, a review. Food Research International, 49(1), 170–183. https://doi​.org​/10​.1016​/j​ .foodres​.2012​.07​.034 Liu, X., Chen, X., Wang, H., Yang, Q., ur Rehman, K., Li, W., & Zhang, J. (2017). Dynamic changes of nutrient composition throughout the entire life cycle of black soldier fly. PLOS ONE, 12(8), e0182601. Loh, A. M. B., Njowe, K. Y., & Sop, M. K. (2017). Preliminary proximate composition and mineral contents of five edible insects from Cameroon. International Journal of Current Microbiology and Applied Sciences, 6(3), 1984–1995.

4  •  Nutrient Content and Functionalities of Edible Insects  81 Lonnerdal, B. (2000). Dietary factors influencing zinc absorption. The Journal of Nutrition, 130(5), 1378S–1383S. Makkar, H. P., Tran, G., Heuzé, V., & Ankers, P. (2014). State-of-the-art on use of insects as animal feed. Animal Feed Science and Technology, 197, 1–33. Madibela, O. R., Seitiso, T. K., Thema, T. F., & Letso, M. (2007). Effect of traditional processing methods on chemical composition and in vitro true dry matter digestibility of the Mophane worm (Imbrasia belina). Journal of Arid Environments, 68(3), 492–500. Manditsera, F. A., Luning, P. A., Fogliano, V., & Lakemond, C. M. (2019). Effect of domestic cooking methods on protein digestibility and mineral bioaccessibility of wild harvested adult edible insects. Food Research International, 121, 404–411. Mariod, A. A. (Ed.). (2020). African edible insects as alternative source of food, oil, protein and bioactive components. Springer Nature. McClements, D. J. (2004a). Food emulsions: Principles, practices, and techniques. CRC Press. McClements, D. J. (2004b). Protein-stabilized emulsions. Current Opinion in Colloid & Interface Science, 9(5), 305–313. https://doi​.org​/10​.1016​/j​.cocis​.2004​.09​.003 Megido, R. C., Poelaert, C., Ernens, M., Liotta, M., Blecker, C., Danthine, S., Bindelle, J. (2018). Effect of household cooking techniques on the microbiological load and the nutritional quality of mealworms (Tenebrio molitor L. 1758). Food Research International, 106, 503–508. Meyer-Rochow, V. B. (1975). Can insects help to ease the problem of world food shortage. Search, 6(7), 261–262. Mishyna, M., Martinez, J.-J. I., Chen, J., & Benjamin, O. (2019). Extraction, characterization and functional properties of soluble proteins from edible grasshopper (Schistocerca gregaria) and honey bee (Apis mellifera). Food Research International, 116, 697–706. https://doi​.org​/10​.1016​/j​.foodres​.2018​.08​.098 Mishyna, M., Martinez, J.-J. I., Chen, J., Davidovich-Pinhas, M., & Benjamin, O. (2019). Heat-induced aggregation and gelation of proteins from edible honey bee brood (Apis mellifera) as a function of temperature and pH. Food Hydrocolloids, 91, 117–126. https://doi​.org​/10​.1016​/j​.foodhyd​.2019​.01​.017 Mitsuhashi, J. (2010). The future use of insects as human food. Forest Insects as Food: Humans Bite Back, 115, 122. Mlcek, J., Rop, O., Borkovcova, M., & Bednarova, M. (2014). A comprehensive look at the possibilities of edible insects as food in Europe–a review. Polish Journal of Food and Nutrition Sciences, 64(3), 147–157. Moghadasian, M. H., & Shahidi, F. (2017). Fatty acids. In International encyclopedia of public health (second edition) (pp. 114–122). Academic Press. Mundi, S., & Aluko, R. (2012). Physicochemical and functional properties of kidney bean albumin and globulin protein fractions. Food Research International, 48(1), 299–306. Mwangi, M. N., Oonincx, D., Stouten, T., Veenenbos, M., Melse-Boonstra, A., Dicke, M., & van Loon, J. J. A. (2018). Insects as sources of iron and zinc in human nutrition. Nutrition Research Reviews, 31(2), 248–255. https://doi​ .org​/10​.1017​/s0954422418000094 Nadathur, S., Wanasundara, J. P., & Scanlin, L. (2016). Sustainable protein sources. Academic Press. Nicolai, T., Britten, M., & Schmitt, C. (2011). β-Lactoglobulin and WPI aggregates: Formation, structure and applications. Food Hydrocolloids, 25(8), 1945–1962. https://doi​.org​/10​.1016​/j​.foodhyd​.2011​.02​.006 Nongonierma, A. B., & FitzGerald, R. J. (2017). Unlocking the biological potential of proteins from edible insects through enzymatic hydrolysis: A review. Innovative Food Science and Emerging Technologies, 43, 239–252. https://doi​.org​/10​.1016​/j​.ifset​.2017​.08​.014 Nowak, V., Persijn, D., Rittenschober, D., & Charrondiere, U. R. (2016). Review of food composition data for edible insects. Food Chemistry, 193, 39–46. Nowakowski, A. C., Miller, A. C., Miller, M. E., Xiao, H., & Wu, X. (2022). Potential health benefits of edible insects. Critical Reviews in Food Science and Nutrition, 62(13), 3499–3508. Oibiokpa, F. I. (2017). Nutrient and antinutrient compositions of some edible insect species in Northern Nigeria. Fountain Journal of Natural and Applied Sciences, 6(1), 9–24. Oibiokpa, F. I., Akanya, H. O., Jigam, A. A., Saidu, A. N., & Egwim, E. C. (2018). Protein quality of four indigenous edible insect species in Nigeria. Food Science and Human Wellness, 7(2), 175–183. Okagu, O. D., Verma, O., McClements, D. J., & Udenigwe, C. C. (2020). Utilization of insect proteins to formulate nutraceutical delivery systems: Encapsulation and release of curcumin using mealworm protein-chitosan nano-complexes. International Journal of Biological Macromolecules, 151, 333–343. Onyeike, E. N., Ayalogu, E. O., & Okaraonye, C. C. (2005). Nutritive value of the larvae of raphia palm beetle (Oryctes rhinoceros) and weevil (Rhyncophorus pheonicis). Journal of the Science of Food and Agriculture, 85(11), 1822–1828. Orkusz, A. (2021). Edible insects versus meat—Nutritional comparison: Knowledge of their composition is the key to good health. Nutrients, 13(4), 1207. Paul, D., & Dey, S. (2014). Essential amino acids, lipid profile and fat-soluble vitamins of the edible silkworm Bombyx mori (Lepidoptera: Bombycidae). International Journal of Tropical Insect Science, 34(4), 239–247.

82  Edible Insects Processing for Food and Feed Payne, C. L. R., Scarborough, P., Rayner, M., & Nonaka, K. (2016a). Are edible insects more or less 'healthy' than commonly consumed meats? A comparison using two nutrient profiling models developed to combat over- and undernutrition. European Journal of Clinical Nutrition, 70(3), 285–291. https://doi​.org​/10​.1038​/ ejcn​.2015​.149 Payne, C. L. R., Scarborough, P., Rayner, M., & Nonaka, K. (2016b). A systematic review of nutrient composition data available for twelve commercially available edible insects, and comparison with reference values. Trends in Food Science and Technology, 47, 69–77. https://doi​.org​/10​.1016​/j​.tifs​.2015​.10​.012 Pearce, K., & Kinsella, J. (1978). Emulsifying properties of proteins: Evaluation of a turbidimetric technique. Journal of Agricultural and Food Chemistry, 26(3), 716–723. Pino Moreno, J., & Ramos-Elorduy, J. (2002). Pragmatic uses of cochineal (Homoptera Dactylopiidae). Paper presented at the Ethnobiology and biocultural diversity: Proceedings of the 7th International Congress of Ethnobiology, Athens, GA, October 2000. Post, M. (2018). Proteins in cultured beef. In Proteins in Food Processing (pp. 289–298). Elsevier. Purschke, B., Meinlschmidt, P., Horn, C., Rieder, O., & Jäger, H. (2018). Improvement of techno-functional properties of edible insect protein from migratory locust by enzymatic hydrolysis. European Food Research and Technology, 244, 999–1013. Raheem, D., Raposo, A., Oluwole, O. B., Nieuwland, M., Saraiva, A., & Carrascosa, C. (2019). Entomophagy: Nutritional, ecological, safety and legislation aspects. Food Research International, 126, 19. https://doi​.org​/10​ .1016​/j​.foodres​.2019​.108672 Rahnamaeian, M., Cytryńska, M., Zdybicka-Barabas, A., Dobslaff, K., Wiesner, J., Twyman, R. M., & SchmidHempel, P. (2015). Insect antimicrobial peptides show potentiating functional interactions against Gramnegative bacteria. Proceedings of the Royal Society B: Biological Sciences, 282(1806), 20150293. Rumpold, B. A., & Schlüter, O. (2015). Insect-based protein sources and their potential for human consumption: Nutritional composition and processing. Animal Frontiers, 5(2), 20–24. Rumpold, B. A., & Schlüter, O. K. (2013). Nutritional composition and safety aspects of edible insects. Molecular Nutrition & Food Research, 57(5), 802–823. Rumpold, B. A., & Schlüter, O. K. (2013). Potential and challenges of insects as an innovative source for food and feed production. Innovative Food Science & Emerging Technologies, 17, 1–11. https://doi​.org​/10​.1016​/j​.ifset​ .2012​.11​.005 Sales-Campos, H., Reis de Souza, P., Crema Peghini, B., Santana da Silva, J., & Ribeiro Cardoso, C. (2013). An overview of the modulatory effects of oleic acid in health and disease. Mini Reviews in Medicinal Chemistry, 13(2), 201–210. Siddiq, M., Nasir, M., Ravi, R., Dolan, K. D., & Butt, M. S. (2009). Effect of defatted maize germ addition on the functional and textural properties of wheat flour. International Journal of Food Properties, 12(4), 860–870. https://doi​.org​/10​.1080​/10942910802103028 Silverthorn, D. U. (2015). Human physiology. Jones & Bartlett Publishers. Singh, A., Vanga, S. K., Orsat, V., & Raghavan, V. (2018). Application of molecular dynamic simulation to study food proteins: A review. Critical Reviews in Food Science and Nutrition, 58(16), 2779–2789. https://doi​.org​/10​ .1080​/10408398​.2017​.1341864 Singh, H., & Ye, A. (2013). Structural and biochemical factors affecting the digestion of protein-stabilized emulsions. Current Opinion in Colloid & Interface Science, 18(4), 360–370. https://doi​.org​/10​.1016​/j​.cocis​.2013​.04​.006 Sogari, G., Liu, A., & Li, J. (2019). Understanding edible insects as food in western and eastern societies. In Environmental, health, and business opportunities in the new meat alternatives market (pp. 166–181). Sogari, G., Liu, A., & Li, J. (2019). Understanding edible insects as food in Western and Eastern societies. In Environmental, health, and business opportunities in the new meat alternatives market (pp. 166–181). IGI Global. Soloman, M., Ladeji, O., & Umoru, H. (2008). Nutritional evaluation of the giant grassropper (Zonocerus variegatus) protein and the possible effects of its high dietary fibre on amino acids and mineral bioavailability. African Journal of Food, Agriculture, Nutrition and Development, 8(2), 238–251. Song, Y. S., Frias, J., Martinez-Villaluenga, C., Vidal-Valdeverde, C., & de Mejia, E. G. (2008). Immunoreactivity reduction of soybean meal by fermentation, effect on amino acid composition and antigenicity of commercial soy products. Food Chemistry, 108(2), 571–581. https://doi​.org​/10​.1016​/j​.foodchem​.2007​.11​.013 Spranghers, T., Ottoboni, M., Klootwijk, C., Ovyn, A., Deboosere, S., De Meulenaer, B., & De Smet, S. (2017). Nutritional composition of black soldier fly (Hermetia illucens) prepupae reared on different organic waste substrates. Journal of the Science of Food and Agriculture, 97(8), 2594–2600. Spranghers, T., Michiels, J., Vrancx, J., Ovyn, A., Eeckhout, M., De Clercq, P., & De Smet, S. (2018). Gut antimicrobial effects and nutritional value of black soldier fly (Hermetia illucens L.) prepupae for weaned piglets. Animal Feed Science and Technology, 235, 33–42.

4  •  Nutrient Content and Functionalities of Edible Insects  83 Staljanssens, D., Van Camp, J., Herregods, G., Dhaenens, M., Deforce, D., Van de Voorde, J., & Smagghe, G. (2011). Antihypertensive effect of insect cells: In vitro and in vivo evaluation. Peptides, 32(3), 526–530. Stull, V. J., Wamulume, M., Mwalukanga, M. I., Banda, A., Bergmans, R. S., & Bell, M. M. (2018). “We like insects here”: Entomophagy and society in a Zambian village. Agriculture and Human Values, 35(4), 867–883. https:// doi​.org​/10​.1007​/s10460​- 018​-9878-0 Tang, C., Yang, D., Liao, H., Sun, H., Liu, C., Wei, L., & Li, F. (2019). Edible insects as a food source: A review. Food Production, Processing and Nutrition, 1(1), 1–13. Teffo, L. S., Toms, R., & Eloff, J. N. (2007). Preliminary data on the nutritional composition of the edible stink-bug, Encosternum delegorguei Spinola, consumed in Limpopo province, South Africa. South African Journal of Science, 103(11–12), 434–436. Thaiphanit, S., Schleining, G., & Anprung, P. (2016). Effects of coconut (Cocos nucifera L.) protein hydrolysates obtained from enzymatic hydrolysis on the stability and rheological properties of oil-in-water emulsions. Food Hydrocolloids, 60, 252–264. https://doi​.org​/10​.1016​/j​.foodhyd​.2016​.03​.035 Udomsil, N., Imsoonthornruksa, S., Gosalawit, C., & Ketudat-Cairns, M. (2019). Nutritional values and functional properties of house cricket (Acheta domesticus) and field cricket (Gryllus bimaculatus). Food Science and Technology Research, 25(4), 597–605. US Department of Agriculture (USDA). (2015). Database on nutrient content of food. http://www​.ndb​.nal​.usda​.gov (accessed August 23, 2021). US Department of Agriculture (USDA). (2015). Database on nutrient content of food. Retrieved June 9, from http:// www​.ndb​.nal​.usda​.gov Van Huis, A., Dicke, M., & van Loon, J. J. (2015). Insects to feed the world. Journal of Insects as Food and Feed, 1(1), 3–5. Vercruysse, L., Smagghe, G., Beckers, T., & Van Camp, J. (2009). Antioxidative and ACE inhibitory activities in enzymatic hydrolysates of the cotton leafworm, Spodoptera littoralis. Food Chemistry, 114(1), 38–43. Vercruysse, L., Smagghe, G., Herregods, G., & Van Camp, J. (2005). ACE inhibitory activity in enzymatic hydrolysates of insect protein. Journal of Agricultural and Food Chemistry, 53(13), 5207–5211. https://doi​.org​/10​ .1021​/jf050337q Verhoeckx, K. C., van Broekhoven, S., den Hartog-Jager, C. F., Gaspari, M., de Jong, G. A., Wichers, H. J., & Knulst, A. C. (2014). House dust mite (Der p 10) and crustacean allergic patients may react to food containing yellow mealworm proteins. Food and Chemical Toxicology, 65, 364–373. Villaseñor, V. M., Enriquez-Vara, J. N., Urías-Silva, J. E., & Mojica, L. (2021). Edible Insects: Techno-functional properties food and feed applications and biological potential. Food Reviews International, 1–27. https://doi​ .org​/10​.1080​/87559129​.2021​.1890116 Voelker, R. (2019). Can insects compete with beef, poultry as nutritional powerhouses? JAMA, 321(5), 439–441. Wang, R., Wang, T., Feng, W., Wang, Q., & Wang, T. (2021). Rice proteins and cod proteins forming shared microstructures with enhanced functional and nutritional properties. Food Chemistry, 354, 129520. https://doi​.org​ /10​.1016​/j​.foodchem​.2021​.129520 Wang, Z., Li, Y., Jiang, L., Qi, B., & Zhou, L. (2014). Relationship between secondary structure and surface hydrophobicity of soybean protein isolate subjected to heat treatment. Journal of Chemistry, 2014 Article ID 475389, 10 pages. Wierenga, P. A., van Norél, L., & Basheva, E. S. (2009). Reconsidering the importance of interfacial properties in foam stability. Colloids and Surfaces A: Physicochemical and Engineering Aspects, 344(1–3), 72–78. Wu, Q.-Y., Jia, J.-Q., Tan, G.-X., Xu, J.-L., & Gui, Z.-Z. (2011). Physicochemical properties of silkworm larvae protein isolate and gastrointestinal hydrolysate bioactivities. African Journal of Biotechnology, 10(32), 6145–6153. Yang, Y., Tang, H., Zhang, Y., Zhu, F., Lü, P., Yao, Q., & Chen, K. (2018). Research progress on the immune mechanism of the silkworm Bombyx mori. Physiological Entomology, 43(3), 159–168. Yi, H.-y., Chowdhury, M., Huang, Y.-d., & Yu, X.-q. (2014). Insect antimicrobial peptides and their applications. Applied Microbiology and Biotechnology, 98(13), 5807–5822. http://dx​.doi​.org​/10​.1007​/s00253​- 014​ -5792-6 Yi, L., Lakemond, C. M. M., Sagis, L. M. C., Eisner-Schadler, V., van Huis, A., & van Boekel, M. A. J. S. (2013). Extraction and characterisation of protein fractions from five insect species. Food Chemistry, 141(4), 3341– 3348. https://doi​.org​/10​.1016​/j​.foodchem​.2013​.05​.115 Yi, L., Van Boekel, M. A., Boeren, S., & Lakemond, C. M. (2016). Protein identification and in vitro digestion of fractions from Tenebrio molitor. European Food Research and Technology, 242(8), 1285–1297. Zayas, J. F. (2012). Functionality of proteins in food. Springer Science & Business Media. Zhang, C. X., Tang, X. D., & Cheng, J. A. (2008). The utilization and industrialization of insect resources in China. Entomological Research, 38, S38–S47.

84  Edible Insects Processing for Food and Feed Zhou, Z.-F., Ren, Z.-X., Yu, H.-Y., Jia, J.-Q., & Gui, Z.-Z. (2017). Effects of different modification techniques on molecular structure and bioactivity of Bombyx mori pupa protein. Journal of Asia-Pacific Entomology, 20(1), 35–41. Zielińska, E., Baraniak, B., & Karaś, M. (2017). Antioxidant and anti-inflammatory activities of hydrolysates and peptide fractions obtained by enzymatic hydrolysis of selected heat-treated edible insects. Nutrients, 9(9), 970. https://doi​.org​/10​.3390​/nu9090970 Zielińska, E., Baraniak, B., Karaś, M., Rybczyńska, K., & Jakubczyk, A. (2015). Selected species of edible insects as a source of nutrient composition. Food Research International, 77, 460–466. https://doi​.org​/10​.1016​/j​.foodres​ .2015​.09​.008 Zielińska, E., Karaś, M., & Baraniak, B. (2018). Comparison of functional properties of edible insects and protein preparations thereof. LWT, 91, 168–174. https://doi​.org​/10​.1016​/j​.lwt​.2018​.01​.058 Zielińska, E., Karaś, M., & Jakubczyk, A. (2017). Antioxidant activity of predigested protein obtained from a range of farmed edible insects. International Journal of Food Science & Technology, 52(2), 306–312.

Consumer Attitudes towards Insects as Food

5

Marina Carcea, Valentina Narducci, and Valeria Turfani Contents 5.1 Introduction 85 5.2 Traditions and Culture 86 5.3 Psychological Barriers and the Yuck Factor 88 5.4 Sensory and Palatability Aspects 90 5.5 Current Trends in the Use of Insects as Food 95 5.6 Conclusions 99 References 99

5.1 INTRODUCTION Many people around the world eat insects out of choice because of their palatability or because they have an established place in local food cultures (Van Huis et al., 2013). Other people, mostly in Western countries, consider them a nuisance, pests for crops and animals and disgusting to eat, and entomophagy is associated with primitive behaviour. Their nutritional value has been reviewed in Chapter 4, but we can say that, in general, they are interesting sources of proteins, lipids and micronutrients. The most consumed species worldwide belong to the beetle group (Coleoptera) (31%), followed by caterpillars (Lepidoptera) (18%); bees, wasps and ants (Hymenoptera) (14%); grasshoppers, locusts and crickets (Orthoptera) (13%); cicadas, leafhoppers, plant hoppers, scale insects and true bugs (Hemiptera) (10%); termites (Isoptera) (3%); dragonflies (Odonata) (3%); flies (Diptera) (2%); and other orders (5%) (Jongema, 2017). Most edible insects are harvested in the wild, and the concept of farming them for food is relatively new. In some countries there is a tradition of eating insects, whereas in others, insects or insect-containing products are just appearing on the market as a curiosity and people are reluctant to consume them, being culturally new. In this chapter, consumer attitudes towards insects as food will be reviewed, considering traditions and cultures as well as psychological barriers, sensory and palatability aspects and current trends in the use of insects as foods.

DOI: 10.1201/9781003165729-5

85

86  Edible Insects Processing for Food and Feed

5.2  TRADITIONS AND CULTURE Insects have been eaten by humans for tens of thousands of years, and they still retain an important place as traditional food in many parts of the world. Acceptance may lag in some Western nations, but it is widely estimated that insects are currently regularly consumed by about two billion people, around a quarter of the world’s population (Van Huis, 2013). More than 2,000 species of insects (2,140, as reported) are recognized as being consumed by more than 3,000 ethnic groups in up to 130 countries, concentrated mostly in sub-Saharan Africa, Central and South America, or in the Southeast Asia and the Pacific. The countries with the highest consumption of insects as food include the Democratic Republic of Congo, the Republic of Congo, the Central African Republic, Cameroon, Uganda, Zambia, Zimbabwe, Nigeria, and South Africa. Nevertheless, there are 11 countries in Europe which are entomophagous (Ramos-Elorduy, 2009; Gahukar, 2020). Human beings have a long history of interactions with insects stretching back many thousands of years. Archaeological studies report that 10,000 years ago, hunters and gatherers ate bugs to survive (Sutton, 1995; De Foliart, 1997; Van Itterbeeck, 2012; Van Huis, 2017). In Europe, the first references to entomophagy come from Ancient Greece. Aristoteles left proof of the practice of eating cicadas in his Historia Animalium (384–322 BC), where he says that cicadas are most delicious when eaten as a chrysalis, whereas as adults, males taste better but females taste better after mating when they are full of eggs. Many other documents show how usual it was to eat insects in those times: Diodorus Siculus (200 BC) called people from Ethiopia “Acridophagi” because of their diet based on grasshoppers and locusts (family Acrididae). The Roman author Pliny the Elder mentions in his Naturalis Historia (Pliny the Elder, 1969) that Roman aristocrats loved to eat beetle larvae reared on flour and wine. In the book of Leviticus from the Bible (11:22), the Jews were said to be permitted to eat “the locust of any kind, the bald locust of any kind, the cricket of any kind, and the grasshopper of any kind”. In the Gospel after Matthew (3:4), St. John the Baptist is said to have survived on locusts and honey when he lived in the desert. In Asia, Chinese literature usually refers to entomophagy and the use of insects in traditional medicine. Despite the fact that insects have always been an essential element in the human diet since the beginning of time and they continue to be eaten in different countries around the world, at a certain point they started to be abandoned in modern Western societies, especially in Europe and in the United States. The most probable reason for that is linked with the origin of agriculture and livestock. The Fertile Crescent, an historical region containing western territories of Asia, the Nile Valley and the Nile Delta, is considered the birthplace of agriculture and, secondarily, of livestock (Western Neolithic Revolution). From this moment on, agriculture and animal husbandry started to spread towards Europe and later to America, so that eventually they replaced hunting and gathering of resources as the main food sources. As a consequence, the consumption of insects was replaced by the consumption of meat, especially from larger herbivorous and omnivorous animals, which, in addition, offered a wider variety of products: fur and leather, feathers, lactic products, traction power and a new means of transport. Thus, agriculture and livestock husbandry became common practices all over Europe as they related to more stable food sources. Animal hunting and the consumption of insects are both very dependent on seasonality to obtain food, so they shifted to the background and started to be considered primitive practices (Ramos-Elorduy, 2009). As the industrial revolution commenced in England in the eighteenth century, agriculture became even more efficient and yields increased even more: the industrial sector grew as a result of this, and people moved to urban areas to find a job. All over Europe, urbanisation was a huge trend and, even today, it is ongoing. The growing impact of agriculture on people’s lives also led to an increased fear of a poor harvest due to pests, plagues and bad weather. Insects increasingly were seen as a huge threat to food provision and health, and therefore the idea of them would trigger a lot of negative feelings (Satterthwaite et al., 2010).

5  •  Consumer Attitudes towards Insects as Food  87 Latin American food culture has a rich history of eating insects which dates back centuries. Bernardino de Sahagún (1557) describes in his work Historia de las cosas de Nueva España how the eggs of water bugs were considered delicacies at the court of ancient Aztec kings. Nowadays, this insect caviar is still being eaten and is referred to as escamoles (Parsons, 2010). Still today, Latin American food culture includes the consumption of a large variety of insects. In Latin America, approximately 50 million people (about 10% of the total population) are indigenous peoples. Insects contribute significantly to the food security of the majority of the autochthonous societies in the northwest Amazon (5%–7% of total protein intake during the year and 12%–26% during certain months of the year) (Costa-Neto, 2016). Similarly, between rural communities in remote areas of Ecuador and Colombia and in the bushy areas of the High Andes, entomophagy is still a substantial part of the diet for most tribes. In Colombia, leafcutter ants are seen as a delicacy. They are known as hormigas culonas (big-bottomed ants), and their gastronomic and economic value is comparable to that of caviar and truffle mushrooms (Milton, 1984; Onore, 1997; Van Huis et al., 2013). In Mexico, 535 types of edible insects are presently consumed in regions situated in the centre, south, and southeast of the country as part of the traditional Mexican food culture. During the Jumil Festival, people harvest edible stinkbugs in the woods to consume them raw or alternatively roasted, fried or even ground with tomato, onion and chili to fill tacos. Locusts, known as chapulines, are harvested and eaten during the rainy season (Van Huis et al., 2013). The chapulines are often toasted or processed into a chili and served in a tortilla. Especially in urban areas, they are a popular street food, sold in many street stands and small restaurants, and they are increasingly found on the menus of expensive restaurants (Cohen et al., 2009). The practice of entomophagy has been commonplace in many Asian countries from 2000 BC on, and presently 349 insect species are eaten in 29 Asian countries (Ramos-Elorduy, 2006). In Vietnam, Laos and Thailand, insects are the main source of meat for the poor (Yhoung-aree J., 2010). Mainly large insects are preferred, such as grasshoppers, butterfly larvae and large bugs, but people also consume ants, wasps, bees and dragonflies. The largest Asian consumer of insects is Thailand, which alone has 194 edible insect species, with a very lively and ancient gastronomic tradition both in cities and in rural areas. 164 insect species are readily available at markets or from several pushcarts in Bangkok including grasshoppers, bamboo worms, silk larvae, crickets, weevils, ant queens and scorpions. Many bars in Thailand serve fried bugs alongside their libations. Crickets, grasshoppers and a variety of worms are all considered favourite snacks in this part of the world (Sirimungkararat et al., 2010; Raheem et al., 2019). Edible insects have historically been consumed in China for more than 2,000 years owing to their medicinal and trophic value (Feng et al., 2018). Insects are still used as health food in China, an example being caterpillar fungus, which is believed to enhance immunity and have anti-cancer properties (Wei et al., 2021). Currently, 324 species are documented that are either edible or associated with entomophagy in China (Feng et al., 2018). Japan has a long-lasting traditional culture of eating wild insects. There are at least 117 insect species traditionally used as food in this country (Mitsuhashi, 1997). Nowadays, insects are particularly appreciated by old people, but they are above all consumed with the family during end-ofyear festivities and given as gift to relatives or friends who have come from the city to the countryside, or sold as snacks in ordinary supermarkets and sometimes restaurants (Payne et al., 2019). In Africa, 300 million people habitually and traditionally consume insects. Edible insects play a very important role as a natural resource that is used particularly in months of food shortage. Approximately more than 524 species of edible insects are consumed, which satisfy more than 50% of the protein intake requirement (Hlongwane et al., 2020). The most consumed species are butterflies, followed by beetles, locusts and hymenoptera (bees, wasps, ants), termites and fewer flies. The largest consumers are found in Central Africa, in particular in the Democratic Republic of Congo, with 300 g caterpillars consumed per capita per week (Kelemu et al., 2015). In Europe, and among non-aboriginal Americans, Canadians, Australians and New Zealanders, entomophagy remains rare or even a taboo (Shelomi, 2015).

88  Edible Insects Processing for Food and Feed With the increase of agricultural practices, as explained before, it is likely that people in Europe and Western countries have begun to regard insects with contempt as many of them are harmful for agriculture and for human health, amongst other reasons. This, together with economic and social development, has led to an aversion to edible insects, with entomophagy increasingly seen as a “primitive peoples” practice (Costa-Neto, 2015). However, in the last decade, in some European countries, consumers have begun to show an interest in new food products that use insects as ingredients in an unrecognizable form.

5.3 PSYCHOLOGICAL BARRIERS AND THE YUCK FACTOR Despite early references to entomophagy in the Old Continent, it was not until after the end of the nineteenth century that Europeans, and more largely Westerners, became acquainted with the term entomophagy, which refers to the practice of eating insects and was officially coined by the British entomologist, Holt (Shelomi, 2015). Regardless of his efforts, edible insects were never eaten extensively by his fellow countrymen, highlighting the particularity of entomophagy in the Western world. It has to be noted here that insects were never truly a part of Western food cultures. Consuming a certain insect species is a markedly local issue, determined by the species, its life cycle and availability as well as the cultural environment in which the given species is consumed, being accepted by the society to a high or to a low degree (Dobermann et al., 2017; Payne et al., 2019). For these reasons the negative attitude towards eating insects has increased over time, and insects are virtually synonymous with nuisance (mosquitoes, flies, termites). Some insects may accidentally end up in meals: the U.S. Food and Drug Administration reports that there may be 60 fragments of insect in 100 g of chocolate, for example, and certain insects are also transmitters of disease (Van Huis et al., 2013; Vanhonacker et al., 2013; Looy et al., 2014; Deroy et al., 2015). These could be some reasons why European society refuses to accept edible insects as food (Yen, 2009). Unusual novel foods like insects are generally not considered to be appropriate as food in cultures where they are not commonly eaten, and are often rejected for reasons other than their intrinsic sensory properties. This is also related to the fact that it has been forbidden to sell insects as food in the EU for the last 20 years and until very recently (EC, 1997). The purpose of the legislation is to protect consumers from unknown hazards such as allergies, poisons and infections. However, some countries in the EU, such as the Netherlands, Belgium, France, and Denmark, have interpreted the law in a less strict manner, enabling the selling as well as rearing of insects to a certain extent (Berg et al., 2017). These social, cultural and legislative barriers, together with food neophobia – the fear of new, novel and unfamiliar foods – helped to increase the “yuck” factor. Disgust is one of the seven universal emotions and arises as a feeling of aversion towards something offensive (Looy et al., 2014; La Barbera et al., 2018). Food neophobia in relation to insects can be reduced through familiarity, as seen in studies where acceptance became higher when insects were presented in familiar carriers such as muffins, cookies, chocolate and chips (Caparros-Megido et al., 2014; Tan et al., 2016; Gurdian et al., 2022). Food neophobia and disgust are recognized as being at the origin of rejection of entomophagy, disgust being the most important factor. Neophobia and disgust are not the same thing, although they are often confused: in fact, a familiar food can arouse disgust and a new food does not necessarily arouse disgust. Both neophobia and disgust are evolutionarily linked to avoiding health risks due to eating something potentially harmful, and at the same time they are linked to a culture-dependent learning mechanism (Pliner and Hobden, 1992; De Foliart, 1999; Hartman and Siegrist, 2018; La Barbera et al., 2018; Wendin and Nyberg, 2021). Food neophobia is defined as the reluctance to eat and/or avoidance of novel foods. A learning process leading from the unknown to the familiar is required to overcome this reluctance, but what is ordinarily familiar or unknown is linked to the reference culture. A scale to measure food neophobia was proposed by Pliner and Hobden in 1992 and it is still in use. Disgust is recognized as a primary emotion (one that is universally found in all human beings). Psychology has discovered that disgust (with consequent

5  •  Consumer Attitudes towards Insects as Food  89 behaviours) can be mediated by the existence of unconscious, culturally learned associations of thought. Coming to food disgust, Hartman and Siegrist (2018) found that it can be aroused by different elements and developed a scale with sub-scales to measure it in all its components. Now, the practice of entomophagy arouses disgust in many Western people, but at least two billion people in other parts of the world adopt it. La Barbera et al. (2018) studied how Westerners’ disgust towards the use of insects as food occurs through the existence of culturally learned, unconscious associations of thought between insects and more directly disgusting objects such as dirt, matter in decomposition, faeces and microbes. The important conclusion is that not only a familiarization process is needed to accept insects as novel food, but such unconscious associations also need to be overcome. This implies that attention should be paid in designing first experiences of contact with novel insect food in such a way that they don’t produce disgust, which would reinforce the rejection. The information/familiarization process can only be slow and gradual, and it may take even an entire generation. The process will have to include the introduction to the market of foods containing insects together with a range of communication and food education actions. It is well documented that first tastings have a not negligible probability of developing disgust if the insect is offered in a recognizable form, i.e., directly as such (whole) or as an ingredient in large pieces identifiable as insect parts. It is more effective to propose familiar and appreciated foods (such as biscuits, snack bars, and meatballs) containing finely ground insects or insect flour as an ingredient. In this way, the familiar food is recognized as such at first sight and the sensory characteristics of the new ingredient (flavour, aroma, texture) accompany those, known and appreciated, of the traditional ingredients (La Barbera et al., 2018). Therefore, in order to promote entomophagy, the disgust factor must be addressed. Disgust can be monitored through the Food Disgust Scale (FDS), a self-report measure that enables the assessment of an individual’s emotional disposition to react with disgust to certain food-related (offensive) stimuli (Hartmann and Siegrist, 2018). Negative attitudes towards invertebrates are a deep-seated, visceral response among Western peoples, they arise mainly out of cultural/social construction and low exposure to insects (Looy and Wood, 2006). Positive sensory experiences play a necessary role in the process of learning to accept a food, but they are inadequate when unusual and culturally inappropriate foods are involved. As curious tasting does not imply acceptance, curiosity can be a key driver to tasting insects for the first time (Verneau et al., 2016; Sogari et al., 2018). This trigger, however, does not entail that consumers will be willing to eat insects repeatedly if the product is not considered to be tasty. The availability, or at least the interest, of Western consumers towards the introduction of insects as food in the diet is very low, but slowly increasing. Verbeke found in 2015 that only 3% of a Flemish consumer sample indicated they were “definitely willing or ready” to adopt insects as a foodstuff, while another 16% claimed to be “willing or ready”. This makes 1 out of 5 people, indicating an interest in entomophagy which could allow the establishment of a small starting market (Verbeke, 2015). Two years before, another study had found only a 5% availability to try insects as foods in a sample of consumers with similar socio-demographic characteristics and from the same study region (Vanhonacker et al., 2013). Verbeke found that readiness to accept entomophagy in the Flemish consumers from his study was higher in males than in females, in younger than in older people, in people not having a strong attitude towards meat, in people who are open to trying novel food and in people who are interested in the environmental impact of their food choices (Verbeke, 2015). Similar features were found in other studies on consumers of different Western countries (Collins et al., 2019; Tuccillo et al., 2020; Herbert and Beacon, 2021; Moruzzo et al., 2021). Making Western consumers familiar with and accepting of insects as food is a slow and gradual process that will need the introduction to the market of novel foods containing insects and at the same time the implementation of a wide range of communication and food education actions (La Barbera et al., 2018). To this end, some authors have explored and proposed possible approaches and precautions (La Barbera et al., 2018; Collins et al., 2019; Wendin and Nyberg, 2021). Providing more information can improve the likelihood that people will want to try entomophagy. People have poor information about the nutritional benefits of insect food, its sustainability or the range

90  Edible Insects Processing for Food and Feed of possible products and their taste. However, knowledge of nutritional and sustainability benefits is not sufficient (La Barbera et al., 2018; Collins et al., 2019; Tuccillo et al., 2020; Moruzzo et al., 2021; Wendin and Nyberg, 2021). The presence of visible insect parts in foods is a strongly negative attribute (La Barbera et al., 2018; Collins et al., 2019; Mishyna et al., 2020; Moruzzo et al., 2021; Wendin and Nyberg, 2021). Insect food processed to resemble familiar products is more palatable and acceptable to many people. As the market develops and more people are familiarized, passing through positive experiences, products with whole or recognizable insects may become more accepted. For now, the choice experiments indicate that manufacturers may prefer to consider products where the nature of the food is not apparent. Consumer expectations influence the evaluation of the textural properties of unfamiliar food. Mishyna et al. (2020) report that people who expected to experience the crispy texture of mealworms evaluated this texture positively, whereas those who expected a meat-like texture evaluated mealworms negatively. However, Sogari et al. (2018) report that the neophobic attitude towards cricket jelly decreased if individuals’ perceptions of sensory properties changed positively after tasting.

5.4  SENSORY AND PALATABILITY ASPECTS Edible insects possess a wide range of organoleptic properties (taste, flavour, texture, colour). These mainly depend on insect species and development stage (egg, larva, pupa, male or female adult). In addition, the feed that insects fed on, or were fed if reared, also influences taste and flavour. Finally, the cooking method can make a difference in the taste, flavour, colour and texture, not to mention the addition of seasonings (Ramos-Elorduy, 1998; Van Huis et al., 2013; Kouřimská et al., 2016; Elhassan et al., 2019; Mishyna et al., 2020). The taste and flavour of edible insects is generally described as nutty, lemony, shrimp-like and fatty, this last due to their commonly high lipid content (Ramos-Elorduy, 1998). However, very different taste, aroma and flavour notes are reported from both consumers and sensory analyses, going from sweet to savoury, from mild to spicy or sour, from fruits and vegetables to fish and meat; for a detailed list of sensory properties of edible insects according to the available literature see Table 5.1. The flavour and aroma can be linked to substances produced by the species at certain stages of their life, such as pheromones in bees (which, however, are washed away by blanching in boiling water) and formic acid in ants, which is responsible for sourness. The high lipid content of many edible insects makes them susceptible to rancidity if not properly stored (Elhassan et al., 2019). Texture ranges from crispy to soft, passing through various grades such as chewy, plump and juicy. It is generally crisp for insects having a hard exoskeleton, and soft in soft-skinned ones (Mishyna et al., 2020). However, some insect parts, like the legs and wings of grasshoppers, locusts and crickets, can be negatively perceived in the mouth (Sogari et al., 2018), and it is recommended that they be removed before consumption due to the risk of intestinal constipation (Van Huis et al., 2013). Insect taste and texture can be modified by the processing method used. Slow wet heating of honeybee brood leads to gentle coagulation of the proteins, which yields a plump and soft, yet solid texture, together with a herbal, vegetable and nutty flavour (but the texture becomes hard and grainy if they are steamed or boiled quickly), whereas dry heating (roasting or frying) of honeybee brood produces a crispy texture and a meaty flavour (Bruun et al., 2016). Raw water boatmen have a fishy flavour that becomes more shrimp-like when they are dried (Ramos-Elorduy, 1998). Cricket broth made with crickets which were alive before cooking was found to have higher acidity (pH) and higher perception of saltiness and umami than broth made with crickets which were frozen before cooking (Farina, 2017). Sensory properties of edible insects are also summarized in Table 5.1. In the Western world, sensory tests conducted with panels of experts and with consumers have shown that traditional products (hamburgers, pasta, baked goods, bars) modified with the addition of a percentage

5  •  Consumer Attitudes towards Insects as Food  91 TABLE 5.1  Sensory properties of edible insects TYPE OF INSECT

DEVELOPMENT STAGE

PROCESSING

TASTE, AROMA AND FLAVOUR Sweet Sweet Ants

Adults

-

Sweet

Termites

Adults

-

Sweet Savoury Savoury

Leaf footed bugs

Sour Sour

Sour Bitter Umami Umami

Umami, broth Umami, broth Umami, broth

Mild, neutral (takes on flavour of condiments) Mild Mild Nutty Nutty

REFERENCE

-

Kouřimská and Adámková, 2016; Ramos-Elorduy, 1998 Kouřimská and Adámková, 2016; Ramos-Elorduy, 1998 Ramos-Elorduy, 1998

Termites Macrotermes bellicosus, Macrotermes subhyalinus, Nasutitermes triodiae

-

.

Elhassan et al., 2019

Red wood ant Formica rufa Black ant Polyrhachis vicina Jet ant Lasius fuliginosus Ants Black ant Polyrhachis vicina

-

-

Elhassan et al., 2019; Ramos-Elorduy, 1998

Adults -

-

Mishyna et al., 2020 Elhassan et al., 2019

Mealworm Tenebrio molitor, Alphitobius diaperinus Cricket Acheta domesticus Grasshopper Locusta migratoria Termites Macrotermes bellicosus, Macrotermes subhyalinus, Nasutitermes triodiae

-

-

Elhassan et al., 2019

Adults

-

Elhassan et al., 2019

Adults

-

Elhassan et al., 2019

-

-

Elhassan et al., 2019

Cricket Acheta domesticus Grasshopper Locusta migratoria

Adults

-

Adults

-

Elhassan et al., 2019; Ramos-Elorduy, 1998 Elhassan et al., 2019

Ants

Adults

-

Kouřimská and Adámková, 2016; Ramos-Elorduy, 1998

92  Edible Insects Processing for Food and Feed TABLE 5.1 (CONTINUED)  Sensory properties of edible insects TYPE OF INSECT

DEVELOPMENT STAGE

Nutty, hazelnut, Brazil nut

Honey bee broods

Larvae and pupae

Nutty

Mealworm Tenebrio molitor, Alphitobius diaperinus Termites

Nutty

Nutty

PROCESSING

REFERENCE Bruun Jensen et al., 2016

Larvae

Raw or slow wet heat (steamed) -

Adults

Dry heat

Termites Macrotermes nigeriensis

Adults

Termites Macrotermes bellicosus, Macrotermes subhyalinus, Nasutitermes triodiae Wasps

-

Mild fried or roasted in their oil .

Kouřimská and Adámková, 2016; Ramos-Elorduy, 1998 Igwe et al., 2011; Ramos-Elorduy, 1998

-

-

Cricket Acheta domesticus Grasshopper Locusta migratoria

Adults

-

Kouřimská and Adámková, 2016; Ramos-Elorduy, 1998 Elhassan et al., 2019

Adults

-

Elhassan et al., 2019

Larvae of darkling beatles (mealworms)

Larvae

-

Raw corn

Caterpillar of smoky wainscots

Larvae

-

Corn

Cricket Acheta domesticus Grasshopper Locusta migratoria

Adults

-

Kouřimská and Adámková, 2016; Ramos-Elorduy, 1998 Kouřimská and Adámková, 2016; Ramos-Elorduy, 1998 Elhassan et al., 2019

Adults

-

Elhassan et al., 2019

Treehoppers

-

-

Ramos-Elorduy, 1998

Leaf footed bugs Nopal worms Mealybugs

-

-

-

-

Mushrooms

Cockroaches

-

-

Vegetables

Grasshopper Locusta migratoria

Adults

-

Ramos-Elorduy, 1998 Ramos-Elorduy, 1998 Kouřimská and Adámková, 2016; Ramos-Elorduy, 1998 Kouřimská and Adámková, 2016; Ramos-Elorduy, 1998 Elhassan et al., 2019

Nutty

Pine seeds

Nuttiness Nuttiness Cereals Wholemeal bread

Cereal Vegetables From avocado to fried zucchini Pumpkin Fried potatoes Fried potatoes

Elhassan et al., 2019; Mishyna et al., 2020

Elhassan et al., 2019

(Continued)

5  •  Consumer Attitudes towards Insects as Food  93 TABLE 5.1 (CONTINUED)  Sensory properties of edible insects TYPE OF INSECT Legumes Kidney beans Fruit Apple

DEVELOPMENT STAGE

PROCESSING

REFERENCE

Agave worms

Larvae

-

Ramos-Elorduy, 1998

Striped shield bugs (stinkbugs)

-

-

Red wood ant Formica rufa Jet ant Lasius fuliginosus Ants Grasshopper Locusta migratoria Giant water bug Lethocerus indicus

-

-

Kouřimská and Adámková, 2016; Ramos-Elorduy, 1998 Elhassan et al., 2019

Adults Adults

-

Elhassan et al., 2019 Mishyna et al., 2020 Elhassan et al., 2019

Adult

Salted boiled

Kiatbenjakul et al., 2014, Kiatbenjakul et al., 2015

Dragonfly larvae and other aquatic insects

Larvae

-

Caviar

Eggs of water boatman and backswimmers

Eggs

-

Fish Shrimps Herring

Water boatmen Water boatmen Caterpillars of erebid moths

Adults Adults Larvae

Fresh Dried -

Shrimp

Grasshopper Locusta migratoria

Adults

-

Kouřimská and Adámková, 2016; Ramos-Elorduy, 1998 Kouřimská and Adámková, 2016; Ramos-Elorduy, 1998 Ramos-Elorduy, 1998 Ramos-Elorduy, 1998 Kouřimská and Adámková, 2016; Ramos-Elorduy, 1998 Elhassan et al., 2019

Agave worms Tree worms Larvae of wooddestroying beatles Cricket Acheta domesticus Honey bee broods

Larvae Larvae Larvae

-

Adults

-

Ramos-Elorduy, 1998 Ramos-Elorduy, 1998 Kouřimská and Adámková, 2016 Elhassan et al., 2019

Agave worms

Larvae and pupae Larvae

Dry heat (fried or roasted) -

Bruun Jensen et al., 2016 Ramos-Elorduy, 1998

Ants

-

-

Mishyna et al., 2020

Lemon Lime Lemon, lime Fruit Pineapple, fruity, green apple, banana, and floral aroma Fish Fish

Meat (roasted pork rinds, meaty, chicken) Cracklings Pork rinds Fatty brisket with skin Chicken Meaty, chickeny, bacon, liver Spicy Charred Charred lemon

(Continued)

94  Edible Insects Processing for Food and Feed TABLE 5.1 (CONTINUED)  Sensory properties of edible insects TYPE OF INSECT Fatty, buttery, milky Fatty

Fatty

Buttery, milky

Maillard reaction products Maillard reaction products Maillard reaction products Maillard reaction products, toasty Unique flavour

TEXTURE Crumbly Crusty, hard and coarse Crusty, coarse, grainy and hard Crunchy

Crispy and crunchy Plump and soft but solid; chewy if harvested at advanced pupa stage

DEVELOPMENT STAGE

PROCESSING

REFERENCE

Larvae of wood destroying beatles

Larvae

-

Termites Macrotermes bellicosus, Macrotermes subhyalinus, Nasutitermes triodiae Honey bee broods

Adults

Dry heat

Larvae and pupae

Lyophilised and powdered

Haber et al., 2019

Cricket Acheta domesticus Grasshopper Locusta migratoria Honey bee broods

Adults

-

Elhassan et al., 2019

Adults

-

Elhassan et al., 2019

Larvae, pupae

Bruun Jensen et al., 2016

Giant water bug Lethocerus indicus

Adult

Raw or slow wet heat (steamed) Frozen fresh or salted boiled

Cricket Acheta domesticus Grasshopper Locusta migratoria Black ant Polyrhachis vicina Termites Macrotermes bellicosus, Macrotermes subhyalinus, Nasutitermes triodiae Honey bee broods

Adults

-

Elhassan et al., 2019

Adults

-

Elhassan et al., 2019

Honey bee broods

Larvae, pupae

-

Kouřimská and Adámková, 2016; Ramos-Elorduy, 1998 Elhassan et al., 2019; Ramos-Elorduy, 1998

Kiatbenjakul et al., 2014, Kiatbenjakul et al., 2015

Elhassan et al., 2019

Adults

Dry heat

Elhassan et al., 2019

Larvae, pupae

Dry heat (fried or roasted) Raw or slow wet heat (steamed)

Bruun Jensen et al., 2016 Bruun Jensen et al., 2016

5  •  Consumer Attitudes towards Insects as Food  95 of insect flour often obtain high scores in comparison not only with plant-based variants, but also with the original products. The percentage of insect meal incorporated has an effect on the results of the tests, but the percentage of insect meal must not forcedly remain low; it depends on the product. In hamburgers, for example, it can reach 30%, while baked products with 5% insect flour are more appreciated than those containing 20% (Mishyna et al., 2020).

5.5 CURRENT TRENDS IN THE USE OF INSECTS AS FOOD Though the science of edible insects is still in its infancy, nevertheless the number of research studies and of publications on the use of insects as food has greatly increased in the last 10 years based on the assumption that insects can provide food at low environmental cost and can be easy to rear. In countries where eating insects is part of their culinary traditions, such as in many Asian, African and South American countries, insects are eaten whole, grilled, stir fried, roasted, blanched or cooked in soups or stews, as snacks or main meals (Bomolo et al., 2017, 2019; Ebenebe et al., 2017; Ghosh et al., 2017; Lautenschläger et al., 2017a, 2017b; Okia et al., 2017; Turpin and Si, 2017; Cruz-Labana et al., 2018; Feng et al., 2018; Manno et al., 2018; Bbosa et al., 2019; Hurd et al., 2019; Roberts, 2019; Van Itterbeeck, 2019; Gahukar, 2020). They are also dried (via sun, oven, freeze, fluidized bed, spray and microwave drying) and ground into powder, sometimes mixed with salt and spices to flavour dishes or used as food colorant (cochineal) (Melgar-Lalanne et al., 2019). In Europe very few cases of traditional and still surviving use of insects as food are documented: insect larvae (cheese fly, Piophila casei) are let grow into other foods such as cheese to give it a special taste and texture, like a traditional cheese in the Sardinia (casu martzu, rotten cheese) and Corsica islands (see http://www​.sardegnaagricoltura​.it​/documenti​/14​_43​_20070607153029​.pdf ), or insects are let soak in alcoholic beverages to give them a distinctive flavour, such as in the case of homemade ant-flavoured aquavit in Sweden (Svanberg and Berggren, 2019). In recent years, in countries without a tradition of insect consumption and where processed foods are more consumed by the population, several food industries have started experimenting with insects, and over the years hundreds of products have appeared on the market (see also bugbu​​rger.​​se​/gu​​ide​/t​​he​-bi​​g​-lis​​ t​-of-​​edibl​​e​-ins​​​ect​-p​​roduc​​ts/),​mainly offered on the Internet. Some brands have started with producing products which are characterized by looking like common, well known and broadly accepted foods such as bread, muffins, cookies, pasta or granola bars, where a small amount of insect powder is part of the ingredients. Consumers do not see the insects so it is not difficult to accept this kind of product, and the texture and taste might not be greatly influenced, although the product composition might be affected. Grain-based products have been preferred for the incorporation of insect powders and seem a promising field due to the wide product range and their popularity. As far as pasta products are concerned, Biró et al. (2019) developed a silkworm-enriched buckwheat pasta with up to 10 g silkworm for 100 g of pasta, which was enough to enrich pasta from a nutritional and a sensorial point of view. Duda et al. (2019) studied the quality and nutritional and textural properties of durum wheat pasta enriched with cricket powder at 5%, 10% and 15% levels. Ҫabuk and Yilmaz (2020) fortified the traditional Turkish egg pasta (eriște) with mealworm and grasshopper flour and legume flours. A mixture of millet and cricket flour at levels of 5% and 10% was also used by Jakab et al. (2020) to increase the nutritional quality of dry wheat pasta. Bread and bakery products are also a very popular vehicle for the introduction of insects. Alemu et al. (2017) found that bun products with medium amounts (5%) of cricket flour were preferred to no or high amounts (10%) of cricket flour. De Oliveira et al. (2017) produced a nutritionally enriched wheat bread with cinereus cockroach and showed that bread with a 10% insect flour addition differed little from white and whole wheat bread. Haber et al. (2019) investigated the use of grasshopper powder in bread baking and concluded that it is possible to include up to 100 g/kg of this powder without alterations

96  Edible Insects Processing for Food and Feed in sensorial aspects. Cricket powder was also added to wheat flour at levels of 10% or 30% by Osimani et al (2018) to produce bread which underwent technological, microbiological and nutritional evaluation and they concluded that bread enriched with 10% cricket powder showed a discrete global liking by untrained panelists. Cricket flour–containing buns were produced and sensory evaluated in a study by Pambo et al. (2018). Galli et al. (2020) studied lactic acid bacteria isolated from cricket powder’s spontaneous fermentation as potential starters for cricket-wheat bread production. Cappelli et al. (2020) studied the rheological properties and characteristics of bread manufactured with cricket and mealworm powder at 5%, 10% and 15% levels of substitution. Powdered mealworm and buffalo worm larvae were used by Gaglio et al. (2021) to functionalize sourdough Italian-style (ciabatta) bread. Montevecchi et al. (2021) evaluated the technological and nutritional parameters of bread fortified with black soldier fly prepupae at 0.2% and 0.4% levels. Protein fortified bread was produced by Roncolini et al. (2019) by including mealworm powder into bread doughs at 5% and 10% substitution levels, and the textural, microbiological, nutritional and sensory features of bread were studied, whereas Gonzales et al. (2018) investigated the performance of a 5% addition of cricket, black soldier flies and mealworms in bakery products. Igual et al. (2021) evaluated amino acid and protein bio-accessibility in bread enriched with lesser mealworm and mealworm. Khatun et al. (2021) investigated the impact of supplementing wheat flour with cricket flour or paste at different levels (5%–15%) on the rheological and textural properties of baked chapatti. The high protein content in cricket powder prompted Da Rosa Machado et al. (2019) to successfully use it in the manufacturing of gluten-free breads instead of other protein sources of vegetable origin. Also, Kowalczewski et al. (2019, 2021a, 2021b) studied the use of cricket powder in gluten-free bread manufacturing. Other popular grain-based foods are bars, extruded snacks and crackers. Ribeiro et al. (2019) employed defatted freeze-dried edible crickets in cereal and dry fruit bars and evaluated their nutritive value and sensory profile: their results suggest that the characteristic flavour of crickets seem to be associated with their lipid content. Chocolate bars with insect flour seem to be appreciated by consumers (Cicatiello et al., 2020) as well as energy bars of two different compositions with the addition of cricket flour and puff-pastry bars sprinkled with whole roasted mealworm larvae (Adámek et al., 2020). Caporizzi et al. (2018) reported the use of 3D printing as a particularly suitable technique in the production of novel foods with cereal and insect flour. Severini et al. (2018) employed 3D printing technology to obtain snacks with a designed geometry from wheat flour dough enriched with ground larvae of yellow mealworms as a novel source of protein, and the research group of Azzollini et al. (2018) produced extruded cereals made of wheat flour and ground yellow mealworm larvae at 10% and 20% levels and studied the effects of formulation and process conditions on their microstructure, texture and digestibility. Akullo et al. (2018) manufactured crackers enriched with dried and ground crickets and soldier and winged termites and found that crackers enriched with 5% winged termite had good nutrient and sensory qualities. Similarly, Ardoin et al. (2021) formulated and evaluated snack crackers with increasing levels of cricket powder (5%, 10%, 15% and 20%) in substitution of wholewheat flour. Igual et al. (2020) evaluated the effects of enrichment with different quantities of house cricket powder on the quality of extruded corn snacks. Insects and pea powder have also been used by García-Segovia et al. (2020) as alternative protein and mineral sources in extruded snacks, as well as by Ribeiro et al. (2021), who formulated extruded snacks with corn flour and house cricket powder. Wendin et al. (2021) added mealworm in different amounts to crisps. Tao et al. (2017) investigated the feasibility of incorporating edible insect flours (cricket or locust) at 10% or 15% addition levels in an extruded rice product and concluded that those levels of addition give good results from technological and sensorial points of view. Other bakery products where wheat flour has been partially substituted with insect powder are cookies, biscuits and muffins: Castro Delgado et al. (2020) manufactured chocolate chip cookies with 15% and 30% levels of cricket flour and assessed consumer acceptability in three countries, whereas Awobusuyi et al. (2020) studied the consumer acceptance of biscuits supplemented with a sorgum-insect meal at 20%–40%–60% levels and Lucchese-Cheung et al. (2021) examined the sensorial perception of cookies made with mealworm flour. Cookies enriched with silkworm pupae were developed by Torres et al.

5  •  Consumer Attitudes towards Insects as Food  97 (2022). Burt et al. (2020) also studied the acceptance of using cricket flour and Zielińska et al. (2021) of cricket and mealworm flour as a substitute for wheat flour in muffins, whereas Smarzyński et al. (2021) studied shortcake biscuits made with cricket flour. Roncolini et al. (2020) experimented with mealworm powder as a novel baking ingredient for manufacturing high-protein, mineral-dense rusks. Akande et al. (2020) assessed mulberry silkworm pupae and African palm weevil larvae as alternative protein sources in snack filling. Hernández Toxqui et al. (2021) used yellow warm larvae flour to develop functional ice cream, whereas David-Birman et al. (2022) used silkworm pupae powder. The production of meat alternatives or meat analogue products is on the rise following consumers’ interest in flexitarianism, vegetarianism and veganism (Kim et al., 2022). Moreover, the possibility of using insects as extenders in meat products is considered an opportunity to develop more sustainable products (Karnjanapratum et al., 2022; Pintado and Delgado-Pando, 2020). Kim et al. (2017) investigated the effect of the addition of house cricket flour addition on the physicochemical and textural properties of meat emulsion under various formulations and concluded that cricket flour possesses the necessary physical properties to be used as an alternative non-meat ingredient within emulsified meat products. Scholliers et al. (2020) experimented with insect and pork proteins in hybrid meat products, whereas Kim et al. (2020) studied the effect of interaction between mealworm protein and myofibrillar protein (pork ham) in meat emulsion systems. Stoops et al. (2017) studied the microbial dynamics during the production and storage of minced meat-like products from two species of mealworm larvae. Smetana et al. (2018) applied twin screw high-moisture extrusion to a mixture of protein concentrates (insect with concentration of 15%–50% dry matter and soy) and water to produce fibrous meat analogues with texture and protein composition like meat. Smetana et al. (2019) used high-moisture extrusion with insect and soy protein concentrates and found that the inclusion of 15%–40% of insects could imitate the texture of meat. The addition of cricket powder to pork pâté at 2% was shown by Smarzyński et al. (2019) to give a product of high attractiveness for consumers. Insects have also been experimented with in the production of beverages. Tello et al. (2021) aimed to use mealworm larvae to develop an alternative to bovine milk with a lower environmental impact and similar nutritional profile. Torres-Acosta et al. (2021) supplemented two types of mezcal (a distilled alcoholic drink from Mexico) with two different insect species (mealworm and Central American locust) and provided evidence of antioxidant potential from insects that remain after storage. Insects have also been used in the manufacturing of improved formulations of complementary foods (Agbemafle et al., 2020; Aboge et al., 2021; Adepoju and Ajayi, 2021; Mekuria et al., 2021), and microencapsulation of insect powder has also been experimented to improve appearance and other organoleptic characteristics (Sánchez et al., 2021). Akande et al. (2022) studied the possibility of substituting powdered milk with migratory locust powder as a cheap and sustainable protein source in peanut-based ready-to-use therapeutic foods used in the treatment of malnutrition. Besides flour, another intermediate product that can be readily incorporated in products by the food industry is paste, and to this end a mealworm paste free of testable exoskeleton particles was successfully manufactured and studied by De Smet et al. (2019), whereas a whole new range of products can be obtained by fermentation. Castro Lopez et al. (2020) provided an overview of the available literature on fermentation applied to obtain new insect-based products because evidence has suggested that alternative technologies, in particular fermentation, could be used to obtain diverse insect-based ingredients/products with unique properties, as confirmed also by Lee et al. (2021), who studied fermented mealworm as a novel prebiotic. Fermented edible insects are also reviewed by Kewuyemi et al. (2020) and indicated as promoters of food security in Africa. The next step in the evolution of insect-containing foods are products which start to show some parts of insects but in which they are not yet presented whole. This implies that insects are more acceptable to consumers; examples can be insects eggs or mixed insect spices. The last step in the evolution of insect-containing foods are products to be developed in the long term, which involve a radical change in insect consumption: insects are shown in the food, they are the

98  Edible Insects Processing for Food and Feed main ingredients and consumers look for them and are happy to eat them. Examples can be insect burgers, frozen insects with vegetables, insect balls and new fermented foods made from insects or insect larvae (Hwang et al., 2019). As already mentioned, 3D printing technologies have been advocated to render ingredients such as insects more acceptable and appealing to consumers (Lupton and Turner, 2018). More recently, components such as chitin, proteins, fats and oils have been extracted from insects and used in foods (Cheseto et al., 2020; Delicato et al., 2020). As regards insect proteins, in addition to their nutritional value, they present technological-functional properties that can be used to create specific applications in innovative food products or insect-based dietary supplements (Fasolin et al., 2019). The black soldier fly has been identified as a promising source for the sustainable production of proteins, lipids and bioactive substances (Müller et al., 2017; Caligiani et al., 2018). Miron et al. (2019) evaluated the possibility of extracting water-soluble proteins from black soldier fly larvae, and Mintah et al. (2019) studied the functionality and antioxidative capacity of proteins and hydrolysates extracted from the same source. Soetemans et al. (2019) studied the use of organic acids to improve the fractionation of the larvae juice into lipid and protein enriched fractions. Yu et al. (2017) prepared protein hydrolysates from mealworm powder by enzymatic hydrolysis using five different proteases, and the hydrolysates were then tested for their antioxidant activities, whereas Yi et al. (2017) studied the effect of pH and NaCl on the extraction yield of water-soluble proteins from yellow mealworm. Ndiritu et al. (2019) studied the effect of NaCl and pH on the functional properties of edible cricket protein concentrate whereas Hall et al. (2017) studied the functional properties of tropical banded cricket protein hydrolysates. Hall et al. (2018) studied the effect of enzymatic hydrolysis on the bioactive properties and allergenicity of cricket protein. Purschke et al. (2018) investigated the recovery and characterization of soluble proteins from migratory locust and also the enzymatic hydrolysis of migratory locust protein flour as a method to improve the techno-functional properties of this flour. Clarkson et al. (2018) studied the potential of extracted migratory locust protein fractions as value-added ingredients. Chatsuwan et al. (2018) characterized and studied the functionality and antioxidant activity of water-soluble proteins extracted from silkworm pupae. Mishyna et al. (2019) extracted, characterized and studied the functional properties of soluble proteins from edible grasshopper and honeybee. Kim et al. (2020) studied the improvement of the functional properties of insect protein solutions by use of transglutaminase as a cross-linking agent, and Mishyna et al. (2021) recently reviewed the techno-functional properties of edible insect proteins and the effects of processing. Recently, Luna et al. (2021) produced cricket protein hydrolysates by means of enzymatic hydrolysis and used them to formulate corn tortillas and tortilla chips with 20% hydrolysates content. Insects are also a novel source of lipids for a range of applications. Oils are the main co-product of insect protein extraction and they present very interesting characteristics depending on the origin and feed source of the producing insect; in several cases, the level of unsaturated fatty acids is high (Berezina, 2017; Dreassi et al., 2017). Depending on the processing approach, a defatting step can also be crucial for protein extraction (Jeong et al., 2021). Purschke et al. (2017) contributed to the development of fractionation processes for the production of standardised insect material for incorporation into food and feed (similarly to fat and protein rich raw material of vegetable origin, e.g. soy) by experimenting with the supercritical CO2 extraction of edible insect oil from yellow mealworm larvae. Sipponen et al. (2018) presented a dry fractionation technology for upgrading house crickets and yellow mealworm larvae by extraction with supercritical carbon dioxide to obtain oil, followed by separation into fine and coarse fraction by air classification, each possessing a different texture and flavour. Matthäus et al. (2019) studied the exploitation, properties and refining of fat obtained by cold pressing of black soldier fly larvae, whereas Tzompa-Sosa et al. (2019) studied four insect oils extracted by means of an aqueous-based oil extraction method and concluded that mealworm oils and cricket oil have characteristics desirable for table oils and as food ingredients. Smetana et al. (2020) studied the processing, sustainability and design of insect margarine derived from insect biomass of mealworm and black soldier fly. Amarender et al. (2020) analysed optimum extraction methodologies for lipid extraction and isolation of proteins from whole spray-dried cricket powder. Insect oil deodorization can help in disguising its taste in food products (Tzompa-Sosa et al., 2021, 2022).

5  •  Consumer Attitudes towards Insects as Food  99

5.6 CONCLUSIONS Insects can be consumed as such or mixed with other ingredients in the formulation of popular foods such as bakery products, pasta, snacks, beverages, etc. The idea in this latter case is that insects can contribute to improving the chemical composition of known products, but at the same time it is easy to camouflage their presence or their taste. In fact, studies suggest that consumers’ acceptance of edible insects can be enhanced by processing and blending them with familiar food products; selling them in attractively designed packaging; and sensibilizing and educating consumers about insect eating in connection with nutritional and sustainability considerations, but also by changing the current legislation. Legal rules on the use of insects as food vary across the world. In the EU, the United States and Canada, insects are novel and legally treated as such. For example, the EU has authorized the commercialization of the mealworm, of the migratory locust and only very recently of the house cricket. The exploitation of insects in Western countries is still in its infancy but an improving legislation as well as scientific and technological advancements can support the growing insect industry. Interest in the adoption of insect food as an alternative protein source has increased, although in Western societies, consumers still react with disgust. Insects also have the potential to produce high quality food ingredients (lipids, proteins and their hydrolysates, chitin and other health-promoting substances). The influence of processing conditions on the technological and nutritional quality of insect raw materials and the evolution of product quality during the storage of foods with insects as an ingredient are also new areas of study. There is also a need to establish tight protocols to validate product identity in the developing insect food market, and accurate labelling of content identity is important to gain consumers’ confidence. A challenge encountered in the use of insects as food is also the assessment of their safety, and consideration should be made of microbiological contamination, toxicological hazards such as chemical hazards and antinutrients and allergenicity issues related to different exposures, including injection, ingestion, inhalation and skin contact (see Chapter 11).

REFERENCES Aboge, D.O.; Orinda, M.A.; Konyole, S.O. 2021. Effect of partial substitution of soybean flour with cricket flour on the nutritional composition, in vitro-protein digestibility and functional properties of complementary porridge flour. International Journal of Tropical Insect Science, 42: 1137–1145. Adámek, M.; Adámková, A.; Kouřimská, L. et al. 2020. Sensory evaluation and acceptance of food made of edible insects. Potravinarstvo Solovak Journal of Food Sciences, 14: 921–928. Adepoju, O.T.; Ajayi, K. 2021. Consumption pattern and acceptability of winged termites (Macroterme bellicosus)enriched infant complementary foods in Ekiti State, Nigeria. International Journal of Tropical Insect Science, 41(3): 2039–2050. Agbemafle, I.; Hadzi, D.; Amagloh, F.K.; Zotor, F.B.; Reddy, M.B. 2020. Nutritional, microbial and sensory evaluation of complementary foods made from blends of orange-fleshed sweet potato and edible insects. Foods, 9(9): art. no. 1225. Akande, O.A.; Falade, O.O.; Badejo, A.A.; Adekoya, I. 2020. Assessment of mulberry silk worm pupae and African palm weevil larvae as alternative protein sources in snack fillings. Heliyon, 6(4): art no. e03754. Akande, O.A.; Oluwamukomi, M.; Osundahunsi, O.F.; Ijarotimi, O.S.; Mukisa, I.M. 2022. Evaluating the potential of utilising migratory locust powder (Locusta migratoria) as an alternative protein source in peanut-based ready-to-use therapeutic foods. Food Science and Technology International, January 2022. Akullo, J.; Nakimbugwe, D.; Obaa, B.B.; Okwee-Acai, J.; Agea, J.G. 2018. Development and quality evaluation of crackers enriched with edible insects. International Food Research Journal, 25(4): 1592–1599.

100  Edible Insects Processing for Food and Feed Alemu, M.H.; Olsen, S.B.; Vedel, S.E.; Kinyuru, J.N.; Pambo, K.O. 2017. Can insects increase food security in developing countries? An analysis of Kenian consumer preferences and demand for cricket flour buns. Food Security, 9(3): 471–484. Amarender, V.R.; Kanika, B.; Aaron, D.; Sanjeewa, G. 2020. Lipid and protein extraction from edible insects – Crickets (Gryllidae). LWT-Food Science and Technology, 125: art. no. 109222. Ardoin, R.; Marx, B.D.; Boeneke, C.; Prinyawiwatkul, W. 2021. Effects of cricket powder on selected physical properties and US consumer perceptions of whole-wheat snack crackers. International Journal of Food Science and Technology, 56(8): 4070–4080. Awobusuyi, T.D.; Pillay, K.; Siwela, M. 2020. Consumer acceptance of biscuits supplemented with a sorghum-insect meal. Nutrients, 12(4): art. no. 895. Azzollini, D.; Derossi, A.; Fogliano, V.; Lakemond, C.M.M.; Severini, C. 2018. Effects of formulation and process conditions on microstructure, texture and digestibility of extruded insect-riched snacks. Innovative Food Science and Emerging Technologies, 45: 344–353. Bbosa, T.; Tamale Ndagire, C.; Muzira Mukisa, I.; Fiaboe, K.K.M.; Nakimbugwe, D. 2019. Nutritional characteristics of selected insects in Uganda for use as alternative protein sources in food and feed. Journal of Insect Science, 19(6): 23. Berezina, N. 2017. Insects: Novel source of lipids for a fan of applications. OCL-Oilseeds and Fats, Crops and Lipids, 24(4): art. no. D402. Berg, J.; Wendin, K.; Langton, M.; Josell, A.; Davidsson, F. 2017. State of the art report: Insects as food and feed. Annals of Experimental Biology, 5(2): 37–46. Biró, B.; Fodor, R.; Szedljak, I.; Pásztor-Huszár, K.; Gere, A. 2019. Buckwheat-pasta enriched with silkworm powder: Technological analysis and sensory evaluation. LWT – Food Science and Technology, 116: art. no. 108542. Bomolo, O.; Niassy, S.; Chocha, A. et al. 2017. Ecological diversity of edible insects and their potential contribution to household food security in Haut-Katanga province, Democratic Republic of Congo. African Journal of Ecology, 55(4): 640–653. Bomolo, O.; Niassy, S.; Tanga, C.M. et al. 2019. The value chain of the edible caterpillar Elaphrodes lactea Gaede (Lepidoptera: Notodontidae) in the Miombo forest of the Democratic Republic of the Congo. Journal of Ethnobiology and Ethnomedicine, 15(1): art. no. 39. Bruun, J.A.; Evans, J.; Jonas-Levi, A.; Benjamin, O.; Martinez, I.; Dahle, B.; Roosh, N.; Lecocq, A.; Foley, K. 2016. Standard methods for Apis mellifera brood as human food. Journal of Apicultural Research, 56, 1–28. Burt, K.G.; Kotao, T.; Lopez, I.; Koeppel, J.; Goldstein, A.; Samuel, L.; Stopler, M. 2020. Acceptance of using cricket flour as a low carbohydrate, high protein, sustainable substitute for all-purpose flour in muffins. Journal of Culinary Science and Technology, 18(3): 201–213. Ҫabuk, B.; Yilmaz, B. 2020. Fortification of traditional egg pasta (eriște) with edible insects: Nutritional quality, cooking properties and sensory characteristics evaluation. Journal of Food Science and Technology, 57(7) 2750–2757. Caligiani, A.; Marseglia, A.; Leni, G. et al. 2018. Composition of black soldier fly prepupae and systematic approaches for extraction and fractionation of proteins, lipids and chitin. Food Research International, 105: 812–820. Caparros-Megido, R.; Sablon, L.; Geuens, M. et al. 2014. Edible insects acceptance by Belgian consumers: Promising attitude for entomophagy development. Journal of Sensory Studies, 29(1), 14–20. Caporizzi, R.; Derossi, A.; Severini, C. 2018. Cereal-based and insect-enriched printable food: From formulation to postprocessing treatments. Status and perspectives. In Fundamentals of 3D Food Printing and Applications, eds. Godoi, F.; Bhandari. B.; Prakash, S.; Zhang, M., 93–116. Academic Press, London, United Kingdom. Cappelli, A.; Oliva, N.; Bonaccorsi, G.; Lorini, C.; Cini, E. 2020. assessment of the rheological properties and bread characteristics obtained by innovative protein sources (Cicer arietinum, Acheta domesticus, Tenebrio molitor): Novel foods or potential improvers for wheat flour? LWT – Food Science and Technology, 118: art. no. 108867. Castro Delgado, M.; Chambers, E.; Carbonell-Barrachina, A.; Noguera Artiaga, L.; Vidal Quintanar, R.; Burgos Hernandez, A. 2020. Consumer acceptability in the USA, Mexico and Spain of chocolate chip cookies made with partial insect powder replacement. Journal of Food Science, 85(6): 1621–1628. Castro López, C.; Santiago-López, L.; Vallejo-Cordoba, B. et al. 2020. An insight to fermented edible insects: A global perspective and prospective. Food Research International, 137: art. no. 109750. Chatsuwan, N.; Puechkamut, Y.; Pinsirodom, P. 2018. Characterization, functionality and antioxidant activity of water-soluble proteins extracted from Bombix mori L. Current Applied Science and Technology, 18(2): 83–96. Cheseto, X.; Baleba, S.B.S.; Tanga, C.M.; Kelemu, S.; Torto, B. 2020. Chemistry and sensory characterization of a bakery product prepared with oils from African edible insects. Foods, 9(6): art. no. 800. Cicatiello, C.; Vitali, A.; Lacetera, N. 2020. How does it taste? Appreciation of insect-based snacks and its determinants. International Journal of Gastronomy and Food Science, 21: art. no. 100211. Clarkson, C.; Mirosa, M.; Birch, J. 2018. Potential of extracted Locusta migratoria protein fractions as value-added ingredients. Insects, 9(1): art​.n​o. 20.

5  •  Consumer Attitudes towards Insects as Food  101 Cohen, J.H.; Sánchez, N.D.M.; Montiel-ishinoet, F.D. 2009. Chapulines and food choices in rural Oaxaca. Gastronomica: The Journal of Food and Culture, 9(1): 61–65. Collins, C.M.; Vaskou, P.; Kountouris, Y. 2019. Insect food products in the Western world: Assessing the potential of a new ‘Green’ market. Annals of the Entomological Society of America, 112(6), 518–528. Costa-Neto, E.M. 2015. Anthropo-entomophagy in Latin America: An overview of the importance of edible insects to local communities. Journal of Insects as Food and Feed, 1(1): 17–23. Costa-Neto, E.M. 2016. Edible insects in Latin America: Old challenges, new opportunities. Journal of Insects as Food and Feed, 2(1): 1–2. Cruz-Labana, J.D.; Crosby-Gálvan, M.M.; Delgado-Alvarado, A.; Alcántara-Carbajal, J.L.; Cuca-García, J.M.; Tarango-Arámbula, L.A. 2018. Nutritional content of Liometopum apiculatum Mayr larvae (“escamoles”) by vegetation type in north-central Mexico. Journal of Asia-Pacific Entomology, 21(4): 1239–1245. Da Rosa Machado, C.; Thys, R.C.S. 2019. Cricket powder (Grillus assimilis) as a new alternative protein source for gluten-free breads. Innovative Food Science and Emerging Technologies, 56: art. no. 102180. David-Birman, T.; Romano, A.; Aga, A.; Pascoviche, D.; Davidovich-Pinhas, M.; Lesmes, U. 2022. Impat of silkworm pupae (Bombix mori) powder on cream foaming, ice cream properties and palatability. Innovative Food Science and Emerging Technologies, 75: art. no. 102874. Delicato, C.; Schouteten, J.J.; Dewettinck, K.; Gellynck, X.; Tzompa-Sosa, D.A. 2020. Consumers’ perception of bakery products with insect fat as partial butter replacement. Food Quality and Preference, 79: art. no. 103755. De Foliart, G.R. 1997. An overview of the role of edible insects in preserving biodiversity. Ecology of Food and Nutrition, 36(2–4): 109–132. https://​doi​.org​/10​.1080​/03670244​.1997​.9991510 De Foliart, G.R. 1999. Insects as food: Why the Western attitude is important. Annual Review of Entomology, 44: 21–50. https://doi​.org​/10​.1146​/annurev​.ento​.44​.1​.21 De Oliveira, L.M.; Da Silva Lucas, A.J.; Cadaval, C.L.; Mellado, M.S. 2017. Bread enriched with flour from cinereous cockroach (Nauphoeta cinerea). Innovative Food Science and Emerging Technologies, 44: 30–35. De Smet, J.; Lenaerts, S.; Borremans, A.; Scholliers, J.; Van Der Borght, M.; Van Campenhout, L. 2019. Stability assessment and laboratory scale fermentation of pastes produced on a pilot scale from mealworms (Tenebrio molitor). LWT – Food Science and Technology, 102: 113–121. Deroy, O.; Reade, B.; Spence, C. 2015. The insectivore’s dilemma, and how to take the west out of it. Food Quality and Preferences, 44: 44–55. Dobermann, D.; Swift, J.A.; Field, L.M. 2017. Opportunities and hurdles of edible insects for food and feed. Nutrition Bulletin, 42: 293–308. Dreassi, E.; Cito, A.; Zanfini, A.; Materozzi, L.; Botta, M.; Francardi, V. 2017. Dietary fatty acids influence the growth and fatty acid composition of the yellow mealworm Tenebrio molitor (Coleoptera: Tenebrionidae). Lipids, 52(3): 285–294. Duda, A.; Adamczak, J.; Chelminska, P.; Juszkiewicz, J.; Kowalczewski, P. 2019. Quality and nutritional /textural properties of durum wheat pasta enriched with cricket powder. Foods, 8(2): art. no. 46. Ebenebe, C.I.; Amobi, M.I.; Udegbala, C.; Ufele, A.N.; Nweze, B.O. 2017. Survey of edible insect consumption in South-Eastern Nigeria. Journal of Insects as Food and Feed, 3(4): 241–252. EC No 258/97 Regulation of the European Parliament and of the Council of 27 January 1997 concerning novel foods and novel food ingredients. Elhassan, M.; Wendin, K.; Olsson, V.; Langton, M. 2019. Quality aspects of insects as food—Nutritional, sensory, and related concepts. Foods, 8: 95. Farina, M.F. 2017. How method of killing crickets impact the sensory qualities and physiochemical properties when prepared in a broth. International Journal of Gastronomy and Food Science, 8: 19–23. Fasolin, L.H.; Pereira, R.N.; Pinheiro, A.C. et al. 2019. Emergent food proteins-Towards sustainability, health and innovation. Food Research International, 125: art. no. 108586. Feng, Y.; Chen, X.-M.; Zhao, M.; He, Z.; Sun, L.; Wang, C. -Y.; Ding, W. -F. 2018. Edible insects in China: Utilization and prospects. Insect Science, 25(2): 184–198. Gaglio, R.; Barbera, M.; Tesoriere, L. et al. 2021. Sourdough ciabatta bread enriched with powdered insects: Physicochemical, microbiological and simulated intestinal digesta functional properties. Innovative Food Science and Emerging Technologies, 72: art. no. 102755. Gahukar, R.T. 2020. Edible insects collected from forests for family livelihood and wellness of rural communities: A review. Global Food Security, 25: 100348. Galli, V.; Venturi, M.; Pini, N.; Granchi, L. 2020. Technological feature assessment of lactic acid bacteria isolated from cricket powder’s spontaneous fermentation as potential starters for cricket-wheat bread production. Foods, 9(9): art. no. 1322. García-Segovia, P.; Igual, M.; Noguerol, A.T.; Martínez-Monzó, J. 2020. Use of insects and pea powder as alternative protein and mineral sources in extruded snacks. European Food Research and Technology, 246(4): 703–712.

102  Edible Insects Processing for Food and Feed Ghosh, S.; Lee, S.-M., Jung, C.; Meyer-Rochow, V.B. 2017. Nutritional composition of five commercial edible insects in South Korea. Journal of Asia-Pacific Entomology, 20(2): 686–694. Gonzalez, M.C.; Garzon, R.; Rosell, C. 2018. Insects as ingredients for bakery goods. A comparison study of H. illucens, A. domestica and T. molitor flours. Innovative Food Science & Emerging Technologies, 51: 205–210. Gurdian, C.E.; Torrico, D.D.; Li, B.; Prinyawiwatkul, W. 2022. Effects of tasting and ingredient information statement on acceptability, elicited emotions, and willingness to purchase: A case of pita chips containing edible cricket protein. Foods, 11: 337. Haber, M.; Mishyna, M.; Martinez, J.J.I.; Benjamin, O. 2019. The influence of grasshopper (Schistocerca gregaria) powder enrichment on bread nutritional and sensorial properties. LWT – Food Science and Technology, 115: art. no. 108395. Hall, F.G.; Jones, O.G.; O’Haire, M.E.; Liceaga, A.M. 2017. Functional properties of tropical banded cricket (Gryllodes sigillatus) protein hydrolysates. Food Chemistry, 224: 414–422. Hall, F.; Johnson, P.E.; Liceaga, A. 2018. Effect of enzymatic hydrolysis on bioactive properties and allergenicity of cricket (Gryllodes sigillatus) protein. Food Chemistry, 262: 39–47. Hartmann, C.; Siegrist, M. 2018. Development and validation of the food disgust scale. Food Quality and Preference, 63: 38–50. Herbert, M.; Beacom, E. 2021. Exploring consumer acceptance of insect-based snack products in Ireland. Journal of Food Products Marketing, 27(6): 267–290. Hernández Toxqui, A.G.; Ramírez Ramírez, J.; Pino Moreno, J.M.; Talamantes Gómez, J.M.; Angeles Campos, S.C., Ramírez Orejel, J.C. 2021. Development of nutraceuticals ice creams using flour yellow worm larvae (Tenebrio molitor), chia (Salvia hispanica) and quinoa (Chenopodium quinoa). Frontiers in Veterinary Science, 8: art. no. 629180. Hlongwane, Z.T.; Slotow, R.; Munyai, T.C. 2020. Nutritional composition of edible insects consumed in Africa: A systematic review. Nutrients, 12: 2786. https://doi​.org​/10​.3390​/nu12092786 Hurd, K.J.; Shertukde, S.; Toia, T. 2019. The cultural importance of edible insects in Oaxaca, Mexico. Annals of the Entomological Society of America, 112(6): 552–559. Hwang, D.; Lee, W.S.; Goo, T. -W., Park, S. -W.; Yun, E. –Y. 2019. Effect of pretreatment with paste and sauce extract made using Tenebrio molitor larvae on ethanol damaged HepG2 cells. Entomological Research, 49(12): 509–518. Igual, M.; García-Segovia, P.; Martínez-Monzó, J. 2020. Effect of Acheta domesticus (house cricket) addition on protein content, colour, texture, and extrusion parameters of extruded products. Journal of Food Engineering, 282: art. no. 110032. Igual, M.; García-Segovia, P.; Martínez-Monzó, J. 2021. Amino acids and protein in vitro bio-accessibility from edible insect and pea protein enriched bread. Journal of Insects as Food and Feed, 7(6): 1001–1009. Igwe, C.; Ujowundu, C.; Nwaogu,L.; Okwu, G.N. 2011. Chemical analysis of an edible African termite, Macrotermes nigerianensis; a potential antidote to food security problem. Biochemistry and Analytical Biochemistry, 1: 1. Jakab, I.; Tormási, J.; Dhaygude, V.; Mednyánszky, Z.S.; Sipos, L.; Szedljak, I. 2020. Cricket flour-laden millet flour blends’ physical and chemical composition and adaptation in dried pasta products. Acta Alimentaria, 49(1): 4–12. Jeong, M.-S.; Lee, S.-D.; Cho, S.-J. 2021. Effect of three defatting solvents on the techno-functional properties of an edible insect (Grillus bimaculatus) protein concentrate. Molecules, 26(17), art. no. 5307. Jongema, Y. 2017. List of edible insects of the world. https://www​.wur​.nl​/en ​/ Research​-Results​/Chair​-groups​/ Plant​ -Sciences​/ Laboratory​-of​-Entomology​/ Edible​-insects​/ Worldwide​-species​-list​.htm (accessed on April 14, 2022). Karnjanapratum, S.; Kaewthong, P.; Indriani, S.; Petsong, K.; Takeungwongtrakul, S. 2022. Characteristics and nutritional value of silkworm (Bombix mori) pupae-fortified chicken bread spread. Scientific Reports 12(1): art​.n​o. 1492. Kelemu, S.; Niassy, S.; Torto, B. et al. 2015. African edible insects for food and feed: Inventory, diversity, commonalities and contribution to food security. Journal of Insects as Food and Feed, 1: 103–119. Kewuyemi, Y.O.; Kesa, H.; Chinma, C.E.; Adebo, O.A. 2020. Fermented edible insects for promoting food security in Africa. Insects, 11(5): art. no. 283. Khatun, H.; Van Der Borght, M.; Akhtaruzzaman, M; Claes, J. 2021. Rheological characterization of chapatti (Roti) enriched with flour or paste of house crickets (Acheta domesticus). Foods, 10(11): art​.n​o. 2750. Kiatbenjakul, P.; Intarapichet, K.-O.; Cadwallader, K.R. 2014. Identification of potent sulfur-containing odorants in scent glands of edible male giant water bug, Lethocerus indicus (Lep. and Serv.). Flavour and Fragrance Journal, 29(2): 107–113. Kiatbenjakul, P.; Intarapichet, K.-O.; Cadwallader, K.R. 2015. Characterization of potent odorants in male giant water bug (Lethocerus indicus Lep. and Serv.) an important edible insect of Southeast Asia. Food Chemistry, 168: 639–647.

5  •  Consumer Attitudes towards Insects as Food  103 Kim, H.-W.; Setyabrata, D.; Lee, Y.; Jones, O.G.; Kim, Y.H.B. 2017. Effect of house cricket (Acheta domesticus) flour addition on physicochemical and textural properties of meat emulsion under various formulations. Journal of Food Science, 82(12): 2787–2793. Kim, T.-K.; Yong, H.I.; Jang, H.W.; Kim, Y.-B.; Choi, Y.-S. 2020. Functional properties of extracted protein from edible insect larvae and their interaction with transglutaminase. Foods, 9(5): art. no. 591. Kim, T.-K.; Lee, M.H.; Yong, H.I. 2020. Effect of interaction between mealworm protein and myofibrillar protein on the rheological properties and thermal stability of the prepared emulsion systems. Foods, 9(10): art. no. A4. Kim, T.-K.; Yong, H.I.; Cha, J.Y.; Park, S.-Y., Jung, S.; Choi, Y.-S. 2022. Drying-induced restructured jerky analog developed using a combination of edible insect protein and textured vegetable protein. Food Chemistry, 373: art. no. 131519. Kouřimská, L.; Adámková, A. 2016. Nutritional and sensory quality of edible insects. NFS Journal, 4: 22–26. Kowalczewski, P.Ł.; Walkowiak, K.; Masewicz, Ł. et al. 2019. Gluten-free bread with cricket powder-mechanical properties and molecular water dynamics in dough and ready product. Foods, 8(7): art. no. 240. Kowalczewski, P.Ł.; Gumienna, M.; Rybicka, I. et al. 2021a. Nutritional value and biological activity of gluten-free bread enriched with cricket powder. Molecules, 26(4): art. no. 1184. Kowalczewski, P.Ł.; Walkowiak, K.; Masewicz, Ł. et al. 2021b. LF NMR spectroscopy analysis of water dynamics and texture of gluten-free bread with cricket powder during storage. Food Science and Technology International, 27(8): 776–785. La Barbera, F.; Verneau, F.; Amato, M.; Grunert, K. 2018. Understanding Westerners’ disgust for the eating of insects: The role of food neophobia and implicit associations. Food Quality and Preference, 64: 120–125. Lautenschläger, T.; Neinhuis, C.; Kikongo, E.; Henle, T.; Förster, A. 2017a. Impact of different preparations on the nutritional value of the edible caterpillar Imbrasia epimethea from Northern Angola. European Food Research and Technology, 243(5): 769–778. Lautenschläger, T.; Neinhuis, C.; Monizi, M. et al. 2017b. Edible insects of Northern Angola. African Invertebrates, 58(2): 55–82. Lee, S.J.; Kwon, G.T.; Chung, Y.H. et al. 2021. Study of fermented mealworm (Tenebrio molitor L.) as a novel prebiotic for intestinal microbiota. Journal of the Korean Society of Food Science and Nutrition, 50(6): 543–550. Looy, H.; Wood, J.R. 2006. Attitudes toward invertebrates: Are educational “bug banquets” effective? The Journal of Environmental Education, 37: 37–48. Looy, H.; Dunkel, F.V.; Wood, J.R. 2014. How then shall we eat? Insect-eating attitudes and sustainable foodways. Agriculture and Human Values, 31: 131–141. Lucchese-Cheung, T.; Kluwe de Aguiar, L.A.; Spers, E.E.; De Lima, L.M. 2021. The Brazilians’ sensorial perceptions for novel food-cookies with insect protein. Journal of Insects as Food and Feed, 7(3): 287–299. Luna, G.C.; Martin-Gonzalez, F.S.; Mauer, L.J.; Liceaga, A.M. 2021. Cricket (Acheta domesticus) hydrolysates’ impact on the physicochemical, structural and sensory properties of tortillas and tortilla chips. Journal of Insects as Food and Feed, 7(1): 109–120. Lupton, D.; Turner, B. 2018. Food for the future? Consumer responses to the idea of 3D-printed meat and insect based foods. Food and Foodways, 26(4): 269–289. Manno, N.; Estraver, W.Z.; Tafur, C.M. et al. 2018. Edible insects and other chitin-bearing foods in ethnic Peru: Accessibility, nutritional acceptance and food-security implications. Journal of Ethnobiology, 38(3): 424–447. Matthäus, B.; Piofczyk, T.; Katz, H.; Pudel, F. 2019. Renewable resources from insects: Exploitation, properties and refining of fat obtained by cold pressing from Hermetia illucens (Black soldier fly) larvae. European Journal of Lipid Science and Technology, 121(7): art. no. 1800376. Melgar-Lalanne, G.; Hernández-Álvarez, A.-J.; Salinas-Castro, A. 2019. Edible insects processing: Traditional and innovative technologies. Comprehensive Reviews in Food Science and Food Safety, 18: 1166–1191. Mekuria, S.A.; Kinyuru, J.N.; Mokua, B.K.; Tenagashaw, M.W. 2021. Nutritional quality and safety of complementary foods developed from blends of staple grains and honey bee larvae (Apis mellifera). International Journal of Food Science, art. no. 5581585. Milton, K. 1984. Protein and carbohydrate resources of the Maku Indians of Northwestern Amazonia. American Anthropologist, 86(1): 7–27. Mintah, B.K.; He, R.; Dabbour, M.; Xiang, J.; Agyekum, A.A.; Ma, H. 2019. Technofunctional attribute and antioxidative capacity of edible insect protein preparations and hydrolysates thereof: Effect of multiple mode sonochemical action. Ultrasonics Sonochemistry, 58: art. no. 104676. Miron, L.T.; Postma, R.P.; Bosch, G.; Eppink, M.H.M. 2019. Preliminary evaluation of aqueous protein extraction from black soldier fly larvae (Hermetia illucens L.). Industrial Biotechnology, 15(6): 365–369.

104  Edible Insects Processing for Food and Feed Mishyna, M.; Martinez, J.-J.I.; Chen, J.; Benjamin, O. 2019. Extraction, characterization and functionl properties of soluble proteins from edible grasshopper (Schistocerca gregaria) and honey bee (Apis mellifera). Food Research International, 116: 697–706. Mishyna, M.; Chen, J.; Benjamin, O. 2020. Sensory attributes of edible insects and insect-based foods – Future outlooks for enhancing consumer appeal. Trends in Food Science & Technology, 95: 141–148. Mishyna, M.; Keppler, J.K.; Chen, J. 2021. Techno-functional properties of edible insect proteins and effects of processing. Current Opinion in Colloid and Interface Science, 56: art. no. 101508. Mitsuhashi, J. 1997. Insects as traditional foods in Japan. Ecology of Food and Nutrition, 36(2–4): 187–199. Moruzzo, R.; Mancini, S.; Boncinelli, F.; Riccioli, F. 2021. Exploring the acceptance of entomophagy: A survey of Italian consumers. Insects, 12: 123. Montevecchi, G.; Licciardello, F.; Masino, F.; Miron, L.T.; Antonelli, A. 2021. Fortification of wheat flour with black soldier fly prepupae. Evaluation of technological and nutritional parameters of the intermediate doughs and final baked products. Innovative Food Science and Emerging Technologies, 69: art. no. 102666. Müller, A.; Wolf, D.; Gutzeit, H.O. 2017. The black soldier fly, Hermetia illucens – A promising source for sustainable production of proteins, lipids and bioactive substances. Zeitschrift für Naturforschung- Section C Journal of Biosciences, 72(9–10): 351–363. Ndiritu, A.K.; Kinyuru, J.N.; Gichuhi, P.N.; Kenji, G.M. 2019. Effects of NaCl and pH on the functional properties of edible crickets (Acheta domesticus) protein concentrate. Journal of Food Measurement and Characterization, 13(3): 1788–1796. Okia, C.A.; Odongo, W.; Nzabamwita, P.; Ndimubandi, J.; Nalika, N.; Nyeko, P. 2017. Local knowledge and practices on use and management of edible insects in Lake Victoria basin, East Africa. Journal of Insects as Food and Feed, 3(2): 83–93. Onore, G. 1997. A brief note on edible insects in Ecuador. Ecology of Food and Nutrition, 36(2–4): 277–285. Osimani, A.; Milanović, V.; Cardinali, F. et al. 2018. Bread enriched with cricket powder (Acheta domesticus): A technological, microbiological and nutritional evaluation. Innovative Food Science and Emerging Technologies, 48: 150–163. Pambo, K.O.; Okello, J.J.; Mbeche, R.M.; Kinyuru, J.N., Alemu, M.H. 2018. The role of product information on consumer sensory evaluation, expectations, experiences and emotions of cricket-flour-containing buns. Food Research International, 106: 532–541. Parsons, J.R. 2010. The pastoral niche in Pre-Hispanic Mesoamerica. In Pre-Columbian Foodways: Interdisciplinary Approaches to Food, Culture and Markets in Ancient Mesoamerica, eds. Staller, J., Carrasco, M., 109–136. Springer, New York. Payne, C.; Caparros Megido, R.; Dobermann, D.; Frédéric, F.; Shockley, M.; Sogari, G. 2019. Insects as food in the Global North – The evolution of the entomophagy movement. In Edible Insects in the Food Sector, eds. Sogari, G., Mora, C., Menozzi, D. Springer, Cham. Pintado, T.; Delgado-Pando, G. 2020. Towards more sustainable meat products: Extenders as a way of reducing meat content. Foods, 9(8): art. no. 9081044. Pliner, P.; Hobden, K. 1992. Development of a scale to measure the trait of food neophobia in humans. Appetite, 19: 105–120. Pliny the Elder. 1969. Natural History Books, 8–11. Harvard University Press, Cambridge, MA. Purschke, B.; Stegmann, T.; Schreiner, M.; Jäger, H. 2017. Pilot-scale supercritical CO2 extraction of edible insect oil from Tenebrio molitor L. larvae – Influence of extraction conditions on kinetics, defatting performance and compositional properties. European Journal of Lipid Science and Technology, 119(2): art. no. 1600134. Purschke, B.; Meinlschmidt, P.; Horn, C.; Rieder, O.; Jäger, H. 2018. Improvement of techno-functional properties of edible insect protein from migratory locust by enzymatic hydrolysis. European Food Research and Technology, 244(6): 999–1013. Raheem, D.; Carrascosa, C.; Oluwole, O.B.; Nieuwland, M.; Saraiva, A.; Millán, R.; Raposo, A. 2019. Traditional consumption of and rearing edible insects in Africa, Asia and Europe. Critical Reviews in Food Science and Nutrition, 59(14): 2169–2188. Ramos-Elorduy, J. 1998. Creepy Crawly Cuisine: The Gourmet Guide to Edible Insects. Park Street Press, Rochester, Vermont. Ramos-Elorduy, J. 2006. Threatened edible insects in Hidalgo, Mexico and some measures to preserve them. Journal of Ethnobiology and Ethnomedicine, 2: 51. Ramos-Elorduy, J. 2009. Anthropo-entomophagy: Cultures, evolution and sustainability. Entomological Research, 39: 271–288. Ribeiro, J.C.; Lima, R.C.; Maia, M.R.G. et al. 2019. Impact of defatting freeze-dried edible crickets (Acheta domesticus and Gryllodes sigillatus) on the nutritional value, overall liking and sensory profile of cereal bars. LWT – Food Science and Technology, 113: art. no. 108335.

5  •  Consumer Attitudes towards Insects as Food  105 Ribeiro, L.; Cunha, L.M.; García-Segovia, P.; Martínez-Monzó, J., Igual, M. 2021. Effect of the house cricket (Acheta domesticus) inclusion and process temperature on extrudate snack properties. Journal of Insects as Food and Feed, 7(7): 1117–1129. Roberts, M.S. 2019. Sustainability and traditional livelihood systems in Northern Lao PDR with an emphasis on edible insects. International Review of Modern Sociology, 45(2): 93–126. Roncolini, A.; Milanović, V.; Cardinali, F. et al. 2019. Protein fortification with mealworm (Tenebrio molitor L.) powder: Effect on textural, microbiological, nutritional and sensory features of bread. PLOS ONE, 14(2): art. no. e0211747. Roncolini, A.; Milanović, V.; Aquilanti, L. 2020. Lesser mealworm (Alphitobius diaperinus) powder as a novel baking ingredient for manufacturing high-protein mineral-dense snacks. Food Research International, 131: art. no. 109031. Sánchez, M.; Gómez, C.; Avendaño, C. et al. 2021. House fly (Musca domestica) larvae meal as an ingredient with high nutritional value: Microencapsulation and improvement of organoleptic characteristics. Food Research International, 145: art. no. 110423. Satterthwaite, D.; McGranahan, G.; Tacoli, C. 2010. Urbanization and its implications for food and farming. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences, 365(1554): 2809–2820. Scholliers, J.; Steen, L.; Fraeye, I. 2020. Gelation of a combination of insect and pork proteins as affected by heating temperature and insect:meat ratio. Food Research International, 137: art. no. 109703. Severini, C.; Azzollini, D.; Albenzio, M.; Derossi, A. 2018. On printability, quality and nutritional properties of 3D printed cereal based snacks enriched with edible insects. Food Research International, 106: 666–676. Shelomi, M. 2015. Why we still don’t eat insects: Assessing entomophagy promotion through a diffusion of innovations framework. Trends in Food Science & Technology, 45(2): 311–318. Sipponen, M.H.; Mäkinen, O.E., Rommi, K. et al. 2018. Biochemical and sensory characteristics of the cricket and mealworm fractions from supercritical carbon dioxide extraction and air classification. European Food Research and Technology, 244(1): 19–29. Sirimungkararat, S.; Saksirirat, W.; Nopparat, T. et al. 2010. Edible products from eri silkworm (Samia ricini D.) and mulberry silkworm (Bombyx mori L.) in Thailand. In Forest Insects as Food: Humans Bite Back, eds Durst, P.B.; Johnson, D.V.; Leslie, R.N.; Shono, K., 189–200. FAO, Bangkok, Thailand. Smarzyński, K.; Sarbak, P.; Musiał, S.; Jezowski, P.; Piatek, M.; Kowalczewski, P.T. 2019. Nutritional analysis and evaluation of the consumer acceptance of pork pâté enriched with cricket powder-preliminary study. Open Agriculture, 4(1), 159–163. Smarzyński, K.; Sarbak, P.; Kowalczewski, P.Ł. et al. 2021. Low-field NMR study of shortcake buiscuits with cricket powder and their nutritional and physical characteristics. Molecules, 26(17): art. no. 5417. Smetana, S.; Ashtari Larki, N.; Pernutz, C. et al. 2018. Structure design of insect-based meat analogs with highmoisture extrusion. Journal of Food Engineering, 229: 83–85. Smetana, S.; Pernutz, C.; Toepfl, S.; Heinz, V.; Van Campenhout, L. 2019. High moisture extrusion with insect and soy protein concentrates: Cutting properties of meat analogues under insect content and barrel temperature variations. Journal of Insects as Food and Feed, 5(1): 29–34. Smetana, S.; Leonhardt, L.; Kauppi, S.-M.; Pajic, A.; Heinz, V. 2020. Insect margarine: Processing, sustainability, and design. Journal of Cleaner Production, 264: art. no. 121670. Soetemans, L.; Uyttebroek, M.; D’Hondt, E.; Bastiaens, L. 2019. Use of organic acids to improve fractionation of the black soldier fly larvae juice into lipid- and protein-enriched fractions. European Food Research and Technology, 245(10): 2257–2267. Sogari, G.; Menozzi, D.; Mora, C. 2018. Sensory-liking expectations and perceptions of processed and unprocessed insect products. International Journal on Food System Dynamics, 9: 314–320. Stoops, J.; Vandeweyer, D.; Crauwels, S. et al. 2017. Minced meat-like products from mealworm larvae (Tenebrio molitor and Alphitobius diaperinus): Microbial dynamics during production and storage. Innovative Food Science and Emerging Technologies, 41: 1–9. Sutton, M. 1995. Archaeological aspects of insect use. Journal of Archaeological Method and Theory, 2(3): 253–298. Svanberg, I.; Berggren, Ǻ. 2019. Ant schnapps for health and pleasure: The use of Formica rufa L. (Hymenoptera: Formicidae) to flavour aquavit. Journal of Ethnobiology and Ethnomedicine, 15(1): art. no. 68. Tan, H.S.G.; van den Berg, E.; Stieger, M. 2016. The influence of product preparation, familiarity and individual traits on the consumer acceptance of insects as food. Food Quality and Preference, 52: 222–231. Tao, J.; Davidov-Pardo, G.; Burns-Whitmore, B.; Cullen, E.M.; Li, Y.O. 2017. Effects of edible insect ingredients on the physicochemical and sensory properties of extruded rice products. Journal of Insects as Food and Feed, 3(4): 263–278.

106  Edible Insects Processing for Food and Feed Tello, A.; Aganovic, K.; Parniakov, O.; Carter, A.; Heinz, V.; Smetana, S. 2021. Product development and environmental impact of an insect-based milk alternative. Future Foods, 4: art. no. 100080. Torres-Acosta, R.I.; Moreno-Ramírez, Y.D.R.; García-García, L.D. et al. 2021. Insects with phenolics and antioxidant activities to supplement Mezcal: Tenebrio molitor L. and Schistocerca piceifrons Walker. Southwestern Entomologist, 46(3): 625–635. Torres, K.S.; Sampaio, R.F.; Ferreira, T.H.B.; Argondoña, E.J.S. 2022. Development of cookie enriched with silkworm pupae (Bombix mori). Journal of Food Measurement and Characterization, 16: 1540–1548. Tuccillo, F.; Marino, M.G.; Torri, L. 2020. Italian consumers’ attitudes towards entomophagy: Influence of human factors and properties of insects and insect-based food. Food Research International, 137(11): 109619. Turpin, M.; Si, A. 2017. Edible insect larvae in Kaytetye: Their nomenclature and significance. Journal of Ethnobiology, 37(1): 120–140. Tzompa-Sosa, D.A.; Yi, L.; van Valenberg, H.J.F., Lakemond, C.M.M. 2019. Four insect oils as food ingredient: Physical and chemical characterisation of insect oils obtained by an aqueous oil extraction. Journal of Insects as Food and Feed, 5(4): 279–292. Tzompa-Sosa, D.A.; Dewettinck, K.; Gellynck, X.; Schouteten, J.J. 2021. Replacing vegetable oil by insect oil in food products: Effect of deodorization on the sensory evaluation. Food Research International, 141: art. no. 110140. Tzompa-Sosa, D.A.; Dewettinck, K.; Gellynck, X.; Schouteten, J.J. 2022. Consumer acceptance towards potato chips fried in yellow mealworm oil. Food Quality and Preference 97: art. no. 104487. Van Huis, A. 2017. Edible insects and research needs. Journal of Insects for Food and Feed, 3 (1): 3–5. Van Huis, A.; Van Itterbeeck, J.; Klunder, H.; Mertens, E.; Halloran, A.; Muir, G.; Vantomme, P. 2013. Edible insects: Future prospects for food and feed security. In Forestry Paper 171. FAO, Rome. Van Itterbeeck, J.; Van Huis, V. 2012. Environmental manipulation for edible insect procurement: A historical perspective. Journal of Ethnobiology and Ethnomedicine, 8(3): 3. Van Itterbeeck, J.; Rakotomalala Andrianavalona, I.N.; Rajemison, F.I. et al. 2019. Diversity and use of edible grasshoppers, locusts, crickets, and katydids (Orthoptera) in Madagascar. Foods, 8(12): art. no. 666. Vanhonacker, F.; Van Loo, E.J.; Gellynck, X.; Verbeke, W. 2013. Flemish consumer attitudes towards more sustainable food choices. Appetite, 62: 7–16. Verbeke, W. 2015. Profiling consumers who are ready to adopt insects as a meat substitute in a Western society. Food Quality and Preference, 39(1): 147–155. Verneau, F.; La Barbera, F.; Kolle, S.; Amato, M.; Del Giudice, T.; Grunert, K. 2016. The effect of communication and implicit associations on consuming insects: An experiment in Denmark and Italy. Appetite, 106: 30–36. Wei, Y.; Zhang, L.; Wang, J. et al. 2021. Chinese caterpillar fungus (Ophiocordyceps sinensis) in China: Current distribution, trading, and futures under climate change and overexploitation. Science of the Total Environment, 755(Pt1): 142548. Wendin, K.M.E.; Nyberg, M. 2021. Factors influencing consumer perception and acceptability of insect-based foods. Current Opinion in Food Science, 40: 67–71. Wendin, K.; Berg, J.; Jönsson, K.I. et al. 2021. Introducing mealworm as an ingredient in crisps and pȃtés – sensory characterization and consumer liking. Future Foods, 4: art. no. 100082. Yen, A.L. 2009. Edible insects: Traditional knowledge or Western phobia? Entomological Research, 39: 289–298. Yhoung-aree, J. 2010. Edible insects in Thailand: Nutritional values and health concerns. In Forest Insects as Food: Humans Bite Back, eds. Durst, P.B.; Johnson, D.V.; Leslie, R.N.; Shono, K., 201–216. FAO, Bangkok, Thailand. Yi, L.; Van Boekel, M.A.J.S.; Lakemond, C.M.M. 2017. Extracting Tenebrio molitor protein while preventing browning: Effect of pH and NaCl on protein yield. Journal of Insects as Food and Feed, 3(1): 21–31. Yu, M.-H., Lee, H.-S.; Cho, H.-R.; Lee, S.-O. 2017. Enzymatic preparation and antioxidant activities of protein hydrolysates from Tenebrio molitor larvae (mealworm). Journal of the Korean Society of Food Science and Nutrition, 46(4): 435–441. Zielińska, E.; Pankiewicz, U.; Sujka, M. 2021. Nutritional, physiochemical, and biological value of muffins enriched with edible insects flour. Antioxidants, 10(7): art. no. 1122.

Edible Insect Farming Rimsha Naseem, Waqar Majeed, Mian Muhammad Awais, Muhammad Noman Naseem, Naureen Rana, and Uzma Ramzan

6

Contents 6.1 Introduction 108 6.1.1 Entomophagy: Wild Harvesting to Insect Farming 108 6.2 Mini-Livestock: An Advantageous Farming Choice 110 6.2.1 Environmental Aspects 110 6.2.1.1 Life cycle assessment 110 6.2.1.2 Footprints of land and water 111 6.2.1.3 Greenhouse gas emissions 111 6.2.1.4 Minimize pesticides 111 6.2.1.5 Biodegradation of waste material 111 6.2.1.6 Resource inventory for insect farming 113 6.2.1.7 Feeding requirements 113 6.2.1.8 Energy consumption 113 6.2.1.9 Edible insects services for the agriculture system 115 6.2.1.10 Feed conversion ratio 115 6.2.1.11 Economical aspects 115 6.2.1.12 Transport 116 6.3 Types of Insect Farming 116 6.3.1 Traditional Insect Farming 116 6.3.2 Indoor Insect Farming 117 6.3.2.1 Mulberry silkworm 117 6.3.2.2 House cricket 117 6.3.2.3 Yellow mealworm 120 6.3.2.4 Black soldier fly (BSF) 120 6.3.2.5 Housefly (HF) 121 6.3.3 Outdoor Insect Farming 121 6.3.3.1 Grasshoppers 121 6.3.3.2 Palm weevil or Sago larvae 121 6.3.3.3 Bamboo caterpillar 122 6.3.3.4 Weaver ants 122 6.4 Cost of Cultivation 123 6.5 Challenges and Way Forward 126 Acknowledgment 126 References 126 DOI: 10.1201/9781003165729-6

107

108  Edible Insects Processing for Food and Feed

6.1 INTRODUCTION Human beings have always taken nature for granted, for all its resilience. Earth supported the lives of some 2.6 billion people about 70 years ago. In 1987, population growth had risen to around 5 billion people worldwide. The current population of 7.7 billion people is expected to increase to 9.7 billion by 2050 and 11 billion by 2100 (Hite and Seitz 2021). When production and demand grow exponentially, it puts pressure on the ecosystem and results in food scarcity, degradation of agricultural land, an increase in food prices, and disparity of resources among the people. Population pressure and exploitation of natural resources are the major dilemmas at stake in food security (FAO 2013a; Bélanger and Pilling 2019). If this condition persists, about 827 million people will starve (Van Huis 2013). This alarming statistic draws the world’s attention to seeking alternative food resources that are cost-effective and rich in nutrients. Food resources like livestock and agricultural products seem to be insufficient to feed our ever-growing human population. Therefore, important action must be taken worldwide to change the feeding behavior of societies, improve traditional farming practices (agriculture and livestock), and explore affordable, nutritious alternative food sources (Gahukar 2014). Insect farming is perceived as a promising substitute for coping with the food crisis (Van Huis 2015), offering high quality, cheap sources of beneficial proteins, lipids, amino acids, and various types of micro and macronutrients. This can address the current gloomy outlook of food security (Defoliart 1995; Rumpold and Schlüter 2013a, 2013b).

6.1.1 Entomophagy: Wild Harvesting to Insect Farming In the past, entomophagy was common in tropical or subtropical parts of the world, such as Southeast Asia and sub-Saharan Africa, and non-insect-eaters were mainly restricted to temperate zones, such as Europe, Russia, and northern parts of North America (Waltner-Toews 2017), where insects only used as a diet for farmed animals, e.g., chickens, pigs, and especially fish as a source of high-quality protein. Due to apparent nutritional, ecological, ethical, and economic benefits, the ancient negative prejudice toward insect eating is fading in Europe and European-driven societies (Van Huis 2013). Approximately 1,000–2,200 species of insects are being used as dietary components in 130 countries by 2 billion people (Jongema 2017; Tao and Li 2018). Even though most people still oppose insects as food, but consumption is rising in Western countries (Yen 2015). Figure 6.1 shows the total number of insect species consumed by peoples in different biogeographical regions (Naseem et al. 2021). Approximately 95% of insect species are captured from the wild, which could be a serious concern from a health point of view and responsible for a decrease in insect biodiversity. Moreover, increased agricultural practices, urbanization, and desertification have contributed to a decrease in the wild insect harvest (Looy, Dunkel, and Wood 2014; Yen 2015). Recorded data showed that many species like Xyleutes redtembacheri, Comadia redtembacheri (agave worm), Liometopum apiculatum (Navajo reservation ant), and Scyphophorusacu punctatus (agave weevil) are the victims of overharvesting (Morales-Ramos, Rojas, and Shapiro-Ilan 2014). On the other hand, pesticides and insecticides on crops harm many species such as Locusta migratoria and Anaphe panda (Cerritos 2009; Van Huis 2013). Semi-cultivation or outdoor farming accounts for 6% of all types of insect farming and is mostly practiced in Papua New Guinea, sub-Saharan Africa, Indonesia, Malaysia, Amazon basin, and Thailand. The prominent species used for semi-cultivation are mopane caterpillars, palm weevil grubs, and the eggs of aquatic bugs (Gerda et al. 2001). In semi-cultivation, insects are farmed in their natural habitat, with certain limitations pertaining to the availability, abundance, location, and season of resources. Now the world can look forward to a better version of insect cultivation utilizing indoor farming, which corresponds to 2% of insect farming types and ensures the maximum availability of insects under controlled conditions (Zhu et al. 2012; Nadeau et al. 2015; Yen 2015; Raheem et al. 2019; Feng et al. 2009).

6  •  Edible Insect Farming  109

356

Nearctic

97 754

Palearctic

African

618

Oriental

107

Neotropical Australian

471 FIGURE 6.1  Total number of edible insect species in different biogeographical regions. (Reprinted with permission from Springer Nature [Naseem et al. 2021].)

FIGURE 6.2  Summary of edible insect farming potentials based on socio-economical aspects.

Insect farming is practiced in many regions for different purposes due to its sustainability and effective nutrition. In Asian countries, Thailand seems to pioneer for edible insect farming based on medium and large-scale farms for mass production, processing, transport, and marketing. Approximately 20,000 small farms run in Thailand and successfully produce house crickets, mulberry silkworms, palm weevil

110  Edible Insects Processing for Food and Feed larvae, yellow mealworms, grasshoppers, and locusts (Hanboonsong, Jamjanya, and Durst 2013; Dossey, Tatum, and McGill 2016). In Cambodia, Laos, and Vietnam, crickets are popular for insect farming. Gonimbrasia belina (Mopane caterpillars) are available all year, raised and utilized by 70% of Congolese people and 95% of people in the Central African Republic (Kenis et al. 2014). In China, a total of 324 species belonging to different orders are consumed (Chen, Feng, and Chen 2009; Zhu et al. 2012), but mass production of insects has not been established successfully (Marone 2016). In Africa, the majority of the population is undernourished; it is mainly sub-Saharan Africa that is grappling with food insecurity, and a twin track approach consisting of agriculture and insect farming is being used to reduce hunger. In East Africa, insect farming has gained remarkable interest due to its sustainability as a food system, high economic potential, and the scarcity of arable land and water (Van Huis 2013). Collaborations between African and European enterprises on value chain manufacturing in the growing, processing, distribution, and consumption of edible insects are being developed in Africa (Halloran and Münke 2014; Halloran et al. 2016). In Western countries, insect farming is primarily a family-run business which focuses on zoo and pet animal feed, but some farms produce food for human consumption in small amounts. (FAO 2013a). The most farmed cricket species are Gryllus bimaculatus, Acheta domesticus, and Gryllodes sigillatus, while the most farmed mealworm species are T. molitor, Z. morio, and A. diaperinus. Some other species such as L. migratoria belong to locusts, G. mellonella is a known species of waxworm, and maggots of house fly, sun beetles, and cockroaches are farmed in the Netherlands and the United States (FAO 2013a). The cricket pet feed industry in the United States is labelled as a multimillion-dollar enterprise, with up to 50 million crickets produced per week (Weissman et al. 2012). For family livelihood, stink bug species are a valuable source of nutrition and income for native people America (Cerritos 2009). Mechanical harvesting is common in Thailand and Mexico to produce red grasshopper and weaver ants (Ocophylla smaragdina Fab.) respectively (Offenberg 2011). Insect farms are rare in Western countries, while the practice of farming insects has been going on in China for over 5,000 years. In North America, no farms grew insects specifically for food before 2012, but many farms grew mainly crickets and mealworms to produce feed for pets and fishing bait. Many USA and Canadian insect farms have been heavily focused on robotics, mechanization, automation, sensor technology, and data aggregation to iterate quickly towards the insect farms of the future (Shockley et al. 2018). In Mexico, between 2012 and 2017, only a few start-up farms were working on the domestication of traditionally consumed insects.

6.2  MINI-LIVESTOCK: AN ADVANTAGEOUS FARMING CHOICE Insect farming is considered an efficient way of sustenance that ensures ecological, economic, and environmental sustainability for humans compared to other livestock farming. It contributes good quality proteins and other essential dietary components in food and feeding materials despite all its disgusting factors. Insect farming is exactly represented by the saying “great outlay for little cost”. If insects become a profitable dietary component for humans, large quantities of insects need to be produced continuously. This means that both farming and processing need to be highly advanced and approachable (Berggren, Jansson, and Low 2018). A detailed description is given in Figure 6.2.

6.2.1 Environmental Aspects 6.2.1.1 Life cycle assessment Life cycle assessment (LCA) is a complete tool for estimating the effect of the environment on goods and services provided by the relevant organism, product, or system (Curran 2016). LCA was first applied to

6  •  Edible Insect Farming  111 the farming of insects for food in 2012 and for feed in 2015 (Van Zanten et al. 2015; Oonincx and de Boer 2012; Roffeis et al. 2015). Insect farming is carried out efficiently based on the species, life cycle, habitat, and host plant. Tenebrio molitor, Musca domestica, and Hermetia illucens are the only three species of insects to have been the main focus of an LCA as food and feed. The functional unit is the necessary constituent of life cycle analysis and is based on mass, nutrient, or economy, e.g., what type of product is obtained from the insect and what fraction of this product (proteins or micro-nutrients) is beneficial to humans or animals as compared to another substitutional product (de Vries, van Middelaar, and de Boer 2015; de Vries and de Boer 2010).

6.2.1.2 Footprints of land and water According to an estimate, two-thirds of the world’s population will be destitute of freshwater sources by 2025 (FAO 2013a), and about 1.8 billion people have to live in regions with an insufficient water supply. About 70% of the water is used in livestock farming and different agricultural practices (Doreau, Corson, and Wiedemann 2012). The measure of water footprint or the amount of freshwater by volume is necessary for FCE calculations. Recorded data by (Miglietta et al. 2015) showed that mealworm fostering required less water than any other livestock. Yellow mealworm and lesser mealworm can be reared on 2 L/kg of water or even on organic biowastes (Ramos-Elorduy et al. 2002), while livestock rearing requires 2200 L/kg of water, which is considered a large amount of water in comparison with mini-livestock (Chapagain and Hoekstra 2003). Comparative analysis of various species for insect farming is shown in Figure 6.3. Studies showed that commercial farming of insects (L. migratoria and T. molitor) requires minimal land use (Oonincx and de Boer 2012).

6.2.1.3 Greenhouse gas emissions Greenhouse gas (GHG) emissions from livestock farming have a larger effect on global warming, and one of the possible advantages of implementing insect production schemes may be a decrease in these emissions. It is well known that certain insects produce a significant amount of methane. However, there is relatively little detail known on GHG emissions from insect farming. Oonincx et al. (2010) studied the greenhouse gas and ammonium emissions associated with T. molitor, A. domesticus, and L. migratoria. They observed that CO2 gas is extremely reliant upon different metamorphic stages, alimentation, temperature, and level of activity, as shown in Figure 6.3. Agriculture and dairy farms are notorious for discharging CO2 (14.8 kg) (Steinfeld et al. 2006), CH4 (31%), NO2 (65%), and two-thirds of the ammonia (Steinfeld et al. 2006) in the environment (Fiala 2008; Van Huis 2013; Sachs 2016).

6.2.1.4 Minimize pesticides Insect-eating helps to minimize pesticide consumption and many harmful chemicals that would become part of the diet indirectly. Edible insect species also aid in the biological control of pests by acting as pests of economically valuable plants that are traditionally controlled with insecticides. However, they would be largely caught artificially with extra profits bringing in. As a result, the number of mating adults would be controlled, limiting the future generation of pests and the usage of pesticides reduces, the resistance that insects against control will be reduced (Cerritos 2009; Kouřimská and Adámková 2016). In the meantime, organically botanic goods can be provided to approach best for the Integrated Pest Management practices (IPM) (DeFoliart 1997).

6.2.1.5 Biodegradation of waste material In low and middle-income countries, the proper disposal of waste is still challenging. Usually, the disposal of waste without proper treatment can have negative impacts on environment and can cause major health issues (Hoornweg 2014). Up to 85% of domestic wastes is degradable, and every year, 1.3 billion

112  Edible Insects Processing for Food and Feed

FIGURE 6.3  Comparison chart between edible insects species and other livestock based on different parameters like (A) digestible body mass %, (B) FCR (feed conversion ratio), (C) GHG (greenhouse gas) emission based on body mass gain, (D) ammonia production g/kg mass gain, (E) GWP (global warming potential) based on CO2 gas production per kilogram of meat production, (F) energy consumption, (G and H) land and water use respectively. (Reprinted with the permission of Elsevier [Gahukar 2016].)

6  •  Edible Insect Farming  113 tons of food waste are generated, resulting in a financial loss of approximately 1 billion dollars annually (FAO 2013b). The underuse of organic waste management results in the huge loss of valuable nutrients. As a result, there is an urgent need to develop multiple potential food waste recycling and recovery techniques to minimize and mitigate the negative impacts of food waste. Biowaste processing with fly larva is a recent invention that is gaining momentum. Eco-friendly and sustainable management of waste is among the urgent issues for countries that want to provide the best possible lifestyle for their populations. Composting is one of the most modern methods for recycling waste food in most African countries. This is a time-consuming process, and compost is the only product of this process. Black soldier fly (BSF) larvae do not require specific microorganisms and help in organic waste management. BSF take only two weeks to produce natural or biofertilizer used for crop cultivation, and larvae biomass acts as a viable protein source in livestock feed. This fly species is also famous for being the best alternative for recycling organic waste and ensuring environmental safety as it reduces greenhouse gas emissions (Li et al. 2011; Dzepe et al. 2021).

6.2.1.6 Resource inventory for insect farming This analysis mainly corresponds to the consumption of resources (construction of facilities feed) used by insects during their different life stages. The scale of production and intensity determines the housing requirements and other facilities, which may vary between regions. In warm climates, small-scale insect production is done using simple structures; for example, in Thailand, Gryllus bimaculatus and A. domesticus are farmed using AAC concrete structures with tin roofs and open sides, which characterize a low impact system (Halloran and Münke 2014). In comparison with Thailand, Canada and the Netherlands use advanced systems for insect production that resemble with livestock farming. The future of insect production is difficult to predict, but one factor is always a matter of concern: whether the production system is economically advantageous or not. In the case of small-scale insect farming, upscaled production will need to be treated on a case-by-case basis.

6.2.1.7 Feeding requirements Productivity and turnover are high in insect production systems, and these are two fundamental aspects of any production system. Unlike other animals, insects can be farmed and fed using less expensive and more sustainable sources (Miech et al. 2016). The majority of consumed species prefer herbivory (Van Huis 2015, 2013), although in Thailand some species of crickets are fed on chicken proteins to minimize harvesting time. Tenebrio molitor (mealworms) are normally reared on wheat flour or wheat bran added with yeast, skimmed milk powder, and soybean flour (Makkar et al. 2014). In some cases, the nutrition composition of insects may be altered by altering their feed; for example, crickets fed with cassava tops and weeds have higher methionine contents than crickets fed with common chicken feed (Miech et al. 2017). Current studies also explained the use and benefits of organic side streams as the feed for mealworms and crickets that proved the most promising path for producing both species economically with no competition (Halloran et al. 2016; Van Zanten et al. 2015). In the livestock sector, feed optimization is an important feature of efficiency. For this purpose, high formula feeds have been designed with precise nutrition that resulted in maximum and quick growth of animals but did not prove to be a sustainable feed source. It is hard to predict the future of formulated feed products for insects, and it depends on producer interest (Halloran et al. 2016). However, nutrient requirements like carbohydrates, lipids, amino acids, vitamins, and minerals have been studied and are described in Table 6.1.

6.2.1.8 Energy consumption The poikilothermic nature of insects helps to change their core body temperature with environmental fluctuations, resulting in high metabolic rates compared to other birds and animals (National Research

Glucose+ Maltose+ Sucrose+ Fructose+ Galactose+ Arabinose++ Ribose++ Xylose++ Galactose++

CARBOHYDRATES

Phospholipids+++ Linolenic+++ Linoleic (Pfa)+++

LIPIDS

PROTEINS Globulins*** Lipoproteins*** Nucleoprotein*** Insoluble proteins

MACRO-NUTRIENTS

Leucine Isoleucine Valine Threonine Lysine Arginine Methionine Histidine Phenylalan Tryptophan Tyrosine Proline Serine Cysteine Glycine Aspartic acid

AMINO ACIDS (***)

Ascorbic acid* Ergocalsiferol* Cholecalsiferol* Retinol* Phyloquinone* α and βcarotene* B1: Thiamin** B2: Riboflavin** B3: Nicotinamide** B4: Choline** B5:Pantothenicacid** B6: Pyridoxine** B12: Cobalamine**

VITAMINS

STEROLS Ergosterol Cholesterol Phytosterols

MICRO-NUTRIENTS

TABLE 6.1  Summary of macro-nutrients and micro-nutrients required for insect species during farming (Morales-Ramos, Rojas, and Shapiro-Ilan 2014; Varelas 2019), (+): absorbed and metabolized by insects; (++): absorbed but not metabolized by insects; (+++): not synthesized by insects; (***): synthesized by insects; (*): lipid soluble; (**): water-soluble vitamins

114  Edible Insects Processing for Food and Feed

6  •  Edible Insect Farming  115 Council 2011). Many studies explained that the high metabolisms of insects favor the use of waste or other by-products for insect growth that could not be used for livestock (Van Huis 2013). Temperature fluctuations also regulate the growth of the same species. For example, the optimal temperature for the growth of A. domesticus is 30°C with the accomplishment of the life cycle in 8 days, while the same species at 18°C took 8 months to grow. This means a long life cycle leads to a high demand for feed and water inputs (Ayieko et al. 2010). In a study in the Netherlands, surplus heat generation by metabolism had a significant impact on the growth of larvae in the mealworm production system. This heat could also be used to grow smaller, more heat-demanding larvae (Oonincx and de Boer 2012). Insect rearing systems have not been mechanically optimized, and they could greatly depend upon manual labor (Rumpold and Schlüter 2013b). Geographical locations also influenced insect farming; for example, tropical climatic conditions are suited for A. domesticus rearing because of relatively high temperatures (Oonincx and de Boer 2012). In comparison with insects, animals require more technically equipped and energy intensive input systems for temperature regulation (Halloran et al. 2017).

6.2.1.9 Edible insects services for the agriculture system Edible insects in the agriculture system provide provisional services (food and income), regulating services (pollination and biological control), and supporting services (soil formation, nutrient recycling, water infiltration, and water retention). In Mexican maize fields, edible grasshoppers (Sphenarium purpurascens) are harvested. Edible insects are plentiful by-products of low-intensity agriculture, and they are a typically seasonal and protein-rich food. Some farmed insects, such as weaver ants, provide help in biological control by acting as predators for more than 50 species of pests (Way and Khoo 1992). In farming, process ants work as sustainable ecological units and are inexpensive in terms of cost. Their collection and processing require little energy, and they are consumed locally and/or sold for a profit. Many edible insects are threatened by agricultural intensification, which includes mechanization, tree clearance, and pesticide use (Payne and Van Itterbeeck 2017).

6.2.1.10 Feed conversion ratio Feed conversion ratio (FCR) is a pivotal pillar in considering and choosing alternative food resources for the future. FCR is an efficiency evaluation tool in commercial livestock production (Nakagaki and Defoliart 1991; Ramos-Elorduy 2008, 2009; Oonincx and de Boer 2012; Van Huis 2013; Van Huis and Oonincx 2017). FCR studies on the protein conversion ratios of different insect species showed economically viable options to meet energy demands (Collavo et al. 2020; Offenberg 2011). Lundy and Parrella (2015) explained that a grain-based diet for the production of A. domesticus had better FC efficiency when compared to livestock species, as described in Figure 6.3. Protein density and composition are considered the primary elements for growth rates and efficiencies, as described by (Oonincx et al. 2015). Insects do not need much energy to maintain their body temperature, and this is an important reason for their higher FCRs and growth. FCR is also linked with temperature, environmental, feeding, and nutritional conditions; e.g., in the case of B. mori, these conditions mark its capacity to digest, absorb, and convert the consumed feed into body mass (Rahmathulla et al. 2005). The edible portion of insects is another influential factor for production system than conventional livestock (Nakagaki and Defoliart 1991).

6.2.1.11 Economical aspects Insect farming does not require a large start-up cost, and not many resources are required to harvest such species. For farmed species, a diverse range of low-cost fodders is used with high energy transfer. Interestingly, T. molitor has an ECI (efficiency of conversion of ingested food) value of 53%–73%,

116  Edible Insects Processing for Food and Feed although, in the case of most animals, it is recorded as almost 40% (Morales-Ramos and Rojas 2015). Many families in Southeast Asia and India practice insect farming (Halloran et al. 2017; van Huis and Oonincx 2017). As insect farming needs labor from production to sale, it opens up new windows for employment. Therefore, mini-livestock farming is better for increasing regional income (Tang et al. 2019). Another benefit of insect consumption is product availability in terms of quality and quantity. It is also economical when compared to other livestock and vegetable proteins (Van Der Werf and Salou 2015).

6.2.1.12 Transport It is difficult to say whether insects will entail different transport resources than other livestock products. Production is currently much more localized than livestock production, a global industry with long transport distances. Moreover, some companies in Europe produce freeze-dried products from commercially produced insects. Transportation is frequently described as an important category in the agri-food industry. By reducing the density of the insect products, freezing and drying can potentially reduce cost influences linked with transportation (Roy et al. 2009).

6.3  TYPES OF INSECT FARMING 6.3.1 Traditional Insect Farming Several tribal communities, for example in Africa and Thailand, raise and harvest insects in traditional ways (Raheem et al. 2019; FAO 2013a). People have to travel long distances to capture insects for wild harvesting because it is the only means of sustenance. There is a need to enhance these procedures so that natives can rely on the insect for the survival of their families as well as to make a good living and improve the conditions in which they live. Wild harvesting of insect fauna depends upon various climatic factors, seasonal factors, insect abundance, habitat, predators, and bushfire. Biodiversity conservation is considered a vital tool for sustainable wild harvesting in the case of the non-availability of the primary host, and alternative plant resources have a great impact on the survival rate of entomofauna. The Tasar silkworm, Eri silkworm, and Muga silkworm are three wild silkworm species that are successfully farmed on wild host plants (Gahukar 2014). In Zambia, the population of mopane caterpillars can be managed by protecting and monitoring the host tree species from bushfire (Chidumayo and Mbata 2002). Additionally, the overexploitation of edible insect species can be controlled and managed by the forestry community with the help of native people (Bhatia and Yousuf 2013). Several initiatives have been taken to support or conserve the biodiversity of insects in forests, such as making an inventory of new host plants to maintain the natural ecological balance. In contrast to traditional wild collection, insects can also be farmed under controlled conditions, which increase opportunities for employment in rural and city zones (Gray 2014). Hanboonsong, Jamjanya, and Durst (2013) suggested that cricket farms run on a small scale in primary schools could serve as inexpensive and valid protein sources to cope with malnutrition. In Africa and Thailand, insects can be reared with pumpkin and other vegetables at times of fluctuating environmental conditions. In rural areas, insect farming is a better way of improving livelihoods because of sustainable and low-cost practice (Gahukar 2011b, 2011a). It can be practiced with the help of women and children to increase rural food stock in the form of edible insect species (Kim, Kim, and Oh 2008). By integrating traditional and ecological knowledge, the monetary value of insects can be determined (Losey and Vaughan 2006). In conclusion, traditional farming of edible insects helps establishing a good income source for resource-poor producers. There is a need for proper training and basic know-how in rearing insects to meet these goals.

6  •  Edible Insect Farming  117

6.3.2 Indoor Insect Farming Indoor farming of edible insects has emerged as a more efficient, appealing, and sustainable alternative to wild harvesting for large-scale insect production (Raheem et al. 2019; Nadeau et al. 2015). Developed countries such as America and Europe, where insect farming was not adopted traditionally, have now increased their market share by combining its economic and nutritional aspects. Indeed, insect-based food and feed products are equally beneficial for developed and developing countries (Cadinu et al. 2020). For the rearing of mulberry worm in China, Korea, and India, indoor farming is preferably done using local materials such as bamboo pools, nylon nets, and the construction of mud walls (Feng et al. 2009; Sathe, Ghodake, and Sathe 2011; Kim, Kim, and Oh 2008). Similarly, in Japan, for the rearing of Vespula sp., wooden hive boxes are used (Nonaka 2009), while in Thailand, insect species are reared in cement tanks, concrete tanks, cylinders, plastic boxes or drawers, or wood boxes enclosed by plastic sheet, as shown in Figure 6.4 (Hanboonsong, Jamjanya, and Durst 2013), and jars made of clay are used for the collection of stingless bees in Mexico (Ramos-Elorduy 2009). In India, many bees like Apis dorsata and Apis cerana indica Fb. have been successfully farmed by (Ronghang and Ahmed 2010). Indoor farming is a practical choice because it is efficient and needs only a small area for establishment. Indoor cultivation of insects requires the proper photoperiod, temperature, relative humidity, quality feed, and deterrence of parasites and diseases. Silkworm mass production is done using fortified feed with proteins, vitamins, and essential nutrients (Gahukar 2014). Commercially beneficial species like house crickets and yellow mealworms are reared in Europe and North America for pet feed. In Thailand, grasshoppers, crickets, and many other types of insects are successfully farmed by 20,000 entrepreneurs (Hanboonsong, Jamjanya, and Durst 2013).

6.3.2.1 Mulberry silkworm Bombyx mori L. is a domesticated, monophagous, and indoor farmed insect because of its commercial importance and the high demand for its larvae and pupae in the food and feed industry. India is the second most recognized country for its well-established silkworm industry. Rearing is done using mostly single tray or multilayered trays supported by bamboo poles in cement or mud houses. Among the different types of silkworms, like bivoltine, multivoltine, and F1, the bivoltine strain is considered the best fit in indoor farming due to its high larval weight and fecundity (Areerat et al. 2021). Usually, nylon mesh or cement houses are considered suitable and cheaper for rearing the larvae of the mulberry worm (Gahukar 2014). Successful rearing of B. mori has been done on chopped leaves and shoots of Morus alba. The selection of foliage and leaf determines the rate of production. Cultivation practices, plant nutrition, and soil type affect the moisture level in plant leaves, which ultimately affects the rearing of the silkworm. Shootreared larvae showed 50% more development with less cost than those reared on leaves, which showed 15% development. The use of dry leaves in rearing beds lessens the chance of pathogen attack (Tayal and Chauhan 2017). Biofortification must be done to avert the effects of pathogens with the assistance of different types of herbal tonics, jaggery solutions, and protein powders (Malik and Reddy 2007; Bhede, Lande, and Pathrikar 2013; Amala Rani et al. 2011). The rearing spectrum of silkworms is described in Figure 6.6 A.

6.3.2.2 House cricket Crickets and other orthopterans are considered the preferred edible insects among Europeans from the acceptance point of view. They have a high protein profile range between 6.25% and 77.13%, which is an important dietary requirement; a high amount of chitin consumption has a significant impact on the microbiota of the human gut (Rumpold and Schlüter 2013a, 2013b; Stull et al. 2018; Kulma et al. 2020). In addition, the high amount of polyunsaturated fatty acids in crickets lowers the risk of many heart diseases such as coronary artery disease (Paul et al. 2017). Thailand is famous for its cricket production;

118  Edible Insects Processing for Food and Feed

FIGURE 6.4  Types of breeding containers: (a) concrete cylinder, (b) concrete block, (c) plywood box (d) plastic drawers. (Reprinted with the permission of FAO data copyrights [Hanboonsong, Jamjanya, and Durst 2013].)

approximately 22,340 cricket farmers have been recorded as being involved either in large or small scale cricket farming. Initially, three crickets (Gryllus bimaculatus, Teleogryllus testaceus, T. occipitalis) were used for farming, but later, Acheta domesticus was found to be most effective in farming and production (Hanboonsong, Jamjanya, and Durst 2013). A. domesticus (the house cricket) is well recognized as a traditional and most promising food source due to its rich and attractive dietary profile. It has a low feed conversion ratio compared to other farmed animals, which is why it is a trending farmed inset in Thailand, Lao, and Vietnam (Fernandez-Cassi et al. 2019). Crickets are made up of 77.13% protein content, 25–33 g/100 g fat content, and 40% fatty acid content as dry matter (Pastell et al. 2021). In Finland, indoor cricket-rearing facilities consist of plastic boxes with a top lid having a plastic net to allow airflow and prevent condensation. In Thailand, four different types of rearing facilities – concrete cylinder pens, concrete block pens, plywood boxes, and plastic drawers – are used for cricket rearing, as presented in Figure 6.5. Small-scale concrete cylinders (80 ×

6  •  Edible Insect Farming  119

FIGURE 6.5  Cricket-rearing sheds at different sites in the northeast: (a) Loei Province, (b) Buri Ram Province, (c) Maha Sarakham Province, (d) Khon Kaen Province. (Reprinted with the permission of FAO [Hanboonsong, Jamjanya, and Durst 2013].)

120  Edible Insects Processing for Food and Feed 50 cm) with 20 to 150 pens per unit are used for farming due to this system’s low cost and easy availability. Concrete block pens are also extensively used for medium and large-scale farming. The facility size depends upon space, usually comprising rectangular and interconnected pens of 1.2 m long, 2.4 m wide and 0.6 m high, having a capacity of 25 to 30 kg per pen. Overcrowding or overheating are the major drawbacks that lead to outbreaks of diseases and pests such as mites in concrete blocks. Plywood boxes 1.2 × 2.4 × 0.5 meters in size are made of plywood or gypsum boards, with a 20–30 kg/pen capacity. These boxes are easy to clean and heat resistant, but are less stable in cold, damp, or hot conditions than concrete block pens. In the plastic drawer system, three to four drawers of crickets stacked on top of each other take up very little space and are appropriate for small or medium-scale farming. Boxes measuring 0.8 × 1.8 × 0.3 meters have a 6–8 kg/box cricket-rearing capacity (Hanboonsong, Jamjanya, and Durst 2013). Dietary composition or feed has a strong impact on the rearing of crickets. A diet rich in protein, particularly chicken protein, and enriched with omega 3 has been found to be the most effective feed. Before harvesting, crickets are fed with different vegetables like pumpkins, morning glory leaves, watermelons, and cassava leaves to improve their taste and minimize the cost of a high protein diet (Lundy and Parrella 2015).

6.3.2.3 Yellow mealworm Tenebrio molitor L. is a commonly reared insect in Europe and is considered an important species for large-scale biomass conversion into proteins. This species can be used due to its high protein contents and resistance to harmful components like pesticides, mycotoxins, and heavy metals (Bordiean et al. 2020). It is a highly promising and popular species due to its high nutritious profile of essential amino acids similar to soybean and fish meal and serves as an alternative livestock feed source. A mealworm diet in poultry feed enhances digestibility and growth rate (De Marco et al. 2015). In 5 g wooden or plastic containers with holes in the lids for aeration, 500 beetles can be farmed (Hanboonsong, Jamjanya, and Durst 2013). Numerous diets, such as canola oil, peanut oil, soy protein, dry egg, and dry potato flour, are effective nutritional supplements for indoor rearing (Morales-Ramos et al. 2013). To shorten larval development time, reduce insect mortality, and increase weight gain. A diet high in protein obtained from yeast appears to be superior to feed used by commercial breeders (Van Broekhoven et al. 2015). The FCR of yellow mealworms is higher than conventional livestock when provided with the optimal diet. A high protein content in the larval diet ensures a higher survival rate for the larvae and reduces the developmental time (Oonincx et al. 2015). Yellow mealworm is very likely farmed on a huge scale as food and several beneficial compounds like proteins, fatty acids, chitin and amino acids, or other beneficial components in the near future.

6.3.2.4 Black soldier fly (BSF) Hermetia illucens larvae are detritivores, are holometabolous, and can survive on various organic substrates as food sources. BSF larvae (black soldier fly) have the ability to convert organic waste into biofertilizers used to improve soils. The larval biomass is used to provide proteins and dietary fats for animal feed in medium to large scale production systems. The production system of BSF is classified into two categories. In the first system, natural substrate is provided to BSF females for egg laying; this system is mainly used by hobby gardeners and individual farmers. The second system deals with separate adult rearing in special cages for egg production. Rearing of BSF adults is somewhat difficult task and requires cages, greenhouses, or rearing rooms, while the larval population can be maintained in trays, bays, and digesters (Diener et al. 2011). Indoor farming of BSF larvae is done in circle-shaped containers 30 cm × 12 cm in size, which are covered by net to protect them from predators. Approximately four seeded containers can be used to rear 500 larvae, which use organic material like fruit waste as feed. During their larval stage, the BSF larvae feed voraciously on various types of organic waste and convert carbohydrates and nitrogen compounds into secondary protein foodstuffs (Nyakeri et al. 2019). The survival rates of black soldier fly larvae are higher, about 77% for those reared on vegetable waste and 47% for those reared on household waste (Dzepe et al. 2021).

6  •  Edible Insect Farming  121

6.3.2.5 Housefly (HF) The ecology and biology of the house fly (Musca domestica) have been studied many times and they are mostly used in the production of animal feed (Makkar et al. 2014). The HF is usually considered a carrier of several human and animal diseases, that’s why it is less popular than other species for insect farming (Greenberg 1973). At a small scale, HF production is carried out in China and Ghana (Charlton et al. 2015). The use of housefly larvae for the biodegradation of manure provides an opportunity to reduce waste disposal, whereas its biomass can also be used as a protein-rich animal feed. The rearing of house fly larvae is done on animal manures and maize bran, livestock and poultry wastes, a mixture of egg content and wheat bran. The nutritional value of M. domestica larva meal was comparable to that of most high protein feed ingredients. Larva meal was 60% protein with a well-balanced amino acid profile and 20% fat with 57% monounsaturated fatty acids and 39% saturated fatty acids. The larva meal was deficient in omega-3 fatty acids (Hussein et al. 2017).

6.3.3 Outdoor Insect Farming In semi-cultivation, insects are not kept in captivity because of their natural availability to native people in particular habitats. This facilitates conservation practices and enhances food security in less privileged areas. Some cottage methods are practiced by locals in the Amazon basin, Papua New Guinea, sub-Saharan Africa, Malaysia, Thailand, and Indonesia for the rearing of palm weevil grubs. Mopane and bamboo caterpillar termites are successfully farmed in tropical Africa and Thailand, respectively. In Mexico, eggs of aquatic bugs are obtained by semi-cultivation (Gahukar 2016). In some parts of South America, insects are reared in integration with local fruits and vegetables. Termite semi-cultivation on sorghum and different cereals is popular in Western Africa (Farina, Demey, and Hardouin 1991).

6.3.3.1 Grasshoppers Grasshoppers are well known for their large amount of biomass production throughout the year (Blásquez, Moreno, and Camacho 2012). The chemical constituents of grasshoppers mainly depend on its food habits; for example, members of the Pyrgomorphidae, Romaleidae, and Acrididae families are graminivores (Finke and Oonincx 2013), while the members of the Tettigoniidae family obtain their nutrients from small insects and the soft parts of plants filled with sap (Capinera 2005). Other environmental factors like the intensity of light, temperature, and humidity influence the sex and life stages of grasshoppers. Grasshoppers are mainly collected from maize and paddy fields with the help of direct handpicking at night or in the early morning using nets and cloths (Paul et al. 2016). Rearing boxes for grasshoppers are shown in Figure 6.6 C.

6.3.3.2 Palm weevil or Sago larvae Rhynchophorus ferrugineus belongs to the family Curculionidae and is also known as the red-stripe palm weevil, the Asian palm weevil, and the Sago palm weevil. Palm weevil larvae are naturally farmed on sago palm tree trunks (Metroxylon sagu, M. rumphii) covered with tree barks in Thailand, Malaysia, Indonesia, and Papua New Guinea (Chinarak, Chaijan, and Panpipat 2020). Indoor farming of the sago palm weevil is also carried out in plastic containers provided with ground palm stalk and pig feed. The life cycle of the larvae is completed in 40–45 days in a palm trunk; however, for indoor farming, the life cycle of palm weevil larvae is completed in 25–30 days with the production of 1–2 kilograms of larvae per container (Gahukar 2016; Hanboonsong, Jamjanya, and Durst 2013). Sago palm weevil farming is highly economical as only US $0.4 is required for 1 kg production, and 400–600 kg of larvae earn a total income of US $2,625–3,937 (Hanboonsong, Jamjanya, and Durst 2013).

122  Edible Insects Processing for Food and Feed

FIGURE 6.6  Farming of different edible insect species (A) Mulberry silkworm farming, (B) ant farming, (C) rearing cages for grasshoppers. (Reprinted with the permission of Elsevier [Gahukar 2016].)

6.3.3.3 Bamboo caterpillar The bamboo caterpillar belongs to the family Crambidae and lives in bamboo groves in Thailand, Myanmar, Yunnan Province in China, and northern Laos. Omphisa fuscidentalis, which feeds on bamboo plants, belongs to Dendrocalamus sp. and Thyrsostachys sp. The caterpillars live and feed inside the bamboo shoot and are removed by the farmers by cutting the clumps of bamboo. Removing the caterpillars from the infested internodes is a recent farming method. This method consists of cutting a rectangular hole (9 × 13 cm), preferably from the upper part of the young shoot. The caterpillars congregate in internodes after 45–60 days and remain there for 8 months. Bamboo shoots are covered with nylon net cages to allow the natural mating process to increase production. According to an estimate, 1 kg caterpillar packaged in plastic boxes sells on the market for US $6–8. Some private food companies supply cooked caterpillar in sealed containers which can be used or eaten after being heated in a microwave oven (Hanboonsong, Jamjanya, and Durst 2013).

6.3.3.4 Weaver ants Oecophylla smaragdina Fabricius (weaver ant) is famous as an efficient biocontrol agent because it preys on various insect pests. These ants have been consumed as food in Southeast Asia for centuries, as well as being used as bird feed and in conventional medicine. The ant colonies are reared on mango trees by constructing nests of intertwined leaves with the help of larval silk. All trees with the same insect

6  •  Edible Insect Farming  123 colonies are connected with the help of a nylon string of 7 mm in diameter. Cat food plus 35% sucrose solution is used to feed the ants 16 times between the first feeding (Offenberg and Wiwatwitaya 2010). Ants are harvested once a year via a bag or basket attached to one end of a long bamboo pole with strings. The nest is shaken after a hole is poked into the tip and both larvae and pupae fall into the bag. The contents of the bag are collected in a plastic container supplied with tapioca from cassava and rice flour to stop ants escaping. Adult ants can climb up from the container via a branch or stick inserted into it. In home gardens, ant farming is also carried out using water and food scraps. An average of 300–400 grams of larvae and pupae are collected per day, and a collector can earn US $8–15 from this amount. An increase in market demand decreases the population of ants in the wild, which is a major constraint of overfarming (Hanboonsong, Jamjanya, and Durst 2013). An illustration of the ant farming system is shown in Figure 6.6 B.

6.4  COST OF CULTIVATION The ever-growing human population puts pressure on available resources, and it has become the primary cause of starvation. The hunger monster can be controlled by using the insect as the savior. Insect production is promising to fill the protein deficit that can be foreseen (Henchion et al. 2017) because insect-based proteins seem the best alternative to plant-based or conventional meat sources. Approximately 2,000 integrated insect farms are registered that produce 7,500 tons of red ants, grasshoppers and crickets per year. An estimated $5,000 can thus be added annually to people’s income, which is double that of crop cultivation (Gray 2014). For insect farming start-ups, the perception of economic figures and profit gain are necessary ingredients. For the successful production of insects, a species-specific approach should be used based on place, environment, habitat, and the seasonal availability of insects (Mancuso, Pippinato, and Gasco 2019). The insect industry has been growing in Europe and America and is famous for producing T. molitor, H. illucens, and A. domesticus (Dossey, Tatum, and McGill 2016). Besides this, many cottage industries are also run by small entrepreneurs in developing countries for financial support (Macombe et al. 2019). According to a survey by the IPIFF (International Platform for Food and Feed), each year 6,000 tons of insects are produced for food and feed, and this is projected to rise to up to 5 million tons by 2030. Production cost is always calculated based on harvesting cycles per year or per season, the cost of breeding nurseries and cages, and feeding materials. Economic figures for the cultivation of different insect species such as crickets, yellow mealworms, and black fly have been analyzed and reported (Niyonsaba et al. 2021). H. illucens production is mainly done using biodegraded materials and requires no special infrastructure or labor for rearing. However, the calculated figure for the production of black flies on the basis of feed, energy, and labor ranges from $903 to $17,632 per ton of dried larvae. Similarly, the cultivation cost and feed margins of A. domesticus were $8,625–$15,587.64 per ton of fresh larvae (Niyonsaba et al. 2021). The cost per harvesting cycle was 4,682 TBH (Thai baht) or $139.02. Besides harvesting adult crickets, two other valuable products (cricket eggs and fertilizers from waste) can be obtained from cricket farms. An estimated production cost for palm weevil larvae per harvesting cycle is 5,910 TBH or $175. The summarized data evaluates that the production cost for insects is far lower than that for conventional livestock. Insect farmers will earn more with little investment and enhance their livelihood. Based on growing insect farming potential globally, the insect market will reach $1.4 billion by 2024 (Research and Markets 2019). Information on edible insect farming methods, edible species, life spans, seasonal availability, feeding requirements, developmental stages at which insects are consumed, and rearing cages and applications is given in Table 6.2.

45–55

30–45

40–60

13–15

3–6

Bombyx mori

Acheta domesticus

Tenebrio molitor

Hermetia illucens

Musca domestica

House cricket

Yellow meal worm

Black soldier fly

House fly

SPECIES

LIFE SPAN (DAYS)

Silkworm

COMMON NAME

Indoor

Indoor

Indoor

Indoor, wild harvesting

Indoor

FARMING METHOD

Whole year

-

Whole year

Whole year

Whole year

SEASONAL AVAILABILITY

Chicken protein, vegetables, watermelon, and cassava leaves Spent grains, beer yeast, bread and biscuit remains, potato steam peelings, dried grains, maize distillers Fruit waste and chicken manure, compost tea, catering waste, food scraps Fresh dairy cattle manure, maize bran

Mulberry leaves, crushed or chopped

FEED

Larvae, pupae

Larvae

Larvae

Adult

Larvae

DEVELOPMENTAL STAGE

Fly cages

Circular plastic container

Nylon mesh or permanent houses in cement Concrete cylinders, plywood boxes, plastic drawers Wood and plastic containers

REARING BEDS

Human and animal food

Zoo, fish, livestock, and pet animal feed, human food

Protein extraction, food for humans and pets

Animal feed, human food

APPLICATION

(Continued)

Hussein et al. 2017

Dzepe et al. 2021

Houben et al. 2020

Lundy and Parrella 2015

Van Huis 2013; Pal and Roy 2014

REFERENCE

TABLE 6.2  Summary of edible insect farming methods, edible species, life spans, seasonal availability, feeding requirements, developmental stages at which insects are consumed, and rearing cages and applications

124  Edible Insects Processing for Food and Feed

40–45

45–60

Rhynchophorus ferrugineus

Omphisa fuscidentalis

Oecophylla smaragdina

Palm Weevil

Bamboo caterpillar

Weaver ants

365

Oxya spp.

SPECIES

LIFE SPAN (DAYS)

Grasshopper

COMMON NAME

Semicultivation, wild harvesting

Semicultivation

Semicultivation, indoor farming

Semicultivation, wild harvesting

FARMING METHOD

-

Whole year

One generation per year

SEASONAL AVAILABILITY

Cat food and 30% sucrose solution

Fresh inner pulp of bamboo

Trunk of sago palm trees, pig feed

Maize and rice paddy sap, small insects

FEED

Adult, larvae, pupae, eggs

Larvae

Larvae, pupae

Adult

DEVELOPMENTAL STAGE

Nets or mango plant

Internodes of bamboo plants

Trunk of Sago palm trees, plastic trays

Rice paddy fields

REARING BEDS

Medicinal use and human food

Human food

Nutrient recycling, protein source, human food Protein source for humans

APPLICATION

Hanboonsong, Jamjanya, and Durst 2013; Chinarak, Chaijan, and Panpipat 2020 Pal and Roy 2014; Hanboonsong, Jamjanya, and Durst 2013 Itterbeeck 2014

Pemberton 1994

REFERENCE

TABLE 6.2 (CONTINUED)  Summary of edible insect farming methods, edible species, life spans, seasonal availability, feeding requirements, developmental stages at which insects are consumed, and rearing cages and applications

6  •  Edible Insect Farming  125

126  Edible Insects Processing for Food and Feed

6.5  CHALLENGES AND WAY FORWARD In the wild, insects are most vulnerable to habitat loss and natural enemies. The insect species in the wild can be conserved by habitat management, biodiversity protection with the help of local and practical knowledge (Mbata, Chidumayo, and Lwatula 2002). Sustainable and post-harvest managements ensure the safety and benefits of wild harvesting (Rumpold and Schlüter 2013b, 2013a). Consumer acceptance is a major challenge in insect farming. Usually, people of modern civilizations do not accept insects as food due to some disgusting factors associated with them. Public awareness is helpful to change the negative attitudes of people by using electronic media and organizing shows, exhibitions, and street plays about insect-based diets. Improvements in marketing strategies, modification of harvesting, processing, and packing techniques will also help draw the attention of consumers to insects (Offenberg 2011). Insect farming is only conducted for a few species despite the huge biodiversity of insects. To initiate the farming of new species, one should know about the ecology, habitat, environment, and dietary constituents of edible insect species. A complete knowledge of nutritional requirements for rearing helps improve growth and production efficacy by preparing low-cost insect feed. The cottage industry must be adapted to a commercial scale. For this purpose, smart, efficient, and circular business models should be developed. Technical guidance and government assistance bring huge change to promote insect farming at a commercial scale. This step should be implemented urgently because most banks and funding agencies have no awareness about the business patterns of insect farming and potential income to be gained. Some agencies like FAO, UNDP, and IFAD work to promote insect farming and eating as a potential, alternative source of proteins and other essential nutrients. To increase the number of entrepreneurs, efficient business strategies and proper atomization should be developed for mass rearing (Van Der Werf and Petit 2002). There is a crucial need to improve the legislation on livestock farming and introduction of new products in the market. No product should be introduced to the market without screening. The evaluation process should consist of testing, confirmation, nutritional assessments, storage materials, and food chain. Good hygienic conditions are obligatory in the processing of insect species because insects may carry the pathogens of different bacterial and viral diseases (Tanada and Kaya 2012). To prevent or reduce exposure to contamination, a healthy insect diet should conform to described standards (Mbata, Chidumayo, and Lwatula 2002). Finally, insects can serve as a sustainable and easy to approach option to fight the challenge of food security. There is a need to introduce cost-effective production and management practices. In the future, the insect industry can be projected across the globe by promoting international trade and excellent business models. By the inclusion of more professional and government agencies, gross domestic production can be increased.

ACKNOWLEDGMENT The authors wish to express their gratitude to Hanboonsong et al. 2013, Gahukar 2016, Naseem et al. 2021 and the copyright permission agencies to grant permission for figures material reuse/reprint. We also acknowledge the reviewers and well-wishers who always help in approaching scientific research.

REFERENCES Amala Rani, G., C. Padmalatha, R. Soma Raj, and A. J. A. Ranjith Singh. 2011. “Impact of Supplementation of Amway Protein on the Economic Characters and Energy Budget of Silkworm Bombyx mori L.” Asian Journal of Animal Sciences 5 (3). Knowledgia Review: 190–95. doi:10.3923/ajas.2011.190.195.

6  •  Edible Insect Farming  127 Areerat, Sathita, Pipatpong Chundang, Chalermpol Lekcharoensuk, and Attawit Kovitvadhi. 2021. “Possibility of Using House Cricket (Acheta domesticus) or Mulberry Silkworm (Bombyx mori) Pupae Meal to Replace Poultry Meal in Canine Diets Based on Health and Nutrient Digestibility.” Animals 11 (9). Multidisciplinary Digital Publishing Institute: 2680. doi:10.3390/ani11092680. Ayieko, M. A., V. Oriamo, and I. A. Nyambuga. 2010. Processed Products of Termites and Lake Flies: Improving Entomophagy for Food Security within the Lake Victoria Region. African Journal of Food, Agriculture, Nutrition and Development 10 (2): 2085–2098. Bélanger, Julie, and Dafydd Pilling. 2019. The State of the World’s Biodiversity for Food and Agriculture. The State of the World’s Biodiversity for Food and Agriculture. Food and Agriculture Organization of the United Nations (FAO). doi:10.4060/ca3129en. Berggren, A. Jansson, and M. Low. 2018. “Using Current Systems to Inform Rearing Facility Design in the Insectas-Food Industry.” Journal of Insects as Food and Feed 4 (3). Wageningen Academic Publishers: 167–70. doi:10.3920/JIFF2017.0076. Bhatia, Narendra Kumar, and Mohammad Yousuf. 2013. “Forest Insect Industry in Collaborative Forest Management: An Overview.” International Journal of Industrial Entomology 27 (1). Korean Society of Sericultural Science: 166–79. doi:10.7852/ijie.2013.27.1.166. Bhede, B. V., U. L. Lande, and D. T. Pathrikar. 2013. “Evaluation of Different Mulberry Varieties for Commercial Rearing of Silkworm Hybrid in Maharashtra.” Journal of Entomological Research 37 (1). Malhotra Publishing House: 17–24. Blásquez, Julieta Ramos-Elorduy, José Manuel Pino Moreno, and Víctor Hugo Martínez Camacho. 2012. “Could Grasshoppers Be a Nutritive Meal?” Food and Nutrition Sciences 03 (02). Citeseer: 164–75. doi:10.4236/ fns.2012.32025. Bordiean, Anna, Michał Krzyżaniak, Mariusz J. Stolarski, Stanisław Czachorowski, and Dumitru Peni. 2020. “Will Yellow Mealworm Become a Source of Safe Proteins for Europe?” Agriculture (Switzerland) 10 (6). Multidisciplinary Digital Publishing Institute: 1–30. doi:10.3390/agriculture10060233. Broekhoven, Sarah Van, Dennis G. A. B. Oonincx, Arnold van Huis, and Joop J. A. van Loon. 2015. “Growth Performance and Feed Conversion Efficiency of Three Edible Mealworm Species (Coleoptera: Tenebrionidae) on Diets Composed of Organic by-Products.” Journal of Insect Physiology 73. Elsevier: 1–10. doi:10.1016/j. jinsphys.2014.12.005. Cadinu, Lorenzo A., Paolo Barra, Francesco Torre, Francesco Delogu, and Fabio A. Madau. 2020. “Insect Rearing: Potential, Challenges, and Circularity.” Sustainability 12 (11). Multidisciplinary Digital Publishing Institute: 4567. Capinera, John L. 2005. Encyclopedia of Entomology. Choice Reviews Online. Vol. 42. Springer Science & Business Media. doi:10.5860/choice.42-4375. Cerritos, René. 2009. “Insects as Food: An Ecological, Social and Economical Approach.” CAB Reviews: Perspectives in Agriculture, Veterinary Science, Nutrition and Natural Resources 4 (027). CABI: 1–10. doi:10.1079/PAVSNNR20094027. Chapagain, A. K., and A. Y. Hoekstra. 2003. Virtual Water Trade: A Quantification of Virtual Water Flows between Nations in Relation to International Trade in Livestock and Livestock Products. Virtual Water Trade: Proceedings of the International Expert Meeting on Virtual Water Trade. UNESCO-IHE Delft. Charlton, A. J., M. Dickinson, M. E. Wakefield, E. Fitches, M. Kenis, R. Han, F. Zhu, et al. 2015. “Exploring the Chemical Safety of Fly Larvae as a Source of Protein for Animal Feed.” Journal of Insects as Food and Feed 1 (1). Wageningen Academic Publishers: 7–16. doi:10.3920/JIFF2014.0020. Chen, Xiaoming, Ying Feng, and Zhiyong Chen. 2009. “Common Edible Insects and Their Utilization in China: INVITED REVIEW.” Entomological Research 39 (5). Wiley Online Library: 299–303. doi:10.1111/j.1748-5967.2009.00237.x. Chidumayo, E. N., and K. J. Mbata. 2002. “Shifting Cultivation, Edible Caterpillars and Livelihoods in the Kopa Area of Northern Zambia.” Forests Trees and Livelihoods 12 (3). Taylor & Francis: 175–93. doi:10.1080/1472 8028.2002.9752423. Chinarak, Khanittha, Manat Chaijan, and Worawan Panpipat. 2020. “Farm-Raised Sago Palm Weevil (Rhynchophorus ferrugineus) Larvae: Potential and Challenges for Promising Source of Nutrients.” Journal of Food Composition and Analysis 92. Elsevier: 103542. doi:10.1016/j.jfca.2020.103542. Collavo, Alberto, Robert H. Glew, Yung-Sheng Huang, Lu-Te Chuang, Rebecca Bosse, and Maurizio G Paoletti. 2020. “House Cricket Small-Scale Farming.” Ecological Implications of Minilivestock 27. New Hampshire Science Publishers: 533–58. doi:10.1201/9781482294439-33. Curran, Mary Ann. 2016. “Life Cycle Assessment in the Agri-Food Sector: Case Studies, Methodological Issues, and Best Practices.” The International Journal of Life Cycle Assessment 21. Springer. doi:10.1007/ s11367-015-0977-5.

128  Edible Insects Processing for Food and Feed DeFoliart, G. R. 1995. “Edible Insects as Minilivestock.” Biodiversity and Conservation 4 (3). Springer: 306–21. doi:10.1007/BF00055976. DeFoliart, G. R. 1997. “An Overview of the Role of Edible Insects in Preserving Biodiversity.” Ecology of Food Nutrition 36 (2–4). Taylor & Francis: 109–32. doi:10.1080/03670244.1997.9991510. Diener, Stefan, Nandayure M. Studt Solano, Floria Roa Gutiérrez, Christian Zurbrügg, and Klement Tockner. 2011. “Biological Treatment of Municipal Organic Waste Using Black Soldier Fly Larvae.” Waste and Biomass Valorization 2 (4). Springer: 357–63. doi:10.1007/s12649-011-9079-1. Doreau, Michel, Michael S. Corson, and Stephen G. Wiedemann. 2012. “Water Use by Livestock: A Global Perspective for a Regional Issue?” Animal Frontiers 2 (2). Oxford University Press: 9–16. doi:10.2527/af.2012-0036. Dossey, A. T., J. T. Tatum, and W. L. McGill. 2016. “Modern Insect-Based Food Industry: Current Status, Insect Processing Technology, and Recommendations Moving Forward.” In Insects as Sustainable Food Ingredients, 113–52. Elsevier. doi:10.1016/b978-0-12-802856-8.00005-3. Dzepe, Daniel, Paulin Nana, Hervé Mube Kuietche, Janaina Mitsue Kimpara, Ornela Magatsing, Timoléon Tchuinkam, and Rousseau Djouaka. 2021. “Feeding Strategies for Small-Scale Rearing Black Soldier Fly Larvae (Hermetia illucens) as Organic Waste Recycler.” SN Applied Sciences 3 (2). Springer: 1–9. doi:10.1007/ s42452-020-04039-5. FAO. 2013a. Edible Insects. Future Prospects for Food and Feed Security. Food and Agriculture Organization of the United Nations. Vol. 171. Food and Agriculture Organization of the United Nations. FAO. 2013b. Food Wastage Footprint. FAO. Farina, L., F. Demey, and J. Hardouin. 1991. “Production de Termites Pour l’aviculture Villageoise Au Togo.” Tropicultura 9 (4): 181–87. Feng, Ying, Min Zhao, Zhao He, Zhiyong Chen, and Long Sun. 2009. “Research and Utilization of Medicinal Insects in China.” Entomological Research 39 (5). Wiley Online Library: 313–16. doi:10.1111/j.1748-5967.2009.00236.x. Fernandez-Cassi, X., A. Supeanu, M. Vaga, A. Jansson, S. Boqvist, and I. Vagsholm. 2019. “The House Cricket (Acheta domesticus) as a Novel Food: A Risk Profile.” Journal of Insects as Food and Feed 5 (2). Wageningen Academic Publishers: 137–57. doi:10.3920/JIFF2018.0021. Fiala, Nathan. 2008. “Meeting the Demand: An Estimation of Potential Future Greenhouse Gas Emissions from Meat Production.” Ecological Economics 67 (3). Elsevier: 412–19. doi:10.1016/j.ecolecon.2007.12.021. Finke, Mark D., and Dennis Oonincx. 2013. “Insects as Food for Insectivores.” In Mass Production of Beneficial Organisms: Invertebrates and Entomopathogens, 583–616. Elsevier. doi:10.1016/B978-0-12-391453-8.00017-0. Gahukar, R. T. 2011a. “Entomophagy and Human Food Security.” International Journal of Tropical Insect Science 31 (3). Cambridge University Press: 129–44. doi:10.1017/S1742758411000257. Gahukar, R. T. 2011b. “Food Security in India: The Challenge of Food Production and Distribution.” Journal of Agricultural and Food Information 12 (3–4). Taylor & Francis: 270–86. doi:10.1080/10496505.2011.588932. Gahukar, R. T. 2014. “Impact of Major Biotic Factors on Tropical Silkworm Rearing in India and Monitoring of Unfavourable Elements: A Review.” Sericologia, 150. Gahukar, R. T. 2016. “Edible Insects Farming: Efficiency and Impact on Family Livelihood, Food Security, and Environment Compared with Livestock and Crops.” In Insects as Sustainable Food Ingredients, 85–111. Elsevier. doi:10.1016/b978-0-12-802856-8.00004-1. Gerda, H., R. Martinez, N. Briceno, L. Pizzoferrato, P. Manzi, M. Tommaseo Ponzetta, O. Marin, and M. G. Paoletti. 2001. “Palm Worm: (Rhynchophorus palmarum) Traditional Food in Amazonas, Venezuela - Nutritional Composition, Small Scale Production and Tourist Palatability.” Ecology of Food and Nutrition 40 (1). Taylor & Francis: 13–32. doi:10.1080/03670244.2001.9991635. Gray, Denis D. 2014. “Edible Insects a Boon to Thailand’s Farmers.” pp. 1–11. https://phys.org/news/2014-08-edibleinsects-boon-thailand-farmers.html. Greenberg, B. 1973. Flies and Disease. Vol II. Biology and Disease Transmission. Princeton University Press. Halloran, A., Y. Hanboonsong, N. Roos, and S. Bruun. 2017. “Life Cycle Assessment of Cricket Farming in NorthEastern Thailand.” Journal of Cleaner Production 156. Elsevier: 83–94. doi:10.1016/j.jclepro.2017.04.017. Halloran, A., N. Roos, R. Flore, and Y. Hanboonsong. 2016. “The Development of the Edible Cricket Industry in Thailand.” Journal of Insects as Food and Feed. doi:10.3920/JIFF2015.0091. Halloran, Afton, and Christopher Münke. 2014. Discussion Paper: Regulatory Frameworks Influencing Insects as Food and Feed. FAO, 1–39. Hanboonsong, Yupa, Tasanee Jamjanya, and Patrick B Durst. 2013. “Six-Legged Livestock: Edible Insect Farming, Collec on and Market in Thailand. Food and Agriculture Organization of the United Nations Regional Office for Asia and the Pacific. Bangkok, 2013.” Rap Publication 3: 69. Henchion, Maeve, Maria Hayes, Anne Maria Mullen, Mark Fenelon, and Brijesh Tiwari. 2017. “Future Protein Supply and Demand: Strategies and Factors Influencing a Sustainable Equilibrium.” Foods 6 (7). Multidisciplinary Digital Publishing Institute: 53.

6  •  Edible Insect Farming  129 Hite, Kristen A, and John L Seitz. 2021. Global Issues: An Introduction. John Wiley & Sons. Hoornweg, D., and P. Bhada-Tata. 2014. “What a Waste : A Global Review of Solid Waste Management Cb.” World Bank 15 (November). World Bank: 1–2. Houben, David, Guillaume Daoulas, Michel Pierre Faucon, and Anne Maïmiti Dulaurent. 2020. “Potential Use of Mealworm Frass as a Fertilizer: Impact on Crop Growth and Soil Properties.” Scientific Reports 10 (1). Nature Publishing Group: 1–9. doi:10.1038/s41598-020-61765-x. Huis, Arnold Van. 2013. “Potential of Insects as Food and Feed in Assuring Food Security.” Annual Review of Entomology 58. Annual Reviews: 563–83. doi:10.1146/annurev-ento-120811-153704. Huis, Arnold Van. 2015. “Edible Insects Contributing to Food Security?” Agriculture and Food Security 4 (1). BioMed Central: 1–9. doi:10.1186/s40066-015-0041-5. Huis, Arnold Van, and Dennis G. A. B. Oonincx. 2017. “The Environmental Sustainability of Insects as Food and Feed. A Review.” Agronomy for Sustainable Development 37 (5). Springer: 1–14. doi:10.1007/ s13593-017-0452-8. Hussein, Mahmoud, Viju V. Pillai, Joshua M. Goddard, Hui G. Park, Kumar S. Kothapalli, Deborah A. Ross, Quirine M. Ketterings, et al. 2017. “Sustainable Production of Housefly (Musca domestica) Larvae as a Protein-Rich Feed Ingredient by Utilizing Cattle Manure.” PLOS ONE 12 (2). Public Library of Science San Francisco, CA: e0171708. doi:10.1371/journal.pone.0171708. Itterbeeck, Joost Van. Prospects of semi-cultivating the edible weaver ant Oecophylla smaragdina. PhD thesis, Wageningen University, Wageningen, NL. pp. 1–111. Jongema, Y. 2017. List of Edible Insect Species of the World. Laboratory of Entomology, Wageningen University. Kenis, M., N. Koné, C. A. A. M. Chrysostome, E. Devic, G. K. D. Koko, V. A. Clottey, S. Nacambo, and G. A. Mensah. 2014. “Insects Used for Animal Feed in West Africa.” Entomologia 2 (October): 107–14. doi:10.4081/ entomologia.2014.218. Kim, Sun Am, Kyung Moon Kim, and Boung Jun Oh. 2008. “Current Status and Perspective of the Insect Industry in Korea.” Entomological Research 38 (SUPPL. 1). Wiley Online Library: S79–85. doi:10.1111/j.1748-5967.2008.00178.x. Kouřimská, Lenka, and Anna Adámková. 2016. “Nutritional and Sensory Quality of Edible Insects.” NFS Journal 4. Elsevier: 22–26. doi:10.1016/j.nfs.2016.07.001. Kulma, M., V. Tumovå, A. Fialovå, and L. Koufimskå. 2020. “Insect Consumption in the Czech Republic: What the Eye Does Not See, the Heart Does Not Grieve Over.” Journal of Insects as Food and Feed 6 (5). Wageningen Academic Publishers: 525–35. doi:10.3920/JIFF2020.0020. Li, Qing, Longyu Zheng, Hao Cai, E. Garza, Ziniu Yu, and Shengde Zhou. 2011. “From Organic Waste to Biodiesel: Black Soldier Fly, Hermetia Illucens, Makes It Feasible.” Fuel 90 (4). Elsevier: 1545–48. doi:10.1016/j. fuel.2010.11.016. Looy, Heather, Florence V. Dunkel, and John R. Wood. 2014. “How Then Shall We Eat? Insect-Eating Attitudes and Sustainable Foodways.” Agriculture and Human Values 31 (1). Springer: 131–41. doi:10.1007/ s10460-013-9450-x. Losey, John E., and Mace Vaughan. 2006. “The Economic Value of Ecological Services Provided by Insects.” BioScience 56 (4). American Institute of Biological Sciences: 311–23. doi:1​0.164​1/000​6 -356​8(200​6)56[​311:T​ EVOES​]2.0.​CO;2.​ Lundy, Mark E., and Michael P. Parrella. 2015. “Crickets Are Not a Free Lunch: Protein Capture from Scalable Organic Side-Streams via High-Density Populations of Acheta domesticus.” PLOS ONE 10 (4). Public Library of Science: e0118785. doi:10.1371/journal.pone.0118785. Macombe, C., S. Le Feon, J. Aubin, and F. Maillard. 2019. “Marketing and Social Effects of Industrial Scale Insect Value Chains in Europe: Case of Mealworm for Feed in France.” Journal of Insects as Food and Feed 5 (3). Wageningen Academic Publishers: 215–24. doi:10.3920/JIFF2018.0047. Makkar, Harinder P. S., Gilles Tran, Valérie Heuzé, and Philippe Ankers. 2014. “State-of-the-Art on Use of Insects as Animal Feed.” Animal Feed Science and Technology 197. Elsevier: 1–33. doi:10.1016/j. anifeedsci.2014.07.008. Malik, Firdose Ahmad, and Y Srinivasa Reddy. 2007. “Role of Mulberry Nutrition on Manifestation of Post Cocoon Characters of Selected Races of the Silkworm, Bombyx mori L.” Sericologia 47 (1): 63–86. Mancuso, Teresina, Liam Pippinato, and Laura Gasco. 2019. “The European Insects Sector and Its Role in the Provision of Green Proteins in Feed Supply.” Quality - Access to Success 20 (S2). Romanian Society for Quality Assurance: 374–81. Marco, M. De, S. Martínez, F. Hernandez, J. Madrid, F. Gai, L. Rotolo, M. Belforti, et al. 2015. “Nutritional Value of Two Insect Larval Meals (Tenebrio molitor and Hermetia illucens) for Broiler Chickens: Apparent Nutrient Digestibility, Apparent Ileal Amino Acid Digestibility and Apparent Metabolizable Energy.” Animal Feed Science and Technology 209. Elsevier: 211–18. doi:10.1016/j.anifeedsci.2015.08.006.

130  Edible Insects Processing for Food and Feed Marone, P. A. 2016. “Food Safety and Regulatory Concerns.” In Insects as Sustainable Food Ingredients, 203–21. Elsevier. doi:10.1016/b978-0-12-802856-8.00007-7. Mbata, K. J., E. N. Chidumayo, and C. M. Lwatula. 2002. “Traditional Regulation of Edible Caterpillar Exploitation in the Kopa Area of Mpika District in Northern Zambia.” Journal of Insect Conservation 6 (2). Springer: 115–30. doi:10.1023/A:1020953030648. Miech, P., Å. Berggren, J. E. Lindberg, T. Chhay, B. Khieu, and A. Jansson. 2016. “Growth and Survival of Reared Cambodian Field Crickets (Teleogryllus testaceus) Fed Weeds, Agricultural and Food Industry by-Products.” Journal of Insects as Food and Feed 2 (4). Wageningen Academic Publishers: 285–92. doi:10.3920/ JIFF2016.0028. Miech, P., J. E. Lindberg, Å. Berggren, T. Chhay, and A. Jansson. 2017. “Apparent Faecal Digestibility and Nitrogen Retention in Piglets Fed Whole and Peeled Cambodian Field Cricket Meal.” Journal of Insects as Food and Feed 3 (4). Wageningen Academic Publishers: 279–88. doi:10.3920/JIFF2017.0019. Miglietta, Pier Paolo, Federica De Leo, Marcello Ruberti, and Stefania Massari. 2015. “Mealworms for Food: A Water Footprint Perspective.” Water (Switzerland) 7 (11). MDPI: 6190–6203. doi:10.3390/w7116190. Morales-Ramos, Juan A., and M. Guadalupe Rojas. 2015. “Effect of Larval Density on Food Utilization Efficiency of Tenebrio molitor (Coleoptera: Tenebrionidae).” Journal of Economic Entomology 108 (5). Oxford University Press: 2259–67. doi:10.1093/jee/tov208. Morales-Ramos, Juan A., M. Guadalupe Rojas, and David I. Shapiro-Ilan. 2014. Mass Production of Beneficial Organisms: Invertebrates and Entomopathogens. Mass Production of Beneficial Organisms: Invertebrates and Entomopathogens. Academic Press. Morales-Ramos, Juan A., M. Guadalupe Rojas, David I. Shapiro-Llan, and W. Louis Tedders. 2013. “Use of Nutrient Self-Selection as a Diet Refining Tool in Tenebrio molitor (Coleoptera: Tenebrionidae).” Journal of Entomological Science 48 (3). Georgia Entomological Society, Inc.: 206–21. doi:10.18474/0749-8004-48.3.206. Nadeau, Luc, Isaac Nadeau, Frank Franklin, and Florence Dunkel. 2015. “The Potential for Entomophagy to Address Undernutrition.” Ecology of Food and Nutrition 54 (3). Taylor & Francis: 200–208. doi:10.1080/03670244.20 14.930032. Nakagaki, Barbara J., and Gene R. Defoliart. 1991. “Comparison of Diets for Mass-Rearing Acheta domesticus (Orthoptera: Gryllidae) as a Novelty Food, and Comparison of Food Conversion Efficiency with Values Reported for Livestock.” Journal of Economic Entomology 84 (3). Oxford University Press: 891–96. doi:10.1093/jee/84.3.891. Naseem, Rimsha, Waqar Majeed, Naureen Rana, Elmo Borges de Azevedo Koch, and Muhammad Rehan Naseem. 2021. “Entomophagy: An Innovative Nutritional and Economic Navigational Tool in Race of Food Security.” International Journal of Tropical Insect Science 41. Springer: 2211–21. National Research Council. 2011. “Guide for the Care and Use of Laboratory Animals - Committee for the Update of the Guide for the Care and Use of Laboratory Animals, Institute for Laboratory Animal Research.” Guide for the Care and Use of Laboratory Animals 327 (3): 220. Niyonsaba, H. H., J. Hohler, J. Kooistra, H. J. van der Fels-Klerx, and M. P. M. Meuwissen. 2021. “Profitability of Insect Farms.” Journal of Insects as Food and Feed 7 (5). Wageningen Academic Publishers: 923–34. doi:10.3920/JIFF2020.0087. Nonaka, Kenichi. 2009. “Feasting on Insects.” Entomological Research 39 (5). Wiley Online Library: 304–12. doi:10.1111/j.1748-5967.2009.00240.x. Nyakeri, E. M., M. A. Ayieko, F. A. Amimo, H. Salum, and H. J. O. Ogola. 2019. “An Optimal Feeding Strategy for Black Soldier Fly Larvae Biomass Production and Faecal Sludge Reduction.” Journal of Insects as Food and Feed 5 (3). Wageningen Academic Publishers: 201–13. doi:10.3920/JIFF2018.0017. Offenberg, Joachim. 2011. “Oecophylla Smaragdina Food Conversion Efficiency: Prospects for Ant Farming.” Journal of Applied Entomology 135 (8). Wiley Online Library: 575–81. doi:10.1111/j.1439-0418.2010.01588.x. Offenberg, Joachim, and Decha Wiwatwitaya. 2010. “Sustainable Weaver Ant (Oecophylla smaragdina) Farming: Harvest Yields and Effects on Worker Ant Density.” Asian Myrmecology 3 (1): 55–62. Oonincx, Dennis G. A. B., and Imke J. M. de Boer. 2012. “Environmental Impact of the Production of Mealworms as a Protein Source for Humans – A Life Cycle Assessment.” PLoS ONE 7 (12). Public Library of Science San Francisco: e51145. doi:10.1371/journal.pone.0051145. Oonincx, Dennis G. A. B., Joost van Itterbeeck, Marcel J. W. Heetkamp, Henry van den Brand, Joop J. A. van Loon, and Arnold van Huis. 2010. “An Exploration on Greenhouse Gas and Ammonia Production by Insect Species Suitable for Animal or Human Consumption.” PLoS ONE 5 (12). Public Library of Science San Francisco: e14445. doi:10.1371/journal.pone.0014445. Oonincx, Dennis G A B, Sarah Van Broekhoven, Arnold Van Huis, and Joop J A van Loon. 2015. “Feed Conversion, Survival and Development, and Composition of Four Insect Species on Diets Composed of Food by-Products.” PLOS ONE 10 (12). Public Library of Science: e0144601.

6  •  Edible Insect Farming  131 Pal, Partha, and Spandita Roy. 2014. “Edible Insects: Future of Human Food–A Review.” International Letters of Natural Sciences 21: 1–11. Pastell, Helena, Sandra Mellberg, Tiina Ritvanen, Marja Raatikainen, Satu Mykkänen, Jarkko Niemi, Ilkka Latomäki, and Gun Wirtanen. 2021. “ How Does Locally Produced Feed Affect the Chemical Composition of Reared House Crickets (Acheta domesticus)?” ACS Food Science & Technology 1 (4). ACS Publications: 625–35. doi:10.1021/acsfoodscitech.0c00083. Paul, Aman, Michel Frederich, Rudy Caparros Megido, Taofic Alabi, Priyanka Malik, Roel Uyttenbroeck, Frederic Francis, et al. 2017. “Insect Fatty Acids: A Comparison of Lipids from Three Orthopterans and Tenebrio molitor L. Larvae.” Journal of Asia-Pacific Entomology 20 (2). Elsevier: 337–40. doi:10.1016/j. aspen.2017.02.001. Paul, Aman, Michel Frederich, Roel Uyttenbroeck, Priyanka Malik, Sandrino Filocco, Aurore Richel, Stephanie Heuskin, Taofic Alabi, Rudy Caparros Megido, and Thierry Franck. 2016. “Nutritional Composition and Rearing Potential of the Meadow Grasshopper (Chorthippus parallelus Zetterstedt).” Journal of Asia-Pacific Entomology 19 (4). Elsevier: 1111–16. Payne, Charlotte L. R., and Joost Van Itterbeeck. 2017. “Ecosystem Services from Edible Insects in Agricultural Systems: A Review.” Insects 8 (1). Multidisciplinary Digital Publishing Institute: 24. doi:10.3390/ insects8010024. Pemberton, Robert. 1994. “The Revival of Rice-Filled Grasshoppers as Human Food in South Korea.” Pan-Pacific Entomologist 70 (4): 325–27. Raheem, Dele, Conrado Carrascosa, Oluwatoyin Bolanle Oluwole, Maaike Nieuwland, Ariana Saraiva, Rafael Millán, and António Raposo. 2019. “Traditional Consumption of and Rearing Edible Insects in Africa, Asia and Europe.” Critical Reviews in Food Science and Nutrition 59 (14). Taylor & Francis: 2169–88. doi:10.1080 /10408398.2018.1440191. Rahmathulla, V. K., Rufaiel S. Z. Haque, M. T. Himantharaj, G. S. Vindya, and R. K. Rajan. 2005. “Food Ingestion, Assimilation and Conversion Efficiency of Mulberry Silkworm, Bombyx mori L.” International Journal of Industrial Entomology 11 (1). Korean Society of Sericultural Science: 1–12. Ramos-Elorduy, Julieta. 2008. “Energy Supplied by Edible Insects from Mexico and Their Nutritional and Ecological Importance.” Ecology of Food and Nutrition 47 (3). Taylor & Francis: 280–97. doi:10.1080/03670240701805074. Ramos-Elorduy, Julieta. 2009. “Anthropo-Entomophagy: Cultures, Evolution and Sustainability.” Entomological Research 39 (5). Wiley Online Library: 271–88. doi:10.1111/j.1748-5967.2009.00238.x. Ramos-Elorduy, Julieta, Ernesto Avila González, Alma Rocha Hernández, and José Manuel Pino. 2002. “Use of Tenebrio molitor (Coleoptera: Tenebrionidae) to Recycle Organic Wastes and as Feed for Broiler Chickens.” Journal of Economic Entomology 95 (1). Oxford University Press: 214–20. doi:10.1603/0022-0493-95.1.214. Roffeis, M., B. Muys, J. Almeida, E. Mathijs, W. M. J. Achten, B. Pastor, Y. Velásquez, A. I. Martinez-Sanchez, and S. Rojo. 2015. “Pig Manure Treatment with Housefly (Musca domestica) Rearing-An Environmental Life Cycle Assessment.” Journal of Insects as Food and Feed 1 (3). Wageningen Academic Publishers: 195–214. doi:10.3920/JIFF2014.0021. Ronghang, R., and R. Ahmed. 2010. “Edible Insects and Their Conservation Strategy in Karbi Anglong District of Assam, North East India.” The Bioscan 2: 515–22. Roy, Poritosh, Daisuke Nei, Takahiro Orikasa, Qingyi Xu, Hiroshi Okadome, Nobutaka Nakamura, and Takeo Shiina. 2009. “A Review of Life Cycle Assessment (LCA) on Some Food Products.” Journal of Food Engineering 90 (1). Elsevier: 1–10. doi:10.1016/j.jfoodeng.2008.06.016. Rumpold, Birgit A., and Oliver K. Schlüter. 2013a. “Nutritional Composition and Safety Aspects of Edible Insects.” Molecular Nutrition and Food Research 57 (5). Wiley Online Library: 802–23. doi:10.1002/mnfr.201200735. Rumpold, Birgit A., and Oliver K. Schlüter. 2013b. “Potential and Challenges of Insects as an Innovative Source for Food and Feed Production.” Innovative Food Science and Emerging Technologies 17. Elsevier: 1–11. doi:10.1016/j.ifset.2012.11.005. Sachs, Jeffrey D. 2016. The Age of Sustainable Development. Revista Finanzas y Politica Economica. Vol. 8. Columbia University Press. doi:1​0.147​18/re​vfina​nzpol​iteco​n.201​6.8.2​.1. Sathe, T. V., T. Nishad, Dattatray Ghodake, and Dr. Asawari Sathe. 2011. “Sucking Insect Pests and Medicinal Value of Tulsi Ocimum Sanctum L. (Lamiaceae).” Indian Journal of Applied Research 4 (3): 31–33. doi:10.15373/2249555x/mar2014/102. Shockley, Marianne, Julie Lesnik, Robert Nathan Allen, and Alicia Fonseca Muñoz. 2018. “Edible Insects and Their Uses in North America: Past, Present and Future.” In Edible Insects in Sustainable Food Systems, 55–79. Springer. doi:10.1007/978-3-319-74011-9_4. Steinfeld, Henning, Pierre Gerber, Tom Wassenaar, Vincent Castel, Mauricio Rosales, and Cees de Haan. 2006. Livestock’s Long Shadow: Environmental Issues and Options. Renewable Resources Journal. Vol. 24. Food & Agriculture Org. doi:10.1071/AN15729.

132  Edible Insects Processing for Food and Feed Stull, Valerie J., Elijah Finer, Rachel S. Bergmans, Hallie P. Febvre, Colin Longhurst, Daniel K. Manter, Jonathan A. Patz, and Tiffany L. Weir. 2018. “Impact of Edible Cricket Consumption on Gut Microbiota in Healthy Adults, a Double-Blind, Randomized Crossover Trial.” Scientific Reports 8 (1). Nature Publishing Group: 1–13. doi:10.1038/s41598-018-29032-2. Tanada, Yoshinori, and Harry K. Kaya. 2012. Insect Pathology. Academic Press. doi:10.1016/ B978-0-08-092625-4.50001-7. Tang, Chufei, Ding Yang, Huaijian Liao, Hongwu Sun, Chuanjing Liu, Lanjun Wei, and Fanfan Li. 2019. “Edible Insects as a Food Source: A Review.” Food Production, Processing and Nutrition 1 (1). Springer: 1–13. doi:10.1186/s43014-019-0008-1. Tao, Jaynie, and Yao Olive Li. 2018. “Edible Insects as a Means to Address Global Malnutrition and Food Insecurity Issues.” Food Quality and Safety 2 (1). Oxford University Press: 17–26. doi:10.1093/fqsafe/fyy001. Tayal, Mukesh K., and T. P. S. Chauhan. 2017. “Silkworm Diseases and Pests.” In Industrial Entomology, 265–89. Springer. doi:10.1007/978-981-10-3304-9_9. Varelas, Vassileios. 2019. “Food Wastes as a Potential New Source for Edible Insect Mass Production for Food and Feed: A Review.” Fermentation 5 (3). Multidisciplinary Digital Publishing Institute: 81. doi:10.3390/ fermentation5030081. Vries, M. de, and I. J. M. de Boer. 2010. “Comparing Environmental Impacts for Livestock Products: A Review of Life Cycle Assessments.” Livestock Science 128 (1–3). Elsevier B.V.: 1–11. doi:10.1016/j.livsci.2009.11.007. Vries, M. de, C. E. van Middelaar, and I. J. M. de Boer. 2015. “Comparing Environmental Impacts of Beef Production Systems: A Review of Life Cycle Assessments.” Livestock Science 178. Elsevier: 279–88. doi:10.1016/j. livsci.2015.06.020. Waltner-Toews, D. 2017. Eat the Beetles!: An Exploration Into Our Conflicted Relationship with Insects. ECW Press. Way, M. J., and K. C. Khoo. 1992. “Role of Ants in Pest Management.” Annual Review of Entomology 37 (1): 479–503. doi:10.1146/annurev.ento.37.1.479. Weissman, David B., David A. Gray, Hanh Thi Pham, and Peter Tijssen. 2012. “Billions and Billions Sold: Pet-Feeder Crickets (Orthoptera: Gryllidae), Commercial Cricket Farms, an Epizootic Densovirus, and Government Regulations Make for a Potential Disaster.” Zootaxa 3504 (3504). Magnolia Press: 67–88. doi:10.11646/ zootaxa.3504.1.3. Werf, Hayo M. G. Van Der, and Jean Petit. 2002. “Evaluation of the Environmental Impact of Agriculture at the Farm Level: A Comparison and Analysis of 12 Indicator-Based Methods.” Agriculture, Ecosystems and Environment 93 (1–3). Elsevier: 131–45. doi:10.1016/S0167-8809(01)00354-1. Werf, Hayo M. G. Van Der, and Thibault Salou. 2015. “Economic Value as a Functional Unit for Environmental Labelling of Food and Other Consumer Products.” Journal of Cleaner Production 94. Elsevier: 394–97. doi:10.1016/j.jclepro.2015.01.077. Yen, A. L. 2015. “Insects as Food and Feed in the Asia Pacific Region: Current Perspectives and Future Directions.” Journal of Insects as Food and Feed 1 (1). Wageningen Academic Publishers: 33–55. doi:10.3920/JIFF2014.0017. Zanten, Hannah H. E. Van, Herman Mollenhorst, Dennis G. A. B. Oonincx, Paul Bikker, Bastiaan G. Meerburg, and Imke J. M. De Boer. 2015. “From Environmental Nuisance to Environmental Opportunity: Housefly Larvae Convert Waste to Livestock Feed.” Journal of Cleaner Production 102. Elsevier: 362–69. doi:10.1016/j. jclepro.2015.04.106. Zhu, Feng Xiang, Wei Ping Wang, Chun Lai Hong, Ming Guang Feng, Zhi Yong Xue, Xiao Yang Chen, Yan Lai Yao, and Man Yu. 2012. “Rapid Production of Maggots as Feed Supplement and Organic Fertilizer by the Two-Stage Composting of Pig Manure.” Bioresource Technology 116. Elsevier: 485–91. doi:10.1016/j. biortech.2012.04.008.

Startups Simona Grasso and Matteo Bordiga

7

Contents 7.1 Introduction 134 7.2 Case Studies 137 7.2.1 Food-Based Startups 137 7.2.1.1 Mighty Cricket 137 7.2.1.2 Illegal Oats 139 7.2.1.3 Jiminy’s 139 7.2.2 Technology-Based Startups 140 7.2.2.1 Aspire Food Group 140 7.2.2.2 BeoBia (The Bug Factory) 141 7.2.2.3 FarmInsect 142 7.3 Insect-Focused Foodtech Startups in Europe 143 7.3.1  Ÿnsect 144 7.3.2  Ÿnsect Human Nutrition & Health (the Dutch Food Branch of Ÿnsect) 145 7.3.3 Innovafeed 146 7.3.4 Protix Biosystems 146 7.3.5 nextProtein 147 7.3.6 Nextalim 148 7.3.7 Nasekomo 149 7.3.8 Hexafly 150 7.3.9 Entocycle 151 7.3.10 Hargol FoodTech 152 7.3.11 BetaHatch 152 7.3.12 Grubbly Farm 152 7.3.13 Plento 153 7.3.14 Insectta 153 7.3.15 Protenga 153 7.3.16 Bugsolutely 153 7.3.17 Magalarva 153 7.3.18 Entobel 153 Further Reading 153

DOI: 10.1201/9781003165729-7

133

134  Edible Insects Processing for Food and Feed

7.1 INTRODUCTION Entomophagy, or the act of eating insects, is nothing new. More than 2 billion people use insects in their meals now, and humans have been eating them for thousands of years. Insect eating has evolved into a lucrative industry that is expanding quickly. By 2030, the market for edible insects could be worth more than $6 billion, which suggests that items made from insects might become commonplace at your neighborhood market. Compared to cattle, insects emit 80 times less methane since they require a lot less room, food, water, shelter, and general upkeep. Crickets need one gallon of water, but a pound of beef uses 1,850 gallons and a pound of chicken uses 500 gallons. These environmental advantages can translate into better, more lucrative economic practices: raising smaller animals requires less room and less water and feed. It is possible to fight climate change and be fiscally prudent by altering Western diets. Insects are no longer available in Western food markets; instead, animal-based protein is preferred. This has left behind a whole industry for meals made from insects, along with all of the health and environmental advantages they offer. Insects are assuming a more dominant position in the global protein market as demands from climate change increase and these issues become more significant to consumers. Small, nimble, risk-taking entrepreneurs have been paving the way while big companies like Nestlé and PepsiCo are gradually entering the market by doing research and development (R&D). While the Asia-Pacific area continues to be the largest market for edible insects, insects are increasingly gaining popularity in the food and technological scenes in Europe, especially in Eastern European and Scandinavian nations. Even though the United States has lagged behind in adopting and funding entomophagy, startups are springing up all over the place and extending their reach. In terms of value, a recent market research estimate predicts that the market for edible insects will grow at a CAGR (Compound Annual Growth Rate) of 28.3% from 2022 to 2030 and will reach $9.6 billion by that time. The market for edible insects is anticipated to grow at a CAGR of 31.1% between 2022 and 2030, reaching 3,139,035.10 tonnes. To meet the increased food needs of the population, the existing level of food production must be doubled. Finding environmentally friendly, sustainable, and nutrient-dense food sources would be necessary for this effort. Due to their high nutritional value and ability to meet the human demand for food, edible insects could be a fantastic option in this situation. The COVID-19 pandemic posed a number of difficulties for the food industry, particularly for producers of meat products around the world. The production of meat products has faced numerous difficulties, including the risk of carrying on with distribution, transportation, and other supply chain operations; a shortage of workers; and delays in product development. The expansion of the food sector has been further constrained by the rescheduling of state funding projects and private investment finance. According to estimates, these variables will have an impact on the meat products business and increase demand for alternative protein sources, such as edible insects. According to data released by the U.S. Department of Agriculture, between March and April 2020, the amount of frozen pork in storage decreased by 4%, and slaughter rates decreased by 25%. It is still too early, according to many news sources, to proclaim a food crisis. Consumers will have much fewer options for meat and protein, according to some economists. These various elements might have a global snowball impact. Insects could replace the void created by U.S. meat producers during COVID-19 because they are rich in protein, vitamins, and minerals. Insect type (crickets, mealworms, black soldier flies, buffalo worms, grasshoppers, ants, silkworms, cicadas, and other edible insects), application (food and beverages and feed), end use (human consumption and animal nutrition), and geography are the main categories used to segment the edible insect market. Additionally, this report assesses market rivals and does regional and national market analysis. In terms of product, the insect powder market is anticipated to expand at the greatest CAGR from 2022 to 2030. Some of the key factors driving the expansion of this market include rising health and

7 • Startups  135 wellness trends, the rise in the number of health clubs and fitness facilities that offer insect powder, the emergence of several startups that make insect protein bars and shakes, and busy lifestyles that call for foods that are both highly nourishing and practical, like insect powder. According to insect type, the cricket segment is anticipated to hold the greatest market share for edible insects in 2022. The high nutritional value and ease of farming and processing of crickets, their inclusion in a variety of recipes and food products, and the expanding demand for cricket-based food products such as protein powders, bars, and snacks are all factors in the growth of the cricket market. The food and beverage application is anticipated to experience the highest CAGR over the projection period of 2022 to 2030. This market is expanding quickly due to a number of factors, including the increasing demand for high-quality alternative protein and amino acid sources among end users, the increased consumption of processed whole insects as food, and the global food shortage. According to end use, the human eating sector is anticipated to hold the greatest market share for edible insects in 2022. The significant market share of this sector is attributed to three factors: the growing demand for insect-based food products to feed the expanding global population; the high nutritional value of insects in human nutrition; and the growing demand for environmentally friendly alternative sources of protein (Tables 7.1–​​7.5). Geographically speaking, it is anticipated that North America will see the most CAGR during the projection period. The growing desire for environmentally friendly, protein-rich food products in this region is the cause of the rising demand for edible insect foods. The main drivers behind the expansion of the edible insect business in North America are growing consumer familiarity with insects as food, a decline in food neophobia, and a change in views toward insects both generally and as food.

TABLE 7.1  Insect-focused foodtech startups and related nationalities (snacks and fat & oils sectors). Snacks: Combined funding $59M Chirps Chips Jimini’s Entis Plento Ento Swarm Protein Small Giants (Crické) Grubs Up Enorm Sens Exo Bear Protein Totolines Gricha Hey Planet Insnack Leap Cricket Protein

USA France Finland Singapore Malaysia Germany UK Australia Denmark UK USA Germany Mexico Mexico Denmark Germany Australia

Fat & Oils: Combined funding $0.1M BiteBack Insect Millibeter

USA Belgium

136  Edible Insects Processing for Food and Feed TABLE 7.2  Insect-focused foodtech startups and related nationalities (animal feed & aquaculture sector). Animal feed & aquaculture: Combined funding $760m Ÿnsect AgriProtein Enterra Feed FreezeM Proteinea Nasekomo Innovafeed Protix Biosystems Nutrition Technologies FlyFarm NextProtein VakSea Tebrio Insectta Inseco FarmInsect GmbH

France UK Canada Israel UK Bulgaria France Netherlands Malaysia France USA Spain Singapore South Africa Germany

TABLE 7.3  Insect-focused foodtech startups and related nationalities (meal alternatives and biotechnics & genetics sectors). Meal alternatives: Combined funding $0.5m Kreca Ento-Food Thailand Unique Edible Bug Shop Bugsolutely Bugfoundation Bold Foods Burgs Foods Kric8 Eat Grup Tinyfoods Isaac nutrition Zirp DieWurmfarm

Netherlands Thailand Australia Thailand Germany Germany Netherlands UK UK Netherlands Germany Austria Austria

Biotechnics & Genetics: Combined funding $1M Beta Bugs Nutrinsect Insect Technology Group

UK Italy UK

7 • Startups  137 TABLE 7.4  Insect-focused foodtech startups and related nationalities (organic fertilizers and pet food sectors). Organic fertilizers: Combined funding $32m Entobel Midgard Insect Farm Entofood Hexafly Beta Hatch Jord Producers Goterra

Singapore Canada Malaysia Ireland USA USA Australia

Pet Food: Combined funding $3.9M HiProMine Entoma petfood Reglo Ofrieda Onto Tomojo

Poland Denmark France Germany Russia France

Some of the key players operating in the edible insect market include are reported below as case studies.

7.2  CASE STUDIES 7.2.1 Food-Based Startups 7.2.1.1 Mighty Cricket Headquarters: St Louis, Missouri Mighty Cricket combines finely crushed crickets in basic and flavored protein powders, a range of oatmeal, pancake/waffle mixes, and chocolate bars in addition to offering appealing packaging and appetizing food selections. Its products use the Acheta domesticus species, sometimes known as the house cricket, and offer a sizable supply of protein and general nutrients without losing flavor. Mighty Cricket are made up of a group of enthusiastic, environmentally concerned foodies from all over the world. They were established in St. Louis, Missouri, and support sustainable living by promoting delicious food, valued cultures, and healthy lifestyles. Sarah Schlafly built a vision to create a clean and equitable protein supply to sustain the planet throughout her 10 years of employment in the food industry. Sarah started Mighty Cricket in 2018 after obtaining experience in corporate marketing, entrepreneurship, and accounting. Sarah has so far used the group’s great talents and ambitions to bootstrap the business. This little group has had a significant impact on the future of food in a short period of time. The most recently developed kind of protein is cricket flour, which is created by roasting and coarsely grinding insects (this is also known as cricket powder or acheta protein).

138  Edible Insects Processing for Food and Feed TABLE 7.5  Insect-focused foodtech startups and related nationalities (insect powder & flour, farming & breeding and farm management software sectors). Insect powder & flour: Combined funding $15m Nutrinsectos Cric Griyum Proti-farm Hargol FoodTech Flying spark The Cricket Bakery Hoppa Grilo Protein GaiaFood Cricket One Asia

Mexico Costa Rica Mexico Netherlands Israel Israel Australia Australia Australia Netherlands Vietnam

Farming & Breeding: Combined funding $38M EntoCube Entomo Farms Tiny Farms Protenga Micrento Entocycle Cricket Lab Nextalim Entomics

Finland Canada USA Singapore Czech Republic UK Thailand France UK

Farm Management Software: Combined funding $0.8M Cogastro Vandalsoft BugBox

Lithuania South Korea Estonia

The primary use of cricket flour in food is to replace meat because it offers a higher concentration of the same essential nutrients, including easily digestible complete protein, vitamin B12, iron, and a potent prebiotic fiber known as chitin. Additionally, it is much more environmentally friendly and sustainable. In other words, the most sustainable and complete protein source to date is cricket powder. The protein products from Mighty Cricket are created to have incredible flavor and meet the demands of consumers who care about their health and the environment. Their offerings include protein powders, chocolate bars, and high protein oats. Their entire line of products is devoid of dairy, gluten, soy, eggs, and peanuts. The products are extremely sustainable, low in sugar, non-GMO, and free of preservatives and antibiotics. Products Energy chocolate, oatmeal, protein powder, swag.

7 • Startups  139

7.2.1.2 Illegal Oats Headquarters: Denver, US By incorporating insect protein into granola and calling it "ento-granola," this company strives to make edible insects accessible for eating. The Illegal Oats brand, which mostly uses mealworms, is hitting the market in an effort to broaden consumer acceptance of other edible insects besides simply crickets. The Illegal story Clare Whetzel (founder) first heard about the competition when she learned that her university was looking for business ideas that addressed sustainability-related problems. Clare recalled a podcast she had listened to on people eating insects, where she had heard that since humans have been on the planet, people have been eating insects because they are full of nutrients like protein, good fats, calcium, and iron. It is more crucial than ever to eat sustainably with an estimated 9 billion people on the planet by 2030 and an ever-increasing demand on natural resources. Clare wanted to blend the advantages that insects have for our environment and our health into wonderful foods that we already eat. So, Illegal Oats was formed as an idea. Infused with natural mealworm powder and made with the best organic ingredients, Illegal Oats creates original granola. Ento-granola and ento-granola bites are made by Illegal Oats. Entomophagy, or eating insects, is referred to by the term "ento." Granola from Illegal Oats contains organic insect larvae powder. This protein-rich snack's ingredients are all organic and gluten-free. What’s illegal about oats? The company is named Illegal Oats because of its entirely unique ingredient: mealworms. Illegal Oats ento-granola is incredibly tasty and rich in protein and other necessary nutrients, embracing the most sustainable and healthy source of protein by including ento-powder into each recipe, thus representing the meal of the future. Products Orange-cranberry ento-granola (7 grams of protein per half cup): This granola is flavorful and full of nutrients and protein. The distinct zesty orange flavor and the sweet cranberry flavor work well together to create an unforgettable taste. Orange cranberry granola from Illegal Oats tastes great for breakfast, lunch, dinner, or a snack. Ingredients: Oats, mealworm powder, pecans, cranberries, maple syrup, coconut oil, orange zest, cinnamon, salt.

7.2.1.3 Jiminy’s Headquarters: Berkeley, California Aimed at people who can't stand the idea of eating insects themselves but still want to partake in sustainable food consumption, Jiminy's is a brand of pet food. Jiminy's, a dog treat company founded in 2016, uses insect protein in place of animal protein. According to its website, feeding cats and dogs accounts for 25% to 30% of the environmental impact of meat consumption in the United States. According to the firm, by switching to Jiminy's dog treats in 2020, dog owners helped conserve more than 218 million gallons of water and avoided the emission of 20.5 million grams of greenhouse emissions. Insects can be a wholesome, sustainable snack for dogs just like they can be for humans. They produce sustainable pet food and treats at Jiminy's utilizing cricket protein. Products from Jiminy provide pets with nutrition and encourage long-term planet stewardship. The highest quality components, carefully picked to improve the health and happiness of pets, are used to create human-grade products. The goal is to give exceptional nutrition and sustainability, not only to match other products on the market in quality.

140  Edible Insects Processing for Food and Feed Great products: Jiminy's concentrates on making foods that are both very nutritious and environmentally friendly. Cricket protein is rich in protein, contains all nine essential amino acids, and offers additional advantages such as high levels of omega acids, iron, and fiber. Impact on the environment: Jiminy's products are based on insect proteins, beginning with crickets. Insects use land and water more efficiently than traditional protein sources, require less feed, and emit fewer damaging greenhouse gases. Dog food: Jiminy's sustainable, hypoallergenic dog chow derived from insect protein is becoming increasingly popular among dog owners who care about the environment. The critical amino acids that dogs need to live healthy, active lives have been found to be present in insects, making them a highquality, digestible protein source. Treats: Dogs absolutely adore Jiminy's tasty, ecological, and hypoallergenic dog treats produced from insect protein. Protein, nutrients, vitamins, and fiber are all naturally present in this foodstuff. In actuality, it's a premium digestible protein that is brimming with the vital amino acids canines need to live long, active lives.

7.2.2 Technology-Based Startups Even though insect farming uses few resources, it can be labor intensive, which makes it challenging to grow such enterprises. The effort and expense involved are increased by the fragile nature of hatching, feeding numerous times each day, harvesting, and cleaning. The cost of insect protein is anticipated to be more competitive than that of animal protein as insect farming becomes more autonomous. These three startups are using technology to reduce the need for manpower.

7.2.2.1 Aspire Food Group Headquarters: Austin, Texas; London, Canada Aspire is developing "smart-farming" technology to reduce the financial and labor costs of insect farming. It is the parent company of the brand of protein bars made from edible insects called Exo, which it bought in 2018. It introduced the first robotic cricket farm in the world in 2017. Aspire is able to continuously monitor, track data, and investigate its insects thanks to sensors, autonomous robotics, centralized distribution systems, and custom assemblies. Recently, the company raised $16.8 million to build insect farms in Canada. The company has already established weevil and cricket farms in the United States, Ghana, and Mexico. Tradition Through ground-breaking innovations that are in keeping with our largest stakeholder, the earth, they are establishing the traditions of tomorrow. Aspire is leading a trend to generate incredibly high-quality protein with little impact on the environment. They envision a world where planet-based nutrition, which is better for people, animals, and plants, is widely available. Technology They are technological trailblazers, developing the world's most advanced, dense, wholesome, and environmentally friendly protein production system. Primary technology icon They cultivate insects that have a similar protein quality to meat and an environmental impact that is closer to plants using cutting-edge technology (such as robotics and automated data collection). Their technology is first class thanks to their dedication to constant innovation, creative rethinking, and exceptional nutrition.

7 • Startups  141 Smart farming: From hatch to batch Technology icon precision farming By using proprietary sensor technology and the internet of things (IoT) to collect real-time data on insects from hatch to batch, they put the "smart" in "smart farming." Technology icon modeling They use sophisticated modeling simulations, analytic methods, and data approaches to optimize procedures and advance farming technology. Technology icon zero waste To ensure that their farm meets strict sustainability criteria, they are installing zero-waste methods. Scalability To farm insects from hatch to harvest, they are creating autonomous robotics, centralized distribution systems, and custom assemblies. They can iterate on agricultural methods quickly and accurately because they use their own technologies. Their farms may be established anywhere and reliably produce dependable, affordable yield thanks to the process standardization and farm modularity. Lotte Confectionery Co. of South Korea has made a significant investment in Canada's Aspire Food Group, a manufacturer of edible insect protein, totaling over 10 billion won ($8 million). Lotte said in a statement on Thursday that, starting with the investment made through a local venture capital fund, it would like to work with the Canadian alternative protein business on everything from technical alliances to product development. As an emerging source of protein, edible insect products are of great interest to Lotte Confectionery, according to a statement from the business. One of its new playing grounds was the insect protein business. Aspire Food Group, a pioneer in producing cricket protein powder, was founded in 2016. In the first part of this year, Ontario, Canada, will have built the largest plant in the world for producing insect proteins. The investment follows a late 2021 agreement for joint research and development between Ynsect, a company specializing in insect protein research, in France and the research center of the Lotte Group. For the first time ever, Ynsect has made a mass breeding smart factory for mealworms commercially available. Due to their high protein content, edible insects are regarded as being very nourishing. Food businesses are rushing to enter the edible insect market as a result of their rapid growth and reproduction rates. Insects are largely utilized as animal feed, but over the next ten years, they have virtually unlimited growth potential as a source of protein, according to Lotte. Because an insect farm consumes only about one-eighth as much water and emits only about onethird as much CO2 as a typical cattle farm, they are also viewed as an environmentally benign food source. According to Global Market Insights Research and the South Korean Ministry of Agriculture, Food, and Rural Affairs, the edible insect market will increase from $112 million in 2019 to $710 million by 2024. Crickets, grasshoppers, and mealworms have not yet gained widespread acceptance in South Korea as food, with the size of the relevant market being only 30 billion won annually. In 2016, CJ Cheiljedang Corp. began collaborative research on edible insects with the South Korean edible insect research institute; however the company has not yet created any pertinent items.

7.2.2.2 BeoBia (The Bug Factory) Headquarters: London, UK A U.K.-based firm has developed a method for people to establish mealworm farms in their homes using kitchen scraps. Customers feed the mealworms by emptying waste into BeoBia "eco-growing pods,"

142  Edible Insects Processing for Food and Feed which produce a "constant supply of inexpensive, sustainable, and nutrient-rich pet feed." Mealworms naturally compost, and since the pods also create fertilizer, the idea has the potential to revolutionize the management of food waste, animal feed, and plant development all at once. BeoBia is Irish Gaelic for "food for life." They chose this name because it represents their purpose to produce and eat food sustainably without endangering the health of the earth. They consider insects to be essential to life since they are not only nutritious and delectable but also consume a much less amount of the planet's limited resources. Traditionally, 70% of the world's arable land is used for the production of meat. Since insects may be vertically farmed, raising them is a more effective alternative. Re_ is an insect-growing pod that gives users the ability to produce their own sustainable and healthful protein. Re_ is a straightforward and user-friendly product that comes in five modular trays that may be positioned anywhere in the house. Re_ makes it possible to harvest 100–300 grams of mealworms at a time. Mealworms provide approximately 54% protein in addition to a wealth of important vitamins, minerals, and antioxidants. A really sustainable approach is for the customer to feed mealworms with leftover fruit and vegetable waste and to use their frass (mealworm dung) as plant fertilizer.

7.2.2.3 FarmInsect Headquarters: Bergkirchen, Germany Through its user-friendly platform that guides farmers through the insect production process, the company, which was established in Germany in 2019, seeks to rethink agricultural techniques. Farmers without prior insect-harvesting experience can easily and rapidly pick it up thanks to FarmInsect technology. Insect farming produces feed for cattle, which in turn generates waste that may be utilized in insect farming, creating a sort of circular economy. When compared to the price of conventional animal feed, this closed-loop method can help farmers save up to 20% while enabling them to be self-sufficient and environmentally friendly. How a query turned into FarmInsect What are the greatest factors contributing to climate change? The company's founders realized that one of the main sources of CO2 emissions was agriculture. The question that followed was "And what can be done about it?" They understood this to be their area of expertise, and, when insects were allowed as animal feed in the EU in 2017, that here is where they could make a difference. Thomas and Wolfgang had previously established one with great success. And they were aware that the climate catastrophe required immediate attention. They quit their jobs as a result, and they started FarmInsect. That's why they left their companies and founded FarmInsect. Beneficial insects 30% reduction in costs: They are trying to cut costs as prices rise everywhere. They provide top quality at affordable pricing! Customers can save up to 30% on the price of traditional animal feed when using their locally produced feed. Independent prices Insects and waste are crisis-unaware. No matter what is happening on the global markets, it is still viable to produce insects as animal feed and profit from cheaper costs. Secure supply chains Globalization has led to a rise in the complexity and unpredictability of supply chains. Not with them; their supply chains are short, open, and simple, allowing for higher levels of security to be guaranteed.

7 • Startups  143 Tangible benefits • The 52 immunological peptides found in the black soldier fly represent the organism's "super strength" against sickness. It is the same for both humans and livestock: it is what it eats. • 50% less CO2 emissions. • 50% reduction in CO2 emissions. • In comparison to soy feed and imported fishmeal, feeding insects results in a 50% reduction in CO2 emissions, which helps fight the climate problem. Solution for the feed issue Europe's farmers face a challenge. They must import up to 90% of their feed from outside the EU due to a lack of protein. And those have dire repercussions: For the production of fishmeal, the seas are overfished and rain forests are removed. These steps to obtain protein feed are significantly hastening climate change. The answer might be to use locally and sustainably produced insects as agricultural feed. System of fattening Depending on the circumstance, they either install the highly automated fattening system on the farmer's property directly or provide him the facility's fattened larvae. The idea behind it is pretty straightforward: after the larvae are removed from the climate chamber, food is added. The box is automatically grabbed by the robot, and the compost and larvae are separated using a sieve. As soon as the boxes are full of feed, the robot sets them back down. The larvae are manually inserted. Young larvae & repurchase When the farmer's system is installed, they send him new animals every week. These are many black soldier fly breeding lines that are perfectly suited to the local diet. They've already increased feed conversion by 50% in just two years as a result. If the larvae are well-behaved, the farmer can feed himself off of them or sell them back to the business, making the most efficient use of the larvae with the least amount of waste. Larvae used as pet food When all the fattened larvae are taken into account, one portion goes to the farmers and the other to the producers of pet foods (pet food). They sustainably and regionally manufacture the larval food. It's crucial to note that all young larvae and larvae are grown in Germany using sustainable local agriculture for feed.

7.3  INSECT-FOCUSED FOODTECH STARTUPS IN EUROPE The practice of eating insects, or entomophagy, is anticipated to significantly alter the way food is produced. According to a survey, the consumption of food in Europe is thought to be responsible for up to 30% of all greenhouse gas emissions. The carbon footprint of meat is larger than that of plant-based foods. Insects are a viable alternative and a sustainable food source, and they have gained popularity as a result of the negative environmental effects of animal agriculture and meat production. In contrast to cattle, insects can be raised in great quantities without the need for a lot of resources like feed, water, or land. Insects are more resource-efficient and contain significant quantities of iron, zinc, and vitamin B12 in addition to having about 60% more protein than beef and 43% more than chicken. Meat worms, black army flies, crickets, and grasshoppers are among the most often utilized insects in food technology. Here is a list of rapidly expanding food tech firms in Europe that are focusing on insects and aiming to realize a sustainable and circular economy. Insects can be a fantastic source of protein for both animals and humans.

144  Edible Insects Processing for Food and Feed

7.3.1  Ÿnsect Headquarters: Evry, France A leader in the world of insect breeding and the production of premium ingredients for plant and animal nutrition, Ÿnsect is based in Paris. The start-up promises to create a more environmentally friendly protein source in place of soy, which consumes a lot of soil, land, and water to manufacture. Since 2016, this French firm has run a vertical farm; currently, a second facility called Ÿnfarm, with a goal of producing over 100,000 tons of insects annually, is being built near Poulainville. Insect, a company founded in 2011, wants to turn insects into a significant, premium component of pet, aquaculture, and plant food. The business provides a sustainable, all-natural response to the rising protein need around the world. In order to create "farm hills" (Fermilières®), low-footprint vertical farms utilized for Molitor breeding—small common beetles known as mealworms— Ÿnsect leans on innovative patented technology that is covered by 25 patents. At the Cleantech Open Global Forum in San Francisco in 2014, the business took home the People's Choice award. And more recently, Ÿnsect began to establish itself as a world authority on alternative protein. Through a number of investors headquartered in France, Belgium, the United Kingdom, Hong Kong, and Singapore, the company raised 110 million euros in 2019. By the end of October 2020, Ÿnsect had finished its series C funding round with about 310 million euros in investments, bringing its total funding from international investors to about 355 million euros since its creation. In 2020, Ÿnsect received more recognition when it was included in the Global Cleantech 100 list, a directory of the most creative and promising businesses that are expected to have a significant impact on the market and the development of international industries in the next five to ten years. Ÿnsect has been growing and processing insects into quality components for animal nutrition since its founding in 2011. We create new techniques for mass-producing insects and automate procedures for turning them into the highest-quality raw materials. Our expertise is unmatched in the globe, making Ÿnsect an indispensable partner for those working in the food business as well as for research facilities, financiers, and government organizations. The Molitor is a great bug for high-quality nourishment. We coined the term coleoculture, just as the fishing industry did with aquaculture. This phrase refers to beetle breeding, the most significant aspect of insect biodiversity. Mealworms have been naturally occurring around the planet for a very long time, including in Europe. Its social nature and nocturnal habits make breeding easier. Scientists have studied this insect in great detail, and fans have started growing it all over the world. Tenebrio molitor and Alphitobius diaperinus, popularly known as Molitor mealworm and Buffalo mealworm, were two of the beetle species that Ÿnsect selected. These insects have been cultivated for pet markets, zoos, fish feed, and currently for human food for many years. Mealworm includes elements necessary for the health of people, animals, and plants, including as vitamins and omega-6 polyunsaturated fatty acids, and is made up of 72% high-quality proteins. Our insects can be ground into flour and utilized as protein-rich ingredients in a variety of human food products, including pasta, bread, biscuits, and burgers. Why Mealworms? • It is the only insect protein that significantly benefits animal development and health, with farmed fish mortality reduced by 40% compared to natural diet. • Up to a 35 percent increase in output without fertilization can be achieved using mealworm frass as a natural fertilizer for plants and crops. • It is a highly digestible premium ingredient. • The young insects are fed and have a few weeks of healthy growth. 95 percent of the Molitor larvae are steamed, sterilized, and then converted without the use of chemicals into proteins and premium oil

7 • Startups  145 when they reach maturity. The remaining 5% mature into adults and procreate to maintain the juvenile population.

7.3.2  Ÿnsect Human Nutrition & Health (the Dutch Food Branch of Ÿnsect) Headquarters: Ermelo, The Netherlands Nutrition that works for a healthy lifestyle and the environment The way we eat will totally shift over the coming years. We can ensure both a healthy way of life and a healthy planet by using effective, sustainable solutions. It is made possible by AdalbaPro, the world's first ingredient line made from insects for the food and beverage industry, which was developed by nsect Human Nutrition & Health, the Dutch food division of nsect. Insect Human Nutrition & Health has developed products called AdalbaPro Fiber Textured Insect Protein (FTIP) and AdalbaPro Insect Protein Concentrate 80 (IPC80) that make using insect components in meals the new standard. The source of AdalbaPro is the resilient bug with great nutritional content known as the smaller mealworm, Alphitobius diaperinus. In comparison to other types of protein, it is not only incredibly nourishing but also a lot more sustainable. considerably less land, water, and feed are needed than with other options. This environmentally friendly method can be used to produce and improve food and drinks. such as pasta, baking supplies, sports nutrition, meat, and meat substitutes. Fiber with Textured Insect Protein, AdalbaPro Fiber Textured Insect Protein (FTIP) has a high protein content and is a good source of iron. FTIP is a fantastic substitute for traditional meat products because of its distinctive texture and ability to cook and brown like meat. AdalbaPro FTIP's flavor and color can be changed, providing countless customization options. AdalbaPro Insect Protein Concentrate 80 The soluble protein and nutrient-rich powder AdalbaPro Insect Protein Concentrate (IPC) is perfect for blending into drinks and shakes. IPC80 provides a comprehensive amino acid profile, a high level of digestible protein, and a wealth of vital vitamins and minerals. It has a mild, nutty flavor that makes it perfect for use in the fortification of a variety of foods and beverages. AdalbaPro Fiber Powder AdalbaPro Fiber Powder is a fiber-rich powder that also includes a range of vitamins, minerals, and proteins. The flavor of AdalbaPro Fiber Powder is subdued and nutty. FP is ideal for a wide range of applications due to its sweetness and high nutritional qualities. AdalbaPro Buffalo Mealworm Oil At room temperature, AdalbaPro Buffalo Mealworm Oil is a stable oil that is semi-solid. It is heat stable and has a great fatty acid composition (1:1:1 ratios of MUFA, PUFA, and SFA). The flavor of AdalbaPro BMO is mildly nutty. Due to its superior composition and qualities, the substance can be used in a wide range of applications. AdalbaPro Whole & Defatted Buffalo Powder AdalbaPro WBP and DBP are nutrient-rich food powders made from whole and defatted buffalo. AdalbaPro WBP and AdalbaPro DBP provide a wide range of nutrients, such as a full amino acid profile, dietary fiber, healthy fats, and a wealth of vital vitamins and minerals, such iron and vitamin B12. Both WBP and DBP have a mild, nutty flavor and high nutritional qualities, making them appropriate for a wide range of applications. In comparison to WBP, DBP has 30% more protein and 50% less fat.

146  Edible Insects Processing for Food and Feed

7.3.3 Innovafeed Headquarters: Paris, France Specifically for animal feed and aquaculture, InnovaFeed is a biotech firm engaged in the generation of a new source of protein from insect rearing (Hermetia illucens). The business's model enables it to put insects at the center of its agri-food system. To address the burden on natural resources and the need for a more responsive and sustainable method, it uses an industrial scale insect rearing procedure. For a fair price, InnovaFeed offers natural, sustainable protein to the fish farming industry. The global insect industry's highest production capacity By 2030, 10 additional sites will be operational, up from two in 2020. With the inauguration of its Nesle plant in November 2020, just 18 months after the foundation was set, Innovafeed has shown the scalability of its approach and is currently in an accelerated phase. Nesle, the world's largest operating insect producing facility This production facility, which began operating in Nesle in November 2020, shares space with Tereos (a manufacturer of starch) and Kogeban (a biomass plant), providing a singular example of industrial cooperation. It will make it possible for Tereos to locally value its byproducts while reducing the energy requirements connected with their processing, and it will allow Innovafeed to obtain a supply of highquality raw materials. Innovafeed may use Kogeban's waste energy thanks to the co-location. The co-location initiative for Innovafeed will save CO2 emissions by 57,000 tonnes annually. The pilot site in Northern France is called Gouzeaucourt. In the midst of the biggest European deposit of agricultural and agri-food byproducts, Gouzeaucourt, France, the first Innovafeed facility was officially opened in October 2017. The deployment of our manufacturing capacity to begin promoting our products has been made possible by this production unit, which values regional agricultural byproducts for insect rearing. To ensure the finest tracability, our production facility is GMP+ certified. Decatur will house Innovafeed's upcoming US manufacturing facility. By partnering with ADM, Innovafeed will recreate this special industrial symbiosis model in the United States at the Decatur plant, the biggest corn processing facility in the world, building on the knowledge gained in France. Through interconnected infrastructure connecting the two locations, co-products made from corn at ADM Decatur will be locally recycled and used to feed insects. Additionally, Innovafeed will be able to employ 27 MW of unused residual energy from the ADM process thanks to this manufacturing strategy.

7.3.4 Protix Biosystems Headquarters: Dongen, The Netherlands Dongen-based Insects are raised by Protix for use in sustainable meat, fish, and eggs. It produces sustainable elements like proteins and fats by breeding Black Soldier Fly (Hermetia illucens) larvae. Customers of Protix use these nutrients as components in their fish and animal feed. The company asserts that these components are more environmentally friendly than soy or fishmeal. The company claims that Protix's insects are fed on organic waste from the food sector. The insects are then utilized as food components in a variety of diets for fish, fowl, and other animals. Food residues are allegedly reused and added back to the food chain in a secure and genuinely circular system, according to Protix, which claims that this allows the company to close the food cycle. According to the company, utilizing insects as a protein substitute helps avoid overfishing and forest destruction for soy farming. Protix, founded in 2009, has grown to be the top supplier of insect ingredients, sourcing a wide variety of components from the Black Soldier Fly. By creating components from insects, Kees Aarts and Tarique Arsiwalla launched Protix in 2009 with the goal of promoting a sustainable food system. The firm

7 • Startups  147 has a proven track record of market introductions and innovation. Protix is able to produce consistently and at the proper quality because to cutting-edge control systems, artificial intelligence, genetic improvement programs, and robots. Global A-players like Buhler and Hendrix Genetics are partners with Protix. The World Economic Forum has given Protix the Technology Pioneer award. The Erasmus Centre for Entrepreneurship has named Protix as the fastest-growing Dutch firm with a social goal for 2020. Protix recently won the Dutch Innovation award 2020 for being the most innovative Dutch business. Sustainable innovation Protix makes investments in technology that fosters long-term expansion. Innovations in breeding facilities and machinery are crucial. Protix developed cutting-edge technologies and conducts business on an industrial scale. Science is at the heart of everything a company does, with a heavy emphasis on engineering and research to continuously further improve quality, controllability, efficiency, and overall competitiveness. The European Fund for Regional Development is providing financial assistance for this project: OPZuid Products ProteinX: One of their premier pet food and aqua feed products. The most sensible and environmentally beneficial animal proteins come from sources that are insects. High-quality lipids, micronutrients, and amino acids are present in Protix ProteinX to naturally improve the animal's health. In many dry and wet pet food and aquaculture applications, it completely replaces conventional protein while also bringing functional advantages and greater palatability. The pet food industry strives for a more sustainable feeding system while pets develop in harmony with nature. Aquaculture uses ProteinX successfully, from hatchery diets to grower diets for salmon, trout, shrimp, and other species. Applications will increase further in light of the EU's 2017 approval of the use of insect protein in aquaculture. LipidX: A beneficial and long-lasting energy source for numerous species. A major component of functional diets that focus on gut health is lipidX. Due to its high concentration of quickly metabolized medium-chain fatty acids, Protix LipidX, a pure insect lipid, offers a rapid source of energy. Protix LipidX, among other things, contains 40% lauric acid, which is renowned for its antibacterial qualities in the gastrointestinal tract. A particularly beneficial energy source for young animals with digestive issues and poor nutrient absorption is Protix LipidX. The enterprise lessens its environmental impact as the animals develop in good condition. High in lauric acid (40%) PureeX: A fresh component for tasty wet pet treats. All the best elements of ProteinX and LipidX are combined in Protix PureeX. Insects are processed fresh for PureeX, producing a component with exceptional nutrient absorption, palatability, and digestibility. For pets with digestive issues, Protix PureeX makes the ideal ingredient in wet pet food. It is beneficial for diets aimed at improving intestinal health and hypoallergenic. Without compromising flavor or health, the fresh puree contains a sustainable source of energy, helping to create a sustainable food system.

7.3.5 nextProtein Headquarters: Paris, France By exploiting insect protein as a feedstock, nextProtein was established to assist address the shortage of land and resources. The business creates insect-based protein that has the same nutritional content as conventional protein sources, has a very little carbon footprint, and costs substantially less. The procedure utilized by nextProtein has received EU approval for use in pet food, aquaculture, and as a feedstock for

148  Edible Insects Processing for Food and Feed the pig and poultry sectors. NextProtein, a company founded in 2015 by Syrine Chaalala and Mohamed Gastli, uses insect protein as a feedstock to address resource and land limitations. Like the majority of startups, nextProtein got its start in a garage in 2014. It was created as a result of the founders' personal and professional aspirations, who were looking for ways to alter agricultural food production while maintaining their ability to coexist and work together. An alternative to wasteful and unsustainable agricultural systems can be created by combining the organic cycles of nature with the scalable efficiency of technology, according to Syrine, an Emergency Operations Specialist at the UN Food and Agriculture Organization, and Mohamed, a Grammy-nominated chemical engineer. Syrine's prior employment in some of the world's poorest countries and Mohamed's aspirations for entrepreneurship served as the impetus. On a show, a locust plague ravaged crops and cattle pastures while the population went without food in Madagascar, which had just experienced natural calamities and a food scarcity. The pair realized their aim of more sustainable industrial farming operations by exploiting the life cycle of the Black Soldier Fly, which feeds on organic waste, after seeing the amazing capacity of insects to produce protein, which could feed cattle and hence the population. nextProtein Dry protein powder is used as an ingredient in animal feed for aquaculture, pet food, and other uses. nextOil A lipid product that can be used as an animal feed component for aquaculture. nextGrow A natural fertilizer for use in agriculture derived from the organic excrement and frass of the Black Soldier Fly.

7.3.6 Nextalim Headquarters: Poitiers, France Poitiers-based Insects have been utilized by NextAlim to convert food waste into proteins for the animal food and green chemistry industries. It creates methods for commercial insect farming that yield organic fertilizers and byproducts like proteins, oil, and other insect derivatives. Black soldier fly larvae are used in the company's breeding process to extract nutrients from organic waste and metabolize them into lipids, proteins, and other byproducts, conserving water, land, and other resources. NextAlim is a forerunner in the insect farming sector Since its establishment in 2014, NextAlim has amassed extensive experience in the entire process, from feed preparation to larvae growth and processing into proteins. As of right now, NextAlim specializes in BSF breeding and genetics. Their area of expertise is large-scale industrial neonate multiplication. NextAlim stands out among breeders as a player: • • • •

After eight years of R&D and technical development, industry maturity was attained. High capacity for manufacture. Ability to offer rearing guidance to consumers to guarantee best use of offered products. Delivering young animals to customers' locations through proprietary supply chains.

The local, mid-scale, low-carbon BSF larva offer is being developed by NextAlim: Since being acquired by La Compagnie des Insectes Industrie in May 2022, NextAlim has developed strong relationships with agricultural unions and cooperatives throughout France to support the creation of a network of entomo-bioreactors in five regional agricultural sectors that will feed local byproducts to BSF larvae, allowing the bioconversion process to occur in place rather than requiring the raw materials to travel several hundred kilometers. They develop the market's first low carbon larva offer in collaboration with

7 • Startups  149 breeders by converting abandoned low energy chicken farm buildings into bioreactors for the growth of our insects, fulfilling the environmental promises of entomoculture for feed. Black Soldier Fly eggs: Black Soldier Flies lay fresh eggs every day, which are manufactured at the NextAlim Poitiers facility, where their quality is closely inspected. Black Soldier Fly neonates: Freshly hatched neonates are carefully inspected for vitality before being packaged in special solutions and sent the same day to our customers, where they can be utilized for rearing. 5 to 7 days old larvae: They are raised from neonates that are seeded on premium feeding substrate. When necessary, a sorting step is conducted at the conclusion of the 5- or 7-day grow-out phase in order to remove the frass and give consumers calibrated, clean juvenile larvae.

7.3.7 Nasekomo Headquarters: Sofia, Bulgaria Nasekomo manufactures animal meals from organic waste. By reintroducing insects to people's plates, which are their natural diets and have incredible nutritional benefits and immunity boosters, the company claims to raise the productivity, health, and wellbeing of farmed animals. The Bulgarian business changes the way people eat by making higher-quality proteins more affordable for everyone. As well as reducing greenhouse gas emissions, food waste, and the loss of renewable resources like clean water, marine life stocks, and arable lands, Nasekomo also promotes environmental sustainability. Technology Fully automated technology that was created entirely in-house is scalable and incredibly economical. Innovations The R&D department works diligently, both independently and in partnership with top academic institutions and business partners, to enhance the insect industry and Nasekomo's position within it. We already have a number of innovations on the route to patent protection because of its current emphasis on genetics and automation. Base of Operations Facilities are in a prime position to benefit from the abundance of production materials that are consistently in low demand because they are the first insect rearing business in South-East Europe. Network Long-term partnerships with businesses all throughout the value chain, as well as with partners in technology and other fields, give them access to a consistent supply of supplies and a regular flow of customers. Strategy By fusing knowledge of insect breeding with ongoing development of unique technologies, they hope to establish themselves as the leading player in the insect market and offer customers high-valued goods and services. Whole Dried Larvae Used by poultry farmers to naturally feed birds and makers of wild bird feed to cover or coat fat balls. The protein and energy that birds need during the winter are found in whole dried larvae. Birds naturally eat insects, so there is nothing better for them to do than peck at some larvae. Although they can also be mixed, ground, inside the fat ball, whole dried larvae serve as a coating agent for fat balls. Whole dried

150  Edible Insects Processing for Food and Feed larvae can be picked up off the ground by chickens and eaten. It promotes their natural behavior, lowers stress levels, and prevents animal fights. Larvae that have been dried provide a potent mix of nutrients to augment their regular diet.Defatted Protein Meal Used by pet food producers to create hypoallergenic, healthful, and sustainable goods for dogs and cats as well as by aquafeed producers to provide a highly digestible and sustainable protein source alternative to fish meal. Protein content in defatted protein meals ranges from 50 to 65 percent, depending on the degree of defatting. There is a very low risk of allergies because it is a very uncommon and unusual source of protein for pets. Defatted protein meal is a premium ingredient for pets and is simple to incorporate into dry food and treats due to its high protein digestibility (almost 90%), presence of chitin, minerals, and micronutrients. For both parents and pets, it has an excellent flavor and a pleasing aroma. It offers fish meals with a very comparable amino acid profile the greatest nutritional and environmental alternative. Many fish species naturally consume insects, which can give them essential nutrients like chitin and antibacterial peptides. Enhanced FCR, weight gain, and feed intake are all benefits of using insect protein meal in aquaculture, as well as better health and animal welfare (measurable through blood markers, animal gut health modulator, reduction of antibiotic, back to naturality, reduction of mortality rate and diseases issues). Insect Oil Used by producers of chicken and pig feed to add a useful source of fat to starter feed diets for animals and pet food makers to create healthy, useful goods for dogs and cats. Refined insect oil is a great substitute for coconut oil or other vegetable oils since it contains about 50% lauric acid, a potent natural antioxidant that helps to preserve pets' immune systems. Due to the availability of Omega 6 and 9, it can be utilized as a natural palatability enhancer (used to coat kibble and treats) and a good source of energy. In terms of nutrition, the post-weaning stage is crucial for piglets. Insect oil has health advantages (supporting the digestive system) and improves general animal welfare, which can help to lessen the problems caused by food changes (stress control, decreasing mortality rate and the use of antibiotics). Lauric acid is a potent antibacterial and antiviral agent that is present. In place of soybean oil or other vegetable oils, insect oil can serve as a useful addition to the diet of hens. The high concentration of medium-chain fatty acids, including lauric acid, can boost their immune system, have a favorable impact on their gut microbiome, and increase productivity by giving them a valuable source of energy derived from an insect larvae, which is a natural diet for poultry. Organic Certified Insect Fertilizer Used by farmers to create organic fertilizers based on insect frass or as a direct soil amendment and fertilizer. Castings from Black Soldier fly larvae and food fibers are used to make fertilizer. It is the ideal option for nourishing and repairing soil. Fertilizer aids in producing healthy crops and fortifies plants' inherent defenses because it is abundant in organic matter, vital minerals, and chitin. Our fertilizer is excellent for growing vegetables since it has N, P, and K levels at 3-3-3. Black Soldier Fly Eggs Employed by businesses involved in the bioconversion of organic waste and BSF insect farms. The most important aspect of insect farming may be biology. They work with living things that are extremely sensitive to environmental factors. After hatching, a gram of eggs can hold up to 40.000 neonates. BSF eggs are the result of years of research to strengthen the genetic pool, and they will soon be supported by genetic advancements made by the Fly Genetic firm, a partnership between Nasekomo and the leading animal genetics company, Grimaud Group. Eggs can assist establish a new Black Soldier Fly colony, support an established one to increase production, and also schedule daily needs to support bioconversion units and maintain the conversion flow.

7.3.8 Hexafly Headquarters: Meath, Ireland Hexafly, an Irish agritech business, raises insects to feed fish, fertilize plants, and provide protein to foods and medications. Black soldier flies, which are a low-value waste stream, are transformed into bioplastics,

7 • Startups  151 fertilizers, and valuable feeds by Hexafly's sustainable insect farming approach. Numerous businesses in the medicinal, cosmetic, animal feed, food, and plant nutrition industries use the insect-derived goods that Hexafly provides. Black Soldier Fly, a microscopic insect that has the potential to feed the entire world, is the only focus of Ireland's vertical insect farm. Their commitment to the Black Soldier Fly has led to technological developments that make it possible to feed plants and animals a diet that is very nutritious, sustainable, and natural. Innovative Solutions From Ireland They generate sustainable, nutrient-dense protein using cutting-edge bioconversion technology to feed the hard-working plants and animals on our world. Hexameal Their insect protein is a naturally occurring, high-protein animal feed that is sustainably produced. It is ideal for hypoallergenic pet diets and contains a wealth of essential amino acids and highly digestible proteins. A natural and sustainable alternative to soy and fishmeal is hexameal. Improved animal intestinal health and a rise in immunological defenses are further advantages of food. Hexafrass Natural protection from disease and pests provided by nature. It has been demonstrated to lessen the impact of aphids because it contains bio-active insect chitin, which not only improves soil health but also actively encourages root growth and plant growth. Hexaoil Insect oil is a great high-energy animal feed because it includes easily digested fatty acids. It is a sustainable alternative to many coconut and palm oils and has a composition that is almost comparable to coconut oil. It provides a highly advantageous blend of good fats, unsaturated fats, and other necessary elements. Its high lauric acid concentration improves immunity and intestinal health. Hexagrubs The ideal high-protein, calcium-rich diet for local fish, hedgehogs, birds, and reptiles. It is ideal for hypoallergenic pet diets and contains a wealth of essential amino acids and highly digestible proteins.

7.3.9 Entocycle Headquarters: London, UK A London-based company called Entocycle raises insects with the goal of protecting and restoring the environment. It was started to stop the industrial exploitation of the natural environment by farming insects as a sustainable substitute for soy and fishmeal. The business is constructing a brand-new facility for insect farming that will be utilized to breed black soldier flies. Food scraps will be fed to these flies to produce sustainable insect protein. Optimum, effective insect farming technology To enable precise Black Soldier Fly larvae counting and dosing, they have created a high-performance, unique neonate counter. Story Keiran Whitaker (Founder) had the good fortune to live on sun-drenched beaches in some of the world's most breathtaking locations during his five years as a scuba diving instructor. But on his travels, he saw firsthand the harm that our ailing food production system is doing to the environment. One day he made the decision to step in and created a more effective and long-lasting method of feeding the globe. “Nature has spent 150 million years perfecting how to feed animals. It’s called insects.”

152  Edible Insects Processing for Food and Feed The perfect natural ingredient for animals Whole insects offer a complete protein that is comparable in nutritional value to meat and fish since it contains all essential amino acids, healthy fats, and key minerals including calcium, iron, potassium, magnesium, phosphorus, and zinc. Natural and wholesome • The larvae of black army flies are rich in protein, vital amino acids, lipids, and minerals. • They have a low environmental footprint. • Using a lot less water and area, insects generate protein far more effectively than farmed animals. UK-based farms • In the UK, insects are entirely farmed. Scalable and effective insect farming technology • Accurate Black Soldier Fly larvae counting and dosing are made possible by a leading neonate counter. • Counted and batched larvae that are ready for production are delivered via a complete BSF breeding system. • Entocycle offers a team of skilled engineers who have a history of completing plants from conception to commissioning. • They can help with the following facets of constructing a new BSF facility: – Feasibility studies – Factory design – Project management – Colony build-up

7.3.10 Hargol FoodTech Hargol FoodTech is based in Israel and produces products based on grasshopper for both human and animal nutrition. They sell a protein powder with 72% proteins, omega 3s, minerals and vitamins. From a farming point of view, they claim to be sustainable because they have a year-round vertical farming production with innovative cage infrastructure to increase density. They operate four different facilities in Israel and they claim to be the first company in the world to commercially develop and raise edible locusts.

7.3.11 BetaHatch BetaHatch is a startup based in Washington. They cultivate and process mealworms to use them as poultry, acquaculture, swine and pet feed products. They currently sell mealworm meal, oil and whole dried mealworms. The company also sells an organic crop fertiliser made from insect waste. The dry fertiliser is supposed to stimulate plant growth and positively impact soil health. Due to their closed system, they claim to operate a zero waste protein production.

7.3.12 Grubbly Farm Grubbly Farm is based in Atlanta and it grows and processes black soldier fly larvae as chicken feed and pet food. As chicken feed they sell starter grower feed, layer feed and chicken snack for all ages. For dogs they sell snacks. Their insects recycle food waste before being oven dried and then blended with other ingredients such as vegetables and grains.

7 • Startups  153

7.3.13 Plento Plento is a joint venture between Singapore's Asia Insect Farm Solutions and Thailand's cricket pasta maker Bugsolutely. Plento is developing chips that mix plant-based proteins and insect proteins, usingvegetables such as mung bean and chickpea together with cricket flour and mealworm flour.

7.3.14 Insectta Insectta is the first urban insect farm in Singapore. They rear black soldier flies and convert food waste to insect products to be used as pet feeds and organic plant fertilisers. They sell both dried and live larvae, eggs, fertiliser and a seed starting kit. They have recently invested in the extraction ot chitosan and they are currently working on the production of semiconductors and probiotics.

7.3.15 Protenga Protenga is based in Singapore and transforms agricultural and food waste into insect products for fish, poultry, pets and crops using black soldier flies. They produce a protein meal, oil and frass (an organic fertiliser made from insect waste) and they have a pet food division.

7.3.16 Bugsolutely Based in Bangkok, Thailand, this company makes cricket pasta, using 20% cricket flour and 80% wheat flour. Thanks to the cricket addition, the pasta has increased amounts of protein, omega 3s, vitamins and minerals.

7.3.17 Magalarva Based in Indonesia, this company uses food waste and industry by-products to feed their black sodier fly larvae. They produce protein for animal feed, pet food, oil and an organic fertiliser.

7.3.18 Entobel Entobel is based in Singapore but has the production facilities in Vietnam. They convert low value biomass into functional ingredients. They produce a meal from defatted and dried black soldier fly larvae, as well as oil and a soil fertiliser.

FURTHER READING http://nextprotein​.co/ http://www​.ynsect​.com​/en/ http://www​.ynsect​.com​/en/ https://aliainsectfarm​.it​/chi​-siamo/ https://app​.dealroom​.co​/ lists​/14805

154  Edible Insects Processing for Food and Feed https://aspirefg​.com/ https://betahatch​.com/ https://bugfactory​.co​.uk/ https://eatsens​.com/ https://eatsmallgiants​.com​/why​-edible​-insects/ https://escalon.services/blog/startups-focused-on-edible-insects-take-flight-around-the-world/ https://farminsect​.eu​/en/ https://grubblyfarms​.com/ https://hargol​.com/ https://hexafly​.com ​/about https://hexafly​.com ​/products https://innovafeed​.com ​/en ​/our​-story/ https://jiminys​.com/ https://magalarva​.com/ https://mightycricket​.com/ https://nasekomo​.life​/about/ https://nasekomo​.life​/impact/ https://nasekomo​.life​/products/ https://plentofoods​.com/ https://protix​.eu​/products​_by​_protix/​# proteinx https://protix​.eu​/protix​-journey​/protix​-history/ https://sg​.style​.yahoo​.com ​/insect​-start​-ups​-southeast​-asia​-sustainable​-diet​-future​- 055602330​.html https://siliconcanals​.com ​/news​/startups​/insect​-focused​-food​-tech​-startups​-europe/ https://susinchain​.eu/ https://us​.acrofan​.com ​/detail​.php​?number​= 678662 https://www​.bugsolutely​.com/ https://www​.entobel​.com/ https://www​.entocycle​.com/ https://www​.globenewswire​.com ​/en ​/news​-release​/2022​/05​/25​/2450095​/0​/en ​/ Edible ​-Insects​-Market​-Worth​-9​- 6​ -Billion​-by​-2030 ​-Exclusive​-Report​-by​-Meticulous​-Research​.html https://www​.greenbiz​.com​/article​/these​- 6​-startups​-are​-creating​-buzz​-around​-edible​-insects https://www​.illegaloats​.com/ https://www​.insectta​.com/ https://www​.nextalim​.com/ https://www​.nextalim​.com ​/products https://www​.protenga​.com/ https://ynsect​-food​.com/

Mass Production Technologies

8

Waqar Majeed, Rimsha Naseem, Naureen Rana, Hammad Ahmad Khan, Uzma Ramzan, Sobia Kanwal, Elmo Borges de Azevedo Koch, Nazia Ehsan, Muhammad Sarfraz Ahmed, and Muhammad Naveed Contents 8.1 Introduction 8.2 Characteristics of Insects for Automated Rearing 8.2.1 Distinct Features of Insect Farming 8.2.2 Different Species of Insects for Food and Feed 8.2.3 Other Applications 8.3 General Methodology for Mass Production 8.3.1 Feed Principles for the Mass Production of Insects 8.3.1.1 Solid-feed 8.3.1.2 Semisolid feed 8.3.1.3 Liquid feed 8.4 Feed and Nutritional Requirements for Insect Rearing 8.4.1 Macronutrients 8.4.2 Micronutrients 8.4.3 Plant Material 8.4.4 Laboratory Diet 8.5 Equipment and Mechanization for Insect Mass Rearing 8.5.1 Production and Operation Management 8.5.2 Rearing Area 8.5.3 Feeding and Watering 8.5.4 Separation and Sorting 8.5.5 Cleaning Room 8.5.6 Dung Area 8.6 Production and Processing Technologies by Species 8.6.1 Black Soldier Fly 8.6.1.1 Adult colony and its management 8.6.1.2 Mating and oviposition 8.6.1.3 Production of larvae and its maintenance 8.6.1.4 The feed used for rearing 8.6.1.5 Costs and quality maintenance 8.6.1.6 Impact of different factors on the growth of BSFL

156 157 157 158 158 159 160 160 161 161 161 161 162 163 164 164 164 165 166 166 167 167 168 168 168 169 170 171 171 171

DOI: 10.1201/9781003165729-8

155

156  Edible Insects Processing for Food and Feed 8.6.2 Crickets 171 8.6.2.1 Production of crickets worldwide 171 8.6.2.2 Rearing units 171 8.6.2.3 Diets and feeds 172 8.6.2.4 Environmental conditions 173 8.6.2.5 Reproduction 173 8.6.3 Mealworm 174 8.6.3.1 Rearing 174 8.6.3.2 Feed 175 8.6.4 Housefly 175 8.6.4.1 The feed and its maintenance in rearing 176 8.6.4.2 Process of production 177 8.6.5 Waxworm 178 8.6.5.1 Development of larvae and their diet maintenance 178 8.6.5.2 Rearing and reproduction 179 8.7 Environmental Control and Conditions 180 8.7.1 Physical Factors 180 8.7.1.1 Light, temperature and location 180 8.7.2 Mechanical factors 181 8.7.2.1 Filtration system 181 8.7.2.2 Panels and pads for evaporation 181 8.7.2.3 Humidifiers 181 8.8 Basic Needs for the Supply Chain System 182 8.8.1 Feed 182 8.8.2 Farms and Farmed Species 182 8.8.3 Transportation, Storage, and Distribution 183 8.8.4 Processing and Manufacturing Infrastructure 183 8.9 Challenges 184 Acknowledgement 185 References 185

8.1 INTRODUCTION Insects are internationally accepted and traditionally eaten as food in many areas worldwide (Murefu et al. 2019), ensuring nutritional value for humans (Zielińska, Karaś, and Baraniak 2018). The history of using insects as food is a long one. Different historical documentation and its uses are highlighted throughout history. Silkworm species were probably being used in China as early as 4,000 years ago. However, the development of farming of insect species extended as food sources and other services such as silk and honey production (Bombyx mori and Apis mellifera, respectively). Before the 18th century, honey from honeybees was the most usable source of sugar. Sericulture history seems to be very old (4500–3000 BC) because of its easy culture. The traditional practices were started about 7,000 years ago (Ramos-Elorduy 2009). Insects collection of wild has been adapted with different materials and techniques, but these are less efficient for mass production. Insects eating collection and rearing cost over the calories remained a debate and opportunity-based habit. The most suitable region for the faster growth and reproduction trend is tropical because of the sustained humidity and temperature over the year. Not unexpectedly, tropical and subtropical areas are the most durable entomophagous areas. On the other hand, insects have high feed conversion ratios, reproduce, and grow easily (Smil 2002). The average edible portion is 40% of beef, chicken (58%), pork (55%), and orthoptera species (80%) (Nakagaki and Defoliart 1991). Insects have little

8  •  Mass Production Technologies  157 environmental impact relative to conventional animal production but maintaining a good environment for insect farming is very little understood (Dobermann, Swift, and Field 2017). The large context of insects production as food and feed is not a widespread activity. Little knowledge is accessible, and production state of art remains rudimentary in contrast with technology production for other livestock farming. Mass production has become the most concerning issue in the present scenario because farming other than insects produces the bulk of the mass in a day. Still, many industries worldwide are producing large-scale edible insect and their products. By comprising the nutritional value of insects and their products with other sources are always leading. It needs to extend the setup of industries and farms, so food security and health food issues may be resolved. Some tactics are necessary to make this business more profitable and sustainable such as cheap insect feed production, mechanization, building infrastructure, intensively cultured rooms, and other processes. The term entomophagy, or eating insects, is no longer hidden in the modern world, despite its historical use (Müller et al. 2016). Today, around 2.5 billion individuals consume insect species as a necessary component of their diet (Van Huis 2016). Insects have an adequate organic food supply for the world’s development and emerging regions to meet the monstrous problems of starvation and undernutrition (Mishyna, Chen, and Benjamin 2020). Insects are often used as animal feed due to their high nutritional content and natural diets (Rumpold and Schlüter 2013; Veldkamp and Bosch 2015). Yellow mealworms, black soldier fly larvae, termites, and grasshoppers are commercial insects used as feed (Dobermann, Swift, and Field 2017). These insects take the place of costly feed; for example, the house fly is used as a source of protein for fish (Awoniyi, Aletor, and Aina 2003). A broiler diet containing 20%, 25%, or 30% soybeans might be substituted with 0%, 5%, or 9% yellow mealworms (Ramos-Elorduy et al. 2002). Housefly larvae (10–15%) used to feed broilers have a high rate of growth (Hwangbo et al. 2009). A fish farmer employed house fly larvae to feed Nile tilapia, black soldier flies to feed Salmo salar, and mealworms to feed catfish, which outperformed any commercial market feed product (Roncarati et al. 2015; Lock, Arsiwalla, and Waagbø 2016; Wang et al. 2017). Due to their particular qualities, such as well off supplies of proteins, fats, flavour, and fast energy sources (Finke and Oonincx 2013). Many experimental analyses characterize insects as a major source of protein in animal feed. Both standards of health and safety operating procedures (SOPs) are mainly concerned with the mass processing of insects as food. In the present era use of sterile insect technique (SIT) and many other useful techniques has developed other than insects as a food source. These cultures are probably the most developed and mass-reared. This chapter summarizes the scientific approaches of mass production of some insects species as food and feed, their production methods depending on the technologies, and production information to deliver an image of futuristic edible insects rearing industry.

8.2  CHARACTERISTICS OF INSECTS FOR AUTOMATED REARING In the present era, more than 2000 species of insects are edible and different products as a part of food and feed (Ramos-Elorduy 2009; Naseem et al. 2021), although a minimum quantity is used for farming (Table 8.1).

8.2.1 Distinct Features of Insect Farming Insects have frequently been considered small livestock superior to conventional livestock production, representing 40% of global GDP (Steinfeld et al. 2006). Insect farming is economically significant since

158  Edible Insects Processing for Food and Feed TABLE 8.1  Number and percentage of edible insects orders ORDER Coleoptera Lepidoptera Hymenoptera Orthoptera Hemiptera Isoptera Diptera Blattodea Others

EDIBLE SPECIES Beetles Caterpillars Bees, Wasps Grasshoppers, Locusts, Crickets Cicadas, Leafhoppers, Plant hoppers Termites Flies Cockroaches

NUMBER 659 362 321 278 237 59 37 37 45

PERCENTAGE (%) 32.38 17.79 15.77 13.66 11.65 2.90 1.82 1.82 2.21

it provides an abundant source of revenue in many developing countries, including Southeast Asia and South Africa. They act as a divergent economic tool because of their culture in urban, semi-urban, and rural areas (Oonincx and de Boer 2012). Some of the distinctive features of insects farming are elaborated in (Figure 8.1).

8.2.2 Different Species of Insects for Food and Feed The most common insects which are being used to produce food and feed comprise Jamaican field cricket (Gryllus assimilis), House cricket (Acheta domesticus), Locusts (Locusta migratoria), Field cricket (Gryllus bimaculatus), Meal worm (Tenebrio bimaculatus), European honeybee (Apis mellifera), Silkworm (Bombyx mori), Housefly (Musca domestica), Palm weevil (Rhynchophonus ferrugineus and Rhynchophonus Phoenicis), Green rose chafer (Cetonia aurata), Common bottle fly (Lucilia sericata) and Waxworm (Galleria mellonella) (Table 8.2).

8.2.3 Other Applications Insects are also harvested for pollination and agricultural pests control. On a medium scale, ladybugs, bees, and lacewings production have also been raised. Some insects, including aphids, are often raised to feed insects and any of such insects that are not edible (Scudder 2009; Cock et al. 2017). For research purposes, some of the species are used, such as Drosophila melanogaster and Anastrepha fraterculus in

FIGURE 8.1  Key features of insects farming. (Reprinted with permission from Springer Nature [Naseem et al. 2021].)

8  •  Mass Production Technologies  159 TABLE 8.2  Comparison of insects species used as food and feed NAME Jamaican field cricket House cricket Locusts Field cricket Meal worm European honeybee Silkworm Housefly Palm weevil Palm weevil Green rose chafer Common bottle fly Waxworm

SCIENTIFIC NAME Gryllus assimilis Acheta domesticus Locusta migratoria Gryllus bimaculatus Tenebrio bimaculatus Apis Mellifera Bombyx mori Musca domestica Rhynchophonus ferrugineus R. Phoenicis Cetonia aurata Lucilia sericata Galleria mellonella

FEED

FOOD ✔ ✔ ✔

✔ ✔ ✔ ✔ ✔

✔ ✔ ✔

✔ ✔ ✔

genetic experiments (Yamanaka, Rewitz, and O’Connor 2013; Cladera et al. 2014) and Manduca sexta in biomedical research (Venter and Schneemann 2008) and also different other services are provided by insects (Freelance 2019). Different societies worldwide have used many insect species medicinally (Costa-Neto 2005; Dossey 2010). Insect-consuming cultures appear to combine them with different health advantages besides diet (Raheem et al. 2019). Caterpillar fungus, for example, is expected to have immunostimulatory and anti-cancer properties (Chen et al. 2009). The Periplaneta americana (American cockroach), marketed in China and Korea for baldness treatment, is one of the only species reported to be developed for medicinal purposes as mass. However, in many communities, other insects are used to treat different diseases. Lucilia sericata (Diptera) was used for maggot therapy and allantoin larvae secretions were used as osteomyelitis therapy. Honeybee venom has multiple uses, including inflammation, discomfort and asthma (Kampmeier and Irwin 2009). However, insects may be commonly used in medicines to mediate impact on antimicrobial and antifungal peptides, insulin synthesis, antioxidant enzymes and ACE (angiotensin-converting enzyme) (Mlcek et al. 2014; de Castro et al. 2018). The development of vaccines and some other important protein, such as the baculovirus vaccine, is one of the fascinating uses of insects in medicine. After a moth larva is injected with a baculovirus, a vaccination or protein may be obtained from the insect’s hemolymph. (Madhan, Prabakaran, and Kwang 2010). Recent studies have established blood glucose-reducing agents, which leads to the production of silkworm powder as a diabetic medicine in Korea. The Chinese Ministry of Health and the State Food and Drug Administration have also documented such gain (Belluco et al. 2013; Nongonierma and FitzGerald 2017). In summary, identifying physiologically significant compounds in edible insects is a possible and important use for medicinal purposes (Han et al. 2017). However, considering the present absence of empirical evidence, more study is warranted to validate the advantages of using insects as food or medication.

8.3  GENERAL METHODOLOGY FOR MASS PRODUCTION The growth of the insect food industry is deficient in verifiable knowledge and the essential components of the food supply chain compared to other food production systems focused on conventional domestic animals (Cortes Ortiz et al. 2016; Dobermann 2017; Van Huis 2017). It is also recognized that insect farming skills and the design of facilities are urgently needed as wild harvests give way to mass rearing in the entomophagy industry (Miech et al. 2016; Van Huis 2013).

160  Edible Insects Processing for Food and Feed Livestock is currently one of the fastest-growing agricultural subsectors in most developed countries, increasing demand for livestock products (Thornton 2010; Makkar et al. 2014). Nevertheless, many developed countries have deficits in the availability of feed to meet the demand for livestock. A significant factor in addressing this shortfall may be new alternative feed resources. FAO addressed fruit and vegetable waste usage to process other value-added products as feed for livestock and substrates (Wadhwa and Bakshi 2013). Insects are a key substitute to protein sources in South-East Asia, Central to West Africa, and Central to South America. However, western customers are very skeptical and disgusting about introducing insect-based food and insect-derived proteins as part of their diet (Vanhonacker et al. 2013). Barsics et al. (2017) documented that the consumption of insects is already a common or even typical habit for several communities worldwide. In western cultures, though, insects are mostly deemed and unpretentious, even in dire circumstances. Insects do not only add protein and significant amino acids to human diets but also fatty acids. Moreover, as already stated, they are effective for rearing, and waste from their rearing is suitable as an organic fertilizer within the biorefinery definition (Verbeke et al. 2015). In addition to being grown using agricultural waste, they need to feed six time less than livestock. Additional theories are that insects have a higher feed conversion ratio, low water and energy consumption, and are a healthy source of important protein for feeding animals (Nugroho and Nur 2018). In reality, in aquaculture and poultry, insects are already used as feed. More than 2000 insect species seem to be an interesting edible resource with health benefits worldwide. These protein-rich insects provide an acceptable choice rather than classical forms of protein, lowering feed costs and waste (Naseem et al. 2021; Borrelli et al. 2017). A nutritious diet consisting of organic by-products may be as effective as industrial breeders’ diets to develop mealworm species’ efficient development (van Broekhoven et al. 2015). A food-based organic diet has been critical for larval development, mass density, and colony maintenance (Morales-Ramos and Rojas 2015). Recycling and transforming low-quality, plant-based waste into high-quality energy, protein, and fat food will take a reasonably short time with mealworms. In addition, the omnivorous household cricket Acheta domesticus can be provided with a wide variety of organic products, which makes it possible to grow six or seven generations per year in a system (Makkar et al. 2014). The use of insect meals in animal feed involves mass insect processing. Currently, there are no clear information available for highquality insect processing, in particular for the insect breeding on biowaste, organic side stream substrates (Makkar et al. 2014; Bosch et al. 2019), and for the rearing of black soldiers, three basic methods are used, such as poultry manure, pig manure and food wastes.

8.3.1 Feed Principles for the Mass Production of Insects Some fundamental tactics are important for the production of insects, such as raw materials and improved feed production. Additional values, feed processing activities and multiple feed presentations will be required to balance the nutritional needs of farmed insects with the nutrients provided by these feedstocks (powder, pellets, semi-moist, liquid). In addition, feed compositions that are less costly and less adversely harmful should be explored. Makkar et al. (2014) discussed the elaborated material regarding the feeding composition of insects as feed. Insects can be classified as solid and liquid feeders (Resh 2003); liquid or liquid feed mouth parts are more suitable for suckers. Still, most liquid eaters can liquefy solid foods through extra-oral digestion (Cohen 2003). Solid feeders can chew and reduce large solid food particles into ingestible pieces (Morales-Ramos, Rojas, and Coudron 2013). Following is the discussion of liquid, semi-solid and solid feeds.

8.3.1.1 Solid-feed Feed in the form of mash is the most cost-effective form to produce. This involves grinding and mixing of all the raw materials for the nutritious needs of insects. No more compaction and heat procedure with feed

8  •  Mass Production Technologies  161 is carried out; thus, energy costs for preparing the feed are minimal relative to extruded and pellets feed. While on the contrary, raw products are also segregated owing to shipping and storage. The nutritional quality of the feed, particularly the intake of insect feed by individuals, may be affected. Pelleting is by far the most popular processing of feed production. Raw feed ingredients are dosed to the ground, mixed, formed into pellets in a pelletizer, and then cooled (Abdollahi, Ravindran, and Svihus 2013; Tumuluru 2019). Solid feeds can also be extruded. Extrusion manufacturing includes finely molding, mixing, preconditioning, boiling and cooling the mash, and shaping with the aid of an extruder. The benefits of the extraction method include cooking advanced ingredients, inactivating certain antioxidant factors (ANF) and feed sterilization (Bensky 2019; Mościcki 2011).

8.3.1.2 Semisolid feed A semi-moist is a feed in which the humidity limits are variable. The content of water activity in these feeds can vary from 0.65 to 0.85 is more significant. The amount of water activity in a product affects its shelf life. To prevent microbial or fungal food deterioration, sugars, salts, or glycols modify the durability. Semi-moist feed texture, produced by pelleting and extrusion, is softer than solid feeds. The insects must bite or chew tiny pieces of this meal to consume it. Mouth-sucking insects can easily consume semisolid foods.

8.3.1.3 Liquid feed As such, liquid feeds are not made in certain instances where livestock are fed liquid foods, milk feed, or milk feed substitutes. These are powder that dissolves quickly when mixed with hot water. In either case, the ingredients are soluble or ground into a fine powder. The spray-drying liquid solutions may produce particular feeds, produce free-flowing powders, full nutritionally healthy feed, or a concentrate mixed with other ingredients, probably on the farm.

8.4  FEED AND NUTRITIONAL REQUIREMENTS FOR INSECT REARING Most insects are omnivores in their natural habitats which are farmed worldwide for nutrition. Omnivores exhibit high dietary diversity and may eat the food in diverse contexts. However, their dietary preferences are dynamic and challenging to classify. Since several dietary sources are used, the proper proportion of each product is eaten in nature is impossible to ascertain (Morales-Ramos, Rojas, and Coudron 2013). The dietary versatility of omnivorous insects, on the other side, allows them to grow and reproduce on substandard diets. This food diversity will make them easier to depend on low-value food supplies, making them suitable for large-scale farmers. Although a balanced diet can make the production industry of its high level (Won and Kwon 2009; Van Huis 2013; van Broekhoven et al. 2015).

8.4.1 Macronutrients Carbohydrate is considered one of the best sources of energy. It is necessary for the production of polysaccharides, amino acids and chitin found in arthropods exoskeleton. Monosaccharides and disaccharides are the most abundant sugars involved in arthropod feeding. Sugars are often used to refer monosaccharides (galactose, glucose and fructose) and disaccharides (maltose and sucrose). Depending on the kind of arthropod, the needs for each particular sugar differ. Many insects can absorb and metabolize glucose and

162  Edible Insects Processing for Food and Feed fructose, but certain monosaccharides, like ribose, arabinose, galactose and xylose, are not metabolized despite being rapidly absorbed. Specialist feeders cannot digest sucrose and maltose, which are disaccharides, while generalist feeders can. Herbivores and predators are examples of generalist feeders, whereas parasitoids and certain mites are examples of specialized feeders (Cohen 2003). Predatory arthropods can obtain carbohydrates from prey in the form of the disaccharide (trehalose) and the polysaccharide (glycogen), and these two main forms are reserves in insects. However, the amylase in insects can digest starch, which is also present in many non-prey food sources consumed by omnivore predators (Morales-Ramos, Rojas, and Coudron 2013). Lipids are structurally important components of the cell membrane. They also serve as a barrier to water conservation in the arthropods cuticle which offer an effective method for storing and delivering metabolic energy under prolonged demands. Polyunsaturated fatty acids, such as linolenic acids and linoleic, are important for the nutrition of insects. Insects synthesize them in minor quantities. The synthesis of polyunsaturated fatty acids by insects has been proven in certain species, while limited capability is found in others, including cockroaches, mosquitoes, and aphids. Derivatives of polyunsaturated fatty acids, known as eicosanoids, stimulate oviposition in crickets and may be important for reproduction in all insects (Švácha 2013; Morales-Ramos, Rojas, and Shapiro-Ilan 2013; Morales-Ramos, Rojas, and Coudron 2014). In the fat body of insects, phospholipids are synthesized. Phospholipids are essential for the transfer of lipids, the production of vitellin and other lipoproteins (Morales-Ramos, Rojas, and Coudron 2013). Insects can readily ingest and utilize phospholipids, an excellent supply of polyunsaturated fatty acids (Švácha 2013). Proteins are categorized as globulins, insoluble proteins, lipoproteins, and nucleoproteins based on their solubility and function. Globulins contain hormones, enzymes, antibodies and protein. Nucleoproteins are linked with ribosomes and nucleic acids (Morales-Ramos, Rojas, and Coudron 2013). Lipoproteins are usually transport proteins and insoluble proteins and these passive compounds sometimes referred to as structural proteins (Feingold and Grunfeld 2000). Protein is a necessary component for the growth and development of all organisms (especially insects). Increasing the protein content of yellow mealworms, T. molitor in food, reduces development time (Morales-Ramos et al. 2010) and improves weight gain (van Broekhoven et al. 2015). The ten essential amino acids should be present in arthropod diets i.e., methionine, lysine, phenylalanine, arginine, isoleucine, tryptophan, valine, leucine, histidine, and threonine (Cohen 2003; Lundgren 2009). These ten amino acids are considered essential because insects are unable to synthesize them. Other amino acids that can be produced but only in small amounts or that need a significant amount of energy to be made may also be necessary for the diet of insects (Švácha 2013; Lundgren 2009). Sclerotin is composed of tyrosine, which is an important component needed in large amounts during moulting. (Hopkins and Kramer 1992). Aspartic and glutamic acid, cysteine, serine, and glycine are considered important for the growth of silkworms (Švácha 2013); on the other hand, Proline is essential in the sugar metabolism during flight (Carter et al. 2006; Lundgren 2009). In recent years, a “carcass study” of the proportion of amino acids, fats, or glycogen in arthropods has given valuable information for evaluating food’s nutritional content and consumption in relation to arthropods. The researchers concluded that the proportion of essential amino acids in each insect is of very much importance. Lindig, Hedin, and Poe (1981) concluded that analyzing the insect carcass might give information on the types of protein to include in an artificial diet that are responsible for healthy growth. Wider use of this technique to incorporate additional nutrients like carbohydrates and fats may be beneficial.

8.4.2 Micronutrients Vitamins, often known as growth factors, are classified based on their solubility in lipids and water. They are often needed for insect development since they are unable to synthesize them. Insects cannot synthesize sterols and therefore are necessary nutrients across all their members. Sterols perform several essential functions in insect physiology, intracellular membrane components,

8  •  Mass Production Technologies  163 hormone precursors, surface wax components, and lipoprotein carriers. Insects produce cholesterol based sterols, while plant phytosterols and fungus ergosterol are also important sources of sterols (Cohen 2003). Vitamins C and B complex are included in the water-soluble category, such as thiamin, niacin and nicotinamide, folic acid, pantothenic acid, riboflavin, pyridoxine, biotin, and cobalamins. The B vitamins are essential in the diets of all insects because they act as cofactors for enzymes, excluding B12, which is generally not essential (Švacha 2013; Cohen 2003). Predatory arthropods may take B complex vitamins through their prey, but non-prey items such as fungus can contain significant amounts of B vitamin (Lundgren 2009). Vitamin C has a considerable role in molting, and it is present more in herbivorous insects than in entomophagous insects. Vitamin C is also an antioxidant that may help detoxify and protect against microbial infestation (Cohen 2003). Choline and Inositol are phospholipid components. In addition to its structural function in phospholipids, choline has a role in spermatogenesis and oogenesis and is presumably necessary for all insects. Inositol in Coleoptera is recognized to be necessary and has a function in the neurological system. Carotenoids (A), calciferol (D), Retinol and phyloquinone are lipidsoluble vitamins (K). In insects, A and E vitamins play their role in pigment formation and reproduction. Vitamins A and E act as antioxidants (Cohen 2003). In non-prey foods like pollen and cereals, fat-soluble vitamins are typically present in large amounts (Lundgren 2009). While certain elements such as phosphorus, sulfur, nitrogen and iron may be derived from organic sources, other vital growth and reproductive elements (mineral) cannot be bio-synthesized from inorganic sources (minerals). Twenty-four elements are known to be necessary for life. They are listed below: Calcium, nitrogen, oxygen, chlorine, iron, hydrogen, phosphorous, sodium, silicone, magnesium, carbon, sulfur, potassium, iodine, manganese, copper, molybdenum, cobalt and zinc. According to their ionic charge, elements may be classified into cation (+) and anion (−). Metals such as magnesium, iron, potassium, sodium, manganese, zinc, calcium, and copper are included in cations. Phosphorus sulfur, iodine, chloride, and fluoride are included in the anions. Minerals are compounds made up of cations and anions. Some minerals are important for insects’ growth, such as phosphorus, potassium, copper, zinc, magnesium, sodium, and manganese. In many enzymes pathways, iron is essential, notably in DNA production. Calcium is needed to a lower degree in arthropods than invertebrates, although necessary for muscular excitation (Cohen 2003). For insects, trace elements have rarely been identified by nutrition or tissue levels (Cohen 2003; Nation, Sr. 2021). Therefore, trace element concentrations were rarely investigated for their effect on insect productivity. This contrasts sharply emphasis on protein, fat, and carbohydrates. However, trace elements’ nutritional and physiological significance has been well demonstrated in insects (Lavilla et al. 2010), along with their function in controlling gene expression via transcription factors (Zinke et al. 2002). The main cause for this gap in trace element knowledge and evaluation was the difficulty of detecting and precisely measuring concentrations. Recent advancements with ICPMS (inductively coupled plasma mass spectrometry) have given a great solution to determine several trace elements simultaneously at a range of low levels, making it the perfect technique for dealing with insects. Information on the dynamics between food trace element levels and food insect retention levels was a starting point for using ICPMS in trace elements in predatory heteropteran (Coudron et al. 2012). In addition, trace elements have a unique feature that is not produced or destroyed but of an essential physiological function, making them distinct from proteins, lipids, and carbohydrates. As explained below, these characteristics may help establish connections between diet and fitness using trace elements to make them good biomarkers. The natural diet is derived only from natural sources such as plant species or plant products. Insects are grown in the synthetic diet, and semi-synthetic and these dietary groups are divided into plant material and laboratory diets. More information about these diets is available in (Meyer-Rochow et al. 2021).

8.4.3 Plant Material The nutritional quality of host plants feeding insects influences the nutrient content of food-producing insects. When 10, 20, and 50% of Vermiwash (water washings of Earthworm cocoons) was sprayed on mulberry leaves and fed to fifth instar of B. Mori larvae, substantial increases were seen on the level

164  Edible Insects Processing for Food and Feed of dosage-dependent of fat, carbohydrates, and proteins (Purusothaman et al. 2012). In one case study, Rhynchophous phoenicis larvae were reared on the four organic substrates e.g., split watermelon, raw papaya, sugarcane tops, split pineapple, normal contents of fat, protein, and carbohydrates were found (Ebenebe et al. 2017). They found that split pineapple may be chosen as a possible source of protein and fluids for feeding larvae. Quaye et al. (2018) evaluated four diets based alone on palm yolk oil combined with banana and pineapple or millet residues. In R. phoenicis larvae fed on oil palm yolk, the most significant protein content (32 % and fibre (8.4 %) were observed. In field-collected sixth instar nymphs, Rutaro et al. (2018) utilized the inflorescence of four plants. They discovered that it had a high content (21%) of essential fatty acids than a low (12-13%) in less diverse diets. The overall lipid content and weight of grasshoppers did not differentiate across diets, although the fatty acid composition could be affected by food intake in an insect.

8.4.4 Laboratory Diet Laboratory or artificial diets have significant benefits over natural plant material in producing silkworms since they are semi-synthetic or synthetic and may be used for various insects. They assist in rearing insects that differ little from individuals and are subsequently made available for bioassays or other reasons whenever necessary (Meyer-Rochow, Ghosh, and Jung 2019; Meyer-Rochow et al. 2021). The artificial diet for R. differens was formulated using rice head, chicken egg busters, finger millet head seeds, simsim cake, wheat bran, sorghum seed heads, germinated finger millets, shea butter and dog pellets. More diverse diets have led to a higher PUFA and linoleic acid concentration. The content of fatty acids differs considerably across diets. The researchers concluded that the necessary fatty acid contents might be enhanced by feeding hopper to highly diversified meals, especially when it comes to mass rearing (Rutaro et al. 2018). Ghaly (2009) prepared the diet by weight of dry components (wheat bran + corn flour + dried powder yeast + whole wheat flour combined by weight 3:3:1) and of liquid component (honey mixed + glycerin by weight 1:1). In a 1:1 ratio, the two components were then combined. It was observed that diet contained Gonimbrasia belina and Anthoaera zambezina, was superior to plant material.

8.5  EQUIPMENT AND MECHANIZATION FOR INSECT MASS REARING 8.5.1 Production and Operation Management One of the major challenge in processing of food-producing insects involves striking balance which is the key to success between mechanization, labor, automation, expenditure, and productivity. In addition, technologies will considerably increase the elements involved in insect development, associated efficiency, production costs such as feed, irrigation, handling, collecting, washing, refining, packaging, and storage (Figure 8.2). The present edible-insect farms are not highly automated at a significant level, with two aspects: Overpriced items are accessible on the market and most of them are poor or varied quality. Insects can be only used as a protein replacement if growers can provide vast biomass, a consistent supply, a high standard, and reasonably priced meals for meat and fishmeal. Food insect farming could offer significant amounts at a cost per kilo competitive with other animal-related commodities once at least 80% of the industrialized production processes. The mechanism may encourage the partial substitute of manual farm labor, boost production, boost commodity quality, supply performance, and quality. An automated rearing

8  •  Mass Production Technologies  165

FIGURE 8.2  Operation and production management sketch of insects farming. (Modified/Reprinted with permission from Elsevier [Cortes Ortiz et al. 2016].)

FIGURE 8.3  Rearing is of mealy worm (Tenebrio bimaculatus), stacked boxes are overlapped above each other in a trolley to allow movement in the room. (Modified/Reprinted with permission from Elsevier [Cortes Ortiz et al. 2016].)

method of mechanization, automation, and information can be used in large-scale insect farming competitively and efficiently. The monitoring system can be enhanced to improve performance by managing the control system of mating, feeding, oviposition rate, environment, life cycle, and microbial control. The choice of equipment used in industrial mechanization is determined by the species being reared, size, location of farm, economic position, and the most relevant environment in the region.

8.5.2 Rearing Area The resting area is very much important in the efficient production of insect production. For certain insects that can move vertically, leap or fly, 3-dimensional room for rowing can enable greater density, relying on just two-dimension flat spaces (cardboard separators, egg cage material, or a longer-lasting grid of some material or other). Reduce the distance between trays containing insects and feed while balancing the

166  Edible Insects Processing for Food and Feed need for air and temperature control for those confined to the ground of a flat surface (fly and mealworms larvae, etc.). Furthermore, feed/substratum depth is vital in such insects’ trays to not add too much waste because certain insect species can use food and living room at a certain depth. Boxes may be stored in multi-level racks packed with many raising boxes as possible to reduce spatial use of kilograms of insect generated minimizing the room required to farm maximum insects. In certain situations, it is possible to use stackable boxes or a pallet to facilitate free mobility across the rising field (Figure 8.3). An example of a trolley raising space to hold the boxes of a mealy worm species. In the rearing sector, the climate control device should be able to sustain acceptable environmental conditions.

8.5.3 Feeding and Watering Feed handling is an important component of the entire system for any kind of farm. This also applies to industrialized food insect farming (Figure 8.2). To minimize feed spoilage, insect farmers must serve on time (pickup and delivery: one or two time in a week) because of the perishable nature of certain feeds utilized, particularly those having high content of water, like those made from non-marketable food items. Spoilage is less important, when feed formulated by using dry material like wheat flour, yeast, spent grains and cake of dry oilseed. Even yet, only necessary feed should be provided, and excessive or wasted feed should be avoided, especially because feed is frequently contaminated by insect faeces, saliva, secretions, and other environmental contaminants. It’s rare for feed to be recycled or reused once it’s been supplied to farmed insects, but it’s possible in some well-designed and highly efficient systems for species that require extremely dry circumstances. Since the type, amount, format, and speed of feed production rely significantly on the species being cultivated and for its lifestyle, it is essential to employ a feeding and identification system that delivers the appropriate food at the right time and right stage. The tower silos may be utilized with mechanical downloaders, with many distribution units and arms to reach each shell in a mixer wagon; a central computer robotically directs the wagons to locate the arms to be supplied. A computer software utilizes an array for each location on the shell to guide the wagon about the location in which feed will be delivered. Computerized systems may be updated by technical production observations utilizing wireless communication networks through enterprise resource planning and tablets or connected techniques. This data would be limited to a set of value-scale parameters that might assist in making objective decisions. Data from different elements (humidity, temperature etc.) might be sent to a computer in a computerized insect rearing system that can be assessed using comparisons of input, microclimate, population growth, analysis of incidents and historical data. Thus, a PID controller (an instrument used to regulate temperature, flow, pressure, speed, and other process variables in industrial control applications) system may take autonomous choices that reverse it into warmer rooms, development production speed up, and vice versa. In fresh conditions collected biomass or serious alterations, which reduce performance or incidence of illnesses. One of the greatest benefits of insect species production is lower water requirements than vertebrate animals. For certain species, fresh vegetables and fruit may supply water directly in their diet, while others need a small quantity of water, regardless of diet (if the feed they’re using is dry). A low-pressure water system must be established to supply the amount of water required for each raising container for those who require water separate from their food. In most circumstances, more is not needed to ensure that the insects do not drown in it, that the water not flood the insect habitat, and that excessive microbial development does not occur.

8.5.4 Separation and Sorting As with other agricultural products, traceability of edible insects, from farm to table, is important to ensure safety and track risks or disease spread to their source. The harvesting of each new product of

8  •  Mass Production Technologies  167 insects species must be carefully cataloged according to the raw materials used for feeding and water utilized (Figure 8.2). This data should be used to follow the insects through each stage of the production process, including quality inspections at each important point in the system. There should be pre-defined and selfcontained mechanisms that allow for the tracking of a product’s history, location, and travel through the supply chain at any given moment. Radio-frequency identification (RFID) tags that provide a minimum of information can be used to describe all goods or batches: date and time of batch commencement, lots of raw material utilized, parenting precedence, life stages, density, date, cleaning, and feed elements. An integrated traceability management system is possible to have individual control of each manufacturing lot, inventory control and ending product, incident detection, complete monitoring of lot, and critical points of system. In this way, the quality and safety of items can be ensured. The same management system can take manufacturing orders, handle packing lines, and remotely control production using conventional bar code issuance, validation, and reading. These bar codes are essential for the traceability and usage of standards in the food sector. For such machines (Figure 8.2) for handling cricket species like A. domesticus or G. assimilis, a mechanical arm is applied to each box placed on the conveyor belt, which takes the rearing substrates out of the boxes. It collects crickets gently to the area where they are softly brushed with a circular, soft brush. The suction is designed to collect the cricket without harming it. The crickets are then trapped in a net receiver where they may be handled, weighed, and transferred for further operations. Throughout the separating phase, the feed conversion is automatically estimated by measuring the quantity of feed taken by the insects: the weight gain and production of frass by insects, respectively. After the separation process, simple electrical logger cells can record pre- and post-feeding insect weights.

8.5.5 Cleaning Room In certain instances, insect washing, or rinsing is required. It may even be desired after harvest (although this will be unnecessary in many cases, mainly if they are instantly frozen or heated upon harvesting). In some cases, insects may be raised in the same container as the animal throughout harvest; therefore, removing insects from the container is unnecessary before harvesting. However, in some cases, rinsing or cleaning of collected insects, or moving them across containers before harvesting throughout their life cycle, may be useful. In such circumstances, both boxes should be sent to a high-pressure cleaning following separation, packed with fresh substratum, and returned to the separation location (Figure 8.2). For all rearing boxes, 166 insects species can be produced for easy management and mechanically durable materials such as organic feeding supplies which would be washed numerous times. Clean raising boxes with cleansed insects and new raising substrate. Edible insects are separated by age and scale and returned to the rearing area or packed for transfer into the processing center. Dead insects and garbage are treated regularly by an official entity for cleanliness purpose.

8.5.6 Dung Area After producing and completing all the edible insects’ processes, dung/manure then becomes the most concerning factor. Residues should be collected, processed, and stored for additional uses such as fertilizers and other animals feed in a separate area from the rearing and cleaning (Figure 8.2). Furthermore, by using carbon-rich feedstocks as a straw, various applications can stabilise insect species’ frass. In contrast, the amount of carbon depends upon the frass of insect species. It is moistened and maintained pH of approximately 7.5, piled, and turned carefully after 48 hours (low and rise of temperature becomes the main factor during this process). Then, processed excrement of insect larvae is packed using a flour machine and bags for the shipment.

168  Edible Insects Processing for Food and Feed

8.6  PRODUCTION AND PROCESSING TECHNOLOGIES BY SPECIES 8.6.1 Black Soldier Fly Compared with traditional animal or poultry production, insect farming has been regarded as a promising option because it demands little land and water. The primary operation of the system in BSF is the preprocessing of the substrate material of waste feed. The black soldier fly, H. illucens, has numerous unusual characteristics, establishing it as a model strategy for waste reduction as a source of protein. Black soldier fly prepupae may be fed for various species, including catfish, rainbow trout, chickens, and swine (St-Hilaire et al. 2007; Sealey et al. 2011). Black soldier fly larvae have a high concentration of lipids and oils that can feed animals (Makkar et al. 2014; Henry et al. 2015; Spranghers et al. 2017). Except for Antarctica, this species may be found on almost all major continents, abundantly present in tropical and temperate regions. Due to the prevalence of black soldier fly larvae in waste, other insect pests, such as the house fly, are often excluded. Additionally, larvae of black soldier fly can decrease pathogens found in waste, such as Escherichia coli and Salmonella spp (Erickson et al. 2004; Liu et al. 2008). However, it should be noticed that the production and capacity for reducing various wastes are specific to the population (Zhou et al. 2013). Establish a colony from a local population if interested in black soldier fly industrialization. This technique frequently eliminates the possibility of utilising a strain of black soldier fly from another part of the world that is susceptible to the local environmental circumstances (Cortes Ortiz et al. 2016). BSF has the complete metamorphosis stages (holometabolous) and has a life cycle of approximately 45 days (Ferrarezi et al. 2016). The color of BSF at the larval stage is white, while it becomes black at the getting stage of adult, which makes a clear difference in both stages (Sheppard et al. 2002) (Figure 8.4). This section will study the black soldier fly’s life history, mass protein and fat production usage. Furthermore, life cycle, adult and larval colony maintenance, mating, and oviposition have been discussed. Still, unfortunately, we must say that all the implementations has not been enhanced on the large-scale production, although experimentation has been done in the laboratories.

8.6.1.1 Adult colony and its management Too little is understood about the actions of adult black soldiers. The main cause for this is their shortened lifetime and their tendency to spend much of their adult lives in less stable environments. Rarely males come back to development sites; there is no evidence that this species is eaten or drank by adults.; thus, females are only present at such areas to oviposit. Adults are less susceptible to spreading infections from diverse substrates they encounter to consume, making them more hygienic than other flies. Females die soon after laying their eggs. As a result, dead adults collect at the cage bottom. Few details on the black soldier fly’s adult life history are available. On the other hand, it is critical for mass colonial development. Males usually first appear with the females after two days (Tomberlin, Sheppard, and Joyce 2002), and colonies are maintained in the cages of 1 mm3. Additionally, sunshine influences mating activity (Tomberlin and Sheppard 2002). When sunlight is limited in some regions and its strength is reduced, this aspect can be extremely constraining during the winter. Attempts have been made, however, to address this constraint by the development of artificial light systems. Additionally, temperature and humidity must be considered (Holmes, Vanlaerhoven, and Tomberlin 2012). Temperatures below 27°C cause adults to become less active, translating to less oviposition rates and mating couples (Tomberlin and Sheppard 2001).

8  •  Mass Production Technologies  169

FIGURE 8.4  Life cycle of H. Illucens.

FIGURE 8.5  Black soldier fly A) Mating/Love cage B) Egg-laying of female. (Modified/Reprinted with permission from Eawag/Sandec [Dortmans et al. 2017].)

8.6.1.2 Mating and oviposition The reproduction of H. illucens is biphasic (i.e., mating and oviposition) and has been studied partially (Tomberlin and Sheppard 2002). Breeding continues year-round in the tropics under normal circumstances but is limited to a few generations in warm temperate areas (Tomberlin and Sheppard 2001). Seasonal variation influences mating behavior, most notably the shortening of days and the correlated light intensity (Park et al. 2010). Male territorial activity has been reported to escalate into a fight when a male congener approaches, whereas females initiate breeding efforts (Tomberlin and Sheppard 2001). Adults have been observed forming aggregation sites where males attempt to conquer and mate with females in flight.

170  Edible Insects Processing for Food and Feed Due to this action, colonies must maintain in cages measuring a minimum of 1 m3 and some illustrations of its different stages are presented in (Figure 8.5). Females can lay eggs in an oviposition site two days after mating if volatile organic compounds are released from nearby decaying organic matter (Tomberlin, Sheppard, and Joyce 2002; Zhou et al. 2013). Eggs are deposited in dry interstices near a supply of moist food consumed by future larvae (Tomberlin, Sheppard, and Joyce 2002). Environmental conditions influence the reproduction steps. For example, Tomberlin and Sheppard (2002) revealed that the intensity of light controls mating periods, while oviposition was regulated by humidity and temperature (Tomberlin and Sheppard 2002). Presently, efforts have resulted in systems capable of generating roughly 60% of mating under natural light conditions (Zhang et al. 2010). Temperatures over 26°C were found to support oviposition, allowing for the development of artificial breeding systems to produce H. illucens during the year in temperate regions. Rearing in an artificial environment generally requires breeding techniques in a controlled environment deprived of sunlight (Park et al. 2010). Oonincx et al. (2016) demonstrated that unique wavelengths of LED lightening (LED ratio UV: B: G = 1: 1: 3) stimulated the eyes of BSF, ensuring survival and egg viability, while Heussler et al. (2018) reported that LED light was the perfect lightening for promoting flies longevity in smaller breeding cages in comparison to halogen and fluorescent lamps (Oonincx et al. 2016; Heussler et al. 2018).

8.6.1.3 Production of larvae and its maintenance Gradually, as the number of BSF adults ovipositing increased, the amount of prepupa harvested increased as well. As a result, the amount of non-BSF flies sighted on the medium decreased gradually. It is likely that when the larvae hatched, they released species-specific chemicals that attracted more females to the attractant. This resulted in increased egg production, which increased the prepupa yield (Nyakeri et al. 2017). Additionally, few mature larvae developed initially may have matured and pupated to become adults, boosting BSF numbers enough to outcompete other fly species. It is assumed that the avoidance of other fly species in the medium is due to the larvae of black soldier fly (Erickson et al. 2004). Because BSF larvae excrete antimicrobial compounds (Park et al. 2010) that enhance discourage of other fly species in the medium (Erickson et al. 2004; Zheng et al. 2013). The variability in larvae harvested may be described by variations in the frequency and quality of odors formed by decomposing food, as well as the nutritional quality of the substrate materials (Nyakeri et al. 2017). As eggs hatched approximately after four days, the temperature can impact the viability and hatching time (Tomberlin, Sheppard, and Joyce 2002; Holmes, Van Laerhoven, Tomberlin 2010). Eggs normally should be kept on cardboard in a tiny jar (500 mL transparent plastic cup) enclosed with a towel of paper and safe with a rubber band. In the colony larval room, the cub can be placed on a shelf. This area must be maintained at a constant temperature of approximately 27°C, relative humidity of 60-70%, and an L:D ratio of 14:10. Temperature is a dynamic characteristic which should be carefully monitored because it has an impact on larval growth (Tomberlin, Adler, and Myers 2009). In the past Gainesville stable diet (Gainesville fly diet is a dry mixture of pelleted peanut hulls) was used. After adding water, the diet remains for 30 minutes before processing to ensure proper pellet breakdown. The diet (then re-mixed, and eggs were added) was used as feed due to its high production at a minimum cost, and this feed was later on fed to the colony of larvae. The value of food given to the larvae for 2-4 days depends on its number. After the larvae have digested the initial feeding allocation(s), they would be transferred to a larger container 24 cm long, 12 cm deep, and 13 cm broad, and fed libitum until it is about half-filled. From this onward, the larva can be a shifted to a big container having a size long (76 cm), deep (12 cm), wide (45 cm) and let them remain until pupa stage (Cortes Ortiz et al. 2016). When approximately 40% of the larvae turn into pupae, their feeding is stopped (Tomberlin, Sheppard, and Joyce 2002). The container should be covered and secured with gauze to prevent emerging adults from moving out of it. Any kind of infection has the potential to cause a major die-off in the trough. Because all age groups are mixed, determining the quality of the product may be challenging. Additionally, a substantial amount of capacity is needed to operate such buildings. This constraint may also be a source of concern when it comes to heating and cooling costs.

8  •  Mass Production Technologies  171

8.6.1.4 The feed used for rearing BSF can consume diverse types of waste materials. They can develop on the manure of cattle’s (Myers et al. 2014), municipal waste (Diener et al. 2011), kitchen wastes such as raw vegetables and fruits (Nguyen, Tomberlin, and Vanlaerhoven 2013, 2015) and human wastes (Banks, Gibson, and Cameron 2014). Additionally, previous research has shown that when plant based substrates are used for BSF larvae, they develop faster than animal substrates (St-Hilaire et al. 2007; Tomberlin, Adler, and Myers 2009). Although black soldiers fly larvae may eat various waste products and the type of waste can affect their development time. As a result, during the formulation of waste stream for larvae to increase productivity and output, care must be taken. Probiotics could be one possible alternative.

8.6.1.5 Costs and quality maintenance The costs of maintaining an adult colony are extremely low. If necessary, adults should be watered; though, an automated misting process will eliminate the need for additional staff. The true costs of colony management are removing dead individuals, controlling oviposition, and reintroducing adults in colony. After discussing production manuals, it is mandatory to ensure the quality of larval and pupae. Because it feeds on various waste materials that can contaminate the quality of insects production (such as microbial and heavy metal contamination can be a question if they are not maintained). On the other hand, there is a need to ensure the nutritional value during the production process.

8.6.1.6 Impact of different factors on the growth of BSFL Researchers described different factors involved during the rearing and processing of BSF, such as physical, chemical, and others. BSF has the potential of manipulating acidic and basic conditions and showed resistance to different environments. The feeding system exerts influence on the growth parameter of BSF, while substrate pH does not affect mortality. Batch-fed BSFL has more growth than daily fed BSFL (Meneguz, Gasco, and Tomberlin 2018). The influence of different factors and their features has been listed in (Figure 8.6).

8.6.2 Crickets 8.6.2.1 Production of crickets worldwide The Indian cricket, G. sigillatus and the house cricket, A. domesticus, are two most commercially produced cricket species in the United States. Many farmers of cricket have recently adopted Gryllidae assimilis, Jamaican field cricket, to substitute colonies of A. domesticus damage that caused by the densovirus infection (Weissman et al. 2012). In united states crickets are produced at commercial level at least for 65 years. For exotic species including amphibians, mammals, reptiles and birds, they are often utilised as food and fish bait for fish species. In the previous decade, United States developed markets for as cricket based food source for human beings. Crickets are dried and crushed into fine powder form or in course meal (often called “cricket flour”) which are added to various industrial food items. The methods of crickets production are unreliable and labor-intensive, despite the fact that crickets are produced in the US for decades. The rearing mechanisms are performed manually, and no proper equipment has been designed to accelerate the procedure.

8.6.2.2 Rearing units In order to expand the surface area, crickets are produced in big boxes that are 36–60 cm deep and filled within an array of substances, such as egg cartons or shipping dividers. Rearing boxes made up of a

172  Edible Insects Processing for Food and Feed

FIGURE 8.6  Impacts of different factors on the BSFL growth performance.

variety of materials, such as wood box or cardboard, polyethylene, metal, and fiberglass. Top layers are preferable since crickets find it difficult to climb the walls. Cardboard or Wooden boxes must be coated with a substance which is hydrophobic, like a plastic bag, to prevent the moisture from cricket waste or watering devices from destroying the box, particularly during usage. To prevent escapes, the top edges of wooden or cardboard boxes must always be coated with Teflon, smooth substances like cello tape. The dimensions vary from 50 to 160 cm in length and 40 to 100 cm in width. Larger boxes provide more rearing surface, reducing number of units to use, but larger boxes require more work to handle. Although, the handling of smaller boxes is easy but need more units and maintenance workers. Water is supplied using industrial poultry watering systems that have been modified with sponges to avoid drowning. Some farmers have created their irrigation systems, which need less management. However, automated watering systems have generated leaks, flooding the raising unit and drowning the crickets. The present crickets rearing systems are closed with solid bottoms that do not allow the outflow of extra water. Future automated watering system designs should address this issue, as well as improve the rearing boxes.

8.6.2.3 Diets and feeds Food supply is carried by different method like filling material stack in paper plates or shallow trays and at the top of egg cartons. Mixtures of cricket feed are commercially available, but many farmers create their unique feed blends. Patton (1967) determined that meals comprising 32-47% carbohydrate, 3.2-5.2% fat and 20-30% protein were most successful in rearing A. domesticus. Despite this, many commercialized cricket feed mixtures claim crude protein levels of up to 20%. The ingredient’s cost is a barrier to creating the commercially sustainable diet compositions. Nakagaki and Defoliart (1991) concluded that conventional feeds of chicken produced cricket biomass at significantly reduced cost than diet. The composition of protein was 22% utilized by Nakagaki and Defoliart (1991) in chicken feed. Increasing demand of crickets as food or feed requires the production of diets that are more effective, cheaper and healthier. This should aim to make underused biomass like wasted grains, ethanol operations, algae, food industry by-products presently unused biomass like different grasses (i.e., switchgrass).

8  •  Mass Production Technologies  173

8.6.2.4 Environmental conditions Crickets need greater temperatures for rearing than other frequently produced insects. The optimum temperature range for A. domesticus development and growth is between 32 and 35°C. Temperatures in cricket farming regions have been reported to vary between 29.5 and 33°C. In A. domesticus, the incubation period for egg lasts 10 to 14 days and development of nymph remains for 6 to 8 weeks (Patton 1978). The optimum conditions of relative humidity for A. domesticus development and growth are unknown, although Clifford et al. (1977) describe range of 25-50 percent as acceptable. Few commercial producers describe utilizing relative humidity levels ranging from 33 to 45% in their cultures. According to Krieg (1998), relative humidity of 55% is sufficient to keep crickets from collecting near the water source while keeping the food dry and unsuitable for microbe development.

8.6.2.5 Reproduction Adult crickets are moved to fresh cages to reproduce. The use of cardboard box separators in place of egg cartons provides a more rearing area for adult crickets. Cardboard box partitions give large open areas for the adult crickets to walk. Water and food are kept at the top of the partitions. Fertilization occurs in trays or pans loaded with a wet oviposition medium and put on top of box separators. Females prefer moist surface area for oviposition with sufficient depth to hide the eggs. Substrates used for oviposition comprise vermiculite, cotton, peat moss, sand, and coconut husk. To obtain a better oviposition substrate, the above-mentioned components are mixed in different portions. The substrate of oviposition for cricket must be flexible enough to allow female to burrow with ovipositor while still retaining moisture. In US, the utmost oviposition medium in crickets farm is peat moss. The oviposition material is placed in shallow containers of (5 to 7.5 cm deep) trays and the mature crickets are acclimated to it for 48-72 hours. As previously stated, the development of an egg takes place about 10-14 days (Clifford et al. 1977). Near hatching the oviposition trays are transferred to fresh rearing boxes having water and food which maintained in same climatic factors until the most eggs hatch. When the first instar develops in the box, the oviposition food sources, water, and trays are taken off the box. 1st instar batches are separated into new rearing units. The volume is used to estimate the number of first instars in each batch (34 ml equals 10,000 and 12,000 first instars). Weight may also be used to predict the number of initial instars. The average weight of 1st instar is around 500 µg. In this way first 200 instar crickets’ weight is about 1 g. Across the commercial cricket farms the density of rearing differ significantly. Some farmers estimate concentrations of 500 to 750 crickets in a 28.3 L rearing area. The (30×30×5 cm) size six stacking egg cartons yields this amount of rearing space. According to Lundy and Parrella (2015) estimations of 1,800 cm2 area available on each 30×30 cm egg carton, has the rearing space area equal to 10,800 cm2. Considering these calculations, the volume of cricket nymphs per area raised in commercialized farms of crickets in US ranged between 4 to 7 nymphs/dm2, equal to 100 cm2. Other farms use cardboard partitions instead of egg carton material, which could allow for harvesting in less time. Adult crickets that are used for the reproduction of colonies are often marketed as fish bait. When females’ reproductive production is reduced, adults are captured and packaged for sale. Commonly, all watering, oviposition and feeding systems are disconnected from the rearing box while harvesting adults crickets nymphs. The large cardboard partitions are removed by shuddering the immature crickets into the rearing box. After determining the average volume filled by 1000 adult crickets, adult cricket counting is determined by volume, as stated for the 1st instars. Crickets are captured from cartons by enabling them to Mount paper or Wood containers, which are then stirred to cause crickets to drop in a tube linked with volume measurement box. Crickets are typically marketed in thousands and supplied in boxes with paper wrapping material. Crickets should be sold frozen and priced per pound for food or feed component in the market. Assume the crickets are being gathered for the insect-based food sector. In that case, they are usually refrigerated and packaged in food-grade paper bags before being placed into cartons, which are subsequently stored in refrigerators. For shipment, the boxes are usually placed on standard-sized boxes and wrapped in situ before being transported to the treating site by refrigerated truck. A farm that serves the insect-based food sector would have adequate freezer (Walk-in) space to store entire pallets.

174  Edible Insects Processing for Food and Feed

8.6.3 Mealworm The mealworm Tenebrio molitor beetle is a European native that is now found all over the world. It has a considerably longer lifespan than HI or M. domestica, varying from 280 to 630 days (Makkar et al. 2014). The T. molitor (TM) lifecycle has two basic stages: mealworm and adult. At room temperature, larvae hatch in 10-12 days and develop in 8-20 stages over 3-4 months. Mature larvae are pale yellow-brown in appearance and may survive up to 18 months. They are usually 20-32 mm in length and weigh 130-160 mg. If the temperature is decreased, the pupal stage may take up to 20 days. Adult TM lives 2-3 months (Makkar et al. 2014). TM and its larvae are pests of grain, flour, and food markets. They are omnivores and quickly recycle organic food waste into a high-calorie protein- and fat-rich diet (van Broekhoven et al. 2015). Mealworms are high in protein (47-60%) and vital amino acids. They also have a high-fat content (31-43%). Except for lauric and linoleic acids, which are much lower in TM and the fatty acid content is similar to that of the M. domestica larvae diet. Minerals such as P, K, Na, Mg, Fe, and Cu are abundant in mealworms. However, they have lower levels of Ca and Mn than HI (Makkar et al. 2014). Mealworms were first commercially produced in the United States of America in the 1950s. Mealworms were first used as fishing bait, but they were quickly adopted as a source of nutrition for Zoo animals and exotic pets (Tran, Heuzé, and Makkar 2015). Mealworms have lately shown their ability to partially substitute fishmeal in the diets of several fish species (Li et al. 2020). The utilization of substantial TM meal contents has also received a positive response. Mealworms, for example, are extremely appealing and may replace up to 40% of the fishmeal component in African catfish diets without a noticeable decrease in growth performance or feed efficiency ratio. Furthermore, diets that replaced fishmeal up to 80% of the time showed excellent development results. Rainbow trout have also been successfully fed TM meal (Makkar et al. 2014), replacing up to 50% of their nutritional requirements (Ajani, Nwanna, and Musa 2004). Interesting findings were also achieved for marine carnivorous species like seabream (Ajani, Nwanna, and Musa 2004) and European sea bass (Wang et al. 2017), where fish development was not affected substantially by fishmeal substitutions of 25%. It is possible to feed small European sea bass 25% of TM larval diet for six weeks and significantly reduce inflammation while also improving anti-parasite activity (van Broekhoven et al. 2015). It is also worth noting that the inclusion of a full-fat TM larvae meal does not negatively affect most quality traits of the fish flesh of rainbow trout, except for the fatty acid profile (Gasco et al. 2018).

8.6.3.1 Rearing Mealworms are traditionally produced using trays to retain larvae as they grow and adults to reproduce the colony. The most often used tray size is 65 L x 50 W x 15 H cm, lightweight and deep enough to keep larvae and adults confined (Cortes Ortiz et al. 2016). In recent years, MoralesRamos et al. (2012) introduced the screens tray bottom system to fall below frass particles. Adult holding trays may utilize bottom screens to let early instars fall into the shallow tray at the bottom to be collected. T. molitor is grown using stackable pots with nylon screens at the bottom. Larvae are kept in containers on a diet of wheat bran and other supplements. Five or more containers may be stacked on a trolley for transport (Figure 8.7 A). The holes in the screens are 500 m (0.5 mm) in size, such that only a little amount of food is lost due to frass particles passing through. With each new generation of larvae, a new container is filled with frass particles, which descend continually until they reach a collecting container at the bottom, finishing up in a pile. Mite infestations can be managed, and larval food intake can be observed more precisely by eliminating frass from containers. Changes in the colony’s consumption of food may be utilized to improve productivity or to identify issues. The opening of the screen will accommodate larvae up to the fifth instar based on measurements taken from the head capsule. Before transporting the

8  •  Mass Production Technologies  175 larvae to the screened containers, place them in containers with solid bottoms for 4-5 weeks at 25–28°C to prevent them from drowning. In addition to harvesting offspring from adult holdings, screen bottoms may be utilized. Female T. molitor oviposits on the trays’ bottoms, bonding the eggs together with a sticky substance. Eggs are frequently stuck to food particles which are seemingly insurmountable obstacles to remove without causing damage. To avoid egg cannibalism in a tray system, trays must be changed on a regular basis to remove adults from the oviposited substrate. Adult T. molitor is placed in a container with a screen on the bottom that sits on a shallow tray top (Figure 8.7 B). It has an opening of 850 mm, which allows tiny larvae to reach the bottom collecting tray. First to sixth instars may pass through the aperture of the screen with hardly any food loss (MoralesRamos and Rojas 2015). A number of variables may impact mealworm rearing systems’ productivity, and techniques must be developed to identify and manage these aspects. Larval densities sustained throughout production is one of these variables. Many tenebrionid species’ pupation is hampered or delayed by high larval numbers. Increased larval density may also affect growth rates and decrease pupal mass, resulting in tenebrionids developing more slowly, dying more often, and engaging in cannibalism. The physiological condition of larvae may be impacted as a result of increased crowding.​

8.6.3.2 Feed Compared to a control diet utilized by commercial mealworm producers, larvae on diets higher in protein had greater survival and shorter development times. In contrast, those on a diet LPHS had poorer survival and longer development times. According to many studies, high-protein diets accelerate mealworm development, especially those produced from yeast (J. A. Morales-Ramos et al. 2010; van Broekhoven et al. 2015). Their adult stage can be produced by supplementing a simple diet while larvae development is required wheat bran (J. A. Morales-Ramos et al. 2010, 2011; Morales-Ramos, Rojas, and Shapiro-Ilan 2014). Van Broekhoven et al. (2015) described that the supplementation containing slices of carrot, potato and cabbage are most commonly used in the USA. Wheat bran may be devoid of minerals, including vitamins, vital fatty acids, and sterols, but vegetable supplements may help to make up for those deficiencies. On the other hand, it can also balance the requirement of water in the diet. Adding uncooked vegetables may also serve as a water supply for larvae. T. molitor diets must have their nutritional ratios balanced. T. molitor’s survival and growth rates can be enhanced by supplementing with protein, although this will increase the expense of the food. It is preferable to supplement high-protein by-products (J. A. Morales-Ramos et al. 2011; Morales-Ramos, Rojas, and Shapiro-Ilan 2014; van Broekhoven et al. 2015). Eating a lot of calories reduces an insect’s capacity to fight against diseases. T. molitor, for particular, becomes more susceptible to entomopathogenic nematodes when its diet is higher in lipid (Shapiro-Ilan et al. 2008; Krams et al. 2015).

8.6.4 Housefly The housefly (Musca domestica) is the most promising farm insect in China. Because housefly maggots can grow on a range of substrates, they can be utilised to convert wastes into beneficial biomass high in protein and fat and suitable for animal feed, with the residues being used as organic fertilizer. Fly larvae, which are particularly promising among insects, can be created inexpensively and rapidly using organic waste material (Pastor et al. 2015). Housefly (Musca domestica) is considered as the replacement of fishmeal (Fasakin, Balogun, and Ajayi 2003; Ogunji, Schulz, and Kloas 2008), poultry feed (Adeniji 2007; Pieterse and Pretorius 2013), and management of waste matter (Čičková et al. 2015; Roffeis et al. 2015). In West Africa, house-fly larvae are frequently produced by exposing organic substrates to natural adult flies (Kenis et al. 2014). Larvae only take 3 to 4 days to mature in tropical climates. The housefly is a holometabolous insect, which completes all the stages like egg, larva, pupa and adult. Its life is impacted by the availability of food and a favourable temperature. A female lays 4-6

176  Edible Insects Processing for Food and Feed

FIGURE 8.7  Mealworm production trays (A) for larvae (B) for adults. (Modified/Reprinted with permission from Elsevier [Cortes Ortiz et al. 2016].)

FIGURE 8.8  Life cycle of Musca domestica.

hatches consisting of 75-150 eggs. Its lifespan is about 15 to 30 days, and the female is ready to mate when it is three days old, and the male can mate after its emergence. The color of house fly eggs is white; when entering the pupal stage, maggots develop dark color, hard outer shell, other stages of color formation, and the life cycle is described in (Figure 8.8).

8.6.4.1 The feed and its maintenance in rearing The water content for fly larvae in the standard rearing is always 65-70%. Because of the high-water content of animal manures, wheat bran or rice bran must also be added to the manures to change water content. Animal manures are fermented to kill any heat-sensitive bacteria present (Zhu et al. 2012; Chen et al. 2014). Adult houseflies are typically reared on a diet of 50% milk powder, 50% glucose, whereas there are many alternative media compositions for the rearing of larvae such as pig dung, wheat bran and chicken manure (Chen and Feng 2002). Hogsette (1992) described manure-free feeding compositions for housefly larvae. In the 1990s, Ralston Purina (St. Louis, Missouri) produced a commercial diet (CSMA) that was popular in the United

8  •  Mass Production Technologies  177 States, but this formulation is no longer available. A 33 percent wheat bran, 27 percent meal of alfalfa, and 40 percent granules of brewer’s yeast were used in this composition (Hogsette 1992) Adding cornmeal while removing yeast granules resulted in effective and cheaper diet (89.5 percent yield): 50 percent wheat bran, 30 percent alfalfa meal, and 20 percent cornmeal. Case study: Fresh dairy cow dung was collected directly from the Cornell Teaching Dairy Barn floor and brought to the laboratory for instant experiments kept at 4°C and utilized for 2-3 days while keeping its 85 % moisture contents. The experiments related to larval growth and degradation of waste were accomplished inside the environmental chamber (L: 300 cm, × H: 250 cm × W: 300 cm) with a controlled temperature of 28℃, RH of was about 70%. This experiment was done in 12 hours ambient light photoperiod and 12 hours darkness. Without eliminating the larvae from this chamber, egg hatching, larval development, and behaviour were observed. The experimental goals determined the duration of incubation and the kind of the assessments. The egg suspension, as produced above, was uniformly dusted over the waste for larva feeding nutritional analysis; 50 mg (10 ml suspension) was utilized for 250 g of dung in plastic containers (L: 30 cm x H: 8 cm x W: 12 cm). Larvae were collected before pupation at the third instar after 5-6 days of development, washed with water, dried, and kept at -20°C (Hussein et al. 2017). The adult life span of house fly larvae fed with food waste was 30-50 days under conditions of temperature 25.0 to 35.0°C and relative humidity 50.0 to 60.0 percent. The oviposition phase is 10 to 25 days. The parameters that influence the rate of oviposition of adult flies are adult feedstuff, the density of adult flies and oviposition bait (Čičková et al. 2015). The density of housefly adults in this research ranges from 30.0 to 50.0 thousand, and daily egg production may range from 2.00 to 6.00 g/day, allowing for a followup investigation. The housefly eggs may hatch 6 to 10 hours late, and it is the most important stage of the housefly existence (Zhu et al. 2015). The larval stage’s development degree influences the size of an individual, efficiency of reproduction and even life duration. Furthermore, the ratio of culture substrate had a favorable impact on the volume and output of 50 larvae. The protein level of dishes trash is greater than that of basic meals, and these proteins may stimulate animal growth and development (Titi et al. 2000; Wen, Wang, and De Clercq 2016). The contents of crude protein and crude fat in the culture substrate rose as the percentage of dish waste. Therefore, the power and consumption efficiency in the substrate increased, which became favorable to the development and nutritional consumption of housefly larvae.

8.6.4.2 Process of production The flow process system (Figure 8.9) is given as an example of the rearing system of the housefly. The rearing room designed to produce houseflies contains (25-32°C, RH = 65-75%), to provide aeration, 8-h light, and prevent escape. A lengthy pipe is inserted into the fly cages, and water is carried from water tank, where the adult flies may consume the liquid. Another method is to place an inverted cup (full of water) on a filter paper-wrapped plate. Water is constantly soaking into the paper. Adult densities of approximately 5625 individuals per cubic meter of production chamber area are optimum (1:1 sex ratio). When adult houseflies reach the age of three days, a patch of cotton fabric is kept over moist wheat bran (water content 60-70%) to serve as an oviposition medium for egg collecting. The fragrance of fermented wheat bran attracts adult flies to these oviposition locations. Each day, between 09:00 (9:00 am) and 16:00 (4:00 pm), eggs are collected for 5-10 days. Adult oviposition duration can be prolonged by supplementing the adult diet with chicken eggs, ultimately increasing eggs production. After collection, fly eggs are deposited on the surface of the diet medium in containers at a density of approximately 1 g of eggs per kilogram of food medium. After 5 or 6 days, at a temperature of 25-32°C and a relative humidity (RH) of 65-75 percent, the eggs hatch into larvae. Several techniques are used to remove the larvae from the food medium, including negative phototaxy and self-escape, decreasing oxygen concentration and screen mesh (Jia 2007). Within the raising chamber, mature larvae may also be combined with dry wheat bran for pupation, which appears after 1-2 days at 25-30°C temperature and RH of 65-75%. The pupae are moved to the adult rearing rooms. When larvae are collected from the medium for use in a specific function, they can be processed in any

178  Edible Insects Processing for Food and Feed

FIGURE 8.9  Flow chart process scheme for the production of housefly. (Modified/Reprinted with permission from Elsevier [Cortes Ortiz et al. 2016].)

way necessary, such as dehydrating in microwaves to make larval powder. Dry maggots can be stored sealed for 4-6 months at room temperature (25°C). The cost of growing housefly larvae is estimated using the feeding medium’s composition, rearing scale, and labour costs. According to the retail pricing on September 15, 2015, the projected price of dried maggots is about 9.42 USD/kg. Like other insects, the study of dietary formulation in the housefly remains an available domain for research. There is considerable demand for the usage of industrial and agricultural products, as there is tremendous versatility in the consumption of houseflies-based food.

8.6.5 Waxworm The larvae of larger wax moths, G. mellonella, are commonly referred to as waxworms. The larger wax moth is a major pest in Apis mellifera L. hives, eating honey, pollen, and larvae in active bee colonies. In contrast, bee colony mortality is not because of the wax moths (Annand 2008). G. mellonella has been raised for a wide range of purposes for over 80 years, such as experimental bioassays, biological control agents like egg parasitoids, entomopathogenic nematodes, insect predators, tachinid parasitoids, pet reptile food, fishing bait and other entomopathogens. Waxworms have been grown commercially and marketed in the United States for at least 60 years as bait for fishing and exotic pets. G. mellonella larval development takes 28 to 30 days at 20 to 30°C, while the pupae development takes 6 to10 days at a similar temperature (Krams et al. 2015). Soon after mating, the adult female begins oviposition. Depending on adult density, rh factor, temperature, and larvae feeding. A single female lay 1450 to 1950 eggs in her lifetime (Marston, Campbell, and Boldt 1975). Adults can live up to 12 days without eating; however, most females reach oviposition potential 7 days after hatching.

8.6.5.1 Development of larvae and their diet maintenance G. mellonella’s natural diet comprises of honeycombs bee wax, bee brood pollen and honey. The G. mellonella may only develop on bee wax (Krams et al. 2015), and larvae fed on pollen with bee brood

8  •  Mass Production Technologies  179 grow faster and gain more mass (Annand 2008). G. mellonella does not need bee’s wax for development; in addition, the diet shortens the development period and enhance growth rate, owing to the metabolic water production. Due to the high cost of beeswax, attempts have been undertaken to remove it from G. mellonella diet compositions. Haydakm (1936) introduced the first artificial diet composition exempt of bee’s wax, which comprised of glycerin (226 g), honey (273 g), baker’s yeast (45 g), cornflour (182 g), equal quantities (91 g) of powdered skim milk, whole wheat flour and wheat bran. Balázs (1958) found that a meal consisting of bee’s wax (175 g), maize flour (220 g), baker’s yeast (75 g) and equal quantities (110 g) of powdered milk, glycerin, honey, wheat flour and wheat grain improved survivability of larvae. Beck (1960) discovered that a diet consisting of honey (236 g), baby cereal (Pablum) (321 g), brewer’s glycerin (208 g), yeast (94 g), and water resulted in reduced wax (47 g) and improved larval survival (94 ml). Dutky et al. (1962) removed honey from a composition that included infant food (Pablum) (440 g), sucrose (180 g), glycerin (210 g), water (170 ml) and a vitamin solution (1 mL). Marston and Campbell (1973) could replace the expensive baby cereal strategy with CSMA insect pest composition. Previous conventional mass production systems for G. mellonella were designed for the bulk production of biocontrol agents and relied on small dietary changes. Commonly, larvae in the 1st and 2nd stages were reared in jars or cylindrical containers constructed of polystyrene, high-density polyethylene, and glass. The container size varies from 1 to 5 L and is often adjusted with screened bottom and top or sometimes on both angles. Rearing jars are half-filled with dietary material and G. mellonella eggs at a 1 mg/12-24 g diet ratio (Gross 1994). The larvae must be reared at temperatures ranging from 28 to 30°C, with relative humidity ranging from 60 to 75 percent, and in complete darkness. Larvae are allowed to develop for two weeks in these jars before being transferred to wide trays with a capacity at least three times that of the cylindrical jars. If necessary, a diet is introduced at that stage (Marston, Campbell, and Boldt 1975). Larvae grow in pans for next 2 weeks before being removed from the food through sifters or a filter. Larvae may now be collected for sale or frozen for preservation.

8.6.5.2 Rearing and reproduction For the supply of adults to the reproductive colonies, completely mature larvae in some trays or cylinders are allowed to finish their development and pupate. On the cylinder’s edges, larvae spin cocoons or sheets of cardboard may be put vertically in the tray’s middle. In the cardboard gaps, larvae will pupate. To preserve darkness, pans containing cocoons are put into emergence boxes of impenetrable substance and plywood. The dimensions of the emergence box may vary. Approximately its size is 101 cm wide, 71 cm deep and 121 cm high (Marston, Campbell, and Boldt 1975). To hold several pans, emergence boxes may have numerous shelves within. The door of the emergency cage is coated with the insulating wrapper. To enable the emerging adult to go into the rearing cages, translucent tubes are linked to them from the top of the emerging cage. Light flowing through the transparent tube attracts moths. Similarly, larvae may be allowed to grow completely within cylindrical jars. Larvae move to the container’s top and spin cocoons to form pupate. Breeding cylinders should have extra ventilation to evaporate the moisture accumulated before pupation as the larvae empty their gut tracts. Cocoons may be retrieved from the container’s top and put into the emergence boxes (King and Hartley 2016). G. mellonella adults are nocturnal and only show activeness at night. Adults are reared in the circumstances at a temperature 24°C, relative humidity 78-80%, and 14 hours of photoperiod. Fluorescent natural white lights provide illumination with moderate intensity. Moths shed their scales regularly due to their tiny size and low weight, becoming airborne. Scale levels in the environment that are too high may be dangerous to people (Reneccke 2009) and must be regulated. Hartley et al. (1977) define a scale collector as a network of elastic tubes connected to the top surfaces of each oviposition boxes via a central PVC tube. An air blower/ vacuum pump is used to move air from the main tube to the filter. To prevent moths from being sucked into the air circulation network, each tube linking oviposition cages to the air circulation network is equipped with a metal screen. Davis and Jenkins (1995) developed a filtering system to control scaling throughout the rearing chamber. The system consists of two continuously operating

180  Edible Insects Processing for Food and Feed filtration machines that draw air in five locations along opposing sides at low elevations. The air is routed through three filters: an exterior polyester pad, a pleated panel filter (DQP) to remove bigger particles, and a pocket filter (Viledon MF-90) capable of eliminating granules as small as 1 µm. The filtered air is exhausted through a damper located on units top, ensuring a consistent circulation throughout the chamber (Cortes Ortiz et al. 2016). The size and intricacy of the oviposition boxes vary. They are often square boxes of PVC tubing, high-density polyethene (HDPE) framing, metal framework, and wood with screens. The substance of the screen may be nylon, polypropylene, metal, polyester, and the size of the aperture can range from 0.4-300 mm. The substrates of oviposition may also differ. In one configuration, the oviposition substrate is provided from the top of the cage via weather-stripping coated perforations that hang inside the cage. The oviposition medium was made of plastic tarps treated with a fine droplet of 10% sucrose over water and was covered on both sides with granulated sugar. After the sugar had fully dried, the sheets were inserted through the cage openings for moth treatment. The eggs were retrieved by cleaning the sheets in freshwater with sustainable detergents and then putting the water through a normal No. 60 sieve (Marston, Campbell, and Boldt 1975). The oviposition medium was waxed paper, as well as the accordion-folded sheets that were used to deliver it into oviposition cages. Scraping eggs off sheets was the method used to gather them (King 1979; Cortes Ortiz et al. 2016). The oviposition cages are intended to rear Diatraea grandiosella Dyar, the southwestern corn borer; nevertheless, the method is extremely applicable for several species of Lepidoptera, particularly G. mellonella. Cages have 320 mm mesh hardware fabric walls and are composed of steel angle sheathing (which can be replaced with aluminum or another material). The top of the cage is equipped with identical openings to those mentioned previously for inserting sheets of waxed paper as oviposition media. This cage also has a bottom hole for introducing pupae for the emerging of adults within the cage. The dimensions of the cage are 64 × 64 × 64 cm and have space for more than 1500 adults. This method has been used in several techniques for mass rearing Helicoverpa zea, Spodoptera frugiperda and Heliothis virescens (Frank M. Davis 2009).

8.7  ENVIRONMENTAL CONTROL AND CONDITIONS 8.7.1 Physical Factors 8.7.1.1 Light, temperature and location Physical variables, such as light wavelength and intensity, may influence arthropods growth and reproduction. The light wavelength influences some insects, and some are not taking the light necessary for their development. T. molitor and Z. morio do not require light for development, while M. domestica and H. illucens are needed. In different studies, it has been noted that light-dependent species reproduce more when exposed to sunlight. Choosing a location is critical when it comes to insect farming and building facilities. Land and space tend to be more costly in densely populated areas. A high level of efficiency and precision are required in manufacturing systems. Air filtration is required to prevent harmful chemicals from entering the system, and pests are usually more resistant to pesticides since highly intensive systems may be crushed quickly. With plenty of room and ideal weather (like in tropical regions), room acclimatization isn’t necessary. It is possible to establish an insect production company in a variety of places. Areas with chilly winters and hot summers, such as the United States, will need heating and cooling equipment, respectively. After determining the optimum temperature for insect growth, reproduction and understanding the area’s climate conditions, it is possible to estimate energy requirements by calculating the difference between extreme (maximum and minimum) external temperatures and the ideal for insect rearing.

8  •  Mass Production Technologies  181

8.7.2 Mechanical factors 8.7.2.1 Filtration system Many acclimatization factors must be addressed during production process. The system’s efficiency might be compromised if one of the components was chosen incorrectly. This may quadruple energy usage compared to a system that rapidly increases room humidity. Because a single valve failure may cause the system to fail which claims that all components are vital. The next sections address several mass rearing acclimatization systems (Krauss, Steffan-Dewenter, and Tscharntke 2003). Filters are mostly used to remove volatile solids. Filters are often disregarded in production systems, yet they are vital in keeping the system isolated from the outside. Pesticides and biologicals, as well as entomophagous and entomopathogenic organisms, are the major concerns. Parasites, predators, or competing species pose little risk to the rearing integrity, and they may develop and spread to cause economic losses. The exterior air typically includes entomopathogenic microorganisms, including fungus, bacteria, and viruses. Once in the rearing system, they may cause severe harm or even death (Weissman et al. 2012). When it comes to contamination prevention, mechanical filters are the most cost-effective option. Among the essential filters, these nets can keep rats, birds, and other powerful creatures from entering the ventilation system since they may be made of various materials. Options abound and decision-making is influenced by facility requirements, cost, particle size retention, ease of cleaning, and replacement. Mechanical filters have the drawback of increasing dirt buildup because of load losses in the air movement caused by friction. Filters like this help to remove hazardous substances from the air by absorbing them. Pesticides, organic solvents, and other airborne pollutants may be removed using activated carbon (Ao and Lee 2004). As the activated carbon collects particles, the filter’s activity degrades with time, necessitating a replacement after a specific number of operating hours. Activated carbon is sandwiched between two layers of porous material in certain filters. This filter type is capable of retaining solid particles as well as volatile chemicals. Even so, particle buildup will reduce the filter’s useful life; in this case, a prefilter is recommended.

8.7.2.2 Panels and pads for evaporation Animal farms and greenhouses with more traditional climate control utilize the panel system for cooling. For the system to function, water is used to constantly wet evaporative pads within a wall cover. A constant airflow is forced through the moistened panels by fans mounted on the opposite wall, which serve as exhausts and remove the room’s air. Water evaporation reduces the temperature of the air by transferring energy to it, and as a result, the air becomes moist, raising the relative humidity. The pads design provides a laminar flow arrangement, promoting air circulation, making it a popular choice for humidifying and cooling air before entering the rearing area. As an evaporative panel, the device works flawlessly.

8.7.2.3 Humidifiers Water is evaporated and then mixed with the air in various ways by humidifiers. Because healthy growth, prevention of illnesses and malformations need an appropriate atmospheric humidity level, because humidity is critical in insect production. However, a humidity-increasing device lowers the temperature of the air as well. Air conditioning systems or steam water may be used to cool space before being brought in from the outside. This device can warm the rearing chamber while simultaneously pumping huge quantities of water into the environment. In the case of a steam fault, an alert or cutoff mechanism is required to prevent overheating. Cold-weather establishments may utilize water boilers to generate vapor of heat and humidity rooms using this method. Dehumidifiers may be used to reduce humidity in the rearing room environment in certain situations. Still, they are costly to operate and are most likely used in producing insects for human consumption. Air ventilation should be used as much as possible to reduce the amount of surplus humidity. Fans are essential for a turbulent flow to break down stratification and evenly distribute fresh air in the rearing chamber. Water condenses in the evaporator of air conditioners, which then

182  Edible Insects Processing for Food and Feed removes moisture from the air. While air conditioner dehumidification is effective only when the outside air exchange ratio is low and its effect may be overlooked when the outside air exchange ratio is high. In an electrostatic precipitator, the incoming particles are negatively charged, and a positively charged plate attracts them. They are very effective in removing germs and other pollutants from the air. Air purification systems that use electrostatic precipitator filters are particularly suggested in a rearing house (Cortes Ortiz et al. 2016).

8.8  BASIC NEEDS FOR THE SUPPLY CHAIN SYSTEM 8.8.1 Feed Insect farmers frequently use grain/corn-based feed formulation at commercial scale for fishing bait and pet food sectors, and this feed is comparable in nutritional value to chicken feed. However, some of these sectors use organic feed to grow insects. Most organic insects do not have an official certification; according to certain farms, insects are grown specifically for human consumption. Crickets are given human-grade food and milled grains and, in some situations, are classified as post-consumer food waste. Companies define their standards depending on their marketing objectives because no legal definitions, standards, or certifications define the insect food grade. However, it is unclear if this is necessary from a legal standpoint or to ensure the safety and quality of insects raised for human food in general. It’s possible but unlikely that older farms produce insects as live pet/reptile feed or fishing bait in indoor farming and insect-friendly feeding standards. Given the necessity for gluten-free food items and the ability of insect powders or meals to provide a high-protein substitute, gluten-free feed may be necessary for insect farmers to utilise whenever possible. Similarly, when demand for organically produced protein increases, demand for edible insects raised on organic feed rises proportionally. There are no certified organic edible insect farms, despite claims by certain industries that the edible insects they offer fed a certified organic diet. Insect farming is a rapidly growing industry to produce livestock feed, especially fish, pig, and chicken feed, mostly based on biomaterials with a minimum cost like food waste, animal parts (fish offal, etc.), animal waste, and agricultural byproducts with low quality, for many flies’ farm especially black soldier fly. Many insects, particularly flies, can convert the trash into healthy food. When the integration of fly farms with other animal farms or facilities of meat processing occurs, there is some debate about fly larvae’s potential to decrease overall levels of microorganisms. The fed waste as feed ultimately increases the cleanliness. Using pre-and postconsumer organic waste, algae, or non-food crop biomass such as fast-growing grasses and other plants as a source of insect nutrition could be a largely untapped, which is environmentally beneficial source of insect nutrition. While many home-level insect growers regularly use food waste such as fruit, grain and vegetable. Edible insect farms may safely utilise these insect feed component sources on a large scale, decreasing the costs of edible insects while also increasing their already low environmental impact. At an industrial scale, a system to maintain the quality and quantity of the food waste stream would allow edible insect species to be fed with more nutrients and reduce the risk of foodborne infections (this strategy is not devised by any insect farm yet).

8.8.2 Farms and Farmed Species Generally, we need a huge number of edible insects’ farms on a large scale. A major problem is the need for innovation of insect’s farms on a large scale, enabling the farmer to realize the anticipated financial

8  •  Mass Production Technologies  183 advantages of producing animals that convert 10 times more effective feed or food than cattle and less area is required to feed them. As previously stated, existing insect farming methods rely mainly on human labor, which is expensive in industrialized nations and a major factor in the continuing high costs of edible insects in food. More study is required to enhance methods that will preserve quality and insect health while boosting automation and life cycle shortening. Restrictions addressing land usage, especially for farms located in the urban areas, and some other agricultural laws, notably in North America and Europe, are virtually non-existent for all elements of insect farming, along with regulations regarding insects as food. Furthermore, the government has no taxes, subsidies, or other financial support mechanisms in various parts of the world, generally, for edible insect farms. In that case, tax structures encourage food production; the sector will be pushed even farther toward long-term sustainability. On the market, insects provide the cheapest animal-based protein. We need far bigger insect farms, higher quantity and variety of insects generated by such farms throughout the world. The mealworm and crickets are the bases of rapidly growing insect-producing farms. Some of the industry’s shortcomings are combined by representing insects diversity and lack of scales among commercially available insects from the indoor condition of industrial insect farms. Food production based on insects suffers from higher costs per unit than the other products from farm to fork. The whole insects processing industry consists of small companies and farms, or the end products form of processed insects. It can be said that larger and bigger insect farms with a wider variety of species than crickets, waxworms, and mealworms are part of the answer. Farms with more varied offers may generate new income streams from higher-end goods, allowing them to cater to a broader range of preferences. Many restaurants, for example, may want to provide grasshoppers, gigantic water bugs, or other unusual dishes containing cricket and mealworm-based meals. The existing startups in this sector will expand more efficiently with an increased (therefore cheaper cost per unit), more reliable and constant supply. Simultaneously, larger mainstream food farms will become increasingly interested in incorporating this new commodity into their products.

8.8.3 Transportation, Storage, and Distribution Generally, food-based insect and edible insects industries will rely heavily on current infrastructure in the mainstream food sector to get started. Because a strong infrastructure and procedures for the transportation, distribution of processed food, and raw storage already exist. Farms based on insects can use this resource forever with minimal changes as it develops. When does it make sense to export life vs frozen, dry vs wet, processed vs raw? This is an essential issue for an insect farmer or processor to consider. As with most other food component comedies, the coup de gras for the most efficient shipping, transporting, and storing food ingredients based on insects is powdered and dry. Dry products get more benefits through their extended shelf life, low cost, and lightweight.

8.8.4 Processing and Manufacturing Infrastructure Producers of Insect-based food presently use two main techniques to produce the products: • Working with food manufacturing sectors to identify the optimal methods and, as a result, outsource manufacture after the required product result is determined. • Use of commercial-grade equipment to make the food items. Each of these potential approaches has its own set of benefits. Companies may depend on industry expertise to create high-quality goods in the initial outsourcing scenario. Every method has its advantages. Companies may depend on industries expertise to create high-quality goods in initial outsourcing. Though

184  Edible Insects Processing for Food and Feed learning and time are required to produce the high-quality products to sell may be more in the second scenario, the financial expenses may be lower at first, which is frequently beneficial for the startup using little or no money. Many companies in the present insect-based food industry make their products by hand in small commercial kitchens. For the predictable future, a small newly formed insect-based food company at an important scale will entirely depend on existing services, processing equipment and contract manufacturers compared with older, larger and more advanced food sectors. The industrialised world already has an excellent infrastructure of contract manufacturers, equipment, and facilities with the financial means to access it. When small food companies expand, contract manufacturers, copackers, other fee-based processing and packaging firms become essential. A commercial kitchen may not fulfil the product demand as a business develops. The firm may not be well-financed and big enough to construct its processing facility, costing several million (USD). Most of these manufacturers exist, but many do not allow insects inside their facilities for several reasons. Some rely on strict restrictions like limited product range (for example, some process only starches, plant material but not process protein), allergens and kosher status. Still, many are interested in having insect facilities.

8.9 CHALLENGES Insect farming for food and feed production is still in its infancy, despite the significant attention it has garnered recently and a few experts’ excellent research and analytical work. Fundamental and applied research is very close to engineering and industrial applications because of the high relationship between their beginnings and development. The fast increase in the world’s population may be the driving factor behind the quick advancement of science and technology and the growing attention of entrepreneurs and business leaders. According to demographic estimates, the world’s population will soar above 9 billion people in 2050. Scientists say global food and feed production would have to be increased by feeding these people. That means dealing with the shortage of land for farming and gardening, the depletion of ocean fishing resources, and the negative effects of climate change on food and feed production in both developed and developing nations. All of this has obstacles that must be conquered. Food and feed production must be re-evaluated and modified by humans if they are to survive. Going beyond improving manufacturing efficiency and reducing waste is required. Modern civilizations, in particular, seem to need new sources of food and feed. Insects may be a logical and lucrative way to satisfy future food and feed needs. Even yet, using insects as food or feed has several issues on many levels. Some of these problems are already apparent, but more will become apparent as research into insect farming continues and the practice grows more widespread. The acceptability of insects and insect-based goods in contemporary cultures is the first major issue. Modern cultures, especially in the West, vary about eating food containing components derived from insects. Even though humans have consumed edible insects in different parts of the world since antiquity revulsion is lessened when feeding is made with components derived from insect biomass. As a result, steps must be made to increase knowledge about insects’ food and feed production possibilities. This social issue has seen some progress, but much more required. International agencies, food and feed producers, and political decision-makers all need to be engaged in this process. The correct assessment of the variables involved in insect rearing is the second major issue. Insect farming will need much research before the conventional method of harvesting for edible insects. Given the topic’s breadth, only a genuinely interdisciplinary approach can ensure substantial development in a reasonable amount of time. Advances in understanding insect health, development rate, nutritional profile, capacity to convert agricultural wastes and by-products into usable biomass will need specialized skills in rearing conditions and insect diet. Insect-derived nutrients and micronutrients, including amino acids, lipids, chitin, and metallic elements, have similar optimum usage difficulties.

8  •  Mass Production Technologies  185 The intense insect rearing itself is the third major challenge. Mechanization, automation, and manual labor all need to be balanced fairly. The majority of insect farming techniques need a lot of manual work. Having a large crew is essential when raising animals to large sizes since only a few phases are mechanized. As a result, compared to alternative feed sources like soy meal, goods currently on the market are still expensive. There are still several steps that need to be optimized. When selecting a new insect, several factors must be considered. Researchers and entrepreneurs have collaborated on a few instances, resulting in substantial advances and the filing of patent applications. Regulation, scale production, high inputs, and commodities management are all problems linked to insect farming.

ACKNOWLEDGEMENT The authors wish to express their gratitude to Cortes Ortiz et al., 2016, Naseem et al., 2021, Dortmans et al., 2017 and the copyright permission agencies for taking hints or modified materials. We also acknowledge the reviewers and well-wishers who always help in approaching the scientific research.

REFERENCES Abdollahi, M. R., V. Ravindran, and B. Svihus. 2013. “Pelleting of Broiler Diets: An Overview with Emphasis on Pellet Quality and Nutritional Value.” Animal Feed Science and Technology 179(1–4). Elsevier: 1–23. Adeniji, A. A. 2007. “Effect of Replacing Groundnut Cake with Maggot Meal in the Diet of Broilers.” International Journal of Poultry Science 6(11). Citeseer: 822–25. Ajani, E. K., L. C. Nwanna, and B. O. Musa. 2004. “Replacement of Fishmeal with Maggot Meal in the Diets of Nile Tilapia, Oreochromis niloticus.” World Aquaculture Magazine 35(1): 52–54. Annand, N. 2008. “Small Hive Beetle Management Options.” Primefacts 764(764): 1–7. Ao, C. H., and S. C. Lee. 2004. “Combination Effect of Activated Carbon with TiO2 for the Photodegradation of Binary Pollutants at Typical Indoor Air Level.” Journal of Photochemistry and Photobiology, Part A: Chemistry 161(2–3). Elsevier: 131–40. doi:10.1016/S1010-6030(03)00276-4. Awoniyi, T. A. M., V. A. Aletor, and J. M. Aina. 2003. “Performance of Broiler - Chickens Fed on Maggot Meal in Place of Fishmeal.” International Journal of Poultry Science 2(4). Asian Network for Scientific Information: 271–74. doi:10.3923/ijps.2003.271.274. Balázs, A. 1958. “Nutritional and Nervous Factors in the Adaptation of Galleria mellonella to Artificial Diet.” Acta Biologica Hungarica 9: 47–69. Banks, Ian J., Walter T. Gibson, and Mary M. Cameron. 2014. “Growth Rates of Black Soldier Fly Larvae Fed on Fresh Human Faeces and Their Implication for Improving Sanitation.” Tropical Medicine and International Health 19(1). Wiley Online Library: 14–22. Barsics, Fanny, Rudy Caparros Megido, Yves Brostaux, Catherine Barsics, Christophe Blecker, Eric Haubruge, and Frédéric Francis. 2017. “Could New Information Influence Attitudes to Foods Supplemented with Edible Insects?” British Food Journal 119(9). Emerald Publishing Limited: 2027–39. doi:10.1108/BFJ-11-2016-0541. Beck, S. D. 1960. “Growth and Development of the Greater Wax Moth, Galleria mellonella (L.).” Wisconsin Academy of Sciences, Arts and Letters 49: 137–48. Belluco, Simone, Carmen Losasso, Michela Maggioletti, Cristiana C. Alonzi, Maurizio G. Paoletti, and Antonia Ricci. 2013. “Edible Insects in a Food Safety and Nutritional Perspective: A Critical Review.” Comprehensive Reviews in Food Science and Food Safety 12(3). Wiley Online Library: 296–313. doi:10.1111/1541-4337.12014. Bensky, Alan. 2019. “Technologies and Applications.” Short-Range Wireless Communication. Wood Head Publishing, Cambridge, UK. doi:10.1016/b978-0-12-815405-2.00014-2. Borrelli, Luca, Lorena Coretti, Ludovico Dipineto, Fulvia Bovera, Francesca Menna, Lorenzo Chiariotti, Antonio Nizza, Francesca Lembo, and Alessandro Fioretti. 2017. “Insect-Based Diet, a Promising Nutritional Source, Modulates Gut Microbiota Composition and SCFAs Production in Laying Hens.” Scientific Reports 7(1). Nature Publishing Group: 1–11.

186  Edible Insects Processing for Food and Feed Bosch, G., H. H. E. van Zanten, A. Zamprogna, M. Veenenbos, N. P. Meijer, H. J. van der Fels-Klerx, and J. J. A. van Loon. 2019. “Conversion of Organic Resources by Black Soldier Fly Larvae: Legislation, Efficiency and Environmental Impact.” Journal of Cleaner Production 222. Elsevier: 355–63. doi:10.1016/j. jclepro.2019.02.270. Broekhoven, Sarah van, Dennis G. A. B. Oonincx, Arnold van Huis, and Joop J.A. van Loon. 2015. “Growth Performance and Feed Conversion Efficiency of Three Edible Mealworm Species (Coleoptera: Tenebrionidae) on Diets Composed of Organic By-Products.” Journal of Insect Physiology 73. Elsevier: 1–10. doi:10.1016/j. jinsphys.2014.12.005. Carter, Clay, Sharoni Shafir, Lia Yehonatan, Reid G. Palmer, and Robert Thornburg. 2006. “A Novel Role for Proline in Plant Floral Nectars.” Naturwissenschaften 93(2). Springer: 72–79. doi:10.1007/s00114-005-0062-1. Castro, Ruann Janser Soares de, André Ohara, Jessika Gonçalves dos Santos Aguilar, and Maria Aliciane Fontenele Domingues. 2018. “Nutritional, Functional and Biological Properties of Insect Proteins: Processes for Obtaining, Consumption and Future Challenges.” Trends in Food Science and Technology 76. Elsevier: 82–89. doi:10.1016/j.tifs.2018.04.006. Chen, Bor Yann, Kai Wei Lin, Yu Min Wang, and Chia Yi Yen. 2009. “Revealing Interactive Toxicity of Aromatic Amines to Azo Dye Decolorizer Aeromonas hydrophila.” Journal of Hazardous Materials 166(1). Elsevier: 187–94. doi:10.1016/j.jhazmat.2008.11.030. Chen, J., G. Zhao, and R. Han. 2014. “Mass Production of Housefly (Musca domestica) for Waste Reutilization and as Feed for Chickens and Shrimps.” In: International Conference “Insects to Feed the World,” Wageningen University, Wageningen, Netherlands, 14–17. Chen, X., and Y. Feng. 2002. “Review on Nutritional Value of Edible Insects.” Chinese Forestry Science and Technology 1(2): 54–59. Čičková, Helena, G. Larry Newton, R. Curt Lacy, and Milan Kozánek. 2015. “The Use of Fly Larvae for Organic Waste Treatment.” Waste Management 35. Elsevier: 68–80. doi:10.1016/j.wasman.2014.09.026. Cladera, Jorge L., Juan C. Vilardi, Marianela Juri, Laura E. Paulin, M. Cecilia Giardini, Paula V. Gómez Cendra, Diego F. Segura, and Silvia B. Lanzavecchia. 2014. “Genetics and Biology of Anastrepha fraterculus: Research Supporting the Use of the Sterile Insect Technique (SIT) to Control this Pest in Argentina.” BMC Genetics 15(2). BioMed Central: 1–14. doi:10.1186/1471-2156-15-S2-S12. Clifford, Craig W., Richard M. Roe, and J. P. Woodring. 1977. “Rearing Methods for Obtaining House Crickets, Acheta domesticus, of Known Age, Sex, and Instar.” Annals of the Entomological Society of America 70(1). Oxford University Press: 69–74. doi:10.1093/aesa/70.1.69. Cock, Matthew J. W., Sean T. Murphy, Moses T. K. Kairo, Emma Thompson, Rebecca J. Murphy, and Antonio W. Francis. 2017. “Erratum to: Trends in the Classical Biological Control of Insect Pests by Insects: An Update of the BioCat Database (BioControl, (2016), 61, 4, (349–363), 10.1007/S10526-016-9726-3).” BioControl 62(4). Springer: 579–80. doi:10.1007/s10526-017-9821-0. Cohen, Allen Carson. 2003. Insect Diets: Science and Technology. CRC Press. Cortes Ortiz, J. A., A. T. Ruiz, J. A. Morales-Ramos, M. Thomas, M. G. Rojas, J. K. Tomberlin, L. Yi, R. Han, L. Giroud, and R. L. Jullien. 2016. “Insect Mass Production Technologies.” In: Insects as Sustainable Food Ingredients, 153–201. Elsevier. doi:10.1016/b978-0-12-802856-8.00006-5. Costa-Neto, Eraldo Medeiros. 2005. “Entomotherapy, or the Medicinal Use of Insects.” Journal of Ethnobiology 25(1). BioOne: 93–114. doi:1​0.299​3/027​8-077​1(200​5)25[​93:EO​TMUO]​2.0.C​O;2. Coudron, T. A., L. C. Mitchell, R. Sun, J. D. Robertson, N. V. Pham, and H. J. R. Popham. 2012. “Dietary Composition Affects Levels of Trace Elements in the Predator Podisus maculiventris (Say) (Heteroptera: Pentatomidae).” Biological Control 61(2). Elsevier: 141–46. doi:10.1016/j.biocontrol.2012.01.011. Davis, F. M. 2009. “Insect Rearing Productión Systems, A Case Study: The Southwestern Corn Borer.” In: Principles and Procedures for Rearing High Quality Insects. Mississippi State University Press. Starkville, MS, 307–35. Davis, F. M., and J. N. Jenkins. 1995. “Management of Scales and Other Insect Debris: Occupational Health Hazard in a Lepidopterous Rearing Facility.” Journal of Economic Entomology 88(2). Oxford University Press: 185– 91. doi:10.1093/jee/88.2.185. Diener, Stefan, Nandayure M. Studt Solano, Floria Roa Gutiérrez, Christian Zurbrügg, and Klement Tockner. 2011. “Biological Treatment of Municipal Organic Waste Using Black Soldier Fly Larvae.” Waste and Biomass Valorization 2(4). Springer: 357–63. doi:10.1007/s12649-011-9079-1. Dobermann, D. 2017. “Insects as Food and Feed: Can Research and Business Work Together?” Journal of Insects as Food and Feed 3(2). Wageningen Academic Publishers: 155–60. doi:10.3920/JIFF2016.0040. Dobermann, D., J. A. Swift, and L. M. Field. 2017. “Opportunities and Hurdles of Edible Insects for Food and Feed.” Nutrition Bulletin 42(4). Wiley Online Library: 293–308. doi:10.1111/nbu.12291.

8  •  Mass Production Technologies  187 Dortmans, Bram, Stefan Diener, Bart Verstappen, and Christian Zurbrügg. 2017. “Black Soldier Fly Biowaste Processing.” In: Black Soldier Fly Biowaste Processing. A Step-by Step Guide. Swiss Agency for Development and Cooperation (SDC) and the Swiss State Secretariat for Economic Affairs (SECO), 100. Dossey, Aaron T. 2010. “Insects and Their Chemical Weaponry: New Potential for Drug Discovery.” Natural Product Reports 27(12). Royal Society of Chemistry: 1737–57. doi:10.1039/c005319h. Dutky, S. R., J. V. Thompson, George E. Cantwell, et al. 1962. “A Technique for Mass Rearing the Greater Wax Moth (Lepidoptera: Galleriidae).” Proceedings of the Entomological Society of Washington 64(1): 56–58. Ebenebe, C. I., V. O. Okpoko, A. N. Ufele, and M. I. Amobi. 2017. “Survivability, Growth Performance and Nutrient Composition of the African Palm Weevil (Rhyncophorus phoenicis Fabricius) Reared on Four Different Substrates.” Journal of Bioscience and Biotechnology Discovery 2(1): 1–9. doi:10.31248/jbbd2016.018. Erickson, Marilyn C., Mahbub Islam, Craig Sheppard, Jean Liao, and Michael P. Doyle. 2004. “Reduction of Escherichia coli O157:H7 and Salmonella enterica serovar Enteritidis in Chicken Manure by Larvae of the Black Soldier Fly.” Journal of Food Protection 67(4). International Association for Food Protection: 685–90. doi:10.4315/0362-028X-67.4.685. Fasakin, E. A., A. M. Balogun, and O. O. Ajayi. 2003. “Evaluation of Full-Fat and Defatted Maggot Meals in the Feeding of Clariid Catfish Clarias gariepinus Fingerlings.” Aquaculture Research 34(9). Wiley Online Library: 733–38. doi:10.1046/j.1365-2109.2003.00876.x. Feingold, Kenneth R., and Carl Grunfeld. 2000. “Introduction to Lipids and Lipoproteins.” In: Endotext. Ferrarezi, Rhuanito, Lorenzo Cannella, Abdel Rahman Ahmed Nassef, and Donals S. Bailey. 2016. Alternative Sources of Food for Aquaponics in the U.S. Agricultural Experiment Station, 34. Finke, Mark D., and Dennis Oonincx. 2013. “Insects as Food for Insectivores.” In: Mass Production of Beneficial Organisms: Invertebrates and Entomopathogens, 583–616. Elsevier. doi:10.1016/B978-0-12-391453-8.00017-0. Freelance, Christopher B. 2019. “To Regulate or Not to Regulate? The Future of Animal Ethics in Experimental Research with Insects.” Science and Engineering Ethics 25(5). Springer: 1339–55. doi:10.1007/s11948-018-0066-9. Gasco, Laura, Francesco Gai, Giulia Maricchiolo, Lucrezia Genovese, Sergio Ragonese, Teresa Bottari, and Gabriella Caruso. 2018. “Feeds for the Aquaculture Sector Current Situation and Alternative Sources.” In: Feeds for the Aquaculture Sector. Springer International Publishing, 1–103. Ghaly, A. E. 2009. “The Use of Insects as Human Food in Zambia.” OnLine Journal of Biological Sciences 9(4): 93–104. doi:10.3844/ojbsci.2009.93.104. Gross, Harry R. 1994. “Mass Propagation of Archytas marmoratus (Diptera: Tachinidae).” Environmental Entomology 23(1). Oxford University Press: 183–89. Han, Rachel, Jungyoung Tiffany Shin, Jin Kim, Yong Seok Choi, and Young Wook Kim. 2017. “An Overview of the South Korean Edible Insect Food Industry: Challenges and Future Pricing/Promotion Strategies.” Entomological Research 47(3). Wiley Online Library: 141–51. doi:10.1111/1748-5967.12230. Hartley, G. G., C. W. Gantt, E. G. King, and D. F. Martin. 1977. “Equipment for Mass Rearing of the Greater Wax Moth and the Parasite Lixophaga Diatraeae.” ARS-S-US Agricultural Research Service 164: 1–4. Haydakm, Mykola H. 1936. “Is Wax a Necessary Constituent of the Diet of Wax Moth Larvae?” Annals of the Entomological Society of America 29(4). Oxford University Press: 581–88. Henry, M., L. Gasco, G. Piccolo, and E. Fountoulaki. 2015. “Review on the Use of Insects in the Diet of Farmed Fish: Past and Future.” Animal Feed Science and Technology 203(1). Elsevier: 1–22. doi:10.1016/j. anifeedsci.2015.03.001. Heussler, Carina D., Andreas Walter, Hannes Oberkofler, Heribert Insam, Wolfgang Arthofer, Birgit C. SchlickSteiner, and Florian M. Steiner. 2018. “Influence of Three Artificial Light Sources on Oviposition and HalfLife of the Black Soldier Fly, Hermetia illucens (Diptera: Stratiomyidae): Improving Small-Scale Indoor Rearing.” PLOS ONE 13(5). Public Library of Science: e0197896. Hogsette, Jerome A. 1992. “New Diets for Production of House Flies and Stable Flies (Diptera: Muscidae) in the Laboratory.” Journal of Economic Entomology 85(6). Oxford University Press: 2291–94. Holmes, L. A., S. L. Van Laerhoven, and J. K. Tomberlin. 2010. “Lower Temperature Threshold for Black Soldier Fly (Diptera: Stratiomyidae) Egg and Adult Eclosion.” In: Lower Temperature Threshold for Black Soldier Fly (Diptera: Stratiomyidae) Egg and Adult Eclosion. American Academy of Forensic Sciences, 1–310. Holmes, L. A., S. L. Van Laerhoven, and J. K. Tomberlin. 2012. “Relative Humidity Effects on the Life History of Hermetia illucens (Diptera: Stratiomyidae).” Environmental Entomology 41(4). Oxford University Press: 971–78. Hopkins, Theodore L., and Karl J. Kramer. 1992. “Insect Cuticle Sclerotization.” Annual Review of Entomology 37(1). Annual Reviews: 273–302.

188  Edible Insects Processing for Food and Feed Hussein, Mahmoud, Viju V. Pillai, Joshua M. Goddard, Hui G. Park, Kumar S. Kothapalli, Deborah A. Ross, Quirine M. Ketterings, et al. 2017. “Sustainable Production of Housefly (Musca domestica) Larvae as a ProteinRich Feed Ingredient by Utilizing Cattle Manure.” PLOS ONE 12(2). Public Library of Science: e0171708. doi:10.1371/journal.pone.0171708. Hwangbo, J., E. C. Hong, A. Jang, H. K. Kang, J. S. Oh, B. W. Kim, and B. S. Park. 2009. “Utilization of House Fly-Maggots, a Feed Supplement in the Production of Broiler Chickens.” Journal of Environmental Biology 30(4): 609–14. Jia, S. F. 2007. “Study on Maggot Separation, Processing and Preservation.” Feeding Research 7: 36–37. Kampmeier, Gail E., and Michael E. Irwin. 2009. “Commercialization of Insects and Their Products.” In: Encyclopedia of Insects, 220–27. Elsevier. doi:10.1016/B978-0-12-374144-8.00068-0. Kenis, M., N. Koné, C. A. A. M. Chrysostome, E. Devic, G. K. D. Koko, V. A. Clottey, S. Nacambo, and G. A. Mensah. 2014. “Insects Used for Animal Feed in West Africa.” Entomologia 2(October): 107–14. doi:10.4081/ entomologia.2014.218. King, E. G. 1979. “Production of the Tachinid Lixophaga Diatraeae on Its Natural Host, the Sugarcane Borer, and on an Unnatural Host, the Greater Wax Moth.” Agricultural Research. Southern Region, Science and Education Administration. King, E. G., and G. G. Hartley. 2016. “Galleria mellonella.” In: Encyclopedia of Parasitology, Springer Berlin, Heidelberg, 1096–1096. doi:10.1007/978-3-662-43978-4_1240. Krams, Indrikis, Sanita Kecko, Katariina Kangassalo, Fhionna R. Moore, Eriks Jankevics, Inna Inashkina, Tatjana Krama, Vilnis Lietuvietis, Laila Meija, and Markus J. Rantala. 2015. “Effects of Food Quality on Trade-Offs among Growth, Immunity and Survival in the Greater Wax Moth Galleria mellonella.” Insect Science 22(3). Wiley Online Library: 431–39. Krauss, Jochen, Ingolf Steffan-Dewenter, and Teja Tscharntke. 2003. “How Does Landscape Context Contribute to Effects of Habitat Fragmentation on Diversity and Population Density of Butterflies?” Journal of Biogeography 30(6): 889–900. doi:10.1046/j.1365-2699.2003.00878.x. Krieg, N. R. 1998. “Acheta domesticus.” Hilgardia 289: 288–89. Lavilla, Isela, Genma Rodríguez-Liñares, Josefina Garrido, and Carlos Bendicho. 2010. “A Biogeochemical Approach to Understanding the Accumulation Patterns of Trace Elements in Three Species of Dragonfly Larvae: Evaluation as Biomonitors.” Journal of Environmental Monitoring 12(3). Royal Society of Chemistry: 724–30. Li, Yanxian, Trond M. Kortner, Elvis M. Chikwati, Ikram Belghit, Erik Jan Lock, and Åshild Krogdahl. 2020. “Total Replacement of Fish Meal with Black Soldier Fly (Hermetia illucens) Larvae Meal Does Not Compromise the Gut Health of Atlantic Salmon (Salmo salar).” Aquaculture 520. Elsevier: 734967. doi:10.1016/j. aquaculture.2020.734967. Lindig, O. H., P., A. Hedin, and W. E. Poe. 1981. “Amino Acids in Pecan Weevil, Southwestern Corn Borer and Tarnished Plant Bug, and at Their Feeding Sites.” Comparative Biochemistry and Physiology. Part A: Physiology 68(2). Elsevier: 261–63. Liu, Qiaolin, Jeffery K. Tomberlin, Jeff A. Brady, Michelle R. Sanford, and Ziniu Yu. 2008. “Black Soldier Fly (Diptera: Stratiomyidae) Larvae Reduce Escherichia coli in Dairy Manure.” Environmental Entomology 37(6). Oxford University Press: 1525–30. Lock, E. R., T. Arsiwalla, and R. Waagbø. 2016. “Insect Larvae Meal as an Alternative Source of Nutrients in the Diet of a Tlantic Salmon (S Almo Salar) Postsmolt.” Aquaculture Nutrition 22(6). Wiley Online Library: 1202–13. Lundgren, Jonathan G. 2009. “Nutritional Aspects of Non-prey Foods in the Life Histories of Predaceous Coccinellidae.” Biological Control 51(2). Elsevier: 294–305. Lundy, Mark E., and Michael P. Parrella. 2015. “Crickets Are Not a Free Lunch: Protein Capture from Scalable Organic Side-Streams via High-Density Populations of Acheta domesticus.” PLOS ONE 10(4). Public Library of Science: e0118785. doi:10.1371/journal.pone.0118785. Madhan, Selvaraj, Mookkan Prabakaran, and Jimmy Kwang. 2010. “Baculovirus as Vaccine Vectors.” Current Gene Therapy 10(3). Bentham Science Publishers: 201–13. Makkar, Harinder P. S., Gilles Tran, Valérie Heuzé, and Philippe Ankers. 2014. “State-of-the-Art on Use of Insects as Animal Feed.” Animal Feed Science and Technology 197. Elsevier: 1–33. doi:10.1016/j.anifeedsci.2014.07.008. Marston, N., B. Campbell, and P. E. Boldt. 1975. Mass Producing Eggs of the Greater Wax Moth, Galleria mellonella (L.). Technical Bulletin. Department of Agriculture. Agricultural Research Service, US Department of Agriculture. Marston, Norman, and Barbara Campbell. 1973. “Comparison of Nine Diets for Rearing Galleria mellonella.” Annals of the Entomological Society of America 66(1). Oxford University Press: 132–36. Meneguz, Marco, Laura Gasco, and Jeffery K. Tomberlin. 2018. “Impact of PH and Feeding System on Black Soldier Fly (Hermetia illucens, L; Diptera: Stratiomyidae) Larval Development.” PLOS ONE 13(8). Public Library of Science: e0202591.

8  •  Mass Production Technologies  189 Meyer-Rochow, Victor Benno, Ruparao T. Gahukar, Sampat Ghosh, and Chuleui Jung. 2021. “Chemical Composition, Nutrient Quality and Acceptability of Edible Insects Are Affected by Species, Developmental Stage, Gender, Diet, and Processing Method.” Foods 10(5). Multidisciplinary Digital Publishing Institute: 1036. doi:10.3390/ foods10051036. Meyer-Rochow, Victor Benno, Sampat Ghosh, and Chuleui Jung. 2019. “Farming of Insects for Food and Feed in South Korea: Tradition and Innovation.” Berliner und Munchener Tierarztliche Wochenschrift 132(5–6). Schlütersche Verlagsgesellschaft GmbH & Co. KG: 236–44. doi:10.2376/0005-9366-18056. Miech, P., J. E. Berggren, T. Lindberg, B. Chhay, B. Khieu, and A. Jansson. 2016. “Growth and Survival of Reared Cambodian Field Crickets (Teleogryllus testaceus) Fed Weeds, Agricultural and Food Industry By-Products.” Journal of Insects as Food and Feed 2(4). Wageningen Academic Publishers: 285–92. doi:10.3920/ JIFF2016.0028. Mishyna, Maryia, Jianshe Chen, and Ofir Benjamin. 2020. “Sensory Attributes of Edible Insects and Insect-Based Foods–Future Outlooks for Enhancing Consumer Appeal.” Trends in Food Science and Technology 95. Elsevier: 141–48. Mlcek, Jiri, Otakar Rop, Marie Borkovcova, and Martina Bednarova. 2014. “A Comprehensive Look at the Possibilities of Edible Insects as Food in Europe - A Review.” Polish Journal of Food and Nutrition Sciences 64(3): 147–57. doi:10.2478/v10222-012-0099-8. Morales-Ramos, J. A., M. G. Rojas, D. I. Shapiro-Ilan, and W. L. Tedders. 2011. “Self-Selection of Two Diet Components by Tenebrio molitor (Coleoptera: Tenebrionidae) Larvae and Its Impact on Fitness.” Environmental Entomology 40(5): 1285–94. doi:10.1603/EN10239. Morales-Ramos, J. A., M. G. Rojas, D. I. Shapiro-Nan, and W. L. Tedders. 2010. “Developmental Plasticity in Tenebrio molitor (Coleoptera: Tenebrionidae): Analysis of Instar Variation in Number and Development Time under Different Diets.” Journal of Entomological Science 45(2). Georgia Entomological Society, Inc.: 75–90. doi:10.18474/0749-8004-45.2.75. Morales-Ramos, Juan A., and M. Guadalupe Rojas. 2015. “Effect of Larval Density on Food Utilization Efficiency of Tenebrio molitor (Coleoptera: Tenebrionidae).” Journal of Economic Entomology 108(5). Oxford University Press: 2259–67. doi:10.1093/jee/tov208. Morales-Ramos, Juan A., M. Guadalupe Rojas, Sasha Kay, David I. Shapiro-llan, and W. Louis Tedders. 2012. “Impact of Adult Weight, Density, and Age on Reproduction of Tenebrio molitor (Coleoptera: Tenebrionidae).” Journal of Entomological Science 47(3). Georgia Entomological Society, Inc.: 208–20. doi:10.18474/0749-8004-47.3.208. Morales-Ramos, Juan A., M. Guadalupe Rojas, and Thomas A. Coudron. 2013. “Artificial Diet Development for Entomophagous Arthropods.” In: Mass Production of Beneficial Organisms: Invertebrates and Entomopathogens, 203–40. Elsevier. doi:10.1016/B978-0-12-391453-8.00007-8. Morales-Ramos, Juan A., M. Guadalupe Rojas, and David I. Shapiro-Ilan. 2014. Mass Production of Beneficial Organisms: Invertebrates and Entomopathogens. Academic Press. Mościcki, Leszek. 2011. Extrusion-Cooking Techniques: Applications, Theory and Sustainability. John Wiley & Sons. doi:10.1002/9783527634088. Müller, A., J. Evans, C. L. R. Payne, and R. Roberts. 2016. “Entomophagy and Power.” Journal of Insects as Food and Feed 2(2). Wageningen Academic Publishers: 121–36. doi:10.3920/JIFF2016.0010. Murefu, T. R., L. Macheka, R. Musundire, and F. A. Manditsera. 2019. “Safety of Wild Harvested and Reared Edible Insects: A Review.” Food Control 101. Elsevier: 209–24. doi:10.1016/j.foodcont.2019.03.003. Myers, Heidi M., Jeffery K. Tomberlin, Barry D. Lambert, and David Kattes. 2014. “Development of Black Soldier Fly (Diptera: Stratiomyidae) Larvae Fed Dairy Manure.” Environmental Entomology 37(1). Entomological Society of America: 11–15. Nakagaki, Barbara J., and Gene R. Defoliart. 1991. “Comparison of Diets for Mass-Rearing Acheta domesticus (Orthoptera: Gryllidae) as a Novelty Food, and Comparison of Food Conversion Efficiency with Values Reported for Livestock.” Journal of Economic Entomology 84(3). Oxford University Press: 891–96. doi:10.1093/jee/84.3.891. Naseem, Rimsha, Waqar Majeed, Naureen Rana, Elmo Borges de Azevedo Koch, and Muhammad Rehan Naseem. 2021. “Entomophagy: An Innovative Nutritional and Economic Navigational Tool in Race of Food Security.” International Journal of Tropical Insect Science 41(3). Springer: 2211–21. Nation, Sr., James L. 2021. Insect Physiology and Biochemistry. CRC Press. doi:10.1201/b18758. Nguyen, Trinh T. X., Jeffery K. Tomberlin, and Sherah Vanlaerhoven. 2013. “Influence of Resources on Hermetia illucens (Diptera: Stratiomyidae) Larval Development.” Journal of Medical Entomology 50(4). Oxford University Press: 898–906. Nguyen, Trinh T. X., Jeffery K. Tomberlin, and Sherah Vanlaerhoven. 2015. “Ability of Black Soldier Fly (Diptera: Stratiomyidae) Larvae to Recycle Food Waste.” Environmental Entomology 44(2). Oxford University Press: 406–10.

190  Edible Insects Processing for Food and Feed Nongonierma, Alice B., and Richard J. FitzGerald. 2017. “Unlocking the Biological Potential of Proteins from Edible Insects through Enzymatic Hydrolysis: A Review.” Innovative Food Science and Emerging Technologies 43. Elsevier: 239–52. doi:10.1016/j.ifset.2017.08.014. Nugroho, Rudy Agung, and F. M. Nur. 2018. “Insect-Based Protein: Future Promising Protein Source for Fish Cultured.” In IOP Conference Series: Earth and Environmental Science, 144:12002. IOP Publishing. Nyakeri, E. M., H. J. Ogola, M. A. Ayieko, and F. A. Amimo. 2017. “An Open System for Farming Black Soldier Fly Larvae as a Source of Proteins for Smallscale Poultry and Fish Production.” Journal of Insects as Food and Feed 3(1). Wageningen Academic Publishers: 51–56. doi:10.3920/JIFF2016.0030. Ogunji, Johnny, Carsten Schulz, and Werner Kloas. 2008. “Growth Performance, Nutrient Utilization of Nile Tilapia Oreochromis niloticus Fed Housefly Maggot Meal (Magmeal) Diets.” Turkish Journal of Fisheries and Aquatic Sciences 8(1). Citeseer: 141–47. Oonincx, D. G. A. B., N. Volk, J. J. E. Diehl, J. J. A. van Loon, and G. Belušič. 2016. “Photoreceptor Spectral Sensitivity of the Compound Eyes of Black Soldier Fly (Hermetia illucens) Informing the Design of LEDBased Illumination to Enhance Indoor Reproduction.” Journal of Insect Physiology 95. Elsevier: 133–39. doi:10.1016/j.jinsphys.2016.10.006. Oonincx, Dennis G. A. B., and Imke J. M. de Boer. 2012. “Environmental Impact of the Production of Mealworms as a Protein Source for Humans – A Life Cycle Assessment.” PLoS ONE 7(12). Public Library of Science, San Francisco, USA: e51145. doi:10.1371/journal.pone.0051145. Park, Kwan-Ho, Won-Tae Kim, Sang-Beom Lee, Young-Cheol Choi, and Si-Kab Nho. 2010. “Seasonal Pupation, Adult Emergence and Mating of Black Soldier Fly, Hermetia illucens (Diptera: Stratiomyidae) in Artificial Rearing System.” International Journal of Industrial Entomology 21(2). Korean Society of Sericultural Science: 189–91. Pastor, B., Y. Velasquez, P. Gobbi, and S. Rojo. 2015. “Conversion of Organic Wastes into Fly Larval Biomass: Bottlenecks and Challenges.” Journal of Insects as Food and Feed 1(3). Wageningen Academic Publishers: 179–93. doi:10.3920/JIFF2014.0024. Patton, R. L. 1967. “Oligidic Diets for Acheta domesticus (Orthoptera: Gryllidae).” Annals of the Entomological Society of America 60(6). Oxford University Press: 1238–42. Patton, R. L. 1978. “Growth and Development Parameters for Acheta domesticus.” Annals of the Entomological Society of America 71(1). Oxford University Press: 40–42. Pieterse, E., and Q. Pretorius. 2013. “Nutritional Evaluation of Dried Larvae and Pupae Meal of the Housefly (Musca domestica) Using Chemical-and Broiler-Based Biological Assays.” Animal Production Science 54(3). CSIRO Publishing: 347–55. Purusothaman, S., S. Muthuvelu, U. Balasubramanian, and P. Murugesan. 2012. “Biochemical Analysis of Mulberry Leaves (Morus alba L.) and Silkworm Bombyx mori Enriched with Vermiwash.” Journal of Entomology 9(5). Academic Journals Inc.: 289–95. Quaye, Bernard, Comfort Charity Atuahene, Armstrong Donkoh, Benjamin Mensah Adjei, Obed Opoku, and Michael Adu Amankrah. 2018. “Nutritional Potential and Microbial Status of African Palm Weevil (Rhynchophorus phoenicis) Larvae Raised on Alternative Feed Resources.” American Scientific Research Journal for Engineering, Technology, and Sciences (ASRJETS) 48(1): 45–52. Raheem, Dele, Conrado Carrascosa, Oluwatoyin Bolanle Oluwole, Maaike Nieuwland, Ariana Saraiva, Rafael Millán, and António Raposo. 2019. “Traditional Consumption of and Rearing Edible Insects in Africa, Asia and Europe.” Critical Reviews in Food Science and Nutrition 59(14). Taylor & Francis: 2169–88. doi:10.1080 /10408398.2018.1440191. Ramos-Elorduy, Julieta. 2009. “Anthropo-Entomophagy: Cultures, Evolution and Sustainability.” Entomological Research 39(5). Wiley Online Library: 271–88. doi:10.1111/j.1748-5967.2009.00238.x. Ramos-Elorduy, Julieta, Ernesto Avila González, Alma Rocha Hernández, and José Manuel Pino. 2002. “Use of Tenebrio molitor (Coleoptera: Tenebrionidae) to Recycle Organic Wastes and as Feed for Broiler Chickens.” Journal of Economic Entomology 95(1). Oxford University Press: 214–20. doi:10.1603/0022-0493-95.1.214. Reneccke, J. P. 2009. “Health and Safety Issues in the Rearing of Arthropods.” In: Principles and Procedures for Rearing High Quality Insects, Mississippi State University Press. Starkville, MS, 71–85. Resh, V. H. 2003. Encyclopedia of Insects. Carde RT (Ed.). Academic Press. Roffeis, M., B. Muys, J. Almeida, E. Mathijs, W. M. J. Achten, B. Pastor, Y. Velásquez, A. I. Martinez-Sanchez, and S. Rojo. 2015. “Pig Manure Treatment with Housefly (Musca domestica) Rearing-an Environmental Life Cycle Assessment.” Journal of Insects as Food and Feed 1(3). Wageningen Academic Publishers: 195–214. doi:10.3920/JIFF2014.0021. Roncarati, A., L. Gasco, G. Parisi, and G. Terova. 2015. “Growth Performance of Common Catfish (Ameiurus melas Raf.) Fingerlings Fed Mealworm (Tenebrio molitor) Diet.” Journal of Insects as Food and Feed 1(3). Wageningen Academic Publishers: 233–40. doi:10.3920/JIFF2014.0006.

8  •  Mass Production Technologies  191 Rumpold, Birgit A., and Oliver K. Schlüter. 2013. “Potential and Challenges of Insects as an Innovative Source for Food and Feed Production.” Innovative Food Science and Emerging Technologies 17. Elsevier: 1–11. doi:10.1016/j.ifset.2012.11.005. Rutaro, Karlmax, Geoffrey M. Malinga, Vilma J. Lehtovaara, Robert Opoke, Anu Valtonen, Justus Kwetegyeka, Philip Nyeko, and Heikki Roininen. 2018. “The Fatty Acid Composition of Edible Grasshopper Ruspolia differens (Serville) (Orthoptera: Tettigoniidae) Feeding on Diversifying Diets of Host Plants.” Entomological Research 48(6). Wiley Online Library: 490–98. Rutaro, Karlmax, Geoffrey M. Malinga, Robert Opoke, Vilma J. Lehtovaara, Francis Omujal, Philip Nyeko, Heikki Roininen, and Anu Valtonen. 2018. “Artificial Diets Determine Fatty Acid Composition in Edible Ruspolia differens (Orthoptera: Tettigoniidae).” Journal of Asia-Pacific Entomology 21(4). Elsevier: 1342–49. Scudder, Geoffrey G. E. 2009. “The Importance of Insects.” In: Insect Biodiversity. Science and Society, 7–32. doi:10.1002/9781444308211.ch2. Sealey, Wendy M., T. Gibson Gaylord, Frederic T. Barrows, Jeffery K. Tomberlin, Mark A. McGuire, Carolyn Ross, and Sophie St-Hilaire. 2011. “Sensory Analysis of Rainbow Trout, Oncorhynchus mykiss, Fed Enriched Black Soldier Fly Prepupae, Hermetia illucens.” Journal of the World Aquaculture Society 42(1). Wiley Online Library: 34–45. Shapiro-Ilan, David, M. Guadalupe Rojas, Juan A. Morales-Ramos, Edwin E. Lewis, and W. Louis Tedders. 2008. “Effects of Host Nutrition on Virulence and Fitness of Entomopathogenic Nematodes: Lipid-and ProteinBased Supplements in Tenebrio molitor Diets.” Journal of Nematology 40(1). Society of Nematologists: 13–19. Sheppard, D. Craig, Jeffery K. Tomberlin, John A. Joyce, Barbara C. Kiser, and Sonya M. Sumner. 2002. “Rearing Methods for the Black Soldier Fly (Diptera: Stratiomyidae).” Journal of Medical Entomology 39(4). BioOne: 695–98. doi:10.1603/0022-2585-39.4.695. Smil, Vaclav. 2002. “Eating Meat: Evolution, Patterns, and Consequences.” Population and Development Review 28(4). Wiley Online Library: 599–639. Spranghers, Thomas, Matteo Ottoboni, Cindy Klootwijk, Anneke Ovyn, Stefaan Deboosere, Bruno De Meulenaer, Joris Michiels, Mia Eeckhout, Patrick De Clercq, and Stefaan De Smet. 2017. “Nutritional Composition of Black Soldier Fly (Hermetia illucens) Prepupae Reared on Different Organic Waste Substrates.” Journal of the Science of Food and Agriculture 97(8). Wiley Online Library: 2594–600. doi:10.1002/jsfa.8081. St-Hilaire, Sophie, Craig Sheppard, Jeffery K. Tomberlin, Stephen Irving, Larry Newton, Mark A. McGuire, Erin E. Mosley, Ronald W. Hardy, and Wendy Sealey. 2007. “Fly Prepupae as a Feedstuff for Rainbow Trout, Oncorhynchus mykiss.” Journal of the World Aquaculture Society 38(1). Wiley Online Library: 59–67. doi:10.1111/j.1749-7345.2006.00073.x. Steinfeld, Henning, Pierre Gerber, Tom Wassenaar, Vincent Castel, Mauricio Rosales, and Cees de Haan. 2006. “Livestock’s Long Shadow: Environmental Issues and Options.” Renewable Resources Journal 24. Food & Agriculture Org. doi:10.1071/AN15729. Švácha, P. 2013. “BOOK REVIEW: Chapman RF 1998: The Insects. Structure and Function.” EJE 97(1). EJE: 32. Thornton, Philip K. 2010. “Livestock Production: Recent Trends, Future Prospects.” Philosophical Transactions of the Royal Society of London Series B Biological Sciences 365(1554). The Royal Society: 2853–67. Titi, H. H., M. J. Tabbaa, M. G. Amasheh, F. Barakeh, and B. Daqamseh. 2000. “Comparative Performance of Awassi Lambs and Black Goat Kids on Different Crude Protein Levels in Jordan.” Small Ruminant Research 37(1–2). Elsevier: 131–35. doi:10.1016/S0921-4488(99)00136-4. Tomberlin, Jeffery K., Peter H. Adler, and Heidi M. Myers. 2009. “Development of the Black Soldier Fly (Diptera: Stratiomyidae) in Relation to Temperature.” Environmental Entomology 38(3). Oxford University Press: 930– 34. doi:10.1603/022.038.0347. Tomberlin, Jeffery K., and D. Craig Sheppard. 2001. “Lekking Behavior of the Black Soldier Fly (Diptera: Stratiomyidae).” Florida Entomologist 84(4): 729–30. doi:10.2307/3496413. Tomberlin, Jeffery K., and D. Craig Sheppard. 2002. “Factors Influencing Mating and Oviposition of Black Soldier Flies (Diptera: Stratiomyidae) in a Colony.” Journal of Entomological Science 37(4). Georgia Entomological Society, Inc.: 345–52. doi:10.18474/0749-8004-37.4.345. Tomberlin, Jeffery K., D. Craig Sheppard, and John A. Joyce. 2002. “Selected Life-History Traits of Black Soldier Flies (Diptera: Stratiomyidae) Reared on Three Artificial Diets.” Annals of the Entomological Society of America 95(3). Oxford University Press: 379–86. doi:1​0.160​3/001​3-874​6(200​2)095​[0379​:SLHT​OB]2.​0.CO;​2. Tran, Gilles, Valérie Heuzé, and Harinder P. S. Makkar. 2015. “Insects in Fish Diets.” Animal Frontiers 5(2). Oxford University Press: 37–44. doi:10.2527/af.2015-0018. Tumuluru, Jaya Shankar. 2019. “Pelleting of Pine and Switchgrass Blends: Effect of Process Variables and Blend Ratio on the Pellet Quality and Energy Consumption.” Energies 12(7). Multidisciplinary Digital Publishing Institute: 1198. doi:10.3390/en12071198. Van Huis, A. 2013. “Potential of Insects as Food and Feed in Assuring Food Security.” Annual Review of Entomology 58. Annual Reviews: 563–83. doi:10.1146/annurev-ento-120811-153704.

192  Edible Insects Processing for Food and Feed Van Huis, A. 2016. “Edible Insects Are the Future?” Proceedings of the Nutrition Society 75(3). Cambridge University Press: 294–305. doi:10.1017/S0029665116000069. Van Huis, A. 2017. “Edible Insects and Research Needs.” Journal of Insects as Food and Feed 3(1). Wageningen Academic Publishers: 3–5. doi:10.3920/JIFF2017.x002. Vanhonacker, Filiep, Ellen J. Van Loo, Xavier Gellynck, and Wim Verbeke. 2013. “Flemish Consumer Attitudes towards More Sustainable Food Choices.” Appetite 62. Elsevier: 7–16. doi:10.1016/j.appet.2012.11.003. Veldkamp, Teun, and Guido Bosch. 2015. “Insects: A Protein-Rich Feed Ingredient in Pig and Poultry Diets.” Animal Frontiers 5(2). American Society of Animal Science: 45–50. doi:10.2527/af.2015-0019. Venter, P. A., and A. Schneemann. 2008. “Recent Insights into the Biology and Biomedical Applications of Flock House Virus.” Cellular and Molecular Life Sciences 65(17). Springer: 2675–87. doi:10.1007/s00018-008-8037-y. Verbeke, Wim, Thomas Spranghers, Patrick De Clercq, Stefaan De Smet, Benedikt Sas, and Mia Eeckhout. 2015. “Insects in Animal Feed: Acceptance and Its Determinants among Farmers, Agriculture Sector Stakeholders and Citizens.” Animal Feed Science and Technology 204. Elsevier: 72–87. doi:10.1016/j.anifeedsci.2015.04.001. Wadhwa, M., and S. P. M. Bakshi. 2013. “Utilization of Fruit and Vegetable Wastes as Livestock Feed and as Substrates for Generation of Other Value-Added Products.” RAP Publication 2013–04 (FAO) 4: 67. Wang, L., J. Li, J. N. Jin, F. Zhu, M. Roffeis, and X. Z. Zhang. 2017. “A Comprehensive Evaluation of Replacing Fishmeal with Housefly (Musca domestica) Maggot Meal in the Diet of Nile Tilapia (Oreochromis niloticus): Growth Performance, Flesh Quality, Innate Immunity and Water Environment.” Aquaculture Nutrition 23(5). Wiley Online Library: 983–93. doi:10.1111/anu.12466. Weissman, David B., David A. Gray, Hanh Thi Pham, and Peter Tijssen. 2012. “Billions and Billions Sold: Pet-Feeder Crickets (Orthoptera: Gryllidae), Commercial Cricket Farms, an Epizootic Densovirus, and Government Regulations Make for a Potential Disaster.” Zootaxa 3504(3504). Magnolia Press: 67–88. doi:10.11646/ zootaxa.3504.1.3. Wen, Zongguo, Yuanjia Wang, and Djavan De Clercq. 2016. “What Is the True Value of Food Waste? A Case Study of Technology Integration in Urban Food Waste Treatment in Suzhou City, China.” Journal of Cleaner Production 118. Elsevier: 88–96. Won, Ran, and OhSeok Kwon. 2009. “Relationship between Diet Composition and the Fecundity of Musca domestica.” Entomological Research 39(6). Wiley Online Library: 412–15. Yamanaka, Naoki, Kim F. Rewitz, and Michael B. O’Connor. 2013. “Ecdysone Control of Developmental Transitions: Lessons from Drosophila Research.” Annual Review of Entomology 58. Annual Reviews: 497–516. doi:10.1146/ annurev-ento-120811-153608. Zhang, Jibin, Ling Huang, Jin He, Jeffery K. Tomberlin, Jianhong Li, Chaoliang Lei, Ming Sun, Ziduo Liu, and Ziniu Yu. 2010. “An Artificial Light Source Influences Mating and Oviposition of Black Soldier Flies, Hermetia illucens.” Journal of Insect Science 10(1). Oxford University Press. doi:10.1673/031.010.20201. Zheng, Longyu, Tawni L. Crippen, Leslie Holmes, Baneshwar Singh, Meaghan L. Pimsler, M. Eric Benbow, Aaron M. Tarone, Scot Dowd, Ziniu Yu, and Sherah L. Vanlaerhoven. 2013. “Bacteria Mediate Oviposition by the Black Soldier Fly, Hermetia illucens (L.),(Diptera: Stratiomyidae).” Scientific Reports 3(1). Nature Publishing Group: 1–8. Zhou, Fen, Jeffery K. Tomberlin, Longyu Zheng, Ziniu Yu, and Jibin Zhang. 2013. “Developmental and Waste Reduction Plasticity of Three Black Soldier Fly Strains (Diptera: Stratiomyidae) Raised on Different Livestock Manures.” Journal of Medical Entomology 50(6). Oxford University Press: 1224–30. doi:10.1603/ME13021. Zhu, Feng Xiang, Wei Ping Wang, Chun Lai Hong, Ming Guang Feng, Zhi Yong Xue, Xiao Yang Chen, Yan Lai Yao, and Man Yu. 2012. “Rapid Production of Maggots as Feed Supplement and Organic Fertilizer by the Two-Stage Composting of Pig Manure.” Bioresource Technology 116. Elsevier: 485–91. doi:10.1016/j. biortech.2012.04.008. Zhu, Feng Xiang, Yan Lai Yao, Su Juan Wang, Rong Guang Du, Wei Ping Wang, Xiao Yang Chen, Chun Lai Hong, Bing Qi, Zhi Yong Xue, and Hong Quan Yang. 2015. “Housefly Maggot-Treated Composting as Sustainable Option for Pig Manure Management.” Waste Management 35. Elsevier: 62–67. doi:10.1016/j. wasman.2014.10.005. Zielińska, Ewelina, Monika Karaś, and Barbara Baraniak. 2018. “Comparison of Functional Properties of Edible Insects and Protein Preparations Thereof.” LWT - Food Science and Technology 91. Elsevier: 168–74. doi:10.1016/j.lwt.2018.01.058. Zinke, Ingo, Christina S. Schütz, Jörg D. Katzenberger, Matthias Bauer, and Michael J. Pankratz. 2002. “Nutrient Control of Gene Expression in Drosophila: Microarray Analysis of Starvation and Sugar-Dependent Response.” EMBO Journal 21(22). John Wiley & Sons, Ltd: 6162–73. doi:10.1093/emboj/cdf600.

Insect Farming for Feed Case Study

9

Marco Meneguz and Sihem Dabbou Contents 9.1 Introduction 194 9.2 Strategy of the Company 195 9.2.1 Modular Approach 195 9.2.2 Local Approach 196 9.2.3 Energetic Approach 196 9.2.4 Contract Approach 197 9.3 The Black Soldier Fly 197 9.3.1 Biology 197 9.3.2 Why Black Soldier Fly? 198 9.4 Products and Services of BEF Biosystems 199 9.5 Insect Farming 200 9.5.1 Side Characteristics 200 9.5.2 Feeding System 201 9.6 Insect Mass Production Technologies 201 9.6.1 Cages for Reproduction 201 9.6.2 Nursery 201 9.6.3 Fattening System 203 9.7 Environmental Impact of Our Plant 203 9.8 Strategies of Industry Marketing 206 9.9 Investments in the Insect Sector 207 9.10 Future Trends in Insect Products 207 9.11 Future Prospects for BEF Biosystems 207 9.12 Conclusion 208 References 208

DOI: 10.1201/9781003165729-9

193

194  Edible Insects Processing for Food and Feed

9.1 INTRODUCTION How much will the demand for animal proteins have risen by 2050? Estimations about population growth predict that by 2050 there will be over 9 billion people to feed on planet Earth. This rapid growth leads to a situation of food insecurity, particularly in developing countries. Globally, around one-third of all the food produced for human consumption in the world (1.3 billion tons of edible food) is lost and wasted across the entire supply chain every year (FAO, 2019). Fruit and vegetable processing by- and co-products contribute to about 44% of global food waste (Salim et al., 2017) and are considered a major problem in terms of the efficiency and sustainability of food supply chains. These wastes are considered the largest contributor to resource loss, such as land, water, and energy use, as well as biodiversity loss (De Baan et al., 2013). They also lead to greenhouse gas emissions (IPCC, 2015) and pose a potential threat to the environment and economy (FAO, 2019). The recycling, reducing and re-use of fruit and vegetable processing by- and co-products streams is one of the priority objectives of the European Union in order to improve resource efficiency across the economy and to achieve sustainable food systems (European Commission, 2017). To support European businesses and consumers in making the transition to a stronger and more circular economy (CE), the European Commission adopted an ambitious new CE Package (European Commission, 2015). This package includes legislative proposals on waste management and waste avoidance, with long-term targets to reduce landfill and increase recycling and reuse. The development of innovative circular-system solutions and technologies have been proposed, but their sustainability has not been assessed. The transformation of wastes into value-added products can contribute to reducing the ecological and water footprint associated with crop cultivation and give new economic and environmental perspectives (Gassara et al., 2013; Pagana et al., 2014). Insects have the ability to convert biogenic waste streams into high-value protein products with advantageous feed conversion efficiency. This bioconversion constitutes a new approach and an interesting example of sustainable CE (European Commission, 2019). Insect larvae requires the establishment of an appropriate nutritional strategy, and they can be successfully reared on a number of different organic substrates (fruit and vegetable waste, manure, cereals, etc.), thereby transforming waste into nutrient-rich raw materials (Makkar et al., 2014; Meneguz et al., 2018). According to Regulation (EU) 2017/893 (European Commission, 2017), only vegetable waste can be used as feed if the insects are intended for food production, except for a few animal protein sources: milk, eggs, honey, rendered fat and non-ruminant blood products. It is prohibited to feed insects with animal by-products like slaughterhouse products, manure, or waste streams like household waste (European Commission, 2009, 2016, Federal Agency for the Safety of the Food Chain, 2019). The feeding of insect larvae drastically reduces waste and the harvested larvae can be used as valuable raw materials in the animal feed industry. This technology has received attention in recent years due to the business opportunities it offers, which simultaneously address several challenges of modern society: hygiene issues arising from the lack of waste management, unemployment in urban areas, and an increased demand for sustainable feed for the ever-growing aquaculture and aviculture sectors. The positive outputs of this procedure are several: reduction of waste management costs (landfilling, anaerobic digestion, incineration), lower resource use than other protein and fat productions and a value gain from the sale of insect derived products. As a result, the feed industry will need to find new strategies to maintain the process sustainably along the whole supply chain. Since July 2017, new EU legislation allows the feeding of processed animal protein (PAP) from insects to aquaculture animals. In this context, the insect farming industry for animal feed is a relatively novel activity and could play an important role in closing the loop ‘from fork to farm’ by transforming food losses and waste

9  •  Insect Farming for Feed  195 into an additional supply of protein feed for livestock and aquaculture. In fact, insects are part of the natural diet of chickens and fish, which are highly motivated to interact with them and consume them (Ipema et al., 2020). Hermetia illucens and Tenebrio molitor are considered innovative protein sources in animal nutrition (Gasco et al., 2020). In fact, many species of freshwater fish are insectivorous, so insects are already part of their main diet. Recently, it has been discovered that low doses of insect meal can also act as gut prebiotics, making them an alternative to antibiotics. Waithanji et al. (2020) has demonstrated that using insects for feed production is likely to solve the current challenge in the development of a sustainable and productive supply chain in both the aquaculture and poultry sectors. The insect farming industry is already present in Europe today, producing more than 5,000 tons of insect protein for pet foods and aquaculture feed on the basis of by-products from the agri-food industry (IPIFF, 2019). The Italian insect feed and food market is projected to grow. The overall growth of the insect rearing sector in Italy has been largely attributed to the region’s insect feed sector. Nowadays, small companies are producing insects for the pet food market as food for reptiles, birds and amphibians. Moreover, new small companies are starting to include insect meal in recipes for dog and cat food. In this scenario, one of the companies that is currently converting its production according to CE principles is considered a case study. BEF Biosystems (BEF) is a company founded in 2017 in Torino, Piedmont (Italy) with the aim of breeding black soldier flies (Hermetia illucens, L.; [BSF]). The idea behind BEF is to create some small-scale plants on Italian territory. This chapter will try to analyse the BEF approach, the goals for the future of the insect market, and possible future developments for insect utilisation in the European feed market.

9.2  STRATEGY OF THE COMPANY The BEF’s idea is to have a small-scale plant on Italian territory using local vegetable by-products. The insect farm is called BugsFarm™. There are two kinds of BugsFarm™; one involves breeding larvae and adults, and the other only larvae. The customers for this model are biogas plant owners, as we explain in the next paragraph, and feed-mill producers. The strategy of the company can be divided into four points: • • • •

Modular approach Local approach Energetic approach Contract approach

9.2.1 Modular Approach The Bioconverter, a patented machine for breeding BSF larvae, is the innovative aspect of the BugsFarmTM approach (Figure 9.1). The Bioconverter is a module that can allow the treatment of up to 500 kg of byproducts and the production of 100 kg of larvae per cycle. It is a highly technological machine that allows the main parameters for BSF breeding – temperature, humidity, and carbon dioxide – to be measured remotely. The modularity of this technology allows workers to avoid direct contact with the larvae, food, and environmental conditions, reducing worker and microbiological risks.

196  Edible Insects Processing for Food and Feed

FIGURE 9.1  Bioconverter modules in the greenhouse at the first Italian BugsFarm™. (Courtesy of Giuseppe Tresso.)

9.2.2 Local Approach The main business model of BEF is to work at small and medium scales, depending on the territory and heterogeneity of the Italian situation. The size and dimension of the plant (BugsFarm™) are modulated on a small number of by-products. The by-products are selected from the area near the BugsFarm. Furthermore, the size of the BugsFarm™ is modulated to meet the minimum requirement for a fast and easy authorization process. Moreover, the customers of the BugsFarm™ produce live larvae at a local scale and the BEF service collects the larvae produced, reducing the cost of substrate transport and enabling local production.

9.2.3 Energetic Approach The key to the success of the insect industry is to have low cost heating for the different life cycle stages of BSF. The idea of BEF is to have a connection to a biogas plant. Italian biogas plants are estimated to generate more than 1600 drivers in 2020 (Istat, 2020). Biogas production is regulated by the European Directive of 2008 (Dir. 2008/98/EC). Biogas plants have three main products: gas, energy and heat (under hot water accumulation). BugsFarm™ can connect to a biogas plant through the hot water circuit before the heat sink, thus exploiting the heat from the biogas plant and increasing the circularity of energy production. Furthermore, the BugsFarm™ includes solar panel coverage to reduce energy consumption and include a green, sustainable energy intake.

9  •  Insect Farming for Feed  197 The breeding modules consume a small amount of energy compared to the highly automated systems used in vertical farming.

9.2.4 Contract Approach The main circular approach should include a legal contract between the components of the network created and managed by BEF. The network approach is regulated by Italian Law of April 9, 2009, n. 33 on the agricultural network contract. Therefore, the following parts must be defined in the network contract: • The objectives of innovation and increasing the competitive capacity of the participants, as well as the methods agreed upon to measure progress towards these objectives • The specific objectives that constitute the prerequisite for identifying the activities necessary for the achievement of the general objectives • A network programme that outlines the enunciation of the rights and obligations of each participant, as well as the methods for achieving the common purpose • The methods for distributing the common agricultural product BEF acts as the lead partner of the network contract and manages the input and output of the production chain. The other partners in the network are the customers: the biogas plant owners, the feed millers, and the by-product producers.

9.3  THE BLACK SOLDIER FLY 9.3.1 Biology The black soldier fly (Hermetia illucens L.), is a Diptera (Stratiomydae family) originated in the tropical, subtropical and warm temperate zones of America and is currently naturalised worldwide (Wang and Shelomi, 2017). In optimal environmental conditions, the BSF larvae hatch after four days, and one larva can process from 25 to 500 mg of substrate/daily. The life cycle of the BSF encompasses different stages: egg, larva, prepupa, pupa and adult. The eggs are elongated in shape and yellow-white in color. The female usually lays a clutch of eggs in a dark place near organic matter. Females usually lay between 546 and 1,505 eggs (Booth and Sheppard, 1984) in a single clutch. The eggs hatch in 102 to 105 hours at 24°C (Booth and Sheppard, 1984) while at 27°C they hatch in ≈72 hours. Higher temperatures than 27°C do not further accelerate the duration of incubation. Once the eggs hatch, the larvae find whatever waste they can and immediately start to consume it. The BSF larva is a very resilient organism and has the ability to extend its life cycle under unfavourable conditions. The larval stage is the only stage during which the BSF feeds. During this time of larval development, they accumulate enough fat reserves and proteins to allow the larvae to pupate (Dortmans et al., 2021). The larvae pass through five larval stages or instars in a variable length of time (from a few weeks to a few months), depending on the environmental conditions, in particular the temperature (Tomberlin et al., 2009), the quantity and quality of the substrate (Makkar et al., 2014; Oonincx et al., 2015, Meneguz et al., 2018) and the BSF strains (Zhou et al., 2013). The larvae are pale white in colour with a small black head containing their mouthparts (Newton, 2005).

198  Edible Insects Processing for Food and Feed

FIGURE 9.2  Black soldier fly prepupae. (Courtesy of Marco Meneguz.)

FIGURE 9.3  Products of a BugsFarm™. (Courtesy of Marco Meneguz.)

During the last stage, the larvae stop eating and reach the final larval stage, called the prepupa (Figure 9.2). During the prepupa stage, the larva replaces its mouthpart with a hook-shaped structure and becomes dark brown to charcoal grey in colour (Dortmans et al., 2021). The prepupae leave the substrate for a high, safe, clean space in a stage called “self-harvesting” that removes an otherwise labor-intensive step from their farming (Wang and Shelomi, 2017). During the pupal stage, a series of metamorphoses modifies the internal parts of the body, which takes between 7 days and 3 months (diapause) in temperate regions (Holmes et al., 2013). Pupation ends when the fly emerges from its pupal shell, which takes less than 5 minutes. When adults hatch from the exuviae, they use their time to mate and lay eggs before they die. The adult does not feed since its digestive apparatus is atrophied. Mating occurs only in specific singular environmental conditions; the flies mate in the air and then fall to the ground (Tomberlin and Sheppard, 2001).

9.3.2 Why Black Soldier Fly? The black soldier fly, especially in the larval stage (BSFL), is one of the most promising insect species for livestock feed. BSFL meal and oil are already considered an animal-grade alternatives to fish meal and fish oil used to feed carnivorous fish, poultry, and pigs (Gasco et al., 2019) due to their high protein and lipid content. BSFL has not only demonstrated its potential for feed production in the aquaculture and

9  •  Insect Farming for Feed  199 poultry sector (Gasco et al., 2019) but also its ability to bioconvert waste from agricultural and farming processes into soil organic fertilizer (frass and exuviae of BSFL; leftovers; Klammsteiner et al., 2019). Therefore, it is possible to use BSFL leftovers as a compost-like amendment due to the presence of a diverse range of microorganisms. From the perspective of the bioeconomy, the potential of this insect seems to be huge in terms of implementing a sustainable conversion for the feed and food supply chains. BSFL processing can reduce the cost of waste transport and space requirements for landfills if the mechanism is applied at the end of biowaste generation (Dortmans et al., 2021). The benefits of using insect-based feed also include economic aspects: the market price of BSF decreased from 5.6 $/kg in 2016 to 2.5 $/kg in 2018 (Ferrer Llagostera et al., 2019). This tendency to lower prices results in increased competitiveness against common feeding products, which attracts producers and stakeholders. As is the case for other insect species, the chemical composition of BSFL depends on the rearing substrates and on the larval development stage (Barragan-Fonseca et al., 2017). The BSF live larvae contain a high amount of moisture, with a dry matter (DM) content of about 30% as fed. The crude protein (CP) and ether extract (EE) contents of BSF live larvae are around 15% and 12% DM, respectively (Veldkamp and van Niekerk, 2019). In order to be able to manage fat and protein sources separately in feed formulations, BSF manufacturers have started to produce not only a full fat BSFL meal, but also partially and highly defatted BSFL meals (Schiavone et al., 2017). Besides its nutritional value, BSFL meal is rich in bioactive compounds, including chitin, medium chain fatty acids (C6-12) and antimicrobial peptides. The amino acid (AA) profile of BSFL is slightly influenced by substrate, meaning that it is always particularly rich in lysine (6%–8% of CP content). Compared to soybean meal, BSFL reared on swine manure presents a similar AA profile in terms of lysine, leucine, phenylalanine, and threonine, while higher amounts of alanine, methionine, histidine, and tryptophan have been observed (Barragan-Fonseca et al., 2017). The fatty acid (FA) profile of BSFL is mainly influenced by the FA profile of the substrate where the larvae are reared (Meneguz et al., 2018). Generally, the amount of saturated FA (SFA) in BSFL is around 58%–72%, while the monounsaturated fatty acid (MUFA) and polyunsaturated fatty acids (PUFA) content represent 19%–40% of the total fat content (Barragan-Fonseca et al., 2017). Concerning the environmental effects of climate change, BSFL emits low levels of greenhouse gases (GHG) (Oonincx et al., 2015). Furthermore, only a minimal amount of land is required to produce 1 kilogram of protein (Salomone et al., 2017). This becomes important in terms of a circular economy, as insects can be raised in the same area as the livestock supply chain, thereby reducing the costs and environmental impact associated with transportation.

9.4  PRODUCTS AND SERVICES OF BEF BIOSYSTEMS The main input products for the BugsFarm™ are: • A full plant for BSF larvae rearing (a bioconverter as needed, a system for grinding the material, and storage for feed) • Neonate BSF for the biogas plant owner • A formulated diet for the BSF larval fattening process The main output products (Figure 9.3) from the BugsFarm™ are: • Live larvae (the main market being poultry and layer hens) • Dry larvae (the main market being the pet food market) • Frass (the main market being the floro nursery soil producers and fertilizer for the main crops)

200  Edible Insects Processing for Food and Feed High insect biomass can be economically produced by utilising appropriate breeding technologies and effective habitat management systems that are based on the biology and habitat characteristics of target insect species, including factors such as diet, temperature, light/illumination, humidity, ventilation, rearing containers, and water facilities (Dzepe et al., 2021; Chia et al., 2018).

9.5  INSECT FARMING 9.5.1 Side Characteristics There are two main approaches to BSF breeding: the centralised plant and the diffused plant approach. • Centralised plant approach: this insect feeding system requires a huge amount of collected substrates (up to hundreds of thousands of tons) in a single location, as well as a significant initial investment to achieve the break-even point. This is the main approach used in northern Europe, where many stakeholders from France, the Netherlands, and Germany recoup significant investments of around 50–140 million of euros each (allaboutfeed​.n​et). The centralised approach to insect breeding requires highly technological machinery and highly educated of workers for the maintenance and management of the plant. The issues required for management and the biological issues connected to high insect density are mitigated by the fact that this type of plant allows for higher production with reduced costs at the end of the process. This approach does not allow the recovery of substrates from small producers because of logistic-cost constraints. The range of action of this type of plant covers several hundred kilometers around the collection site. • Diffused-plant approach: one of the most difficult aspects of breeding BSF is reducing the cost of substrate recollection. The diffused plant approach requires a small amount of substrate (few miles tons) compared to a centralised one. The supply of the substrates requires little effort and also allows the substrate to be collected from small producers. This kind of approach avoids the dispersion of these resources, allows the recovery of nutrients and increases the circularity of the system. Moreover, local communities can contribute to the development of local-feed production and waste management, as reported in Kenia, where 80,000 BSF producers are located on the national territory (roc​kefe​ller​foun​datio​n​.org). Similar approaches are described in Europe by stakeholders from the United Kingdom, Germany, and Italy. The diffused-plant approach involves a small plant using plastic boxes or containers; in the case study, the bioconverter is the module used for breeding larvae and can produce 500 kg of diet per cycle. BEF Biosystems presents a case study of its own plant, where the approach is based on a diffused plant connected by a network that manages the plants. The BEF plant is based in the local community and was founded on agricultural tradition. The collection points for the substrate used in the insect diet are located not far from the BugsFarm™, just a few kilometres away. The supply of different substrates allows for the recovery and reuse of waste products from agricultural and agri-industrial production that are usually disposed of or used for compost production at a cost to the public community. The bioconverter units are positioned in lines, and the horizontal position is the main component of this insect rearing system. The worker can move around using the classic tractor, making it easier to operate. All the main work is simplified to allow low labour intensity. The system can be connected to a biogas plant that produces hot water. This system reduces heating costs to zero and increases cogeneration of the biogas plant, increasing the circularity of the system. No competition arises from this approach between the two production systems (biogas and insect breeding); instead, synergy and coworking are important for this approach. The BEF’s BugsFarm™ allows to no-knowledge property of a biogas plant to have a full plant with a capacity of 2,000 tons of substrates and residues bioconversion using BSF.

9  •  Insect Farming for Feed  201 The main bioconverter management operations can be summarised as feeding, “seeding” of larvae, and harvesting once the substrate has been bioconverted and the larvae have grown.

9.5.2 Feeding System Black soldier flies are reared around the world using organic wastes such as manure (wastes of pigs, poultry or cattle) (Xiao et al., 2018; Miranda et al., 2020), rotting fruits and vegetables (Jucker et al., 2017), winery and brewery by-products (Jucker et al., 2017; Meneguz et al., 2018), coffee bean pulp (OspinaGranobles and Carrejo-Gironza, 2021), distillers’ grains (Chia et al., 2018), fish offal (St‐Hilaire et al., 2007) and food wastes (Nguyen et al., 2015). European legislation is more restrictive compared to regulations in other parts of the world, and the precautionary principle is a milestone for European regulations. The precautionary principle means that the safety and health of products for humans, animals, and vegetation must be monitored. Regulation n.1017/2017 EC lists the main raw materials that are permitted for use as feed in European countries. The new list includes the raw material that can be used for rearing insect and BSFL. Only by-products and residues derived from the production of vegetables, milk and eggs can be used for feeding insects for the purposes of producing feed. The use of animal, catering, restaurant, manure, and organic waste from recycling collection is not allowed for BSF feeding (European Commission, 2017). The main products that can be collected on Italian territory, considering the case study, are byproducts and residues from fruit and vegetable processing. Moreover, brewery and winery by-products can be used, as reported by Meneguz et al. (2018). Cattaneo et al. (unpublished data) reported around 2,000 tons of by-product suitable for BSF breeding available in the vicinity of the first BugsFarm™. The mix of the substrate has been studied for a long time by the in-house research centre at BEF Biosystems, and a complete and balanced diet is now used for rearing BSF larvae. The diet allows the larvae to reach a good size for harvesting.

9.6  INSECT MASS PRODUCTION TECHNOLOGIES The main technologies used for BSF rearing are related to different approaches and applications of BSF products. The production of BSF is divided into three stages of the cycle.

9.6.1 Cages for Reproduction The cages used for reproduction have different sizes and shapes and usually require a flight space (Figure 9.4; Figure 9.5).​ Black soldier fly adults do not require feeding during their short lives, but this can be extended to two weeks using water administration (May, 1961; Macavei et al., 2020). The eggs are collected using a wooden stick that allows the females to lay eggs on the dark sides of the holes (Booth and Sheppard, 1984). As reported by May (1961), the eggs are yellow-white in colour and have an elongated shape. The eggs need a certain temperature (27°C) and relative humidity (70%) for hatching in 72h (Tomberlin et al., 2002). The climatic chamber must have a control system to manage and regulate the temperature and humidity of the air and a ventilation system is required to maintain good air quality.

9.6.2 Nursery After the hatching period, the neonate larvae have a size near to 0.1 mm (Figure 9.6; 9.7) and require a food source immediately to begin growing. Depending on diet and environmental conditions, the larvae

202  Edible Insects Processing for Food and Feed

FIGURE 9.4  a, b. Black soldier fly cages for adult reproduction at the BEF Biosystems research center. (Courtesy of Marco Meneguz.)

FIGURE 9.5  Black soldier fly adults mating at the BEF Biosystems resaerch center. (Courtesy of Marco Meneguz.)

9  •  Insect Farming for Feed  203 will pass through the first three stages within the first five days. The BEF breeding system creates optimal conditions (27°C and 70% RH) (Chia et al., 2018) during the first stages of life and produces the best neonate larvae for the next period. The bioconverter is used as a nursery and ensures high levels of control and monitoring. Once the larvae reach the end of this phase of the cycle, they must be collected and sieved using a vibro wave, which is used for traditional seed sieving​​.

9.6.3 Fattening System The small larvae must to be transferred to the bioconverter unit for the fattening process, where the machine can be loaded with 500 kg of substrates collected from the local area. The programmable logic controller (PLC) of the Bioconverter machine uses a complex algorithm to control temperature and ventilation. The cycle takes seven days to bioconvert the larvae’s diet, and at the end, the larvae will be collected for further sieving process before the processing of the dry larvae for fat extraction.

9.7  ENVIRONMENTAL IMPACT OF OUR PLANT BEF Biosystems performed a preliminary analysis of the global warming potential (GWP) of the whole process from substrate to larval killing. As can be seen from the GWP analysis using the Sankey diagrams (unpublished data), the components that generate a greater emission of climate-altering gases are energy uses (electrical + thermal) and substrate production and processing. Energy use accounts for 36% to 52% of the total emissions, while the production and processing of substrates ranges from 34.6% for waste from fresh-cut fruits and vegetables to 54.3% for brewery by-products and 41.3% for by-products from legume production. The amount of diesel used by tractors for business operations (e.g. collection and distribution of substrates, harvesting of larvae, cleaning, etc.) and the materials used contribute to a lesser extent than the other three components seen previously and are both always below 3% in all three scenarios. This is mainly due to the possibility of amortising environmental impacts over the years. This is achieved through high annual productivity (100 kg of larvae per cycle with up to 52 cycles per year) and limited use of the farm tractor. The tractor only needs to run for a few hours (about three) to complete all operational needs, such as adding the substrate and larvae at the beginning of the cycle and sieving residual frass and fattened larvae at the end. As far as electricity is concerned, the predominant factor is the consumption due to the cold room for killing larvae, which absorbs 4.5 kWh and operates for the entire time of the production cycle, requiring 52.1% of the energy absorbed by the plant. This impact is hardly avoidable or amortisable since cold rooms are used to kill the larvae before their subsequent processing (e.g. flour production, bagging, dehydration, extractions, etc.) as well as containing the most perishable substrates. For their part, the bioconverters contribute to 27.8% of the emissions due to the use of electricity for the ventilation system, which is essential to maintain a correct internal humidity and, at the end of the cycle, to guarantee an optimal reduction of the content of substrate water, a fundamental step is to ensure efficient separation of the larvae from the growth material. As far as thermal energy is concerned, the heat must be sufficient to warm the adult cages and bioconverters to a temperature that is suitable for the growth of larvae. This temperature ranges between 27–40°C. A method devised by the company that allows the reduction of both operating costs and environmental impact is to associate BugsFarms with companies that treat by-products or specially cultivate substrates for the production of biogas. Biogas production is an anaerobic digestion process through which methane and carbon dioxide are mainly produced from organic material (corn, urban waste, manure, etc.).

204  Edible Insects Processing for Food and Feed

FIGURE 9.6  One gram of black soldier fly neonate larvae at