Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Latin-American Seeds - Agronomic, Processing and Health Aspects

Download as pdf or txt
Download as pdf or txt
You are on page 1of 465

Latin-​American Seeds

Over the last few years, Latin-​American seeds have gained increased importance
(in part due to the increased demand for gluten-​free foods). Worldwide demand
for Latin-​American seeds and grains has risen in high proportion. In parallel,
research into seeds and grains from this region in all relevant fields has inten-
sified. Latin-​American Seeds: Agronomic, Processing and Health Aspects
summarizes recent research into Latin-​American crops regarding their agro-
nomic and botanical characteristics, composition, structure, use, production,
technology, and impact on human health.
Latin-​American cultivars studied here are included in the groups of cereals,
pseudo-​cereals, oilseeds, and legumes that are used in a great variety of innova-
tive and traditional foods. The main crops that are covered in this book are
Latin-​American maize (Zea mays L.), amaranth (Amaranthus spp), quinoa
(Chenopodium quinoa Willd), kañiwa (Chenopodium pallidicaule Aellen), chia
(Salvia hispanica L.), sacha inchi (Plukenetia volubilis), and legumes such as black
turtle and common beans (Phaseolus vulgaris) and tarwi (Lupinus mutabilis).

Key Features:
• Contains updated information about recent research work on Latin-​
American crops.
• Includes a variety of Latin-​American plant species that are used in a
great variety of innovative and traditional foods.
• Addresses a wide range of topics related to agronomy, plant physiology,
and nutritional and technological properties, processing, fractionation,
and development of new products for human health.
The book also provides information about milling, fractioning, and oil extraction,
and on the elaboration of various derived traditional and novel food products.
The nutritional and health implications of these Latin-​American crops are also
widely addressed.
Food Biotechnology and Engineering
Series Editor:
Octavio Paredes-​López

Volatile Compounds Formation in Specialty Beverages


Edited by Felipe Richter Reis and Caroline Mongruel Eleutério dos Santos

Native Crops in Latin America: Biochemical, Processing, and


Nutraceutical Aspects
Edited by Ritva Repo-​Carrasco-​Valencia and Mabel C. Tomás

Starch and Starchy Food Products: Improving Human Health


Edited by Luis Arturo Bello-​Perez, Jose Alvarez-​Ramirez, and Sushil Dhital

Bioenhancement and Fortification of Foods for a Healthy Diet


Edited by Octavio Paredes-​López, Oleksandr Shevchenko, Viktor Stabnikov, and
Volodymyr Ivanov

Latin-​American Seeds: Agronomic, Processing and Health Aspects


Edited by Claudia Monika Haros, María Reguera, Norma Sammán, and
Octavio Paredes-​López

For more information please visit https://www.routledge.com/Food-Biotechnology-


and-Engineering/book-series/CRCFOOBIOENG
Latin-​American Seeds
Agronomic, Processing and
Health Aspects

Edited by
Claudia M. Haros, María Reguera,
Norma Sammán, and Octavio Paredes-​López
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 Claudia M. Haros, María Reguera, Norma Sammán, and Octavio Paredes-​López
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.copyri​ght.
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: Haros, Claudia Monika, 1966– editor.
Title: Latin-American seeds : agronomic, processing and health aspects /
Claudia Monika Haros, María Reguera, Norma Sammán, and Octavio Paredes-López.
Description: First edition. | Boca Raton, FL : CRC Press, 2023. |
Series: Food biotechnology and engineering |
Includes bibliographical references and index.
Identifiers: LCCN 2022040886 (print) | LCCN 2022040887 (ebook) |
ISBN 9780367531454 (hbk) | ISBN 9781032367392 (pbk) |
ISBN 9781003088424 (ebk)
Subjects: LCSH: Crops–Latin America. | Crops–Health aspects–Latin America.
Classification: LCC SB99.L29 L38 2023 (print) |
LCC SB99.L29 (ebook) | DDC 338.1098–dc23/eng/20221118
LC record available at https://lccn.loc.gov/2022040886
LC ebook record available at https://lccn.loc.gov/2022040887
ISBN: 9780367531454 (hbk)
ISBN: 9781032367392 (pbk)
ISBN: 9781003088424 (ebk)
DOI: 10.1201/​9781003088424
Typeset in Kepler Std
by Newgen Publishing UK
Contents

Series Preface ix
Preface xiii
About the Editors xv
Contributors xix

1 Agronomical Characterization of Latin-​American


Crops 1
Sven-​Erik Jacobsen, Angela Miranda, Doris Chalampuente-​Flores,
Juan Pablo Rodríguez, Luis Rodolfo Montes Osorio, Alain P. Bonjean,
Nete Kodahl, César Tapia Bastidas and Marten Sørensen

2 Genotype and Environment as Key Factors


Controlling Seed Quality in Latin-​American
Crops 53
Luisa Bascuñán-​Godoy, María Reguera, Ángel Mujica,
Néstor Fernández del Saz, Carolina Sanhueza, Catalina Castro,
José Ortiz, Gabriel Barros, José Delatorre-Herrera, Karina B. Ruiz,
Teodoro Coba de la Peña, Enrique Ostria, Enrique Martínez,
and Susana Fischer

3 The Outlook for Latin-​American Crops:


Challenges and Opportunities 91
Nieves Fernández-​García, Inmaculada Román-​García,
and Enrique Olmos

v
vi Contents

4 Structure and Composition of Latin-​American


Crops 119
Barbara Borczak, José María Coll Marqués,
Octavio Paredes-​López, and Claudia M. Haros

5 Latin-​American Crops in Gluten-​Free


Applications 161
Silvia V. Melgarejo-​Cabello, Jehannara Calle-​Domínguez,
María Alejandra Giménez, Claudia M. Haros, and
Ritva Ann-​Mari Repo-​Carrasco-​Valencia

6 Fractionation of Seeds or Grains 205


Marianela Capitani, Adriana Scilingo, Edgardo Calandri,
María Alejandra Giménez, Marcela Lilian Martínez,
Vanesa Ixtaína, Nancy Chasquibol Silva, M. Carmen Pérez Camino,
Natalia Bassett, Víctor Delgado-​Soriano, Ritva Ann-​Mari
Repo-​Carrasco-​Valencia, Claudia M. Haros, and Mabel Tomás

7 Food Uses of Selected Ancient Grains 259


Claudia M. Haros, Marcela Lilian Martínez,
Bernabé Vázquez Agostini, and Loreto A. Muñoz

8 Nutritional Composition, Bioactive and


Anti-​Nutritional Compounds of Latin-​American
Crop Grains 303
Norma Sammán, María C. Rossi, Sonia Calliope, and
Ritva Ann-​Mari Repo-​Carrasco-​Valencia

9 Contributions from Latin-​American Grains to


Nutrition and Health 341
Carla Motta, Norma Sammán, and Isabel Castanheira
Contents vii

10 Ingredients of High Nutritional Value Obtained


from Latin-​American Crops through
Biotechnology 371
Manuel Oscar Lobo, Ana Laura Mosso, María Dolores Jiménez,
and Norma Sammán

11 Current Position of Legislation on Latin-​American


Grains and Its Regional Socioeconomic Impact 401
María Dolores Jiménez, Ana Laura Mosso, Claudia M. Haros,
and Norma Sammán

Index 427
Series Preface

BIOTECHNOLOGY –​ OUTSTANDING FACTS


The beginning of agriculture started about 12,000 years ago and ever since has
played a key role in food production. We look to farmers to provide the food we
need but at the same time, now more than ever, to farm in a manner compatible
with the preservation of the essential natural resources of the earth. Additionally,
besides the remarkably positive aspects that farming has had throughout his-
tory, several unintended consequences have been generated. The diversity of
plants and animal species that inhabit the earth is decreasing. Intensified crop
production has had undesirable effects on the environment (e.g., chemical con-
tamination of groundwater, soil erosion, exhaustion of water reserves). If we do
not improve the efficiency of crop production in the short term, we are likely to
destroy the very resource base on which this production relies. Thus, the role of
so-​called sustainable agriculture in the developed and underdeveloped world,
where farming practices are to be modified so that food production takes place
in stable ecosystems, is expected to be of strategic importance in the future; but
the future has already arrived.
Biotechnology of plants is a key player in these scenarios of the 21st cen-
tury. Nowadays, molecular biotechnology in particular is receiving increasing
attention because it has the tools of innovation for agriculture, food, chem-
ical, and pharmaceutical industries. It provides the means to translate an
understanding of, and ability to modify, plant development and reproduction
into enhanced productivity of traditional and new products. Plant products
from seeds, fruits, and plant components and extracts are being produced with
better functional properties and longer shelf life; and they need to be assimilated
into commercial agriculture to offer new options to small, and more than small,
industries and finally to consumers. Within these strategies it is imperative to
select crops with larger proportions of edible parts as well, thus generating less
waste; it is also imperative to consider the selection and development of more
environmentally friendly agriculture.
The development of research innovation for products is progressing; how-
ever, the constraints of relatively lengthy delivery times to reach market, intellec-
tual property rights, uncertain profitability of products, consumer acceptance,
and even caution and fear with which the public may view biotechnology are

ix
x Series Preface

tempering the momentum of all but the most determined efforts. Nevertheless,
it appears uncontestable that the food biotechnology of plants and microbials is
and will emerge as a strategic component for providing the food crops and other
products required for human well-​being.

FOOD BIOTECHNOLOGY AND ENGINEERING SERIES /​


OCTAVIO PAREDES-​LÓPEZ, PH.D, SERIES EDITOR
The Food Biotechnology and Engineering Series aims to address a range of
topics around the edge of food biotechnology and the food microbial world. In
the case of foods, it includes agrigenomics, molecular biology, genetic engin-
eering, metabolic aspects, omic sciences, chemistry, nutrition, medical foods
and health ingredients, and processing and engineering with traditional-​and
innovative-​based approaches. Environmental aspects and strategies for pro-
ducing green foods cannot be left aside. At the world level there are foods and
beverages produced by different types of microbial technologies to which this
series will give attention. It will also consider genetic modifications of microbial
cells to produce nutraceutical ingredients, and advances of investigation into
the human microbiome as related to diets.

LATIN-​AMERICAN SEEDS: AGRONOMIC,


PROCESSING AND HEALTH ASPECTS. EDITORS:
CLAUDIA M. HAROS, MARÍA REGUERA, NORMA
SAMMÁN, AND OCTAVIO PAREDES-LÓPEZ
This book is divided into eleven chapters dealing with the agronomic
characteristics of crops from the Latin-​American region; seed quality from crops
as influenced by genotype and environmental factors; challenges and oppor-
tunities facing Latin-​American seeds; characteristics and composition of crops
and their seeds; gluten-​free applications in products from these crops; main
components of their grains such as types of flour, oil, proteins, fibers, and starch;
food uses of ancient grains; bioactive compounds, peptides, and anti-​nutritional
factors present in these crops; potential and outstanding contributions from
these regional crops in nutrition and health at world level; components and
ingredients from these crops for nutrition and nutraceutical purposes; the last
chapter reports on current legislation for the rational use of Latin-​American
seeds, and their updating as compared to some advanced experiences in other
countries from other areas of all continents.
Series Preface xi

This multi-​author publication has been produced by 52 scientists and tech-


nical experts who come from many academic institutions and organizations
located in Latin America: Argentina, Chile, Cuba, Ecuador, Guatemala, Mexico,
and Peru; in Europe: Denmark, France, Portugal, and Spain; and in the Arab
world: the United Arab Emirates.
This book is of interest to academicians, food scientists, food technologists,
biotechnologists, agricultural specialists at universities, technical people in pri-
vate and public organizations, and baccalaureate and graduate students.
Thanks are due to the editors, professors Norma Sammán, María Reguera,
and Claudia M. Haros, and to all of the authors for their excellent contributions
to this book and to Sofia Blanco Haros for her contribution towards the writing
of the Alt Text. Acknowledgments are also due to the editorial staff of CRC Press,
especially to Mr. Stephen Zollo and to Ms. Laura Piedrahita.
Preface

The authors and editors of this book belong to the CYTED Ia ValSe-​Food Group
(Iberoamerican Valuable Seeds, Ref. 119RT0567) and Chia-​Link Network. This
book gathers recent investigation and experience of Ia ValSe-​Food Group on
Latin-​American valuable seeds and other crops, and provides comprehensive
and up-​to-​date knowledge within several relevant fields of food science and
agriculture.
The main crops included in this book are Latin-​American maize (Zea mays
L.), amaranth (Amaranthus spp), quinoa (Chenopodium quinoa Willd), kañiwa
(Chenopodium pallidicaule Aellen), chia (Salvia hispanica L.), sacha inchi
(Plukenetia volubilis), and legumes such as black turtle (Phaseolus vulgaris) and
tarwi (Lupinus mutabilis).
These seeds are still underutilized and their cultivation is reduced but, in
recent years, worldwide demand for them has been rising to a remarkable level,
resulting in an increasing trend in their production. They have therefore become
widely recognized for their excellent nutritional and nutraceutical values by
food scientists and producers. They contain high‐quality proteins, peptides
with outstanding nutraceutical and medicinal properties, some polysaccharides
with useful functional performance, and important quantities of minerals,
vitamins, and several bioactive compounds; additionally, they are all gluten free,
which makes them suitable for people suffering from various gluten intolerances
or with a low preference for the consumption of this component. For these
reasons, interest in Latin-​American crops has increased substantially which has
resulted in intensification of research projects in different geographical regions
of the world.
This book summarizes the large number of recent investigations performed
on these seeds and provides knowledge within several of the relevant fields
of food science and agronomy, and their impact on the prevention of various
chronic diseases. Along with information on their origin and on their centers
of distribution, botanical characteristics, production and utilization, structure,
and chemical composition, a good level of attention is also directed toward their
interesting components as indicated above. This publication also includes infor-
mation about milling, fractioning, and oil extraction, and on the elaboration of

xiii
xiv Preface

various derived traditional and novel food products. The nutritional and health
implications of these Latin-​American crops are widely addressed.
We hope that this book can contribute to raising the scientific importance
of these cultivars and at the same time increasing their potential uses in high-​
quality human nutrition; the latter is fundamental to corresponding health and
strengthening the immunological system of all consumers. In view of most of
these food treasures’ role as natural vaccines, their sustainable use, as classified
by the United Nations, is a key element for the preservation of their genetic bio-
diversity, a basic tool for the present and future of food availability of world-​wide
populations.
The Editors
About the Editors

Claudia M. Haros, PhD, graduated Bachelor of Chemistry from the School of


Exact and Natural Sciences, University of Buenos Aires (UBA), Argentina, in
1990. She earned an MSc in Bromatology and Food Technology (1992); and an
MSc in Biology Analysis (1997) from UBA. She earned her PhD in Chemistry
(UBA,1999). From 1991–​2003, she worked as university professor in the Organic
Chemistry Department, Food Science and Technology Area of UBA. From 1991–​
1999 she was Research Assistant in the Cereals and Oilseeds Group, Department
of Industrial Chemistry, UBA. Later, from 2000–​2002 she worked in Spain as
a visiting professor in the Cereal Group of the Institute of Agrochemistry and
Food Technology (IATA) in Valencia. In 2003, she was a postdoc fellow at the
Department of Food Microbiology, Institute of Animal Reproduction and Food
Research (CENEXFOOD-​ EU), Polish Academy of Science, Olsztyn, Poland.
From 2003–​ 2004 she received an award for working with Prof. Sandberg
of the Department of Chemical and Biological Engineering, Life Science
Division, University of Chalmers, Gothenburg, Sweden. In 2005 she became an
Associate Researcher for the Spanish Council for Scientific Research (CSIC)
in the framework of a Ramón y Cajal Programme. Since 2008 she has been a
Senior Scientist at CSIC and continues her investigation into the Cereal Group,
Department of Food Science of IATA. Since the early stages of her career, she has
mainly been engaged in research in the cereal science and technology field. The
major theme in Dr. Haros’s research is the utilization of different strategies to
improve the nutritional and/​or functional value of cereals, their ingredients, and
products, with the emphasis on starch, minerals, and proteins. These strategies
include use of different physical, biochemical or biological treatments during
the cereal milling process; development of new cereal by-​products by including
novel ingredients; use of new starter phytase producers for regulating content
and composition of lower myo-​inositol phosphates in food with clear nutritional
and health benefits. The isolation of chemical components of cereals/​pseudo-​
cereals such as starches, proteins, fibers, or oils by different processes such as
dry milling, wet milling, and cold germ pressing, among others, is one of the
main topics of Dr. Haros’s career. Her investigations include determination of
bioaccessibility/​bioavailability of minerals, glycaemic index and nutrient inputs
according to Dietary Reference Intakes/​Adequate Intakes of nutrients. The

xv
xvi About the Editors

ultimate objective is to identify dietetic solutions and innovation to prevent


diseases and to improve consumers’ well-​being and health. She was Leader of
the International Chia-​Link Network (2015–​2021) (www.chial​ink.es/​) and from
2018 has been the Leader of la ValSe-​Food Group (Iberoamerican Valuable
Seeds-​Food, with participation of research groups/​companies from 12 countries
from Iberoamerica (www.cyted.org/​es/​val​se_​f​ood). She has been involved as
participant or leader in numerous national (Argentina, Chile, Spain, Poland, and
Sweden), European, and international research projects related to cereals. She
developed extensive activity in transferring knowledge through close cooperation
with stakeholders mainly dealing with the inclusion of cereals/​pseudo-​cereals/​
ancients crops in the human diet through their characterization and utilization
to develop innovative and sensory accepted cereal-​ based products. Today
she belongs to the International Board of the Latam Food Innovation Hub in
pursuing diffusion of her research activities in a business forum (www.lat​amfo​
odin​nova​tion​hub.com/​).

María Reguera, PhD, graduated Bachelor of Environmental Sciences and


completed her PhD in Plant Biology at the Universidad Autónoma de Madrid
(UAM, Madrid, Spain), earning the Doctoral Excellence Award in 2009. This work
contributed to the understanding of the role of boron in symbiotic nitrogen
fixation in legumes. Later, her postdocs at the Department of Plant Sciences
at the University of California Davis (UCD, US) (2010–2014) and as Juan de la
Cierva Fellow (UAM, Spain) (2015–2017) at the Department of Biology at UAM
focused on analyzing the impact of abiotic stress on crops. She also participated
as Lecturer of undergraduate courses and taught different postgraduate
courses. As Juan de la Cierva postdoctoral fellow at UAM (2015–2017) led
different research lines and was the PI of a competitively funded international
project working on quinoa studying the impact of agroecological conditions
on the nutritional properties of seeds. Currently, she holds a non-permanent
assistant professor position as Ramon y Cajal researcher at UAM leading, as PI,
a research line studying the biological mechanisms underlying plant responses
to changing environmental conditions and their impact on seed quality in
quinoa and other emerging crops while she continues her work on plant boron
nutrition. She maintains active collaboration with different international and
national research groups and with entities that belong to the private sector and
contributes actively to scientific conferences. She has organized two scientific
meetings on quinoa, participates in I+ D evaluation activities (as a referee of
High-impact factor journals and as a member of several international research
committees, including the European Commission), in outreach activities, and
has also combined her research work with intense teaching and mentoring
activity. Altogether, her scientific career gives evidence of her knowledge
About the Editors xvii

working on a diversity of crops and plant model systems, her achievements in


understanding the regulation of key agronomic traits (including grain yield or
seed quality under various abiotic stresses), and has proven results extending
basic research findings on applied agronomic research.

Dr. Norma Sammán is Chemistry Engineer, Doctor in Food Technology,


and Full Professor at the National Universities of Jujuy and Tucumán.
She has extensive experience in the chemical, nutritional, and functional
characterization of food and the development of new food products. She
gained important achievements in the academic, research, and transfer fields.
Since 1999 she has been a member of the Regional Academic Committee of the
Doctorate in Network in Food Science and Technology and has contributed
to the training of human resources through the development of national and
international courses and directing more than 20 doctoral theses. She was
Director of the Jujuy Research and Transfer Center (CIT Jujuy) UNJu-​CONICET
(2012–​2018) and President (2003–​2009, 2015–​2018) of LATINFOODS (Latin
American Food Composition Network), and since 2018 she has been the
Geographical Representative of the network for South America.
She has contributed to the capacity and standard development in food com-
position. Her research activities are currently oriented to the study of Andean
crops in Northwest Argentina with the main objective of promoting sustainable
productive activities. (i) The research group she directs relieved agricultural
producers in relation to their productive and economic–​social characteristics
to detect potentialities. Organizational proposals accepted by the communities
involved and compatible with their worldview, values, culture, and history were
made. (ii) The biodiversity of Andean crops was preserved through the reinsertion
of genetic material and characterization of the different varieties. This material
was distributed to agricultural cooperatives to encourage their cultivation and
conservation. (iii) Processed foods are developed based on Andean products,
using low-​cost technologies applicable to the region, which allow high reten-
tion of nutrients, bio-​functional compounds, and their sensory characteristics.
She has published numerous scientific articles, book chapters, and technical or
policy-​guidance documents and has received awards for her work in the field of
food and nutrition.

Octavio Paredes-​López, PhD, earned his Bachelor’s degree in Biochemical


Engineering, and a Master’s in Food Science and Technology at the National
Polytechnic Institute in Mexico City; a Master’s in Biochemical Engineering
at the Czech Academy of Sciences; and a PhD at the University of Manitoba,
Department of Plant Sciences, Winnipeg, Manitoba, Canada. He has competed
postdoctoral and research stays in several universities and institutes in countries
xviii About the Editors

such as the USA, Canada, France, UK, Germany, Switzerland, and Brazil. He has
published more than 245 scientific papers in refereed international journals,
55 reviews and book chapters, is author/​editor/​coeditor of 10 books, and 125
articles on the role of science in popular newspapers mainly in Mexico and some
in the USA and France; in total more than 435 works. His scientific publications
have been on basic and applied aspects of foods (i.e, fruits, cereals, plant foods,
molecular biology, genetically modified organisms (GMOs); on microbial
topics (i.e., genetic modifications, fermentation technology, overexpression of
secondary compounds); and on biotechnology. Students: he has graduated over
100 students doing research in his laboratory; some of them from abroad. He
has received all Mexican awards of his area of research including the highest
recognition: Mexican Award in Science. Emeritus Professor and Emeritus
National Researcher. International recognition: Institute of Food Technologists
Fellow USA; WK Kellogg International Food Security Award; Nestlé Award;
Academy of Sciences of the Developing World –​Agriculture Award, Trieste, Italy.
Ordre National du Mérite, Republique Française, Paris, France. Vice President
and President of the Mexican Academy of Sciences. Doctor Honoris causa of
three Mexican universities. Doctor Honoris causa of the University of Manitoba,
Canada, where he earned his PhD.
Contributors

Gabriel Barros Sonia Calliope


Universidad de Concepción Universidad Nacional de Jujuy
Concepción, Chile and Consejo Nacional de
Investigaciones Científicas y
Luisa Bascuñán-​Godoy Técnicas (CONICET)
Universidad de Concepción San Salvador de Jujuy, Argentina
Concepción, Chile
Marianela Capitani
Natalia Bassett Argentina–​CONICET, CCT Tandil
Universidad Nacional de Jujuy Buenos Aires, Argentina
and Consejo Nacional de
Investigaciones Científicas y Isabel Castanheira
Técnicas (CONICET) National Institute of Health Doutor
San Salvador de Jujuy, Argentina Ricardo Jorge
Lisbon, Portugal
Alain P. Bonjean
Bonjean & Associates SAS Catalina Castro
Orcines, France Universidad de Concepción
Concepción, Chile
Barbara Borczak
University of Agriculture Doris Chalampuente-​Flores
Krakow, Poland Universidad Santiago de Compostela,
Spain / Universidad
Edgardo Calandri Técnica del Norte, Ecuador
Instituto de Ciencias y Tecnología de
los Alimentos Córdoba -​ICyTAC Nancy Chasquibol Silva
(CONICET) Universidad de Lima,
Cordoba, Argentina Lima, Peru

Jehannara Calle-​Domínguez Teodoro Coba de la Peña


Instituto de Investigaciones para la Centro de Estudios Avanzados en
Industria Alimenticia (IIIA) Zonas Áridas (CEAZA)
La Habana, Cuba La Serena, Chile

xix
xx Contributors

José María Coll Marqués Vanesa Ixtaína


Instituto de Agroquímica y Universidad Nacional de La Plata
Tecnología de Alimentos (UNLP)
(IATA-​CSIC) Buenos Aires, Argentina
Paterna, Valencia, Spain
Sven-​Erik Jacobsen
José Delatorre-Herrera Quinoa Quality ApS
Universidad Arturo Prat Regstrup, Denmark
Iquique, Chile
María Dolores Jiménez
Víctor Delgado-​Soriano Universidad Nacional de Jujuy
Universidad Nacional Agraria La and Consejo Nacional de
Molina (UNALM) Investigaciones Científicas y
Lima, Peru Técnicas (CONICET)
San Salvador de Jujuy, Argentina
Néstor Fernández del Saz
Universidad de Concepción Nete Kodahl
Concepción, Chile University of Copenhagen
Frederiksberg C, Denmark
Nieves Fernández-​García
Centro de Edafología y Biologia Manuel Oscar Lobo
Aplicada del Segura -​Consejo Universidad Nacional de Jujuy and
Superior de Investigaciones Consejo Nacional de Investigaciones
Científicas (CEBAS-​CSIC). Científicas y Técnicas (CONICET)
Murcia, Spain San Salvador de Jujuy, Argentina

Susana Fischer Enrique Martínez


Universidad de Concepción Centro de Estudios Avanzados en
Concepción, Chile Zonas Áridas (CEAZA)
La Serena, Chile
María Alejandra Giménez
Universidad Nacional de Jujuy Marcela Lilian Martínez
and Consejo Nacional de Instituto Multidisciplinario
Investigaciones Científicas y de Biología Vegetal (IMBIV,
Técnicas (CONICET) CONICET-​UNC)
San Salvador de Jujuy, Argentina Córdoba, Argentina

Claudia M. Haros Silvia V. Melgarejo-​Cabello


Instituto de Agroquímica y Tecnología Universidad Nacional Agraria La
de Alimentos (IATA) Molina
Paterna, Valencia, Spain Lima, Peru
Contributors xxi

Ángela Miranda Enrique Ostria


Instituto de Ciencia y Tecnología Centro de Estudios Avanzados en
Agrícolas (ICTA) Zonas Áridas (CEAZA)
Ciudad de Guatemala, Guatemala La Serena, Chile

Luis Rodolfo Montes Osorio Octavio Paredes-​López


Universidad de San Carlos de Centro de Investigación y de Estudios
Guatemala Avanzados (CINVESTAV)
Ciudad de Guatemala, Guatemala Guanajuato, Mexico

Ana Laura Mosso M. Carmen Pérez Camino


Universidad Nacional de Jujuy Instituto de la Grasa (IG-​CSIC)
and Consejo Nacional de Seville, Spain
Investigaciones Científicas y
Técnicas (CONICET) María Reguera
San Salvador de Jujuy, Argentina Universidad Autónoma de Madrid
Madrid, Spain
Carla Motta
National Institute of Health Doutor Ritva Ann-​Mari
Ricardo Jorge Repo-​Carrasco-​Valencia
Lisbon, Portugal Universidad Nacional Agraria La Molina
Lima, Peru
Ángel Mujica
Universidad Nacional Altiplano Juan Pablo Rodríguez
Puno, Peru International Center for Biosaline
Agriculture
Loreto A. Muñoz Dubai, UAE
Universidad Central de Chile
Santiago, Chile Inmaculada Roman-​García
Centro de Edafología y Biología
Enrique Olmos Aplicada del Segura –​Consejo
Centro de Edafología y Superior de Investigaciones
Biología Aplicada del Científicas (CEBAS-​CSIC).
Segura -​Consejo Superior Murcia, Spain
de Investigaciones
Científicas (CEBAS-​CSIC). María Constanza Rossi
Murcia, Spain Universidad Nacional de Jujuy
and Consejo Nacional de
José Ortiz Investigaciones Científicas y
Universidad de Concepción Técnicas (CONICET)
Concepción, Chile San Salvador de Jujuy, Argentina
newgenprepdf

xxii Contributors

Karina B. Ruiz Marten Sørensen


Universidad Arturo Prat University of Copenhagen
Iquique, Chile Frederiksberg C, Denmark

Norma Sammán César Tapia Bastidas


Universidad Nacional de Jujuy INIAP, Estación Experimental Santa
and Consejo Nacional de Catalina, Quito, Ecuador
Investigaciones Científicas y
Técnicas (CONICET) Mabel Tomás
San Salvador de Jujuy, Argentina Universidad Nacional de La Plata
(UNLP)
Carolina Sanhueza Buenos Aires, Argentina
Universidad de Concepción
Concepción, Chile Bernabé Vázquez Agostini
Universidad Nacional de Jujuy
Adriana Scilingo and Consejo Nacional de
Universidad Nacional de La Plata Investigaciones Científicas y
(UNLP) Técnicas (CONICET)
Buenos Aires, Argentina San Salvador de Jujuy, Argentina
Chapter 1

Agronomical Characterization
of Latin-​American Crops
Sven-​Erik Jacobsen1, Angela Miranda2,
Doris Chalampuente-​Flores3, Juan Pablo Rodríguez4,
Luis Rodolfo Montes Osorio5, Alain P. Bonjean6,
Nete Kodahl7, César Tapia Bastidas8 and Marten Sørensen7
1Quinoa Quality ApS, Regstrup, Denmark.
2Instituto de Ciencia y Tecnología Agrícolas

(ICTA), Ciudad de Guatemala, Guatemala​


3Universidad Santiago de Compostela, Spain /​Universidad

Técnica del Norte, Ecuador


4International Center for Biosaline Agriculture, Dubai, UAE,
5Universidad de San Carlos de Guatemala, Ciudad de Guatemala,

Guatemala
6Bonjean & Associates SAS, Orcines, France
7University of Copenhagen, Thorvaldsensvej

Frederiksberg C, Denmark
8INIAP, Estación Experimental Santa Catalina,

Panamericana sur km 1, Quito, Ecuador

CONTENTS
1.1 The Importance of Agro-​biodiversity 2
1.2 The Andean Region of Latin America 3
1.3 Ancient Techniques for Improving Food Security 5
1.4 Limitations of Andean and Other Latin American Crops 7
1.5 Selected Crops with Massive Potential for the Future 7
1.5.1 Quinoa (Chenopodium quinoa Willd.) 7
1.5.1.1 Introduction 7
1.5.1.2 Limitations –​Breeding 9

DOI: 10.1201/9781003088424-1 1
2 Characterization of Latin-American Crops

1.5.1.3 Future Potential 9


1.5.2 Kañawa (Chenopodium pallidicaule Aellen) 10
1.5.2.1 Introduction 10
1.5.2.2 Limitations –​Breeding Needs 13
1.5.2.3 Future Potential 13
1.5.3 Andean lupine (Lupinus mutabilis Sweet) 14
1.5.3.1 Introduction 14
1.5.3.2 Limitations 17
1.5.3.3 Future Potential 18
1.5.4 Common Bean (Phaseolus vulgaris L.) 19
1.5.4.1 Introduction 19
1.5.4.2 Limitations 21
1.5.5 Chia (Salvia hispanica L.) 23
1.5.5.1 Introduction 23
1.5.5.2 Limitations 24
1.5.5.3 Future Potential 25
1.5.6 Andean maize (Zea mays ssp. mays L.) 25
1.5.6.1 Introduction 25
1.5.6.2 Limitations 27
1.5.6.3 Future Potential 27
1.5.7 Sacha Inchi (Plukenetia volubilis L.) 28
1.5.7.1 Introduction 28
1.5.7.2 Limitations and Breeding Opportunities 31
1.5.7.3 Future Potential 32
1.6 General Discussion 32
References 33

1.1 THE IMPORTANCE OF AGRO-​BIODIVERSITY


Dependence on a few crops has negative consequences for ecosystems, food
diversity, and health. Food monotony increases the risk of micronutrient defi-
ciency. A strategy based on utilizing existing plant agro-​biodiversity offers much
promise when the objective is to feed the world’s growing population within the
coming decades (Jacobsen et al. 2013). The discussion is which selection cri-
teria should be implemented to ensure the highest level of sustainability and
reliability for the biodiversity-​based strategy of crop development (Jacobsen
et al. 2015).
When considering current global challenges, a reformation of our food
systems is imperative to ensure food security, mitigate climate change, and
1.2 The Andean Region of Latin America 3

alleviate malnutrition. In this regard, underutilized crops may be essential tools


that can provide agricultural hardiness and reduced need for external inputs,
climate resilience, diet diversification, and improved income opportunities for
smallholders.
Few neglected and underutilized species (NUS) have made their way to export
markets worldwide (Akinnifesi et al. 2008). They are interesting for increasing
agro-​biodiversity, and they often have high nutritional value and can be an essen-
tial source of micronutrients, protein, energy, and fibre, contributing to food and
nutrition security. Many of these crops require relatively low inputs, and they
can be grown on marginal lands and easily intercropped or rotated with staple
crops and fit easily into agroecological practices. As frequently adapted to mar-
ginal conditions, they have the unique ability to tolerate or withstand stresses,
making production systems more sustainable and climate-​resilient (Li and
Siddique 2018).
There are examples of NUS that have become important commercial crops.
For example, quinoa and lentil are two NUS crops that attracted global interest
despite being important mainly in subsistence agriculture, either local or region-
ally (Li and Siddique 2018). The Latin-​American crops described in this chapter
represent some of the most promising cultivars for the future. Their highlights
are listed in Table 1.1.

1.2 THE ANDEAN REGION OF LATIN AMERICA


Reassessing neglected and underutilized crops for maintaining regional food
security and improving human nutrition worldwide might be an excellent
opportunity to recover forgotten crops at risk of extinction (Leidi et al. 2018;
Jacobsen et al. 2013, 2015). For example, global initiatives like the declaration of
2013 as the international year of quinoa (Bazile et al. 2016; FAO 2013) boosted
public knowledge of an important seed crop already adopted by vegetarian,
vegan, and gluten intolerant consumers, and in general, consumers interested in
sustainable food and reducing meat consumption. However, many other crops
whose valuable diversity is maintained by local producers and consumers lay far
behind in world awareness (Gahukar 2014; Hernández Bermejo and Leon 1994).
The Andean region is an important part of Latin America, and the most
fragile. The challenge of increasing Andean crop production while protecting
the fragile Andean environment and its endemic biodiversity lies mainly in
developing organic conservation agriculture, considering various techniques.
It should include the use of reduced or no tillage for increasing soil fertility
and reducing purchased inputs, maintain soil cover for minimizing erosion,
4 Characterization of Latin-American Crops

TABLE 1.1 HIGHLIGHTS OF SELECTED LATIN-​AMERICAN CROPS


Crop species –​origin Origin Highlights
Quinoa (Chenopodium Andean region High nutritional quality
quinoa) High content of micronutrients such
as iron, calcium, and zinc
High protein quality
Climate proof, tolerant to drought and
saline soils
Ideal for plant-​based food
Increasing market demand
Kañawa Andean region High level of tolerance to cold and
(Chenopodium frost
pallidicaule) High content of iron
High protein quality
Andean lupine Andean region Highest protein content among
(Lupinus mutabilis) pulses, similar to soybean
Dual-​purpose (protein and oil)
Potential for alkaloid use
Common bean S. Mexico, N Part of the traditional diet
(Phaseolus vulgaris) Central America Well-​known product with high market
demand
Maize, Andean maize S. Mexico, N Well-​known product
(Zea mays) Central America, High global demand
Andean region Many uses
Chia (Salvia S. Mexico, N High nutritional quality
hispanica) Central America Increasing world market
Sacha inchi Tropical S. High nutritive value
(Plukenetia volubilis) America Exceptional oil composition
Good sensory acceptability
Well suited for cultivation
Numerous potential applications in
gastronomy, medicine, and cosmetics
Broad adaptation
Source: Authors’ own work.

and adopt improved crop rotations, including the use of neglected and
underutilized crops.
The cold and rural highland of the Peruvian-​Bolivian-​Chilean altiplano, and
specifically the productive agricultural land close to Lake Titicaca, on the border
between Peru and Bolivia, has played an essential role in the domestication of
1.3 Ancient Techniques for Improving Food Security 5

crops of the pre-​Hispanic Andean world, as well as in situ conservation of genetic


variability.

1.3 ANCIENT TECHNIQUES FOR IMPROVING


FOOD SECURITY
Tahuantinsuyo, the Inca territory that covered the current area of Peru, Bolivia,
Chile, Ecuador, and Argentina (approx. AD 1100−1542), was a highly developed
country until the invasion and Spanish conquest in the 16th century. The
different pre-​Hispanic Andean cultures like Wari, Chimus, Colleagues, Pukara,
Tiwanaku, Lupaka, and finally the Incas, answered with technological strategies
and wisdom to the challenges of agricultural production in the altiplano, with
altitudes up to 4500 m above sea level (a.s.l.), characterized by frost, drought,
floods, hailstorms, and high solar radiation. They developed a social and eco-
nomic structure adapted to the extreme climatic and edaphic conditions, cre-
ating sound systems of managing soil and water linked to a cosmovision of the
Andean world (Paz Silva 1993).
A range of pre-​Hispanic modifications of the climate have been developed.
They were constructed and used to offset the adverse climate to enable culti-
vation of quinoa and other Andean crops. Crop production in these extreme
conditions is only possible because of these ancient structures. Therefore, crop
production at this high altitude is purely organic, utilizing animal manure, gen-
etic diversity, and diverse technologies to secure food production.
The ethnic diversity of pre-​Inca cultures is characterized by various agri-
cultural skills and knowledge invented by these cultures (Erickson 1985). The
technologies, such as terraces, high beds, artificial lakes, boundary layers, and
stonewalls, allowed them to modify the environment, especially on the altiplano
(Figure 1.1).
The Pukara developed the qochas (artificial lakes) to produce food in cold
and plane conditions with an excess of humidity. The Lupaka, situated at the
shores of Lake Titicaca, constructed waru warus (high beds) to produce food,
mainly quinoa and kañawa. These crop species are highly resistant to adverse
climatic conditions prevalent in the Andean highlands. Andenes (terrazas) were
constructed in mountain areas, cultivating Andean mountain slopes to reduce
frost risk. Andenes, waru warus, and qochas are still in use by indigenous com-
munities of the altiplano.
A unique production system developed for specific conditions in the southern
altiplano of the salt desert of Bolivia enabled quinoa to be grown under harsh
conditions, where no other crops can be grown.
The Aynoka is a collection in situ of germplasm of the crop, for instance,
quinoa, and related species, between which crossings may occur, maintaining
6 Characterization of Latin-American Crops

A B

Figure 1.1 (A) Quinoa on the altiplano of Bolivia (left, Sven-​Erik Jacobsen, right, Freddy
Chipana) (B) Waru waru (high beds) in Puno, Peru. Source: Authors’ own work.
Notes: A) The image on the left shows two men standing in a quinoa crop plantation with
a snow-topped mountain in the background.
B) The image on the right shows the traditional high beds of the altiplano with a lake
behind them.

genetic variation. Aynokas are systems of farmers’ organization of the altiplano


for multiple purposes, such as food security, rational soil management, pest
control, conservation, and use of genetic diversity, and management of crop
production in different altitudes (Ichuta and Artiaga 1986). In situ conservation
is related to both the cultivated crops as well as wild relatives. The systems are
widely distributed in the Andean region under different names such as mandas,
laymes, etc. (Mujica and Jacobsen 2000).
In situ conservation is practiced only in a few places, due to lack of techno-
logical progress, such as tractors, modern varieties for the market, and lack of
farmers’ support and stimuli, accentuated by the economic crisis and extreme
poverty of many of the rural communities who carry out these systems of culti-
vation and preservation of genetic resources. When systems are abandoned, it
may lead to losing Andean culture and identity.
The wild relatives have particular importance, often seen on fields’ borders
and in areas regarded as sacred where conservationist farmers conserve
them. In addition, the diversity of wild relatives is used for food, medicine,
and rituals, especially in times of drought and other climatic adversities in
the altiplano.
The ancient technologies are all in danger, and in many places, structures
are abandoned. It is important that institutions in charge of preserving gen-
etic resources succeed in changing this development; otherwise, they could
disappear in the near future, together with genetic diversity (Mujica and
Jacobsen 1998).
1.5 Selected Crops with Massive Potential for the Future 7

1.4 LIMITATIONS OF ANDEAN AND OTHER LATIN


AMERICAN CROPS
Since colonial times, there has been a lack of native Andean crops due to devalu-
ation and rejection of crops, and some were even prohibited from cultivation.
It led to a reduction in their consumption, later exacerbated by considering it
“Indian food” or “poor man’s food” (Horton 2014). The challenge is recognizing
that Andean grain and pulse crops contribute to food sovereignty (Mercado
et al. 2018).
Selection and breeding efforts have been sparse of Andean crops in general.
For this reason, they have remained at a marginal level with regard to produc-
tion, market, and consumption. They have often served as essential subsistence
crops but have had difficulties reaching domestic and international markets to
create income for the rural population, agro-​industries, and the country.
Borlaug and his colleagues (Borlaug 2000) initiated the first green revolu-
tion, successfully increasing yields under rising use of fertilisers and pesticides.
However, this agronomical revolution created several environmental problems,
which should now be corrected, focusing on sustainable production, soil fer-
tility, and increasing agro-​biodiversity, using the vast range of underutilized
crops with defined potential.
A sustainable green revolution no. 2 should utilize a wider diversity of
crops so that food production can benefit from a broader set of species, each
adapted for specific marginal conditions. Breeding techniques may be classical
or molecular, including CRISPR-​Cas-​based systems (i.e. Clustered Regularly
Interspaced Short Palindromic Repeats), and TILLING (Targeting Induced Local
Lesions IN Genomes), and eventually, Genetically Modified Organisms (GMO)
(López-​Marqués et al. 2020).

1.5 SELECTED CROPS WITH MASSIVE POTENTIAL


FOR THE FUTURE
The selected crops are different types of crops with different virtues. They are all
highly valuable crops, potentially contributing to an improved agro-​biodiversity
worldwide and improved food diversity and food security.

1.5.1 Quinoa (Chenopodium quinoa Willd.)


1.5.1.1 Introduction
One of the oldest existing crops
It is not known when quinoa was brought to Europe, so the crop literally
remained unknown outside Andean countries until the late 1970s, first
8 Characterization of Latin-American Crops

introduced to the USA. Later studies of adaptation were initiated in England,


Denmark, and the Netherlands, followed by a further spread with trials in Europe
and areas with a temperate to subtropical climate (Jacobsen 2017).
Until recently, no commercial production of quinoa was found outside the
Andes. Nowadays, quinoa is cultivated, or at least tested, in many countries
(>120) and the potential for further expansion of global production is high
(Alandia et al. 2020). However, there is still little scientific information avail-
able regarding the factors limiting quinoa cultivation. Sellami et al. (2021)
reviewed scientific papers dealing with quinoa from 2000 to 2020, finding only
33 publications. The main topics addressed were variety, deficit irrigation,
water quality, fertilization, and sowing date, with very few focusing on tillage
and weed control.

Origin of quinoa
Historically, quinoa has been continuously selected for new environments in
the Andean region, as it was spread gradually from its centre of origin around
the Titicaca Lake between Peru and Bolivia. Distribution from the lake went
both north, to Ecuador, Colombia, and Venezuela, and south, to Chile and
Argentina, east, from the highlands down to valleys, and west, to coastal regions
of the Andean countries. The process, however, was slow due to environmental

Figure 1.2 (A) Quinoa (Vikinga cv.) in the field (B) Titicaca plants showing the panicles
(end stages of seed development). Source: Authors’ own work.
Notes: A) The image on the left shows an extensive quinoa field with trees and houses at
the back of the field.
B) The image on the right shows multiple quinoa panicles in which seeds are developed.
1.5 Selected Crops with Massive Potential for the Future 9

variability and the unstable climatic conditions in the Andean region (Bertero
et al. 2004; FAO 2013).
Lately, tropical species of quinoa have been adapted to temperate regions of
the USA and Europe, and even in northern Europe it is now possible to grow
quinoa successfully (Figure 1.2).

1.5.1.1.1 Chenopodium species
The Chenopodium genus includes around 250 species from all over the world. It
is considered one of the most nutritious genera due to its protein and dietary
fibre content and healthy fat, ash, and minerals (Repo-​Carrasco et al. 2003).
The oldest domesticated Chenopodium species, identified to date, is the South
American quinoa developed in the Andes mountains c.7500 years ago (Pearsall
1992). It reached North America c.AD 1200.
In Northern Europe, Chenopodium album L. with common names fat-​hen
and lamb’s quarters, which is a global weed species, was a secondary crop in
Denmark during the Iron Age (1200 BC–​AD 400) (Stokes and Rowley-​Conwy
2002), where both the seeds and the leaves were consumed. Chenopodium album
was also used as a pasture for milking cows in Denmark under World War II
(1940−1945), as farmers found that it secured good milk production in a situ-
ation where there was a lack of quality foodstuffs.

1.5.1.2 Limitations –​ Breeding
To our knowledge, modern molecular tools such as GMO and CRISP/​Cas9 are not
used in quinoa. Much research work has been performed using these molecular
tools in many different crops, however, they have hardly been accepted by con-
sumers in Europe so far, and, in addition, they are subjected to strict regula-
tion by governments (López-​Marqués et al. 2020). Our suggestion is to combine
molecular and genetic techniques, such as TILLING, with traditional breeding
together with other molecular techniques, not considered GMOs, in order to
breed improved quinoa cultivars that can be commercialized without further
regulation limitation or consumption acceptance problems.

1.5.1.3 Future Potential
Quinoa, a new world staple
Quinoa is a new crop, presently being tested in various parts of the world.
According to the Food and Agriculture Organization (FAO), quinoa is regarded
as a new world staple and predicted to spread fast across the globe (FAO 2013).
New study sites and commercial productions, apart from South and North
America and Europe, are seen in central Asia, i.e. the Aral Sea basin (Khaitov
et al. 2020) and Africa, i.e. Zimbabwe (Muziri et al. 2020).
10 Characterization of Latin-American Crops

In order to adapt the tropical crop species quinoa to European conditions,


day-​length neutral varieties have been developed. These varieties have been
tested and grown successfully in various countries and climate zones worldwide.
Quinoa is being established in various countries globally (Bazile et al. 2016), such
as China with 12,000 ha, being the third-​largest producer of quinoa in the world
(Xiu-​shi et al. 2019). Quinoa was tested with success in several harsh environ-
ments, with results mainly from 2−4 t ha-​1 in Iran, Turkey, and UAE (Razzaghi et al.
2020). Good results were seen also in Africa, e.g. Morocco (Nanduri et al. 2019),
Burkina Faso (Alvar-​Beltran et al. 2019), and Egypt (Adel 2020). In Asia aside from
in China, quinoa has also been tested in Bhutan (Katwal and Bazile 2020).

1.5.1.3.1 Global demand
Due to the increasing global demand for quinoa, both as an Andean export com-
modity and for agricultural development purposes, there is enormous interest
in testing quinoa for growing under various environmental and geographical
conditions.
Quinoa is mainly grown in the Andean countries Bolivia and Peru, with an
area of 112,000 ha in Bolivia and 65,000 ha in Peru. The estimated area in Europe
is 8000 ha, with the largest producers being Spain and France, with produc-
tion also in the rest of Europe from Sweden in the north to Italy in the south.
China has introduced quinoa and developed different varieties, which are grown
around the country on 13,000 ha. In Africa there is mainly small-​scale production
in several countries, such as Morocco, Egypt, Zambia, and Kenya. Total global
quinoa production is 200,000 ha, of which around 88% is produced in the two
cited Andean countries (Jacobsen, personal communication, December 2022).
As the market widens, there is great potential for production of quinoa also
outside the Andes. The estimated potential area in Europe is 2 million ha. For
example, yield levels in Northern Europe are normally 1−2 t ha,1 but with poten-
tially higher yields as demonstrated in plot trials (Jacobsen 1997, 2017; Jacobsen
and Christiansen 2016; Jacobsen et al. 1994, 2010).

1.5.2 Kañawa (Chenopodium pallidicaule Aellen)


1.5.2.1 Introduction
Cañahua, cañihua, kañawa, or kañiwa (C. pallidicaule) is an Andean species
belonging to the Amaranthaceae family. Kañawa, as well as quinoa, is a
gluten-​free achene type grain (Gorinstein et al. 2015). Saponin content is low
in Cañahua. Protein content is 15–​18%, with high protein quality like quinoa
(Moscoso-​ Mujica et al. 2017; Rodriguez et al. 2020). C. pallidicaule plants
produce several dry, single-​seeded fruits from small inflorescences distributed in
many branches (Mamani 2016a). Before harvest, seed losses are 15–​35% due to
1.5 Selected Crops with Massive Potential for the Future 11

Figure 1.3 Kañawa plants in the field. Source: Authors’ own work.
Note: The image shows different flowered kañawa plant varieties, planted in rows showing
the color diversity (purple, green, and yellow).

shattering occurring between flowering and physiological maturity (Rodriguez


et al. 2017). Diminutive flowers are distributed in lateral inflorescence branches
that are a feature of both cultivated and wild Amaranthaceae species (Spehar and
Santos 2002). In Bolivia and Peru, three growth habit forms are known, ‘Saihua
(Chuqhu)’, ‘Lasta (Thasa)’, and ‘Pampalasta (mamakañawa or illamanku)’ (see
Table 1) (Quispe 2004; Mamani 2016a). The cultivated Saihua and Lasta kañawa
growth habit forms produce a fair amount of grains and are widespread in the
highland regions (Maydana 2010; Mamani 2016b; Rodriguez et al. 2017, 2020).
The third growth habit form, Pampalasta, is a wild ecotype (Estaña and Muñoz
2012; Serrano 2012; Mamani 2016a). This wild kañawa is known for its early
ripening grains (Mamani 2016). The glycaemic index is low (Mangelson et al.
2019). A problem is the irregular grain maturation on the branches, which leads
to seed shattering, indicating its status as partially domesticated (Rodriguez
et al. 2017, 2020) (Figure 1.3).

1.5.2.1.1 Domestication/​cultivation history, including present


cultivation
Kañawa and quinoa cultivation are related, and both belong to the genus
Chenopodium with c.150 species. Kañawa originated in the Andean highlands
12 Characterization of Latin-American Crops

of Peru and Bolivia. The Tiwanaku culture accomplished its semi-​domestication


in Bolivia’s Collao highland. For millennia, chenopod species were cultivated as
food crops in North, Central, and South America. For example, the pit goosefoot
(C. berlandieri Moq. subsp. jonesianum Bruce Smith) has been being cultivated
by the native population in the eastern part of North America since before the
domestication of maize (Zea mays L.) (Bruno 2006, 2014; Fritz et al. 2017).
Kañawa is reported as having been cultivated for 7000 years as a staple crop
in pre-​Incan societies (Gade 1970). Communities around the Titicaca Lake basin
between Peru and Bolivia are the origin of kañawa as observed from collections
of landraces (Estaña and Muñoz 2012; Serrano 2012; Mamani 2016b). Bolivia
holds an extensive collection of kañawa accessions with 801 entries and Peru
with 341 (Mamani 2013, 2016a; Mamani et al. 2018; Mangelson et al. 2019).
Quinoa is well known and scaled-​up worldwide (Alandia et al. 2020), while
kañawa is only traded in small quantities from Peru and Bolivia (Rodriguez
et al. 2020). In Bolivia, kañawa is grown in the high-​altitude regions or alti-
plano ecozone of departments of La Paz, Oruro, Cochabamba, Potosi, Tarija;
and in Peru throughout the Sierra in Cusco, Puno, Ayacucho, Junín, and Huaraz
(Bravo et al. 2010; Rojas et al. 2010; Mamani et al. 2018). Peru is the first pro-
ducer. However, breeding is limited, and the few available varieties for scaled-​up
production derive from selection efforts (Rodriguez and Mamani 2016; Mamani
2016b).
Export of kañawa remains marginal compared to quinoa (Carrasco and Soto
2010; Rodriguez et al. 2010; Rodriguez et al. 2020). Producers of this crop have
consolidated into clusters, around La Paz, Oruro, and Cochabamba in Bolivia,
supplying urban markets (Mamani 2016b; Rodriguez et al. 2020; INEI 2020). In
Peru, producers in Puno, Cusco, and Juliaca commercialize kañawa (Giuliani
et al. 2012; INEI 2020).
Until 2019, Peru had higher yields than Bolivia, but in general, it is low, c.500 kg
ha.1 Cultivation of kañawa is still practiced with traditional methods (Rodriguez
et al. 2020). Barley (Hordeum vulgare L.) and oats (Avena sativa L.) are grown in
preference to kañawa because they are attractive as alternative feeds for dairy
cattle (Mamani 2016b), albeit that pilot research demonstrated that kañawa
does constitute a good quality alternative feed for minor livestock (Calle 2004;
Cortez Quispe 2016).

1.5.2.1.2 Geographical distribution
Currently, Bolivia and Peru are the main producers and exporters of kañawa
(Gade 1970; Estaña and Muñoz 2012; Rodriguez et al. 2020). Kañawa is not
cultivated outside the high-​altitude regions of Bolivia (north of Oruro and
near Cochabamba) and Peru (Ayaviri, Puno, Cusco, Ayacucho, and Junin).
Geographically, the distribution of kañawa is not as extensive as for quinoa
(Estaña and Muñoz 2012; Mamani 2016b). Kañawa collections are small and
1.5 Selected Crops with Massive Potential for the Future 13

maintained in germplasm banks (Galluzzi and Lopez Noriega 2014: Bonifacio


2019). Kañawa is a partially domesticated Chenopodium species, which provides
nutritional food security to many Andean farmers’ households in the highland
region (3500–​4200 m a.s.l.) of both countries. In the Andean region kañawa is
grown by subsistence farmers due to its tolerance to frost, drought, and salinity
(Rodriguez et al. 2017, 2020).

1.5.2.1.3 Uses
In rural households, farmers toast the grains, and they are then milled in a flat
stone mill. The flour is known in Peru as ‘cañihuacu’ and as ‘pito’ in Bolivia. Its
local popularity as grain is due to its high nutritional profile (Repo-​Carrasco-​
Valencia 2011). In addition, kañawa flour demonstrated a high capacity to mix
with cereal flours improving the dough quality for bakery products (Pisfil 2017;
Zegarra 2018; Chambi 2019).

1.5.2.2 Limitations –​Breeding Needs


Genetic erosion is reported for some traditional varieties in Bolivia (Serrano
2012; Mamani 2016a). As a result, farmers have replaced kañawa with forages
to feed dairy cattle. Furthermore, kañawa has received little or no com-
mercial breeding attention resulting in registered cutivars, and only native
landraces exist.
Using Cobalto-​60 drums and Gamma radiation, the induced mutation was
used in seeds to provoke changes in plant morphology and, if successful, reduce
seed shattering. Therefore, this needs to be studied and tested with a broad
range of genotypes.
Genetic improvement of kañawa should be addressed by obtaining var-
ieties according to current climate change while maintaining product quality
requirements for local and export customers.

1.5.2.3 Future Potential
Recently the genome of kañawa was sequenced (Mangelson et al. 2019),
meaning that it can be used as a genetic model to understand the evolution of
Chenopodium spp., particularly the cultivated C. quinoa. In addition, develop-
ment of the genome sequence can help and support the improvement of kañawa
through breeding.
The grain has excellent potential as a nutritive food source with a low-​
glycaemic index and high in iron and zinc. Studies in Peru and Bolivia have
demonstrated that flour mixed with wheat or amaranth can be transformed
into nutritive food products such as cookies (Pisfil 2017; Zegarra 2018; Chambi
2019; Coila 2019; Šver et al. 2019). Kañawa has been introduced as an integrated
element in the Complementary School Food Program, but its availability needs
to be scaled up and supported to improve nutritional security at the rural and
14 Characterization of Latin-American Crops

urban household levels. Kañawa may also be used for biodegradable films or
packages (Ramirez 2016; Salas 2017; Yanapa Velasquez 2018).
Soil salinity is expanding in many parts of the world, especially in high-​
altitude regions, and kañawa has demonstrated that it can tolerate high soil
salinity (Rodriguez et al. 2020). Therefore, the crop should be promoted for intro-
duction as an alternative for alleviating soil salinity.
Further development of the crop ought preferably to be accomplished by util-
izing native varieties, where, for instance, early maturity is present.
Kañawa can now be found in non-​domestic markets, e.g., in the USA, where
it is marketed as baby quinoa. Use of such a denomination is not, however,
recommended, as kañawa should instead be marketed on its own virtues.

1.5.3 Andean lupine (Lupinus mutabilis Sweet)


1.5.3.1 Introduction
The history of this species as an Andean subsistence crop demonstrates its
potential as a crop for low-​input agriculture in temperate climates (Cowling
et al. 1998). The selection activities of Andean farmers have represented the only
means of domestication, giving rise to semi-​domesticated forms characterized
by indehiscent pods, large seeds, multi-​colored flowers, and highly branched
architecture (Clements et al. 2008).
It is a robust crop that can be grown in poor soils and dry climates (Williams
1993), and which stands out for its great potential in soil recovery (De Ron et al.
2017). Furthermore, in addition to its high alkaloid content (4.5 g 100g-​1 DW),
the crop has high resistance to microbial infections and insect attacks (Gueguen
and Cerletti 1994; Ciesiolka et al. 2005). However, the presence of alkaloids in the
seeds and the relatively low yields (800−1300 kg ha-​1) have strongly limited the
expansion of the crop (Tapia 2015; Galek et al. 2017). Therefore, efforts have been
made to re-​establish lupine as a crop in South America and adapt it to Europe’s
conditions (Caligari et al. 2000).
The Andean lupine is characterized by the highest grain quality of all
cultivated lupines, presenting an oil content similar to that of soybeans (Glycine
max (L.) Merr.) (Gulisano et al. 2019). Numerous studies investigating the nutri-
tional profile and potential applications of this pulse have found a wide range of
possible products ranging from proteins, oils, and food additives to cosmetics,
medicines, and bio-​pesticides. In addition, the nutritional advantages make it
an ideal product for the transition from meat-​intensive diets to diets based on
vegetable proteins (Gulisano et al. 2019).

1.5.3.1.1 Origin, Diversification, and Domestication


The oldest evidence of cultivated Andean lupine is related to the seeds found
in tombs of the Nazca culture and representations in Tiahuanaco ceramics in
1.5 Selected Crops with Massive Potential for the Future 15

Peru (FAO 1992; Tapia and Fries 2007). The oldest archaeological evidence of
domesticated L. mutabilis seeds has been found in the Mantaro Valley in cen-
tral Peru and dates to about AD 200. The use of Restriction site Associated DNA
Sequencing (RAD-​Seq) in the analysis of this archaeological material confirms
that L. mutabilis was first domesticated in the Cajamarca region (northern Peru)
from the wild progenitor L. piurensis C.P.Sm. Demographic analysis suggests
that L. mutabilis separated from its parent around 2600 BC and suffered a bottle-
neck in domestication, with subsequent rapid population expansion as it was
cultivated in the Andes (Atchison et al. 2016). L. mutabilis is reported in eastern
South America, from Colombia to northern Argentina, and with a wide altitud-
inal range from 1500 to 3800 m a.s.l. (Jacobsen and Mujica 2008).
In the Andean region, 83 species have been identified; the wild relatives that
show diversity and variability found in the Andean lupine are the following
species: Lupinus ananeanus Ulbr., L. aridulus C.P.Sm., L. ballianaus C.P.Sm.,
L. chlorolepis C.P.Sm., L. condensiflorus C.P.Sm., L. cuzcensis C.P.Sm., L. dorae
C.P.Sm., L. eriocladus Ulbr., L. gibertianus C.P.Sm., L. macbrideianus C.P.Sm.,
L. microphyllus Desr., L. paniculatus Desr., L. sufferuginous Rusby, L. tarapacencis
C.P.Sm. and L. tomentosus DC. (Jacobsen and Mujica 2006).

Figure 1.4 Andean lupine (A) plant and (B) seeds and pods. Source: Authors’ own
work.
Notes: A) The image on the left shows a long stem lupine flower with small petals colored
in blue, white, and yellow, and palmately lobed leaves within a lupin field. In the back-
ground, some trees are visible.
B) The image on the right shows many lupine seeds on graph paper which, together with
the rulers in the image, indicate size of approximately 1 cm. These seeds are labeled as
DCH-​094. Also, two lupine dark-​colored pods appear below the seeds and a bunch of pods
(dark color) to the right of the seeds.
16 Characterization of Latin-American Crops

1.5.3.1.2 Botanical description
Three geographically separated morpho-​types of the Andean lupine have been
suggested based on the considerable genetic and morphological variability and
wide ecological adaptation in the Andean zone: (a) L. mutabilis, lupino (northern
Peru and Ecuador), of more prolific branching, very late, greater hairiness of leaves
and stems, some ecotypes behave as biennials, tolerant to anthracnose; (b) the
Andean lupine, tarwi (central and southern Peru), scarcely branched, moderately
late, somewhat tolerant to anthracnose; and (c) the Andean lupine, Tauri (high-
lands of Peru and Bolivia), smaller (1–​1.40 m) with a developed main stem, very
early, susceptible to anthracnose (Tapia et al. 1980; Gross 1982; Tapia 2015).
High diversity seed characteristics include shape (lenticulate to spherical),
primary and secondary seed (seed coat, i.e., testa) color, as well as distribution
patterns, can range from pearly white to solid black and includes beige, yellow,
brown, dark brown, and those in between such as brownish green and greyish
green. Most of the seeds have a secondary color distribution in darker shades of
the primary one; the secondary color distribution also varies among a wide range
of patterns, such as brown-​shaped, crescent, mottled, or spotted, which can be
expressed either as solitary or in combination (Falconí 2012; Tapia 2015) colors.
The presence of considerable variation in germplasm is shown by different
phenotypic traits, such as a wide range of growth periods, branching patterns,
color and shape of grains and flowers, as well as flowering times. Both the inter-​
simple sequence repeat (ISSR) and single-​sequenced repeat (SSR) markers have
revealed broad genetic diversity among L. mutabilis lines (Galek et al. 2007;
Chirinos-​Arias et al. 2015), which could be related to the mixed pollination
system that the species has, and which could explain the presence of the testa
color (Chirinos-​Arias et al. 2015).

1.5.3.1.3 Environments where lupine is grown


The requirement for lupine is variable, depending on soil, temperature, and
wind. It grows well in temperatures from 20 to 25ºC; grain development is
optimal below 9.5ºC (night temperature), a condition that occurs in the high
Andean region (Gross 1982); the early ecotypes of Puno-​Peru need 450 mm of
precipitation per crop cycle, while the late ecotypes require 600−700 mm (Tapia
2015). Lupine prefers sandy loam soils, with a thick, deep texture, a balance of
nutrients, and good drainage with a pH of 5−7 (Meneses 1996; FAO 2000; Jacobsen
and Mujica 2006). The seedling stage is susceptible to frost (-​4°C); the higher the
temperature, the greater the growth and development; at less than 0ºC, develop-
ment and evapotranspiration are inhibited (Tapia and Fries 2007).

1.5.3.1.4 Uses
The Andean lupine can be consumed directly as a snack (Villacrés et al. 2003)
and as an ingredient in different products such as fresh salads, soups, cakes,
1.5 Selected Crops with Massive Potential for the Future 17

cookies, bread, hamburgers, and baby food (Cremer 1983; Ruales et al. 1988;
Villacrés et al. 2003; Güémes-​Vera et al. 2008). New uses of lupine are related to
the extraction of oil and production of vegetable milk, yoghurt, and obtaining
flours and by-​products for animal feed (Tapia 1982; Peralta and Villacrés 2015).
Alkaloids such as lupunine and sparteine present in the leaves, stem and
seeds of lupine plants were traditionally used, in combination with paico
(Dysphania ambrosioides (L.) Mosyakin & Clemants, syn. Chenopodium
ambrosioides L., a wild relative of quinoa) to repel pests on potato crops such as
the Andean weevil (Premnotrypes spp.), and on quinoa the Kona Kona (Eurysacca
quinoae P.), the primary pest of quinoa. In livestock, it is used to control internal
and external parasites, practices that are disappearing due to the promotion of
industrial agrochemicals (Canahua and Román 2016).
Lupine seeds are used for consumption after debittering (Cremer 1983;
Villacres et al. 2000), reducing the alkaloid content to 0.02% for people and 0.4 to
0.6% for pigs, ruminants, and poultry (Tapia and Fries 2007; Caicedo and Peralta
2000). To eliminate antinutritive substances (alkaloids), a hydro-​thermal process
was carried out, which consists of hydrating the dry grain and soaking it for 12 to
14 h, after which it is cooked for 30 to 40 min. Finally, the seed is left in a stream
of continuous drinking water for three or four days, or in circulating water from
the river or streams between seven to ten days (Caicedo et al. 2001).

1.5.3.2 Limitations
1.5.3.2.1 From an agronomic point of view
Limitations are the lack of early maturity and high-​yielding genotypes, lack of
locally adapted genotypes (Caligari et al. 2000), and lack of good quality seed
(Mercado et al. 2018).
There has been limited plant breeding taking place to increase yield and
seed quality and decrease susceptibility to diseases such as anthracnose
(Colletotrichum gloeosporioides). Lack of advanced biotechnological methods
in genetics, molecular cytogenetics, and tissue culture have limited the pos-
sibility of exploiting natural variability and performing distant crosses and
haploidization of material from reproduction (Gulisano et al. 2019).
Domestication and history of reproduction are fragmented in time and space,
and participatory approaches with farmers to select local ecotypes are missing.

1.5.3.2.2 From a nutritional point of view


The Andean lupine has a bitter taste due to the high content of alkaloids, limiting
direct consumption for both humans and animals (Guerrero 1987). Hence, it is
necessary to improve the debittering processes (Mercado et al. 2018).
Specific data about the consumption of the Andean lupine or tarwi are
needed, which will help make decisions and take actions that allow the promo-
tion of greater consumption (Mercado et al. 2018).
18 Characterization of Latin-American Crops

1.5.3.3 Future Potential
The Andean lupine, known locally as ‘tarwi’ or ‘chocho’ (L. mutabilis Sweet), has
been cultivated, processed, and consumed for at least 1500 years, and whose
genetic variability has adapted to many microclimates (Williams 1993). Even
before the Spanish conquest, this crop played an essential role in high Andean
production systems and fed the indigenous population (FAO 2016). Among
legumes, lupine is characterized by its high protein content of good quality,
suitability for environmentally robust production, and potential health benefits
(Lucas et al. 2015).
In the Andean region, the annual per capita consumption varies. In Ecuador,
it is 4−8 kg person-​1, much higher than in Bolivia (0.2 kg person-​1) and Peru
(0.5 kg person-​1). In 2017, production did not meet domestic demand in Ecuador,
reporting a deficit of around 6000 tons (Mazón 2018).
The gastronomic versatility and nutritional qualities of this legume crop,
combined with the work carried out for more than 20 years by both public and
private entities in technological innovation, post-​harvest, added value, quality
seed, and improved varieties, have renewed interest in this crop (Horton 2014;
Nájera 2015; Márquez 2016).
The combined use of germplasm and modern approaches to broaden the gen-
etic base could help the introgression of desirable adaptive traits for environ-
ments adapted to certain latitudes (Gulisano et al. 2019). Future work should
aim to develop bitter lines, with a sufficient level of alkaloids in vegetative tissues
to decrease the presence of pathogens (Wink 1991); and sweet lines, which will
be easier to process for consumption. Also, integrated pest and disease man-
agement programs are required to improve farmers’ production systems (Mina
et al. 2018).
Converting production into a dual-​purpose alternative (protein and oil)
similar to soybean could be an economical alternative for the productive and
competitive development of the Andean region (Lucas et al. 2015).
Andean lupine alkaloids could be of commercial importance due to their
pharmacological activity (Ciesiolka et al. 2005; Jiménez-​Martínez et al. 2003).
Furthermore, specific protein isolates and concentrates could be of commer-
cial importance due to their functional properties for the chemical and food
industry (Sathe et al. 1981; Gueguen and Cerletti 1994; Doxastakis 2000; Moure
et al., 2006). In addition, the presence of ferritin (a protein rich in Fe) in the
protein profile (Strozycki et al. 2007) increases the nutritional value of this cul-
ture by offering a safe way to increase the intake of iron in the diet (Zielinska-​
Dawidziak 2015).
Quinolizidine alkaloids (QAs) also have an essential role in the medical field
due to multiple properties such as antiarrhythmic, anti-​inflammatory, diuretic,
and hypotensive effects, among others (Bunsupa et al. 2012). Besides, QAs can
1.5 Selected Crops with Massive Potential for the Future 19

also find application in agriculture as a bio-​stimulant increasing the growth


and yield of other crops (Przybylak et al. 2005) as antibacterial agents (Romeo
et al. 2018) or as biocidal agents that replace synthetic toxins (Bermúdez-​Torres
et al. 2009).

1.5.4 Common Bean (Phaseolus vulgaris L.)


1.5.4.1 Introduction
The traditional production system of common beans is in monoculture with
bush-​type varieties. However, the traditional system known as milpa, which
consists of an intercropping arrangement (maize–​bean) and other vegetables
and tubers, is prevalent in the Highlands but not in all regions of Latin America;
irrigation is generally not used, and the crop depends exclusively on rainfall. The
dry bean seed constitutes one of the primary protein sources for human con-
sumption, providing essential micronutrients such as iron and zinc (Rodríguez-​
Castillo and Fernández-​Rojas 2003; MAGA 2017).

1.5.4.1.1 Origin, diversification, and domestication


The Mesoamerican region, and especially Mexico, is considered the centre
of origin and diversification of beans (Phaseolus spp.), cultivated and wild
(Hernández-​López et al. 2013; Bitocchi et al. 2012, 2017). Common beans are cur-
rently grown in various regions and conditions, from 0–​3000 m a.s.l., mainly by
small producers in Central America, Africa, and Asia, representing 77% of world
production. Central America, and Guatemala and Nicaragua are the largest pro-
ducers (550 and 390 million kg per year, respectively) (Secretaría de Economía
2012; OECD 2021).
Latin America is an important centre of diversity for the common bean (Singh
et al. 1991; Beebe et al. 2000; Blair et al. 2013). This diversity should be studied
for the generation of improved cultivars with better seed yield potential. The
common bean is part of Latin-​American people’s main diet, mainly because it is
an essential protein source (Broughton et al. 2003). The improvement of nutri-
tional aspects through breeding, such as bio-​fortification with minerals (iron
and zinc), can create a positive impact on food security (Espinoza-​Moreno et al.
2016; Morais Dias et al. 2018; Orona-​Tamayo et al. 2018; Contin Gomes et al.
2021; OECD 2021).

1.5.4.1.2 Germplasm collections
The National Plant Germplasm System of the United States of America
(NPGS) identifies 81 dry bean species accepted in their online Germplasm
Resources Information Network (GRIN). This number increases to 117 taxa
when including subspecies and varieties. A comprehensive analysis of North
20 Characterization of Latin-American Crops

and Central American species recognizes no less than 36 species, many with
one or more subspecies, five of them of economic importance (FAO 2018). The
Mesoamerican region is considered the main centre of origin and genetic diver-
sity of domesticated and wild species of the genus Phaseolus (Beyra et al. 2004;
Kwak and Gepts 2009; Rossi et al. 2009; Mamidi et al. 2011; Bitocchi et al. 2012).
The germplasm bank at Centro Internacional de Agricultura Tropical (CIAT),
located in Cali, Colombia, one of the international germplasm banks of the
Consultative Group on International Agricultural Research (CGIAR), operates
within the International Treaty Framework on Plant Genetic Resources
for Food and Agriculture (TRFGAA). The TRFGAA is primarily funded by
the Platform of Germplasm Banks coordinated by Global Crop Trust. The
germplasm bank at CIAT maintains the most important collections worldwide
of essential crops that provide carbohydrates and protein in the tropical food
systems: Beans (Phaseolus species, 37,938 accessions), manioc, cassava, tapioca
(Manihot species, 6155 accessions), and forage crops (22,694 accessions). The
CIAT germplasm bank also stores duplicates (security copies) of seeds of world
crops in collaboration with the Nordic Genebank, who are using their under-
ground facilities on the arctic island of Svalbard, midway between Norway and
the North Pole, and in the International Maize and Wheat Improvement Center
(CIMMYT) germplasm bank at Texcoco, Mexico (CIAT 2020). Moreover, the
Institute of Agricultural Science and Technology (ICTA) has national germplasm
collections from different pulse crop species. Additionally, Guatemalan beans
are also in the major world collection, stored in the germplasm bank at CIAT
(FAO 2018).

1.5.4.1.3 Botanical description
Botanically, the genus Phaseolus –​including the common bean species
(P. vulgaris L.) –​belongs to the Leguminosae (Fabaceae) family, Faboideae
(Papilionoideae) subfamily, Phaseoleae tribe and Phaseolinae subtribe. It is an
annual herbaceous plant, and depending on growth habit, it can reach heights
up to 2 m (Fernández et al. 1984).
Common beans can present four growth habits: type I determined (bush
type), type II indeterminate (bush type), type III indeterminate (prostrate), and
type IV indeterminate (climber). Those of determined growth can reach heights
between 30 cm and 90 cm, while the plants with indeterminate habit reach
heights from 50 cm to 3 m (Rosas 2003).
The bean has a primary root and many secondary roots with nodules
developed from an association with the nitrogen-​fixing bacterium Rhizobium. The
leaves are trifoliate. The flowers have a tubular calyx of five sepals, a papilionoid
corolla of unequally sized petals, ten stamens, and a receptive stigma. The flower
color can be white, lilac, purple, or bicolored. The fruits are legumes, also called
1.5 Selected Crops with Massive Potential for the Future 21

pods, and the seeds have two cotyledons. The seeds inside the pods are rich in
protein (Clavijo 1980).

1.5.4.1.4 Production and geographical distribution


Common beans are produced in diverse systems, regions, and environments,
e.g., Latin America, Africa, the Middle East, China, Europe, USA, and Canada. In
Latin America, it is an essential and in the daily-​diet food, especially in Brazil,
Mexico, Central America, and the Caribbean.
In Latin America, bean producers and smallholder farmers, on average, have
2 ha of land to produce this pulse (Fischer and Victor 2014). Common bean pro-
duction in developing countries is considered low input and small-​scale agricul-
ture (Miklas et al. 2006).
Estimated data from FAOSTAT (2021) report an average yield for Mexico,
Guatemala, Honduras, and El Salvador of 862.58 kg ha-​1. Guatemala has the
highest yield with 999.1 kg ha-​1, and Mexico has the lowest yield with 728.3 kg
ha-​1. Despite its importance in some countries’ diets, on the world stage, the
volume of bean production compared to other grains such as maize, wheat, and
rice represents only 1%.

1.5.4.2 Limitations
1.5.4.2.1 Agronomy
The highest production area (8 mill. ha) is found in the Latin-​American region
(CGIAR-​Beanfocus 2019). However, bean production has not expanded along
with increasing demands from a growing population.
The average yield is relatively low, ranging from 600–​1000 kg ha-​1 (Rathna
Priya and Manickavasagan 2020). Yield is affected by heat and drought
(Villarino 2018). Besides, diseases and pests may cause up to complete
loss. Most varieties will take c.65–​110 days to mature, but some landraces
and cultivars may take more than 200 days according to their growth habit
(Katungi et al. 2009).

1.5.4.2.2 Nutrition
The common bean is a good source of protein, but beans also present anti-​
nutritional factors that negatively affect human and animal metabolism; fortu-
nately, these factors are thermolabile and their content may be partially or
totally destroyed during cooking (Cravioto et al. 1945; Bressani 1990; Martín-​
Cabrejas et al. 2004). In addition, following consumption, the flatulence factor
creates rejection of this pulse (Geil and Anderson 1994; Haileslassie et al. 2016).
However, nutrition properties for this pulse can be improved by using different
processing methods, including soaking, de-​hulling, germination, fermentation,
22 Characterization of Latin-American Crops

Figure 1.5 Bean (A) plant and (B) seeds. Source: Authors’ own work.
Notes: A) The image on the left shows a small bean plant with a deltoid shape and with
small white flowers in the center of the plant,which seems to be in a bean field.
B) The image on the right shows oval, bean-​shaped, whitish beans.

and the use of thermal processing methods (Cravioto et al. 1945; Bressani 1990;
Tiwari and Singh 2012).
Storage is critical to maintaining quality. The lack of humidity and tempera-
ture control during storage results in a hard seed coat (testa) and a hard-​to-​cook
seed, which requires a longer cooking time (Coelho et al. 2007). Selected essen-
tial amino acids are lost because of the hard-​to-​cook phenomenon.

1.5.4.2.3 Breeding
Common bean breeding programs have to focus on improved cultivars adapted
to the different crop systems in Latin America. Their studies should focus on dis-
ease resistance and abiotic stress tolerances (Villarino 2018).
Future research should include a strategy that encompasses agronomic
features, post-​harvest management, and marketing since producers in Latin
America have limited access to field supplies and facilities for seed storage.
These issues directly affect the quality of the product for consumption. In add-
ition to agronomic traits, investigations using traditional and new molecular
techniques are needed to generate common bean seeds with improved nutri-
tional and nutraceutical messages.
1.5 Selected Crops with Massive Potential for the Future 23

1.5.5 Chia (Salvia hispanica L.)


1.5.5.1 Introduction
Historically, chia nutlets were the principal food of the Nahuatl (Aztec) cul-
ture of central Mexico. The Jesuit chroniclers considered chia as the third most
important crop of the Aztecs, after maize (Zea mays L.) and beans (Phaseolus
spp.), and before amaranth (Amaranthus cruentus L. and A. hypochondriacus L.).
Moreover, tributes and taxes to the Aztec clergy and nobility were frequently
paid in chia nutlet (Miranda-​Colin 1978; Cahill 2004; Muñoz et al. 2013). In pre-​
Columbian times, chia was an essential crop for the civilizations of that time;
it was cultivated in great quantity by the Aztecs, including nutlets in their diet,
in medicine and art, and even as offerings in pagan religious rituals; it also
functioned as currency and as a tax on the people subjected to this empire.
Mayan culture called it chihaan, which means strong or fortress (Cahill 2003).
Since then, it has remained part of the culinary culture of the ethnic groups of
the region. One popular way was through chia fresco, which, following the intro-
duction of the citrus fruits, is generally an orange or lemon soda to which chia
nutlets are added. In recent years, due to its properties as a superfood, being
a source of Omega-​3, antioxidants, and fiber for human nutrition, it has been
in great demand. As a result, ‘La Chia’ has become established as an emerging
crop in different parts of the world (Pascual-​Villalobos et al. 1997; Beltrán and
Romero 2003; Peiretti and Gai 2009; Jamboonsri et al. 2012; Norlaily et al. 2012;
Muñoz et al. 2013; Bochicchio et al. 2015; FAO/​WHO 2016; López et al. 2017;
Rivera-​Cabrera et al. 2017; Peláez et al. 2019; EFSA 2020).

Figure 1.6 Chia crop (A) plant and (B) seeds. Source: Authors’ own work.
Notes: A) The image on the left shows green chia plants with many leaves and flowers
(with blue petals) in the field, and, at the back of the chia field appear other types of green
plants with different sizes and shapes.
B) The image on the right shows a bulk chia stall accompanied by other plant-​based
products such as onions, tomatoes, purple beans in bags, and chickpeas in bags.
24 Characterization of Latin-American Crops

1.5.5.1.1 Botanical description
Chia (Salvia hispanica L.) is part of the Lamiaceae family, in which are found
other species such as mint (Mentha L. sp.), rosemary (Rosmarinus officinalis
L. syn. Salvia rosmarinus Schleid.), and oregano (Origanum vulgare L.). They are
annual herbs up to 1 m, robust, erect, sparsely branched, often with a woody
base; green stems, sometimes with purple pigmentation. Leaves are petiolate,
ovate, or lanceolate, membranous, with a green upper surface, serrated margins,
and acuminate apex; inflorescence is 4–​10 cm, terminal, in dense racemes;
fruiting calyx acrid, markedly urceolate. The corolla is 8–​9 mm, blue or purple;
stamens appear with exerted anthers; exerted style. Nutlets/​mericarps are 1.75
× 1.2 mm, flattened, greyish-​brown, with dark brown mottling. In the wild state,
they flower and fructify from May to October and can be found in the forests of
tree genera like Ulmus L., Pinus L., Quercus L.) and in open grasslands as well as
in bean and maize fields or cultivated (Bentham and Oersted 1854; Cahill 2003;
Hernández-​Gómez et al. 2008; Lobo Zavala et al. 2011).

1.5.5.1.2 Origin and distribution


Chia is an annual species native to the mountainous areas of western and cen-
tral Mexico, and Guatemala. It is found naturally in oak or pine-​oak forests and
is distributed in semi-​warm and temperate environments of the Transversal
Neovolcanic Axis of the western Mother Sierras and southern Chiapas altitudes
1400 and 2200 m a.s.l. (Cahill 2001, 2003; Ayerza and Coates 2005; Norlaily et al.
2012; Sosa-​Baldivia et al. 2018; Peláez et al. 2019).

1.5.5.1.3 Uses
The most common way to use chia is by soaking the nutlets in water; they form
a mucilaginous-​gelatinous mass used as a flavouring in fruit juices which are
consumed as refreshing drinks. The gelatinous mass of the nutlets can also be
prepared as a pudding. In addition, the germinated seedlings are frequently
eaten in salads, sandwiches, soups, and stews. Due to their mucilaginous prop-
erties, they are commonly sown on clay or other porous material, which is kept
moist. From ground nutlets, flour is obtained that can be used to make bread,
stews, and cakes, generally mixed with cereal flours, and it is an excellent source
of easily digestible proteins and fats. Some other uses were as a revitalizer for
combatants who went to war, and for women preparing for childbirth (Cahill
2003, 2005; Bresson et al. 2009; Capitani et al. 2013; Cabrera and Cerna 2014).

1.5.5.2 Limitations
Chia has great potential as a superfood. The ancient Aztec and Mayan people
knew its beneficial characteristics. Several studies have shown that it can be a
rich source of Omega-​3, easy to use. Unfortunately, not much is known about its
cultivation and agronomic practices when cultivated as a monoculture. There
1.5 Selected Crops with Massive Potential for the Future 25

are few experiences in this regard. In addition, there are no varieties or hybrids
that can be recommended for different conditions. Currently, creole material is
used with a mass selection process, as it is likely unknown how it will behave
in different conditions worldwide. In view of its great demand, developing
breeding programs and agronomic practices are necessary to boost this poten-
tial superfood; fortunately, there are some actual efforts in this direction in a few
countries of Latin America.

1.5.5.3 Future Potential
Chia has great potential as a functional food for modern life. Several studies have
focused on the nutritional potential as a supplement for preventing diseases such
as cancer, diabetes, cardiovascular diseases, inflammatory disorders, and nerves
(Muñoz et al. 2013; Hernández-​Pérez et al. 2021). In addition, its soluble dietary
fiber helps to counteract problems of constipation, diverticula, and colon cancer
(Alvarado Rupflin 2011). To do this, ingest 15 to 25 g of nutlets soaked in water
for 15 min. for 20 days (Bernal et al. 2015). This property caused commercializa-
tion of the nutlets to begin in the 1990s.
It is grown in Argentina, Mexico, Bolivia, Paraguay, and Australia. In 2011–​
2012 Argentina had 35% of the production, with Australia, Mexico, Bolivia, and
Paraguay producing 15% and 3000 ha each (Busilacchi et al. 2015). Nicaragua
and Southeast Asian countries recently joined as producers (Jamboonsri et al.
2012). As a result, world production has proliferated; an example is Nicaragua,
where chia production went from 5000 quintals (=​227 metric tonnes) in 2013 to
180,000 quintals (=​8165 metric tonnes) in 2014 (Miranda 2014). Mexico, which
registered 15 ha harvested in 2006, in 2014 this increased to 16,550 ha. Due to
this, chia has a high potential to become an essential crop for human consump-
tion worldwide (Valdivia-​López and Tecante 2015).

1.5.6 Andean maize (Zea mays ssp. mays L.)


1.5.6.1 Introduction
Maize is one of the world’s top three crops, the two others being rice and wheat,
with a small production under adverse conditions in the highland Andes. The
Andean countries currently cultivate a traditional maize population, frequently
amylaceous, white, or colored, for human consumption, and yellow hybrids pri-
marily for feeding animals. However, they are far from becoming self-​sufficient
and usually import additional quantities of maize grain from other American
countries.

Origin and distribution


The center of origin of maize is Mesoamerica. Most researchers now consider
that maize was domesticated 9000 years ago from a lowland tropical teosinte of
26 Characterization of Latin-American Crops

the Balsas valley of Mexico, Zea mays ssp. parviglumis Iltis & Doebley. This single
domestication occurred in a forest located close to the tropic of Cancer between
700 and 1800 m of elevation with 1500 mm of summer rainfall. In a second step,
this proto-​maize has been crossed above 1500 m through gene-​fluxes from
another teosinte, Zea mays ssp. mexicana (Schrad.) Iltis thus adapted to high-
lands (Bonjean 2020).
From 7000 years ago, human migration dispersed this semi-​domesticated
crop to South America. It seems that Tripsacum introgression happened in this
subcontinent, where teosinte is not present (Grobman Tversqui et al. 1961;
Mangelsdorf 1961). Archaeological discoveries proved that maize was present in
Ecuador from 4400 BC, in Peru between 2800 and 2600 BC, in Colombia between
2745 and 2380 BC, and in Uruguay between 2800 and 2500 BC (Bonjean 2020).
Therefore, South America, with its wide variety of agroclimatic environ-
ments, became an expansive secondary stratified domestication centre where
numerous semi-​ domesticated subgroups emerged, in parallel with similar
processes in Mesoamerica (Kistler et al. 2018). Recent studies (Kistler et al. 2020)
demonstrated that as the second wave of Mesoamerican cultivars reached South
America and intercrossed with Andean maize resulting in more productive
landraces between 2300 and 500 BC before their progenies were introduced back
into Central America around 300 BC.

1.5.6.1.1 Genetic Diversification and Cultivation


Three phenomena, each in its own way, have strongly influenced the cultivation,
early spread, and genetic diversification of maize in Mexico, Mesoamerica, South
America, and the Andean area:

• Around 1500−1200 BC, nixtamalization was developed in


Mesoamerica; a hot-​treatment and aqueous process, which uses lime,
and allows a safe daily consumption of maize grain, making it at the
same time more nutritious for the human population (Cravioto et al.
1945; Staller 2010).
• Since at least 3000 years BC, Mesoamerican farmers were primarily
vegetarians. However, they then invented the milpa system, an inter-
cropping of maize, squash (Cucurbita pepo L.), and bean (Phaseolus
spp.), which limited the use of water, pesticides, and fertilizers, offering
a better nutritional diet than maize alone (Gianoli et al. 2006; Almaguer
González et al. 2015).
• Simultaneously, native populations associated maize with several
gods in their native religions and numerous linked rituals in their cul-
ture, the most common being the production and share of chicha, a
low-​alcoholic drink with great religious and social importance (Pérez-​
Suárez 1997; Kulas 2015) in all of Latin America.
1.5 Selected Crops with Massive Potential for the Future 27

During the pre-​Colombian period, maize was a primary staple food for the
Andean populations together with potato (Solanum tuberosum L.), quinoa
(Chenopodium quinoa), amaranth (Amaranthus cruentus L.), tarwi (Lupinus
mutabilis), and other crops, and also a part of their cosmovision.
During the 16th century, when the Conquistadores invaded the Andean
regions, they forced populations to cultivate wheat and barley and adopt the
Christian religion; they pronounced a ban on some Inca crops such as quinoa and
amaranth involved in previous cults. However, due to its high productivity and
large grain, the European invaders adopted maize. Thus, they introduced it into
Europe, where it progressively replaced millet crops (both Panicum miliaceum
L. and Setaria italic (L.) P. Beauv.). Moreover, due to the better adaptation of
maize than wheat to tropical environments, the Europeans also introduced it in
Africa and Asia during their successive waves of colonial expansion.
Later, at the beginning of the 20th century, maize became hybrid in the USA
(Crown 1998) and genetically modified in the 1990s (Lundmark 2007).

1.5.6.2 Limitations
According to CIMMYT, the average yield in Andean countries is 3.6 t ha-​1
compared to a world average yield of 5.4 t ha-​1, mostly because Andean farmers
do not systematically use improved or hybrid seeds, and they lack financial
funds and technical data. Moreover, mechanization is not always easy to imple-
ment in their mountainous zone. Each year Andean nations import part of the
maize grain they consume.
Andean countries’ increasing food security from maize production would
need better quality, improved genetics, and durable field practices, including irri-
gation and better post-​harvest storage facilities (González 2018) (see Figure 1.7).

1.5.6.3 Future Potential
All Andean countries have maintained a large diversity of traditional maize
populations adapted to various Andean region environments, both in situ with
local farmers and ex-​situ within national and CIMMYT genebanks (Roberts
et al. 1957).
From the beginning of the 1990s, CIMMYT and Andean institutions have
developed a breeding program dedicated to subtropical, mid-​altitude, and high-
land maize (Bjarnason 1994; Galeano et al. 2019). In 2008 CIMMYT, CIAT, and
Harvest Plus also launched a zinc-​enriched maize program to address zinc
deficiency, which is usually characterized by growth retardation, loss of appe-
tite, and impaired immune function, and affects part of the Andean population
(Michail 2019).
The local Andean maize varieties own wide phenotypic diversity (Staller
2016) and include a floury and soft grain with varied colors, shapes, and sizes,
which constitute an attractive breeding reservoir for future selection of staple
28 Characterization of Latin-American Crops

Figure 1.7 Andean maize (A) Plants in Bolivia, and (B) maize ears. Source: Authors’
own work.
Notes: A) The image on the left shows an Andean maize plantation at the back of a sparse
bean field. In the backgound, some tall trees are visible.
B) The image on the right shows many different Andean maize ear types in yellow, black,
and white in which the grains also vary in color within an ear.

food that needs to be protected from genetically modified (GM) maize pollen
fluxes. In addition, their natural pigments and antioxidants are related to health
benefits.
Climate change affects the Andean region (higher temperatures, reduction in
glacier sizes, instability of rainfall and water flow, etc.) as with other parts of the
planet. Second are damage from disease, insects, and other pests increase, and
risks of soil erosion and loss of plant and animal biodiversity. Finally, the warmer
climate also changes cropping systems: maize and potato can now be grown at
higher altitudes than previously, frequently accompanied by the development of
terracing (Skarbø and van der Molen 2016).
In addition, drought is becoming a more common issue. It requires improve-
ment in water management and selection of maize material more tolerant to
abiotic stresses. Technologies such as tilling and gene editing offer hope in this
direction. Agroforestry with native trees may be another solution, which allows
stabilizing mountain slopes and riverbanks.
The challenge of increasing Andean maize production while protecting the
fragile Andean environment and its endemic biodiversity lies mainly in the
development of conservation agriculture.

1.5.7 Sacha Inchi (Plukenetia volubilis L.)


1.5.7.1 Introduction
Plukenetia volubilis is an underutilized oilseed crop native to the Amazon
basin, where humans have utilized it since Incan times. The large seeds contain
1.5 Selected Crops with Massive Potential for the Future 29

Figure 1.8 Sacha inchi plant with fruits. Source: Authors’ own work.
Note: The image shows a Sacha inchi plant with fruit capsules of four lobes and condiform
shaped leaves in the field. At the back of the Sacha inchi field, some short palm trees are
visible in a dark green color.

45–​50% lipid, of which 35.2–​50.8% is α-​linolenic acid (C18:3 n-​3, ω-​3) and 33.4–​
41.0% is linoleic acid (C18:2 n-​6, ω-​6), the two essential fatty acids required by
humans. The seeds also contain 22–​30% protein and have antioxidant properties
(Hamaker et al. 1992; Follegatti-​Romero et al. 2009; Gutiérrez et al. 2011; Maurer
et al. 2012; Chirinos et al. 2013; Cisneros et al. 2014; Triana-​Maldonado et al.
2017). As a result, its excellent nutritional composition and good agronomic
properties have attracted increasing attention in recent years, and cultivation
is expanding.

1.5.7.1.1 Morphology, phylogeny, and distribution


Plukenetia volubilis is a perennial liana with large, oleaginous seeds. The
plants are monecious; the leaves are triangular to ovate with a truncate to
cordate base, palmate venation and basilaminar glands, usually with a small
knob between them. The racemose inflorescence is axillary or terminal with
one to two pistillate flowers situated basally and numerous small, incon-
spicuous, staminate flowers in condensed cymes situated above (Kodahl and
Sørensen 2021). The winged ovary has four carpels, and the style column is
elongate and cylindrical, four-​lobed at the apex. During fruit maturation, the
ovary develops from green and fleshy to brown, woody, and dehiscent. The
30 Characterization of Latin-American Crops

seeds are lenticular, around 1.8 × 0.8 × 1.6 cm in size, and the testa is hard
and brown, with dark brown markings. In cultivation, the fruit is often larger
sized and is five-​or six-​carpellate (Gillespie 1993; Gillespie and Armbruster
1997). The genus Plukenetia L. (Euphorbiaceae) comprises 25 species, several
of which have only recently been described. Circumscription of the genus
has undergone several changes during the last four centuries, but Cardinal-​
McTeague and Gillespie (2020) recently revised the classification. Plukenetia
belongs to tribe Plukenetieae, subtribe Plukenetiinae, and is distinguished
by four-​carpellate ovaries and two extrafloral nectaries situated basally on
the adaxial surface of the leaf blade. The genus is divided into two major
clades: the pinnately veined clade, one primary vein, and the palmately
veined clade, including P. volubilis, with three to five primary veins (Cardinal-​
McTeague and Gillespie 2020).
Plukenetia volubilis is distributed widely in South America and the Lesser
Antilles; it is found in the northern and western parts of the Amazon Basin in
Surinam, Venezuela, Colombia, Ecuador, Peru, Bolivia, and Brazil (Gillespie and
Armbruster 1997). The most common ecological niche for P. volubilis is a moist
to wet lowland forest, but the species complex comprises two morphologic-
ally differing groups: an open savannah and a mid-​elevation species group. The
open savannah group generally has thicker leaf blades and smaller seeds and
fruits, while the mid-​elevation group has narrower leaf blades and differing leaf
base morphology compared with typical P. volubilis. However, further studies
are needed to define species group boundaries better and assess whether the
definition of a new, additional species from within the complex is warranted
(Cardinal-​McTeague and Gillespie 2020).

1.5.7.1.2 Uses
Plukenetia volubilis has traditionally been consumed in Peru and other coun-
tries of Latin America, and has been associated with humans since pre-​
Hispanic times. Artifacts depicting P. volubilis fruits and vines have been
found in Incan burial sites along the coast of Peru, indicating that the plant
may have been cultivated by pre-​Incan cultures 3000–​5000 years ago (Brack
Egg 1999).
The vernacular name ‘Sacha inchi’ is Quechuan and is the most commonly
used name for P. volubilis and other large-​seeded species in the genus. However,
‘Sacha inchik’ or ‘Sacha inchic’ is also used depending on the dialect, and the
meaning is the same; ‘sacha’ can be translated to ‘mountain’ or sometimes to
‘false’ or ‘resembling’, and ‘inchi’ means groundnut/​ peanut. Less common
names for P. volubilis include Sacha Yachi, Sacha Yuchi, Sacha yuchiqui, Yuchi,
sampannankii, suwaa, Correa, amauebe, amui-​ o, maní de Arbol, maní del
monte, and maní Estrella, several of which hint at the nut-​like texture of the seed
(Brack Egg 1999; D. Cacique pers. comm.). Accordingly, some of the most often
1.5 Selected Crops with Massive Potential for the Future 31

mentioned culinary uses are similar to the uses of groundnuts; P. volubilis seeds
are most commonly consumed roasted and salted as a snack but are also used
in confectionery, e.g. dipped in chocolate, or ground to a butter-​like substance,
milled to flour, or used in a large variety of traditional dishes. These include ‘inchi
cucho’ (a spicy, savoury sauce or dip), ‘lechona api’ (plantain porridge), and ‘inchi
capi’ (chicken or beef soup). Likewise, the young leaves are occasionally eaten in
salads or in tea (Flores 2010; Flores and Lock 2013).
However, while P. volubilis has many culinary applications, the most reported
use was shown by an ethnobotanical study performed in San Martín, Peru,
which was related to health (67% of answers) (Rodríguez del-​Castillo et al.
2019). Accordingly, several Peruvian ethnic groups, including the Mayorunas,
Chayuhitas, Shipibas, and Boras, have traditionally used a mixture of P. volubilis
ground seeds and seed oil as a skin cream to rejuvenate and revitalize the skin.
Similarly, the Secoyas, Candoshis, Amueshas, and Cashibos, among others, have
rubbed P. volubilis oil on the skin to relieve muscle pain and rheumatism (Flores
and Lock 2013). In addition, the oil and roasted seeds have been consumed
for cholesterol control, and for cardiovascular and gastrointestinal problems
(Rodríguez del-​Castillo et al., 2019).
With the increasing awareness and popularity of P. volubilis in international
markets in recent years, several new or differently branded products have
also become available, e.g. gourmet oil, protein powder, and encapsulated oil
marketed as a dietary supplement. In addition, roasted and salted, or candied,
seeds are marketed.

1.5.7.2 Limitations and Breeding Opportunities


Plukenetia volubilis is an up-​and-​coming crop, primarily due to the nutritional
composition of the seeds; however, further study of the plant and development
of breeding strategies will probably prove beneficial. Importantly, although
there is a focus on sustainability in the plants’ native range, exploration of
more sustainable management practices will be advantageous, both regarding
biodiversity and climate, but might also improve product quality and provide
opportunities for product branding and marketing. Moreover, even though the
seed oil has been approved for consumption in the European Union, the seeds of
P. volubilis are not yet approved due to a lack of knowledge concerning alkaloid
content and composition in the seeds (EFSA 2020). Studies indicate that thermal
processing significantly reduces the amount of alkaloid compounds in the seeds
(Srichamnong et al. 2018). However, the recent decision from the European Food
Safety Authority (EFSA 2020) nevertheless underlines the need for further docu-
mentation of the safety of consumption of the seeds, including details on poten-
tially allergenic or toxic compounds.
Very little breeding of P. volubilis has been carried out, so the plant is not
considered to be fully domesticated (Vašek et al. 2017). Breeding should be
32 Characterization of Latin-American Crops

aimed at improving the agronomy of the crop, higher yield and improved sen-
sory qualities. Similarly, further exploration of the domestication potential of
other large-​seeded species in the genus would be of interest, both as crops in
their own right but possibly also as material for the development of hybrids
with P. volubilis.

1.5.7.3 Future Potential
Current global challenges include climate change, degradation of land and
environment, population growth, and lack of food security. Accordingly, our
food systems need to be optimized to ensure food security while avoiding
global ecosystem collapse and the subsequent loss of ecosystem services.
It is becoming ever more apparent that upscaling of current agricultural
systems, particularly monocultures, is not the best procedure for this pur-
pose, and alternative strategies are needed ( Jacobsen et al. 2015; Funabashi
2018). Furthermore, neglected and underutilized crops may prove necessary
resources for the reformation of our food production systems by improving,
e.g., climate change resilience, genetic diversity, and the nutritional value of
agricultural products.
Plukenetia volubilis has considerable potential for contributing to these
goals, as the plants can thrive in a broad range of environmental conditions,
have an exceptional nutritional composition, and possibly be an economically
beneficial alternative crop for small-​scale farmers. Further, a wide variety of
cultivars and a high genetic diversity of germplasm is available, providing out-
standing opportunities for further domestication and breeding, ultimately to
establish the integration of P. volubilis in sustainable cultivation systems. Thus,
on a local scale, P. volubilis may aid in food security, alleviate malnutrition, and
provide economic benefits, while simoultaneously being part of the solution to
our global challenges. Consequently, P. volubilis has considerable potential for
further domestication.

1.6 GENERAL DISCUSSION
There is great potential for increasing production of Latin-​American crops, such
as those presented in this chapter. However, it must be done in a sustainable
and wise way. In Ecuador, an increase in demand for Andean lupine intensi-
fied the production system of the crop by using improved varieties and a larger
cultivated area. It caused an escalation in pests and the indiscriminate use of
insecticides (Mina et al. 2018). It is obvious that increased production of new
Andean or Latin-​American crops may lead to new problems, which we must be
prepared for in order to choose the best possible strategies.
References 33

The failure of agricultural development implemented in the Andean highland


and Latin America among other factors has perpetuated poverty and malnutri-
tion in the population. Ancient technologies –​in danger of abandonment –​seem
to be one of the options for increased food production and improvement of the
actual standard of living.
Production of healthy, natural, and nutraceutical food, like the crops
analyzed in this report, associated with environmental modifications, is the
best way to obtain economically acceptable organic production under challen-
ging conditions of climate, soil, and slope. The ancient knowledge of cultiva-
tion suggests rational use of soil, water, and nutrients; integral control of pests
and diseases; and processing and use of foods. We must solve the nutritional
problems that rural communities face throughout the Andes mountains as in
other regions of the Latin-​American and Caribbean subcontinent.
The altiplano of Peru and Bolivia and traditional Latin-​American agricultural
systems can be considered under-​utilized agroecosystems, capable of reverting
the current food deficiency by applying a rational and suitable use of existing
environmental modifications and using the great diversity and variability of
highly nutritious native Andean and Latin American crops.
Specific data about the consumption of Andean and Latin-​American crops
are needed, which will be useful for making decisions and taking actions that
allow their promotion for more consumption. Several crops offer an option for
future food demand, and some of them have been included in this chapter. They
are examples of crops with a great potential for increased use for food locally
and regionally, and also serving as export crops thus creating income for farmers
and in general for the Latin-​American region. The species mentioned here have
high nutritional values, are healthy, climate-​resilient, and tasty, and they may be
important components for more plant-​based foods, replacing meat. These crops
are considered superfoods because of their outstanding nutritional and nutra-
ceutical characteristics; therefore, they should be key components in the diet of
people in Latin America and in the rest of the world.

REFERENCES
Adel, H. 2020. Towards expanding quinoa cultivation in Egypt: The effect of compost and
vermicompost on quinoa pests, natural enemies and yield under field conditions.
Agricultural Sciences 11:191−209. https://​doi.org/​10.4236/​as.2020.112​012.
Akinnifesi, F. K., R. R. B. Leakey, O. C. Ajayi, et al. 2008. Indigenous Fruit Trees in the
Tropics: Domestication, Utilization and Commercialization. CAB International
Publishing, Wallingford, UK.
Alandia, G., J. P. Rodriguez, S.-​E. Jacobsen, D. Bazile, and B. Condori. 2020. Global expan-
sion of quinoa and challenges for the Andean region. Global Food Security 26: 100429.
https://​doi.org/​10.1016/​j.gfs.2020.100429.
34 Characterization of Latin-American Crops

Almaguer González, J. A., H. J. García Ramírez, M. Padilla Mirazo, and M. González Ferral.
2015. La dieta de la milpa: modelo de alimentacion mesoamericana biocompatible.
Secretaria de Salud, Mexico. https://​ali​anza​nahu​aca.org/​2018/​07/​10/​la-​dieta-​de-​la-​
milpa-​mod​elo-​de-​alime​ntac​ion-​mesoam​eric​ana/​.
Alvarado Rupflin, D. I. 2011. Caracterización de la semilla de chan (Salvia hipanica L.) y
diseño de un producto funcional que la contiene como ingrediente. Revista de la
Universidad del Valle Guatemala 23: 43–​49. https://​xdoc.mx/​prev​iew/​cara​cter​izac​
ion-​de-​la-​semi​lla-​del-​chan-​sal​via-​hispan​ica-​l-​y-​diseo-​602f48​2d94​6c8.
Alvar-​Beltran, J., A. Dao, A. D. Marta, et al. 2019. Effect of drought, nitrogen fertilization,
temperature and photoperiodicity on quinoa plant growth and development in the
Sahel. Agronomy 9: 607. https://​doi.org/​10.3390/​agro​nomy​9100​607.
Atchison, G. W., B. Nevado, R. J. Eastwood, N. Contreras-​Ortiz, et al. 2016. Lost crops
of the Incas: Origins of domestication of the Andean pulse crop Tarwi, Lupinus
mutabilis. American Journal of Botany 103(9): 1592–​1606. https://​doi.org/​10.3732/​
ajb.1600​171.
Ayerza, R., and W. Coates. 2005. Chia: Rrediscovering a Forgotten Crop of the Aztecs.
University of Arizona Press, Tucson, AZ.
Bazile, D., S.-​ E. Jacobsen, and A. Verniau. 2016. The global expansion of Quinoa:
Trends and limits. Frontiers in Plant Science 7: 622. https://​doi.org/​10.3389/​
fpls.2016.00622.
Beebe, S., P. W. Skroch, J. Tohme, M. C. Duque, F. Pedraza, and J. Nienhuis. 2000. Structure of
genetic diversity among common bean landraces of Middle American origin based
on correspondence analysis of RAPD. Crop Science 40(1): 264–​273. https://​doi.org/​
10.2135/​crop​sci2​000.4012​64x.
Beltrán, O. M. C., and M. R. Romero. 2003. La chía, alimento milenario. Departamento de
Graduados e Investigación en Alimentos. ENCB. Instituto Politécnico Nacional (IPN),
México.
Bentham, G., and A. S. Oersted. 1854. Labiatae centroamericanae. Videnskabelige
Meddelelser fra Dansk Naturhistorisk Forening i Kjøbenhavn 1853(1–​2): 32–​42.
Bermúdez-​Torres, K., J. M. Herrera, R. F. Brito, M. Wink, and L. Legal. 2009. Activity of
quinolizidine alkaloids from three Mexican Lupinus against the lepidopteran crop
pest Spodoptera frugiperda. BioControl 54: 459–​ 466. https://​doi.org/​10.1007/​s10​
526-​008-​9180-​y.
Bernal, A. E., J. J. Iñaguazo, and B. Chanducas. 2015. Efecto del consumo de chía (Salvia
hispanica) sobre los síntomas de estreñimiento que presentan los estudiantes de
una universidad particular de Lima Este. Revista Cientifica de Ciencias de la Salud
8(2): 8–​24. https://​doi.org/​10.17162/​rccs.v8i2.468.
Bertero, H. D., A. J. de la Vega, G. Correa, S.-​E. Jacobsen, and A. Mujica. 2004. Genotype and
genotype-​by-​environment interaction effects for grain yield and grain size of quinoa
(Chenopodium quinoa Willd.) as revealed by pattern analysis of multi-​environment
trials. Field Crops Research 89: 299−318. https://​doi.org/​10.1016/​j.fcr.2004.02.006.
Beyra, A., and G. Reyes Artiles. 2004. Revisión taxonómica de los géneros Phaseolus y Vigna
(Leguminosea-​Papilionoideae) en Cuba. Anales del Jardín Botánico de Madrid (1979)
61(2): 135–​154. https://​doi.org/​10.3989/​ajbm.2004.v61.i2.41.
Bitocchi, E., L. Nanni, E. Bellucci, et al. 2012. Mesoamerican origin of the common bean
(Phaseolus vulgaris L.) is revealed by sequence data. PNAS 109: E788–​E796. https://​
doi.org/​10.1073/​pnas.110​8973​109.
References 35

Bitocchi, E., D. Rau, E. Belucci, et al. 2017. Beans (Phaseolus ssp.) as a model for
understanding crop evolution. Frontiers in Plant Science 8(122): 1–​21. https://​doi.
org/​10.3389/​fpls.2017.00722.
Blair, M. W., A. J. Cortés, R. Varma Penmetsa, A. Farmer, N. Carrasquilla-​Garcia, and
D. R. Cook. 2013. A high-​throughput SNP marker system for parental polymorphism
screening, and diversity analysis in common bean (Phaseolus vulgaris L.).
Theoretical and Applied Genetics 126(2): 535–​ 548. https://​doi.org/​10.1007/​s00​
122-​012-​1999-​z.
Bochicchio, R., T. D. Philips, S. Lovelli, et al. 2015. Innovative crop productions for healthy
food: The case of chia (Salvia hispanica L.). In The Sustainability of Agro-​Food and
Natural Resource Systems in the Mediterranean Basin, ed. A. Vastola, 29–​45. Springer
Nature AG, Cham, Switzerland. https://​doi.org/​10.1007/​978-​3-​319-​16357-​4_​3.
Bonifacio, A. 2019. Improvement of Quinoa (Chenopodium quinoa Willd.) and Qañawa
(Chenopodium pallidicaule Aellen) in the context of climate change in the high
Andes. International Journal of Agriculture and Natural Resources 46(2): 113−124.
http://​dx.doi.org/​10.7764/​rcia.v46i2.2146.
Bonjean, A. P. 2020. Le maïs, rouleau compresseur américain. Paysans & société 383: 38−46.
https://​doi.org/​10.3917/​pes.383.0038.
Borlaug, Norman E., 2000. Ending world hunger. The promise of biotechnology and
the threat of antiscience zealotry. In Plant Physiology. American Society of Plant
Biologists (OUP).124(2): 487–​490. doi:10.1104/​pp.124.2.487. ISSN 1532-​2548.
Brack Egg, A. 1999. Diccionario Enciclopedico de Plantas Utiles del Peru, p. 400. PNUD,
Cuzco, Peru.
Bravo, R., R. Valdivia, K. Andrade, S. Padulosi, and M. Jager. 2010. Granos andinos: avances,
logros y experiencias desarrolladas en quinua, cañihua y kiwicha en Perú. Bioversity
International, Roma. www.biov​ersi​tyin​tern​atio​nal.org/​e-​libr​ary/​publi​cati​ons/​det​
ail/​gra​nos-​andi​nos-​avan​ces-​log​ros-​y-​exper​ienc​ias-​desarr​olla​das-​en-​qui​nua-​cani​
hua-​y-​kiwi​cha-​en-​peru/​.
Bressani, R. 1990. Chemistry, technology, and nutritive value of maize tortillas. Food
Reviews International 6(2): 225–​264. https://​doi.org/​10.1080/​875591​2900​9540​868.
Bresson, J. L., A, Flynn, and M. Heinonen. 2009. Opinion on the safety of chia seeds (Salvia
hispanica L.) and ground whole chia seeds as a food ingredient. European Food Safety
Authority Journal 996: 1–​26. https://​doi.org/​10.2903/​j.efsa.2009.996.
Broughton, W. J., G. Hernández, M. Blair, S. Beebe, P. Gepts, and J. Vanderleyden. 2003.
Beans (Phaseolus spp.) –​model food legumes. Plant and Soil 252: 55–​128. https://​
doi.org/​10.1023/​ A:1024146710611.
Bruno, M. C. 2006. A morphological approach to documenting the domestication of
Chenopodium in the Andes. In Documenting Domestication: New Genetic and
Archaeological Paradigms, ed. M. A. Zeder, D. G. Bradley, E. Emshwiller, and B .D.
Smith, 32–​45. University of California Press, Berkeley, CA.
Bruno, M. C. 2014. Beyond raised fields: Exploring farming practices and processes
of agricultural change in the ancient Lake Titicaca Basin of the Andes. American
Anthropologist 116(1): 130–​145. https://​doi.org/​10.1111/​aman.12066.
Bunsupa, S., K. Katayama, E. Ikeura, et al. 2012. Lysine decarboxylase catalyzes the
first step of quinolizidine alkaloid biosynthesis and coevolved with alkaloid
production in Leguminosae. Plant Cell 24(3): 1202–​1216. https://​doi:10.1105/​
tpc.112.095​885.
36 Characterization of Latin-American Crops

Busilacchi, H., T. Qüesta, and S. Zuliani. 2015. La chía como una nueva alternativa
productiva para la región pampeana. Agromensajes 41(2): 37–​46. https://​core.ac.uk/​
downl​oad/​pdf/​162568​326.pdf.
Cabrera, J. C., and M. F. Cerna. 2014. Optimizaciòn de la aceptabilidad de un pan integral
de chia (Salvia hispanica L.) mediante la metodologìa de Taguchi. Agroindustrial
Science 4(1): 19–​25. https://​doi.org/​10.17268/​agroind.sci.
Cahill, J. P. 2001. Domestication of chia, Salvia hispanica L. (Lamiaceae). PhD thesis,
University of California, Riverside, CA.
Cahill, J. P. 2003. Ethnobotany of chia, Salvia hispanica L. (Lamiaceae). Economic Botany
57: 604–​618. https://​doi.org/​10.1663/​0013-​0001(2003)057[0604:EOC​SHL]2.0.CO;2.
Cahill, J. P. 2004. Genetic diversity among varieties of chia (Salvia hispanica L.). Genetic
Resources and Crop Evolution 51: 773–​781. https://​doi.org/​10.1023/​B:GRES.000​0034​
583.20407.80.
Cahill, J. P. 2005. Human selection and domestication of chia (Salvia hispanica L.). Journal
of Ethnobiology 25: 155–​174. https://​doi.org/​10.2993/​0278-​0771(2005)25[155:HSA​
DOC]2.0.CO;2.
Caicedo, C., and E. Peralta. 2000. Zonificación potencial, sistemas de producción y
procesamiento artesanal de chocho (Lupinus mutabilis Sweet) en Ecuador. Boletín
Técnico Nº 89. Programa Nacional de Leguminosas. Estación Experimental Santa
Catalina. INIAP-​FUNDACYT-​P-​BID-​206. Quito-​Ecuador. https://​repo​sito​rio.iniap.
gob.ec/​bitstr​eam/​41000/​441/​4/​inia​pscb​t89.pdf.
Caicedo, C., E. Peralta, E. Villacrés, and M. Rivera, 2001. Postcosecha y mercado del chocho
(Lupinus mutabilis Sweet) en Ecuador. INIAP, FUNDACYT, Quito-​Ecuador. https://​
repo​sito​rio.iniap.gob.ec/​bitstr​eam/​41000/​2700/​1/​iniap​scpm​105.pdf.
Caligari, P. D. S., P. Römer, M. A. Rahim, C. Huyghe, J. Neves-​Martins, and E. J. Sawicka-​
Sienkiewicz. 2000. The potential of Lupinus mutabilis as a crop. In Linking Research
and Marketing Opportunities for Pulses in the 21st Century: Proceedings of the Third
International Food Legumes Research Conference, ed. R. Knight, 569–​573. Springer,
DordrechtNetherlands). https://​doi.org/​10.1007/​978-​94-​011-​4385-​1_​54.
Calle, E. 2004. Efecto de la cañahua (Chenopodium pallidicaule Aellen), trito (Triticum
aestivum L.), soya (Glycine max Merr.), germinados en la alimentacion de cuyes en
recria para la prevencion de escorbuto. PhD thesis, Universidad Tecnica de Oruro,
Oruro, Bolivia.
Canahua, A., and P. Román. 2016. Tarwi Leguminosa andina de gran potencial. LEISA
Revista de Agroecología 32(2): 20–​22. www.canun​ite.org/​wp-​cont​ent/​uplo​ads/​2016/​
07/​Leisa_​vol3​2n2.pdf.
Capitani, M. I., S. M. Nolasco, and M.C. Tomás. 2013. Effect of mucilage extraction on the
functional properties of Chia meals. In Food Industry, ed I. Muzzalupo, Ch. 1, 1–​19.
IntechOpen. https://​doi.org/​10.5772/​53171. Available from: www.int​echo​pen.com/​
chapt​ers/​41676.
Cardinal-​McTeague, W. M., and L. J. Gillespie. 2020. A revised sectional classification of
Plukenetia L. (Euphorbiaceae, Acalyphoideae) with four new species from South
America. Systematic Botany 45(3): 507–​536. https://​doi.org/​10.1600/​03636​4420​X159​
3529​4613​572.
Carrasco, D. E., and J. L. Soto. 2010. II. Importancia de los granos andinos. In Granos Andinos.
Avances, Logros y Experiencias Desarrolladas en Quinua, cañahua y Amaranto en
References 37

Bolivia, ed. W. Rojas, J. L. Soto, M. Pinto, M. Jager, and S. Padulosi, 6–​10. Bioversity
International, Rome. https://​cgsp​ace.cgiar.org/​bitstr​eam/​han​dle/​10568/​104​701/​
Granos_​andinos_​avances_​logros_​y_​e xperiencias​_​d es​arro​l lad​a s_​e​n_​qu​inua​_​
ca%C3%B1ahua​_​y_​a​mara​nto_​en_​B​oliv​ia_​1​413.pdf ?seque​nce=​3&isAllo​wed=​y.
CGIAR-​Beanfocus. 2019. Common bean: The nearly perfect food –​The importance of
common bean. http://​ciat-​libr​ary.ciat.cgiar.org/​art​icul​os_​c​iat/​ciat​info​cus/​beanfo​
cus. pdf (accessed December 28, 2020).
Chambi, F. A. 2019. Elaboración de cup-​cakes con sustitución parcial de harina de trigo
con harina de quinua (Chenopodium quinoa), kiwicha (Amaranthus caudatus),
cañihua (Chenopodium pallidicaule) y sustitución de grasa por gomas de linaza
(Linum usitatissimum) y chia (Salvia hispánica). PhD thesis, Universidad Peruana
Unión, Ñaña, Lima. http://​repo​sito​rio.upeu.edu.pe/​han​dle/​UPEU/​2941 (accessed
December 3, 2020).
Chirinos, R., G. Zuloeta, R. Pedreschi, E. Mignolet, Y. Larondelle, and D. Campos. 2013.
Sacha inchi (Plukenetia volubilis): A seed source of polyunsaturated fatty acids,
tocopherols, phytosterols, phenolic compounds and antioxidant capacity. Food
Chemistry 141: 1732–​1739. https://​doi.org/​10.1016/​j.foodc​hem.2013.04.078.
Chirinos-​Arias, M., E. J. Jiménez, and S. L. Vilca-​Machaca. 2015. Analysis of genetic vari-
ability among thirty accessions of Andean Lupin (Lupinus mutabilis Sweet) using
ISSR molecular markers. Scientia Agropecuaria 6: 17–​30. https://​doi.org/​10.17268/​
sci.agrop​ecu.2015.01.02.
CIAT (Centro Internacional de Agricultura Tropical). 2020. Conservación y uso de cultivos.
https://​ciat.cgiar.org/​lo-​que-​hace​mos/​conse​rvac​ion-​y-​uso-​de-​culti​vos/​?lang=​es.
Ciesiolka, D., Gulewicz, P., Martínez-​Villaluenga, C., Pilarski, R., Bednarczyk, M., and
Gulewicz, K. 2005. Products and biopreparations from alkaloid-​rich lupin in animal
nutrition and ecological agriculture. Folia Biologica (Kraków) 53: 59–​66. https://​doi.
org/​10.3409/​173​4916​0577​5789​443.
Cisneros, F. H., D. Paredes, A. Arana, and L. Cisneros-​Zevallos. 2014. Chemical compos-
ition, oxidative stability and antioxidant capacity of oil extracted from roasted seeds
of sacha-​inchi (Plukenetia volubilis L.). Journal of Agricultural and Food Chemistry
62: 5191–​5197. https://​doi.org/​10.1021/​ jf500936j.
Clavijo, P. J. 1980. Resumen general de las principales características agronómicas de
diferentes granos en Colombia. Instituto Interamericano de Ciencias Agrícolas. IICA,
Colombia. https://​repo​sito​rio.iica.int/​bitstr​eam/​han​dle/​11324/​15678/​CDRP21​0415​
16e.pdf ?seque​nce=​1&isAllo​wed=​y.
Clements, J. C., M. S. Sweetingham, L. Smith, G. Francis, G. Thomas, and S. Sipsas. 2008.
Crop improvement in Lupinus mutabilis for Australian agriculture-​progress and
prospects. In Lupins for Health and Wealth. Proceedings of the 12th International Lupin
Conference, Fremantle, Western Australia, 14–​18 September 2008, ed. J. A. Palta, and J.
B. Berger, 244–​250. International Lupin Association, Canterbury, New Zealand. ISBN
(Print) 0864761538.
Coelho, C. M. M., C. de Mattos Bellato, J. C. P. Santos, E. M. M. Ortega, and S. M. Tsai. 2007.
Effect of phytate and storage conditions on the development of the ‘hard-​to-​cook’
phenomenon in common beans. Journal of the Science of Food and Agriculture
87: 1237–​1243. https://​doi.org/​10.1002/​jsfa.2822, www.fao.org/​faos​tat/​en/​?#data/​
QC (accessed December 28, 2020).
38 Characterization of Latin-American Crops

Coila, R. A. 2019. Optimización en la elaboración de galletas utilizando harina de


cañihua (Chenopodium pallidicaule), kiwicha (Amaranthus caudatus) y quinua
(Chenopodium quinoa). PhD thesis, Universidad Nacional de Trujillo, Peru. https://​
1libr​ary.co/​docum​ent/​nzwlm​p1y-​optim​izac​ion-​elab​orac​ion-​galle​tas-​uti​liza​ndo-​
chen​opod​ium-​palli​dica​ule-​ama​rant​hus-​chen​opod​ium.html (accessed November
24, 2020).
Contin Gomes, M. J., H. S. Duarte Martino, and E. Tako 2021. Effects of iron and zinc
biofortified foods on gut microbiota in vivo (Gallus gallus): A systematic review.
Nutrients 13(1): 189. https://​doi.org/​10.3390/​nu1​3010​189.
Cortez Quispe, H. A. 2016. Evaluación de cuatro niveles de polvillo de Qañäwa
(Chenopodium pallidicaule A.) en la alimentación de Cuyes (Cavia porcellus L.) en
crecimiento. Revista Apthapi 1(2): 86–​95. http://​ojs.agro.umsa.bo/​index.php/​ATP/​
arti​cle/​view/​145/​145.
Cowling, W. A., B. J. Buirchell, and M. E. Tapia, 1998. Lupin. Lupinus. Promoting the con-
servation and use of underutilized and neglected crops. International Plant Genetic
Resources Institute [IPGRI], Rome. www.biov​ersi​tyin​tern​atio​nal.org/​filead​min/​user​
_​upl​oad/​onl​ine_​libr​ary/​ publications/​pdfs/​Lupin_​23.pdf.
Cravioto, R. O., R. K, Anderson, E. E. Lockhart, F. de P. Miranda, and R. S. Harris. 1945.
Nutritive value of the Mexican tortilla. Science 102 (2639): 91–​93. https://​doi.org/​
10.1126/​science.102.2639.91.
Cremer, H. D. 1983. Current aspects of legumes as a food constituent in Latin America with
special emphasis on lupines: Introduction. Plant Foods for Human Nutrition 32: 95–​
100. https://​doi.org/​10.1007/​BF01091329.
Crown, J. F. 1998. 90 years ago: The beginning of hybrid maize. Genetics 148(3): 923−928.
www.ncbi.nlm.nih.gov/​pmc/​artic​les/​PMC​1460​037/​pdf/​9539​413.pdf.
De Ron, A., F. Sparvoli, J. Pueyo, and D. Bazile. 2017. Editorial: Protein crops: Food and
feed for the future. Frontiers in Plant Science 8: 105. https://​doi.org/​10.3389/​
fpls.2017.00105.
Doxastakis, G. 2000. Lupin seed proteins. Developments in Food Science 41: 7–​38. https://​
doi.org/​10.1016/​S0167-​4501(00)80004-​7.
EFSA (European Food Safety Authority). 2020. Technical Report on the notification of
roasted seeds from Plukenetia volubilis L. as a traditional food from a third country
pursuant to Article 14 of Regulation (EU) 2015/​2283. European Food Safety Authority
Journal Supporting publication 2020: EN-​1817:17(3): 1–​13. https://​doi.org/​10.2903/​
sp.efsa.2020.EN-​1817.
Erickson, C. L. 1985. Applications of prehistoric Andean Technology: Experiments in
raised field agriculture, Huatta, Lake Titicaca: 1981–​1982. In Prehistoric Intensive
Agriculture in the Tropics, ed. I. S. Farrington, 209–​232. British Archaeological
Reports, Oxford.
Espinoza-​Moreno, R. J., C. Reyes-​Moreno, J. Milán-​Carrillo, J. A. López-​Valenzuela, O.
Paredes-​ López, and R. Gutiérrez-​ Dorado. 2016. Healthy ready-​to-​eat expanded
snack with high nutritional and antioxidant value produced from whole amarantin
transgenic maize and black common bean. Plant Foods for Human Nutrition 71: 218–​
224. https://​doi.org/​10.1007/​s11​130-​016-​0551-​.
Estaña, W., and C. Muñoz. 2012. Variabilidad genética de la Cañihua en las provincias
de Puno. Equipo Tecnico del “Proyecto Mejoramiento de Capacidades Técnico
Productivas para la Competitividad de los Cultivos Andinos de Papa Nativa, Haba
References 39

y Cañihua en la Region Puno”, Puno, Peru. https://​docpla​yer.es/​60307​065-​Cani​hua-​


varia​bili​dad-​genet​ica-​en-​las-​pro​vinc​ias-​de-​puno.html.
Falconí, C. 2012. Lupinus mutabilis in Ecuador with special emphasis on anthracnose
resistance. PhD dissertation. [Wageningen (NL)]: Wageningen University, the
Netherlands.
FAO (Food and Agriculture Organization of the United Nations). 1992. Cultivos marginados,
otra perspectiva de 1492. Rome. www.fao.org/​3/​t06​46s/​t06​46s.pdf (accessed
September 12, 2020).
FAO (Food and Agriculture Organization of the United Nations). 2000. Cultivos andinos
subexplotados y su aporte a la alimentación, 2nd edn. Santiago de Chile-​Chile. www.
fao.org/​temp​ref/​GI/​Reser​ved/​FTP​_​Fao​Rlc/​old/​prior/​sega​lim/​ prodalim/​prodveg/​
cdrom/​contenido/​libro10/​home10.htm (accessed 12 September 2020).
FAO (Food and Agriculture Organization of the United Nations). 2013. The International
Year of Quinoa –​A future sown thousands of years ago. www.fao.org/​qui​noa-​
2013/​en/​.
FAO (Food and Agriculture Organization of the United Nations). 2016. Consumo y
producción de legumbres ha perdido fuerza en América Latina y el Caribe frente
a cultivos más comerciales. www.fao.org/​ameri​cas/​notic​ias/​ver/​es/​c/​455​947/​
(accessed September 12, 2020).
FAO (Food and Agriculture Organization of the United Nations). 2018. Legumbres
pequeñas semillas, grandes soluciones. Ciudad de Panamá. 292 p. ISBN 978-​92-​5-​
131129-​5. www.fao.org/​3/​ca259​7es/​CA259​7ES.pdf.
FAOSTAT (Food and Agriculture Organization of the United Nations). 2021. Food and agri-
culture data. Available at www.fao.org/​faos​tat/​en/​?#data/​QC (accessed December
28, 2020).
FAO/​WHO (Food and Agriculture Organization of the United Nations/​World Health
Organization). 2016. Report joint FAO/​WHO food standards programme. Codex
alimentarius commission. 39th Session. Italy. www.fao.org/​fao-​who-​codexa​
lime​ntar​ius/​sh-​proxy/​en/​?lnk=​1&url=​https%253A%252F%252Fwo​rksp ​ace.fao.
org%252Fsi​tes%252Fco​dex%252FM​eeti​ngs%252​FCX-​718-​48%252​FRep​ort%252FRE​
P16_​PRe.pdf.
Fernández, F., P. Gepts, and M. López. 1984. Etapas de desarrollo de la planta de frijol
común (Phaseolus vulgaris L.). CIAT, Cali, Colombia. ISBN 84-​89206-​54-​6; http://​ciat-​
libr​ary.ciat.cgiar.org/​ciat_​digi​tal/​CIAT/​28093.pdf (accessed December 28, 2020).
Fischer, E. F., and B. Victor. 2014. High-​end coffee and smallholding growers in Guatemala.
Latin American Research Review 49(1): 155–​ 177. https://​doi.org/​10.1353/​
lar.2014.0001.
Flores, D. 2010. Uso Histórico: Sacha Inchi Plukenetia volúbilis L. Proyecto Perubiodiverso,
Perú. https://​repo​sito​rio.promp​eru.gob.pe/​bitstr​eam/​han​dle/​123456​789/​1371/​
Uso_historico_​sacha_​i​nchi​_​201​0_​ke​ywor​d_​pr​inci​pal.pdf ?seque​nce=​1.
Flores, D., and O. Lock. 2013. Revalorizando el uso milenario del sacha inchi (Plukenetia
volubilis L.) para la nutrición, la salud y la cosmética. Revista de Fitoterapia
13(1): 23–​30. www.resea​rchg​ate.net/​prof​i le/​Olga-​Lock/​publ​icat​ion/​271523​
952_ ​ R eva​ l ora​ n do_​ e l_​ u ​ s o_​ m ilenario_​ d el_​ s acha_ ​ i nchi_ ​ P lukenetia_ ​ v olubilis_​
L_ ​ p ara_ ​ l a_ ​ n utricion_​ s alud_​ y _​ c osmetica/​ l inks/​ 5 4cb76460cf2598f7116ed66/​
Revalorando-​el-​uso-​milenario-​d el-​sacha-​inchi-​P lukenetia-​v olubilis-​L-​p ara-​l a-​
nutricion-​salud-​y-​cosmetica.pdf.
40 Characterization of Latin-American Crops

Follegatti-​Romero, L. A., C. A. Piantino, R. Grimaldi, and A.C. Fernando. 2009. Supercritical


CO2 extraction of omega-​ 3 rich oil from Sacha inchi (Plukenetia volubilis L.)
seeds. Journal of Supercritical Fluids 49: 323–​ 329. https://​doi.org/​10.1016/​j.sup​
flu.2009.03.010.
Fritz, G. J., M. C. Bruno, B. S. Langlie, B. D. Smith, and L. Kistler. 2017. Cultigen chenopods
in the Americas: A hemispherical perspective. In Social Perspectives on Ancient Lives
from paleoethnobotanical data, ed. M. P. Sayre, and M. C. Bruno, 55–​75. Springer
International Publishing, Cham, Switzerland. https://​doi.org/​10.1007/​978-​3-​319-​
52849-​6_​3.
Funabashi, M. 2018. Human augmentation of ecosystems: Objectives for food produc-
tion and science by 2045. npj Science of Food 2: 16. https://​doi.org/​10.1038/​s41​
538-​018-​0026-​4.
Gade, D. W. 1970. Ethnobotany of cañihua (Chenopodium-​pallidicaule), rustic seed crop
of altiplano. Economic Botany 24(1): 55–​61. https://​doi.org/​10.1007/​bf0​2860​637.
Gahukar, R. T. 2014. Potential of minor food crops and wild plants for nutritional security
in the Developing World. J Agr Food Inform 15(4): 342−352. https://​doi.org/​10.1080/​
10496​505.2014. 952429.
Galeano, C., M. Nutti, J. Ramírez-​Villegas, et al. 2019. Maize for Colombia: 2030 Vision.
Mexico: Centro Internacional de Mejoramiento de Maíz y Trigo (CIMMYT) and
Centro Internacional de Agricultura Tropical (CIAT). https://​rep​osit​ory.cim​myt.org/​
bitstr​eam/​han​dle/​10883/​20382/​ 61038.pdf ?sequence=​1&isAllowed=​y.
Galek, R., E. Sawicka-​Sienkiewicz, D. Zalewski. 2007. Evaluation of interspecific hybrids of
Andean lupin and their parental forms with regard to some morphological and quan-
titative characters. Fragmenta Agronomica (Poland) 24(2): 81–​87. ISSN: 0860-​4088.
Galek, R., E. Sawicka-​Sienkiewicz, D. Zalewski, S. Stawiński, K. and Spychała. 2017.
Searching for low alkaloid forms in the Andean lupin (Lupinus mutabilis) collection.
Czech Journal of Genetics and Plant Breeding 53: 55–​62. https://​doi.org/​10.17221/​71/​
2016-​CJGPB.
Galluzzi, G., and I. López Noriega. 2014. Conservation and use of genetic resources of
underutilized crops in the Americas—​a continental analysis. Sustainability 6(2):980–​
1017. https://​doi.org/​10.3390/​ su6020980.
Geil, P. B., and J. W. Anderson. 1994. Nutrition and health implications of dry beans: A
review. Journal of the American College of Nutrition 13(6): 549–​558. https://​doi.org/​
10.1080/​07315​724.1994. 10718446.
Gianoli, E., I. Ramos, A. Alfaro-​Tapia, Y. Valdéz, E. R. Echegaray, and E. Yábar. 2006. Benefits
of a maize–​bean–​weeds mixed cropping system in Urubamba Valley, Peruvian
Andes. International Journal of Pest Management 52(4): 283−289. https://​doi.org/​
10.1080/​09670870600796722.
Gillespie, L. J. 1993. A synopsis of Neotropical Plukenetia (Euphorbiaceae) including two
new species. Systematic Botany 18(4): 575–​592. https://​doi.org/​10.2307/​2419​535.
Gillespie, L. J., and W. S. Armbruster. 1997. A contribution to the Guianan Flora:
Dalechampia, Haematostemon, Omphalea, Pera, Plukenetia, and Tragia
(Euphorbiaceae) with notes on subfamily Acolyphoideae. Smithson Contributions to
Botany 86: 1–​48. https://​doi.org/​10.5479/​ si.0081024X.86.
Giuliani, A., F. Hintermann, W. Rojas, et al. 2012. Biodiversity of Andean grains: Balancing
market potential and sustainable livelihoods. Bioversity International, Rome. www.
bioversityinternational.org/​fileadmin/​user_​upload/​online_​library/​publications/​
pdfs/​1635.pdf (accessed November 30, 2020).
References 41

González, X. 2018. Hay sembradas cerca de 500 000 hectareas de maiz en Colombia. Agro
Negocios 12 de septiembre de 2018. www.agron​egoc​ios.co/​agri​cult​ura/​hay-​sembra​
das-​cerca-​de-​500​000-​hectar​eas-​de-​maiz-​en-​colom​bia-​2769​113.
Gorinstein. S., J. Drzewiecki, E. Delgado-​ Licon, et al. 2005. Relationship between
dicotyledone-​ amaranth, quinoa, fagopyrum, soy-​ bean and monocots-​ sorghum
and rice based on protein analyses and their use as substitution of each other.
European Food Research and Technology 221(1–​2): 69–​77. https://​doi.org/​10.1007/​
s00​217-​005-​1208-​2.
Grobman Tversqui, A., W. S. Salhuana, R. Sevilla Panizo, and P. C. Mangelsdorf. 1961. Races
of Maize in Peru, Their Origins, Evolution and Classification. National Academy of
Sciences/​National Research Council, Washington, DC.
Gross, R. 1982. El cultivo y la utilización del Lupinus mutabilis Sweet. FAO, Rome.
Gueguen, J., and P. Cerletti. 1994. Proteins of some legume seeds: Soybean, pea, fababean
and lupin. In New and Developing Sources of Food Proteins, ed. B. J. F. Hudson, 145–​
193. Chapman & Hall, London.
Güémes-​ Vera, N., R. J. Peña-​ Bautista, C. Jiménez-​ Martínez, G. Dávila-​
Ortiz, and G.
Calderón-​Domínguez. 2008. Effective detoxification and decoloration of Lupinus
mutabilis seed derivatives, and effect of these derivatives on bread quality and
acceptance. Journal of the Science of Food and Agriculture 88(7): 1135–​1143. https://​
doi.org/​10.1002/​jsfa.3152.
Guerrero, M. 1987. Algunas propiedades y aplicaciones de los alcaloides del chocho
(Lupinus mutabilis Sweet). In Evento de información y difusión de resultados de
investigación sobre chocho y capacitación en nuevas técnicas de laboratorio, ed. M.
Guerrero, 25–​28. Universidad Técnica de Ambato, Ambato, Ecuador.
Gulisano, A., S. Alves, J. Martins, and L. Trindade. 2019. Genetics and breeding of Lupinus
mutabilis: An emerging protein crop. Frontiers Plant Science. https://​doi.org/​10.3389/​
fpls.2019. 01385.
Gutiérrez, L.-​P., L.-​M. Rosada, and A. Jiménez. 2011. Chemical composition of Sacha Inchi
(Plukenetia volubilis L.) seeds and characteristics of their lipid fraction. Grasas y
aceites 62(1): 76–​83. https://​doi.org/​10.3989/​gya044​510.
Haileslassie, H. A., C. J. Henry, and R. T. Tyler. 2016. Impact of household food processing
strategies on antinutrient (phytate, tannin and polyphenol) contents of chickpeas
(Cicer arietinum L.) and beans (Phaseolus vulgaris L.): A review. International Journal
of Food Science and Technol 51: 1947–​1957. https://​ doi.org/​10.1111/​ijfs.13166.
Hamaker, B. R., C. Valles, R. Gilman, et al. 1992. Amino acid and fatty acid profiles of the
Inca peanut (Plukenetia volubilis). Cereal Chemistry 69(4): 461–​463. www.cereal​sgra​
ins.org/​ publications/​cc/​backissues/​1992/​Documents/​69_​461.pdf.
Hawkes, J. G. 1997. Back to Vavilov: Why were plants domesticated in some areas and not
in others? Paper presented at Conference “Origins of Agriculture and Domestication
of Crop Plants in the Near East”, ICARDA, Aleppo, Syria, 10–​14 May 1997, 1–​3. www.
biov​ersi​tyin​tern​atio​nal.org/​filead​min/​bio​vers​ity/​ publications/​Web_​version/​47/​
ch06.htm.
Hernandez Bermejo, J. E., and J. Leon. 1994. Neglected crops: 1492 from a different perspec-
tive. FAO Plant Production and Protection Series No.26, FAO, Rome. www.fao.org/​ 3/​
t0646e/​t0646e.pdf.
Hernández-​Gómez, J. A., and S. Miranda-​Colín. 2008. Morphological characterization of
chia (Salvia hispanica). Revista Fitotecnia Mexicana 31(2): 105–​113. www.reda​lyc.
org/​artic​ulo.oa? id=​61031203.
42 Characterization of Latin-American Crops

Hernández-​López, V., M. Vargas-​Vasquez, J. Maruaga-​Martínez, S. Hernández-​Delgado,


and N. Mayek-​Pérez. 2013. Origin, domestication and diversification of common
beans: Advances and perspectives. Revista fitotecnia mexicana 32(2): 95–​104. www.sci​
elo.org.mx/​sci​elo.php? script=​sci_​abstract&pid=​S0187-​73802013000200002&lng=​
es&nrm=​iso&tlng=​en.
Hernández-​Pérez, T., M. E. Valverde, and O. Paredes-​López, 2021. Seeds from ancient food
crops with the potential for antiobesity promotion. Critical Reviews in Food Science
and Nutrition. https://​doi.org/​10.1080/​10408​398.2021.1877​107.
Horton, D. 2014. Investigación colaborativa de granos andinos en Ecuador. Fundación
McKnight e Instituto Nacional de Investigaciones Agropecuarias. Quito, Ecuador.
https://​repo​sito​rio.iniap.gob.ec/​ jspui/​bitstream/​41000/​102/​4/​iniapsc315.pdf.
Ichuta, F., and E. Artiaga. 1986. Relación de géneros en la producción y en la Organización
Social en Comunidades de Apharuni, Totoruma, Yauricani-​Ilave. 15−17. BSc thesis in
social science, Universidad Nacional del Altiplano. Puno, Peru.
INEI (Instituto Nacional de Estadisticas e Informatica). 2020. Producción agropecuaria,
según principales productos, 2013−2019. INEI, Peru. www.inei.gob.pe/​media/​
MenuRe​curs​ivo/​publ​icac​ione​s_​di​gita​les/​Est/​Lib1​758/​cap13/​ind13.htm. (accessed
December 18, 2020).
Jacobsen, S.-​E. 1997. Adaptation of quinoa (Chenopodium quinoa) to Northern European
agriculture: Studies on developmental pattern. Euphytica 96: 41−48. https://​doi.org/​
10.1023/​A:100299​2718​009.
Jacobsen, S.-​E. 2017. The scope for adaptation of quinoa in Northern Latitudes of Europe.
Journal of Agronomy and Crop Science 203(6): 603−613. https://​doi.org/​10.1111/​
jac.12228.
Jacobsen, S.-​E., and J. L. Christiansen. 2016. Some agronomic strategies for organic quinoa
(Chenopodium quinoa Willd.). Journal of Agronomy and Crop Science 202(6): 454−463.
https://​doi.org/​10.1111/​jac.12174.
Jacobsen, S.-​E., and A. Mujica. 2006. El tarwi (Lupinus mutabilis Sweet) y sus parientes
silvestres. In Botánica Económica de los Andes Centrales, ed. M. Moraes, B. Øllgaard,
L. P. Kvist, F. Borchsenius, and H. Balslev, 458–​482. Universidad Mayor de San Andrés,
La Paz, Bolivia.
Jacobsen, S.-​E., and A. Mujica. 2008. Geographical distribution of the Andean lupin
(Lupinus mutabilis Sweet). Plant Genetic Resources Newsletter 155: 1–​8. https://​doi.
org/​http://​www2. bioversityinternational.org/​publications/​pgrnewsletter/​article.
asp?lang=​en&id_​article=​1&id_​issue=​155.
Jacobsen, S.-​E., I. Jørgensen, and O. Stølen. 1994. Cultivation of quinoa (Chenopodium
quinoa) under temperate climatic conditions in Denmark. Journal of Agricultural
Science 122: 47–​52. https://​doi.org/​10.1017/​S0021859600065783.
Jacobsen, S.-​E., J. L. Christiansen, and J. Rasmussen, 2010. Weed harrowing and inter-​row
hoeing in organic grown quinoa (Chenopodium quinoa Willd.). Outlook on Agriculture
39: 223–​227. https://​doi.org/​10.5367/​oa.2010.0001.
Jacobsen, S-​ E, M. Sørensen, S. M. Pedersen, and J. Weiner. 2013. Feeding the
world: Genetically modified crops versus agricultural biodiversity. Agronomy
for Sustainable Development 33(4): 651−662. https://​doi.org/​10.1007/​s13​
593-​013-​0138-​9.
Jacobsen, S.-​ E., M. Sørensen, S. M. Pedersen, and J. Weiner. 2015. Using our
agrobiodiversity: Plant-​based solutions to feed the world. Agronomy for Sustainable
Development 35: 1217–​1235. https://​doi.org/​10.1007/​s13593-​015-​0325-​y.
References 43

Jamboonsri, W., T. D. Phillips, R. L. Geneve, J. P. Cahill, and D. F. Hildebrand. 2012. Extending


the range of an ancient crop, Salvia hispanica L. –​a new ω3 source. Genetic Resources
and Crop Evolution 59(2): 171–​178. https://​doi.org/​10.1007/​s10​722-​011-​9673-​x.
Jiménez-​Martínez, C., H. Hernandez, and G. Dávila-​Ortíz. 2003. Lupines: An alternative
for debittering and utilization in foods. In FoodScience and Food Biotechnology,
ed. G. Gutiéerrez-​López, and G. Barbosa-​Cánovas, 233–​252. CRC Press, Boca
Raton, FL.
Katungi, E., A. Farrow, J. Chianu, L. Sperling, and S. E. Beebe. 2009. Common bean in
Eastern and Southern Africa: A situation and outlook analysis. International Centre
for Tropical Agriculture [IITA], Ibadan Nigeria. http://​trop​ical​legu​mes.icri​sat.org/​
wp-​cont​ent/​uplo​ads/​2016/​02/​rso-​com​mon-​bean-​esa.pdf.
Katwal, T. B., and D. Bazile. 2020. First adaptation of quinoa in the Bhutanese moun-
tain agriculture systems. PLoS ONE 15: e0219804. https://​doi.org/​10.1371/​jour​nal.
pone.0219​804.
Khaitov, B., A. A. Karimov, K. Toderich, et al. 2020. Adaptation, grain yield and nutritional
characteristics of quinoa (Chenopodium quinoa) genotypes in marginal environ-
ments of the Aral Sea basin. Journal of Plant Nutrition 44(9): 1365−1379. https://​doi.
org/​10.1080/​01904​167.2020.1862​200.
Kistler, L., S. Y. Maezumi, J. G. de Souza, et al. 2018. Multiproxy evidence highlights a com-
plex evolutionary legacy of maize in South America. Science 362(6420): 1309−1313.
https://​doi.org/​10.1126/​science.aav0207.
Kistler, L., H. B. Thakar, A. M. VanDerwarker, et al. 2020. Archaeological Central American
maize genomes suggest ancient gene flow from South America. PNAS: https://​doi.
org/​10.1073/​pnas.201​5560​117.
Kodahl, N., and M. Sørensen, 2021. Sacha inchi (Plukenetia volubilis L.) –​an underutilized
crop with a great potential. Agronomy 11(6): 1066. https://​doi.org/​10.3390/​agron​
omy1​1061​066.
Kulas, E. ed. 2015. Chicha –​An Andean Idenity − The History and Meaning − Chicha’s
Evolution and the Inca Empire. The Ohio State University. https://​u.osu.edu/​chi​cha/​
chi​cha-​and-​the-​inca-​emp​ire/​.
Kwak, M., and P. Gepts, 2009. Structure of genetic diversity in the two major gene pools
of common bean (Phaseolus vulgaris L., Fabaceae). Theoretical and Applied Genetics
118: 979–​992. https://​ doi.org/​10.1007/​s00122-​008-​0955-​4.
Leidi, E.O., A. Monteros, G. Mercado, et al. 2018. Andean roots and tubers crops as sources
of functional foods. Journal of Functional Foods 51: 86−93. https://​doi.org/​10.1016/​
j.jff.2018.10.007.
Li, X., and K. H. M. Siddique. 2018. Future Smart Food Rediscovering hidden treasures of
neglected and underutilized species for Zero Hunger in Asia, Executive summary, Food
and Agriculture Organization of the United Nations, Bangkok. www.fao.org/​3/​I913​
6EN/​ i9136en.pdf.
Lobo Zavalia, R., M. G. Alcocer, F. J. Fuentes, W. A. Rodriguez, M. Morandini, and M. R.
Devani. 2011. Desarrollo del cultivo de chia en Tucuman, Republica Argentina.
EEAOC Advance Agroindustrial 32(4): 27–​ 30. www.eeaoc.gob.ar/​?publ​icac​ion=​
32-​4-​3.
López, A. X., A. G. Huerta, E. C. Torrez, D. M. Sangerman-​Jarquín, G. O. Rosas, and M. R.
Arriaga, 2017. Chia (Salvia hispanica L.) current situation and future trends. Revista
Mexicana de Ciencias Agrícolas 8(7): 1619–​1631. www.sci​elo.org.mx/​sci​elo.php?scr​
ipt=​sci_​i​ssue​toc&pid=​ 2007093420170007&lng=​en&nrm=​iso.
44 Characterization of Latin-American Crops

López-​Marqués, R. L., A. F. Nørrevang, P. Ache, et al. 2020. Prospects for the accelerated
improvement of the resilient crop quinoa. Journal of Experimental Botany
71(18): 5333−5347. https://​doi.org/​10.1093/​jxb/​eraa285.
Lucas, M. M., F. Stoddard, P. Annicchiarico, et al. 2015. The future of lupin as a protein crop
in Europe. Frontiers in Plant Science 6. https://​doi.org/​10.3389/​fpls.2015.00705.
Lundmark, C. 2007. Genetically modified maize. BioScience 57(11): 996. https://​doi.org/​
10.1641/​ B571115.
MAGA (Ministerio de Agricultura, Ganadería y Alimentación de Guatemala). 2017.
Informe situación del frijol a diciembre 2017: Consumo aparente. Ciudad de
Guatemala, Guatemala. www. maga.gob.gt/​sitios/​diplan/​download/​informacion_​
del_​sector/​informes_​d e_​situacion_​d e_​maiz_​y_​frijol/​2017/​12%20Informe%20
Situaci%C3%B3n%20Del%20Frijol%20Negro%20Diciembre%202017.pdf.
Mamani, E. 2013. Caracterización molecular de 26 accesiones de cañihua (Chenopodium
pallidicaule Aellen) con mayor rendimiento en grano: Altiplano –​Puno. PhD thesis.
Universidad Nacional del Altiplano, Puno, Peru. http://​repo​sito​rio.unap.edu.pe/​han​
dle/​UNAP/​252 (accessed October 17, 2020).
Mamani, F. 2016a. Atlas de biodiversidad genética de cañahua y quinua. ISBN: 978-​99974-​
65-​92-​4. Diseño & Impresiones FLORES. La Paz, Bolivia.
Mamani, F. 2016b. Cultivo de cañahua (Chenopodium pallidicaule Aellen) para la seguridad
alimentaria. ISBN: 978-​ 99974-​65-​
90-​0. Diseño & Impresiones FLORES. La Paz,
Bolivia.
Mamani, F., S. E. Aliaga, A. Bonifacio, D. Torrico, and N. Tapia. 2018. La cañahua grano
milenario de los Andes. Arte dedicado a la producción sostenible. ISBN: 978-​99974-​0-​
353-​7. Diseño & Impresiones FLORES. La Paz, Bolivia.
Mamidi, S., M. Rossi, D. Annam, et al. 2011. Investigation of the domestication of common
bean (Phaseolus vulgaris) using multilocus sequence data. Functional Plant Biology
38: 953–​967. https://​doi.org/​10.1071/​fp11​124.
Mangelsdorf, P. C. 1961. Introgression in maize. Euphytica 10(2): 157−168. https://​doi.org/​
10.1007/​BF0​0022​207.
Mangelson, H., D. E. Jarvis, P. Mollinedo, et al. 2019. The genome of Chenopodium
pallidicaule: An emerging Andean super grain. Applications in Plant Sciences
7(11): e11300. https://​doi.org/​10.1002/​ aps3.11300.
Márquez, C. 2016. La siembra de chocho es más rentable. Revista Líderes. www.
revistalideres.ec/​lideres/​siembra-​chocho-​produccion-​chimborazo.html (accessed
September 12, 2020).
Martín-​Cabrejas, M. A., B. Sanfiz, A. Vidal, E. Mollá, R. Esteban, F. J. López-​Andréu. 2004.
Effect of fermentation and autoclaving on dietary fiber fractions and antinutritional
factors of beans (Phaseolus vulgaris L.). Journal of Agricultural and Food Chemistry
52: 261–​266. https://​doi.org/​10.1021/​ jf034980t.
Maurer, N. E., B. Hatta-​Sakoda, G. Pascual-​Chagman, and L. E. Rodriguez-​Saona. 2012.
Characterization and authentication of a novel vegetable source of omega-​3 fatty
acids, sacha inchi (Plukenetia volubilis L.) oil. Food Chemistry 134: 1173–​1180.
https://​doi.org/​10.1016/​j.foodc​hem.
Maydana, E. 2010. Evaluacion de la produccion de seis variedades de cañahua (Chenopodium
pallidicaule Aellen) con participacion de agricultores en la comunidad de Pacaure
del Municipio de Mocomoco. PhD thesis, Universidad Mayor de San Andres, La Paz,
Bolivia.
References 45

Mazón, N. 2018. El chocho o tarwi como recurso genético de la región andina [Seminario
online]. Quito, Ecuador: Interaprendizaje –​IPDRS. https://​bit.ly/​2rsb​fFF (accessed
September 12, 2020).
Meneses, R. 1996. Las leguminosas en la Agricultura Boliviana. Proyecto Rhizobiología
Bolivia. CIAT-​CIF-​PNLG-​CIFP-​WALL. Cochabamba, Bolivia. pp. 209–​225.
Mercado, G., J. Davalos, IPDRS, Hivos, and Cipca. 2018. Memoria foro virtual: Los caminos
del tarwi y la integración andina: Bolivia, Perú y Ecuador. Bolivia: IPDRS. https://​inter​
apre​ndiz​aje. ipdrs.org/​images/​Documentos2018/​MEMORIACAMINOSDELTARWI_​
14.12.18.pdf (accessed September 12, 2020).
Michail, N. 2019. Colombia inks US$1 million deal for climate-​ smart corn. Food
navigator-​ latam.com on 18 Feb. 2019. www.foodna​viga​tor-​latam.com/​Arti​cle/​
2019/​02/​18/​Maize-​for-​Colom​bia-​part​n ers​hip-​to-​deve​lop-​clim​ate-​cha​nge-​resist​
ant-​corn.
Miklas, P. N., J. D. Kelly, S. E. Beebe, and M. W. Blair. 2006. Common bean breeding for resist-
ance against biotic and abiotic stresses: From classical to MAS breeding. Euphytica
147: 105–​131. https://​doi.org/​10.1007/​s10681-​006-​4600-​5.
Mina, D., Q. Struelens, A. Barragán, and O. Dangles. 2018. El chocho. Un superalimento que
podría convertirse en una amenaza. Revista Nuestra Ciencia 20: 19–​21. www.puce.
edu.ec/​ portal/​wp-​content/​uploads/​2019/​07/​Nuestra-​Ciencia-​n.%C2%BA-​20.pdf.
Miranda, F. 2014. Guía tecnica para el manejo del cultivo de la Chia (Salvia hispánica)
en Nicaragua. GuiagroNicaragua 3: 130–​131 [16 slides]. https://​es.sli​desh​are.net/​
fpmirandasalgado/​manual-​de-​produccion-​de-​chia-​salvia-​hispanica-​40722325.
Miranda-​Colin, S. 1978. Evolución de cultivares nativos de México. Ciencia y Desarrollo
3: 130–​131.
Morais Dias, D., N. Kolba, D. Binyamin, O. Ziv, M. R. Nutti, H. S. Duarte Martino, R. P.
Glahn, O. Koren, and E. Tako, 2018. Iron biofortified Carioca Bean (Phaseolus
vulgaris L.)-​based Brazilian diet delivers more absorbable iron and affects the
gut microbiota in vivo (Gallus gallus). Nutrients 10(12): 1970. http://​dx.doi.org/​
10.3390/​nu1​0121​970.
Moscoso-​Mujica, G., A. Zavaleta, Á. Mujica, M. Santos, and R. Calixto. 2017. Fractionation
and electrophoretic characterization of (Chenopodium pallidicaule Aellen) kanihua
seed proteins. Revista chilena de nutrición 44(2): 144−152. http://​dx.doi.org/​10.4067/​
S0717-​751820​1700​0200​005.
Moure, A., J. Sineiro, H. Domínguez, and J. C. Parajó. 2006. Functionality of oilseed pro-
tein products: A review. Food Research International 39: 945–​963. https://​doi.org/​
10.1016/​j.food​res.2006.07.002.
Mujica, A., and S.-​ E. Jacobsen. 1998. Agrobiodiversidad de las Aynokas de Quinua
(Chenopodium quinoa Willd.) y la seguridad Alimentaria. In Seminario
Agrobiodiversidad en la Región Andina y Amazónica. Lima, 23−24 Nov. 28−29. CCIIA,
UNALM, CLADES y RAE. Lima, Peru.
Mujica, A., and S.-​ E. Jacobsen. 2000. Agrobiodiversidad de las Aynokas de quinua
(Chenopodium quinoa Willd.) y la seguridad Alimentaria. In Proc. Seminario Taller
Agrobiodiversidad en la región andina y amazónica. 23−25 noviembre 1988, ed. C.
Felipe-​Morales, and A. Manrique, 151−156. NGO-​CGIAR. Lima, Peru.
Muñoz, L. A., A. Cobos, O. Diaz, and J. M. Aguilera. 2013. Chia seed (Salvia hispanica): An
ancient grain and a new functional food. Food Reviews International 29(4): 394–​408.
https://​doi.org/​10.1080/​87559129.2013.818014.
46 Characterization of Latin-American Crops

Muziri, T., P. Chaibva, A. Chofamba, et al. 2020. Using principal component analysis to
explore consumers’ perception toward quinoa health and nutritional claims in
Gweru, Zimbabwe. Journal of Food Science and Nutrition 9(2): 1025−1033. https://​
doi.org/​10.1002/​fsn3.2071.
Nájera, S. 2015. ¿Tiene la producción de chochos (L. mutabilis) el potencial de aumentar
los ingresos agrícolas y contribuir a la seguridad alimentaria del Ecuador si es que
sustituye a la producción de soya o maíz? MSc thesis, School of Geociences. University
of Edinburgh.
Nanduri, K. R., A. Hirich, M. Salehi, S. Saadat, and S.-​E. Jacobsen. 2019. Quinoa: A new crop
for harsh environments. In Sabkha Ecosystems. Tasks for Vegetation Science, vol 49,
ed. B. Gul, B. Böer, M. Khan, M. Clüsener-​Godt, and A. Hameed, 301−333. Springer,
Cham. https://​doi.org/​10.1007/​978-​3-​030-​04417-​6_​19.
Norlaily, M. A., K. Y. Swee, Y. H. Wan, K. Boon, W. T. Sheau, and G. T. Soon. 2012. The prom-
ising future of chia, Salvia hispanica L. Journal of Biomedicine and Biotechnology
2012(171956): 1–​9. https://​doi.org/​10.1155/​2012/​171​956.
OECD (Organisation for Economic Co-​Operation and Development). 2021. Chapter 1.
Common Bean (Phaseolus vulgaris). In Novel Food and Feed Safety, Safety Assessment
of Foods and Feeds Derived from Transgenic Crops, Volume 3 Common bean, Rice,
Cowpea and Apple Compositional Considerations. www.oecd-​ilibr​ary.org/​sites/​544d0​
3d6-​en/​index.html?ite​mId=​/​cont​ent/​ component/​544d03d6-​en.
Orona-​Tamayo, D., M. E. Valverde, and Octavio Paredes-​López, 2018. Bioactive peptides
from selected Latin American food crops –​A nutraceutical and molecular approach.
Critical Reviews in Food Science and Nutrition. https://​doi.org/​10.1080/​10408​
398.2018.1434​480.
Pascual-​Villalobos, M., E. Correal, E. Molina, and J. Martínez. 1997. Evaluación y selección
de especies vegetales productoras de compuestos naturales con actividad insecticida.
Centro de Investigación y Desarrollo Agroalimentario (CIDA), Murcia, Spain. www.
inia.es/​sites/​fro​ntbo​otst​rap/​Pages/​PageNo​tFou​ndEr​ror.aspx?req​u est​Url=​www.
inia.es/​gcont​rec/​proyec​tos/​res​ulta​dos-​97/​agric​ola/​sc94-​039.pdf.
Paz Silva, L. J. 1993. Philosophy for the development of Andean Ecosystems. In Andean
Agro Ecosystem: Problems, Limitations and Perspectives. Annals of the international
workshop on Andean agro ecosystem, 30 March–​2 April, Lima, Peru, pp. 11−29.
https://​agris.fao.org/​agris-​sea​rch/​ search.do?recordID=​QP9300027.
Pearsall, D. 1992. The origins of plant cultivation in South America. In The Origins of
Agriculture. An International Perspective, Eds. C. Wesley Cowan & P. Jo Watson ,173–​
205). Washington, DC/​London: Smithsonian Institution Press
Peiretti, P. G., and F. Gai, 2009. Fatty acid and nutritive quality of chia (Salvia hispanica L.)
seeds and plant during growth. Animal Feed Science and Technology 148(2–​4): 267–​
275. https://​doi.org/​10.1016/​j.anifeedsci.2008.04.006.
Peláez, P., D. Orona-​Tamayo, S. Montes-​Hernández, M. Elena Valverde, O. Paredes-​López,
and A. Cibrián-​Jaramillo, 2019. Comparative transcriptome analysis of cultivated
and wild seeds of Salvia hispanica (chia). Scientific Reports 9: 9761. https://​doi.org/​
10.1038/​s41​598-​019-​45895-​5.
Peralta, E., and E. Villacrés, 2015. 100 recetas prácticas usando quinua, chocho y
amaranto. Publicación miscelánea Nº 421. Programa Nacional de Leguminosas y
Granos Andinos y Departamento de Nutrición y Calidad. Estación Experimental
References 47

Santa Catalina. Instituto Nacional de Investigaciones Agropecuarias (INIAP).


Quito, Ecuador. https://​repo​sito​rio.iniap.gob.ec/​ bitstream/​41000/​2727/​1/​
iniapscpm421.pdf.
Pérez Suárez, T. 1997. El dios del maíz en Mesoamérica. Arqueología Mexicana
5(25): 44–​55.
Pisfil, C. A. 2017. Optimización del nivel de sustitución de la harina de trigo por harina de
quinua, cañihua y kiwicha en la elaboración de pan panini precocido. PhD thesis,
Universidad Nacional Pedro Ruiz Gallo, Peru. http://​repo​sito​rio.unprg.edu.pe/​han​
dle/​UNPRG/​1301 (accessed 20 October, 2020).
Przybylak, J. K., D. Ciesiolka, W. Wysocka, et al. 2005. Alkaloid profiles of Mexican wild lupin
and an effect of alkaloid preparation from Lupinus exaltatus seeds on growth and
yield of paprika (Capsicum annuum L.). Industrial Crops Products 21: 1–​7. https://​doi.
org/​10.1016/​j.indcrop.2003.12.001.
Quispe, E. 2004. Morfología y variabilidad de las cañahuas cultivadas (Chenopodium
pallidicaule Aellen). Ing. Agr. thesis, Universidad Mayor de San Simon, Cochabamba,
Bolivia.
Ramirez, S. 2016. Amido e farinha de cañihua (Chenopodium pallidicaule): extração,
caracterização e desenvolvimento de filmes biodegradáveis. Master’s disserta-
tion, Faculdade de Zootecnia e Engenharia de Alimentos, University of São Paulo,
Pirassununga, Brazil. https://​doi.org/​10.11606/​D.74.2016.tde-​12082​016-​104​942
(accessed November 30, 2020).
Rathna Priya, T. S., and A. Manickavasagan. 2020. Common Bean. In Pulses, ed. A.
Manickavasagan, and P. Thirunathan, 77−97. Springer International, Cham. https://​
doi.org/​10.1007/​978-​3-​030-​41376-​7_​5.
Razzaghi, F., M. R. Bahadori-​Ghasroldashti, S. Henriksen, A. R. Sepaskhah, and S.-​E.
Jacobsen. 2020. Physiological characteristics and irrigation water productivity of
quinoa (Chenopodium quinoa Willd.) in response to deficit irrigation imposed at
different growing stages − A field study from Southern Iran. Journal of Agronomy and
Crop Science 206(3): 390−404. https://​doi.org/​10.1111/​jac.12392.
Repo-​Carrasco, R., C. Espinoza, and S.-​E. Jacobsen. 2003. Nutritional value and use
of the Andean crops quinoa (Chenopodium quinoa) and kañiwa (Chenopodium
pallidicaule). Food Reviews International 19(1–​2): 179–​189. https://​doi.org/​10.1081/​
fri-​120018​884.
Repo-​Carrasco-​Valencia, R. 2011. Andean indigenous food crops: Nutritional value and
bioactive compounds. PhD thesis, Department of Biochemistry and Food Chemistry,
University of Turku, Finland. www.utu​pub.fi/​bitstr​eam/​han​dle/​10024/​74762/​Repo-​
Carra​sco-​Valen​cia-​Diss2​011.pdf ? sequence=​1.
Rivera-​Cabrera, F., M. Medina-​Valdez, C. Pelayo-​Zaldívar, et al. 2017. Phytochemical com-
position of Salvia hispanica L. extracts and their satiety effect. Revista Mexicana de
Ingeniería Química 6(1): 47–​53. www.reda​lyc.org/​pdf/​620/​6204​9878​006.pdf.
Roberts, L. M., U. J. Grant, R. Ramirez E., W. H. Hatheway, and D. L. Smith. 1957. Races of
Maize in Colombia. NAS/​NRC, 510, Washington, DC www.ars.usda.gov/​ARSUs​erFi​
les/​50301​000/​Races_​of_​Maize/​RoM_​Colombia_​0_​Book.pdf.
Rodriguez, J. P., and F. Mamani, 2016. Participatory breeding and gender role in cañahua, an
NUS Andean Grain crop in Bolivia. In Gender, Breeding and Genomics –​Case Studies.
Workshop held 18–​21 October 2016, CGIAR Gender and Agriculture Research
48 Characterization of Latin-American Crops

Network, Nairobi. www.resea​rchg​ate.net/​publ​icat​ion/​323053515_​Participatory_​


breeding_​and_​gender_​role_​in_​canahua_​a_​NUS_​A​ndea​n_​Gr​ain_​crop​_​in_​Boli​via.
Rodriguez, J. P., E. Maydana, R. Flores, N. Calancha, M. Flores, and F. Mamani. 2010. Seguridad
Alimentaria y revalorización de la Cañihua (Chenopodium pallidicaule): Caso de
estudio en comunidades del Altiplano Norte de Bolivia. In Memorias 3er Congreso
Mundial de la Quinua. FCAPV-​UTO, Oruro, Bolivia.
Rodriguez, J. P., M. Aro, M. Coarite, et al. 2017. Seed shattering of Cañahua (Chenopodium
pallidicaule Aellen). Journal of Agronomy and Crop Science 203(3): 254–​267. https://​
doi.org/​10.1111/​jac.12192.
Rodriguez, J. P., S.-​ E. Jacobsen, M. Sørensen, and C. Andreasen. 2020. Cañahua
(Chenopodium pallidicaule): a promising new crop for arid areas. In Emerging
Research in Alternative Crops, Environment & Policy 58 ed. A. Hirich, R. Ragab, R.
Choukr-​Allah et al., chap. 9, 221–​243. Springer International, Cham. https://​doi.org/​
10.1007/​978-​3-​319-​90472-​6_​9.
Rodríguez, J. P., H. Rahman, S. Thushar, and R. K. Singh, 2020. Healthy and resilient cereals
and pseudo-​cereals for marginal agriculture: Molecular advances for improving
nutrient bioavailability. Frontiers in Genetics. February 27, 2020. https://​doi.org/​
10.3389/​fgene.2020.00049.
Rodríguez-​Castillo, L., and X. E. Fernández-​Rojas, 2003. Los frijoles (Phaseolus vulgaris): Su
aporte a la dieta del costarricense. Acta Médica Costarricense 45(3): 120–​125. http://​
actamedica.medicos.cr/​.index.php/​Acta_​Medica/​article/​view/​110/​93
Rodríguez del-​Castillo, A.M., G. Gonzalez-​Aspajo, M. F. Sánchez-​Márquez, and N. Kodahl.
2019. Ethnobotanical knowledge in the Peruvian Amazon of the neglected and
underutilized crop Sacha Inchi (Plukenetia volubilis L.). Economic Botany 73(2): 281–​
287. https://​doi.org/​10.1007/​s12231-​019-​09459-​y.
Rojas, W., J. L. Soto, M. Pinto, et al. 2010. Granos andinos: avances, logros y experiencias
desarrolladas en quinua, cañahua y amaranto en Bolivia. Bioversity International,
Roma.
Romeo, F. V., S. Fabroni, G. Ballistreri, S. Muccilli, A. Spina, and P. Rapisarda. 2018.
Characterization and antimicrobial activity of alkaloid extracts from seeds of
different genotypes of Lupinus spp. Sustainability 10(3): 788. https://​doi.org/​10.3390/​
su1​0030​788.
Rosas, J. C. 2003. El cultivo de frijol común en América Tropical. Carrera de Ciencia y
Producción Agropecuaria. Escuela Agrícola Panamericana/​Zamorano. Honduras.
https://​bdigital.zamorano. edu/​bitstream/​11036/​2424/​1/​prueba%2009.pdf.
Rossi, M., E. Bitocchi, E. Bellucci, et al. 2009. Linkage disequilibrium and population struc-
ture in wild and domesticated populations of Phaseolus vulgaris L. Evolutionary
Applications 2: 504–​522. https://​ doi.org/​10.1111/​j.1752-​4571.2009.00082.x.
Ruales, J., P. Pólit, and B. M. Nair. 1988. Nutritional quality of blended foods of rice, soy and
lupins, processed by extrusion. Food Chemistry 29: 309–​321. https://​doi.org/​10.1016/​
0308-​8146(88)90046-​5.
Salas, L. M. 2017. Produção e caracterização de filmes biodegradáveis a base do
pseudocereal canihua (Chenopodium pallidicaule). PhD dissertation, Universidade
Estadual de Campinas, Campinas, SP. www.repo​sito​rio.unic​amp.br/​han​dle/​REPO​
SIP/​330​342 (accessed November 24, 2020).
Sathe, S. K., S. S. Deshpande, and D. K. Salunkhe. 1981. Functional properties of lupin
seeds (Lupinus mutabilis) proteins and protein concentrates. Journal of Food Science
47: 491–​502. https://​doi.org/​10.1111/​j.1365-​2621.1982.tb10110.x.
References 49

Secretaría de Economía. 2012. Análisis de la cadena de valor del frijol. Dirección General
de Indrustrias Básicas, Estados Unidos Mexicanos. www.econo​mia.gob.mx/​files/​
comunidad_​ n egocios/​ i ndustria_​ c omercio/​ a nalisis_ ​ c adena_ ​ v alor_ ​ f rijol.pdf
(accessed December 28, 2020).
Sellami, M. H., C. Pulvento, and A. Lavini. 2021. Agronomic practices and performances of
quinoa under field conditions: A systematic review. Plants 10(1): 72. https://​doi.org/​
10.3390/​plants10010072.
Serrano, R. 2012. Distribución de la diversidad genética y etnobotánica de cañahua
(Chenopodium pallidicaule Aellen) en las comunidades del Altiplano Norte.
PhD thesis. Facultad de Agronomía, Universidad Mayor de San Andres, La Paz,
Bolivia. https://​repo​sito​rio.umsa.bo/​bitstr​eam/​han​dle/​ 123456789/​8002/​T-​1652.
pdf ?sequence=​1&isAllowed=​y.
Singh, S. P., R. Nodari, and P. Gepts. 1991. Genetic diversity in cultivated common
bean: I. Allozymes. Crop Science 31(1): 19–​23. https://​doi.org/​10.2135/​crop​sci1​991.
0011183X003100010004x.
Skarbø, K., and K. van der Molen. 2016. Maize migration: Key crop expands to higher
altitudes under climate change in the Andes. Climate and Development 8(3): 245−255.
https://​doi.org/​10.1080/​17565529.2015.1034234.
Sosa-​Baldivia, A., G. Ruiz-​Ibarra, R. R. Robles de la Torre, R. Robles Lopez, and A. Montufar
Lopez, 2018. The chia (Salvia hispanica): Past, present and future of an ancient
Mexican crop. Australian Journal of Crop Science 12(10): 1626−1632. https://doi.
org/10.21475/​ajcs.18.12.10.p1202.
Spehar, C. R., and R. L. D. Santos. 2002. Quinoa BRS Piabiru: Alternative for diversification
of cropping systems. Pesquisa Agropecuária Brasileira 37(6): 889–​893. https://​doi.
org/​10.1590/​s0100-​204x20​0200​0600​020.
Srichamnong, W., P. Ting, P. Pitchakarn, O. Nuchuchua, and P. Temviriyanukul. 2018. Safety
assessment of Plukenetia volubilis (Inca peanut) seeds, leaves, and their products.
Food Science and Nutrition 6(4): 962−969. https://​doi.org/​10.1002/​fsn3.633.
Staller, J. E. 2010. Maize cobs and cultures: History of Zea mays L. Springer, Berlin. https://​
doi.org/​10.1007/​978-​3-​642-​04506-​6.
Staller, J. E. 2016. High altitude maize (Zea mays L.) cultivation and endemism in the Lake
Titicaca Basin. Journal of Botany Research 1(1): 8−21. https://​doi.org/​10.36959/​
771/​556.
Stokes, P., and P. Rowley-​Conwy. 2002. Iron Age Cultigen? Experimental return rates for
fat hen (Chenopodium album L.). Environmental Archaeology 7(1): 95−99. https://​doi.
org/​10.1179/​env.2002.7.1.95.
Strozycki, P. M., A. Szczurek, B. Lotocka, M. Figlerowicz, and A. B. Legocki. 2007. Ferritins
and nodulation in Lupinus luteus: Iron management in indeterminate type
nodules. Journal of Experimental Botany 58: 3145–​3153. https://​doi.org/​10.1093/​
jxb/​erm152.
Švec, I., M. Hruskova, R. Kapacinskaite, and T. Hofmanova. 2019. Effect of quinoa
(Chenopodium quinoa) and cañahua wholemeals (Chenopodium pallidicaule) on
pasting behaviour of wheat flour. Advances in Food Science and Engineering 3(1): 1−8.
https://​dx.doi.org/​10.22606/​afse.2019.3100.
Tapia, M. 1982. Proceso agroindustrial del tarwi (Lupinus mutabilis). In Actas de la
Conferencia Internacional del Lupino, 58−62. Asociación Internacional del Lupinu.
Torremolinos, Spain. Servicio de Publicaciones Agrarias, Ministerio de Agricultura,
Pesca y Almentación, Madrid, Spain.
50 Characterization of Latin-American Crops

Tapia, M. E. 2015. El tarwi, lupino Andino. Tarwi, tauri o chocho. Corporación gráfica
Universal, Lima, Peru. http://​fadv​amer​ica.org/​wp-​cont​ent/​uplo​ads/​2017/​04/​
TARWI-​espa​nol.pdf.
Tapia, M. E., and A. M. Fries, 2007. Guía de campo de los cultivos andinos. FAO, ANPE-​
Peru. https://​runama​qui.fr/​wp-​cont​ent/​uplo​ads/​2020/​07/​FAO-​Los-​culti​vos-​andi​
nos-​docume​nto-​compl​eto.pdf.
Tapia, M. E., S.A. Mujica, and A. Canahua. 1980. Origen, distribución geográfica y sistemas
de producción de la quinua. In I Reunión sobre genética y fitomejoramiento de la
quinua, A1–​A8. PISCA-​UNTA-​IBTA-​IICA-​CIID, Publicacion − Universidad Nacional
Tecnica del Altiplano (Peru). Puno, Peru.
Tiwari, B. K., and N. Singh. 2012. Pulse Chemistry and Technology. Royal Society of
Chemistry, Cambridge, UK. ISBN 978-​1-​84973-​331-​1.
Triana-​Maldonado, D. M., S. A. Torijano-​ Gutiérrez, and C. Giraldo-​ Estrada. 2017.
Supercritical CO2 extraction of oil and omega-​3 concentrate from Sacha inchi
(Plukenetia volubilis L.) from Antioquia, Colombia. Grasas y Aceites 68(1): 1−11.
https://​doi.org/​10.3989/​gya.0786​161.
Valdivia-​López, M. Á., and A. Tecante, 2015. Chia (Salvia hispanica): A review of native
Mexican seed and its nutritional and functional properties. Advances in Food and
Nutrition Research 75: 53−75. https://​doi.org/​10.1016/​bs.afnr.2015.06.002.
Vašek, J., P. H. Čepková, I. Viehmannová, M. Ocelák, D. Cachique Huansi, and P. Vejl. 2017.
Dealing with AFLP genotyping errors to reveal genetic structure in Plukenetia
volubilis (Euphorbiaceae) in the Peruvian Amazon. PLoS ONE 12(9): e0184259.
https://​doi.org/​10.1371/​ journal.pone.0184259.
Vavilov, N.I. 1935. Theoretical basis for plant breeding. Vol. 1. Moscow. Origin and
Geography of Cultivated Plants. In The Phytogeographical Basis for Plant Breeding,
ed. Y. A. Ovchinikov (trans. D. Love). Vol. 1, 316−366. Cambridge University Press,
Cambridge, UK.
Villacrés, E., C. Caicedo, and E. Peralta. 2000. Diagnóstico del procesamiento artesanal,
comercialización y consumo del chocho. In Zonificación Potencial, Sistemas de
Producción y Procesamiento Artesanal del Chocho (Lupinus mutabilis Sweet) en
Ecuador, 24–​41. Quito, Ecuador.
Villacrés, E., E. Peralta, and M. Alvarez. 2003. Chochos en su punto. In Chochos. Recetarios.
Disfrute Cocinando con Chochos, ed. E. Peralta, Publicación Miscelánea 118: 1–​53.
INIAPFUNDACYT, Quito, Ecuador.
Villarino, E. 2018. Here’s how to do bean breeding the climate-​smart way. https://​bigd​
ata.cgiar.org/​heres-​how-​to-​do-​bean-​breed​ing-​the-​clim​ate-​smart-​way/​ (accessed
December 28, 2020).
Williams, J. T. ed. 1993. Underutilized Crops: Pulses and Vegetables. Chapman and Hall,
London.
Wink, M. 1991. Plant breeding: Low or high alkaloid content. In Proceedings of the 6th
International Lupin Conference, 25–​30 Nov 1990, ed. D. von Bayer, 326–​334. Temuco,
Pucón, Chile.
Xiu-​shi, Y., Q. Pei-​you, G. Hui-​min, and R. Gui-​xing. 2019. Quinoa industry development
in China. Ciencia e Investigación Agraria 46: 208−219. https://​doi.org/​10.7764/​rcia.
v46i2.2157.
References 51

Yanapa Velasquez, L. L. 2018. Elaboración de biopelículas para envasado de alimentos a


partir de quitosano y cañihua (Chenopodium pallidicaule). Thesis. Universidad
Nacional del Altiplano. Puno, Peru. http://​repo​sito​rio.unap.edu.pe/​han​dle/​UNAP/​
11044 (accessed December 5, 2020).
Zegarra, S. I. 2018. Elaboración de un pan apto para celiacos a base de harina de
Chenopodium pallidicaule y evaluación de su aceptabilidad sensorial. PhD disserta-
tion, Universidad San Ignacio de Loyola, Peru. http://​dx.doi.org/​10.20511/​USIL.the​
sis/​3023 (accessed December 3, 2020).
Zielinska-​Dawidziak, M. 2015. Plant ferritin–​a source of iron to prevent its deficiency.
Nutrients 7(2): 1184–​1201. https://​doi.org/​10.3390/​nu7021​184.
Chapter 2

Genotype and Environment as


Key Factors Controlling Seed
Quality in Latin-​American Crops
Luisa Bascuñán-​Godoy1, María Reguera2, Ángel Mujica3,
Néstor Fernández del Saz1, Carolina Sanhueza1,
Catalina Castro1, José Ortiz1, Gabriel Barros1, José
Delatorre-Herrera4, Karina B. Ruiz5, Teodoro Coba de la Peña6,
Enrique Ostria7, Enrique Martínez7, and Susana Fischer8
1Universidad de Concepción, Concepción, Chile.
2Universidad Autónoma de Madrid, Madrid, Spain.
3Universidad Nacional Altiplano, Puno, Peru.
4Universidad Arturo Prat, Iquique, Chile.
5Universidad Arturo Prat, Iquique, Chile.
6Centro de Estudios Avanzados en Zonas

Áridas (CEAZA), La Serena, Chile.


7Centro de Estudios Avanzados en Zonas

Áridas (CEAZA), La Serena, Chile.


8Universidad de Concepción, Concepción, Chile.

CONTENTS
2.1 Latin American Crops 54
2.2 Major Abiotic Stresses Affecting Seed Quality 55
2.2.1 Drought 56
2.2.2 High Temperatures 56
2.2.3 Salinity 57
2.3 Genotype and Environment Interaction on Crop Productivity
and Seed Quality 58
2.3.1 Chenopodium pallidicaule and Chenopodium quinoa 58
2.3.2 Amaranthus ssp. 63

DOI: 10.1201/9781003088424-2 53
54 Environmental Controls over Seed Quality

2.3.3 Phaseolus vulgaris 65


2.3.4 Zea mays 68
2.3.5 Salvia hispanica 70
2.3.6 Plukenetia volubilis 73
2.3.7 Lupinus mutabilis 73
2.4 Perspectives of Crop Production under Stress Conditions 76
Acknowledgments 76
References 76

2.1 LATIN AMERICAN CROPS


Plants respond to adverse biotic and abiotic conditions within their local envir-
onment, whereby they use their remarkable ability to adjust their physiology to
cope with and acclimate to changing growth conditions.
Climate change threatens agriculture, increasing the loss of soil and redu-
cing areas suitable for crop production systems; by 2050 it will be necessary to
increase plant productivity by 70% to satisfy the needs of an ever-​increasing
world population (FAO 2019). Crop yield is largely determined by climate
conditions, therefore minor deviations from optimal conditions can strongly
penalize yield potential in terms of quantity, but also quality. Considering
the agricultural challenges we face for the next coming decades, a deeper
understanding of the effect of environmental factors on crop growth and devel-
opment could significantly reduce yield losses and improve quality, ensuring
food security worldwide.
The development of new varieties with high and stable genetic potential for
high density crop production is the main challenge for breeding and modern
agronomy. However, there are contrasting views to accomplish this goal. It
has been highlighted that modern cultivars usually achieve higher yields with
plenty of soil nutrients and optimal climate conditions. However, there are sev-
eral reports indicating a better performance of landraces (genotypes adapted to
local regions) compared to modern genetic materials, which generally may pre-
sent low to medium yields. It has been accepted that the main contributions of
landraces are traits related to adaptations to stressful environments, which can
be used in plant breeding programs (Dwivedi et al. 2016).
Since the 1960s, there has been plenty of evidence of a slowdown in yield
improvement rates of major food crops (including rice, wheat and maize) (FAO
2016), and current yield strategies are not enough to meet future requirements
(Lenné and Wood 2011). Considering that climate change will affect the poten-
tial of traditional crops, searching for ancestral crops and landraces with high
genetic diversity and food potential is mandatory. Thus, landraces and wild
2.2 Major Abiotic Stresses Affecting Seed Quality 55

relatives will be important genetic sources for developing new varieties with
higher yields (in terms of quantity and quality) and enhanced stress tolerance.
Latin-​American crops comprise high diversity of plant species as they are
adapted to a wide range of environments; this has resulted in large genetic
pools and, therefore one of the richest repositories worldwide (Bazile et al.
2014). Several cultivars from this region have been widely explored, and many
hybrids and improved varieties have been produced. This is the case of Zea mays
(maize) and Phaseolus vulgaris (pea). However, other crops yet remain to be
further explored, such as Amaranthus sp (amaranto or kiwicha), Chenopodium
pallidicaule (cañahua), Chenopodium quinoa (quinoa), Salvia hispanica (chia),
Plukenetia volubilis (inchi) and Lupinus mutabilis (lupino andino or tarwi), ones
that show great food potential and an unexploited capacity to resist different abi-
otic stresses that make them suitable to grow in marginal environments. Besides,
wild accessions of these species will remain a viable source of germplasm, not
only to maintain high productivity, but also to increase seed quality (FAO 2016).

2.2 MAJOR ABIOTIC STRESSES AFFECTING SEED


QUALITY
Many works have suggested that low protein content, the scarcity of
micronutrients or other nutritional relevant compounds such as antioxidants,
which are consumed from food products coming from staple cultivars, have
resulted in unhealthy diets that are linked to different human pathologies (Kucek
et al. 2015; Fan et al. 2008). On the other hand, ancient plant food resources
could be a good source of healthy and beneficial nutrients (Arzani and Ashraf
2017; Peñas et al. 2014). Nonetheless, how abiotic stresses (such as extreme
temperatures, low water availability, high salts and mineral deficiencies) have
shaped the production quality of ancient crops is something that requires fur-
ther investigation.
The presence of stress at any growth stage can impact crop yield and its nutri-
tional quality. These effects include processes such as impaired gametogenesis,
fertilization, embryogenesis, altered nutrient assimilation and mobilization, and
a reduction in the accumulation of reserves in the endosperm, affecting seed
development and composition (Waqas, et al. 2021). A brief description of the
effect of the three main abiotic stresses enhanced by global warming, which
are able to change plant productivity (in terms of seed quality and quantity) is
presented in this section. Also, at this point it is necessary to clarify that the
terms seed and grain will be distinguished throughout the text, although the fact
they may be related terms, they are different from a botanical standpoint. While
the term seed (consisting of seed coat, endosperm and embryo) will be referred
56 Environmental Controls over Seed Quality

to as an ovule containing an embryo, grain (consisting of bran, endosperm and


germ) will be referred to as fusion of the seed coat and fruit tissues. Additionally,
the term seed quality will be referred to aspects related to human nutritional
consumption rather than their ability to germinate.

2.2.1 Drought
Climate change generates considerable uncertainty about future water avail-
ability. Drought affects 45% of the world’s agricultural land, and forecasts pre-
dict increasing frequency of insufficient precipitation and consequent aridity in
many parts of the world (Bates et al. 2008; IPCC 2018). There is strong scientific
consensus that one of the primary physiological targets of water stress in plants
is photosynthesis (Lawlor and Cornic 2002). Primarily, drought decreases photo-
synthetic rates through the reduction of stomatal conductance, photooxidation
and enzyme damage, thereby decreasing the number of assimilates available
for sink tissues (Sehgal et al. 2018). Besides, drought stress during seed filling
induces embryo abortion, decreases the mobilization of reserves and reduces
the number of amyloplasts. Thus, carbohydrate synthesis (assimilation and
metabolism) in plants is severely altered, affecting quality and yield (Çakir 2004).
Interestingly, drought can induce increased content of molecules with anti-
oxidant capacity in several plant species, as will be discussed throughout this
chapter (Bascunan-​Godoy et al. 2016; Fischer et al. 2017; Laxa et al. 2019). The
physiological basis of this response may vary among species, but in general, it
involves an increase in nutrient remobilization and the induction of antioxidant
metabolism. Within this context, the control of water availability could be used
as a strategy to improve the nutritional quality of several seeds, grains and fruits.

2.2.2 High Temperatures
The world is facing a gradual temperature rise, along with a higher frequency
of heat waves leading to warmer days and nights. Indeed, it is projected that
temperatures will have increased by 2°C by the end of the 21st century (IPCC
2021). Major effects of high temperature on plants (usually ⩾30°C) include the
reduction of life cycles, pollen abortion, kernel shrinkage, reduction of seed
reserves, anther indehiscence and reduced development of the pollen tube.
Physiologically, high temperatures alter processes including photosynthesis
and respiration impairment interfering with enzyme activities, resulting in
chlorophyll degradation, or causing general damage to photosystem II, electron
transfer and energy balance, which is crucial to maintaining cell functioning
(Shah and Paulsen 2003; Prasad et al. 2006; Li et al. 2020). All these events result
in reduced crop yield (Rezaei et al. 2015).
2.2 Major Abiotic Stresses Affecting Seed Quality 57

Reproductive stages are sensitive to elevated temperatures in plants.


Simulated high temperature waves during the vegetative and reproductive
stages in maize showed that the latter stages of development are more sensi-
tive to elevated temperatures. Thus, high temperatures during pre-​and post-​
anthesis reduce CO2 exchange rates (~17%), growth (17–​29%), grain number
(7–​45%),and grain yield (10–​45%) (Waqas et al. 2021). Furthermore, during
flowering, elevated day and/​or night-​time temperatures negatively affect
floret number, silk number, and grain development and quality in rice, canola
and maize (Matsui et al. 2001; Cicchino et al. 2010; Pokharel et al. 2020). Air
temperature above 35°C suppresses rice ovary fertilization and inhibits the
grain filling process, related to a detrimental effect on grain yield (Zhang
et al. 2007).
It should be noted that normally in the field, plants are exposed to a com-
bination of stresses (i.e. heat stress is often associated with drought). This cir-
cumstance worsens the crop’s stress response resulting in larger yield penalties
compared to a single stress situation (Dreesen et al. 2012). Interestingly, heat
and drought stress can disrupt the accumulation of various seed constituents,
primarily starch and proteins, by inhibiting the enzymatic processes of synthesis
(Behboudian et al. 2001; Farooq et al. 2018).

2.2.3 Salinity
Salinity is one of the most deleterious environmental factors limiting crop prod-
uctivity (Shrivastava and Kumar 2015). The main causes of salinity in crops are
irrigation with highly salted groundwater, disintegration, and release of sea-
water in coastal areas, and accumulation of salts in arid/​semiarid regions due
to insufficient leaching of ions (Chinnusamy et al. 2005). Nowadays, nearly 20%
of irrigated agricultural lands are considered saline. Adverse effects of salinity
on plant growth are related to the effects of osmotic stress (due to lower water
availability, similar to what happens under drought and high temperatures) and
ionic stress inducing homeostasis imbalance (Epstein et al. 1980). These two
major effects are accompanied by those caused by oxidative stress and nutrient
imbalance that eventually trigger cytotoxicity and plant growth impairment
(Abobatta 2020).
Salt resistance involves adaptation in order to maintain physiological and
biochemical homeostasis, including structural and molecular adaptation and
salt exclusion mechanisms towards minimizing salt concentrations in cells.
Interestingly, among salt-tolerant species, there are halophytes that naturally
grow and complete their life cycle under salinity conditions, whereas most crops
fail in growth and development and consequently cannot produce yield (Ruiz
et al. 2016).
58 Environmental Controls over Seed Quality

Several crops, including many from the Andean region such as quinoa,
kiwicha or cañahua, can increase the velocity and total percentage of germin-
ation under moderate (100 to 150 mM NaCl, equivalent to 10 to 15 dS m-​1) to high
salinity conditions (Delatorre and Pinto 2009; Razzaghi et al. 2011). However,
growing at high salinity conditions reduces seed value by both, changing the
protein profile and/​or protein content (Aloisi et al. 2016; Ruiz et al. 2016; Fischer
et al. 2017). This could be explained, at least partially, because salinity induces
nutritional plant deficiencies or imbalances due to competition with sodium
(Na+​) or chloride (Cl-​), resulting in decrease in the uptake and transportation
of nutrients including nitrogen, phosphorus, potassium and calcium (Hu and
Schmidhalter 2005).

2.3 GENOTYPE AND ENVIRONMENT INTERACTION


ON CROP PRODUCTIVITY AND SEED QUALITY
While much literature has centered on studying the impact of abiotic stress
on crop yield and the underlying mechanisms, there is scarce information on
the relationship between abiotic stress and grain/​seed quality (even for staple
crops). This section attempts to present an update of the available information
regarding how environmental conditions affect seed quality in the most relevant
and promissory Latin-​American crops.

2.3.1 
Chenopodium pallidicaule and Chenopodium quinoa
Chenopodium quinoa Willd. (quinoa, quinua) (Figure 2.1a, b) and Chenopodium
pallidicaule Aellen (also known as cañahua or cañihua) (Figure 2.1c, d) are two
related annual crops that belong to the family Amaranthaceae. Both species are
native to the Andean altiplano, a high plateau situated at 3500–​4200 m above
sea level between the western and eastern areas of the Andean Cordillera. The
cultivation of cañahua and quinoa dates back more than 7000 years when it was
established in the area as a staple crop by ancient Incan and pre-​Incan soci-
eties. After the Spanish conquest, the cultivation of ancient crops was likely
discouraged due to their association with indigenous cultures and ignorance of
their benefits (Aellen and Just 1943), and became marginal crops for farmers’
subsistence in the Andean region.
The edible seeds of cañahua and quinoa (dicot plants) are not grains, rather
they are achenes ( fruits), composed of a single seed enclosed by an outer peri-
carp (Abdelbar 2018). Seeds of both plants have unique nutritional profiles suit-
able for human consumption. Seeds from C. pallidicaule and C. quinoa are gluten
free, good sources of dietary fiber (Repo-​Carrasco et al. 2009, 2019) and contain
2.3 Genotype and Environment Interaction on Crop Productivity 59

Figure 2.1 Latin-​American crops.


Notes: (a) Chenopodium quinoa Willd (quinoa) at seed filling stage (picture taken in the
northern part of Chile [Vicuña, La Serena, Chile]); (b) dry seeds of a red cultivar of quinoa;
(c) Chenopodium pallidicaule Aellen (cañihua) at flowering stage (note the very small
flowers); (d) dry seeds of cañihua ecotype Kello; (e) Amaranthus ssp. (kiwicha) at flowering
stage; ( f) dry seeds of Amaranthus caudatus (variety Oscar Blanco); (g) Phaseolus vul-
garis (beans) at reproductive stage ( fresh pods in the plant); (h) dry seeds of beans;
(i) Zea mays (maize) during flowering in an established plantation in Peru; (j) diversity of
corn cobs colors including yellow, red and white; (k) Salvia hispanica (chia) at flowering;
(l) variegated seeds of chia, cultivar Oaxaca from Mexico; (m) fresh capsules of
Plukenetia volubilis (sacha inchi); (n) dry capsules of sacha inchi with testa; (o) plantation
of Lupinus mutabilis (tarwi) during flowering in the experimental station of Camacani
(Puno [Peru]); (p) seeds of tarwi, variety SLP-​4. Source: Pictures courtesy of Dr. Enrique
Martinez (Center of Advanced Studies in Arid zones [CEAZA, Chile]), Dr. Alberto
Pedreros (Universidad de Concepcion [Chile), Dr. Grandez de Iquitos (University of
Altiplano [Peru]), Dra. Nancy Chasquibol Silva (Centro de Estudios e Innovación de
Alimentos Funcionales (CEIAF), Universidad de Lima [Peru]) and Dr. Angel Mujica
(University of Altiplano (Peru) and Amaranth Promotion Network [Mexico]).
Notes: Figure 2.1 shows 16 panels with pictures of different plants and seeds.
A) Panel A shows a close look at many star-​shaped seeds that form a panicle in quinoa
with a reddish color.
60 Environmental Controls over Seed Quality

Figure 2.1 (continued)


B) Panel B has two hands showing quinoa seeds belonging to Pandela landrace with
different colors including white and reddish.
C) Panel C is a close look at cañihua plants in green and dark color in the field.
D) Panel D shows seeds of cañihua in an orange color.
E) Panel E shows a dark red panicle of a kiwicha plant (in green) growing in the field.
F) Panel F shows kiwicha whitish seeds.
G) Panel G shows pods of a bean plant.
H) Panel H shows oval shaped white beans with dark spots.
I) Panel I shows maize plants in the field.
J) Panel J shows different types of maize ears dispersed in five rows and in orange to
dark red color.
K) Panel K shows chia flowers in a plant with small blue petals.
L) Panel L shows oval shaped chia seeds in a brownish color and darker spots.
M) Panel M shows a Sacha inchi plant with four or five lobes of fruit capsules.
N) Panel N shows mature, star-​shaped, five lobe Sacha inchi fruit.
O) Panel O shows tarwi plants with panicles with many pink to purple flowers in the
field.
P) Panel P shows bean-​like tarwi seeds of white and black colors and a label in the
center of the picture indicating that these are tarwi seeds.

about 20% protein, with a complete set of essential amino acids (Penarrieta et al.
2008). In addition to high‐quality protein, both seeds contain a wide variety of
other health‐promoting compounds, including antioxidants such as phenols
and flavonoids (Repo‐Carrasco‐Valencia et al. 2010), and lipids, including the
fatty acids Oleic (C18:1) and Linoleic (C18:2) (Khaitov et al. 2020). Both species
are considered Andean superfoods (Mangenson et al. 2019; Repo-​Carrasco-​
Valencia et al. 2019) and quinoa has been chosen as a crop that might contribute
to global food security during the next century (FAO 2019).
Cañahua is a poorly studied species and is considered a partially
domesticated crop whose cultivation is mainly restricted to the Andean
region. Non-​uniform seed ripening and small seed size are the principal agro-
nomic issues that have prevented a more extensive cultivation of cañahua
(Mujica 1994). In contrast, C. quinoa, with different available genotypes and
a higher size of seed, is a more popular crop, cultivated in many areas of the
world, with varieties already developed in North America and Europe (Alandia
et al. 2020).
There is little information available regarding the genetic diversity of cañahua
(Mangelson et al. 2019). In fact, the scientific community is worried about its
in situ preservation, which is crucial for future genetic improvement (plant
breeding). Nevertheless, efforts to preserve the genetic diversity of cañahua
have resulted in the creation of a collection of cañahua germplasm in two banks
located in Peru and Bolivia (Flores 2006; IPGRI et al. 2005).
2.3 Genotype and Environment Interaction on Crop Productivity 61

In contrast, there is much more information about the genetic diversity of


quinoa. Worldwide, there are more than 6000 landraces of quinoa cultivated
by farmers. According to their adaptation capacity to specific agro-​ecological
conditions, cultivars can be classified into five ecotypes: Highlands (or Altiplano
type); Inter-​Andean valleys; Yungas (grown under tropical conditions); Salares
(grown at high altitude salt lake areas and with limited volume of annual rain-
fall [150–​300 mm]); and coastal/​lowlands (rainfall ranges from 500 to 1500 mm
annually) (Martinez et al. 2009; Bazile et al. 2016). In spite of the difficulties in
sequencing a tetraploid organism such as quinoa, its genome sequence was
recently published using a Chilean coastal ecotype (Jarvis et al. 2017) and the
variety “Real” from the altiplano (Zhou et al. 2017).
Importantly, among quinoa landraces, coastal/​lowlands are of particular
importance due to their widely longitudinal geographic range, which allows
greater photoperiod adaptation that makes them highly suitable for quinoa cul-
tivation into different climatic zones (Jacobsen 1997; Bendevis et al. 2014).
The genotype and environment interaction in quinoa has been studied in
different published works. Bertero et al. 2004 tested 24 cultivars of quinoa at
14 sites across three continents during two growing seasons. They found great
differences in seed yield and composition among genotypes through different
environmental growing conditions. The great variability of quinoa seed quality
was supported by the work of Reguera et al. 2018, in which three varieties of
quinoa were adapted to specific agroecological conditions –​ Salcedo-​
INIA
(developed by INIA-​Peru), Titicaca (developed by the University of Copenhagen,
Denmark) and Regalona (developed by Baer, Chile) –​were grown in different
environmental conditions in Spain, Peru and Chile. The results revealed that pro-
tein content, amino acid profiles, mineral composition and phytate amount in
seeds varied depending on cultivation conditions and genotype used, while other
parameters, such as saponin or fiber were stable across locations. Interestingly
all genotypes studied presented higher antioxidant contents when they were
cultivated in the north of Chile.
When analyzing quinoa abiotic stress responses, it should be considered that
one of the major limitations found in quinoa cultivation in the field is the combined
effect of drought and high temperatures. Quinoa is especially sensitive to elevated
temperatures at the reproductive stage, i.e., from the beginning of flowering (pre-​
anthesis) to the end of flowering (post-​anthesis), with the seed filling period
also being a sensitive stage) (Geerts et al. 2008). In general, several reports have
stated that seed yield potential and the content and quality of proteins in quinoa
seeds are reduced under water deficits (Fisher et al. 2013; Bascunan-​Godoy et al.
2016). Working with two lowland quinoas cultivars, Bascunan-​Godoy et al. (2016)
found that water stress increased the accumulation of metabolites that belong to
the ornithine pathway (N metabolism) in both genotypes, Faro and BO78; how-
ever, after a few days of re-​irrigation, only Faro was able to recover metabolite
62 Environmental Controls over Seed Quality

levels to those found in control conditions. Remarkably, this genotype showed


higher drought tolerance and the ability to store betacyanins, which are nitrogen-​
containing pigments with high antioxidant power (Bascunan-​Godoy et al. 2018a).
Interestingly, Fisher et al. 2013 reported an increase in antioxidant capacity
measured using DPPH (2,2-​diphenyl-​2-​picryl-​hydrazyl) while hydric restriction
in Chilean lowland genotypes increased. The authors agree that it is possible to
produce seeds with higher nutritional value when subjecting plants to controlled
water limitation, avoiding penalties on seed yield.
When referring to temperature stress, Lesjak and Calderini 2017, using a low-
land Chilean quinoa genotype, Regalona, in an experiment performed under
controlled conditions, found that the increase of 4°C during the night reduced
grain yield, biomass and grain number. In another experiment, the effect of
temperature on quinoa seed quality was approached under field conditions to
try different sowing dates, therefore modifying the photo-​thermal conditions
during seed-​filling stage (Curti et al. 2020). Differences in sowing dates induced
changes in the lipid content, accompanied by significant variation in fatty acid
concentration. However, the general decrease in lipid content did not similarly
affect the major unsaturated fatty acids among cultivars. A decrease in oleic and
α-​linolenic concentration was observed for almost all cultivars, whereas lino-
leic concentration remained unchanged for some cultivars or even increased
in others. This result again suggests that cultivar-​specific responses to photo-​
thermal conditions during the seed-​filling period are involved in quinoa (Curti
et al. 2020). More recently, work performed by Matias et al. (2021) showed that
when quinoa, grown under Mediterranean field conditions, suffered heat stress,
it eventually affected yields and seed quality in different ways depending on the
length of the cultivar life cycle. Thus, those early maturing varieties are, poten-
tially, better adapted genotypes for areas that are more susceptible to suffering
heat waves when aiming at preserving yields and seed nutritional qualitative-​
related parameters.
Even though quinoa is considered a facultative halophyte (Boindi et al. 2015),
there is a broad gradient of saline tolerance according to the cultivar´s origin
(Delatorre and Pinto 2009; Razzaghi et al. 2011). Quinoa coming from the alti-
plano or those coming from lowlands, two representative quinoa ecotypes, pre-
sent contrasting salinity resilience (Jacobsen and Stølen 1993; Delatorre-​Herrera
and Pinto 2009; Ruiz-​Carrasco et al. 2011). Many quinoa trials tested worldwide
using salinity irrigation have been performed (Hussain et al. 2018, 2020; Roman
et al. 2020; Adolf et al. 2012; Cocozza 2013; Shabala et al. 2013). These reports found
significant interaction between the genotype used and the irrigation conditions
tested for seed yield, biomass and different agronomic traits. This highlights the
great genotypic plasticity of quinoa and also points to the need to assess the
genotypic performance under particular growing conditions. Furthermore,
2.3 Genotype and Environment Interaction on Crop Productivity 63

the impact of salinity on the nutritional quality of quinoa seeds has started being
analyzed in recent years. For instance, Fischer et al. 2017 showed that salinity is
correlated with reduced content of proteins and antioxidants in seeds. Ruiz et al.
(2016), using two contrasting landraces, salares and coastal, revealed interesting
responses to salinity in that belonging to the latter ecotype in terms of growth,
yield and seed quality expressed in enhanced total polyphenol content and anti-
oxidant activities in seeds, accompanied by superior ability to germinate under
saline conditions; however, the total protein content decreased slightly under
salinity. In contrast, in another study conducted by Wu et al. 2016, in which four
quinoa cultivars were grown under six salinity treatments and two levels of fer-
tilization, little variation was found in the protein contents, but changes were
detected in seed density and hardness. Nevertheless, further studies should be
conducted to target the genotypic dependence on salinity stress response in
quinoa.

2.3.2 Amaranthus ssp.


The genus Amaranthus comprises about 400 species of annual or short-​lived
perennial plants collectively known as amaranth. They belong to the family
Amaranthaceae, as does quinoa or cañahua. The historical evidence depicted
that amaranthus domestication and cultivation started about 8000 years ago
by pre-​Columbian civilizations in South and Central America. Amaranths (also
known as Huahtli or Kiwicha) (Figure1e, f) were a staple food for the Aztecs,
Mayans and Incas civilizations (Sauer 1950a, 1950b; Pal and Khoshoo 1972;
Alvarez-​Jubete et al. 2010a). However, after the arrival of and further coloniza-
tion by Europeans, amaranth crops were nearly eradicated and confined to very
few indigenous communities.
Some species produce edible leaves (e.g., Amaranthus tricolor) and gluten
free grains (Amaranthus hypochondriacus, Amaranthus caudatus, Amaranthus
cruentus), being a rich and inexpensive source of daily required proteins,
antioxidants and bioactive compounds (Bressani 1994; Gimplinger et al. 2007;
Alvarez-​Jubete et al. 2010a, 2010b). Additionally, Amaranthus seeds show an
exceptional nutritional profile which includes crude protein 12–​19%, fats 5–​8%,
starch 62–​69%, total sugars 2–​3%, dietary fiber 4–​20% and ash 3–​4% (Alvarez-​
Jubete et al. 2009). Notably, seeds are highly enriched in essential amino acids,
such as lysine (363 –​421 mg/​g N). Regarding fatty acids, amaranth seeds have
high concentrations of the unsaturated fatty acids linoleic and oleic acids (62 and
20% respectively). Other important fatty components are tocopherol (vitamin E)
and tocotrienols. Besides, amaranth seeds have the highest content of minerals,
especially Ca, Mg, Zn and Fe compared with other pseudocereals and cereals
(Alvarez-​Jubete et al. 2010a).
64 Environmental Controls over Seed Quality

Amaranth grows well in temperate and tropical regions. It shows high resili-
ence to various abiotic stresses, either imposed by climate or soil conditions,
especially those encountered in semiarid biomes. Amaranth possesses a C4
photosynthetic pathway, which is an evolved mechanism for CO2 fixation, redu-
cing water loss (Sage et al. 2007). Hence, amaranth plants can perform well
during the vegetative and reproductive stages without significant growth and
yield penalties under heat and drought stress. Flowering starts about 4 to 8
weeks after sowing, but phenological stages vary significantly among species/​
cultivars and soil nitrogen content.
Due to their high resilience plus their nutritional and pharmaceutical
value, there is a large germplasm collection of amaranth crop species. For
example, the United States Department of Agriculture (USDA) and the National
Botanical Research Institute (NBRI) collections include approximately 3000
and 2500 accessions, respectively. However, non-​ domesticated species are
underrepresented in seed banks worldwide (Achigan-​Dako et al. 2014).
Different works have highlighted the outstanding ability of acclimatization
and adaptation to different abiotic stresses of Amaranthus (Liu and Stützel 2004;
Moreno et al. 2017). Under drought stress, several genotypes of amaranth showed
osmotic adjustment capacity and the ability to control water loss by reducing
stomatal conductance and leaf expansion, thus, conferring resistance to severe
drought (Liu and Stützel 2002). However, resistance depends on the genotype.
For example, Liu and Stützel (2004) evaluated the effects of drought stress on
water use efficiency and economic spectrum traits in four genotypes of amar-
anth. The authors found differences in biomass partitioning between roots and
shoots depending on the genotype, but without detecting a significant effect on
water use efficiency. However, Tsutsumi et al. (2017) reported that the water use
efficiency of amaranth crops can vary significantly with genotype, mainly due to
variations in structural, biochemical and physiological components. Probably,
both the physiological C4 pathway and morphological traits together finely
modulate drought resistance in Amaranthus species. Seed resilience against
water deficit is remarkable, as shown by Pulvento et al. (2021). The seed yield
components of Amaranthus hypochondriacus were not significantly affected
by reduced irrigation, nor was seed quality. Although more research is needed,
this is a promising feature projecting intensive use of amaranth plants for food
security in a climate changing world.
Amaranth plants are not only suitable crops for drought prone regions, but
also grow rapidly under high temperatures. Indeed, full genotypic potential is
reached at day/​night temperatures above 25/​15°C, respectively, with optimal
growth at about 30°C (Achigan-​Dako et al. 2014). Nevertheless, much more
research work is needed to fully understand the physiological mechanism of
Amaranthus to cope with heat stress.
2.3 Genotype and Environment Interaction on Crop Productivity 65

Salinity is the third major environmental constraint for plant growth,


which is exacerbated by climate change. Amaranth species are recognized
as glycophytes (Wang et al. 1999). However, amaranth can tolerate salt
concentrations up to 150 mM NaCl. Saucedo et al. (2017) reported that
under salt stress (0.4 to 0.8 M NaCl), Amaranthus cruentus and wild relatives
accumulated protective proteins (LEA) in leaves, stems and roots. Agronomical
practices such as seed priming have demonstrated significant improvement in
Amaranthus caudatus seed germination under severe salt stress (up to 200 mM
NaCl) (Moreno et al. 2017).
Responses to combined abiotic stress are far from being well understood in
Amaranthus species. Pulvento et al. (2021) studied drought and salinity effects
in grains of Amaranthus hypochondriacus under field conditions during 3 years
in Italy (Pulvento et al. 2021). Individually, differential water supply (irrigation
time, irrigation volume) did not induce changes in yield and quality, but salinity
reduced seed yield by 55%, compared to control conditions. However, it was
remarkable that the quality of amaranth seeds was preserved under combined
drought-​salinity treatments and no significant differences in starch, protein
and lipids were found (Pulvento et al. 2021). Interestingly, salinity induced a
significant and remarkable increase in β‐cyanin, β‐xanthin, betalain, total
carotenoids, β‐carotene, ascorbic acid, total polyphenolic content, total fla-
vonoid content and total antioxidant capacity in leaves of Amaranthus tri-
color. Hence, it is highly probable to find similar results in Latin-​American
Amaranthus crop species.
There is no doubt about the nutritional quality and the potential of discovering
new bioactive compounds in Amaranthus. Nowadays, Amaranthus presents a
great research opportunity to properly understand the genetic control of physio-
logical responses and resilience to environmental stresses. Also, plant biologists
have the challenge of studying the gaps in biological processes against combined
stresses, enhanced yield and harvesting index, while keeping the nutritional
quality of Amaranthus.

2.3.3 
Phaseolus vulgaris
Phaseolus vulgaris L. (beans) (Figure1 g, h) belongs to the Fabaceae family, which
comprises species displaying a wide variety of forms: trees, shrubs and herbs,
including many with a climbing growth habit. Common beans are the most fre-
quently produced and consumed worldwide legume, with high commercial value
(Broughton et al. 2003). Like other legumes, common beans can fix atmospheric
nitrogen (N2) through the symbiotic fixation process with rhizobia (Lugtenberg
and Kamilova 2009), thus allowing a reduced use of chemical fertilizers and pro-
moting more sustainable agriculture.
66 Environmental Controls over Seed Quality

Although the dried seeds of P. vulgaris are low in methionine and cyst-
eine, these are an important source of vegetal protein for millions of people,
supplementing those amino acids which are lacking in diets based on maize, rice
or other cereals (Broughton et al. 2003; Wortmann 2006). Beans are especially
rich in lysine and tryptophan, iron, copper and zinc; and beneficial antioxidants
and flavonoids (Wortmann 2006). Due to extensive plant-​ breeding efforts,
P. vulgaris is cultivated in several agroecological environments and comprises
numerous cultivars that differ in their growth habitat as well as seed size and
color (Purseglove 1968; Singh et al. 1991a).
The wild bean was domesticated several times in the pre-​Colombian era in
both Mesoamerica (Mexico) and the Andean zone, but phylogenetic and popula-
tion structure analyses have confirmed the origin of P. vulgaris in Mesoamerica
(Bitocchi et al. 2012). Mesoamerican and Andean populations are distinguished
by yield potential, morphology (Singh et al. 1991a,c), isozymes (Singh et al.
1991b), DNA molecular markers (Nodari et al. 1993), as well as by physiological
traits related to photosynthesis (Castonguay et al. 1991). Additionally, there are
other important diversity centers in Brazil, North America, Africa, the Middle
East and Europe. The world collection of cultivated and wild P. vulgaris is held
in the germplasm bank of CIAT (International Center of Tropical Agriculture)
in Cali, Colombia. In 1997, 364 genotypes were identified among the accessions
held in the CIAT collection, and among commonly grown national or regional
cultivars (Beebe et al. 1997).
Many studies considering genotype and environmental factors have been
conducted (Sozen et al. 2018; Suarez et al. 2018; Rainey et al. 2005), con-
cluding that climate changes between years affect yield parameters in dry bean
genotypes.
Regarding abiotic stress, water scarcity is one of the main limiting factors for
common bean crops, negatively affecting seed yield and quality, which might be
related to the fact that the symbiotic nitrogen fixation is rapidly inhibited under
these conditions (Sinclair and Serraj 1995). Local farming practices, domesti-
cation and worldwide spread has entailed the development of a wide variety
of common bean genotypes with a wide range of resistance to water scarcity.
Lopez et al. 2020 studied the molecular basis of differential drought tolerance in
two accessions of common beans (PMB-​0220 and PHA-​0683). They found that
differential regulation of ABA synthesis and signaling related genes among the
two genotypes and control of drought-​induced senescence makes a relevant
contribution to the higher drought resistance level of PHA-​0683 accession. The
effect of water stress on the physical and chemical qualities of the common bean
seeds (variety BRS Realce) was studied in Planaltina, Brazil (Silva et al. 2020).
In this work, water deficit increased the protein and reduced the carbohydrate
and ash contents. Additionally, they found that water stress reduced macro-​and
2.3 Genotype and Environment Interaction on Crop Productivity 67

micro-​mineral content in the grains and changed the physical quality of the
seeds. In spite of the results reported here, other studies have reported a reduc-
tion in protein content under drought (Khalil et al. 2010), but different proteins
respond in various ways.
Under field conditions it is difficult to separate the effect of drought and
high temperatures. Controlled condition experiments conclude that the effect
of high temperatures on bean seeds were more detrimental to seed quality
than was limited rainfall (Muasya et al. 2008). Exposure to high temperatures
during two reproductive growth stages, namely flower bud formation and pod-​
filling, resulted in severe damage (Shonnard et al. 1994) inducing abscission
in numerous flowers and pods (Konsens 1991), reducing the number of pollen
grains, number of viable seeds and yield components (Da Silva et al. 2020).
Suarez et al. 2020 intended to identify genotypic differences in 91 bean
genotypes (derived from interspecific and intraspecific crosses) to high tem-
perature under the environmental conditions of a tropical dry forest ecosystem.
They found that a few Andean lines presented good adaptation capacity to heat
stress. The superior performance under higher temperatures of the identified
genotypes was related to greater ability to partition dry matter, high pollen via-
bility and early growth maturity, resulting in greater grain filling.
Something interesting was the effect of altitude on P. vulgaris, promoting
the content of antioxidants, total phenols, essential amino acids such as valine,
phenylalanine and isoleucine content. The effect of altitude can be linked to the
higher temperature and different light quality, creating a greater stressful con-
dition for the crop that responded by increasing the synthesis of antioxidant
compounds (Nicoletto et al. 2019). The information reported so far must also be
related to the genotype.
P. vulgaris is considered an extremely salt-​sensitive species. Several authors
have approached the effect of sodium chloride (NaCl) on productivity and yield
in contrasting salt-​sensitive genotypes. Assimakopoulou et al. (2015) found
that ‘Corallo’, is a salinity tolerant genotype due to its capacity to sequester Na
in the roots and maintain appropriate K/​Na and Ca/​Na ratios, restricting the
levels of toxic ions in growing shoots. The higher sensitivity to salt of ‘Romano’
cultivar was related to its higher level of Na in the leaves and the reduction of
leaf number and leaf water content. Taibi et al. 2016, on the other hand, also
working in contrasting salinity-​sensitive genotypes, found that the only quali-
tative difference among genotypes was the level of phenolic compounds in
leaves, therefore, the high-​yielding genotype may have increased the activity
of antioxidant enzymes that give better protection against oxidative damage
throughout higher flavonoid and ascorbic acid contents stored under high
salinity. Regarding seed quality, Farooq et al. (2017) found that seed protein
content declines under salt stress, which can be linked with changes in the
68 Environmental Controls over Seed Quality

symbiotic relationship observed under salinity in common beans (Abdi et al.


2015). These works give some clues for selectioning genotypes under saline
stress conditions.

2.3.4 
Zea mays
Maize (Zea mays L.) (Figure1i, j) is one of the leading cereal grains worldwide,
along with wheat and rice (Ali et al. 2014). Maize production areas yield about
850 million tons and the average grain yield is about 5200 Kg ha-​1 (FAO 2016).
Mexico is recognized as the center of origin and domestication of maize. In
fact, maize domestication and growth were already underway in the Tehuacán
valley (Mexico) about 5000 years ago (Vallebueno-​Estrada et al. 2016) and has
generated great genetic diversity through farming and environmental selection
(Perales and Golicher 2014). The huge genetic diversity of maize is reflected in
more than 5500 accessions that exist worldwide ( from more than 40 countries)
(Prasanna 2012). However, it should be noted that commercial lines present a
narrow genetic base (Kasoma et al. 2021). The modest genetic diversity of com-
mercial genotypes has limited variability for value-​added traits. Commercial
maize genotypes are poor in terms of protein content, essential amino acids –​
lysine and tryptophan –​and also have low levels of biological healthy molecules
(Gavicho-​Uarrota et al. 2011). Teosinte (Z. mays ssp. parviglumis Iltis & Doebley)
and tripsacum are the two wild relatives of maize that have been extensively
characterized (Matsuoka et al. 2002). Unlike many hybrid varieties, maize
landraces present high levels of phytochemical compounds such as phen-
olic compounds ( flavonoids and anthocyanins), carotenoids, xanthophylls,
vitamins and dietary fiber, that contribute to human health (Gavicho-​Uarrota
et al. 2011). Nowadays, new lines of maize hybrids have been improved with rele-
vant traits for quality and yielding through gene introgression from landraces
or wild relatives (White et al. 2007).
Numerous articles report tolerance to abiotic stress, including drought, high
temperatures and salinity of different populations, landraces and inbred lines.
However, the fact that about 80% of the total area of maize sown in Mexico is
produced through subsistence farming, indicates that it is a species capable of
withstanding extreme conditions (Bellon et al. 2011). Drought affects various
morpho-​physiological processes, including plant biomass, root length, shoot
length and photosynthesis in maize (Ali et al. 2014, Yamori et al. 2014). The
occurrence of drought stress during the growth period may lead to significant
yield losses due to nutrition and water imbalances (Ashraf et al. 2016). Several
studies involving non-​drought-​tolerant maize hybrids reveal a reduction in
total phenols, hydroxycinnamic acids and antioxidant capacity, while works
on drought-​tolerant maize populations reported the maintenance or increased
2.3 Genotype and Environment Interaction on Crop Productivity 69

level of healthy molecules such as oil, fatty acid composition, protein, and tryp-
tophan content (Ignjatovic-​Micic et al. 2015).
Drought is the major abiotic stress that hinders crop productivity across the
world (Aslam et al. 2015). The impact of drought largely depends on the inten-
sity and duration of water shortage. Mi et al. (2018) studied different levels of
progressive drought on yield formation of maize, reporting that grain yield
was significantly reduced by either the vegetative (18.6–​26.2%) or reproductive
stage (41.6–​46.6%), which was largely caused by the decrease in kernels per
ear. Also, the effect of drought on yield and seed quality is highly dependent on
the sensitivity or genotype tolerance (Anjum et al. 2017). In more drought tol-
erant genotypes, osmoprotectans’ accumulation might play a key role in their
physiological performance and grain biomass trait (Salgado-​Aguilar et al. 2020).
Development of tassel and ear, pollination, fertilization, embryo development,
endosperm development and grain filling are seriously affected by drought stress
in maize (Aslam et al. 2015). Moderate drought produced a decrease in the grain
yield of 33.7% and 62.3% in a drought-​tolerant and a drought-​sensitive maize
variety; respectively (Qi et al. 2010). Additionally, a decrease in soil water con-
tent negatively affects the physiological quality of the maize grains produced,
affecting germination and post-​germination performance (Aslam et al. 2015;
Machado et al. 2020).
Previous reports indicated that high temperatures affect the persistence
and productivity of leaves, and reduced both whole plant dry mass accumula-
tion and grain yield by a shorter duration of grain filling (Wilson et al. 1973).
Thus, high-​temperature stress during the reproductive stage has been the main
obstacle for increasing maize productivity in many places worldwide (Tao et al.
2016). Despite the negative effect of diurnal warming, higher night temperatures
caused an increased gain in kernel weight, resulting from the remobilization of
stored dry mass (Badu-​Apraku et al. 1983).
Warming may affect maize growth and production due to the direct effect
on its structure, biochemical properties, and gas exchange of maize leaves,
resulting in increased net photosynthesis, C:N ratio and soluble sugars (Zheng
et al. 2013). However, at rising temperatures pollen viability and corn silk recep-
tivity resulted in poor seed set and decreased yield (Hussain et al. 2006). Heat-​
stressed maize plants are unable to convert photosynthates into starch in pollen
because key enzymes involved in starch biosynthesis are inhibited at both tran-
scriptional and post-​transcriptional levels (Boehlein et al. 2019). The reduction
in pollen number contributes to distorting the fertilization process (Sánchez
et al. 2014). Additionally, the development of grain filling is accelerated, redu-
cing the period of amyloplast biogenesis and endosperm cell division reducing
grain size (Yang et al. 2017; Waqas et al. 2019). To overcome the negative effect
of higher temperatures on grain yield, hybrid breeding is recommended because
70 Environmental Controls over Seed Quality

hybrid plants have shown a higher capacity to tolerate heat stress in the field
(Hussain et al. 2006).
The effect of salinity on maize yield is also genotype-​dependent, thus deter-
mination of the salt tolerance level in large genetic resources and breeding
populations will be very important to solving the salinity problem (Konuskan
et al. 2017). Exclusion of excessive Na amounts or its storage in vacuoles, is an
important adaptive strategy for maize under salt stress (Farooq et al. 2015).
The response of maize to salinity also varies depending on the developmental
stage, with germination and stand establishment more sensitive than at later
stages (Fortmeier and Schubert 1995; Radic et al. 2007; Farooq et al. 2015;
Konuskan et al. 2017). Despite this, Wang et al. (2016) and Kang et al. (2010)
worked with moderate salinity irrigation (2–​3 g NaCl L−1) in maize, resulting in
normal growing and greater productivity. In fact, Li et al. (2019) reported that
moderate salinity (5 g NaCl·L−1) maintains oil, crude fiber and ash contents in
grains. However, at increased levels of salinity, the excessive uptake of Cl and
Na ions by maize roots leads to severe interference with other essential min-
eral elements, inducing nutritional imbalance (Hasegawa et al. 2000; Karimi
et al. 2005; Turan et al. 2010). The salinity effect on germination and vegetative
growth is very important in determining yield (Konusan et al. 2017). In some
maize genotypes, yield is significantly reduced under salt stress affecting both
grain weight and number (Farooq et al. 2015). In addition, sink limitations and
reduced acid invertase activity in developing grains produce poor grain setting
under saline (Farooq et al. 2015). In fact, Li et al. 2019 working in the Neidan 314,
which is considered a vigorous maize genotype (Ye et al. 2020), found that grain
moisture and starch content decreased under salinity conditions; however, oil
and ash were maintained, and protein and fiber increased. Still, further studies
are needed for in-​depth understanding of the effect of salinity stress on maize
quality.

2.3.5 
Salvia hispanica
Chia (Salvia hispanica L.) (Figure1 k, l) is an annual crop that belongs to the
Lamiaceae family. The area from western Mexico to East-​Central Mexico (with
altitudes between 1400 and 2200 m a.s.l.) has been identified as the center
of its genetic origin (Cahill 2004). Recent contributions using single nucleo-
tide polymorphisms (SNPs) support this information and revealed that most
of the genetic diversity variation of chia remains in wild populations (Pelaez
et al. 2019).
In pre-​Columbian Mesoamerica, chia was cultivated for its edible seeds with
nutritional and therapeutic properties (Cahill 2003); however, after the Spanish
conquest it was dramatically eradicated (Ayerza and Coates 2005a). Chia is a
2.3 Genotype and Environment Interaction on Crop Productivity 71

good source of polyunsaturated fatty acids and has been pointed out as the oil-
seed plant with the highest omega 3 fatty acid content (Cahill 2003, 2004; Ayerza
and Coates 2005a, 2005b, 2009). Additionally, chia contains high amounts of
proteins (up to 26%), dietary fiber (33.9–​39.9%) and high antioxidant activity
because of high levels of tocopherols, phytosterols, carotenoids and phenols
such as chlorogenic and caffeic acids, myricetin, quercetin and kaempferol
(Amato et al. 2015; Reyes-​Caudillo et al. 2008). Additionally, their leaves present
active compounds of nutraceutical, antioxidant and antimicrobial value (Amato
et al. 2015; Elshafie et al. 2018).
Chia is now cultivated and commercialized due to its highly nutritional seeds;
however, genetic studies suggest a slight loss of diversity accompanying domesti-
cation and a near lack of diversity in modern commercial varieties (Cahill 2004).
Chia is a summer crop characterized as a short-​day plant and intolerant to
frost (Jamboonsri et al. 2012; Baginsky et al. 2014). The characteristics of short-​
day flowering species have been considered a problem to expanding their cul-
tivation to a wider range of latitudes. This is because flowers are too late in the
season, and seeds are not able to end maturation before frosting. Nevertheless,
breeding efforts have produced longer-​day flowering genotypes (early flowering)
to extend cultivation of this crop to other climate areas (Jamboonsri et al. 2012).
Most of these new lines are mutants capable of inducing flowering in light cycles
of 13 and 16 h of day length and a few are day-​length insensitive. Certainly, the
in-​depth study of this germplasm including yield, quality and the identification
of seed metabolites has been a main goal.
Genotype and environment (GX E) interaction on chia quality was studied in
a pioneer experiment performed by Ayerza and Coates (2009). They investigated
how seed yield, protein content, oil content and fatty acid composition varies
among three chia selections when planted in three inter-​Andean valleys of
Ecuador, which differ in elevation (affecting chia’s length of growing period).
They found that location mainly affected seed yield, and, to a lesser degree, the
quality (protein and oil contents as well as fatty acid composition). Additionally,
no differences in protein, oil, fiber, amino acids and antioxidant content com-
position were found between Ecuadorian chia genotypes that differed in seed
color (Ayerza 2013).
However, while no great variation in seed quality has been reported in
Mesoamerica, important differences have been observed in studies conducted
in Europe (Italy) comparing commercial and long day/​vs. early flowering
genotypes. Studies carried out by de Falco et al. 2017, 2018 compared the com-
position of commercial chia genotypes (two black, and one white) and long day/​
early flowering genotypes: G3, G8, G17 (mutant genotypes obtained through
treatment with gamma radiation (Jamboonsri et al. 2012)) and W13.1 (derived
from a cross between G8 and a white-​flowered, white-​seeded commercial chia
72 Environmental Controls over Seed Quality

(wild type)) through nuclear magnetic resonance (NMR) spectroscopy and


metabolomics analysis. The metabolic fingerprinting showed great differences
among these populations. Most of the detected metabolites showed larger
variations in the seeds of early flowering genotypes, compared to commer-
cial seeds. In particular, early flowering and white genotypes were reduced in
carbohydrates compared to black genotypes. Interestingly, the highest content
of omega-​3 fatty acids was found in white commercial chia. The early flowering
genotypes showed the highest content of antioxidant metabolites including
caffeoyl derivatives caffeic, chlorogenic and rosmarinic acid; flavonoids
genistein and quercetin, as well as carbohydrates such as sucrose and raffinose
(but not glucose). These results suggest great potential for different purposes
(related to the food industry) among commercial and early flowering genotypes.
Interestingly, a recent study was performed aiming at identifying single nucleo-
tide polymorphisms (SNPs) and simple sequence repeat (SSRs) markers which
can contribute highly to breeding chia towards generating more nutritious seeds
(Pelaez et al. 2019).
Nonetheless, there is little information about the interaction GX E in this
crop and how it affects chia seed quality, which should be further explored in
future experiments. In its zone of origin, chia grows under rainfed or irrigated
conditions (Coates and Ayerza 1996). Silva et al. 2016 worked in the semiarid
conditions in the north of Chile, and using two genotypes black and white
(commercial accessions from Bolivia) found that chia plants induce strat-
egies related to controlling water loss under water shortage. This is consistent
with the results observed by Lovelli et al. 2019 using both commercial and
early flowering genotypes under controlled conditions. There is little infor-
mation on the effect of drought on seed composition, but Silva et al. 2016
reported a decrease in total oil, omega-​3 and linoleic acid in seeds. Besides,
changes in fatty acid compositions were reported by Ayerza et al. 1995, who
performed experiments in five of Argentina’s northwestern locations, differ-
entially influenced by temperature. This work reported that seed compos-
ition, including total oil content and oleic, linoleic, and linolenic fatty acid
concentrations, varied significantly with location. The effect of drought and
temperatures was similarly observed, when Heuer et al. 2002 showed that sal-
inity reduced the oil content and changed the fatty acid composition (palmitic,
stearic, oleic, linoleic and linolenic) of chia seeds. Additionally, Paiva et al.
2018 showed that conductivity levels higher than 4.5 dS m-​1 combined with
temperatures ranging between 20 and 30°C, negatively affected germination,
growth and biochemical components, such as chlorophylls and proteins of
chia seedlings. More studies that involve metabolic changes during the repro-
ductive stage are necessary to establish a more direct relationship regarding
the changes in the plant and seed composition.
2.3 Genotype and Environment Interaction on Crop Productivity 73

2.3.6 
Plukenetia volubilis
Plukenetia volubilis L. ( family Euphorbiaceae) or their name in Quechuan ‘sacha
inchi’ (Figure1 m, n) is a native plant from the Amazon cultivated as early as
3000–​5000 years ago (Bernal and Correa 1992; Brack 1999).
In America, P. volubilis grows in countries such as Peru, Bolivia, Venezuela,
Colombia, Ecuador and Brazil in tropical or subtropical climates, with
temperatures ranging from 10 to 26°C, relative humidity of 78% and altitudes of
at least 1490 m.a.s.l. (Arfini and Antonioli 2013).
Recently, P. volubilis has attracted increasing attention, because of its ole-
aginous seeds that contain an unusual and outstanding nutritional composition
(Kodahl 2020). Depending on extraction methods, most of these studies report
a lipid content of 45–​50%, with a very high proportion of polyunsaturated fatty
acid, greater than 80% (Castaño et al. 2012), including essential fatty acids such
as α-​linolenic and linoleic acids (Hamaker et al. 1992) and an omega-​3: omega-​6
ratio close to 1 (Chirinos et al. 2013). Furthermore, the seeds have high protein
content (between 22 and 30%) and are a good source of antioxidants (Hamaker
et al. 1992; Gutierrez et al. 2011; Ruiz et al. 2013). Kodahl (2020) compared the
content of omega-​3, omega-​6, and antioxidant capacities reported in different
plant species and found that they were significantly higher in P. volubilis
compared to others typically found in the healthy market.
Studies analyzing the genetic diversity of this crop reveal that, unlike
other Latin-​American species, P. volubilis maintains a high degree of diversity
(Corazon-​Guivin et al. 2009). Rodríguez et al. (2010) reported high variability in
morphology, plant production and oil content. Differences among cultivars in
the chemical and phytochemical composition of seeds were also established by
Chirinos et al. (2013) who worked with 16 genotypes from Peru. Results showed
that content of phytochemicals and fatty acid profiles varied according to
genotype.
To date, there is scarce information regarding the effects of environmental
stresses on the productivity and quality of sacha inchi. However, several works
have shown sensitivity to drought and chilling stress in this crop (Luo et al. 2014;
Lei et al. 2014; Tian et al. 2013), and it seems that in association with mycorrhizal
fungi it increased drought tolerance (Tian et al. 2013). Further studies should
focus on elucidating the impact of environmental stresses on the seed quality
of this crop to shed light on variations that might suffer the same nutritional
characteristics.

2.3.7 
Lupinus mutabilis
The genus Lupinus includes almost 200 species, but only four of them play an
important role in agriculture: L. albus, L. angustifolius, L. luteus and L. mutabilis
74 Environmental Controls over Seed Quality

(Gresta et al. 2017). Lupinus mutabilis Sweet (Andean lupin, tauri, tarwi, tarhui,
chocho or kirku) (Figure 1o, p), is an important crop originally from the Central
Andes, with high seed crude protein (32–​ 53%), and oil (13–​ 25%) content
(Gulisano et al. 2017).
It seems that L. mutabilis was separated from its wild progenitor L. piurensis
about 2600 years ago (Gulisano et al. 2019). Andean lupin was first domesticated
in the Cajamarca region (Peru) and cultivated about 1800 years ago in the plains
of South America (Gulisano et al. 2019), being an important plant used in crop
rotation and contributing to soil fertility through nitrogen fixation and phos-
phorus mobilization (Kurlovich et al. 2002).
Germplasm collections are held mainly in Peru, Bolivia and Ecuador, and
smaller collections are also present in Chile and Argentina, among other
countries. Overall, these gene banks hold more than 3000 genotypes of
L. mutabilis (Gulisano et al. 2019 and references therein). The genetic diversity
of the Andean lupin has been addressed by different authors. Guilengue et al.
2020 performed a phenotypic analysis on the yield component of 23 Andean
lupin accessions. They found that the productivity of primary branches was
an important component of total yield accession, as well as the ability to
produce large seeds. The authors highlighted one accession (LM268) that
achieved these characteristics reaching the highest seed production. The gen-
etic diversity revealed in this study, however, prompts further breeding oppor-
tunities outside the altiplane. On the other hand, Berru et al. 2021 compared
seed characteristics and composition in 33 Andean lupine ecotypes (Peruvian)
with other lupine species, including L. albus, L. angustifolius and L. luteus. The
authors found that the Andean lupines had higher protein, lipid and tocoph-
erol content than L. albus and L. angustifolius and presented similar values to
those presented in L. luteus.
Seeds of L. mutabilis are not directly consumed due to the presence
of quinolizidinic alkaloids that give it a bitter flavor. The alkaloid in
higher concentrations are lupanine, followed by tetrahydrorombifoline,
4-​hydroxilupanine, sparteine and 13-​hydroxilupanine (Gross et al. 1988; Ortega-​
David et al. 2010; Castañeda et al. 2008). These substances are important for the
plant since they protect it from phytopathogens and herbivores (Keeler 1976).
Cortes-​Avendano et al. 2020, working in ten ecotypes, found that alkaloids
were influenced by geographical location, likely due to the different climatic
conditions. They evaluated the effect of the aqueous debittering process of seeds
in the profile and levels of quinolizidine alkaloids by gas chromatography and
mass spectrometry (Jacobsen and Mujica 2006). From eight alkaloids identified
before debittering, only small amounts of lupanine and sparteine remained in the
seeds, and no other alkaloids were identified. The debittering of Andean lupines
reduced the level of alkaloids to levels far below the maximal level allowed by
2.3 Genotype and Environment Interaction on Crop Productivity 75

international regulations. These results suggest that L. mutabilis harbors nutri-


tional characteristics suitable for modern food trends.
Regarding the necessary climate conditions for its growth, it is known that the
plant is susceptible to excess humidity and scarce water availability (Jacobsen
and Mujica 2006). Optimal precipitations range from 350 to 850 mm during
their growing cycle; and drought during flowering might stimulate early matur-
ation and cause considerable biomass and seed yield losses (Hardy et al. 1997).
Additionally, it does not tolerate frosts at early stages (Jacobsen and Mujica
2006; Simioniuc et al. 2021).
The effect of water stress on L. mutabilis was analyzed by Carvalho et al. 2004,
2005. The water deficit was imposed 15 days after anthesis, for 20 days. They
found that water stress had little effect on total biomass, but strongly reduced
plant water status, gas exchange and leaf area. Water stress changed the com-
position of all organs in the plant, increasing sugars and oils in stems and sugars
in leaves. Regarding seed quality, water stress increased more than twice the
level of total carbohydrates ( from 50 to 130 g kg-​1 of seed dry weight), modi-
fying the sugar profile, increasing sucrose and reducing raffinose. The water
stress also halved oil content, while protein content was maintained. It would
be interesting to know how the quality of the oils and proteins were affected;
however, these results suggest an impairment in nutrient translocation to seeds,
affecting their final quality.
Another interesting abiotic factor that impacts Andean lupin growth is
freezing. This legume plant species shows low tolerance to spring frost, which
has limited its adaptation to temperate environments (Simioniuc et al. 2021).
Recently, Simioniuc et al. 2021 compared three genotypes of L. mutabilis with
L. albus in a range of freezing temperatures ( from -​2°C to -​10°C) at different
seedling stages. They found greater frost damage effect in the first stage studied
(when cotyledons were breaking through the soil surface). Additionally, they
reported that the LIB222 genotype displayed the highest frost resistance,
related to a higher level of anthocyanin in its tissues. The anthocyanins are
involved in cold and frozen resistance in many plant species due to their
high antioxidant activity (Chalker-​Scott 1999; Ahmed et al. 2015). However,
it is not known so far if these results found at early growing stages could be
extrapolated to other phases of growth (such as flowering or seed filling stage),
or how the quality of the seed may vary under low temperatures (Simioniuc
et al. 2021).
Considering that L. mutabilis is a subsistence species with potential for
growth on marginal lands, more research efforts should be focused on identifi-
cation of genotypes with early maturation and high frost and drought tolerance,
while conserving the high protein and oil content in seeds.
76 Environmental Controls over Seed Quality

2.4 PERSPECTIVES OF CROP PRODUCTION UNDER


STRESS CONDITIONS
Current domesticated crop plants have resulted from a combination of years
of evolution and human selection. It is estimated that more than 30,000 plant
species have been cultivated throughout human history, and of these, approxi-
mately 7000 are crop species (Jacobsen et al. 2015; Kew 2016; Khoshbakht and
Hammer 2008). Nevertheless, nowadays, fewer than 20 species provide most
of the world’s food, with only three crops –​rice, wheat and maize –​providing
60 percent of the world’s food energy intake (Lenné and Wood 2011; FAO 2016).
The majority of the edible plant species in the world are non-​staple foods, often
native species which are still underutilized. In other words, they are used locally
or regionally and very little is known about their biology, or their nutritional and
nutraceutical characteristics (Sogbohossou et al. 2018).
While research efforts are mainly focused on the major staple crops, exploring
underutilized species would reveal untapped potential and may contribute
not only to improving human and animal nutrition, but also to reducing the
loss of biodiversity, contributing to sustainable agriculture in terms of utiliza-
tion of resources, mainly soil and water (Kahane et al. 2013; Mayes et al. 2012;
Sogbohossou et al. 2018; Ulian et al. 2020). This would increase our possibilities
when aiming at coping with climate change through a more environmentally
friendly agriculture that includes exploring resilient species against an ample
range of stresses. In line with this, Latin-​American crops, including their wild
relatives and landraces, developed and adapted to marginal areas and extreme
environments, could make a very valuable contribution. Only if the current
modern breeding technologies are combined with a deeper awareness/​know-
ledge about our agrobiodiversity will we have better chances of meeting the
current food security challenge of feeding 10 billion people by 2050, within the
context of climate change.

ACKNOWLEDGMENTS
This work was supported by grant laValSe-​ Food-​CYTED (Ref. 119RT0567),
Ministerio de Ciencia e Innovación (MICINN, Spain) (PID2019-​105748RA-​I00)
and ANID Fondecyt Regular N° 1211473.

REFERENCES
Abdelbar, O.H. 2018. Flower vascularization and fruit developmental anatomy of quinoa
(Chenpodium quinoa Willd) Amaranthaceae. Annals of Agricultural Sciences
63: 67–​75.
References 77

Abdi, N., I. Hmissi, M. Bouraoui, B. L’taief, and B. Sifi. 2015. Effect of salinity on Common
bean (Phaseolus vulgaris L.)-​Sinorhizobium strain symbiosis. Journal of New Sciences,
Agriculture and Biotechnology 16: 559–​566.
Abobatta, W.F. 2020. Plant responses and tolerance to combined salt and drought stress. In
Salt and Drought Stress Tolerance in Plants. Signaling and Communication in Plants,
edited by M. Hasanuzzaman and M. Tanveer. Cham: Springer. https://​doi.org/​
10.1007/​978-​3-​030-​40277-​8_​2.
Achigan-​Dako, E.G., O.E.D. Sogbohossou, and P. Maundu. 2014. Current knowledge on
Amaranthus spp.: Research avenues for improved nutritional value and yield in leafy
amaranths in sub-​Saharan Africa. Euphytica 197: 303–​317.
Adolf, V.I., S. Shabala, M.N. Andersen, F. Razzaghi, and S.E. Jacobsen. 2012.
Varietal differences of quinoa’s tolerance to saline conditions. Plant and Soil 357:
117–​129.
Aellen, P. and T. Just. 1943. Key of the American species of the genus Chenopodium L.
American Midland Naturalist 30: 46–​76.
Ahmed, N.U., J.I Park, H.J. Jung, Y. Hur, and I.S. Nou. 2015. Anthocyanin biosynthesis for
cold and freezing stress tolerance and desirable color in Brassica rapa. Functional &
Integrative Genomics 15: 383–​394.
Alandia, G., J.P Rodriguez, S.E. Jacobsen, D. Bazile, and B. Condori. 2020. Global
expansion of quinoa and challenges for the Andean region. Global Food Security
26: 100429.
Ali, Q., A. Ali, M. Waseem, A. Muzaffar, S. Ahmed, S. Ali, K.S. Bajwa, M.F. Awan, T.R. Samiullah,
A.I. Nasir, and T. Husnain. 2014. Correlation analysis for morpho-​physiological traits
of maize (Zea mays L.). Life Science Journal 11(12s): 9–​13.
Aloisi, I., L. Parrotta, K. B. Ruiz, C. Landi, L. Bini, G. Cai, S. Bioondi, and S. Del Duca. 2016.
New insight into quinoa seed quality under salinity: Changes in proteomic and
amino acid profiles, phenolic content, and antioxidant activity of protein extracts.
Frontiers in Plant Science 7: 656.
Alvarez-​Jubete, L., E.K. Arendt, and E. Gallagher. 2009. Nutritive value and chemical com-
position of pseudocereals as gluten free ingredients. International Journal of Food
Science and Nutrition 60: 240–​257.
Alvarez-​Jubete, L., E.K. Arendt, and E. Gallagher. 2010a. Nutritive value of pseudocereals
and their increasing use as functional gluten-​free ingredients. Trends in Food Science
and Technology 21: 106–​113.
Alvarez-​Jubete. L., H.H. Wijngaard, E.K. Arendt, and E. Gallagher. 2010b. Polyphenol com-
position and in-​vitro antioxidant activity of amaranth, quinoa and buckwheat as
affected by sprouting and bread baking. Food Chemistry 119: 770–​778.
Amato, M., Caruso, M.C., F. Guzzo, F. Galgano, M. Commisso, R. Bochicchio, R. Labella,
and F. Favati. 2015. Nutritional quality of seeds and leaf metabolites of Chia
(Salvia hispanica L.) from Southern Italy. European Food Research Technology
241: 615–​625.
Anjum, S.A., U. Ashraf, M. Tanveer, I. Khan, S. Hussain, B. Shahzad, A. Zohaib, F. Abbas, M.F.
Saleem, I. Ali, and L.C. Wang. 2017. Drought induced changes in growth, osmolyte
accumulation and antioxidant metabolism of three maize hybrids. Frontiers Plant
Science 8: 69.
Arfini, F., and F. Antonioli. 2013. Sacha inchi. Research about the Conditions for Recognition
of Geographical Indications in Peru. Lima: CRED.
78 Environmental Controls over Seed Quality

Arzani, A. and M. Ashraf. 2017. Cultivated ancient wheats (Triticum spp.): A potential
source of health-​beneficial food products. Comprehensive Reviews in Food Science
and Food Safety 16: 477–​488.
Ashraf, U., M.N. Salim, A. Sher, S.R. Sabir, A. Khan, S. Pan, and X. Tang. 2016. Maize growth,
yield formation and water-​nitrogen usage in response to varied irrigation and
nitrogen supply under semi-​arid climate. Turkish Journal of Field Crops 21: 88–​96.
Aslam, M., M.A. Maqbool, and R. Cengiz. 2015. Drought Stress in Maize (Zea mays L.)
Effects, Resistance Mechanisms, Global Achievements and Biological Strategies for
Improvement. Springer Briefs in Agriculture ISBN 978-​3-​319-​25442-​5.
Assimakopoulou, A., I. Salmas, K. Nifakos, and P. Kalogeropoulos. 2015. Effect of salt
stress on three green bean (Phaseolus vulgaris L.) cultivars. Notulae Botanicae Horti
Agrobotanici Cluj-​napoca. 43: 113–​118.
Ayerza, R. 1995. Oil content and fatty-​acid composition of chia (Salvia hispanica L.) from
5 northwestern locations in Argentina. Journal of the American Oil Chemists Society
72: 1079–​1081.
Ayerza, R. 2013. Seed composition of two chia (Salvia hispanica L.) genotypes which differ
in seed color. Emirates Journal of Food and Agriculture 25: 495–​500.
Ayerza, R. and W. Coates. 2005a. Chia: Rediscovering a Forgotten Crop of the Aztecs. Tucson,
AZ: University of Arizona Press.
Ayerza, R. and W. Coates. 2005b. Effect of ground chia seed and chia oil on plasma total
cholesterol, LDL, HDL, triglyceride content, and fatty acid composition when fed to
rats. Nutrition Research 11: 995–​1003.
Ayerza, R. and W. Coates. 2009. Influence of environment and genotype on crop cycle
and yield; seed protein, oil, and α-​linolenic ω-​3-​fatty acid content of chia (Salvia
hispanica L.). Industrial Crops and Products 30: 321–​324.
Badu-​Apraku, B., R.B. Hunter, and M. Tollenaar. 1983. Effect of temperature during grain
filling on whole plant and grain yield in maize (Zea mays L.). Canadian Journal Plant
Science 63 :357–​363.
Baginsky, C., J. Arenas, H. Escobar, M. Garrido, D. Valero, D. Tello, L. Pizarro, L. Morales,
and H. Silva. 2014. Determinación de fecha de siembra óptima de chia en zonas de
clima desértico y templado mediterráneo semiárido bajo condiciones de riego en Chile.
Universidad de Chile, Facultad de ciencias Agronómicas, Escuela de pregrado.
http://​repo​sito​rio.uch​ile.cl/​bitstr​eam/​han​dle/​2250/​152​812/​Efe​cto-​de-​la-​fecha-​de-​
siem​bra-​en-​el-​rend​imie​nto-​en-​grano-​de-​chia-​%28Sal​via-​hispan​ica-​L%29-​y-​su-​relac​
ion-​con-​el-​crec​imie​nto-​y-​des​arro​llo.pdf ?seque​nce=​1&isAllo​wed=​y
Bascunan-​Godoy, L., M. Reguera, Y.M. Abdel-​Tawab, and E. Blumwald. 2016. Water deficit
stress-​induced changes in carbon and nitrogen partitioning in Chenopodium quinoa
Willd. Planta 243: 591–​603.
Bascunan-​Godoy, L., C. Sanhueza, C.E. Hernández, L. Cifuentes, K. Pinto, R. Álvarez, M.
González-​Teuber, and L.A. Bravo. 2018. Nitrogen supply affects photosynthesis and
photoprotective attributes during drought-​induced senescence in quinoa. Frontiers
in Plant Science 9: 994.
Bates, B.C., Z.W. Kundzewicz, S. Wu, and J.P. Palutikof. 2008. Climate change and water.
Technical Paper of the Intergovernmental Panel on Climate Change, IPCC
Secretariat, Geneva. The American Midland Naturalist, 168.
Bazile, D., E.A. Martínez, and F. Fuentes. 2014. Diversity of quinoa in a biogeographical
island: A review of constraints and potential from arid to temperate regions of Chile.
Notulae Botanicae Horti Agrobotanici Cluj-​Napoca 42: 289–​298.
References 79

Bazile, D., S.E. Jacobsen, and A. Verniau. 2016. The global expansion of quinoa: Trends and
limits. Frontiers in Plant Science 7: 622.
Beebe, S.E., O. Toro, A.V. González, M.I. Chacon, and D.G. Debouck. 1997. Wild–​weed–​
crop complexes of common bean (Phaseolus vulgaris L, Fabaceae) in the Andes of
Peru and Colombia, and their implications for conservation and breeding. Genetic
Resources and Crop Evolution 44: 73–​91.
Behboudian, M.H., Q. Ma, N.C. Turner, and J.A. Palta. 2001. Reactions of chickpea to water
stress: Yield and seed composition. Journal of the Science Food and Agriculture
81: 1288–​1291.
Bellon, M.R., D. Hodsonb, and J. Hellin. 2011. Assessing the vulnerability of traditional maize
seed systems in Mexico to climate change. Proceedings of the National Academy of
Sciences 108(33): 13432–​13437.
Bendevis, M.A., Y. Sun, E. Rosenqvist, S. Shabala, F. Liu, and S.E. Jacobsen. 2014.
Photoperiodic effects on short-​pulse 14C assimilation and overall carbon and
nitrogen allocation patterns in contrasting quinoa cultivars. Environmental and
Experimental Botany 104: 9–​15.
Bernal, Y and J. Correa. 1992. Especies vegetales promisorias del convenio Andrés Bello.
SECAB. 7: 577–​596.
Berru, L.B., P. Glorio-​Paulet, C. Basso, A. Scarafoni, F. Camarena, A. Hidalgo, and A.
Brandolini. 2021. Chemical composition, tocopherol and carotenoid content of
seeds from different Andean Lupin (Lupinus mutabilis) ecotypes. Plant Foods Human
Nutrition 76: 98–​104.
Bertero, H.D., A.J. De la Vega, G. Correa, S.E. Jacobsen, and A. Mujica. 2004. Genotype and
genotype-​by-​environment interaction effects for grain yield and grain size of quinoa
(Chenopodium quinoa Willd.) as revealed by pattern analysis of international multi-​
environment trials. Field Crops Research 89: 299–​318.
Biondi, S., K.B. Ruiz, E.A. Martinez, A. Zurita-​Silva, F. Orsini, F. Antognoni, G. Dinelli, I.
Marotti, G. Gianquinto, S. Maldonado, H. Burrieza, D. Bazile, V.I. Adolf, and S.E
Jacobsen. 2015. Tolerance to saline conditions. In State of the Art Report of Quinoa in
the World in 2013, Chap. 2.3. Paris: FAO and CIRAD.
Bitocchi, E., L. Nanni, E. Bellucci, M. Rossi, A. Giardini, P.S. Zeuli, G. Logozzo, J. Stougaard,
P. McClean, G. Attene, and R. Papa. 2012. Mesoamerican origin of the common bean
(Phaseolus vulgaris L.) is revealed by sequence data. Proceedings of the National
Academy of Sciences 109 : E788–​E796.
Boehlein, S.K., P. Liu, A. Webster, C. Ribeiro, M. Suzuki, S. Wu, J.C. Guan, J.D. Stewart, W.F.
Tracy, A.M. Settles, D.R. McCarty, K.E. Koch, L.C. Hannah, T.A Hennen-​Bierwagen,
and A. Myers. 2019. Effects of long-​term exposure to elevated temperature on Zea
mays endosperm development during grain fill. The Plant Journal 99: 23–​40.
Brack, A. 1999. Diccionario enciclopédico de plantas útiles del Perú. Cuzco: Centro de
Estudios Regionales Andinos –​Bartolomé de las Casas, p. 400.
Bressani, R. 1994. Composition and nutritional properties of amaranth. In Amaranth-​
Biology, Chemistry and Technology, edited by Octavio Paredes-​ Lopez, 185–​ 205.
London: CRC Press.
Broughton, W.J., G. Hernandez, M.W. Blair, S. Beebe, P. Gepts, and J. Vanderleyden. 2003.
Beans (Phaseolus spp.) –​model food legumes. Plant and Soil 252: 55–​128.
Cahill, J.P. 2003. Ethnobotany of chia, Salvia hispanica L. Economy Botany 57: 604–​618.
Cahill, J.P. 2004. Genetic diversity among varieties of chia (Salvia hispanica L.). Genetic
Resources of Crop Evolution 51: 773–​781.
80 Environmental Controls over Seed Quality

Çakir, R. 2004. Effect of water stress at different development stages on vegetative and
reproductive growth of corn. Field Crops Research 89: 1–​16.
Carvalho, I.S., C.P. Ricardo, and M. Chaves. 2004. Quality and distribution of assimilates
within the whole plant of lupines (L. albus and L. mutabilis) influenced by water
stress. Journal of Agronomy and Crop Science 190: 205–​210.
Carvalho, I.S., M. Chaves, and C.P. Ricardo. 2005. Influence of water stress on the chemical
composition of seeds of two lupins (Lupinus albus and Lupinus mutabilis). Journal of
Agronomy and Crop Science 191: 95–​98 doi: 10.1111/​j.1439-​037X.2004.00128.x
Castañeda, B., R. Manrique, F. Gamarra, A. Muñoz, F. Ramos, F. Lizaraso, and J. Martínez.
2008. Probiótico elaborado en base a las semillas de Lupinus mutabilis Sweet (chocho
or tarwi) seeds. Acta Médica Peruana 25: 210–​215.
Castaño, D.L., M. Valencia, E. Murillo, J.J. Mendez, and J.E. Joli. 2012. Fatty acid compos-
ition of Inca peanut (Plukenetia volubilis Linneo) and its relationship with vegetal
bioactivity. Revista Chilena Nutricion 39: 45–​52.
Castonguay, Y. and A.H Markhart III.1991. Saturated rates of photosynthesis in water-​
stressed leaves of common bean and tepary bean. Crop Science 31: 1605–​1611.
Chalker-​ Scott, L. 1999. Environmental significance of anthocyanins in plant stress
responses. Photochemistry and Photobiology 70: 1–​9.
Chinnusamy, V., A. Jagendorf, and J.K. Zhu. 2005. Understanding and improving salt toler-
ance in plants. Crop Science 45:437–​448.
Chirinos, R., G. Zuloeta, R. Pedreschi, E. Mignolet, Y. Larondelle, and D. Campos. 2013.
Sacha inchi (Plukenetia volubilis): A seed source of polyunsaturated fatty acids,
tocopherols, phytosterols, phenolic compounds and antioxidant capacity. Food
Chemistry 141: 1732–​1739.
Cicchino, M., J.I. Rattalino-​Edreira, M. Uribelarrea, and M.E. Otegui. 2010. Heat stress
in field-​grown maize: Response of physiological determinants of grain yield. Crop
Science 50:1438–​1448.
Coates, W., and R. Ayerza. 1996. Production potential of chia in Northwestern Argentina.
Industrial Crops and Products 5: 229–​233.
Cocozza, C., C. Pulvento, A. Lavini, M. Riccardi, R. d’Andria, and R. Tognetti. 2013. Effects of
increasing salinity stress and decreasing water availability on ecophysiological traits
of Quinoa (Chenopodium quinoa Willd.) grown in a Mediterranean-​type agroeco-
system. Journal of Agronomical Crop Science 199: 229–​240.
Corazon-​Guivin, M., D. Castro-​Ruiz, W. Chota-​Macuyama, Á. Rodríguez, D. Cachique,
E. Manco, D. Del-​Castillo, J.F. Renno, and C. García-​Dávila. 2009. Caracterización
genética de accesiones SanMartinenses del banco nacional de germoplasma
de sacha inchi Plukenetia volubilis L. (E.E. El Porvenir—​INIA). Folia Amazónica
18: 23–​31.
Cortés-​Avendaño, P., M. Tarvainen, S. Jukka-​Pekka, P. Glorio-​Paulet, B. Yang, and R. Repo-​
Carrasco-​Valencia. 2020. Profile and content of residual alkaloids in ten ecotypes of
Lupinus mutabilis Sweet after aqueous debittering process. Plant Foods for Human
Nutrition 75: 184–​191.
Curti, R.N., M.D. Sanahuja, S.M. Vidueiros, C.A. Curti, A.N. Pallaro, and H.D. Bertero. 2020.
Oil quality in sea-​level quinoa as determined by cultivar-​specific responses to tem-
perature and radiation conditions. Journal of the Science of Food and Agriculture
100: 1358–​1361.
References 81

Da Silva, D.A., C.A.F. Pinto-​Maglio, E.C.de Oliveira, R.L.D. dos Reis, S.A.M. Carbonell, and
A.F. Chiorato. 2020. Influence of high temperature on the reproductive biology
of dry edible bean (Phaseolus vulgaris L.) Scientia Agricola 77. doi: 10.1590/​
1678-​992X-​2018-​0233
de Falco, B., G. Incerti, R. Bochicchio, T.D. Phillips, M. Amato, and V. Lanzotti. 2017.
Metabolomic analysis of Salvia hispanica seeds using NMR spectroscopy and multi-
variate data analysis. Industrial Crops and Products 99: 86–​96.
de Falco, B., A. Fiore, R. Rossi, M. Amato, and V. Lanzotti. 2018. Metabolomics driven ana-
lysis by UAEGC-​MS and antioxidant activity of chia (Salvia hispanica L.) commercial
and mutant seeds. Food Chemistry 254: 137–​143.
Delatorre-​Herrera, J. and M. Pinto. 2009. Importance of ionic and osmotic components
of salt stress on the germination of four quinua (Chenopodium quinoa Willd.)
selections. Chilean Journal of Agricultural Research 69: 477–​485.
Dreesen, F.E., H.J. De Boeck, I.A. Janssens, and I. Nijs. 2012. Summer heat and drought
extremes trigger unexpected changes in productivity of a temperate annual/​bian-
nual plant community. Environmental and Experimental Botany 79: 21–​30.
Dwivedi, S.L., S. Ceccarelli, M.W. Blair, H.D. Upadhyaya, A.K. Are, and R. Ortiz. 2016.
Landrace germplasm for improving yield and abiotic stress adaptation. Trends in
Plant Science 21: 31–​42.
Elshafie, H.S., L. Aliberti, M. Amato, V. De Feo, and I. Camele. 2018. Chemical composition
and antimicrobial activity of chia (Salvia hispanica L.) essential oil. European Food
Research and Technology 244: 1675–​1682. doi: 10.1007/​s00217-​018-​3080-​x
Epstein, E., J.D. Norlyn, D.W. Rush, R.W. Kingsbury, D.B. Kelly, G.A. Cunningham,
and A.F. Wrona. 1980. Saline culture of crops: A genetic approach. Science 210:
399–​404.
Fan, M.S., F.J. Zhao, S.J. Fairweather-​Taitc, P.R. Poultona, S.J. Dunhama, and S.P. McGrath.
2008. Evidence of decreasing mineral density in wheat grain over the last 160 years.
Journal of Trace Elements in Medicine and Biology 22: 315–​324.
FAO (Food and Agriculture Organization). 2016. Save and Grow in Practice: Maize, Rice,
Wheat. A Guide To Sustainable Cereal Production. www.fao.org/​ag/​save-​and-​grow/​
MRW/​index​_​en.html.
FAO (Food and Agriculture Organization). 2019. FAO: Challenges and Opportunities in a
Global World. Rome. Licence: CC BY-​NC-​SA 3.0 IGO.
Farooq, M., M. Hussai, A. Wakeel, and K.H.M. Siddique. 2015. Salt stress in maize: effects,
resistance mechanisms, and management. A review. Agronomy for Sustainable
Development 35: 461–​481.
Farooq, M., N. Gogoi, M. Hussain, S. Barthakur, S. Paul, N. Bharadwaj, H.M. Migdadi,
S.S. Alghamdi, and K.H.M. Siddique. 2017. Effects, tolerance mechanisms and
management of salt stress in grain legumes. Plant Physiology and Biochemistry
118: 199–​217.
Farooq, M., M. Hussain, M. Usman, S. Farooq, S.S. Alghamdi, and K.H.M. Siddique. 2018.
Impact of abiotic stresses on grain composition and quality in food legumes. Journal
of Agricultural and Food Chemistry 66: 8887–​8897.
Fischer, S., R. Wilckens, J. Jara, and M. Aranda. 2013. Controlled water stress to improve
functional and nutritional quality in quinoa seed. Boletin Latinoamericano y del
Caribe de Plantas Medicinales y Aromáticas 12: 457–​468.
82 Environmental Controls over Seed Quality

Fischer, S., R. Wilckens, J. Jara, M. Aranda, W. Valdivia, L. Bustamante, F. Graf, and I. Obal.
2017. Protein and antioxidant composition of quinoa (Chenopodium quinoa Willd.)
sprout from seeds submitted to water stress, salinity and light conditions. Industrial
Crops and Products 107: 558–​564.
Flores, R. 2006. Evaluación preliminar agronómica y morfológica del germoplasma de
cañahua (Chenopodium pallidicaule Aellen) en la Estación Experimental Belén. Tesis
de Grado. La Paz, Bolivia: UMSA.
Fortmeier, R., and S. Schubert. 1995. Salt tolerance of maize (Zea mays L.): The role of
sodium exclusion. Plant Cell & Environment 18: 1041–​1047.
Gavicho-​Uarrota, V., R. Brasil-​Severino, and M. Maraschin. 2011. Maize landraces (Zea mays
L.): A new prospective source for secondary metabolite production. International
Journal of Agricultural Research 6 :218–​226.
Geerts, S., D. Raes, M. Garcia, J. Vacher, R. Mamani, J. Mendoza, R. Huanca, B. Morales, R.
Miranda, J. Cusicanqui, and C. Taboada. 2008. Introducing deficit irrigation to sta-
bilize yields of quinoa (Chenopodium quinoa Willd.). European Journal of Agronomy
28: 427–​436.
Gimplinger, D.M., G. Dobos, R. Schönlechner, and H.P. Kaul. 2007. Yield and quality of
grain amaranth (Amaranthus sp.) in eastern Austria. Plant Soil and Environment
53: 105–​112.
Gresta, F., M. Wink, U. Prins, M. Abberton, J. Capraro, A. Scarafoni, and G. Hill. 2017. Lupins
in European cropping systems. Legumes in Cropping System 88–​108.
Gross, R., E. Von Baer, R. Koch, L. Marquard, L. Trugo, and M. Wink. 1988. Chemical com-
position of a new variety of the Andean lupin (Lupinus mutabilis cv. Inti) with low
alkaloid content. Journal of Food Composition and Analysis 1: 353–​361.
Guilengue, N., S. Alves, P. Talhinhas, and J. Neves-​Martins. 2020. Genetic and genomic
diversity in a Tarwi (Lupinus mutabilis Sweet) germplasm collection and adaptability
to Mediterranean climate conditions. Agronomy 10: 21.
Gulisano, A., S. Alves, J.N. Martins and L.M. Trindade. 2019. Genetics and breeding of
Lupinus mutabilis: An emerging protein crop. Frontier in Plant Science 10: 1385.
Gutiérrez, L.F., L.M. Rosada, and A. Jiménez. 2011. Chemical composition of Sacha Inchi
(Plukenetia volubilis L.) seeds and characteristics of their lipid fraction. Grasas y
Aceites 62: 76–​83.
Hamaker, B.R., C. Valles, R. Gilman, R.M. Hardmeier, D. Clark, H.H. Garcia, A.E. Gonzales,
I. Kohlstad, M. Castro, R. Valdivia, T. Rodriguez, and M. Lescano. 1992. Amino acid
and fatty acid profiles of the inca peanut (Plukenetia volubilis). Cereal Chemistry
69: 461–​463.
Hardy, A., C. Huyghe, and J. Papineau. 1997. Dry matter accumulation and partitioning,
and seed yield in indeterminate Andean lupin (Lupinus mutabilis Sweet). Australian
Journal of Agricultural Research 48: 91–​102.
Hasegawa, P.M., R.A. Bressan, J.K. Zhu, and H.J. Bohnert. 2000. Plant cellular and molecular
response to high salinity. Annual Review of Plant Physiology and Plant Molecular
Biology 51: 463–​499.
Heuer, B., Z. Yaniv, and I. Ravina. 2002. Effect of late salinization of chia (Salvia hispanica),
stock (Matthiola tricuspidata) and evening primrose (Oenothera biennis) on their oil
content and quality. Industrial Crops and Products 15: 163–​167.
Hu, Y. and U. Schmidhalter. 2005. Drought and salinity: A comparison of their effects on
mineral nutrition of plants. Journal of Plant Nutrition and Soil 168: 541–​549.
References 83

Hussain, M.I., A.J. Al-​Dakheel, and M.J. Reigosa. 2018. Genotypic differences in agro-​
physiological, biochemical and isotopic responses to salinity stress in quinoa
(Chenopodium quinoa Willd.) plants: Prospects for salinity tolerance and yield sta-
bility. Plant Physiology and Biochemistry 129: 411–​420.
Hussain, M.I., A. Muscolo, M. Ahmed, M.A. Asghar, and A.J. Al-​Dakheel. 2020. Agro-​
morphological, yield and quality traits and interrelationship with yield stability in
quinoa (Chenopodium quinoa Willd.) Genotypes under saline marginal environ-
ment. Plants 9: 1763.
Hussain, T., I.A. Khan, M.A. Malik, Z. Ali. 2006. Breeding potential for high temperature tol-
erance in corn (Zea mays L.). Pakistan Journal of Botany 38: 1185–​1195.
Ignjatovic-​Micic, D., J. Vancetovic, D. Trbovic, Z. Dumanovic, M. Kostadinovic, and S.
Bozinovic. 2015. Grain nutrient composition of maize (Zea mays L.) Drought-​
tolerant populations. Journal of Agricultural Food Chemistry 63: 1251–​1260.
IPCC (Intergovernmental Panel on Climate Change). 2018. Progress Report of the Special
Report on Global Warming of 1.5°C. www.ipcc.ch/​site/​ass​ets/​uplo​ads/​2018/​04/​13022​
0180​459-​INF.6-​Repor​tSR-​15.pdf (accessed February 25, 2019).
IPCC (Intergovernmental Panel on Climate Change). 2021. Climate Change 2021: The
Physical Science Basis. www.ipcc.ch/​rep​ort/​ar6/​wg1/​ (accessed August 19, 2021.
IPGRI (International Plant Genetic Resources Institute), PROINPA (Promocion e
Investigacion de Productos Andinos), and IFAD (International Fund for Agricultural
Development). 2005. Descriptores para cañahua (Chenopodium pallidicaule Aellen).
Rome/​La Paz: IFAD; IPGRI; Fundación PROINPA.
Jacobsen, S.E. 1997. Adaptation of quinoa (Chenopodium quinoa) to Northern European
agriculture: Studies on developmental pattern. Euphytica 96: 41–​48.
Jacobsen, S.E., and A. Mujica. 2006. El tarwi (Lupinus mutabilis Sweet) y sus parientes
silvestres. Universidad Mayor de San Andrés, La Paz, 458–​482.
Jacobsen, S.E., and O. Stølen. 1993. Quinoa-​morphology, phenology and prospects for its
production as a new crop in Europe. European Journal of Agronomy 2: 19–​29.
Jacobsen, S.E, M. Sørensen, S.M. Pedersen, and J. Weiner. 2015. Using our
agrobiodiversity: Plant-​based solutions to feed the world. Agronomy for. Sustainable
Development 35: 1217–​1235.
Jamboonsri, W., T.D. Phillips, R.L. Geneve, J.P. Cahill, and D.F. Hildebrand. 2012. Extending
the range of an ancient crop, Salvia hispanica L. –​a new ω3 source. Genetic Resources
and Crop Evolution 59: 171–​178.
Jarvis, D., Y. Ho, D. Lightfoot, S. Schmöckel, B. Li, J. Theo, A. Borm, H. Ohyanagi, K. Mineta,
C.T. Michell, N. Saber, N.M. Kharbatia, R.R. Rupper, A.R. Sharp, N. Dally, B.A.
Boughton, Y.H. Woo, G. Gao, G.W.M. Schijlen, A.A. Momin, S. Negrao, S. Al-​Babili,
C. Gehring, U. Roessner, C. Jung, K. Murphy, S.T. Arold, T. Gojobori, C.G. van der
Linden, E.N. van Loo, E.N. Jellen, P.J. Maughan, and M. Tester. 2017. The genome of
Chenopodium quinoa. Nature 542: 307–​312.
Kahane, R., T.Hodgkin, H. Jaenicke, C. Hoogendoorn, M. Hermann, J.D.H. Keatinge, J.A
Hughes, S. Padulosi, and N. Looney. 2013. Agrobiodiversity for food security, health
and income. Agronomy for Sustainable Development 33: 671–​693.
Kang, Y.H., M. Chen, and S.Q. Wan. 2010. Effects of drip irrigation with saline water on waxy
maize (Zea mays L. var. ceratina Kulesh) in North China Plain. Agricultural Water
Management 97: 1303–​1309.
84 Environmental Controls over Seed Quality

Karimi, G., M. Ghorbanli, H. Heidari, R.A. Khavarinejadand, and M.H. Assareh. 2005. The
effects of NaCl on growth, water relations, osmolytes and ion content in Kochia
prostrate. Biologia Plantarum 49: 301–​304.
Kasoma, C., Shimelis, H., Laing, M.D., Shayanowako, A.I.T. and Mathew, I. 2021. Revealing
the genetic diversity of maize (Zea mays L.) populations by phenotypic traits and
DArTseq markers for variable resistance to fall armyworm. Genetic Resources and
Crop Evolution 68: 243–​259.
Keeler, R.F., Cronin, E.H., and Shupe, J.L. 1976. Lupin alkaloids from teratogenic and
nonteratogenic lupines. IV –​Concentration of total alkaloids, and the teratogen
anagyrine as a afunction of plant part and stage of growthand their relationship to
crooked calf disease. Journal of Toxicology Environmental and Health 1: 899–​908.
Kew, R.B.G. 2016. State of the World’s Plants Report—​ 2016. London: Royal Botanic
Gardens, Kew.
Khaitov, B., A.A. Karimov, K. Toderich, Z. Sultanova, A. Mamadrahimov, K. Allanov, and
S. Islamov. 2020. Adaptation, grain yield and nutritional characteristics of quinoa
(Chenopodium quinoa) genotypes in marginal environments of the Aral Sea basin.
Journal of Plant Nutrition 44: 1365–​1379.
Khalil, S.E. and E.G. Ismael. 2010. Growth, yield and seed quality of Lupinus termisas
affected by different soil moisture levels and different ways of yeast application.
Journal of American Science 6: 141–​153.
Khoshbakht, K., and K. Hammer. 2008. How many plant species are cultivated?. Genetic
Resources and Crop Evolution 55: 925–​928.
Kodahl, N. 2020. Sacha inchi (Plukenetia volubilis L.)—​from lost crop of the Incas to part of
the solution to global challenges? Planta 251: 80.
Konsens, I., M. Ofir, and J. Kigel. 1991. The effect of temperature on the production and
abscission of flowers and pods in snap-​bean (Phaseolus vulgaris L.). Annals of Botany
67: 391–​399.
Konuşkan, Ö., H. Gözübenli, I. Atiş, and M. Atak. 2017. Effects of salinity stress on emer-
gence and seedling growth parameters of some maize genotypes (Zea mays L.).
Turkish Journal of Agriculture 5: 1668–​1672.
Kucek, L.K., L.D. Veenstra, P. Amnuaycheewa, and M.E. Sorrells. 2015. A grounded
guide to gluten: How modern genotypes and processing impact wheat sensitivity.
Comprehensive Reviews in Food Science and Food Safety 14: 285–​302.
Kurlovich, B.S., A.K. Stankevich, and S.I. Stepanova. 2002. The history of lupin domestica-
tion. In Lupins: Geography, Classification, Genetic Resources and Breeding, edited by
B.S. Kurlovich, 147–​164. St. Petersburg: OY International Express.
Lawlor, D.W. and G. Cornic. 2002. Photosynthetic carbon assimilation and associated
metabolism in relation to water deficits in higher plants. Plant, Cell and Environment
25: 275–​294.
Laxa, M., M. Liebthal, W. Telman, K. Chibani, and K. J. Dietz. 2019. The role of the plant
antioxidant system in drought tolerance. Antioxidants 8: 94.
Lei, Y., Y. Zheng, and K. Dai. 2014. Different responses of photosystem I and photosystem
II in three tropical oilseed crops exposed to chilling stress and subsequent recovery.
Trees 28: 923–​933.
Lenné, J.M. and D. Wood. 2011. Agrobiodiversity Management for Food Security. 12–​26.
Wallingford, Oxon: CABI.
References 85

Lesjak, J. and D.F. Calderini. 2017. Increased night temperature negatively affects grain
yield, biomass and grain number in Chilean quinoa. Frontiers in Plant Science 8: 352.
Li, J., J. Chen, J. Jin, S. Wang, and B. Du. 2019. Effects of irrigation water salinity on maize
(Zea may L.): Emergence, growth, yield, quality, and soil salt. Water 11: 2095.
Li, Y.T., W.W. Xu, B.Z. Ren, B. Zhao, J. Zhang, P. Liu, and Z.S. Zhang. 2020. High temperature
reduces photosynthesis in maize leaves by damaging chloroplast ultrastructure and
photosystem II. Journal of Agronomy and Crop Science 206: 548–​564.
Liu, F. and H. Stützel. 2002. Leaf expansion, stomatal conductance, and transpiration
of vegetable amaranth (Amaranthus sp.) in response to soil drying. Journal of the
American Society of Horticultural Science 127: 878–​883.
Liu, F. and H. Stützel. 2004. Biomass partitioning, specific leaf area, and water use effi-
ciency of vegetable amaranth (Amaranthus spp.) in response to drought stress.
Scientia Horticulturae 102: 15–​27.
Lopez, C.M., M. Pineda, and J.M. Alamillo. 2020. Differential regulation of drought
responses in two Phaseolus vulgaris genotypes. Plants 9: 1815.
Lovelli, S., M. Valerio, T.D. Phillips, and M. Amato. 2019. Water use efficiency, photosyn-
thesis and plant growth of Chia (Salvia hispanica L.): A glasshouse experiment. Acta
Physiologiae Plantarum 41: 3.
Lugtenberg, B. and F. Kamilova. 2009. Plant-​growth-​promoting rhizobacteria. Annual
Reviews of Microbiology 63: 541–​556.
Luo, Y.L., Z.L. Su, T.J. Bi, X.L. Cui, and Q.Y. Lan. 2014. Salicylic acid improves chilling toler-
ance by affecting antioxidant enzymes and osmoregulators in sacha inchi (Plukenetia
volubilis). Brazilian Journal of Botany 37: 357–​363.
Machado, F.B., S. De, A.M. David, S.R. Dos Santos, J.C. Figueiredo, C.D. Da Silva, and D.A.
Nobre. 2020. Physiological quality of maize seeds produced under soil water deficit
conditions. Revista Brasileira de Engenharia Agricola e Ambiental 24: 451–​456.
Mangelson, H., D.E. Jarvis, P. Mollinedo, O.M. Rollano-​Penaloza, V.D. Palma-​Encinas, L.R.
Gomez-​Pando, E.N. Jellen, and P.J. Maughan. 2019. The genome of Chenopodium
pallidicaule: An emerging Andean super grain. Applications in Plant Sciences
7(11): e11300.
Martínez, E.A., E. Veas, C. Jorquera, R. San Martín, and P. Jara. 2009. Re-​introduction of
quinoa into arid Chile: Cultivation of two lowland races under extremely low irriga-
tion. Journal of Agronomy and Crop Science 195: 1–​10.
Matías, J., M.J Rodríguez, V. Cruz, P. Calvo, and M. Reguera. 2021. Heat stress lowers
yields, alters nutrient uptake and changes seed quality in quinoa grown under
Mediterranean field conditions. Journal of Agronomy and Crop Science. https://​doi.
org/​10.1111/​jac.12495.
Matsui, T., K. Omasa, and T. Horie. 2001. The difference in sterility due to high temperatures
during the flowering period among japonica-​rice varieties. Plant Production Science
4: 90–​93.
Matsuoka, Y., Y. Vigouroux, M.M. Goodman, J. Sanchez, E. Buckler, and J. Doebley. 2002.
A single domestication for maize shown by multilocus microsatellite genotyping.
Proceedings of the National Academy of Sciences 99: 6080–​6084. d
Mayes, S., F.J. Massawe, P.G. Alderson, J.A. Roberts, S.N. Azam-​Ali, and M. Hermann. 2012.
The potential for underutilized crops to improve security of food production. Journal
of Experimental Botany 63: 1075–​1079.
86 Environmental Controls over Seed Quality

Mi, N., F. Cai, Y.S Zhang, R.P. Ji, S.J. Zhang, and Y. Wang. 2018. Differential responses of maize
yield to drought at vegetative and reproductive stages. Plant Soil and Environment
64: 260–​267.
Moreno, C., C.E. Seal, and J. Papenbrock. 2017. Seed priming improves germination in
saline conditions for Chenopodium quinoa and Amaranthus caudatus. Journal of
Agronomy and Crop Science 204: 40–​48.
Muasya, R.M., W.J.M. Lommen, C.W. Muui, and P.C. Struik. 2008. How weather during
development of common bean (Phaseolus vulgaris L.) affects the crop’s maximum
attainable seed quality. NJAS –​Wageningen Journal of Life Sciences 56 : 85–​100.
Mujica, A. 1994. Andean grains and legumes. In Neglected Crops: 1492 from a Different
Perspective, edited by J. E. Hernández, and J. León, 141–​ 148. vol. FAO Plant
Production and Protection, Series, no. 26. Rome: Food and Agriculture Organization
of the United Nations.
Nicoletto, C., G. Zanin, P. Sambo, and D. L. Costa. 2019. Quality assessment of typical
common bean genotypes cultivated in temperate climate conditions and different
growth locations. Scientia Horticulturae 256: 108599.
Nodari, R.O., S.M. Tsai, R.L. Gilbertson, and P. Gepts.1993. Towards an integrated linkage
map of common bean: 2. Development of an RFLP-​based linkage map. Theoretical
and Applied Genetics 85: 513–​520.
Ortega-​David, E., A. Rodriguez, A. David, and A. Zamora-​Burbano. 2010. Caracterización de
semillas de Lupinus mutabilis sembrado en los andes de Colombia. Acta agronómica
59: 111–​118.
Paiva, E.P., S.B. Torres, T.R.C. Alves, F.V. da S. Sá, M. de S. Leite, and J.L.D. Dombroski. 2018.
Germination and biochemical components of Salvia hispanica; L. seeds at different
salinity levels and temperatures. Acta Scientiarum. Agronomy 40: e39396.
Pal, M. and T.N. Khoshoo. 1972. Evolution and improvement of cultivated Amaranths (v.
inviability, weakness and sterility in hybrids). J Heredity 63: 78–​82.
Peláez, P., D. Orona-​Tamayo, S. Montes-​Hernández, M. Valverde, O. Paredes-​López, and A.
Cibrian-​Jaramillo. 2019. Comparative transcriptome analysis of cultivated and wild
seeds of Salvia hispanica (chia). Scientific Reports 9: 9761.
Peñarrieta, J.M., J.A. Alvarado, B. Åkesson, and B. Bergenståhl. 2008. Total antioxidant
capacity and content of flavonoids and other phenolic compounds in canihua
(Chenopodium pallidicaule): An Andean pseudocereal. Molecular Nutrition and Food
Research 52: 708–​717.
Peñas, E., F. Uberti, C. di Lorenzo, C. Ballabio, A. Brandolini, and P. Restani. 2014.
Biochemical and immunochemical evidences supporting the inclusion of quinoa
(Chenopodium quinoa Willd.) as a gluten-​free ingredient. Plant Foods for Human
Nutrition 69: 297–​303.
Perales, H. and D. Golicher. 2014. Mapping the diversity of maize races in Mexico. Plos One
9: e114657.
Pokharel, M., A. Chiluwal, M. Stamm, D. Min, D. Rhodes, and S.V.K. Jagadish. 2020. High
night-​time temperature during flowering and pod filling affects flower opening,
yield and seed fatty acid composition in canola. Journal of Agronomy and Crop
Science 206: 579–​596.
Prasad, P., K. Boote, L. Allen Jr., J. Sheehy, and J. Thomas. 2006. Species, ecotype and cultivar
differences in spikelet fertility and harvest index of rice in response to high tempera-
ture stress. Field Crops Research 95: 398–​411.
References 87

Prasanna, B.M. 2012. Diversity in global maize germplasm: Characterization and utiliza-
tion. Journal of Biosciences 37: 843–​855.
Pulvento, C., M.H. Sellami, and A. Lavini. 2021. Yield and quality of Amaranthus
hypochondriacus grain amaranth under drought and salinity at various phenological
stages in southern Italy. Journal of the Science of Food Agriculture. doi: 10.1002/​
jsfa.11088.
Purseglove, J.W. 1968. Tropical Crops: Dicotyledons. London: Longmans, pp. 346–​381.
Qi, W., J.W. Zhang, K.J. Wang, P. Liu, and S.T. Dong. 2010. Effects of drought stress on the
grain yield and root physiological traits of maize varieties with different drought tol-
erance. The Journal of Applied Ecology 21: 48–​52.
Radić, V., D. Beatović, and J. Mrđa. 2007. Salt tolerance of corn genotypes (Zea mays
l.) during germination and later growth. The Journal of Agricultural Science
52: 115–​120.
Rainey, K.M., and P.D. Griffiths. 2005. Differential response of common bean genotypes
to high temperature. Journal of the American Society for Horticultural Science
130: 18–​23.
Razzaghi, F., S.H. Ahmadi, V.I. Adolf, C.R. Jensen, S.E. Jacobsen, and M.N. Andersen. 2011.
Water relations and transpiration of quinoa (Chenopodium quinoa Willd.) under
salinity and soil drying. Journal of Agronomy and Crop Science 197: 348–​360.
Reguera, M., C.M. Conesa, A. Gil-​Gómez, C.M. Haros, M.Á. Pérez-​Casas, V. Briones-​Labarca,
L. Bolaños, I. Bonilla, R. Álvarez, K. Pinto, Á. Mujica, and L. Bascuñán-​Godoy. 2018.
The impact of different agroecological conditions on the nutritional composition of
quinoa seeds. Peer J 6: e4442.
Repo-​Carrasco-​Valencia, R. and J.M. Vidaurre-​Ruiz. 2019. Quinoa and Other Andean
Ancient Grains: Super Grains for the Future. Cereal Foods World 6: 1–​10.
Repo-​Carrasco-​Valencia, R., A.A. de La Cruz, J.C. Alvarez, and H. Kallio. 2009. Chemical and
functional characterization of Kaiwa (Chenopodium pallidicaule) grain, extrudate
and bran. Plant Foods for Human Nutrition 64: 94–​101.
Repo-​Carrasco-​Valencia, R., J.K. Hellström, J.M. Pihlava, and P.H. Mattila. 2010. Flavonoids
and other phenolic compounds in Andean indigenous grains: Quinoa (Chenopodium
quinoa), kañiwa (Chenopodium pallidicaule) and kiwicha (Amaranthus caudatus).
Food Chemistry 120: 128–​133.
Reyes-​Caudillo, E., A. Tecante, M.A. Valdivia-​López. 2008. Dietary fibre content and anti-
oxidant activity of phenolic compounds present in Mexican chia (Salvia hispanica
L.) seeds. Food Chemistry 107: 656–​663.
Rezaei, E.E., S. Siebert, and F. Ewert. 2015. Intensity of heat stress in winter wheat—​
Phenology compensates for the adverse effect of global warming. Environmental
Research Letters 10: 024012.
Rodríguez, Á., M. Corazon-​Guivin, D. Cachique, K. Mejia, D. Castillo, J.F. Renno, and
C. Garcia-​Dávila. 2010. Diferenciación morfológica y por ISSR (Inter simple
sequence repeats) de especies del género Plukenetia (Euphorbiaceae) de la
Amazonía peruana: propuesta de una nueva especie. Revista Peruana de Biologia
17: 325–​330.
Roman, V.J., L.A. den Toom, C.C. Gamiz, N. van der Pijl, R.G. Visser, E.N. van Loo, and C.G.
van der Linden. 2020. Differential responses to salt stress in ion dynamics, growth
and seed yield of European quinoa varieties. Environmental and Experimental Botany
177: 104146.
88 Environmental Controls over Seed Quality

Ruiz, C., C. Díaz, J. Anaya, and R. Rojas. 2013. Análisis proximal, antinutrientes, perfil de
ácidos grasos y de aminoácidos de semillas y tortas de 2 especies de Sacha inchi
(Plukenetia volubilis y Plukenetia huayllabambana). Revista de la Sociedad Química
del Perú 79: 29–​36.
Ruiz, K.B., I. Aloisi, V. Canelo, S. Del Duca, P. Torrigiani, H. Silva, and S. Biondi. 2016. Salt
flat versus coastal ecotypes of quinoa: salinity responses in Chilean landraces from
contrasting habitats. Plant Physiology and Biochemistry 101: 1–​13.
Ruiz-​Carrasco, K., F. Antognoni, A.K. Coulibaly, S. Lizardi, A. Covarrubias, E.A. Marínez,
M.A. Molina-​Montenegro, S. Biondi, and A. Zurita-​Silva. 2011. Variation in salinity
tolerance of four lowland genotypes of quinoa (Chenopodium quinoa Willd.) as
assessed by growth, physiological traits, and sodium transporter. Plant Physiology
and Biochemistry 49: 1333–​1341.
Sage, R.F., T.L. Sage, R.W. Pearcy, and T. Borsch. 2007. The taxonomic distribution of
C4 photosynthesis in Amaranthaceae sensu stricto. American Journal of Botany
94: 1992–​2003.
Salgado-​Aguilar, M., T. Molnar, J.L. Pons-​Hernández, J. Covarrubias-​Prieto, J.G. Ramírez-​
Pimentel, J.C. Raya-​Pérez, S. Hearne, and G. Iturriaga. 2020. Physiological and bio-
chemical analyses of novel drought-​tolerant maize lines reveal osmoprotectant
accumulation at silking stage. Chilean Journal of Agricultural Research 80: 241–​252.
Sánchez, B., A. Rasmussen, and J.R. Porter. 2014. Temperatures and the growth and devel-
opment of maize and rice: A review. Global Change Biology 20: 408–​417. doi: 10.1111/​
gcb.12389.
Saucedo, A.L., E. Hernández-​Domínguez, L. Luna-​Valdez, A. Guevara-​García, A. Escobedo-​
Moratilla, E. Bojorquéz-​Velázquez, F. Río-​Portilla, D. Fernández-​Velasco, and A.P.
Barba de la Rosa. 2017. Insights on the structure and function of a Late Embryogenesis
Abundant protein from Amaranthus cruentus: An intrinsically disordered protein
involved in protection against desiccation, oxidant conditions, and osmotic stress.
Frontiers in Plant Sciences 8: 497.
Sauer, J.D. 1950a. Amaranths as dye plants among the pueblo peoples. South-​west. Journal
of Anthropological Sciences 6: 412–​415.
Sauer, J.D. 1950b. The grain amaranths: A survey of their history and classification. Annals
of the Missouri Botanical Garden 37: 561–​632.
Sehgal, A., K. Sita, K.H.M. Siddique, R. Kumar, S. Bhogireddy, R.K. Varshney, B.
HanumanthaRao, R.M. Nair., P.V.V. Prasad, and H. Nayyar. 2018. Drought or/​and
heat-​stress effects on seed filling in food crops: Impacts on functional biochemistry,
seed yields, and nutritional quality. Frontiers in Plant Science 9: 1705.
Shabala, S., Y. Hariadi, and S.E. Jacobsen. 2013. Genotypic difference in salinity tolerance
in quinoa is determined by differential control of xylem Na+​ loading and stomatal
density. Journal of Plant Physiology 170(10): 906–​914.
Shah, N. and G. Paulsen. 2003. Interaction of drought and high temperature on photosyn-
thesis and grain-​filling of wheat. Plant and Soil 257 : 219–​226.
Shonnard, G.C. and P. Gepts. 1994. Genetics of heat tolerance during reproductive develop-
ment in common bean. Crop Science 34: 1168–​1175.
Shrivastava, P. and R. Kumar. 2015. Soil salinity: A serious environmental issue and plant
growth promoting bacteria as one of the tools for its alleviation. Saudi Journal of
Biological Sciences 22: 123–​131.
References 89

Silva, A.N., G.M.L. Ramos, W.Q. Ribeiro, E. Rodrigues, P. Carvalho, C. Andrea, C. Cleo,
and M.A. Vanderlei. 2020. Water stress alters physical and chemical quality in
grains of common bean, triticale and wheat. Agricultural Water Management
231: 106023.
Simioniuc, D.P., V. Simioniuc, D. Topa, M. van den Berg, U. Prins, P.J. Bebeli, and I. Gabur.
2021. Assessment of Andean lupin (Lupinus mutabilis) genotypes for improved frost
tolerance. Agriculture 11: 155.
Sinclair, T.R., and R. Serraj. 1995. Legume nitrogen fixation and drought. Nature Cell Biology
378: 344.
Singh, S.P., P. Gepts, and D.G Debouck. 1991a. Races of common bean (Phaseolus vulgaris,
Fabaceae). Economic Botany 45: 379–​396.
Singh, S.P., R. Nodari, and P. Gepts. 1991b. Genetic diversity in cultivated common bean. I.
Allozymes. Crop Science 31: 19–​23.
Singh, S.P., J.A. Gutiérrez, A. Molina, C. Urrea, and P. Gepts. 1991c. Genetic diversity in
cultivated common bean. II. Marker–​based analysis of morphological and agro-
nomic traits. Crop Science 31: 23–​29.
Sogbohossou, E.O.D., E.G. Achigan-​Dako, P. Maundu, S. Solberg, E.M.S. Deguenon, R.H.
Mumm, I. Hale, A.V. Deynze, and M.E. Schranz. 2018. A roadmap for breeding orphan
leafy vegetable species: A case study of Gynandropsis gynandra (Cleomaceae).
Horticulture Research 5: 2.
Sozen, O., U. Karadavut, H. Ozcelik, H. Bozoglu, and M. Akcura. 2018. Genotype x envir-
onment interaction of some dry bean (Phaseolus vulgaris L.) genotypes. Legume
Research 41: 189–​195.
Suarez, J.C., J.A. Polanía, A.T. Contreras, L. Rodriguez, L. Machado, C. Ordoñez, S. Beebe,
and I M. Rao. 2020. Adaptation of common bean lines to high temperature
conditions: Genotypic differences in phenological and agronomic performance.
Euphytica 216: 1–​20.
Suarez-​Salazar, J.C., J.A., Polania, A.T. Contreras-​Bastidas, L. Rodríguez Suárez, S. Beebe,
and I.M. Rao. 2018. Agronomical, phenological and physiological performance of
common bean lines in the Amazon region of Colombia. Theoretical and Experimental
Plant Physiology 30: 303–​320.
Taibi, K., F. Taibi, L.A. Abderrahim, A. Ennajah, M. Belkhodja, and J.M. Mulet. 2016. Effect
of salt stress on growth, chlorophyll content, lipid peroxidation and antioxidant
defence systems in Phaseolus vulgaris L. South African Journal of Botany 105: 306–​312.
Tao, Z.Q., Y.Q. Chen, C. Li, J.X. Zou, P. Yan, S.F. Yuan, X. Wu, and P. Sui. 2016. The causes and
impacts for heat stress in spring maize during grain filling in the North China Plain
-​A review. Journal of Integrative Agriculture 15 : 2677–​2687.
Tian, Y.H., Y.B. Lei, Y. Zheng, and Z.Q. Cai. 2013. Synergistic effect of colonization with
arbuscular mycorrhizal fungi improves growth and drought tolerance of Plukenetia
volubilis seedlings. Acta Physiologiae Plantarum 35: 687–​696.
Turan, M.A., A.H.A. Elkarim, N. Taban, and S. Taban. 2010. Effect of salt stress on growth
and ion distribution and accumulation in shoot and root of maize plant. African
Journal of Agricultural Research 5: 584–​588.
Tsutsumi, N., M. Tohya, T. Nakashima, and O. Ueno. 2017. Variations in structural, bio-
chemical, and physiological traits of photosynthesis and resource use efficiency in
Amaranthus species (NAD-​ME-​type C4). Plant Production Science 20: 300–​312.
90 Environmental Controls over Seed Quality

Ulian T., M. Diazgranados, S. Pironon, S. Padulosi, L. Udayangani, L. Davies, M.J.R. Howes,


J.S Borrell, I. Ondo, O.A. Perez-​Escobar, S. Sharrock, P. Ryan, D. Hunter, M.A. Lee, C.
Barstow, L. Luczaj, E. Mattana, et al. 2020. Unlocking plant resources to support food
security and promote sustainable agriculture. Plants, People, Planet 2: 421–​445.
Vallebueno-​Estrada, M., I. Rodríguez-​Arévalo, A. Rougon-​Cardoso, J. Martínez González,
A. García Cook, F. Montiel, and J.P. Vielle-​Calzada. 2016. The earliest maize from San
Marcos Tehuacán is a partial domesticate with genomic evidence of inbreeding.
Proceedings of the National Academy of Sciences of the United States of America
113: 14151–​14156.
Wang, Q.M., Z.L. Huo, L.D. Zhang, J.H. Wang, and Y. Zhao. 2016. Impact of saline water
irrigation on water use efficiency and soil salt accumulation for spring maize in arid
regions of China. Agricultural Water Management 163: 125–​138.
Wang, Y., Y. Meng, H. Ishikawa, T. Hibino, Y. Tanaka, N. Nii, and T. Takabe. 1999.
Photosynthetic adaptation to salt stress in three-​color leaves of a C4 Amaranthus
tricolor. Plant and Cell Physiology 40: 668–​674.
Waqas, M.A., X. Wang, S.A. Zafar, M.A. Noor, H.A. Hussain, A.M. Nawaz, and M. Farooq.
2021. Thermal stresses in maize: Effects and management strategies. Plants 10: 293.
White, P.J., L.M. Pollak, and S.A. Duvick. 2007. Improving the fatty acid composition of corn
oil by using germplasm introgression. Lipid Technology 19: 35−38.
Wilson, J.H., M.S.J. Clowes, and J.C.S. Allison. 1973. Growth and yield of maize at different
altitudes in Rhodesia. Annals of Applied Biology 73: 11–​84.
Wortmann, C.S. 2006. Phaseolus vulgaris L. (common bean). In PROTA 1: Cereals and
Pulses/​Céréales et légumes secs, edited by M. Brink and G. Belay. Wageningen,
Netherlands: PROTA. https://​uses.plant​net-​proj​ect.org/​f/​index.php?title=​Phas​eolu​
s_​vu​lgar​is_​-​_​haric​ot_​s​ec_​(PROTA)&mobil​eact​ion=​togg​le_​v​iew_​desk​top) Accessed
December 07, 2022.
Wu, G., A.J. Peterson, C.F. Morris, and K.M. Murphy. 2016. Quinoa seed quality response to
sodium chloride and sodium sulfate salinity. Frontiers in Plant Science 7: 790.
Yamori, W., K. Hikosaka, and D.A. Way. 2014. Temperature response of photosynthesis in
C3, C4, and CAM plants: Temperature acclimation and temperature adaptation.
Photosynthesis Research 119: 101–​117.
Yang, H., T. Huang, M. Ding, D. Lu, and W. Lu. 2017. High temperature during grain filling
impacts on leaf senescence in waxy maize. Agronomy Journal 109: 906–​916.
Ye, X., M. Zhang, M. Zhang, and Y Ma. 2020. Assessing the performance of maize (Zea
mays L.) as trap crops for the management of sunflower broomrape (Orobanche
cumana Wallr.). Agronomy10: 100.
Zhang, G.L., L.Y. Chen, S.T. Zhang, G.H. Liu, W.B. Tang, Z.Z. He, and M. Wang. 2007. Effects
of high temperature on physiological and biochemical characteristics in flag leaf of
rice during heading and flowering period. Scientia Agricultura Sinica 40: 1345–​1352.
Zheng, Y., M. Xu, R. Shen, and S. Qiu. 2013. Effects of artificial warming on the structural,
physiological, and biochemical changes of maize (Zea mays L.) leaves in northern
China. Acta Physiologia Plantarum 35: 2891–​2904.
Zou, C., A. Chen, L.H. Xiao, H.M. Muller, P. Ache, G. Haberer, M.L. Zhang, W. Jia, P. Deng,
R. Huang, D. Lang, F. Li, D.L. Zhan, X.Y. Wu, H. Zhang, J. Bohm, R.Y. Liu, S. Shabala,
R. Hedrich, J.K. Zhu, and H. Zhang. 2017. A high-​quality genome assembly of quinoa
provides insights into the molecular basis of salt bladder-​based salinity tolerance
and the exceptional nutritional value. Cell Research 27: 1327–​1340.
Chapter 3

The Outlook for Latin-​American


Crops
Challenges and Opportunities

Nieves Fernández-​García1, Inmaculada


Román-​García1, and Enrique Olmos1
1Centro de Edafología y Biología Aplicada del Segura -​

Consejo Superior de Investigaciones Cientifícas


(CEBAS-​CSIC). Espinardo, Murcia, Spain.

CONTENTS
3.1 Introduction 92
3.2 Quinoa (Chenopodium quinoa Willd) 94
3.2.1 Quinoa Production and Market 95
3.2.2 Present and Future Uses of Quinoa 95
3.3 Amaranthus: Kiwicha (Amaranthus caudatus L.), and
Amaranth (A. cruentus L. and A. hypochondriacus L.) 98
3.3.1 Present and Future Uses of Amaranthus 98
3.3.2 New Regions of Cultivation 99
3.4 Chia (Salvia hispánica L.) 100
3.4.1 State of the Art in Chia Crops 100
3.4.2 New Regions of Cultivation 102
3.4.3 Present and Future Uses of Chia 103
3.5 Tarwi (Lupinus mutabilis) 104
3.6 Sacha Inchi (Plukenetia volubilis L.) 105
3.7 Maize (Zea mays L., Variety Purple) 106
3.8 Kaniwa (Chenopodium pallidicaule Aellen) 107
3.9 Common Bean (Phaseolus vulgaris) 108

DOI: 10.1201/9781003088424-3 91
92 Latin-American Crops’ Challenges and Opportunities

3.10 Conclusions 109


Acknowledgments 110
References 110

3.1 INTRODUCTION
There is no single, easy solution to world hunger and poverty. Proposals for the
future of the European Commission’s common agricultural and food policy
stress the reduction in meat consumption and production as a starting point
(Schiermeier, 2019), given that a growing share of grain production is used for
animal feed.
Of the thousands of known edible plant species, only three cereals—​
maize, wheat and rice, as the main cultivated plants in the world—​contribute
to nearly 60% of the calories and proteins consumed by humans from
vegetables (FAO, 2000). These three crops are considered the backbone of
agriculture, playing a critical role in global food provision (FAO, 2000, 2016,
2017).
This dependence on just a few crops makes current cropping systems
highly vulnerable to projected impacts of climate change and accompanying
extreme weather events. Climate change poses a considerable threat to
global food security, so expanding our food sources by integrating so-​called
emerging or alternative crops can be a sustainable solution. Thousands of
neglected and underutilized species with balanced amino acid and micro-
nutrient profiles are known of, which can help address climate change and
reduce hunger. These crops include quinoa, amaranth, chia, sacha inchi,
tarwi and kaniwa. Implementation of these underutilized crops could pro-
vide us with a more diversified agricultural system and improve nutritional
quality and climate resilience of future crop systems. Currently, these emer-
ging crops are considered “the food of the future for humanity” by the Food
and Agriculture Organization of the United Nations (FAO) and the World
Health Organization (WHO), as they are able to replace animal proteins
and provide bioactive compounds; they also contain gluten-​free proteins
(FAO, 2017).
Latin-​American crops such as quinoa, amaranth and chia are alternative,
profitable and strategic crops. Healthy lifestyles and healthy eating are on the
rise, and we are increasingly concerned about what we eat, filling our shopping
baskets with more real food and less ultra-​processed food. This trend, which
is currently associated with healthy and organic diets, has led to a notable
increase in the sale of these “superfoods” of Latin-​American origin in recent
years. These foods are highly appreciated for their nutritional properties,
3.1 Introduction 93

Population (billion)

2100 10.9

2050 9.7
Year

2030 8.5

2017 7.8

1990 5.3

0 2 4 6 8 10 12

Figure 3.1 World Population 2019: Wall Chart (ST/​ ESA/​ SER.A/​434). Revision
Produced by United Nations Department of Public Information. Source: United Nations,
Department of Economic and Social Affairs, Population Division (2019). World
Population Prospects -​Population Division -​United Nations. https://​pop​ulat​ion.
un.org/​wpp/.​

and some are consumed as alternatives to animal products like milk (OECD/​
FAO, 2019).
On the other hand, the FAO estimates that by 2050 there will be 2 billion more
inhabitants on the planet, and that this increase will be accompanied by a pro-
gressive but rapid loss of arable land (Figure 3.1). What agrifood measures are on
the horizon to meet global food demand? Could Latin-​American food systems
feed the world (Pastor-​Pazmiño, 2017)?
In addition, the world’s potential land for food production is threatened by
land degradation, affected by soil erosion, depletion and contamination due to
both climate change and human misuse (EEA, 2019). This situation represents a
major loss that will translate into a loss of global capacity to produce food and
supply our growing world population. In order to be able to go into the future
with the certainty that we will have food for the entire population, the vast
majority of specialists considers precision agriculture to be necessary. When
and where to sow, and what to sow, are two of the essential questions yet to be
answered (Castillo, 2015).
These degraded lands are likely to end up in a process of desertification. For
this reason, there is an obligation to find alternative crops that are able to both
94 Latin-American Crops’ Challenges and Opportunities

cover nutritional demand and grow in adverse soils. Crops like quinoa, amaranth
and chia are able to grow in adverse areas with conditions such as saline soils,
high temperatures, etc. That is why we can ask ourselves the question: Could
Latin-​American crops feed the world?
The outstanding properties of Andean crops are the main reason why their
consumption has been maintained for millennia in certain civilizations, from
ancient times to their expansion to global recognition today. Many of these crops
were underestimated by Spanish conquerors, who considered them as irrelevant
foods, and were thus replaced by crops with which they were already familiar,
such as wheat (Atchison, 2016).
Within the Latin-​American crops considered “superfoods”, in addition to
quinoa, maize, chia and amaranth, we also find crops that are less well known,
such as sacha inchi, black bean or tarwi. These lesser known crops are also
valued for their high percentage of proteins, high levels of essential amino acids
and high concentrations of antioxidants, and some of them are also gluten free.

3.2 QUINOA (CHENOPODIUM QUINOA WILLD)


Quinoa (Chenopidium quinoa Willdenow), of the Chenopodiaceae family and
known as quinua (Quechua), jopa (Aymara), suba (Chibcha), quinhua (Mapuche)
and quinoa (Spanish), is native to the Andes. It originated in Peru and Bolivia,
around Lake Titicaca, and was probably domesticated in several areas simultan-
eously (Vietmeyer, 1989). With approximately 7,000 years of cultivation, quinoa
is one of the oldest crops in the Andean region, where ancient civilizations such
as the Tiahuanacota and Inca contributed to its domestication and conserva-
tion throughout time (Jacobsen, 2003). Quinoa and potatoes seem to have been
the staple foods of many ancient highland civilizations. On the arrival of the
Spaniards, quinoa was replaced by cereals despite being a local staple food at
the time.
On archeological digs, quinoa has been discovered in tombs in Tarapacá,
Calama and Arica in Chile, and in different regions of Peru (Tapia, 2014). At the
time of the Spanish arrival, quinoa was well developed technologically and was
widely distributed within and beyond Inca territory. The first Spaniard to note
the cultivation of quinoa was Pedro de Valdivia who, on noticing the planted
crops around Concepción, recorded that the native Indians also sowed quinoa
among other plants for food (Tapia, 2014).
Quinoa is an annual broad-​leaved, dicotyledonous herb usually standing
about 1-​3 m high (Mujica, 1992). The woody central stem is either branched or
unbranched depending on the genotype, sowing density and environmental
conditions, and it may be green, red or purple. The inflorescences emerge from
3.2 Quinoa (Chenopodium quinoa Willd) 95

the higher part of the plant or from the axils of the stem, and they have a central
axis from which a secondary (amarantiform) and/​or tertiary axis with flowers
(glomeruliform) emerges. Depending on the cultivar, the seeds are on average
about 2 mm in diameter and they come in a range of different colors; they can be
white, red or even black (Rojas, 2003).

3.2.1 Quinoa Production and Market


The consumption of quinoa has increased significantly in recent years, mainly
in China and EU countries such as France, Germany and Spain. According to
a Euromonitor study (Mascaraque, 2018), consumption of quinoa and other
Andean grains is mainly trend driven and health motivated.
Quinoa production showed a remarkable increase after 2013, which was
declared the year of quinoa by the FAO (Bazile et al., 2015). Worldwide quinoa
production increased steadily until 2015, then declined and was maintained in
the following years, possibly due to the fall in the price of quinoa seeds in 2015
(Figure 3.2).
The main producers worldwide are Peru, Bolivia and Ecuador (Figure 3.3).
Strikingly, quinoa production in Spain has grown exponentially in recent years: the
country went from no production 10 years ago to being the main producer in
Europe by 2019 (17,217 ton). Currently, according to the Spanish Ministry of
Agriculture (MAPA, 2020), 6,638 ha are dedicated to quinoa cultivation in Spain,
and Andalusia is the main production region (93% of the total).

3.2.2 Present and Future Uses of Quinoa


Archaeological and ethnobotanical data show that quinoa has been cultivated
in the Andean regions of South America for thousands of years. Quinoa remains
have been found in various contexts, including burial grounds, hearths, storage
structures, human digestive tracts and coprolites (Lopez et al., 2011). This
indicates that quinoa was not used only for food, but also in religious rites.
Quinoa is traditionally consumed in three different forms: (1) as a whole seed,
(2) in soups or (3) as pitu; a kind of toasted refined flour. It is mainly consumed in
the form of seed and soup in the Andean regions and as pitu in the Plateau. Both
the seed and quinoa flour have also traditionally been used for bread making and
fried or cooked products (Angeli et al., 2020). Another ancient way of consuming
quinoa is in puffed form. To make puffed quinoa, the seed is treated through an
extrusion process using high temperatures and pressure. Flour can also be used
to make pasta, chips and tortillas ( flatbreads). Moreover, in recent years, the
consumption of sprouts has increased exponentially. Sprouting seeds are now
common on supermarket shelves, and they are appreciated as a rich source of
newgenrtpdf
96
Latin-American Crops’ Challenges and Opportunities
Quinoa worlwide production 2010 -2019 Average price U.S. dollars per kg 2010-2019
250000 8
6.74
Production in metric tons

7
193822
200000 186147
6
159647 161415
148720 146735
150000 5 4.36
118175 3.58
4
97862 3.15 3.21
2.96
100000 80069 84644 3 2.32
1.66 1.78
2
50000
1

0 0
2010 2011 2012 2013 2014 2015 2016 2017 2018 2019 2010 2011 2012 2013 2014 2017 2018 2019 2020

Figure 3.2 (left panel) Worldwide evolution of production). Source: FAO, 2019.
Figure 3.2 (right panel) Price of quinoa between 2010 and 2019. Source: Prospectiva, 2020; FAO Worldwide, 2010–​2019; Tridge blog
©Statista, 2021.
3.2 Quinoa (Chenopodium quinoa Willd) 97

Quinoa production worldwide 2010-2019, by country


250000
Production in metric tons

200000

150000

100000

50000

0
2010 2011 2012 2013 2014 2015 2016 2017 2018 2019

Bolivia Ecuador Peru

Figure 3.3 Main quinoa producers worldwide: Peru, Bolivia and Ecuador. Source:
Worldwide; FAO, 2010 to 2019. Data were obtained from Statista. www.fao.org/​faos​
tat/​en/​#data/​QCL.

healthy phytochemicals and for their sensory characteristics. They are rich in
nutrients like minerals, amino acids and vitamins, in addition to nutraceuticals,
and they contain low levels of anti-​nutritional substances like tannin, lectin
and galactoside compared to non-​germinated seeds. The most common form of
sprout consumption is in salads as a food supplement.
Nowadays, a large number of sprouting seeds are consumed, such as sun-
flower, broccoli and chia. Nutritional analysis of sprouts shows that sprouting
quinoa seeds are being consumed in countries like Peru. On the other hand,
quinoa seeds can be fermented to obtain beer, which is gluten-​free and therefore
suitable for consumption by people with coeliac condition. The food industry
is producing more and more quinoa-​derived products, including pasta and
breakfast products rich in quinoa, as well as snacks, energy bars and ready-​to-​
eat meals.
New uses for quinoa include the extraction of starch granules and their
application in emulsifying systems. The starch granule is a biopolymer that is
abundant in the seeds and inexpensive to extract. It is both biodegradable and
biocompatible, which makes it a suitable candidate to replace synthetic and inor-
ganic particles used as Pickering stabilizers. Among the starch-​based Pickering
emulsions, octenylsuccinate quinoa starch (OSQS) shows interesting emulsi-
fying properties, mainly due to the small size of the starch granule compared to
that obtained from other species (Li et al., 2020).
98 Latin-American Crops’ Challenges and Opportunities

3.3 AMARANTHUS: KIWICHA (AMARANTHUS


CAUDATUS L.), AND AMARANTH (A. CRUENTUS L.
AND A. HYPOCHONDRIACUS L.)
The genus Amaranthus is considered an ancient crop that has been cultivated
from 5,000–​7,000 years ago (Sauer, 1967). It comprises approximately 60 species
that grow in many regions of the world (Saunders and Becker, 1984). Most
amaranth species are native to America, mainly to Mesoamerica and South
America, with only 15 species originating in Asia, Africa, Australia and Europe
(Martinez-​Lopez et al., 2020). The main species cultivated for the characteristic
of the seeds are A. caudatus, A. cruentus and A. hypochondriacus, which have
been domesticated in the Andean region of South America and Mesoamerica,
respectively (Sauer, 1967). These amaranth species are herbaceous annual
plants that were domesticated in prehistoric times (Sauer, 1976). In Tehuacan,
Puebla, Mexico, archaeological remains were found indicating that A. cruentus
was already being cultivated around 4000 BC, and that A. hypochondriacus was
grown around AD 500 (Sauer, 1976; Jacoben and Mujica, 2003). The oldest arch-
aeological evidence of A. caudatus dates back 2,000 years and was found in nor-
thern Argentina (Hurnziker and Planchuelo, 1971). The inflorescence consists of
large branches of solid red, yellow, green, or variable colors. Likewise, the seeds
are extremely variable in color. Amaranth has small seeds of between 1 and
1.5 mm in diameter.
Interest in amaranth has increased progressively since the 1980s, when
the US National Academy of Sciences developed a research program entitled
“Underexploited Tropical Plants with Promising Economic Value”, which
included 36 promising species with the greatest potential for cultivation
(Berghofer and Schoenlechner, 2002; Caselato-​Sousa and Amaya-​Farfan, 2012;
Orona-​Tamayo and Paredes-​López, 2017; Soriano-​Garcia et al., 2018). Since
then, a major effort has been made to study the nutritional and antioxidant
characteristics of amaranth and its cultivation possibilities in different parts of
the world (Park et al., 2020).

3.3.1 Present and Future Uses of Amaranthus


Amaranth is a versatile seed, and is considered one of the most nutritious
pseudo-​cereals that is high in gluten-​free protein. Amaranth is also used in
sweets or drinks, in fresh leaf salads, popped seeds, and as a source of natural
red dye, animal feed and for obtaining oils and cosmetic products (López-​Mejía
et al., 2014; Repo-​Carrrasco-​Valencia, 2011). The Amaranth seed can be roasted
or milled into flour, and it can be consumed on its own or included in other crop
products such as bread, cakes, muffins, pancakes, biscuits, dumplings, noodles,
pasta and crackers (Cardenas-​Hernandez, 2016).
3.3 Amaranthus: Kiwicha (Amaranthus caudatus L.) 99

Moreover, the incorporation of amaranth flour in pasta processing has


been found to produce a low level of stickiness that is more desirable for pasta
products (Chillo et al., 2008; Fiorda et al., 2013). Guardianelli et al. (2021) have
studied mixing amaranth flour from germinated and non-​germinated seeds with
wheat flour for bread. Wheat flour withstood the inclusion of 25% germinated
amaranth seeds without substantial changes in bread quality. Additionally,
roasted amaranth flour has the potential to replace wheat flour in the prep-
aration of acceptable gluten-​free noodles (Beniwal et al., 2019). Singh and Liu
(2021) compared noodles made from amaranth flours with wheat flour noodles
and demonstrated that roasting improved cooking results and the texture of
cooked noodles.
Amaranth is also consumed as a vegetable throughout Asia, Africa, the
Caribbean, Greece and Mexico. Amaranth leaves are usually picked fresh for use
as greens in salads or steamed, boiled or fried in oil. Cooked greens can be used
as a side dish, in soups or as an ingredient (FAO, 2020).

3.3.2 New Regions of Cultivation


Currently, amaranth is widely cultivated and consumed throughout India, Nepal,
China, Indonesia, Malaysia and the Philippines in Asia; throughout the whole of
Central America and Mexico; and in South and East Africa. Different species of
amaranth are cultivated in different countries. A. hypochondriacus is cultivated in
some countries of South America, Mexico and Asia, while A. caudatus is cultivated
in Argentina, Peru and Bolivia, and A. cruentus is cultivated in Guatemala.
The African continent represents one of the poorest and most malnourished
regions on the planet. African nations in this continent need to develop new
strategies in order to eradicate poverty and child malnutrition in the future.
One of the possibilities is to pay greater attention to underutilized crop species
that have great potential for influencing and improving food security for African
nations and promoting sustainable rural growth and development (Aderibigbe
et al., 2020). Amaranth is an underutilized crop that has great potential in Africa,
given its stress-​tolerant characteristics and ability to grow under extreme situ-
ations, which can help to mitigate the effects of climate change (Aderibigbe et al.,
2020; Alemayehu et al., 2015; Dizyee et al., 2020). This crop was introduced in
sub-​Saharan Africa in the 20th century, mainly A. hypochondriacus, A. cruentus,
A. caudatus, and A. dubius (Dinssa et al., 2016). Farmers in this part of the world
use it for both its leaf and seed. Available data from Tanzania and Kenya show
that 25,548 ha were cultivated in 2017, with a production of approximately 70,000
tons (Ochieng et al., 2019).
The area under kiwicha (amaranth) cultivation in Peru ranged from 2,300 ha
in 2015 to 1,400 ha in 2017, with an average yield of 1,885 kg ha-​1 (Valvas et al.,
2020). Recently, the Ministry of Agriculture and Irrigation of Peru (MINAGRI)
100 Latin-American Crops’ Challenges and Opportunities

has developed a new variety of kiwicha, INIA 442 LA FRONDOSA, resistant to


different pests and able to grow at altitudes of 2,500 to 3,000 m above sea level.
This new variety will increase production to 3.5 ton per ha, increasing the profit-
ability of the crop (MINAGRI). In Mexico, amaranth production in 2019 was about
6,000 tons, much less than those 7,115 produced in 2018 (Panorama Alimentario,
2020). The main producing area in the country is Puebla, accounting for about
half of the amaranth produced. The area dedicated to amaranth cultivation in
Mexico in 2016 was 4,545 ha, and Puebla was the main state with the largest area
dedicated to this crop.

3.4 CHIA (SALVIA HISPÁNICA L.)


Chia (Salvia hispánica L.), of the Lamiaceae family, is native to the central valleys
of Mexico and northern Guatemala, where species of the Labiatae family are
concentrated. Chia seeds began to be used in human food around 3500 BC and
acquired importance as a staple crop in central Mexico between 1500 and 900 BC
(Cahill, 2003; Pelaez et al., 2019). It was one of the main crops of pre-​Columbian
societies, and Aztecs and Mayas used the seeds for the preparation of various
medicines, foods and paintings. The Aztecs received chia seeds as annual
tributes from the people under their rule, and it was used as an offering to the
gods in religious ceremonies (Beltran-​Orozco et al., 2003; Muñoz et al., 2013).
Carl Linnaeus classified chia in 1753. Salvia is a Spanish word meaning “cure,”
and in Latin, hispanica means Spanish; Salvia hispánica thus means “Spanish
plant used to cure or save” (Urbina, 1887; Sharma and Mogra, 2019). The word
“chia” (meaning “oily”) is a Spanish adaptation of chian, or chien in its plural
form, which has its origin in the Nahuatl language of the Aztecs. Chia has been
cultivated since ancient times on the banks of the Grijalva River, the ancient
territory of the Nahuatl Chiapan, which is now the state of Chiapas in modern
Mexico; the plant took its name from the Grijalva River, which means “river of
chia” (Munoz et al., 2013).
Chia (Salvia hispanica L.) is an annual herb that blooms during the summer
months (Alfaro and Silva, 2013; Munoz et al., 2013) and grows up to 1–​2 m in
height. The seeds are oval, smooth and shiny, and mottled with brown, gray, dark
red and white; they are generally found in groups of four. The plant has ribbed and
hairy quadstems (Ayerza and Coates, 2005; Di Sapio et al., 2012; Pelaez et al., 2019;).

3.4.1 State of the Art in Chia Crops


The chia seed has great nutritional qualities. It is rich in antioxidants and an
excellent source of fiber; it has a high protein content and high levels of omega-​3,
3.4 Chia (Salvia hispánica L.) 101

50.0
42
39
40.0 37 36

28
30.0

19 19 17
20.0 16 17
13 13

10.0 06 05
04 03 02 02 02 02 02
02 04 03
0.0
2011 2012 2013 2014 2015 2016 2017 2018 2019

1000 tonnes millions euros value euro/kg

Figure 3.4 Indicative chia imports to Europe from the main producing and supplying
countries: Paraguay, Bolivia, Argentina, Peru, Mexico, Uganda, Chile and Australia.
Source: Eurostat, Market Access Database; The European market potential for chia
seeds CBI.

and is rich in minerals, among other features (Kulczynski et al., 2019; Orona-​
Tamayo et al., 2019). Although the health-​promoting and biologically active
characteristics of chia have been known for many years, it is only in the last few
years that interest in chia consumption has grown from a scientific and com-
mercial point of view (Teoh et al., 2018). The chia market in Europe has been
increasing significantly since 2012. As can be seen in Figure 3.4, chia imports
have grown fourfold in the last eight years.
In recent years, Paraguay has been the main producer of chia in the world
with more than 20,000 tons, which represents almost 50% of the world’s pro-
duction. Other important chia producers are Mexico, Bolivia, Argentina, Peru,
Chile, Uganda and Australia. The world’s largest importer is the United States
(Figure 3.5). Europe’s largest importers of chia are the Netherlands and Germany,
accounting for importations of over 50% (Figure 3.5).
Chia has been considered a novel food by the EU, following the regulation
stipulating that any food that was not consumed to a significant degree within
the European Union before May15, 1997 is to be considered novel. Since January
1, 2018, this region has drawn up a list of authorized novel foods and their
requirements, including chia seeds and chia-​derived oil (Kulczynski et al., 2019).
The following uses for chia are permitted in the EU: (1) as an ingredient in food
in pastry, bread products, yogurts, fruit jellies and dressings; (2) as a nutritional
supplement; and (3) as a base for beverages.
102 Latin-American Crops’ Challenges and Opportunities

Chia seeds import value in 2020


180 168.88
Import value in million U.S. dollars

150

120 112.17

90
62.42 57.52
60 51.37
40.48 36.51 33.04 29.27
30 18.83

Figure 3.5 Chia seeds import value in 2020. Source: Worldwide; UN Comtrade, 2020.
https://​comtr​ade.un.org/​data/​.

3.4.2 New Regions of Cultivation


Chia is a short-​day flowering plant, sensitive to frost at all stages of development
(Bochicchio et al., 2015). The induction of flowering and the transition between
vegetative and generative phases of plant growth are highly dependent on the
length of the day (<12 hours). These characteristics mean that chia cultivation
is limited to latitudes below 25o, close to the equator. Chia grows naturally in
tropical and subtropical areas, with optimum growth between 400 and 2,500 m
above sea level; below 200 m, it does not develop adequately (Orozco et al., 2014).
Chia requires temperatures between 11 and 36oC, and its optimum range is 16–​
26oC (Ayerza and Coates, 2009). Chia grows best in sandy, well-​drained soils
with moderate salinity and with a pH ranging from 6 to 8.5 (Yeboah et al., 2014).
Currently, the main cultivation areas for chia are Paraguay, Mexico, Argentina,
Bolivia, Australia, Colombia, Guatemala and Peru. Chia cultivation is extending
to other regions, however, and in Europe, trials are being carried out in Greece,
France, Germany and Italy (Bochicchio et al., 2015; Gravé et al., 2019; Grimes
et al., 2020; Tabaxi et al., 2016;). Although the results are interesting, it is clear
that there is a need to develop new varieties that can grow in days with a longer
light cycle. Breeders have recently succeeded in developing varieties with low
day-​length dependence, allowing for growth in 12–​15h light conditions (Gravé
et al., 2019). This opens up possibilities for cultivation in regions such as Europe
(Grimes et al., 2020). According to our knowledge, a new variety, named ORURO,
3.4 Chia (Salvia hispánica L.) 103

has been introduced in southwestern France at the Regional Centre for Organic
Agriculture, in Auch. A stable and homogeneous variety selected to bloom in
long summer days, ORURO chia seeds flower early, which allows for produc-
tion in temperate regions (Gravé et al., 2019). In other Latin-​American coun-
tries, such as Chile, the possibility of introducing chia cultivation is also being
evaluated (Baginsky et al., 2016; Silva et al., 2018).

3.4.3 Present and Future Uses of Chia


Nowadays, chia seeds are used whole, ground, and in the form of gel and oil
(Muñoz-​Gonzalez et al., 2019; Zettel and Hitzmann, 2018;). In recent years, chia
gel has been used in the production of various food products, such as low-​fat
sausages (Camara et al., 2020; Pintado et al., 2020). Chia gel has also been used as
an egg substitute in chocolate cakes or as partial fat replacement in pork burgers
(Gallo et al., 2020; Lucas-​Gonzalez et al., 2020). The use of chia gels in these
food products not only reduces the fat and calorie content, but also increases
the amounts of beneficial fatty acids contained in chia oils, such as omega-​3.
However, this substitution should not exceed 25% of the oil or egg in cakes, as
it may affect different product qualities like texture, color or flavor (Borneo
et al., 2010).
Chia flour has also been used to partially replace wheat flour. Borneo et al.
(2010) observed that replacing 7.5% of wheat flour with chia flour was the most
appropriate proportion in terms of nutritional capabilities and organoleptic
qualities. Constatini et al. (2014) observed that replacing 10% of wheat flour
with chia flour had no effect on the final flavonoid content, but it did increase
fat, moisture, polyphenols and dietary fiber, while reducing carbohydrate con-
tent. Other authors have observed that the substitution of 7.5% wheat flour with
ground chia seeds negatively affects the sensory qualities of bread and reduces
consumer acceptance (Kowalski et al., 2020). However, 5% substitution with
ground chia seeds was not found to affect the quality of the bread produced,
which had a higher polyunsaturated fatty acid to saturated fatty acid ratio than
regular wheat bread. The addition of 5% chia flour to white bread can increase
the dietary fiber content by up to 50% (Švec and Hrušková, 2015). Chia flour
can also be used to increase the nutritional quality of gluten-​free foods such as
bread, as these tend to be poorer in minerals and fats (Pellegrini and Agostoni,
2015). Chia flour has also been used in the production of pasta. Oliveira et al.
(2015) observed that the addition of chia flour in pasta production significantly
improved the protein, fiber and mineral content. Similarly, Cota-​Gastelum et al.
(2019) observed that the preparation of pasta with a mixture of wheat semo-
lina (25%), chia flour (10%) and chickpea flour (35%) improved the nutritional
qualities of the pasta as well as its antioxidant properties, increasing consumer
acceptability.
104 Latin-American Crops’ Challenges and Opportunities

The inclusion of chia in dairy products has also been increasing in recent
years. For example, ice cream has been obtained by partially replacing the milk
fat with the olein fraction of chia oil (Ullah et al., 2017). The addition of chia
enhanced the concentration of omega-​3 fatty acids and improved the antioxi-
dant perspectives of ice cream. Chia seed mucilage has also been used as a sta-
bilizer and emulsifier in ice cream (Campos et al., 2016). Chia mucilage has been
added to yogurts, as well, replacing fats and increasing nutritional value (Ribes
et al., 2021). The addition of 2% chia oil to natural yogurts resulted in higher
amounts of unsaturated fatty acids, especially linoleic and alpha-​linolenic acids,
and a high phytosterol content (Derewiaka et al., 2019).
This seed is also a promising source of bioactive peptides which exert bio-
logical functions to promote health, as treatment for or prevention of some
diseases (improvement of the digestive system, stronger bones and muscles,
reduction of risk of heart disease) (Grancieri et al., 2019; Orona-​Tamayo et al.,
2019; Quintal-​Bojórquez et al., 2021). Benefits for the human body of bioactive
peptides are mainly due to the composition of amino acids present in each pep-
tide and the different reactivity. Actually, the most studied properties of bioactive
peptides in human health are about anticancer, antimicrobial, antihypertensive,
antioxidant, anticholesteolemic, anti-​inflammatory and immunomodulatory
diseases (Orona-​Tamayo et al., 2019; Santiago-​Lopez et al., 2016).

3.5 TARWI (LUPINUS MUTABILIS)


Tarwi, a leguminous species (Lupinus mutabilis Sweet) of the Fabaceae family,
also known as tauri, chocho or Andean Lupin, is native to the Andes of Peru,
Bolivia and Ecuador. It has been cultivated in the Andean area since pre-​Incan
times. The main characteristics are the high protein, fat contents and its bitter
taste due to anti-​nutritional alkaloids that make processing necessary before
consumption in seeds; and the plant capacity of fixing atmospheric nitrogen in
the soil due to the association with rhizobia and its ability to adapt to adverse
climatic conditions with minimal demands on soil quality (Gutierrez, 2016).
The wide genetic diversity of tarwi means that there is great variability in its
adaptation to different environmental conditions (soil, temperature, precipi-
tation, etc.). This genetic diversity also means that the seeds’ concentration of
proteins, alkaloids, oils, among other components, also vary widely (Jacobsen
and Mujica, 2006).
Tarwi is considered to be the “Andean soybean” thanks to its high levels of
proteins and oils, macronutrients that can be used in the formulation of various
functional foods to prevent metabolic diseases. Proteins and oils represent more
than half of the seed’s weight. Some of the fatty acids present in tarwi seeds are
3.6 Sacha Inchi (Plukenetia volubilis L.) 105

essential fatty acids for our organism, i.e., we cannot synthesize them ourselves,
so must acquire them in our diet (Jacobsen and Mujica 2006).
Tarwi is an excellent potential source for obtaining protein isolates and bio-
active compounds for the functional food and nutraceutical industry (Jacobsen
and Mujica, 2006; Chirinos-​Arias, 2015; Tapia, 2015). Protein is the most abun-
dant and important macronutrient in this legume (46–​48%), with levels similar to
those found in soybeans; moreover, their concentration can increase after defat-
ting, reaching values of 47–​64% and a digestibility of over 90% (Schoeneberger
et al., 1982; Intiquilla-​Quispe, 2018). This extract can be used to obtain protein
concentrates and isolates for use in gastronomy, the food industry, livestock
feed, and others (Tapia, 2015). Tarwi seeds are raw materials with a high indus-
trial potential in the production of food products such as pastes, flours and oils.
Tarwi paste is used to make drinks, yogurt and jams, while the flour is mainly
used in the production of products like breads, cakes, biscuits, cookies, noodles
and instant soups.
Seed from this crop also has a high fat content, which makes it viable for
industrial oil production; its use in human consumption makes it necessary to
obtain treated grains free of the toxic substances they contain (Ruiz and Sotelo,
2001; Resta et al., 2008).
Flours after protein and oil extraction can be a basic component for the
manufacture of products for the food industry and an efficient use of the crop
(Sosa, 2000; Rosell et al., 2009; Morales et al., 2012).
Unfortunately, this species has not been given the importance it deserves,
partly due to the scarce dissemination of its nutraceutical properties and the
presence of alkaloids that give the seeds a bitter taste, limiting its consumption.
For such reasons, a future objective should be to disseminate the nutraceutical
and medicinal properties of “Andean Lupin” in order to arouse the interest of
other countries in its study and diffusion.

3.6 SACHA INCHI (PLUKENETIA VOLUBILIS L.)


Sacha inchi (Plukenetia volubilis L.), of the Euphorbiaceae family, is also known
as sacha peanut, mountain peanut, Inca nut or Inca peanut (Wang et al.,
2018). Sacha inchi is native to the Peruvian Amazon, and it is also found in the
Colombian, Bolivian and Brazilian Amazon. The fruits of the plant are star-​
shaped where seeds are contained (Kumar et al., 2017, Vasquez-​Osorio et al.,
2017). These seeds show high levels of fat, 41–​43%, and more calories than sugar;
they are also a good source of vitamins, proteins and minerals; better than
most meat products, so they could be a good substitute for meat consumption
(Chirinos et al., 2013; Alvarado, 2014).
106 Latin-American Crops’ Challenges and Opportunities

Andean populations have long consumed this seed in traditional preparations


such as soups, biscuits and children’s food. This is especially true in Peru,
where they obtain oil and flour by various methods, which are used to prepare
food, drinks and snacks that provide omega-​3 and omega-​6 (Vásquez-​Osorio
et al., 2017).
Oil has long been used for food by various ethnic groups in Latin America;
it is now obtaining international recognition for its healthy properties, espe-
cially as an extraordinary source of polyunsaturated fatty acids, one of the most
prospective nutraceutical sources to promote healthy nutrition with extraor-
dinary benefits for human health. Currently, the protein isolate obtained from
the residue after oil extraction has high protein levels (of ∼59%); this residue
may be used in other potential products and it can be further used in a high
number of products. Using waste or by-​products obtained in the preparation of
sacha inchi oil to produce protein isolates can contribute to human health in
the fight against malnutrition, and it can also contribute to global knowledge on
underutilized crops.

3.7 MAIZE (ZEA MAYS L., VARIETY PURPLE)


Maize, or corn (Zea mays L.) is native to the Americas. It began its journey
c.10,000 years ago someplace in Mesoamerica, and from there it spread to the
whole continent. During the 16th and 17th centuries, maize cultivation was
widespread throughout the world, as Spaniards and other Europeans exported
maize from the Americas to Europe. At present, this crop is the third most
important crop (after wheat and rice), and it is produced in most countries of
the world. Maize has great variability in the color of the grain, including white,
yellow, black, purple, blue, red and orange in different varieties (Arteaga et al.,
2016; Kistler et al., 2018).
Zea mays L., a purple variety (purple corn or maize), originated in the Andean
region of what is now Peru, and it has been widely cultivated and consumed
throughout the Andean region of South America, mainly in Peru, Ecuador, Bolivia
and Argentina (Lao et al., 2017). Purple corn has purple episperm seeds (grains)
and cobs, which gives the pigments special characteristics: between 1.5% and
6.0% of the pigments are anthocyanins, which belong to the group of flavonoids
including cianidine-​3-​Glucoside (Guillén-​ Sánchez et al., 2014). Although it
has been consumed in Andean regions for millennia for its many properties,
purple corn is not well known in Europe. It is considered a superfood due to
its high anthocyanin content, and it is grown sustainably and exported all over
the world. Purple corn not only possesses high concentrations of antioxidant
pigments, but also of fiber. Overall, this crop is nutritionally and compositionally
3.8 Kaniwa (Chenopodium pallidicaule Aellen) 107

rich, being similar in the mineral and vitamin content to the yellow corn but
showing higher levels of easily digestible carbohydrates (Ai and Jane, 2016; Lao
et al., 2017). It should be noted that the presence of anthocyanins and other
phenolic compounds that differentiate purple maize from other conventional
maize varieties make it stand out as a good food source with healthy benefits.
However, the information currently available on its benefits is still limited and
more research would be needed to understand its contribution to human health
(Lao et al., 2017).
Consumers can find it canned or vacuum-​packed, although it can also be
consumed in the form of flour or in food supplements.

3.8 KANIWA (CHENOPODIUM PALLIDICAULE AELLEN)


Kaniwa (Chenopodium pallidicaule Aellen), of the Chenopodiaceae family, is also
known as kañiwa, kañawa and cuchi-​quinoa, among other names. The origin
of this pseudo-​cereal is uncertain, but it is almost certainly an Andean native.
Kaniwa is closely related to quinoa, and it was considered a variety of quinoa
until 1929 (Repo-​Carrrasco-​Valencia, 2011). Kaniwa is cultivated in the Andes
of Peru and Bolivia and is characterized by its tolerance to low temperatures; it
grows in extreme environments where other species such as wheat and maize
cannot grow. This crop grows in the extreme altiplano environment, in the
highest areas of Peru and Bolivia on the banks of the Titicaca River, at an altitude
of between 3,800 and 4,100 m above sea level where the average temperature
is less than 10ºC. Here, other staple crops considered essential for global food
security, such as wheat, barley, even quinoa, do not grow as reliably as this plant
(Popenoe, 1989). It is a very important crop for highland farmers; when other
crops fail due to frost, kaniwa still provides food (Repo-​Carrasco-​Valencia, 2011).
Unlike quinoa, kaniwa has very low saponin levels, and it can be used directly
without any prior treatment (Repo-​Carrasco-​Valencia et al., 2009). Its seed does
not contain gluten, so it is of interest to people with coeliac disease; it is also
highly nutritious and rich in nutrients like calcium and iron, and high in protein.
Fortunately, it has a healthy amino acid profile as well.
In kaniwa, proteins (15–​19%) are the second most abundant nutrient after
carbohydrates, and they are of better quality than the proteins found in other
cereals due to a balanced portion of amino acids and the presence of those is
recommended for children and adults (Repo-​Carrasco-​Valencia, 2011). Moreover,
these proteins are particularly rich in lysine (5–​6%), isoleucine and tryptophan
(3.4 and 0.9 g/​100 g of protein, respectively) (Repo-​Carrasco-​Valencia et al.,
2003; Gallego-​Villa et al., 2014). They are the perfect complement to daily protein
intake.
108 Latin-American Crops’ Challenges and Opportunities

The high content of proteins, and outstanding levels of some key amino acids
of these seeds, provide a basis for considering kaniwa as an important crop for
current and future food security

3.9 COMMON BEAN (PHASEOLUS VULGARIS)


Studies about the origin of the common bean indicate a Mesoamerican origin
for these species from which different migratory events extended the distri-
bution of Phaseolus vulgaris towards South America. Their hypothesis on the
origin of the common bean indicates that, from a central area on the western
slopes of the Andes in northern Peru and Ecuador, wild beans dispersed north-
wards (Colombia, Central America and Mexico) and southwards (southern
Peru, Bolivia and Argentina), giving rise to Mesoamerican and Andean gen-
etic variability, respectively (Bitocchi et al., 2003; Chacón-​Sanchez et al., 2005;
Kaplan, 1965).
The common bean, of family Fabaceae, which has been an important source
of livelihood in the Americas since it was first cultivated more than 7,000 years
ago, is now considered one of the most important food legumes, mainly in Africa
and Latin America (Broughton et al., 2003). This legume is often grown by small-​
scale farmers under limiting conditions. Among these limiting conditions is
abiotic stress caused by drought, which currently affects more than 50% of the
world’s bean-​growing areas and can cause production losses of between 10% and
100%. However, according to Polania et al. (2016), a genetic improvement for the
environment for drought resistance (the main abiotic stress factor limiting bean
cultivation) would be in the selection of morphophysiological characteristics in
outstanding genotypes.
On the other hand, this legume –​being a source of bioactive compounds with
the potential to contribute to healthier diets –​is considered a nutraceutical or
functional food (Bressani, 1993). Its antioxidant capacity is determined by the
content of compounds such as polyphenols (including flavonoids), anthocyanins,
and condensed tannins (which contribute to the color of the bean) (Bressani
et al., 1961; Chávez-​Mendoza et al., 2017; Singh et al., 2021).
Although beans are traditionally included in the diet in many Latin-​American
countries and are beneficial for health and disease prevention, a considerable
amount of research is needed to characterize the very diverse genotypes in terms
of their content of bioactive compounds and their effect on consumer health
(Acevedo et al., 1994; Calderon et al., 1992; Suarez-​Martinez et al., 2016). Chavez-​
Mendoza et al. (2017) stressed the need to study wild-​type genotypes for their
nutraceutical qualities to be used as a valuable resource in breeding programs
and strategies to promote their consumption.
3.10 Conclusions 109

Common beans are considered one of the main sources of protein in the
diets of many countries. Cultivation of this legume is mainly carried out by
smallholders under limiting environmental conditions (Buruchara et al., 2011;
Beebe et al., 2013; Keller et al., 2020: Mukankusi et al., 2018). In addition to seed
yield losses related to the challenging environmental conditions where beans
are cultivated, storage of common beans under adverse conditions (including
high temperatures and high humidity) renders them susceptible to a hardening
phenomenon, also known as the hard-​to-​cook (HTC) defect (Bressani, 1993;
Reyes-​Moreno et al., 1993). The HTC defect diminishes the nutritional prop-
erties of this seed due to extension of cooking time affecting the complemen-
tary nutritional role of this legume as well as in diets when it is combined with
cereals. This effect is especially relevant in countries where diet is based on
legumes (as are common beans) and cereals (maize, wheat, or rice). Processing
methods such as soaking, germination, fermentation, and cooking have been
reported as detoxifying the legume seed. Also, soaking prior to cooking softens
the seeds, significantly reducing cooking time (Bressani, 1993; Reyes-​Moreno
et al., 1993).
The importance of considering the cooking time and its relationship with
bean quality has resulted in the development of breeding programs aimed at
improving this qualitative aspect (Carvalho et al., 2017). Incorporating cooking
quality into cultivar selection will allow improvement in the length of cooking
time, thus preserving the nutritional quality of the seeds (Ribeiro et al., 2014).

3.10 CONCLUSIONS
Climate change is a threat to our world pantry, making food security at risk. It is
therefore necessary to transform our food systems by unlocking underutilized
plant resources to support and improve food security, promoting sustainable
agriculture and affordable healthy diets for all.
In the next few years, preservation of biodiversity will be essential. In this
regard, sustainable agricultural productivity is going to be required to meet
growing demand for food that should possess high nutritional quality.
Most of the agricultural species described in this book are not considered
to be among the main staple crops in the world and could be categorized as
neglected and underutilized species. Establishment of these underutilized crops
may contribute to the FAO’s goal of a “world free from hunger and malnutrition,
where food and agriculture contribute to improving the standard of living for all”
in an economical, social, and environmentally sustainable manner, improving
food security and providing essential nutrients thus complementing the current
staple crops (wheat, corn, and rice).
110 Latin-American Crops’ Challenges and Opportunities

In this chapter, we have tried to highlight the agronomic, technological, and


health aspects of some Latin-​American crops such as chia, amaranth, sacha
inchi, tarwi, common bean, maize, kaniwa, and quinoa which could offer natural
food resources with high nutritional value. Indeed, their contribution to food
security has been increasingly recognized worldwide, for present and future
generations. These species are underutilized food crops that indigenous people
have protected and preserved using their traditional knowledge and practices.

ACKNOWLEDGMENTS
This work was supported by “Proyectos estrategicos Ris3Mur (Ref: 2I18SAE00057);
Consejeria de Empleo, Universidades, Empresa y Medio Ambiente de la Región
de Murcia”.

REFERENCES
Acevedo, E., Velazquez-​Coronado, L. and Bressani, R. 1994. Changes in dietary fiber con-
tent and its composition as affected by processing of black beans (Phaseolus vul-
garis, Tamazulapa variety). Plant Foods for Human Nutrition 46: 139–​145.
Aderibigbe, O.R., Ezekiel, O.O., Owolade, S.O., Korese, J.K., Sturm, B. and Henseld, O. 2020.
Exploring the potentials of underutilized grain amaranth (Amaranthus spp.) along
the value chain for food and nutrition security: A review. Critical Reviews in Food
Science and Nutrition DOI: 10.1080/​10408398.2020.1825323.
Ai, Y. and Jane, J. 2016. Macronutrients in corn and human nutrition. Comprehensive
Reviews in Food Science and Food Safety 15 : 581–​598.
Alemayehu, F.R., Bendevis, M.A. and Jacobsen, S.E. 2015. The potential for utilizing the seed
crop amaranth (Amaranthus spp.) in East Africa as an alternative crop to support
food security and climate change mitigation. Journal of Agronomy and Crop Science
201: 321–​329.
Alfaro, F. and Silva, H. 2013. Determinación de umbrales de respuestas fisiológicas e
identificación de mecanismos de tolerancia al déficit hídrico en cuatro accesiones de
Chía (Salvia hispánica L.). Santiago:
Alvarado-​Quiroz, K.D. 2014. Obtención, caracterización fisicoquímica, caracterización
electroforética y digestibilidad del aislado proteico del residuo agroindustrial
de Plukenetia volubilis (Sacha Inchi). Tesis de grado Arequipa-​Perú, Universidad
Católica de Santa María.
Angeli, V., Silva, P.M., Massuela, D.C., Khan, M.W., Hamar, A., Khajehei, F., Graeff-​Hönninger,
S. and Piatti, C. 2020. Quinoa (Chenopodium quinoa Willd.): An overview of the
potentials of the “Golden Grain” and socio-​economic and environmental aspects of
its cultivation and marketization. Foods 9: 216.
Arteaga, M.C., Moreno-​Letelier, A., Mastretta-​Yanes, A., Vázquez-​Lobo, A., Breña-​Ochoa,
A., Moreno-​Estrada, A., Eguiarte, L.E. and Piñero, D. 2016. Genomic variation in
recently collected maize landraces from Mexico. Genomics Data 7: 38–​45.
References 111

Atchison, G.W., Nevado, B., Eastwood, R.J., Contreras-​Ortiz, N., Reynel, C., Madriñán, S.,
Filatov, D.A. and Hughes, C.E. 2016. Lost crops of the Incas: Origins of domestica-
tion of the Andean pulse crop tarwi, Lupinus mutabilis. American Journal of Botany
103(9): 1592–​1606.
Ayerza, R. and Coates, W. 2005. Chia: Rediscovering a forgotten crop of the Aztecs. Tucson,
AZ: University of Arizona. www.jstor.org/​sta​ble/​j.ctv​29sf​ps7.
Ayerza, R. and Coates, W. 2009. Influence of environment on growing period and yield,
protein, oil and α-​linolenic content of three chia (Salvia hispanica L.) selections.
Industrial Crops and Products 30: 321–​324.
Baginsky, C., Arenas, J., Escobar, H., Garrido, M., Valero, N., Tello, D., Pizarro, L., Valenzuela,
A., Morales, L. and Silva, H. 2016. Growth and yield of chia (Salvia hispanica L.) in the
Mediterranean and desert climates of Chile. Chilean Journal of Agricultural Research
76: 255–​264.
Bazile, D. and Baudon, F. 2015. The dynamics of the global expansion of quinoa growing in
view of its high biodiversity. In State of the Art Report of Quinoa in the World in 2013,
edited by D. Bazile, D. Bertero and C. Nieto, 42–​55. Rome: FAO & CIRAD.
Beebe, S., Rao, I., Mukankusi, C., and Buruchara, R. 2013. Improving resource use effi-
ciency and reducing risk of common bean production in Africa, Latin America and
the Caribbean. In Eco-​Eefficiency: From Vision to Reality, edited by C. Hershey and
P. Neate, 117–​134. Colombia: CIAT.
Beltrán-​Orozco, M.C. and Romero, M.R. 2003. La Chía, Alimento Milenario. Mexico:
Departamento de Graduados e Investigación en Alimentos, www.jstor.org/​sta​ble/​
j.ctv​29sf​ps7. Escuela Nacional de Ciencias Biológicas. (ENCB), Instituto Politécnico
Nacional (IPN).
Beniwal, S.K., Devi, A. and Sindhu, R. 2019. Effect of grain processing on nutritional and
physico-​chemical, functional and pasting properties of amaranth and quinoa flours.
Indian Journal of Traditional Knowledge 18: 500–​507.
Berghofer, E. and Schoenlechner, R. 2002. Grain Amaranth. In Pseudocereals and Less
Common Cereals, 219–​260. Berlin: Springer.
Bitocchi, E., Laura Nannia, L., Belluccia, E., Rossia, M., Giardinia, A., Zeulib, P.S., Logozzob,
G., Stougaardc, J., McCleand, P., Attenee, G. and Papaa, R. 2003. Mesoamerican origin
of the common bean (Phaseolus vulgaris L.) is revealed by sequence data. Proceedings
of the National Academy of Sciences 109(14): E788–​796.
Bochicchio, R., Rossi, R., Labella, R., Bitella, B., Permiola, M. and Amato, M. 2015. Effect
of sowing density and nitrogen top-​dress fertilization on growth and yield of Chia
(Salvia hispanica L.) in a Mediterranean environment. Italian Journal of Agronomy
10: 163–​166.
Borneo, R., Aguirre, A. and León, A.E. 2010. Chia (Salvia hispanica L) gel can be used as egg
or oil replacer in cake formulations. Journal of the Academy of Nutrition and Dietetics
110: 946–​949.
Bressani, R. 1993. Grain quality of common beans. Food Reviews International 9(2):
237–​297.
Bressani, R., Elias, L.G. and Navarrete, D.A. 1961. Nutritive value of Central American beans.
IV. The essential amino acid content of samples of black beans, red beans, rice beans,
and cowpeas of Guatemala. Journal of Food Science 26(5): 457–​555.
Broughton, W.J., Hernández, G., Blair, M., Beebe, S., Gepts, P.and Vanderleyden, J. 2003.
Beans (Phaseolus spp.) –​model food legumes. Plant and Soil 252: 55–​128.
112 Latin-American Crops’ Challenges and Opportunities

Buruchara, R., Chirwa, R., Sperling, L., Mukankusi, C., Rubyogo, J.C., Muthoni, R. and Abang,
M.M. 2011. Development and delivery of bean varieties in Africa: The Pan-​Africa
Bean Research Alliance (PABRA) model. African Crop Science Journal 19(4): 227–​245.
Cahill, J. 2003. Ethnobotany of chia, Salvia hispanica L. (Lamiaceae). Economic Botany
57: 604–​618.
Calderon, E., Velasquez, L. and Bressani, R.1992.Comparative-​study of the chemical-​
composition and nutritive-​value of scarlet beans (Phaseolus-​Coccineus) and common
beans (Phaseolus-​Vulgaris). Archivos Latinoamericanos de Nutricion 42(1): 64–​71.
Camara, A.K.F.I., Vidal, V.A.S., Santos, M., Bernardinelli, O.D., Sabadini, E. and Pollonio,
M.A.R. 2020. Reducing phosphate in emulsified meat products by adding chia (Salvia
hispanica L.) mucilage in powder or gel format: A clean label technological strategy.
Meat Science 163: 108085.
Campos, B.E., Ruivo, T.D., da Silva Scapim, M.R., Madrona, G.S. and Bergamasco, R.C. 2016.
Optimization of the mucilage extraction process from chia seeds and application in
ice cream as a stabilizer and emulsifier. Lebensmittel-​Wissenschaft und Technologie
[Food Science and Technology] 65: 874–​883.
Cardenas-​Hernandez, A., Beta, T., Loarca-​Pina, G., Castano-​Tostado, E., Nieto-​Barrera, J.O.
and Mendoza S. 2016. Improved functional properties of pasta: enrichment with
amaranth seed flour and dried amaranth leaves. Journal of Cereal Science 72: 84–​90.
Carvalho, B.L., Patto-​Ramalho, M.A., Cunha-​Vieira, I. and Barbosa-​Abreu, A. F. 2017. New
strategy for evaluating grain cooking quality of progenies in dry bean breeding
programs. Crop Breeding and Applied Biotechnology 17: 115–​123.
Caselato-​Sousa, V.M. and Amaya-​Farfan, J. 2012. State of knowledge on amaranth grain: A
comprehensive review. Journal of Food Science 77(4): 93–​104.
Castillo, J. 2015. Quién nos va a alimentar en el 2050? Foro Económico Mundial [World
Economic Forum] (weforum.org).
Chacón-​Sánchez, M.I., Pickersgill, B. and Debouck, D.G. 2005. Domestication patterns
in common bean (Phaseolus vulgaris L.) and the origin of the Mesoamerican and
Andean cultivated races. Theoretical Applied Genetics 110(3): 432–​444.
Chávez-​Mendoza, C. and Sánchez, E. 2017. Bioactive compounds from Mexican var-
ieties of the common bean (Phaseolus vulgaris): Implications for health. Molecules
22(8): 1360.
Chillo, J., Laverse, P., Falcone, M. and Del Nobile A. 2008. Quality of spaghetti in base
amaranthus wholemeal flour added with quinoa, broad bean and chick pea. Journal
of Food Engineering 84(1): 101–​107.
Chirinos, R., Zuloeta, G., Pedreschi, R., Mignolet, E., Larondelle, Y. and Campos, D. 2013.
Sacha inchi (Plukenetia volubilis): A seed source of polyunsaturated fatty acids,
tocopherols, phytosterols, phenolic compounds and antioxidant capacity. Food
Chemistry 141: 1732–​1739.
Chirinos-​Arias, M.C. 2015. Andean Lupin (Lupinus mutabilis Sweet) a plant with nutra-
ceutical and medicinal potential Tarwi (Lupinus mutabilis Sweet) una planta con
potencial nutritivo y medicinal. Revista Bio Ciencias 3 (3): 163–​172.
Constantini, L., Lukšič, L., Molinari, R., Kreft, I., Bonafaccia, G., Manzi, L. and Merendino, N.
2014. Development of gluten-​free bread using tartary buckwheat and chia flour rich
in flavonoids and omega-​3 fatty acids as ingredients. Food Chemistry 165: 232–​240.
Cota-​Gastelum, A.G., Salazar-​Garcia, M.G., Espinoza-​Lopez, A., Perez-​Perez, L.M., Cinco-​
Moroyoqui, F.J., Martinez-​Cruz, O., Wong-​Corral, F.J. and Del-​Toro-​Sanchez, C.L.
2019. Characterization of pasta with the addition of Cicer arietinum and Salvia
References 113

hispanica flours on quality and antioxidant parameters. Italian Journal of Food


Science 31: 626–​643.
Derewiaka, D., Stepnowska, N., Brys, J., Ziarno, M., Ciecierska, M. and Kowalska, J. 2019.
Chia seed oil as an additive to yogurt. Grasa y Aceites 70: e302.
Di Sapio, O., Bueno, M., Busilacchi, H., Quiroga, M. and Severin, C. 2012. Caracterización
morfoanatómica de hoja, tallo, fruto y semilla de Salvia hispánica L. (Lamiaceae).
Latin American and Caribbean Bulletin of Medicinal and Aromatic Plants 11: 249–​268.
Dinssa, F.F., Hanson, P., Dubois, T., Tenkouano, A., Stoilova, T., Hughes, J. and Keating,
J.D.H. 2016. AVRDC—​the World Vegetable Center’s women-​oriented improvement
and development strategy for traditional African vegetables in sub-​Saharan Africa.
European Journal of Horticultural Science 81: 91–​105.
Dizyee, K., Baker, D., Herrero, M., Burrow, H., McMillan, L., Sila, D.N. and Rich, K.M. 2020.
The promotion of amaranth value chains for livelihood enhancement in East
Africa: A systems modelling approach. African Journal of Agricultural and Resource
Economics –​AFJARE 15: 81–​94.
EEA (European Environment Agency). 2019. Soil, land and climate change. www.eea.eur​
opa.eu/​sign​als/​sign​als-​2019-​cont​ent-​list/​artic​les/​soil-​land-​and-​clim​ate-​cha​nge.
FAO (Food and Agriculture Organization). 2000.Website for Agrobiodiversity: www.fao.
org/​biodi​vers​ity/​index.asp?lang=​en.
FAO (Food and Agriculture Organization). 2010–​2019. Worldwide. Rome: FAO. www.stati​
sta.com/​sta​tist​ics/​486​442/​glo​bal-​qui​noa-​pro​duct​ion/​.
FAO (Food and Agriculture Organization). 2016. Save and Grow in Practice: Maize, Rice and
Wheat. Rome: FAO.
FAO (Food and Agriculture Organization). 2017. The Future of Food and Agriculture –​Trends
and Challenges. Rome: FAO.
FAO (Food and Agriculture Organization). 2020. The Best Thing about Fruits and Vegetables?
Their Diversity! Five Lesser Known but Surprisingly Nutritious Fruits and Vegetables.
www.fao.org/​fao-​stor​ies/​arti​cle/​en/​c/​1364​251.
Fiorda, F.A., Soares Jr., M.S., da Silva, F.A., Grosmann, M.V. and Souto, L.R. 2013.
Microstructure, texture and colour of gluten-​free pasta made with amaranth flour,
cassava starch and cassava bagasse. Food Science and Technology 54(1): 132–​138.
Gallego-​Villa, D.Y., Russo, L., Kerbab, K., Landi, M. and Rastrelli, L. 2014. Chemical and nutri-
tional characterization of Chenopodium pallidicaule (cañihua) and Chenopodium
quinoa (quinoa) seeds. Emirates Journal of Food and Agriculture 26(7): 609.
Gallo, L.R.D., Botelho, R.B.A., Ginani, V.C., de Oliveira, L.D., Riquette, R.F.R. and Leandro,
E.D. 2020. Chia (Salvia hispanica l.) gel as egg replacer in chocolate cakes: Applicability
and microbial and sensory qualities after storage. Journal of Culinary Science and
Technology 18: 29–​39.
Grancieri, M., Martino, H.S.D., and Gonzalez de Mejia, E. 2019. Chiaseed (Salvia hispanica
L.) as a source of proteins and bioactive pep-​tides with health benefits: A review.
Institute of Food Technologists 18(2): 480–​.499.
Gravé, G., Mouloungui1, Z., Poujaud, F., Cerny, M., Pauthe, C., Koumba, I.S: Diakaridja,
N. and Merah, O. 2019. Accumulation during fruit development of components of
interest in seed of Chia (Salvia hispanica L.) cultivar Oruro© released in France.
Oilseeds and Fats Crops and Lipids 26(50): 1–​7.
Grimes, S.J., Capezzone, F., Nkebiwe, P.M. and Grae -​Hönninger, S. 2020. Characterization
and evaluation of Salvia hispanica L. and Salvia columbariae Benth. Varieties for
their cultivation in Southwestern Germany. Agronomy 10: 2012.
114 Latin-American Crops’ Challenges and Opportunities

Guardanielli, L.M., Salinas, M.V. and Puppo, M.C. 2021. Quality of wheat breads enriched
with flour from germinated amaranth seeds. Food Science and Technology
International. https://​doi.org/​10.1177/​108201​3221​1016​577.
Guillén-​Sánchez, J., Mori-​Arismendi. S. and Paucar-​Menacho, L.M. 2014. Características
y propiedades funcionales del maíz morado (Zea mays L.) var. subnigroviolaceo
Characteristics and functional properties of purple corn (Zea mays L.) var.
Subnigroviolaceo. Scientia Agropecuaria 5: 211–​217.
Gutierrez, A., Infantes, M., Pascual, G. and Zamora, J. 2016. Evaluación de los factores en el
desamargado de tarwi (Lupinus mutabilis Sweet). Agroindustrial Science 6: 145–​149.
Hunziker, A. T. and Planchuelo A. M. 1971. Sobre un nuevo hallazgo de Amaranthus
caudatu en tumbas de Argentina. Kurtziana 6: 63–​67.
Intiquilla-​Quispe, A., Flores-​Fernández, C., Jiménez Aliaga, K. and Iris-​Zavaleta, A. 2018.
Capítulo 4: Potencial biotecnológico de las semillas de tarwi. In Lupinus mutabilis
(Tarwi) Leguminosa andina con gran potencial industrial, edited by Iris Zavaleta, 89–​
121. Lima: Fondo Editorial de la Universidad Nacional Mayor de San Marcos.
Jacobsen, S.E. 2003. The worldwide potential for quinoa (Chenopodium quinoa Willd.).
Food Reviews International 19: 167–​177.
Jacobsen, S.E. and Mujica, A. 2003. The genetic resources of Andean grain amaranths
(Amaranthus caudatus L, A. cruentus L. and A. hipochondriacus L) in America. Plant
Genetic Resources Newsletter 133: 41–​44.
Jacobsen, S.E and Mujica, A. 2006. El tarwi (Lupinus mutabilis Sweet.) y sus parientes
silvestres. 25. In Botánica Económica de los Andes Centrales, edited by R. Moraes,
B. Øllgaard, L. P. Kvist, F. Borchsenius and H. Balslev, 458–​482. La Paz: Universidad
Mayor de San Andrés.
Kaplan, L. 1965. Archeology and domestication in American Phaseolus (beans). Economic
Botany 19: 358–​368.
Keller, B., Ariza-​Suarez, D., de la Hoz, J., Aparicio, J.S., Portilla-​Benavides, A.E., Buendia,
H.F., Mayor, V.M., Studer, B. and Raatz, B. 2020. Genomic prediction of agronomic
traits in common Bean (Phaseolus vulgaris L.) under environmental stress. Frontiers
in Plant Science 11: 1001.
Kistler, L., Maezumi, S.Y., de Souza, J.G., Przelomska, N.A.S., Costa, F.M., Oliver Smith, Hope
Loiselle, Ramos-​Madrigal, J., Wales, N., Ribeiro, E.R., Morrison, R.R., Grimaldo, C.,
Prous, A.P., Arriaza, B., Gilbert, M.T.P., Freitas, F.O. and Allaby, R.G. 2018. Multiproxy
evidence highlights a complex evolutionary legacy of maize. South America Science
362(6420): 1306–​1313.
Kowalski, S., Mikulec, A. and Pustkowiak, H. 2020. Sensory assessment and physico-
chemical properties of wheat bread supplemented with chia seeds. Polish Journal of
Food and Nutrition Sciences 70: 387–​397.
Kulczynski, B., Kobus-​ Cisowska, J., Taczanowski, M. and Kmiecik, D. and Gramza-​
Michałowska, A. 2019. The chemical composition and nutritional value of chia
seeds—​Current state of knowledge. Nutrients 11: 1242.
Kumar, B., Smita, K., Cumbal, L. and Debut, A. 2017. Sacha inchi (Plukenetia volubilis L.)
shell biomass for synthesis of silver nanocatalyst. Journal of Saudi Chemical Society
21(1): 293–​298.
Lao, F., Sigurdson, G.T. and Giust, M. 2017. Health Benefits of Purple Corn (Zea mays
L.) phenolic compounds. Comprehensive Reviews in Food Science and Food Safety
16(2): 234–​236. DOI.org/​10.1111/​1541-​4337.12249.
References 115

Li, S., Zhang, B., Li, C., Fu, X. and Huang, Q. 2020. Pickering emulsion gel stabilized by
octenylsuccinate quinoa starch granuleas lutein carrier: Role of the gel network.
Food Chemistry 305: 125476.
Lopez, M.L., Capparelli, A. and Nielsen A. 2011. Traditional post-​harvest processing to
make quinoa grains (Chenopodium quinoa Willd) apt for consumption in Northern
Lipez (Potosı´, Bolivia): Ethnoarchaeological and archaeobotanical analyses.
Archaeological and Anthropological Sciences 3: 49–​70.
López-​Mejía, O.A., López-​Malo, A. and Palou, E. 2014. Antioxidant capacity of extracts
from amaranth (Amaranthus hypochondriacus L.) seeds or leaves. Industrial Crops
and Products 53: 55–​59.
Lucas-​ Gonzalez, R., Roldan-​ Verdu, A., Sayas-​ Barbera, E., Fernandez-​
Lopez, J., Perez-​
Alvarez, J.A. and Viuda-​Martos, M. 2020. Assessment of emulsion gels formulated
with chestnut (Castanea sativa M.) flour and chia (Salvia hispanica L) oil as partial
fat replacers in pork burger formulation. Journal of the Science of Food and Agriculture
100: 1265–​1273.
MAPA (Ministerio de Agricultura, Pesca y Alimentación). 2020. Encuesta sobre superficies y
rendimientos de cultivos. www.mapa.gob.es/​es/​esta​dist​ica/​temas/​estad​isti​cas-​agrar​
ias/​tot​ales​pana​ycca​a202​0_​tc​m30-​553​610.pdf.
Martinez-​ Lopez, A., Millan-​ Linares, M.C., Rodriguez-​ Martin N.M., Millan, F. and
Montserrat-​de la Paz, S. 2020. Nutracetical value of kiwicha (Amaranthus caudatus
L.). Journal of Functional Foods 65: 103735.
Mascaraque, M. 2018. Ancient Grains: From Traditional Staple Food to Superfood.
London: Euromonitor International. https://​blog.euro​moni​tor.com/​anci​ent-​gra​ins-​
from-​trad​itio​nal-​sta​ple-​food-​to-​superf​ood/​.
Morales, P., Rodrigo, V., González M., González, E., Tapia O., Sanhueza, C. and Valenzuela,
B. 2012. Nuevas fuentes dietarias de ácido alfa-​linolénico: una visión crítica. Revista
Chilena de Nutrición 39(3): 79–​87.
Mujica, A. 1992. Granos y leguminosas andinas: cultivos marginados. Roma: Organización
de la Naciones Unidas para la Agricultura y la Alimentación FAO.
Mukankusi, C., Raatz, B., Nkalubo, S., Berhanu, F., Binagwa, P., Kilango, M., Williams, M.,
Enid, K., Chirwa, R. and Beebe, S. 2018. Genomics, genetics and breeding of common
bean in Africa: A review of tropical legume project. Plant Breeding 138(4): 401–​414.
Muñoz, L.A., Cobos, A., Diaz, O. and Aguilera, J.M. 2013. Chia seed (Salvia hispanica): An
ancient grain and a new functional food. Food Reviews International 29: 394–​408.
Muñoz-​ Gonzalez, I., Merino-​ Alvarez, E., Salvador, M., Pintado, T., Ruiz-​Capillas, C.,
Jimenez-​Colmenero, F. and Herrero, AM. 2019. Chia (Salvia hispanica l.) a promising
alternative for conventional and gelled emulsions: Technological and lipid structural
characteristics. Gels 5: 19.
Ochieng, J., Schreinemachers, P., Ogada, M., Dinssa, F.F., Barnos, W. and Mndiga, H. 2019.
Adoption of improved amaranth varieties and good agricultural practices in East
Africa. Land Use Policy 83: 187–​194.
OECD/​ FAO (Organisation for Economic Co-​ operation and Development/​ Food and
Agriculture Organization). 2019. Latin American Agriculture: Prospects and
Challenges, OECD-​FAO Agricultural Outlook 2019-​2028. Paris: OECD Publishing.
Oliveira, M.R., Novack, M.E., Santos, C.P., Kubota, E. and Rosa, C.S. 2015. Evaluation of
replacing wheat flour with chia flour (Salvia hispánica L.) in pasta. Semina: Ciencias
Agrárias 36: 2545–​2553.
116 Latin-American Crops’ Challenges and Opportunities

Orona-​Tamayo, D. and Paredes-​López, O. 2017. Amaranth Part 1—​Sustainable crop for


the 21st century: Food properties and nutraceuticals for improving human health.
Sustainable Protein Sources, 239–​256. Cambridge, MA: Academic Press.
Orona-​Tamayo, D., Valverde, M.E. and Paredes-​López, O. 2019. Bioactive peptides from
selected Latin America crops –​A nutraceutical and molecular approach. Critical
Reviews in Food Science and Nutrition 59(12): 1949.
Orozco, G., Durán, N., González, D., Zarazúa, P., Ramírez, G. and Mena, S. 2014. Proyecciones
de cambio climático y potencial productivo para Salvia hispanica L. en las zonas
agrícolas de México. Revista Mexicana de Ciencias Agrícolas 10: 1831–​1842.
Panorama Alimentario 2020, SIAP. www.inforu​ral.com.mx/​wp-​cont​ent/​uplo​ads/​2020/​11/​
Atlas-​Agro​alim​enta​rio-​2020.pdf.
Park, S.J. Sharma, A. and Lee, H.J. 2020. A review of recent studies on the antioxidant activ-
ities of a third-​millennium food: Amaranthus spp. Antioxidants 9: 1236.
Pastor-​Pazmiño, C., Chonheiro, L. and Wahren, J. 2017. Agriculturas alternativas en
latinoamérica. Tipología, alcances y viabilidad para la transformación social-​
ecológica. Mexico: Fundación Friedrich Ebert.
Pelaez, P., Orona-​Tamayo, D., Montes-​Hernandez, S., Valverde, M.E., Paredes-​Lopez, O.and
Cibrian-​Jaramillo, A. 2019. Comparative transcriptome analysis of cultivated and
wild seeds of Salvia hispanica (chia). Scientific Reports 9: 9761.
Pellegrini, N. and Agostoni, C. Nutritional aspects of gluten-​free products. Journal of the
Science of Food and Agriculture 95: 2380–​2385.
Pintado, T., Ruiz-​Capillas, C., Jimenez-​Colmenero, F.and Herrero, A.M. 2020. Impact of
culinary procedures on nutritional and technological properties of reduced-​fat
longanizas formulated with chia (Salvia hispanica l.) or oat (Avena sativa l.) emulsion
gel. Foods 9: 1847.
Polania, J.A.; Poschenrieder, C.; Beebe, S.and Rao, I.M. 2016. Effective use of water and
increased dry matter partitioned to grain contribute to yield of common bean
improved for drought resistance. Frontiers in Plant Science 7: article 660.
Popenoe, H., King, S.R., Leon, J. and Kalinowski, L.S. 1989. Lost Crops of the Incas: Little-​
Known Plants of the Andes with Promise for Worldwide Cultivation. Washington,
DC: National Research Council.
Quintal-​Bojórquez, N.D.C., Carrillo-​Cocom, L.M., Hernández-​Álvarez, A.J. and Segura-​
Campos, MR. 2021. Anticancer activity of protein fractions from chia (Salvia
hispanica L.). Journal Food Science 86: 2861–​2871.
Repo-​Carrasco-​Valencia, R. 2011. Andean Indigenous Food Crops: Nutritional Value and
Bioactive Compounds. University of Turku, Turku Finlad ISBN 978-​951-​29-​4605-​1.
Repo-​Carrasco-​Valencia, R., Espinoza, C. and Jacobsen, S-​E. 2003. Nutritional value and
use of the Andean crops quinoa (Chenopodium quinoa) and Kañiwa (Chenopodium
pallidicaule). Food Reviews International 19(1–​2): 179–​189.
Repo-​Carrasco-​Valencia, R., Acevedo de La Cruz, A., Icochea Alvarez, J.C. and Kallio, H. 2009.
Chemical and functional characterization of kañiwa (Chenopodium pallidicaule)
grain, extrudate and bran. Plant Foods for Human Nutrition 64(2): 94–​101.
Resta, D., Boschin, G., D’Agostina, A. and Arnoldi, A. 2008. Evaluation of total quinolizidine
alkaloids content in lupin flours, lupin-​based ingredients, and foods. Molecular
Nutrition and Food Research 52(4): 490–​495.
Reyes-​Moreno, C. and Paredes-​Lopez, O. 1993. Hard-​to-​cook phenomenon in common
beans —​A review. Critical Reviews in Food Science and Nutrition 33(3): 227–​286.
References 117

Ribeiro, N.D., Rodrigues, J. de A., Prigol, M., Nogueira, C.W., Storck, L., Gruhn, E.M. 2014.
Evaluation of special grains bean lines for grain yield, cooking time and mineral con-
centration. Crop Breeding and Applied Biotechnology 14(1): 15–​22.
Ribes, S., Pena, N., Fuentes, A., Talens, P. and Barat, J.M. 2021. Chia (Salvia hispanica L.) seed
mucilage as a fat replacer in yogurts: Effect on their nutritional, technological, and
sensory properties. Journal of Dairy Science 104: 2822–​2833.
Rojas, W. 2003. Multivariate analysis of genetic diversity of Bolivian quinoa germplasm.
Food Reviews International 19: 9–​23.
Rosell, C.M., Cortez, G. and Repo-​Carrasco R. 2009. Breadmaking use of Andean crops
quinoa, kañiwa, kiwicha, and tarwi. Cereal Chemistry 86(4): 386–​392.
Ruiz, M.A. and Sotelo, A. 2001. Chemical composition, nutritive value, and toxicology
evaluation of Mexican wild lupins. Journal of Agricultural and Food Chemistry
49: 5336−5339.
Santiago-​ Lopez, L., Hernandez-​ Mendoza, A., Vallejo-​ Cordoba, B., Mata-​ Haro V. and
Gonzalez-​Cordova, A. F. 2016. Food-​derived immunomodulatory peptides. Journal
of the Science of Food and Agriculture 96 (11): 3631–​3641.
Sauer, J.D. 1967. The grain amaranths and their relatives: A revised taxonomic and geo-
graphic survey. Annals of the Missouri Botanical Garden 54(2): 103–​137.
Sauer, J.D. 1976. Grain amaranths Amaranthus spp.(Amaranthaceae). In Evolution of
Crop Plants, edited by N.W. Simmonds, chap. 2: 4–​ 6. London and New York:
Longman.
Saunders, R.M. and Becker, R. 1984. Amaranthus: A potential food and feed resource. In
Advances in Cereal Science and Technology, edited by Y. Pomeranz, 6, 357–​396. St.
Paul, MN: American Association of Cereal Chemists
Schiermeier, Q. 2019. Eat less meat: UN climate-​change report calls for change to human
diet. Nature 572: 291–​292.
Schoeneberger, H., Gross, R., Cremer, H. and Elmadfa, I. 1982. Composition and protein
quality of Lupinus mutabilis. The Journal of Nutrition 112: 70–​76.
Sharma, V. and Mogra, R. 2019. A comprehensive review on chia seeds (Salvia hispanica L.).
International Journal of Chemical Studies 7: 57–​61.
Silva, H., Arriagada, C., Campos-​Saez, S., Baginsky, C., Castellaro-​Galdames, G.and Morales-​
Salinas, L. 2018. Effect of sowing date and wáter availability on growth of plants of
chia (Salvia hispanica L) established in Chile. PLoS ONE 13(9): e0203116.
Singh, J.P., Singh, B. and Kaur A. 2021. Bioactive compounds of legume seeds. In
Bioactive Compounds in Underutilized Vegetables and Legumes, Reference Series in
Phytochemistry, edited by H.N. Murphy and K.Y. Pack, 645–​665. Cham: Springer.
Singh, M. and Liu, S.X. 2021. Evaluation of amaranth flour processing for noodle making.
Journal of Food Processing and Preservation 45 : 1–​9.
Soriano-​García, M., Arias-​Olguín, I.I., Montes, J.P.C., Ramírez, D.G.R., Figueroa, J.S.M.,
Flores-​Valverde, E. and Valladares-​Rodríguez, M.R. 2018. Nutritional functional value
and therapeutic utilization of Amaranth. Journal of Analytical & Pharmaceutical
Research 7: 596–​600.
Sosa, C.A. 2000. Influencia de dos métodos de extracción de un aislado proteico de lupino
(Lupinus mutabilis) en sus propiedades funcionales, Tesis Magis-​tral, Lima: Biblioteca
Agrícola Nacional, Universidad Nacional Agraria La Molina.
Suarez-​Martinez, S.E., Ferriz-​Martínez, R.A., Campos-​Vega, R., Elton-​Puente, J.E., Carbota,
K.T. and García-​Gasca, T. 2016. Bean seeds: Leading nutraceutical source for human
118 Latin-American Crops’ Challenges and Opportunities

health [Semillas del frijol: fuente líder de nutracéuticos para la salud humana].
CyTA –​ Journal of Food 14(1): 131–​137.
Švec, I. and Hrušková, M. 2015. Hydrated chia seed effect on wheat flour and bread techno-
logical quality. Agricultural Engineering International: CIGR Journal Special Issue:
259–​263.
Tabaxi, I., Zervas, G., Tsiplakou, E., Travlos, I.S., Kakabouki, I. and Tsioros, S. 2016. Chia
(Salvia hispanica) fodder yield and quality as affected by sowing rates and organic
fertilization AU –​bilalis, dimitrios. Communications in Soil Science and Plant Analysis
47: 1764–​1770.
Tapia, M.E. 2014. El largo camino de la quinoa: ¿Quiens escribieron su historia? In Estado
del arte de la quinua en el mundo en 2013, 3–​10. Santiago de Chile/​Montpellier,
France: FAO and CIRAD.
Tapia, M.E. 2015. El tarwi, lupino andino –​La Fondazione L’Albero della Vita. Fondo
Ítalo Peruano http://​fadv​amer​ica.org/​wp-​cont​ent/​uplo​ads/​2017/​04/​TARWI-​espa​
nol.pdf.
Teoh, S.L., Lai, N.M., Vanichkulpitak, P., Vuksan, V., Ho, H. and Chaiyakunapruk, N. 2018.
Clinical evidence on dietary supplementation with chia seed (Salvia hispanica L.): A
systematic review and meta-​analysis. Nutrition Reviews 76: 219–​242.
Ullah, R., Nadeem, M. and Imran, M. 2017. Omega-​3 fatty acids and oxidative stability of
ice cream supplemented with olein fraction of chia (Salvia hispanica l.) oil. Lipids in
Health and Disease 16: 34.
United Nations, 2019). World Population Prospects. New York: United Nations Population
Division. https://​pop​ulat​ion.un.org/​wpp/​.
Urbina M. 1887. La chía y sus aplicaciones. La Naturaleza. Sociedad Mexicana de Historia
Natural Tomo 1: 27–​36.
Valvas, R.L.M., Gomez-​Pando, L. and Pinedo-​Taco, R. 2020. Sustainability of the produc-
tion units of the amaranto crop (Amaranthus caudatus L.). Ecosistemas y Recursos
Agropecuarios 7: e2483.
Vásquez-​Osorio, D., Hincapié-​Llanos, G.A., Cardona, M., Jaramillo, D.I. and Vélez Acosta,
L. 2017. Formulación de una colada empleando harina de Sacha Inchi (Plukenetia
Volubilis L.) proveniente del proceso de obtención de aceite. Perspectivas En
Nutrición Humana 19(2): 167–​179.
Vietmeyer, N.D. 1989. Lost Crops of the Incas. National Research Council. Washington,
DC: National Academy Press.
Wang, S., Zhub, F., Kakudac, Y. 2018. Sacha inchi (Plukenetia volubilis L.): Nutritional com-
position, biological activity, and uses. Food Chemistry 265: 316–​328.
Yeboah, T., Owusu, L., Arhin, A. and Kumi, E. 2014. Fighting poverty from the
street: Perspectives of some female informal sector workers on gendered poverty
and livelihood portfolios in Southern Ghana. Journal of Economic and Social Studies
5(1): 239–​267.
Zettel, V. and Hitzmann, B. 2018. Applications of chia (Salvia hispánica L.) in food products.
Trends in Food Science and Technology 80: 43–​50.
Chapter 4

Structure and Composition


of Latin-​American Crops

Barbara Borczak1, José María Coll Marqués2,


Octavio Paredes-​López3, and Claudia M. Haros1
1University of Agriculture, Krakow, Poland
2Instituto de Agroquímica y Tecnología de

Alimentos (IATA-​CSIC), Valencia, Spain


3Centro de Investigación y de Estudios Avanzados (CINVESTAV),

Instituto Politécnico Nal., Unidad Irapuato, Guanajuato, Mexico

CONTENTS
4.1 Introduction 119
4.2 Gross Structural Features 126
4.3 Physical Properties 127
4.4 Kernel Structures 134
4.5 Chemical Composition of Kernels 140
4.5.1 Proteins 140
4.5.2 Carbohydrates 142
4.5.3 Lipids 148
4.6 Conclusions 150
Acknowledgments 151
References 151

4.1 INTRODUCTION
Latin-​American indigenous seeds constitute cereals, pseudo-​cereals, oilseeds,
and legumes. The selected crops are maize from the Latin-​American region

DOI: 10.1201/9781003088424-4 119


120 Structure and Composition of Latin-American Crops

(Zea mays), amaranth (Amaranthus spp), quinoa (Chenopodium spp), kañiwa


(Chenopodium pallidicaule), chia (Salvia hipsanica L.), sacha inchi (Plukentia
volubilis), as well as legumes such as black turtle bean (Phaseolus vulgaris) and
tarwi (Lupinus mutabilis). It is pertinent to mention that some important add-
itional crops also originated in Latin America and are widely distributed around
the world, such as cocoa, sunflower, peanuts, and other type of maize. Below, a
short description is shown of the first eight specific crops described above.
Cereals are mainly represented by maize (Table 4.1). The words maize and
corn are used indiscriminately to refer to the same crop and seed. Based on the
origin of maize (Teocintle, Mexico) and the process of its domestication, two
different cultivars have been developed: Mesoamerican and Andean (Salvador-​
Reyes et al., 2021). There are about 260 very well-​known races, of which 98 are
found in Mesoamerica, while 146 have been found in the Andes (Colombia,
Ecuador, Peru, Brazil, Chile, Argentina, and Bolivia) (Boege, 2009; Serratos, 2012).
The remaining races are found in some regions of South and Central America.
Zea mays is one of the most important staple crops found in the highlands
of the Andean regions of Ecuador, Peru, and Bolivia, between 1,700 and 3,500
m above sea level and is one of the main sources of energy in the diet of the
inhabitants (Zambrano et al., 2021). Zambrano et al. (2021) indicated that in
Ecuador, there are 29 races of maize, 17 found in the highlands, 6 of which are
not well defined (Yánez et al., 2003). In the north of this country, traditional
races called chauchos, huandangos, and mishcas are generally cultivated (soft
yellow kernels). In the central zone, preferably floury white races of maize are
sown, such as blanco blandito, chazo, and guagal. In the south, a type of maize
called zhima (white semi-​dent kernel) is mostly grown (Zambrano et al.,
2021). In Peru, there are 52 races of maize, 32 are grown in the highlands,
and the vast majority corresponds to floury maize. In Bolivia, 40 races were
found of which 31 are cultivated in the highlands (Salhuana, 2004a). There are
a large number of traditional varieties due to the diverse uses for human con-
sumption. The most important native cultivars in this region are karapampa,
kellu, uchupilla, hualtaco, huillcaparu, pasakalla, chuspillo checchi, and kulli
chunkula (Ortiz, 2012). The floury maize is consumed by humans in different
forms: as fresh maize (choclo), dry toasted kernels (tostado, cancha), soups
(chuchuca), drinks (chicha, colada, maicena), breads, and other traditional
preparations (Villacrés et al., 2016). The grain is a great source of energy and
amino acids due to its starch, protein, and oil content. It also contains eight
minerals (K, Fe, Zn, Ca, P, Si, Cu and Mg), nine vitamins (A, B1, B2, B3, B4,
B6, B9, C, and E), some phenolic compounds and anthocyanin, in the case
of black, red, and blue maize (Prasanna et al., 2020; Zambrano et al., 2021)
(Figure 4.1).
4.1 Introduction 121

A B

C D E

Figure 4.1 Races of maize from the Latin-​American region.


Notes: There are five images. The top left image shows many different Andean maize ear
color types, in which the grains also vary in color within an ear. The top right photo shows
many different Andean maize ear color types (white, yellow, purple, and red). The bottom
left image shows pop-​corn maize called “maíz pisingallo” of an orange-​yellow color. The
bottom middle image shows one purple corn. The bottom right image shows two white
maize ears of “Choclo” maize. Source: https://​es.wikipe​dia.org/​wiki/​Arch​ivo:Peruvi​
an_​c​orn.jpg (Source: www.fli​ckr.com/​pho​tos/​jennif​rog/​57252​925/​, Author: Jenny
Mealing);
Image B: https://​es.wikipe​dia.org/​wiki/​Arch​ivo:D%C3%ADa_​Inter​naci​onal​_​de_​
los_​Pueb​los_​Ind%C3%ADgen​as_​(785​2553​230).jpg (www.fli​ckr.com/​pho​tos/​10021​
639@N05/​785​2553​230, Author: Cancillería Ecuador);
Image C: https://​es.wikipe​dia.org/​wiki/​Palo​mita​s_​de​_​ma%C3%ADz#/​media/​Arch​
ivo:Popc​ornC​obs2​007.jpg;
Image D: https://​es.wikipe​dia.org/​wiki/​Zea_​m​ays#/​media/​Arch​ivo:Mai​zmor​ado.
png;
Image E. https://​es.wikipe​dia.org/​wiki/​Varied​ades​_​per​uana​s_​de​_​ma%C3%ADz#/​
media/​Arch​ivo:(Cho​clo)_​fro​m_​Qu​ito.JPG.
122 Structure and Composition of Latin-American Crops

Pseudo-​ cereals of the Latin-​ American regions are mainly composed of


amaranth, quinoa, and kañiwa. They were commonly consumed by Aztecs,
Maya, Incas, and other pre-​Columbian civilizations during pre-​Hispanic times.
However, after the Spanish conquest, their cultivation likely disappeared due
to their association with the religions of indigenous cultures (De la Cruz Torres
et al., 2008; Reguera and Haros, 2017; Mangelson et al., 2019; Orona-​Tamayo
et al., 2019). The genus Amaranthus L. contains about 50–​70 species and 400
varieties (Joshi et al., 2018; Aderibigbe et al., 2022). The main species of amar-
anth cultivated for seeds and most commonly used for human consumption are
A. cruentus in Guatemala, A. caudatus in Peru, and A. hypochondriacus in Mexico
(Bressani, 2003; Reguera and Haros, 2017; Aderibigbe et al., 2022). Quinoa is a
part of the Amaranthaceae family and closely related with amaranths. Quinoa
(Chenopodium quinoa) was a fundamental daily food of the ancient South
American Andean cultures, and is now grown mainly in the Andean countries
of Peru and Bolivia (Reguera and Haros, 2017). Similarly, kañiwa/​cañahua is a
close relative of quinoa. However, unlike quinoa, kañiwa is less bitter because it
contains a smaller amount of saponin (Huamani et al., 2020). As of now, there are
about 30 cultivated and wild kañiwa varieties known (Mangelson et al., 2019).
Kañiwa is grown in Andean highland plateaus at over 4,000 m.a.s.l. and is very
resistant to strict climatic conditions. At present, kañiwa is grown mainly in the
Peruvian and Bolivian altiplano by families for their own consumption and used
as kañiwaco, toasted kañiwa flour. This nutty-​tasting flour is mixed with water
or milk and eaten as breakfast cereal as a great source of calories and nutritive
proteins (Repo-​Carrasco-​Valencia et al., 2009).
Oleaginous seeds include sacha inchi (Plukenetia volubilis L.) and chia (Salvia
hispanica L.). The first is a plant of the Euphorbiaceae family with large seeds
(Wang et al., 2018) (Table 4.1) and used to be described as “wild peanut”, ”moun-
tain peanut”, “Inca peanut”, or “Inca inchi”. It grows in the Amazonian forest
(Gutiérrez et al., 2011) in Surinam, Venezuela, Colombia, Ecuador, Peru, Bolivia,
and Brazil. The genus Plukenetia comprises more than 25 species (Kodahl
and Sørensen, 2021), including P. brachybotrya, P. polyadenia, P. loretensis, and
P. huayllabambana. However, their morphological and physicochemical proper-
ties differ from P. volubilis. Cultivation of sacha inchi dates from pre-​Hispanic
times and most probably constituted part of Incans’ diets (3,000–​5,000 years
ago) (Kodahl and Sørensen, 2021). It grows at altitudes ranging from 200 to 1,500
m (Wang et al., 2018).
Seeds of P. volubilis have been used in traditional cooking in Peru (inchacapi,
lechona api, pururuca, cutacho, inchi cucho, tamales, turron, and chichi). They
can also be consumed in roasted and salted form, in chocolate, pressed as
oil, ground to a buttery substance, or as cooked leaves (Kodahl and Sørensen,
2021). At the same time, chia (Salvia hispanica L.) is a herbaceous plant
4.1 Introduction 123

belonging to the Lamiaceae family (Palma-​Rojas et al., 2017; Hernández-​Pérez


et al., 2021) (Table 4.1). The genus Salvia consists of approximately 900 species
(Knez Hrnčič et al., 2020). The plants can grow to 1–​1.5 m high and the fruits
are indehiscent achene. It originates from west Mexico and northern Guatemala
(Pelaez et al., 2019), although today it is also cultivated in Australia, Bolivia,
Columbia, Peru, Argentina, the USA, and Europe (Palma-​Rojas et al., 2017) with
Mexico being the world’s largest chia producer (Hrnčič et al., 2020). Chia as an
oilseed has great economic potential since it is a source of oil with remarkable
nutritional and nutraceutical characteristics; thus, chia has a brilliant future
for industrial and human purposes (Ayerza, 2009; Orona-​Tamayo et al., 2019;
Hernández-​Pérez et al., 2021).
The legume family (Leguminosae) comprises 643 genera (18,000 species)
which are grouped into 40 tribes and cultivated in either tropical or temperate
climates (Broughton et al., 2003). Legumes native to Latin America that we are
going to deal with in this chapter are Lupinus mutabilis Sweet and Phaseolus
vulgaris also known as black turtle bean, among many others. L. mutabilis is a
plant variously known as ullush, talwish, tauri, tarwi, chocho, lupino or ccquella
(Chirinos et al., 2015). It belongs to genus Lupinus (Table 4.1) with 85–​100 dis-
tinct species and more than 500 specie names (Atchison et al., 2016). It is a rela-
tive of Lupinus albus, Lupinus luteus, and Lupinus angustifolius, three crucial
pulses collected worldwide (Córdova-​Ramos et al., 2020). This legume grows in
the Andes at 2,000 to 3,800 m.a.s.l and is distributed from Colombia to the north
of Argentina (Tapia, 2015, Chirinos et al., 2020). The seeds were found in the
Nazca cultural remains in Peru as possible evidence of their consumption by the
ancient Peruvian cultures Tiahuanaco and Chavin (Castañeda, 1988; FAO, 2010).
Tarwi has a considerable genetic variability within the Andean region, with
different morphological types reported (Gross, 1982) as follows: (1) small plants
of about 60 cm high, no branches, presented in the region of Potosí (Bolivia) and
in the regions above 3,500 m.a.s.l.); (2) branched and tall plants (in the Andean
valleys of Bolivia and Southern Peru), as well as (3) highly branched plants,
taller than 1.8 m and growing in Colombia, Ecuador, and northern Peru (Neves
Martins et al., 2016). Use of tarwi might be limited because of high alkaloid con-
tent (Jacobsen and Mujica, 2006), which results in a bitter taste which is simul-
taneously toxic. However, with the soaking and washing of seeds in water, the
bitter compounds are removed.
The tribe Phaseoleae containing, among others, common beans (P. vulgaris)
is among the most important economic group of legumes in the world. Basal
species (P. microcarpus, P. vulgaris) are part of a complex species that includes
P. acutifolius, P. coccineus, and P. polyanthus (Broughton et al., 2003). The black
bean, also known as black Spanish bean, Tampico bean, and Venezuelan bean,
has its origin in South America with two centers of domestication (Gepts and
124 Structure and Composition of Latin-American Crops

TABLE 4.1 BOTANICAL CLASSIFICATION OF LATIN-​AMERICAN


CROPS AND WHEAT FOR COMPARISON PURPOSES
CEREALS PSEUDO-​CEREALS

Name WHEAT CORN AMARANTH QUINOA


Class Monocotyledoneae Dicotyledoneae
Owpder Poales Caryophyllales
Family Poaceae Amaranthaceae Chenopodiaceae

Subfamily Pooide Bambusoideae Amaranthoideae Chenopodioideae


Tribe Triticeae Andropogoneae Amarantheae Chenopodieae/​
Chenopodiea

Chemopodium
Amaranthus
Triticum
Genus

Zea

Aellen (kanigua/​kanihua/​canihua),
Ch. quinoa Willd, Ch. pallidicaule
T. aestivum, T. aestivum spp.

A. caudatus, A. cruentus, A.

Ch. nuttalia Safford


hypochondriacus
spelta, T. durum

Z. mays
Species

Source: Adapted from Carvajal-​Larena (2019), Chirinos-​Arias (2015), Moreno et al. (2014),


Berghofer and Schöenlechner (2007), Saleem et al. (2016), Kodahl (2020), Kodahl and
Sørensen (2021), Coelho and Salas-​Mellado (2014), Jamshidi et al. (2019),
Knez Hrnčič et al. (2020).
Chenopodium pallidicaule Aellen

KAÑIWA
BLACK
P. acutifolius, P. coccineus, P.
Phaseolus

Phaseolae
Polyanthus, P.vulgaris

Dicotyledoneae
TURTLE BEAN

L.mutabilis and four species as


possible close relatives Fabales
LEGUMES

TARWI

(L. ellsworthianus C.P.Sm.,

Papilionoideae
Genisteae
L. piurensis C.P.Sm., L. praestabilis
Lupinus
Fabaceae or Leguminosae
C.P.Sm., and L. semperfl orens);
Mediterranean species (Lupinus albus
L. and Lupinus luteus
L. Lupinus angustifolius L.)
CHIA

Labiate
Lamiales

S. hispanica Salvia
Mentheae
Lamiaceae/​

Nepetoideae
Magnoliopsida

Benth.

ca. 25 species
INCHI
SACHA

subtribe
OLEAGINOUS SEEDS

P. volubilis, P. brachybotrya, P. polyadenia, Plukentia L.


Plukenetieae

P. loretensis, and P. huayllabambana.


Plukenetiinae
Acalyphoideae
Euphorbiaceae

(Benth) Hutch.,
4.1 Introduction
125
126 Structure and Composition of Latin-American Crops

Dpbouk, 1991). The small seed varieties originated from a small-​seeded wild
type (P.vulgaris var. aborigineus (Burk.) Baudet) in Central America, while the
large-​seeded varieties are from another type (P. vulgaris var. aborigineus) in
the Andean region of South America. Black turtle bean is traditionally used in
Mexican, Caribbean, and South American dishes (Priya and Manickavasagan,
2020) in the form of burrito, feijoada (stew of beans, beef, and pork), gallo pinto
or pabellón criollo (rice, shredded beef, and stewed black beans) (Lim, 2012).
The Andean mountain range in South America is a significant location for crop
domestication intended for human nutrition, even though growing conditions
are demanding with high and varied altitudes (1,500–​4,200 m.a.s.l.) as well as
accompanied by changing climates. It is pertinent to mention that indigenous
people of the Andes domesticated several dozens of separate crop species
(Repo-​Carrasco-​Valencia, 2020; Hernández-​Pérez et al., 2020; Hernández-​Pérez
et al., 2021).
The botanical classification and physicochemical characteristic of Latin-​
American crops are summarized in Tables 4.1, 4.2, and 4.3.

4.2 GROSS STRUCTURAL FEATURES


Latin-​American seeds are classified in this report into different plant crops
(cereals, pseudo-​cereals, legumes, and oilseeds) and they differ between each
other in biological quality proteins, amino acids, polyunsaturated fatty acids,
dietary fiber, and starch content, among other components and traits.
The main parts of maize in general, include: pericarp, aleurone, endosperm,
and germ (Salvador-​Reyes et al., 2021). The outer layer of kernels may be yellow
(Chullpi), black with white spots (Piscorunto), pale yellow (Giant Cuzco), red–​
orange with white stripes (Sacsa), and an intense and homogeneous purple
(Purple). Differences in color are determined by climatic conditions and gen-
etic factors. High altitudes together with lower temperatures induce natural
pigment development, such as anthocyanins (determining blue, purple, and
black colorations). On the other hand, warm climates promote the concentra-
tion of carotenoids (leading to red, yellow, and orange colorations) (Figures 4.1–​
4.2). Piscorunto, Giant Cuzco, Sacsa, and Purple varieties are classified as floury
maizes, because their endosperm is composed exclusively of soft starch. At the
same time, Chullpi is considered a sweet type of maize since it presents part of
its endosperm in a vitreous state (Paliwal et al., 2000). There is also a higher ratio
of starch: protein presented in their endosperm (11:1) compared to vitreous (6:1)
(Wang and Wang, 2019).
In amaranth, quinoa, and kañiwa the bran fraction is higher than traditional
cereals (Bressani, 2003; Reguera and Haros, 2017). They also present greater
4.3 Physical Properties 127

content of protein and fat compared with cereals, whereas the starch content is
lower (Figure 4.3, Table 4.3). Oleaginous plants occur in the form of fruits (sacha
inchi) or grain-​like seeds (chia). The kernels are only covered with a thin layer
in the form of shell and contain a higher amount of protein than cereals and
pseudo-​cereals together with fat and dietary fiber being the greatest among all
Latin-​American crops (Figure 4.4, Table 4.3). The main parts of legumes (tarwi
and black turtle bean) are seed coat, radicle, and cotyledon with the latest having
the greatest share in the whole seed (Figure 4.5). The protein content of tarwi is
remarkable, being one of the greatest among all Andean crops.

4.3 PHYSICAL PROPERTIES
Knowledge about engineering features, mainly the physical properties of crops,
including cereals, pseudo-​cereals, legumes and oilseeds is important for engin-
eers and technicians involved in several harvest and postharvest practices.
Physical characteristics are crucial in investigating the maintenance of crops
in handling operations (Mohsenin, 1986; Suleiman, 2019). Data on physical
properties of Latin-​American crops, including their morphology and size dis-
tribution are essential to proposing equipment and amenities for the following
processes: harvesting, handling, cleaning, conveying, grading, separation, drying,
storing, and processing of crops (Sacilik et al., 2003; Vilche et al., 2003; Reguera
and Haros, 2017). The most important physical and thermal properties of the
grain, oilseed, and other agricultural commodities are seed dimension (length,
width, and thickness), arithmetic and geometric mean diameter, surface area,
volume, sphericity ,and aspect ratio. They also may include the weight of 1,000
grains, bulk, and kernel density, fractional porosity, the static and dynamic coef-
ficient of friction against different materials, as well as the angle of repose, heat
capacity, thermal conductivity and diffusivity, and latent heat of vaporization
(Mohsenin, 1986; Suleiman, 2019).
For example, bulk density influences the capacity of storage and transport
systems, while information about true density is needed for separation and clas-
sification equipment. Seed porosity determines the resistance to airflow during
the aeration and drying of kernels (Reguera and Haros, 2017). Table 4.2 presents
a data sheet of known physical properties of Latin-​American crops compared
to wheat. These properties vary widely, depending on moisture content, variety,
or cultivar in general, year and region of cultivation, as well as temperature of
grains, legumes, and oilseeds (Mohsenin, 1986; Vilche et al., 2003; Reguera and
Haros, 2017; Suleiman, 2019). Latin-​American crops may differ in size and shape
(Figures 4.1–​4.2), and as expected they show different physical and chemical
properties (Tables 4.2–​4.3).
128 Structure and Composition of Latin-American Crops

TABLE 4.2 PHYSICAL PROPERTIES AND GEOMETRIC DIMENSIONS


OF LATIN-​AMERICAN CROPS. (WHEAT WAS PRESENTED FOR
COMPARISON PURPOSES)
Chenopodium
Wheat(a) quinoa
varieties Amaranthus Willd(c-​e)
Parameter Units S/​D/​N*** Maize(b) cruentus(c-​d) L/​M/​S*
Moisture % dm 15.8/​15.8/​ 8.40/​14.44/​ 7.7-​43.9(c) 4.6-​25.8(c)
16.4 12.75/​NR/​ 9.5-​43.6(d) 15.0(d)
12.02 and
11.1

1000-​seed mass G 46.2/​45.9/​ 324.28-​ 0.79(c’) /​1.2(c’’) 2.5-​3.1(c)


40.0 1340.63

Seed weight G NR 50.06-​61.56 NR NR

Hectoliter kg/​ NR NR NR NR
weight 100L
Specific Volume m3/​kg NR NR 0.78-​1.10 x NR
10-​3(c)
Bulk density kg/​m3 791/​789/​ NR 840-​720(c) 747-​667(c)
732 820-​867(c’)
True density(a,c,d)/​ kg/​m3 1104/​ NR 1390-​1320(c) 928-​1188(c)
compacted 1151/​1076
density
Porosity % 0.283/​ NR 0.40-​0.45(c) 0.194-​0.438(c)
0.315/​
0.320
L*(brightness or NR 66.73/​ NR NR
white-​ness), 57.53/​
76.31/​
52.50/​14.72
a*(red-​ness and NR 10.98/​1.62/​ NR NR
green-​ness) 1.64/​17.71/​
2.82
b*(yellow-​ness NR 39.72/​ NR NR
and blue-​ness) 11.58/​
23.83/​
21.41/​0.59
4.3 Physical Properties 129

Chenopodium Plukentia Lupinus Phaseolus


pallidicaule ( f-​g) Salva hispanica L.( f-​h) volubilis(i) mutabilis (j-​k) vulgaris(l-​m)
10.37/​11.79( f) 7.0±0.4(h)/​6.2 ± 0.517 3.3-​8.32(i) 9.99 ± 0.14 11.02(l’);
10.37/​9.61/​9.79/​ (8.19–​13.80)(k) 9.25(l’’)
10.39 ( f ’)
10.7/​10/​.7/​10.7/
7.7/​8.4 ( f ’’)
2.5-​4.0(g) 2.0-​3.5( f ’)/​1.267-​1.387(h) NR 153-​320(k’) 256.2/​
175.9/​
246.1/​
196.3/​
224.4(l)
NR NR NR 0.25±0.06 NR
(0.184-​0.302)(k)
NR NR 0.70-​1.22(i’) NR NR

NR NR NR NR NR

958-​904(g) 729-​458( f ’)/​722(g) NR NR 2.4-​10 (m’)

NR 937-​1075(g) NR NR NR

NR 22.9-​32.8(g) NR NR NR

29.8/​30.5/​29.8(g) 42/​41.9/​41.30(g) NR 61.21±0.10(k’’)/​ 69.9 ± 2.5(m)


(67.2-​87.5)(k)/​
84.59 ± 0.90(j)

11.3/​11.1/​11.9(g) 3.2/​3.8/​3.8(g) NR 2.22±0.10(k’’)/​ 0.6 ± 0(j)


(0.3-​(-​3.4))(k)/​
−2.10 ± 0.44(j)
12.9/​14.1/​14.8(g) 7.4/​8.6/​8.6(g) NR 11.47±0.04(k’’)/​ 5.9±0(m)
(22.0-​33.3)(k)/​
20.32 ± 1.55(j)

(continued)
130 Structure and Composition of Latin-American Crops

TABLE 4.2 (CONTINUED)


Chenopodium
Wheat(a) quinoa
varieties Amaranthus Willd(c-​e)
Parameter Units S/​D/​N*** Maize(b) cruentus(c-​d) L/​M/​S*
Length mm 5.46/​5.37/​ 18.23-​20.36 1.35-​1.50(d) 2.045/​1.889/​
5.38 1.691(d)*
6.51–​8.66(a’)

Width mm 2.56/​2.47/​ 7.93-​17.34 1.22-​1.37(d) 2.015/​1.885/​


2.62 1.689(d)*
3.026–​
3.952(a’)
Thickness mm 2.12/​2.18/​ 5.17-​6.13 0.81-​0.93(d) 0.930/​0.980/​
2.43 0.973(d)*
2.25–​2.79(a’)

Equivalent mm 3.09/​3.06/​ NR 1.10-​1.24(d) 1.394-​1.607(c)


diameter 3.24
Sphericity 0.57/​ NR 0.81-​0.83(d) 0.77-​0.80(c)
0.57/​0.60
57.40–​
69.43%(a’)
Volume mm3 NR NR 0.55-​0.76(d) NR

Equivalence mm2 30.07/​ NR 3.26-​4.04(d) NR


sphere area 29.64/​33.2
Oblate spheroid mm2 0 NR 3.59-​4.43(d) NR
area
Solid of mm2 NR NR 3.60-​4.47(d) NR
revolution area
Arithmetic mm NR NR NR NR
mean dia
Geometric mm 3.36–​4.24(a’) NR NR NR
mean dia,
Surface area S, mm2 47.92–​ NR NR NR
51.65(a’)
Aspect ratio R % NR NR NR NR
4.3 Physical Properties 131

Chenopodium Plukentia Lupinus Phaseolus


pallidicaule ( f-​g) Salva hispanica L.( f-​h) volubilis(i) mutabilis (j-​k) vulgaris(l-​m)
1-​1.2( f)and 1.14(g) 1.99 (min.0.26 7.49-​ 9.64±0.11 10.93/​9.64/​
(min.0.13 max.2.49) ( f ’)/​ 8.92(i’);15-​ (8.86-​10.99)(k)/​ 10.60/​9.39/​
max.1.45) 1.76–​2.42(g) 18(i’’) 9.98±0.64 10.37(l);
9.1±0.9(l’’)
0.92(g) 1.21 (min.0.14-​1.68) 16-​20(i’’) 7.91±0.11 7.05/​6.21/​
(min.0.12max.1.20) ( f ’)/​1.22-​1.36(g) (6.76-​9.39)(k)/​ 6.94/​6.63/​
8.39 ± 0.41 6.97(l);
6.0±0.5(l’’)
0.78(g) 1.01 (min.0.12 7-​8(i’’) 4.95±0.11 4.73/​4.22/​
(min0.12max.1.08) max.1.25) ( f ’)/​ (3.26-​5.72)(k)/​ 5.02/​4.64/​
0.77-​0.84(g) 6.02 ± 0.35 4.41(l);
4.5±0.6(l’’)
NR 1.32-​1.39(g) 13.91-​ NR NR
20.13(i’) NR NR
0.24 (min.0.01 0.41( f ’) NR NR NR
max.0.40)(g) (min.0.01max.0.70)

0.18 (min.0.29 0.88 NR NR NR


max.0.80)(g) (min.0.13max.0.83) ( f ’)/​
1.19-​1.42(g)
NR NR NR NR NR

NR NR NR NR NR

NR NR NR NR NR

0.95(g) 1.40( f) NR NR NR

0.27(g) 0.81( f ’)/​ 1.1-​1.69(g) NR NR NR

2.05(g) 5.29( f ’)/​ 3.78-​7.01(g) NR NR NR

80.70(g) (min.82.75 60.56 NR NR NR


max.92.3) (min.55.3max.67.2)( f ’)/​
48.0-​81.8(g)
(continued)
132 Structure and Composition of Latin-American Crops

TABLE 4.2 (CONTINUED)


Chenopodium
Wheat(a) quinoa
varieties Amaranthus Willd(c-​e)
Parameter Units S/​D/​N*** Maize(b) cruentus(c-​d) L/​M/​S*
Coefficient of 0.41 ± NR NR 0.211-​0.265(c)
friction 0.0(a’) NR
Ply-​wood 0.36/​0.33/​ NR NR 0.145-​0.240(c)
0.36
Galvanized NR NR NR NR
iron(e)/​steel/​
metal sheet(a)
Concrete NR NR NR NR
Angle of repose º NR NR 22.7-​30.6 (c’) 18-​25(c)
Thermal velocity m/​s NR NR NR 0.6-​1.02(c)
Thermal x10 -​ NR NR NR NR
diffusivity 3 mm2/​

s
Electrical µS/​cm NR NR NR NR
conductivity

Notes: (a) Babić et al. (2011); (a’) Shafaei and Kamgar (2017).

(b) Data for five varieties of Peruvian maize grown in the Andean region: Chullpi (Ch),
Piscorunto(P), Giant Cuzco(GC), Sacsa (S) and Purple (P) varieties and yellow corn (Y) from
Salhuana (2004) and Salvador-​Reyes and Clerici (2020).

(c) Abalone et al. (2004); (c’) Ogrodowska et al. (2014); (c’’) Brust et al. (2014).

(d) Abugoch et al. (2009).

(e) Vilche et al. (2003).

( f) Data for two varieties of kaniwa (Cupi, Ramis) from Repo-​Carrasco-​Valencia et al. (2009);
( f ’) Data for four ecotypes of kaŽiwa from the Agronomical Experimental Station-​INIA Salcedo,
Puno, Peru (Kello//​Wila/​Guinda/​Ayara) from Repo-​Carrasco-​Valencia et al. (2010); ( f ’’) Data for
Peruvian ecotypes Roja/​Blanca/​Amarilla/​Illpa-​Inia from La Rosa et al. (2016).

(g) Data from Suleiman et al. (2019) for kaŽiwa and chia as moisture content increased from 10 to
20% d.b.

(h) Ixtaina et al. (2008); Coelho and Salas-​Mellado (2014).

(i) Gutiérrez et al. (2011) and Wanga et al. (2018); (i’)Data for 25 of sacha inchi, Plukenetia volublilis
from Rodrigues et al. (2018); (i’’) Data from Kodahl and Sørensen (2021).
4.3 Physical Properties 133

Chenopodium Plukentia Lupinus Phaseolus


pallidicaule ( f-​g) Salva hispanica L.( f-​h) volubilis(i) mutabilis (j-​k) vulgaris(l-​m)
NR NR NR NR NR

NR 0.28±0.01(g) NR NR NR

NR NR NR NR NR

NR NR NR NR NR
NR 16-​18(g) NR NR NR
NR NR NR NR NR
92.0/​91.0/​91.0 (g) 93.0/​105/​93.0(g) NR NR NR

NR NR NR NR 75.9/​91.6/​
80.5/​82.2/​
80.9(l)
(j) Data for three bitter Lupinus mutabilis genotypes originating from different regions of Peru
(Altagracia from Ancash, Andenes from Cusco, and Yunguyo from Puno) from Córdova-​Ramos
et al. (2020).

(k) Data for 33 lupin ecotypes from different Peruvian regions from Berru et al. (2021) and from
Miano et al. (2015); (k’)Neves Martins et al. (2016); (k’’)Córdova-​Ramos et al. (2020); Mohamed and
Rayas (1995).

(l) Data for five black bean cultivars (Campeiro/​Esplendor/​Grafite/​Supremo/​Valente)


from Vanier et al. (2019); (l’) Data from USDA (2010); Lim (2012); (l’’) Berrios et al.
(1999).
(m) Data for Black beans (Phaseolus vulgaris) var. “San Luis” (BB) from de la Rosa-​
Millán et al. (2019); (m’)Data from Audu and Aremu (2015).
*L/​M/​S: Large˃2.0 mm (27.4%)/​Medium 1.7-​2.0 mm (72.3%)/​Small<1.7 mm
(0.3%);***S/​D/​N: Simonida/​Dragana/​NS40S varieties of Triticum aestivum, NR: not
reported.
134 Structure and Composition of Latin-American Crops

4.4 KERNEL STRUCTURES
As indicated before, structurally, the seeds of Latin-​American crops are composed
of different parts as they belong to different categories (cereals, pseudo-​cereals,
legumes, and oleaginous); these crops are shown in Figure 4.2. Wheat is included
for comparison purposes.
Maize kernel, as indicated before, is divided into layers: pericarp, aleurone,
endosperm, and germ, which reflect in general the structure of common cereals
and pseudo-​cereals (Figure 4.2). Different Latin-​American races of maize differ
in shape and size. Chullpi and Piscorunto kernels present an elongated shape
and low thickness (width 7.93–​9.54 mm, length 18.28–​18.47 mm, and thickness

Pericarp
Endosperm

Germ
Zea mays

Embryo

Tip cap

A B C

Figure 4.2 A. Piscorunto;a B. Chullpi;a C. Giant Cuzco;a D. Sacsa;a E. Purple maize;a


F. Latin-​American maize. Source: (A). Adapted from www.shutt​erst​ock.com/​image-​
photo/​peruv​ian-​chu​lpe-​maiz-​popc​orn-​unpop​ped-​676249​492.
4.4 Kernel Structures 135

Figure 4.2 (continued)


(B). Adapted from www.shutt​erst​ock.com/​image-​photo/​giant-​white-​corn-​bac​kgro​
und-​uses-​166618​307.
(C). Adapted from www.shutt​erst​ock.com/​image-​photo/​pur​ple-​corn-​maize-​isola​
ted-​on-​white-​144​6030​731.
(D). Authors’ own work.
Notes: The figure is composed of four parts, in the following order: structure of the grain,
up close images of three types of maize grains (A, B, and C), details of four different Latin-​
American maize grains (each one from a frontward and backward perspective (D)) and
details of three types of Latin-​American grains with their sizes (E).
Common maize grain illustration of its structure, seen from a cross-​section of the
seed. The image shows a tooth-​like shape surrounded by a structure called the pericarp.
Moreover, the grain shows a tip cap at the bottom. In the upper part of the tip cap there
is a vertical oval shape section called germ and inside of it the embryo is found. Finally, in
the area between the outside of the germ and the pericarp is the endosperm.
A) Up close chullpi maize seeds with an elongated tooth-​like-​shape in a yellow color.
B) Up close image of giant white maize seeds with a rounded tooth-​like-​shape.
C) Up close purple maize seeds with a rounded tooth-​like-​shape; the top half being
purple and the bottom part white.
D) Four different color types of Latin-​American maize (beige, caramel, purple, and an
orange-​yellow color) showing both their frontward and backward views. They have an
approximate range of 0.6 cm–​1 cm.
E) There are three types of maize. The first one has a ball-​like shape, range of 1.2 cm, in a
yellow color. The second one has a ball-​like shape, range of 1 cm, in a deeper yellow color.
The third one has a rounded tooth-​like shape, range of 0.7 cm, and in an orange color.

5.17–​5.74 mm). Meanwhile, Giant Cuzco, Sacsa, and Purple have circular shapes
and different sizes (width 11.37–​17.34 mm, length 12.62–​20.36 mm, and thickness
5.18–​6.13 mm) (Salvador-​Reyes et al., 2021). It has been suggested that the
differences may be linked to evolutionary factors (Paliwal et al., 2000) and that
the evolved races (Giant Cuzco, Sacsa, and Purple) present larger grains and more
homogeneous shapes than the primitive Chullpi and Piscorunto, whose kernels
are smaller and irregularly shaped. The hectoliter weight (50.06–​61.56 kg/100 L)
of Latin-​American maize was lower in comparison to the values for commercial
corn (69 to 75 kg/100 L) (UNALM/​MINAGRI, 2014). This lower density might be
dependent on the strict environmental conditions, plus genetic factors, of the
Andean highlands (i.e., temperature variation, water availability), which might
decrease the period of filling the grain, reducing its final weight and the amount
of cells accumulating starch and proteins (Salvador-​Reyes et al., 2021), giving as
a result a lower protein content.
The endosperm, embryo, and seed coat (pericarp) are the fundamental
structures of pseudo-​ cereals (amaranth, quinoa, and kañiwa). Starch and
proteins are mainly stored in the endosperm, while oil, some proteins, and
minerals occur in embryo. In the pericarp, deposition of mainly cellulose
136 Structure and Composition of Latin-American Crops

and hemicellulose, together with protein and lignin, takes place (Baltensperger,
2003; Reguera and Haros, 2017).
The color of amaranth’s pericarp differs depending on the species and can be
cream, yellow, white, pale brown, dark brown, black, red, or pink (Aderibigbe
et al., 2022). Its surface is glossy, smooth, and a bit straighter with a lens-​like
shape. The length of kernels might be between 1.3 and 1.7 mm, width in the
range of 0.9 to 1.3 mm, while its weight is usually in the frame of 0.6 to 1.0 mg
(Reguera and Haros, 2017). The kernels of amaranth are smaller than cereal
grains (wheat or maize), some pseudo-​cereals (quinoa), Andean lupins, or ole-
aginous seeds (sacha inchi) (Figure 4.3–​4.5, Table 4.2). Quinoa seeds are yellow,

A
Amaranthus spp.

Pericarp
Cotyledons
Endosperm
Shoot Apex
Starchy
Perisperm Procambium Radicle

Pericarp coat

Cotyledons
Radicle
Starchy Perisperm
Endosperm
m
Chenopodium quinoa

Shoot
h tA pex
Apex

C
Chenopodium pallidicaule

Radicle

Starchy Perisperm Cotyledons


Endosperm
Pericarp coat

Figure 4.3 Pseudocereals from Latin-​America: A. Amaranth;a B. Quinoa;a C. Cañahua,


Cañihua o Kañiwa.b Source: Seed structure adapted from Valcárcel-​Yamani and
Lannes (2012) and bBruno et al. (2018).
Images A, B, C: Authors’ own work.
4.4 Kernel Structures 137

Figure 4.3 (continued)


Notes: The figure is composed of three parts, the structure (seen from a cross-​section of
the seed) and images of amaranth (A), quinoa (B) and kañiwa(C).
A)
Illustration of the structure of the pseudocereal amaranth. It has an oval shape,
surrounded by a pericarp, followed by another layer called endosperm, followed by the
germ with two cotyledons and an inner starchy perisperm. Next to the structure, there
is an upclose image and a detailed image (the rounded shape can be observed and the
germ surrounding it) of both white amaranth (approximately 1.1mm) and black amar-
anth (approximately 0.94-​1.3 mm).
B)
Illustration of the structure of the pseudocereal quinoa. It has a round shape, surrounded
by a pericarp coat, followed by another layer called endosperm, followed by the germ
with two cotyledons and an inner starchy perisperm. Next to the structure, there are five
up close images of quinoa and their detailed images: white (approximately 2.5 mm), red
(approximately 2.4 mm), black (approximately 1.3 mm), grey (approximately 1.75 mm)
and beige (approximately 2 mm). Their rounded shape can be observed and the germ
surrounding it.
C)
Illustration of the structure of the pseudocereal kañiwa. It has a round shape, surrounded
by a pericarp coat, followed by another layer called endosperm, followed by the germ with
two cotyledons and an inner starchy perisperm. Next to the structure, there is an up close
image of kañiwa and a detailed image. Their rounded shape can be observed and the germ
surrounding it, they have an approximate range of 1-​1.25mm.

white, orange, red, brown, or black depending on the crop, especially in the case
of wild species; and the pericarp takes on a black tint (Figure 4.3; Saturni et al.,
2010; Reguera and Haros, 2017). Quinoa is a small, dry, one-​seeded fruit with
round-​shaped seeds and diameters that vary between 1.0 and 2.6 mm, together
with 250–​500 seeds per gram (Vilche et al., 2003; Valencia-​Chamorro, 2003;
Reguera and Haros, 2017). The longitudinal section of the seed is round, because
the length and width are roughly equal (Vilche et al., 2003; Reguera and Haros,
2017; Table 4.2).
Kañiwa is not a true cereal equivalent to quinoa. This annual, herbaceous
plant is 0.2–​0.7 m tall and its seeds are 1.0–​1.2 mm long, 0.12–​1.2 mm wide, and
0.12–​1.08 mm thick (Repo-​Carrasco-​Valencia et al., 2009; Figure 4.3, Table 4.2).
The color of pericarp varies depending on the ecotype and can be yellow, brown,
gray, or black (Repo-​Carrasco-​Valencia et al., 2010; La Rosa et al., 2016).
Sacha inchi is characterized by a star-​shaped fruit capsule (3–​5 cm). With
fruit maturation, the color changes from green to blackish brown (Wang et al.,
2018). The fruit capsules contain edible dark brown oval seeds, approximately
1.8×0.8×1.6 cm in size, and the testa is hard and brown, with dark brown
markings (Figure 4.4) (Kodahl and Sørensen, 2021). Chia seeds are generally
138 Structure and Composition of Latin-American Crops

very small. They are about 2 mm long by 1.5 mm wide and less than 1 mm thick,
oval in shape and shiny (Figure 4.4). The color changes from black, gray, or black
spotted to white (Hernández-​Gómez et al., 2008; Knez Hrnčič et al., 2020).
The main parts of L. mutabilis are pericarp (seed coat), cotyledon, radicle, hilar
rin, and hilar fissure (Figure 4.5; Miano et al., 2015). The broad variation exists
for pericarp color, ranging from white to cream to brown and to black (Berru

A
Testa
Perisperm
Salvia hispanica

Endosperm

Cotyledons
Radicle

B
Plukenetia volubilis

Kernel

Inner soft tissue

Outer hard
shelltissue

Figure 4.4 Latin-​American oilseeds: A. Chia;a B. Sacha inchi.b Source: Seed structure
adapted from www.lad​obe.com.mx/​2011/​09/​las-​semil​las-​explo​rado​res-​sol​itar​ios/​
and bhttps://​es.wikipe​dia.org/​wiki/​Pluke​neti​a_​vo​lubi​lis (*). www.shutt​erst​ock.com/​
image-​photo/​sacha-​inchi-​pea​nut-​seed-​isol​ate-​on-​146​0624​984 and www.shutt​erst​
ock.com/​image-​photo/​sacha-​inchi-​isola​ted-​on-​white-​bac​kgro​und-​625267​805.
Images A and B: Authors’ own work.
Notes: The figure is composed of two parts, the structure (seen from a cross-​section of the
seed) and images of chia (A) and sacha inchi (B).
A) Illustration of the structure of chia. It has an oval shape, surrounded by the testa,
followed by another layer called perisperm in the inner section of which is found the
endosperm. At the bottom the two cotyledons are found. Next to the structure, there are
two up-​close images of white (approximate height of 1.85 mm and width of 1.2 mm) and
black chia (approximate height of 2.2 mm and width of 1.4 mm) along with a detailed
image for each.
B) Illustration of the structure of sacha inchi. It has a circular shape, surrounded by
outer hard shell tissue and an inner soft tissue. There are dark brown seeds with oval ribs
of 1.5–​2 cm on the outer layer. Next to the structure, there is an up-​close image along with
a cross-​section of the image (1.3–​1.5 cm).
4.4 Kernel Structures 139

Seed coat
Phaseolus vulgaris

Cotyledons
Radicle
Hipocotyl
Radicle
Epycotil

5 mm
Seed coat B
Hilar rin
Lupinus mutabilis

Hilar fissure

Radicle

Cotyledon

Figure 4.5 Latin-​American legumes: A. Black Turtle Beans;a B. Lupin, Tarwi, Altramuz
or Chocho.b Source: aSeed structure adapted from https://​bota​nica​lgar​den.berke​ley.
edu/​glad-​you-​asked/​seeds-​p1 and bMiano et al. (2015).
Image A: Authors’ own work.
Image B: adapted from https://​en.wikipe​dia.org/​wiki/​Lupinu​s_​mu​tabi​lis.
Notes: The figure is composed of two parts, the structure and images of (A) black turtle
beans (A) and (B) lupin.
A) Illustration of a cross-​section of the structure of black turtle bean. It has a common
bean shape, surrounded by a seed coat, followed by two cotyledons. Next to the structure,
there are two up-​close images of black turtle beans and a detailed image. The first bean
has an approximate height of 1.2 cm and width of 0.7 cm; the second darker bean has an
approximate height of 1.1 cm and width of 0.7 cm.
B) Illustration of the structure of lupin. It has a more rounded bean shape, surrounded
by a seed coat, and one cotyledon (there are two cotyledons, but as it is not a cross-​
section, only one of them can be seen). Next to the structure, there are three different col-
ored detailed images of tarwi (white, white with dark spots, and mostly dark spots). Tarwi
has an approximate height of 0.9 cm and width of 0.7 cm.

et al., 2021). The color of the pericarp influences the color of the corresponding
dish which may have some impact on the processing industry/​consumer prefer-
ence. The seed coat has three cell layers (Miano et al., 2015). The external layer
resistant to water is formed by macrosclereid cells (palisade tissue) containing
lignin, polysaccharides, pectin, calose, quinones, suberin, cutin, and phenols.
The second layer is formed by osteosclereid cells, while the third layer is a
sclerified parenchyma.
The seeds of P. vulgaris are small, glossy, with a dense meaty flavour similar
to that of mushrooms and are nutritionally rich especially in antioxidants (Lim,
140 Structure and Composition of Latin-American Crops

2012). They are mottled black, plump, oblong, or kidney-​shaped: 9.1–​10.93


(length) × 6–​7.05 (width) × 4.22–​5.0 mm (thick) (Table 4.2; Berrios et al., 1999;
Vanier et al., 2019). Some structural characteristics, external and internal, of
P. vulgaris are shown in Figure 4.5 A

4.5 CHEMICAL COMPOSITION OF KERNELS


The proximate composition of Latin-​American seeds is shown in Table 4.3. The
chemical composition of proteins, lipids, carbohydrates, fiber, and bioactive
compounds are well covered in other chapters.

4.5.1 Proteins
The nutritional value of cereals, pseudo-​ cereals, oleaginous seeds, and
legumes is strongly correlated with their protein content and its composition
(Schöenlechner et al., 2008; Reguera and Haros, 2017). Pseudo-​cereals, oilseeds,
and legumes are extraordinarily valuable sources of protein, having a well-​
balanced amino acid set, with a mostly high content of lysine and a sufficient
amount of sulfur-​containing amino acids (Reguera and Haros, 2017). It is worth
mentioning that the amino acid profile of cereals, like maize, plays a remark-
ably nutritional complementary role with that of legumes, like common beans
(Paredes-​López et al., 1984). The proteins in pseudo-​cereals (quinoa, amaranth,
kañiwa), chia, and legumes predominantly consist of globulins and albumins,
and comprise lower content of, or even lack of storage of prolamin proteins,
which constitute the main proteins in cereals, and the detrimental ones in the
case of celiac disease (Alvarez-​Jubete et al., 2010; Sandoval-​Oliveros and Paredes-​
Lopez, 2013; Reguera and Haros, 2017).
Table 4.3 shows variations in the levels of protein content in the different
cultivars of maize crops; interestingly, maize samples from the Andean region
were found to have high levels of protein as reported by Salvador-​Reyes et al.
(2021). Changes in protein content, and in composition in general, is a function
of the specific cultivar, genetically modified crops, nutrients available in the soil,
and environmental influences (Rascon-​Cruz et al., 2004).
Some studies have reported that the Chullpi variety showed the highest pro-
tein content among all kinds included in this work, which might be explained
by the greater amount of protein that vitreous endosperm has compared to the
floury samples (Salvador-​Reyes et al., 2020). All essential amino acids in Latin-​
American maize varieties are presented, with leucine being the most abundant.
The limiting amino acid, as expected, was lysine in all varieties together with
valine and isoleucine in Cuzco and Sacsa. Significant amounts of tryptophan
were detected in Chullpi and Purple maizes, giving them an advantage over trad-
itional maize, in which this amino acid is at a low level (Rascon-​Cruz et al. 2004;
4.5 Chemical Composition of Kernels 141

Badui, 2006). Chapter 8 reports the further and more detailed protein compos-
ition of maize from the Latin-​American region.
Commonly, amaranth is characterized by higher protein, also lysine, content
than cereals (i.e., wheat, maize) or other pseudo-​cereals (quinoa or kañiwa) but
when compared with some Latin-​American crops (oleaginous seeds, legumes), it
may exhibit considerably lower amounts (Table 4.3). The exogenous amino acid
content of amaranth is superior and its mutual proportions are more favorably
balanced than in most cereals (Ballabio et al., 2011; Reguera and Haros, 2017;
Joshi et al., 2018; Segura-​Nieto et al., 1994). Globulin and glutelins are the main
protein fractions, followed by albumin and prolamins. These proteins have
important bioactive peptides which may act as modulators of metabolism and
possess other outstanding biological activities, such as antihypertensive and
antioxidant properties (Orona-​Tamayo and Paredes-​López, 2017). Proteins are
influenced by growing factors and genotype (Schöenlechner et al., 2008; Reguera
and Haros, 2017). Mostly, proteins are deposited in the germ and seed coat
(~65%), while the rest are located in the endosperm (35%) (Saunders and Becker,
1984; Reguera and Haros, 2017). The protein amount of quinoa is greater than
cereals (Table 4.3) and characterized by high biological value (83%), similar to
that of milk protein (Ranhotra et al., 1993; Reguera and Haros, 2017) delivering
considerable amounts of all essential amino acids (Ruales and Nair, 1992;
Abugoch James, 2009; Gonzalez et al., 2012).
Compared to cereal grains, quinoa proteins are particularly rich in lysine,
a limiting amino acid in most cereals. Their balance of essential amino acids
is perfect due to a wider range of amino acids than that found in grains and
legumes, which includes not only higher lysine amount but also higher methio-
nine (Ruales and Nair, 1993; Abugoch James, 2009). Similarly, the protein con-
tent of kañiwa (14.0–​15.7%) is close to quinoa. The protein quality of kañiwa
is exceptional with a balanced essential amino acid composition, including
5–​6% lysine, which is typically limiting in cereal grain crops (Mangelson et al.,
2019; Peńarrieta et al., 2008). The protein composition and amino acid profile of
­amaranth, quinoa, and kañiwa are also extensively presented in Chapter 8.
Oleaginous seeds are a great source of protein, which seems even better than
that of cereals and pseudo-​cereals. The protein content of sacha inchi ranged
between 24.2 and 27.0% (Hamaker et al., 1992; Gutiérrez et al., 2011). According to
Hamaker et al. (1992), leucine (64 mg/​g of protein) is the most abundant essential
amino acid; next are tyrosine, isoleucine, lysine, threonine and valine (55, 50, 43,
43, and 40 mg/​g of protein, respectively). These proteins have sulfur amino acids
(methionine+​cysteine) by 37 mg/​g and phenylalanine by 9 mg/​g (Hamaker et al.,
1992). Albumin (43.7%) is the predominant aqueous soluble protein, followed by
glutelin (31.9%), globulin (27.3%), and prolamin (3.0%) (Sathe et al., 2002).
Chia seeds are rich in protein (16.0–​28.6% d.b.) and the essential amino acids
of seed flour exhibited in general a relatively good balance of them, especially
methionine and cysteine (Sandoval-​Oliveros and Paredes-​López, 2013). The
142 Structure and Composition of Latin-American Crops

main protein fraction corresponded to globulins (52%) which are also a good
source of aromatic amino acids (Sandoval-​Oliveros and Paredes-​López, 2013).
The exogenous aminoacids are arginine, leucine, phenylalanine, valine, and
lysine; and the endogenous amino acids are glutamic and aspartic acids, alanine,
serine, and glycine. The content of amino acid serine is 1.05 g/​100 g, glutamic
acid 3.50 g/​100 g, glycine 0.95 g/​100 g, alanine 1.05 g/​100 g, lysine 0.97 g/​100 g,
and histidine 0.53 g/​100 g of seeds (Ullah et al., 2016). Composition has been
found to be affected by factors such as genotype, environment, and agronomical
input (Jamshidi et al., 2019). Chia seeds do not contain the proteins that make
up gluten, which makes them very valuable for people suffering from celiac dis-
ease. The protein composition and amino acid profile of oleaginous seeds are
also extensively presented in Chapter 8.
The protein content of tarwi is the highest among all Latin-​American crops,
and reaches up to 52.6% d.b. (Table 4.3). It is rich in lysine and cysteine and
comprises two protein classes –​globulins (80% of total protein) and albumins
(Brücher, 1989; Chirinos et al., 2021). Its seeds have low amounts of sulphured
amino acids (methionine and cysteine) when considering human nutrition; how-
ever, its proteins show higher biological values than those of other species of
Lupinus (L. albus, L. angustifolius and L. cosentinii) and are comparable to soybean
proteins. The proteins digestibility is high (>90%) (Neves-​Martins et al., 2016).
The amount of protein (up to 25% d.b.) (Priya and Manickavasagan, 2020) in
black turtle bean is higher than in cereals (Zea mays), pseudo-​cereals (kañiwa,
quinoa, amaranth), and comparable to plants of oleaginous seeds (chia, sacha
inchi) but lower than in tarwi seeds (Table 4.3). The amino acid lysine (1.483%
d.b.) is greater than in other crops, such as maize. However, the proteins are poor
in methionine (0.325% d.b.) and tryptophan (0.256% d.b.) (Lim, 2012). It is per-
tinent to mention that several of the Latin-​American cultivars (e.g., amaranth,
chia, common beans) contain peptides of different lengths and sequence which
once hydrolyzed from proteins by different techniques show a wide variety of
remarkable nutraceutical and medicinal properties, inlcuding roles as vegetable
vaccines, with great commercial potential in the near future (Orona-​Tamayo
and Paredes-​López, 2017). The composition and amino acid profile of Latin-​
American legumes are also extensively described in Chapter 8.

4.5.2 Carbohydrates
The most typical simple sugars of grains are glucose, fructose, arabinose,
and xylose, together with sucrose and maltose (Reguera and Haros, 2017),
which amount is rather small in cereals and pseudo-​ cereals (Berghofer
and Schöenlechner, 2007; Reguera and Haros, 2017). Starch, non-​ starch
polysaccharides and resistant starch are the main polysaccharides in cereals,
whereas non-​ starch polysaccharides found in kernels are predominantly
composed of β-​glucans, cellulose, and hemicelluloses, which are included as part
4.5 Chemical Composition of Kernels 143

of dietary fiber together with resistant starch (Reguera and Haros, 2017). The
amount of dietary fiber in pseudo-​cereals and maize is similar and found to be
like that in other cereals (Table 4.3), together with starch as the most crucial plant
carbohydrate occurring in different shapes and sizes (Valcárcel-​Yamani et al.,
2012; Reguera and Haros, 2017). Oleaginous seeds of chia and sacha inchi con-
tain mainly complex carbohydrates in the form of insoluble fractions (72.4% and
93%, respectively) (Wang et al., 2018; Jamshidi et al., 2019; Kodahl, 2020). Latin-​
American legumes are predominantly composed of complex carbohydrates but
differ in their composition. Black turtle bean contains starch and dietary fiber in
almost equal amounts, while tarwi is composed mainly of oligosaccharides. The
common feature of both legumes is that they are rich in dietary fiber and contain
small amounts of soluble sugars.
Maize is characterized by high amounts of starch, which is the main carbo-
hydrate in its endosperm (Table 4.3). The quantity is higher than in other cereals
and in other Latin-​American crops (pseudo-​cereals, oleaginous seeds, legumes).
Starch granules differ in their shape, size, and distribution depending on the
cultivar –​ Chullpi, Sacsa, and Purple starches are spherical, whereas those of
Piscorunto and Giant Cuzco are irregular and polyhedral. Sacsa had a bimodal
distribution presenting large granules type-​A (>10 μm) and small granules type-​
B (<10 μm), whereas in other varieties only type-​A granules were observed.
Starch granules of Giant Cuzco, Purple, Sacsa, and Chullpi were larger than those
of Piscorunto (Salvador-​Reyes et al., 2021).
The starch percentage of pseudo-​cereals varies between 38.03 and 70.40
(Table 4.3). In amaranth, starch is the main carbohydrate component, but its
content is lower than in cereals (Valcárcel-​Yamani et al., 2012; Reguera and
Haros, 2017; Table 4.3). It is situated in the perisperm. The granules of amaranth
starch (1-​3 µm) are smaller than in other cereals (Berghofer and Schöenlechner,
2007; Valcárcel-​ Yamani et al., 2012; Reguera and Haros, 2017), and with
­polygonal, circular, and elliptical shapes according to their species (López et al.,
1994). Amaranth starch is composed of smaller amylose amounts (0.1–​11.1%)
than those found in other cereals and characterized by normal and waxy types
of starch contained in the same species of amaranth (Stone and Lorenz, 1984;
Schöenlechner et al., 2008; Reguera and Haros, 2017). It has also been reported
that cultivation and environmental conditions of amaranth affect the amylose/​
amylopectin ratio (Stone and Lorenz, 1984). Similarly, in quinoa, starch is also
the most abundant carbohydrate (Table 4.3) with a polygonal shape and a diam-
eter of 1.5–​3.0 µm, being smaller than starch of typical grains (Koziol, 1992; Vega-​
Gálvez et al., 2010; Reguera and Haros, 2017). The amount of amylose in quinoa
grains is smaller (11.0–​12.4%) than that reported in wheat, rice, or maize (Koziol,
1992; Reguera and Haros, 2017).
Starch is the most abundant component, not only in quinoa and amaranth, but
also in kañiwa (Repo-​Carrasco-​Valencia et al., 2010). Granules of starch kañiwa
are in polygonal form with a diameter between 0.7 and 1.3 µm (Luna-​Mercado
144 Structure and Composition of Latin-American Crops

and Repo-​Carrasco-​Valencia, 2021), being smaller when compared to cereal


grains and pseudo-​cereals (amaranth, quinoa). Kañiwa starch is found mainly
in the perisperm surrounded by the embryo and is strongly associated with pro-
tein and other components of the grain. It was found that the amylose content
of kañiwa starch variety Illpa Inia was 6.48% (Luna-​Mercado and Repo-​Carrasco-​
Valencia, 2021). Higher values of amylose in kañiwa starch were also reported
elsewhere (10.7–​17.44%) (Steffolani et al., 2013). More descriptions of the starch
characteristics of maize and pseudo-​cereal starch are explained in Chapter 6.
The carbohydrate composition of oleaginous seeds included in this report
needs more study. Information about the carbohydrate content and composition
of sacha inchi (P. volubilis) seeds is very limited (Kodahl, 2020). According to Wang
et al. (2018), the carbohydrate amount ranged from 12.1–​30.9% and results were
obtained by simply subtracting the moisture, lipid, protein, and ash amounts
from the total weight. Takeyama and Fukushima (2013) reported that the carbo-
hydrate fraction of sacha inchi consists of 18.7% of glucides (Carbohydrate –​(IDF
+​SDF), 72.3% water insoluble dietary fiber (IDF), and 9.0% soluble dietary fiber
(SDF)). However, the exact composition of the fiber and the existence of starch
in the seeds needs further research (Wang et al., 2018; Kodahl, 2020). Even less is
known on carbohydrates composition in chia seeds (16.5–​42.12% d.b.) with no
starch detected (Vrancheva et al., 2019). Most of the literature data deals about
dietary fiber (18–​37.5% d.b.) (Table 4.3; Ayerza, 1995; Ayerza and Coates, 2004;
Ixtaina et al., 2011; USDA, 2011; Da Silva Marineli et al., 2014; Coelho and Salas-​
Mellado, 2014; Amato et al., 2015; García-​Salcedo et al., 2018; Carillo et al., 2018;
Jamshidi et al., 2019; Knez Hrnčič et al., 2020; Hernández-​Pérez et al., 2020). The
dietary fiber of chia is composed mainly of insoluble fractions (>93%) and the
rest (7%) is soluble form (Jamshidi et al., 2019). Interesting is the presence of a
polysaccharide with a molecular weight of 0.8–​2×106 Da (tetrasaccharide with
4-​O-​methyl-​α-​D-​glucoronopyranosyl residues), which is situated in the fruit
exocarp and has the ability to form a highly hydrophilic mucilage (Bochicchio
et al., 2015). White and black-​spotted chia samples from Salta, Argentina showed
a mucilage composition with mannose (1.48 and 1.20%), galactose (3.26 and
3.16%), galacturonic acid (4.60 and 3.58%), glucose (7.41 and 5.88%), arabinose
(9.48 and 9.80%), glucuronic acid (12.74 and 13.72%), xylose (61.04–​63.07%), and
with water capacity absorption 54.24 and 54.03, respectively (Muñoz et al., 2021).
Latin-​American legumes differ in the amount and composition of
carbohydrates. Black turtle bean is mainly composed of complex carbohydrates
(up to 65% d.b.) (Priya and Manickavasagan, 2020). Nearly as much starch (18.17–​
28.06% d.b.) as dietary fiber (17.63–​22.83% d.b.) may be found depending on var-
ieties (Table 4.3; USDA, 2010; Lim, 2012; Vanier et al., 2019). Among total sugars
(2.12% d.b.) (Lim, 2012), monosaccharides, disaccharides, and oligosaccharides
are found. The common oligosaccharides are raffinose, stachyose, and
verbascose. The black turtle starch is granular, round and oval in shape with
newgenrtpdf
TABLE 4.3 CHEMICAL COMPOSITION OF LATIN-​AMERICAN CROPS AND WHEAT (FOR COMPARISON PURPOSES)
Dietary fiber,
CROPS Protein, % d.b. Lipids, % d.b. Starch, % d.b. % d.b. Ash, % d.b.

CEREALS
Wheat 11.6( f:5.70)(c), 1.7(c), 2.0(e), 2.3(h); 61.0(e), 78.4(h) 2.8(h), 6.5(g); 1.4(c), 2.2(h)
11.7(e), 14.3( f:6.25)(h), 13.20(a’) 2.7(a’) 12.20(a’)
maize 7.80/​7.27/​5.10/​ 7.38 and 2.70/​3.93/​3.52/​4.50 74.60/​69.37/​ 4.30/​1.86/​1.12/​ 1.30/​1.46/​1.38/​
5.7( f:6.25)(i) and 4.8(i) 74.80/​70.53 and 1.87 and 3.8(i) 1.50 and 1.40(i)
72.00(i)
PSEUDO-​CEREALS
Amaranthus 14.0( f:5.85)(b), 5.6(c), 5.7(a), 5.9-​6.0(g), 55.1( f), 61.4(a), 11.1( f), 20.6(a); 2.8(a), 2.4(b),
cruentus 14.0-​14.8( f:5.85)(g), 6.0(b), 8.0(e), 8.8( f); 67.3(e) 3.10–​4.20, 2.9(c), 3.3( f),
14.6( f.5.8)( f), 14.9( f:5.70)(c), 5.60–​8.10, 6.10–​ 4.90–​5.00, 2.70–​ 2.4-​2.6(g),2.8-​3.9(s)

4.5
15.2(e), 13.2-​18.2( f:5.85)(s), 7.30; 5.80–​10.90(a’), 4.90(a’), 3.6-​4.4(s)
16.5( f:5.85)(a), 13.80–​21.50; 6.3-​8.1(s)

Chemical Composition of Kernels


15.00–​16.60; 13.10–​21.00(a’)
Chenopodium 11.0( f:5.70)(d) 4.1-​5.8(g), 5.0( f), 5.2(a), 64.2(d), 3.8(h), 6.72(d), 2.7(a), 2.7(d), 3.8(h),
quinoa 12.8-​13.5( f:5.77)(g), 13.3(e), 6.3(h), 7.5(d), 7.5(e); 66.9-​70.4(g), 12.9( f), 14.2(a), 3.3( f), 2.3-​2.5(g);
13.8( f:5.8)( f), 4.0–​7.6(h’) 67.4( f), 69.0(e); 14.6-​19.7(g); 2.0–​7.7(h’)
16.5(h); 9.1–​15.7(h’) 48.5–​69.8(h’) 8.8–​14.1(h’)
Chenopodium 14.41/​14.88 (j); 15.38/​13.29/​ 5.88/​6.96(j); 7.36/​ 52.4/​57.47(j); 11.24/​8.1(j); 5.03/​4.33(m);
pallidicaule 14.72/​14.38 (j’); 15.3/​15.5/​ 6.87/​4.46/​6.66 (j’); 54.07/​52.78/​ 5.33/​7.52/​7.46/​ 3.56/​3.67/​3.38/​
(kañiwa) 14.7(j’’); 15.4/​15.4/​15.7/​13.8 8.5/​8.0/​7.6(j’’); 54.04/​38.03 (j’); 14.37(j’) 3.13 (m’);
(k) ( f:6.25) 7.5/​7.8/​7.5/​3.9 (k) 60.4/​58.5/​62.0(j’’) 5.6/​7.0/​6.0(j’’); 4.6/​4.0/​3.7(m’’)
3.7/​3.7/​3.5/​4.2 (k)
(continued)

145
newgenrtpdf
146
Structure and Composition of Latin-American Crops
TABLE 4.2 (CONTINUED)

Dietary fiber,
CROPS Protein, % d.b. Lipids, % d.b. Starch, % d.b. % d.b. Ash, % d.b.
OLEAGINOUS SEEDS
Salva 28.56 ± 0.235 ( f:6.25)/​
(l) 7.58 ± 0.028(l)/​ (CHO)17.887(l)/​ 39.87 ± 0.757(l) 0.103 ± 0.015(l)/​
hispanica L. 19.78±0.015(l’)/​ 16.06±0.038(l’)/​ 31.46±0.062(l’)/​ (soluble fiber)/​ 4.82±0.025(l’)/​
(chia) 16-​26(l’’)/​18.3 ± 1.613(l’’), 20.30 -​38.60(l’’)/​ 42.12(l’’)/​ 27.88±0.021(l’)/​ 4.3±0.035(l’’)/​
22.7±0.7(t) 32.4±0.214(l’’), 16.5±1.628(l’’)/​ 32.4-​37.50(l’’)/​ 3.7±0.3(t)
32.5±2.7(t) 3.1(t) (other 22.2 ± 0.323(l’’)/​
carbohydrates) 33.5±2.7(t)
Plukentia 24.2-​27 (m),(n)( f:6.25) 33.4-​54.3(m)(n’) 12.1-​30.9 (m),(m’),(m’’) 72.4 (m) 2.7-​6.46 (m)
volubilis (CHO)
(sacha inchi)
LEGUMES
Phaseolus 23.08/​25.36/​23.81/​26.00/​ 1.23/​1.37/​1.22/​1.68/​ 25.50/​18.17/​ 22.83/​17.63/​ 4.25/​3.94/​4.06/​
vulgaris 23.89 ( f:6.25)(r); 1.22(r); 1.42(r’); 1.95(r’’) 28.06/​24.97/​ 20.32/​21.15/​ 4.39/​4.20(r);
(black turtle 21.60(r’); 25.93 ( f:6.25)(r’’) 21.17(r) 19.93(r); 3.60(r’); 4.65(r’’)
bean) 62.36 (CHO) (r’) 15.2(r’)
48.59/​51.67/​
50.56/​46.76/​
50.74 (CHO)(r);
Lupinus 40.87 ± 0.40 16.12 ± 0.14(13.60–​ 29.45 ± 0.43 8.2(q) 3.58 ± 0.06
mubilis (32.03–​46.90)(o)( f:6.25) 18.55) (o) (24.85–​33.90)(t) (2.70–​4.40)(t)
(tarwi) 41.4–​47.7(o’) 15.0–​20.1(o’) 31.65 ± 2.47 (p’) 2.4–​5.2(p)
34.6–​50.2(o’’) 14.3–​23.6(o’’) 4.80 ± 0.21(p’)
32.0–​52.6(p) 13.0–​24.6(p)
47.36±2.80 (p’)( f:6.25) 16.19 ± 1.03(p’)
newgenrtpdf
Notes: d.m. dry basis; f: nitrogen to protein conversion factor used. [Continued]
(a) Alvarez-​Jubete et al.(2010); (a’) Data for A.cruentus, A.hypochondriacus, A.caudatus, respectively from Joshi et al. (2018).
(b) Sanz-​Penella et al. (2013) of A. cruentus.
(c) García-​Mantrana et al. (2014) of A. cruentus and T. aestivum L.
(d) Iglesias-​Puig et al. (2015).
(e) Souci et al. (2000) of A. cruentus, abrased quinoa, F. escutentum and T. aestivum L.
( f) Valcárcel-​Yamani et al. (2012).
(g) Own measurements from Amaranthus spp., Real quinoa and wheat.
(h) Koziol (1992), (h’)Data for different varieties, cultivars and ecotypes of quinoa from Nowak et al. (2016);
(i) Salhuana (2004); Salvador-​Reyes and Clerici (2020) of Chullpi (Ch), Piscorunto(P), Giant Cuzco(GC), and Purple (P) varieties and yellow corn
(Y) as control.
(j) Data for two Peruvian varieties of kaniwa (Cupi/​ Ramis) from Repo-​Carrasco-​Valencia et al. (2009); (j’) Data for four ecotypes of kaŽiwa
from the Agronomical Experimental Station-​INIA Salcedo, Puno, Peru (Kello//​Wila/​Guinda/​Ayara) from Repo-​Carrasco-​Valencia et al.
(2010); (j’’)three Peruvian ecotypes (Chilliwa/​Purple plant/​Red kañiwa Condorsaya) from Huamaní et al. (2020).
(k) Data for Peruvian ecotypes Roja/​Blanca/​Amarilla/​Illpa-​Inia from de La Rosa et al. (2016).

4.5
(l) García-​Salcedo et al. (2018); (l’) Data for Ecuadorian chia seeds from Carillo et al. (2018). (l’’) Data from Ayerza (1995); Ayerza and Coates

Chemical Composition of Kernels


(2004); Ixtaina et al. (2011); USDA, 2011; Da Silva Marineli et al. (2014); Amato et al. (2015); Jamshidi et al. (2019); Coelho and Salas-​Mellado
(2014).
(m) Gutiérrez et al. (2011), Wang et al. (2018); (m’) Ruiz et al. (2013); (m’’) Takeyama, Fukushima (2013).
(n) Hamaker et al. (1992); (n’)Kodahl and Sørensen (2021).
(o) Data for 33 Andean lupin ecotypes from different Peruvian regions from Berru et al. (2021); (o’) Data for in Ecuadorian L. mutabilis from
Schoeneberger et al. (1982); (o’’) Data in Peruvian ecotypes from Caligari et al. (2000).
(p) Carvajal-​Larenas et al. (2016); (p’) Data for three bitter Lupinus mutabilis genotypes originating from different regions of Peru (Altagracia,
from Ancash, Andenes, from Cusco, and Yunguyo, from Puno) from Córdova-​Ramos et al. (2020).
(q) Carvajal-​Larenas et al. (2016).
(r) Data for five black bean cultivars (Campeiro/​Esplendor/​Grafite/​Supremo/​Valente) (Embrapa Rice and Beans, in the city of Santo Antônio
de Goiás, State of Goiás, Brazil) from Vanier et al. (2019); (r’)Data from USDA (2010); Lim (2012); (r’’) Data from Berrios et al. (1999).
(s) Seguera-​Nieta et al. (2018).
(t) Hernández-​Pérez et al. (2020).

147
148 Structure and Composition of Latin-American Crops

indentation. The surface is smooth without fissure (Du et al., 2014). The particle
size varies between 10.0 and 60.3 µm, with a mean granule diameter of 25.3 µm.
The reported amylose content was high and equals 45.4±0.8% (Du et al., 2014). In
the seeds of tarwi, starch is usually absent, and the carbohydrates (29.45–​31.65%)
(Table 4.3.) are mainly oligosaccharides (e.g., stachyose and raffinose) and cell
wall storage polysaccharides (Trugo et al., 2003; Berru et al., 2021). The reported
mean content of soluble sugars in three Andean genotypes (Altagracia, Andenes,
Yunguyo) from different regions in Peru were as follows: glucose (0.30%), fructose
(0.52%), maltose (0.14%), reducing sugars (0.84%), and saccharose (4.34% d.b.)
(Córdova-​Ramos et al., 2020).

4.5.3 Lipids
Maize is characterized by higher lipid content than wheat and lower than the
rest of Latin-​American crops. Lipids are mainly deposited in the germ, but they
are minor compounds of endosperm as for common cereals and take a great
part in the assembly of the starch−protein matrix of grains which strongly
affect the texture of kernels (hardness). Depending on the variety/​ecotype of
common maize, the endosperm might be of vitreous type, richer in free fatty
acids than triacylglycerol or floury structure with no differences between those
two components (Gayral et al., 2015).
The amount of lipids in the Latin-​American maize samples used in this study
by Salvador-​Reyes et al. (2021), is higher than wheat and lower than yellow
maize, and other Latin-​American crops (Table 4.3). The composition of fat in
Latin-​American maize is interesting and contains in average 15% of saturated
fatty acids (mainly palmitic acid), 30% of monounsaturated (oleic) acid, and
55% of polyunsaturated, including the essential fatty acids linoleic (53%) and
α-​linolenic (1.2%) (Salvador-​Reyes et al., 2021). A 100 g portion of whole maize
fulfills the daily recommendation for linoleic essential fatty acid for adults (FAO,
2010). The Chullpi variety presented traces of behenic acid (<1%), not detected in
the other four varieties of Andean maize. The presence of this acid in the human
diet is associated with increased cholesterol concentration in the blood (Cater
and Denke, 2001).
Pseudo-​cereal kernels have an analogous arrangement of lipids in the form of
fat droplets. In amaranth, lipid bodies are situated in the embryo and the endo-
sperm, while in the case of quinoa in the embryo and perisperm (Reguera and
Haros, 2017). Lipid amounts in quinoa, amaranth, and kañiwa are approximately
2–​3 times greater than wheat and maize (Table 4.3; Alvarez-​Jubete et al., 2010;
Reguera and Haros, 2017). Generally, pseudo-​cereal lipids are characterized by a
high content of unsaturated fatty acids (between 75 and 86%) (Valcárcel-​Yamani
et al., 2012; Reguera and Haros, 2017; Huamani et al., 2020). Linoleic acid is the
most important fatty acid of pseudo-​cereals (47.5–​47.8, 48.2–​56.0; 46.9–​48.5%),
4.5 Chemical Composition of Kernels 149

followed by oleic acid (23.7–​32.9, 24.5–​26.7, and 24.8–​25.9%), and palmitic acid
(12.3–​20.9, 9.7–​11.0; 12.9–​13.5%), for amaranth, quinoa, and kañiwa, respect-
ively (Becker, 1994; Valcárcel-​Yamani et al., 2012; Salas et al. 2015; Huamani
et al., 2020; Repo-​Carrasco-​Valencia, 2020). Squalene, an open-​chain triterpene
with many unsaturated bonds is the biochemical precursor of the entire steroid
family presented in amaranth (1.9–​11.2% in its oil). The presence of 3.4–​5.8% of
squalene was also found in quinoa seeds (Valcárcel-​Yamani et al., 2012; Reguera
and Haros, 2017). Despite the high fat content and level of unsaturation, pseudo-​
cereal lipids are generally resistant to oxidation resulting from the presence
of antioxidants such as tocopherols ( Álvarez-​Jubete et al., 2010; Reguera and
Haros, 2017). Pseudo-​cereals could be part of a balanced diet regimen as a result
of their low saturated fatty acid content (Becker, 1994).
The oleaginous seeds such as sacha inchi and chia are of great interest,
because of their exceptionally high lipid content in the form of oil: 35–​60% and
25–​38%, respectively with unique unsaturated fatty acids composition (Ixtaína
et al., 2008; Gutiérrez et al., 2011). Common legumes, such as black turtle bean
(<2% d.b.) are generally low in fat. There are however some exceptions which
include, among others tarwi characterized by high lipid content (19–​24%
d.b.) (Table 4.3). Legumes are comprised mainly of mono-​and polyunsatur-
ated fatty acids, with no cholesterol and saturated fatty acids (Maphosa and
Jideani, 2017).
The lipid content of sacha inchi is very high (33.4–​54.3%) (Gutiérrez et al.,
2011; Wang et al., 2018; Kodahl and Sørensen, 2021) and its composition is unique
(neutral lipids –​97.2%, free fatty acids –​1.2% and phospholipids -​0.8%) (Gutiérrez
et al., 2011). Approximately 77.5–​84.4% polyunsaturated fatty acids (PUFAs), 8.4–​
13.2% monounsaturated fatty acids (MUFAs), and 6.8–​9.1% saturated fatty acids
(SFAs) were found in seeds (Follegatti-​Romero et al., 2009; Maurer et al., 2012;
Chirinos et al., 2013). The predominant fatty acid is α-​linolenic acid (ALA, ω-​3)
(46.8–​50.8%), then linoleic acid (ω-​6, 33.4–​36.2%) and oleic acid (ω-​9, 8.7–​9.6%)
(Guillén et al., 2003; Follegatti-​Romero et al., 2009; Chirinos et al., 2013). Very
low amounts of myristic acid (C14:0), eicosanoic acid (C20:0), and gadoleic acid
(C20:1, ω-​11, 0.16%) were also reported in sacha inchi (Follegatti-​Romer et al.,
2009; Chandrasekaran and Liu, 2015). Lipids in chia seeds constitute a significant
amount (up to 38.60% d.b.) (Table 4.3) (Ayerza, 1995; Ixtaína et al., 2011; Ayerza
and Coates, 2004; Da Silva Marineli et al., 2014; Amato et al., 2015). The main
fatty acids presented by chia are linolenic acid (18:3), linoleic acid (18:2), stearic
acid (18:0), palmitic acid (16:0), and oleic acid (18:1) (Ayerza, 1995; Ayerza and
Coates, 2004; Hernández-​Pérez et al., 2020). ALA constitutes the greatest share
(>60%), making chia one of the best plant-​based sources of omega-​3 (Ayerza and
Coates, 2001).
Together with proteins, lipids comprise (up to 24.6% d.b.) (Table 4.3) more
than half of the seed’s weight of L. mutabilis and are composed of some essential
150 Structure and Composition of Latin-American Crops

fatty-​acids like oleic, linoleic, and linolenic that represent the 40.4%, 37.1%, and
2.9% from the total, respectively (Chirinos, 2015). The oil does not have any
erucic (toxic) acid as in L. albus (Neves-​Martins et al., 2016). Tarwi is also the
only species capable of containing 18% oil content (the minimum for indus-
trial extraction) (Neves-​Martins et al., 2016). The content of lipids in black turtle
bean is very low (1.2–​1.95% d.b.) (Berrios et al., 1999; USDA, 2010; Lim, 2012;
Vanier et al., 2019), being the lowest among all Latin-​American crops and com-
parable to wheat grain (Table 4.3). The fat composition differs depending upon
the bean variety and growth conditions (Priya and Manickavasagan, 2020).
Total saturated fatty acids constitute 0.366%, including myristic (0.001%), pal-
mitic (0.343%), and stearic (0.022%). At the same time, total monounsaturated
fatty acids, mainly in the form of oleic (0.123%) and total polyunsaturated fatty
acids –​0.610%, above all linoleic (0.332%) and linolenic (0.278%) are also pre-
sent (Lim, 2012).
The composition of lipids in Latin-​American crops is extensively described in
Chapter 8.

4.6 CONCLUSIONS
Latin-​American seeds belong to different types of crops: cereals, pseudo-​cereals,
oleaginous seeds, and legumes. They differ between each other not only in terms
of physicochemical properties but also because of their exceptional nutritional
characteristics. They are a great source of high quality proteins (pseudo-​cereals,
oleaginous seeds, and legumes), an excellent source of unsaturated fatty acids
(pseudo-​cereals, oleaginous seeds, legumes), and abundant in dietary fiber (ole-
aginous seeds, black turtle bean). Because of great functional and pro-​healthy
properties, Latin-​American crops constitute a promising future alternative to
already well-​known and cultivated crops intended for human nutrition world-
wide. Despite the above, some data in the literature need completion and fur-
ther research. In other words, more thorough research on physicochemical and
functional properties, better knowledge about the content and composition of
carbohydrates and fiber of different types, and minor components fundamental
to the daily diet present in all Latin-​American genetic materials are still required.
In addition to agronomic types of studies, including traditional and molecular
agricultural techniques, it is necessary to pursue extensive in vitro and in vivo
investigation –​following strict scientific strategies –​to assess the remarkable
nutritional and nutraceutical potential of the Latin-​American crops involved in
this report. It is pertinent to underline that several of these cultivars have been
classified by official international organizations as key food sources of the 21st
century.
References 151

ACKNOWLEDGMENTS
This work was financially supported by the Integrated Program of the University
of Agriculture in Krakow and co-​financed by European Union Funds (No.
POWR.03.05.00-​00-​z222/​17), Food4ImNut Food4ImNut PID2019-107650RB-C21
funded by MCIN/AEI/10.13039/501100011033, Spain.

REFERENCES
Abalone, R., A. Cassinera, A. Gastón, and M. A. Lara. 2004. Some physical properties of
amaranth seeds. Biosystems Engineering 89: 109–​117.
Abugoch, J. L. E. 2009. Quinoa (Chenopodium quinoa Willd.): Composition, chemistry,
nutritional, and functional properties. Advances in Food and Nutrition Research
58: 1–​31.
Aderibigbe, O. R., O. O. Ezekiel, S. O. Owolade, J. K. Korese, B. Sturm, and O. Hensel. 2022.
Exploring the potentials of underutilized grain amaranth (Amaranthus spp.) along
the value chain for food and nutrition security: A review. Critical Reviews in Food
Science and Nutrition 62(3): 656–​669.
Alvarez-​Jubete, L., E. K. Arendt, and E. Gallagher. 2010. Nutritive value of pseudocereals
and their increasing use as functional gluten-​free ingredients. Trends in Food Science
& Technology 21: 106–​113.
Amato, M., M. C. Caruso, F. Guzzo, F. Galgano, M. Commisso, R. Bochicchio, R. Labella,
and F. Favati. 2015. Nutritional quality of seeds and leaf metabolites of Chia
(Salvia hispanica L.) from Southern Italy. European Food Research Technology
241: 615–​625.
Atchison, G. W., B. Nevado, R. J. Eastwood, N. Contreras-​Ortiz, C. Reyne, S. Madriñán, D.
A. Filatov, and C. E. Hughes. 2016. Lost crops of the Incas: Origins of domestica-
tion of the Andean pulse crop tarwi, Lupinus mutabilis. American Journal of Botany
103(9): 1592–​1606.
Audu, S. S., and M. O. Aremu. 2015. Effect of domestic processing on the levels of some
functional parameters in black turtle bean (Phaseolus vulgaris L). Food Science and
Quality Management www.iiste.org, ISSN 2224-​6088 (Paper) ISSN 2225-​0557), 38.
Ayerza, R. 1995. Oil content and fatty acid composition of chia (Salvia hispanica L.) from
five northwestern locations in Argentina. Journal of the American Oil Chemists’
Society 72: 1079–​1081.
Ayerza, R. 2009. The seed’s protein and oil content, fatty acid composition and growing
cycle length of a single genotype of Chia as affected by environmental factors. Journal
of Oleo Science 58(7): 347–​354.
Ayerza, R., and W. Coates. 2001. Omega-​3 enriched eggs: the influence of dietary α-​linolenic
fatty acid source on egg production and composition. Canadian Journal of Animal
Science 81: 355–​362.
Ayerza, R., and W. Coates. 2004. Protein and oil content, peroxide index and fatty acid
composition of chia (Salvia hispanica L.) grown in six tropical and subtropical
ecosystems of South America. Tropical Science 44: 131–​135.
152 Structure and Composition of Latin-American Crops

Babić, Ljiljana., Mirko Babić, Jan Turan, Snežana Matić-​Kekić, Milivoj Radojčin, Sanja
Mehandžić-​Stanišić, Ivan Pavkov, Miodrag Zoranović 2011. Physical and stress–​
strain properties of wheat (Triticum aestivum) kernel. Journal of the Science of Food
and Agriculture 91: 1236–​1243. https://​doi.org/​10.1002/​jsfa.4305.
Badui, S. 2006. Química de los alimentos (4th ed.). Mexico: Editorial Alhambra Mexicana,
pp.140–​145.
Ballabio, C., F. Uberti, C. Di Lorenzo, A. Brandolini, E. Penas, and P. Restani. 2011.
Biochemical and immunochemical characterization of different varieties of amar-
anth (Amaranthus L. ssp.) as a safe ingredient for gluten-​free products. Journal of
Agricultural and Food Chemistry 59: 12969–​12974.
Baltensperger, D. D. 2003. Cereal grains and pseudo-​cereals. In Encyclopedia of Food and
Culture, edited by S.H. Katz and W.W. Weaver. New York: Scribner, URL: www.encyc​
lope​dia.com/​doc/​1G2-​340​3400​119.html. Accessed 03.06.15.
Becker, R. 1994. Amaranth oil: Composition, processing, and nutritional qualities. In
Amaranth Biology, Chemistry, and Technology, edited by O. Paredes-​López, 133–​141.
Boca Raton, FL: CRC Press.
Berghofer, E., and R. Schoenlechner. 2007. Pseudocereals –​An overview. IFS Workshop:
Traditional grains for low environmental impact and good health, Pretoria.
Berriosa, J. D. J., B. G. Swansonb, and W. A. Cheong. 1999. Physico-​chemical character-
ization of stored black beans (Phaseolus vulgaris L.). Food Research International
32: 669–​676.
Berru, L. B., P. Glorio-​Paulet, C. Basso, A. Scarafoni, F. Camarena, A. Hidalgo, and A.
Brandolini. 2021. Chemical composition, tocopherol and carotenoid content of
seeds from different Andean Lupin (Lupinus mutabilis) Ecotypes. Plant Foods for
Human Nutrition 76: 98–​104.
Bochicchio, R., T. D. Philips, S. Lovelli, R. Labella, F. Galgano, A. Di Marisco, M. Perniola,
and M. Amato. 2015. Innovative crop productions for healthy food: The case of chia
(Salvia hispanica L.). In The Sustainability of Agro-​Food and Natural Resource Systems
in the Mediterranean Basin, edited by A. Vastola. New York: Spriger. DOI 10.1007/​
978-​3-​319-​16357-​4_​3.
Boege, E. 2009. Centros de origen, pueblos indígenas y diversificación del maíz. Ciencias
92–​93: 18–​28.
Bressani, R. 2003. Amaranth. In Encyclopedia of Food Sciences and Nutrition, edited by B.
Caballero, 166–​173. Oxford: Academic Press.
Broughton, W. J., G. Hernández, M. Blair, S. Beebe, P. Gepts, and J. Vanderleyden. 2003.
Beans (Phaseolus spp.) –​model food legumes. Plant and Soil 252: 55–​128.
Brücher, H. 1989. Lupinus mutabilis Sweet. In Useful Plants of Neotropical Origin and Their
Wild Relatives, edited by R. Gros, 80–​84. New York: Springer.
Bruno, M. C., M. Pinto, and W. Rojas. 2018. Identifying domesticated and wild kañawa
(Chenopodium pallidicaule) in the archeobotanical record of the Lake Titicaca Basin
of the Andes. Economic Botany 72(2):137–​149.
Brust, J., W. Claupein, and R. Gerhards. 2014. Growth and weed suppression ability of
common and new cover crops in Germany. Crop Protection 63: 1–​8.
Caligari, P. D. S., P. Römer, M. A. Rahim, C. Huyghe, J. Neves-​Martins, and E. J. Sawicka-​
Sienkiewicz. 2000. The potential of Lupinus mutabilis as a crop. In Linking Research
and Marketing Opportunities for Pulses in the 21st Century, 569–​573. Dordrecht:
Springer.
References 153

Carrillo, W., M. Cardenas, C. Carpio, D. Morales, M. Álvarez, and M. Silva. 2018. Content of
nutrients component and fatty acids in chia seeds (Salvia hispanica L.) cultivated in
Ecuador. Asian Journal of Pharmaceutical and Clinical Research 11(2): 1–​4.
Carvajal-​Larenas, F. E. 2019. Nutritional, rheological and sensory evaluation of Lupinus
mutabilis food products –​a review. Czech Journal of Food Sciences 37(5): 301–​311.
Carvajal-​Larenas, F. E., A. R. Linnemann, M. J. R. Nout, M. Koziol, and M. A. A. J. van Boekel.
2016. Lupinus mutabilis: Composition, uses, toxicology, and debittering. Critical
Reviews in Food Science and Nutrition 56: 1454–​1487.
Castañeda, M. 1988. Estudio comparativo de diez variedades de tarwi (Lupinus mutabilis
Sweet.) conducidas en dos ambientes de la Sierra norte y centro del Perú. Tesis
Ingeniero Agrónomo). Universidad Nacional Agraria La Molina, Lima.
Cater, N. B., and M. A. Denke. 2001. Behenic acid is a cholesterol-​raising saturated fatty
acid in humans. The American Journal of Clinical Nutrition 73(1): 41–​44.
Coelho, M. A., and M. de las M. Salas-​Mellado. 2014. Chemical characterization of chia
(Salvia hispanica L.) for use in food products. Journal of Food and Nutrition Research
2(5): 263–​269.
Córdova-​Ramos, J. S., P. Glorio-​Paulet, F. Camarena, A. Brandolini, and A. Hidalgo. 2020.
Andean lupin (Lupinus mutabilis Sweet): Processing effects on chemical compos-
ition, heat damage, and in vitro protein digestibility. Cereal Chemistry 97: 827–​835.
Chandrasekaran, U., and A. Liu. 2015. Stage-​specific metabolization of triacylglycerols
during seed germination of Sacha Inchi (Plukenetia volubilis L.). Journal of the Science
of Food and Agriculture 95: 1764–​1766.
Chirinos-​Arias, M. C. 2015. Andean Lupin (Lupinus mutabilis Sweet) a plant with nutra-
ceutical and medicinal potential. Revista Bio Ciencias 3(3): 163–​172.
Chirinos, R., G. Zuloeta, R. Pedreschi, E. Mignolet, Y. Larondelle, and D. Campos. 2013.
Sacha inchi (Plukenetia volubilis): A seed source of polyunsaturated fatty acids,
tocopherols, phytosterols, phenolic compounds and antioxidant capacity. Food
Chemistry 141: 1732–​1739.
Chirinos, R., E. Cerna, R. Pedreschi, M. Calsin, A. Aguilar-​Galvez, and D. Campos. 2021.
Multifunctional in vitro bioactive properties: Antioxidant, antidiabetic and
antihypertensive of protein hydrolyzates from tarwi (Lupinus mutabilis Sweet)
obtained by enzymatic biotransformation. Cereal Chemistry 98: 423–​433.
Da Silva Marineli, R. É. A. Moraes, S. A. Lenquiste, A. T. Godoy, M. N. Eberlin, and M. R.
Maróstica Jr. 2014. Chemical characterization and antioxidant potential of Chilean
chia seeds and oil (Salvia hispanica L.). LWT Food Science and Technology 59: 1304–​10.
De la Cruz Torres, E., C. Mapes Sánchez, A. Laguna Cerda, J. M. García Andrade, A. López
Monroy, J. González Jiménez, and T. Falcón Bárcenas. 2008. Ancient underutilised
pseudocereals –​potential alternatives for nutrition and income generation. In New
Crops and Uses: Their Role in a Rapidly Changing World, edited by J. Smartt and N.
Haq, 186–​203. Chichester, UK: RPM Print & Design.
de la Rosa-​Millán, J., E. Heredia-​Olea, E. Perez-​Carrillo, D. Guajardo-​Flores, S. Román, O.
Serna-​Saldívar. 2019. Effect of decortication, germination and extrusion on physico-
chemical and in vitro protein and starch digestion characteristics of black beans
(Phaseolus vulgaris L.). LWT -​Food Science and Technology 102: 330–​337.
Du, Shuang-​Kui, Hongxin Jiang, Yongfeng Ai, Jay-​Lin Jane. 2014. Physicochemical proper-
ties and digestibility of common bean (Phaseolus vulgaris L.) starches. Carbohydrate
Polymers 108(8): 200–​205.
154 Structure and Composition of Latin-American Crops

FAO (Food and Agriculture Organization of the United Nations). 2010. Fats and Fatty Acids
in Human Nutrition, Report of an Expert Consultation. FAO Food and Nutrition Paper
91 (Final report).Rome: FAO. www.fao.org/​3/​a-​i19​53e.pdf.
Follegatti-​Romero, L. A., C. A. Piantino, R. Grimaldi, and A. C. Fernando. 2009. Supercritical
CO2 extraction of omega-​3 rich oil from Sacha inchi (Plukenetia volubilis L.) seeds.
Journal of Supercritical Fluids 49: 323–​329.
García-​Mantrana, I., V. Monedero, and M. Haros. 2014. Application of phytases from
bifidobacteria in the development of cereal-​ based products with amaranth.
European Food Research and Technology 238: 853–​-​862.
García-​Salcedo, A. J., O. L. Torres-​Vargas, amd H. Ariza-​Calderón. 2018. Agroindustria y
Ciencia de los Alimentos/​Agroindustry and Food Science. Physical-​chemical charac-
terization of quinoa (Chenopodium quinoa Willd.), amaranth (Amaranthus caudatus
L.), and chia (Salvia hispanica L.) flours and sedes. Acta Agronomica 67(2): 215–​222.
Gayral, M., B. Bakan, M. Dalgalarrondo, K. Elmorjani, C. Delluc, S. Brunet, L. Linossier, M.
H. Morel, and D. Marion. 2015. Lipid partitioning in maize (Zea mays L.) endosperm
highlights relationships among starch lipids, amylose, and vitreousness. Journal of
Agriculture and Food Chemistry 63: 3551−3558.
Gepts, P., and D. Dpbouk. 1991. Origin, domestication, and evolution of the common bean
(Phaseolus vulgaris L.). In Common Beans: Research for Crop Improvement, edited
by A. Van Schoonhoven and O. Voyset, 7–​53. Wallingford: CAB International.
González, J. A., Y. Konishi, M. Bruno, M. Valoy, and F. E. Prado. 2012. Interrelationships
among seed yield, total protein and amino acid composition of ten quinoa
(Chenopodium quinoa) cultivars from two different agroecological regions. Journal
of the Science of Food and Agriculture 92: 1222–​1229.
Gross, R. 1982. El cultivo y utilización del tarwi, L. mutabilis. Rome: FAO.
Guillén, M. D., A. Ruiz, N. Cabo, R. Chirinos, and G. Pascual. 2003. Characterization of Sacha
Inchi (Plukenetia volubilis L.) oil by FTIR spectroscopy and H-​1 NMR. Comparison
with linseed oil. Journal of the American Oil Chemists Society 80: 755–​762.
Gutiérrez, L.-​P., L.-​M. Rosada, and A. Jiménez. 2011. Chemical composition of Sacha Inchi
(Plukenetia volubilis L.) seeds and characteristics of their lipid fraction. Grasas y
Aceites 62(1): 76–​83.
Hamaker, B. R., C. Valles, R. Gilman, R. M. Hardmeier, D. Clark, H. H. García, A. E. Gonzáles,
I. Kohlstad, M. Castro, R. Valdivia, T. Rodriguez, and M. Lescano. 1992. Amino acid
and fatty acid profiles of the Inca peanut (Plukenetia volubilis L.). Cereal Chemistry
69: 461–​463.
Hernández-​Gómez, J. A., S. M. Colín, and A. Penae Lomelí. 2008. Natural outcrossing of chia
(Salvia hispanica L.). Revista. Chapingo Serie Horticultura 14: 331–​337.
Hernández-​Pérez, T., M. E. Valverde, D. Orona-​Tamayo, O. Paredes-​López. 2020. Chia (Salvia
hispanica): Nutraceutical properties and therapeutic applications. Proceedings 53.
doi:10.3390/​proceedings2020053017.
Hernández-​Pérez, T., M. E. Valverde, and O. Paredes-​López. 2021. Seeds from ancient food
crops with the potential for antiobesity promotion. Critical Reviews in Food Science
and Nutrition. https://​doi.org/​10.1080/​10408​398.2021.1877​107.
Huamaní, F., M. Tapia, R. Portales, V. Doroteo, C. Ruiz, and R. Rojas. 2020. Proximate ana-
lysis, phenolics, betalains, and antioxidant activities of three ecotypes of kañiwa
(Chenopodium pallidicaule aellen) from Peru. http://​pha​rmac​olog​yonl​ine.silae.it.
ISSN: 1827-​8620, pp. 229–​236.
References 155

Iglesias-​Puig, E., V. Monedero, and M. Haros. 2015. Bread with whole quinoa flour and
bifidobacterial phytases improve contribution to dietary mineral intake and their
bioavailability without substantial loss of bread quality. LWT-​Food Science and
Technology 60: 71–​77.
Ixtaína, V. Y., S. M. Nolasco, and M. C. Tomás. 2008. Physical properties of chia (Salvia
hispanica L.) seeds. Industrial Crops and Products 28: 286–​293.
Ixtaína, V. Y., M. L. Martínez, V. Spotorno, C. M. Mateo, D. M. Maestri, B. W. K. Diehl, S.
M. Nolasco, and M. C. Tomás. 2011. Characterization of Chia seed oils obtained
by pressing and solvent extraction. Journal of Food Composition and Analysis 24:
166–​74.
Jacobsen, E., and A. Mujica. 2006. El tarwi (Lupinus mutabilis Sweet) y sus parientes
silvestres. In Botánica Económica de los Andes Centrales, 458–​482. La Paz: Universidad
Mayor de San Andrés.
Jamshidi, A. M., A. Ahmadi, R. Bochicchio, and R. Rossi. 2019. Chia (Salvia hispanica L.) as a
novel forage and feed source: A review. Italian Journal of Agronomy 14: 1–​18.
Joshi, D. C., S. Sood, R. Hosahatti, L. Kant, A. Pattanayak, A. Kumar, D. Yadav, and M. G.
Stetter. 2018. From zero to hero: The past, present and future of grain amaranth
breeding. Theoretical and Applied Genetics 131: 1807–​1823.
Knez Hrnčič, M. K., M. Ivanovski, D. Cör, and Ž. Knez. 2020. Chia seeds (Salvia hispanica
L.): An overview—​ phytochemical profile, isolation methods, and application.
Molecules 25(1): 11.
Kodahl, N. 2020. Sacha inchi (Plukenetia volubilis L.)—​from lost crop of the Incas to part of
the solution to global challenges? Planta 251: 80.
Kodahl, N., and M. Sørensen. 2021. Sacha inchi (Plukenetia volubilis L.) is an underutilized
crop with a great potential. Agronomy 11: 1066.
Koziol, M. J. 1992. Chemical composition and nutritional evaluation of quinoa
(Chenopodium quinoa Willd.). Journal of Food Composition and Analysis 5: 35–​68.
La Rosa, R., E. Anaya, Z. Flores, M. Bejarano, L. Brito, and E. Pérez. 2016. Germinación de
Chenopodium pallidicaule aelle “kañiwa” bajo diferentes condiciones de salinidad y
temperatura. The Biologist (Lima) 14(1): 5–​10.
Lim, T. K. 2012. Edible medicinal and non-​medicinal plants. Fruits 2: 815–​848.
Lopez, M. G., L. A. Bello-​Perez, and O. Paredes-​López. 1994. Amaranth carbohydrates. In
Amaranth Biology, Chemistry, and Technology, edited by O. Paredes-​López, 107–​131.
Boca Raton, FL: CRC Press.
Luna-​ Mercado, G. I., and R. Repo-​ Carrasco-​ Valencia. 2021. Gluten-​ free bread
applications: Thermo-​mechanical and techno-​functional characterization of kañiwa
flour. Cereal Chemistry 98: 474–​481.
Mangelson, H., D. E. Jarvis, P. Mollinedo, O. M. Rollano-​Penaloza, V. D. Palma-​Encinas, R.
L. Gomez-​Pando, E. N. Jellen, and P. J. Maughan. 2019. The genome of Chenopodium
pallidicaule: An emerging Andean super grain. Applications in Plant Sciences
7(11): e11300.
Maphosa, Y., and V. A. Jideani. 2017. The role of legumes in human nutrition. In Functional
Food –​ Improve Health Through Adequate Food, edited by Hueda M. Chavarri, 103–​
122. Zagreb: IntechOpen Online ISBN 978-​953-​51-​3439-​8.
Maurer, N. E., B. Hatta-​Sakoda, G. Pascual-​Chagman, and L. E. Rodríguez-​Saona. 2012.
Characterization and authentication of a novel vegetable source of omega-​3 fatty
acids, sacha inchi (Plukenetia volubilis L.) oil. Food Chemistry 134: 1173–​1180.
156 Structure and Composition of Latin-American Crops

Miano, A. C., J. A. García, and P. E. Duarte Augusto. 2015. Correlation between morphology,
hydration kinetics and mathematical models on Andean lupin (Lupinus mutabilis
Sweet) grains. LWT -​Food Science and Technology 61: 290–​298.
Mohamed, A. A., and P. Rayas-​Duarte. 1995. Composition of Lupinus albus. Cereal Chemistry
72: 643–​647.
Mohsenin, N. N. 1986. Physical Properties of Plant and Animal Materials (2nd ed.). New York,
NY: Gordon and Breach Science.
Moreno, M. L, I. Comino I, and C. Sousa 2014. Alternative grains as potential raw material
for gluten-​free food development in the diet of celiac and gluten-​sensitive patients.
Austin Journal of Nutrition and Food Sciences 2(3): 9.
Muñoz, L. A., N. Vera, M. C. Zúñiga, M. Moncada, and C. M. Haros. 2021. Physicochemical
and functional properties of soluble fiber extracted from two phenotypes of
chia (Salvia hispanica L.) seeds. Journal of Food Composition and Analysis 104:
104138.
Neves Martins, J. M., P. Talhinhas, and P. R. de Sousa, 2016. Yield and seed chem-
ical composition of Lupinus mutabilis in Portugal. Revista de Ciências Agrárias
39(4): 518–​525.
Nowak, V., J. Du, and U. Ruth. 2016. Charrondière. Assessment of the nutritional compos-
ition of quinoa (Chenopodium quinoa Willd.). Food Chemistry 193: 47–​54.
Ogrodowska, D., R. Zadernowski, M. Tanska, and S. Czaplicki. 2011. Physical properties of
Amaranthus cruentus seeds from different cultivation regions in Poland. Zywnosc-​
Nauka Technologia Jakosc 18: 91–​104.
Orona-​Tamayo D., O. Paredes-​López. 2017. Amaranth Part 1 -​Sustainable crop for the
21st century: Food properties and nutraceuticals for improving human health.
In Sustainable Protein Sources, edited by S. R Nadathur, J. P. D Wanasundara, and
L. Scanlin, 239–​256. Cambridge, MA: Academic Press.
Orona-​Tamayo, D., M. E. Valverde, and O. Paredes-​López. 2019. Bioactive peptides from
selected Latin American food crops –​A nutraceutical and molecular approach.
Critical Reviews in Food Science and Nutrition, 59(12): 1949–​1975.
Ortiz, A. 2012. Los Maíces en la Seguridad Alimentaria de Bolivia. La Paz: Centro de
Investigación y Promoción del Campesinado (CIPCA).
Paliwal, R. L., G. Granados, H. R. Lafitte, and A. D. Violic. 2000. El maíz en los
trópicos: Mejoramiento y producción. www.fao.org/​d ocrep/​003/​X7650S/​x7650s02.
htm.
Palma-​Rojas, C., C. González, B. Carrasco, H. Silva, and H. Silva-​Robledo. 2017. Genetic,
cytological and molecular characterization of chia (Salvia hispanica L.) provenances.
Biochemical Systematics and Ecology 73 : 16–​21.
Paredes López, O., C. Ordorica-​ Falomir, and F. Guevara-​ Lara. 1984. Las proteínas
vegetales: presente y futuro en la alimentación. In Prospectiva de la Biotenología en
México, edited by R. Quintero Ramírez, 331–​350. Mexico: Conacyt.
Peláez, P., D. Orona Tamayo, S. Montes Hernández, M. Valverde, O. Paredes López, and A.
Cibrián Jaramillo. 2019. Comparative transcriptome analysis of cultivated and wild
seeds of Salvia hispanica (chia). Scientific Reports 9: 9761–​9767.
Peñarrieta, J. M., J. A. Alvarado, B. Åkesson, and B. Bergenståhl. 2008. Total antioxidant
capacity and content of flavonoids and other phenolic compounds in canihua
(Chenopodium pallidicaule): An Andean pseudocereal (en línea). Molecular Nutrition
and Food Research 52: 708–​717.
References 157

Prasanna, B. M., N. Palacios-​Rojas, F. Hossain, V. Muthusamy, A. Menkir, T. Dhliwayo, T.


Ndhlela, F. S. Vicente, S. K. Nair, B. S. Vivek, X. Zhang, M. Olsen, and X. Fan. 2020.
Molecular breeding for nutritionally enriched maize: Status and prospects. Frontiers
in Genetics 10: 1392.
Priya, T. R. S., and A. Manickavasagan. 2020. Common bean. In Pulses Processing and
Product Development, edited by A. Manickavasagan and T. Praveena, 77–​ 98.
New York: SpringerLink. ISBN 978-​3-​030-​41375-​0 ISBN 978-​3-​030-​41376-​7 (eBook)
https://​doi.org/​10.1007/​978-​3-​030-​41376-​7.
Ranhotra, G. S., J. A. Gelroth, B. K. Glaser, K. J. Lorenz, and D. L. Johnson. 1993. Composition
and protein nutritional quality of quinoa. Cereal Chemistry 70: 303–​305.
Rascon-​Cruz, Q., Y. Bohorova, J. Osuna-​Castro, and O. Paredes-​López. 2004. Accumulation,
assembly and digestibility of amarantin expressed in transgenic tropical maize.
Theoretical and Applied Genetics 108(2): 335–​342.
Reguera, M., and C. M. Haros. 2017. Structure and composition of kernels. In Pseudocereals.
Chemistry and Technology (1st ed.), edited by C. M. Haros and R. Schoenlechner, 28–​
48. Oxford: Wiley-​Blackwell.
Repo-​Carrasco-​Valencia, R. 2020. Nutritional value and bioactive compounds in Andean
ancient grains. Proceedings 53(1): 1.
Repo-​Carrasco-​Valencia, R., A. Acevedo de la Cruz, J. C. Icochea Álvarez, and H. Kallio. 2009.
Chemical and functional characterization of kañiwa (Chenopodium pallidicaule)
grain, extrudate and bran. Plant Food for Human Nutrition 64: 94–​101.
Repo-​Carrasco-​Valencia, R., J. K. Hellström, J. M. Pihlava, and P. H. Mattila. 2010. Flavonoids
and other phenolic compounds in Andean indigenous grains: Quinoa (Chenopodium
quinoa), kañiwa (Chenopodium pallidicaule) and kiwicha (Amaranthus caudatus).
Food Chemistry 120: 128–​133.
Rodrígues, H. S., A. Borém, M. S. F. Valente, M. T. G. Lopes, C. D. Cruz, F. C. M. Chaves, and
C. S. Bezerra. 2018. Genetic diversity among accessions of sacha inchi (Plukenetia
volubilis) by phenotypic characteristics analysis. Acta Amazonica 48: 93–​97.
Ruales, J., and B. M. Nair. 1993. Saponins, phytic acid, tannins and protease inhibitors in
quinoa (Chenopodium quimoa, Willd) seeds. Food Chemistry 48: 137–​143.
Ruiz, C., C. Díaz, J. Anaya, and R. Rojas. 2013. Análisis proximal, antinutrientes, perfil de
ácidos grasos y de aminoácidos de semillas y tortas de 2 especies de sacha inchi
(Plukenetia volubilis y Plukenetia huayllabambana). Revista de la Sociedad Química
del Perú 79(1): 29–​36.
Sacilik, K., R. Öztürk, and R. Keskin. 2003. Some physical properties of hemp seed.
Biosystems Engineering 86(2): 191–​198.
Salas, L., D. Tapia, and F. Menegalli. 2015. Biofilms based on canihua flour (Chenopodium
pallidicaule): Design and characterization. Quimica Nova 38(1): 14–​21.
Saleem, Z. M., S. Ahmed, and M. Mohtasheemul Hasan. 2016. Phaseolus vulgaris
Linn.: Botany, medicinal uses, phytochemistry and pharmacology. World Journal of
Pharmaceutical Research 5(11): 1611–​1616.
Salhuana, W. R. 2004. Evaluación de los recursos genéticos de maíz. In Cincuenta años del
Programa Cooperativo de Investigaciones en Maíz (PCIM), 252–​253. Lima: Programa
Cooperativo de Investigaciones en Maíz.
Salvador-​ Reyes, R., M. T. Pedrosa, and S. Clerici. 2020. Review Peruvian Andean
maize: General characteristics, nutritional properties, bioactive compounds, and
culinary uses. Food Research International 130: 108934.
158 Structure and Composition of Latin-American Crops

Salvador-​Reyes, R., A. P. Rebellato, J. A. Lima Pallone, R. A. Ferrari, M. T. Pedrosa, and S.


Clerici. 2021. Kernel characterization and starch morphology in five varieties of
Peruvian Andean maize. Food Research International 140: 110044.
Sandoval-​Oliveros, M. R., and O. Paredes-​López. 2013. Isolation and characterization
of proteins from chia seeds (Salvia hispanica L.). Journal of Agricultural and Food
Chemistry 1: 193−201.
Sanz-​Penella, J. M., M. Wronkowska, M. Soral-​Śmietana, and M. Haros. 2013. Effect of
whole amaranth flour on bread properties and nutritive value. LWT-​Food Science
and Technology 50: 679–​685.
Sathe, S. K., B. R. Hamaker, K. W. Sze-​Tao, and M. Venkatachalam. 2002. Isolation, puri-
fication, and biochemical characterization of a novel water soluble protein from
Inca peanut (Plukenetia volubilis L.). Journal of Agricultural and Food Chemistry
50: 4906–​4908.
Saturni, L., G. Ferretti, and T. Bacchetti. 2010. The gluten-​free diet: safety and nutritional
quality. Nutrients 2: 16–​34.
Saunders, R. M., and R. Becker. 1984. Amaranthus: A potential food and feed resource.
Advance Journal of Food Science and Technology 6: 357–​396.
Schöeneberger, H., R. Gross, H. D. Cremer, and I. Elmadfa. 1982. Composition and protein
quality of Lupinus mutabilis. Journal of Nutrition 112: 70–​76.
Schöenlechner, R., S. Siebenhandl, and E. Berghofer. 2008. Pseudocereals. In Gluten-​Free
Cereal Products and Beverages, edited by E. K. Arendt, and F. Dal Bello, 149–​190.
Oxford: Elsevier.
Segura-​Nieto, M., A. P. Barba de la Rosa, and O. Paredes-​López. 1994. Biochemistry of
amaranth proteins. In Amaranth Biology, Chemistry, and Technology, edited by O.
Paredes-​López, 75–​106. London: CRC Press.
Serratos, J. A. 2012. El origen y la diversidad del maíz en el continente americano.
Mexico: Universidad Nacional Autónoma de México, 4–​12. http://​m.gre​enpe​ace.
org/​mex​ico/​Glo​bal/​mex​ico/​rep​ort/​2012/​9/​GPORI​GENM​AIZ %20final%20web.pdf.
Shafaei, S. M., and S. Kamgar. 2017. A comprehensive investigation on static and dynamic
friction coefficients of wheat grain with the adoption of statistical analysis. Journal
of Advanced Research 8: 351–​361.
Souci, S. W., W. Fachmann, and H. Kraut. 2000. Food Composition and Nutrition Tables.
Stuttgart: Wissenschaft Verlags.
Steffolani, M. E., A. E. León, and G. T. Pérez. 2013. Study of the physicochemical and
functional characterization of quinoa and kañiwa starches. Starch/​Staerke 65(11–​
12): 976–​983.
Stone, L. A., and K. Lorenz. 1984. The starch of amaranth –​Physicochemical properties and
functional characteristics. Starch/​Stärke 36: 232–​237.
Suleiman, R., K. Xie, and K. A. Rosentrater. 2019. Physical and thermal properties of
chia, kañiwa, triticale, and farro seeds as a function of moisture content. Applied
Engineering in Agriculture 35(3): 417–​429.
Takeyama, E., and M. Fukushima. 2013. Physicochemical properties of Plukenetia volubilis
L. seeds and oxidative stability of cold pressed oil (green nut oil). Food Science
Technology Research 19(5): 875–​882.
Tapia, M. 2015. El tarwi, lupino andino. Fondo Ítalo Peruano 15–​16. https://​docpla​yer.es/​
33372​914-​El-​tarwi-​lup​ino-​and​ino.html.
References 159

Trugo, L. C., D. von Baer, and E. von Baer. 2003. Lupin. In Encyclopedia of Food Sciences and
Nutrition (2nd ed.) edited by B. Caballero, 3623–​3629. Oxford: Academic Press.
Ullah, R., M. Nadeem, A. Khalique, M. Imran, S. Mehmood, A. Javid, and J. Hussain. 2016.
Nutritional and therapeutic perspectives of Chia (Salvia hispanica L.): A review.
Journal of Food Science and Technology 53(4): 1750–​1758.
UNALM/​MINAGRI (Universidad Nacional Agraria de la Molina/​Ministerio de Agricultura
del Perú). 2014. Mapa de razas de maíz en el Perú. www.minam.gob.pe/​dive​rsid​adbi​
olog​ica/​wp-​cont​ent/​uplo​ads/​sites/​21/​2014/​02/​razasm​aizp​eru.pdf.
USDA (US Department of Agriculture), Agricultural Research Service. 2010. USDA National
Nutrient Database for Standard Reference, Release 23. Nutrient Data Laboratory
Home Page. www.ars.usda.gov/​ba/​bhnrc/​ndl.
USDA (US Department of Agriculture), Agricultural Research Service. 2011. National
Nutrient Database for Standard Reference, Release 24. Nutrient Data Laboratory
Home Page. Washington, DC: US Department of Agriculture, Agricultural Research
Service. www.ars.usda.gov/​ba/​bhnrc/​ndl.
Valcárcel-​Yamani, B., S. Caetano, and S. Lannes. 2012. Applications of quinoa (Chenopodium
quinoa willd.) and amaranth (Amaranth spp.) and their influence in the nutritional
value of cereal based foods. Food and Public Health 2(6): 265–​275.
Valencia-​Chamorro, S. A. 2003. Quinoa. In Encyclopedia of Food Science and Nutrition,
edited by B. Caballero, 4895–​4902. Oxford: Academic Press.
Vanier, N. L., C. D. Ferreira, I. da Silva Lindemann, J. Pozzada Santos, P. Zaczuk Bassinello,
and M. E. Cardoso. 2019. Physicochemical and technological properties of common
bean cultivars (Phaseolus vulgaris L.) grown in Brazil and their starch characteristics.
Revista Brasileira de Ciências Agrárias 14(3): 56–​75. ISSN 1981-​0997.
Vega-​Gálvez, A., M. Miranda, J. Vergara, E. Uribe, L. Puente, and E. A. Martínez. 2010.
Nutrition facts and functional potential of quinoa (Chenopodium quinoa willd.),
and ancient Andean grain: A review. Journal of the Science of Food and Agriculture
90: 2541–​2547.
Vilche, C., M. Gely, and E. Santalla. 2003. Physical properties of quinoa seeds. Biosystems
Engineering 86: 59–​65.
Villacrés, E., C. Yánez, L. Armijos, M. B. Quelal, and M. Álvarez. 2016. El Despertar
Gastronómico del Maíz: Recetario; Estación Experimental Santa Catalina. Quito,
Ecuador.
Vrancheva, R., L. Krystev, A. Popova, and D. Mihaylova. 2019. Proximate nutritional com-
position and heat-​induced changes of starch in selected grains and seeds. Emirates
Journal of Food and Agriculture 31(9): 718–​724.
Wang, B., and J. Wang. 2019. Mechanical properties of maize kernel horny endosperm,
floury endosperm and germ. International Journal of Food Properties 22(1): 863–​877.
Wang, S., F. Zhu, and Y. Kakuda. 2018. Review Sacha inchi (Plukenetia volubilis L.): Nutritional
composition, biological activity, and uses. Food Chemistry 265: 316–​328.
Yánez, C., J. L. Zambrano, C. Maicedo, V. H. Sánchez, and J. Heredia. 2003. Catálogo de
Recursos Genéticos de Maíces de Altura Ecuatorianos; INIAP, Estación Experimental
Santa Catalina. Quito, Ecuador. https://​repo​sito​rio.iniap.gob.ec/​jspui/​bitstr​eam/​
41000/​43/​1/​ini​apsc​201.pdf.
Zambrano, J. L., C. F. Yánez, and C. A. Sangoquiza. 2021. Maize breeding in the highlands of
Ecuador, Peru, and Bolivia: A review. Agronomy 11: 212–​221.
Chapter 5

Latin-​American Crops in
Gluten-​Free Applications
Silvia V. Melgarejo-​Cabello1, Jehannara Calle-​Domínguez2,
María Alejandra Giménez3, Claudia M. Haros4, and
Ritva Ann-​Mari Repo-​Carrasco-​Valencia1
1Universidad Nacional Agraria La Molina, Peru
2Instituto de Investigaciones para la Industria

Alimenticia (IIIA), La Habana, Cuba


3Universidad Nacional de Jujuy and Consejo

Nacional de Investigaciones Científicas y Técnicas


(CONICET), San Salvador de Jujuy, Argentina
4Universidad Nacional Agraria La Molina, Peru

CONTENTS
5.1 Introduction 162
5.2 Andean Grains 162
5.2.1 Processing of Quinoa 163
5.2.2 Processing of Kañiwa 165
5.2.3 Processing of Kiwicha 165
5.2.4 Processing of Tarwi 167
5.2.5 Gluten-​Free Products 167
5.3 Maize 170
5.3.1 Processing of Maize 170
5.3.1.1 Nixtamalization 170
5.3.1.2 Milling 172
5.3.1.3 Extrusion 174
5.3.1.4 Toasting 178
5.3.2 Gluten-​Free Products 178

DOI: 10.1201/9781003088424-5 161


162 Latin-American Crops in Gluten-Free Applications

5.4 Sacha Inchi and Black Turtle Bean 180


5.4.1 Methods of Processing and Industrial Applications 181
5.4.2 Baked Goods Application 183
5.4.3 Gluten-​Free Products 183
5.5 Chia 185
5.5.1 Processing Chia Seeds 188
5.5.2 Gluten-​Free Foods with Chia 189
5.5.2.1 Bread Products 189
5.5.2.2 Biscuits, Cookies, Cakes and Snacks 190
5.5.2.3 Pasta Products 191
5.6 Conclusions 191
Acknowledgments 192
References 192

5.1 INTRODUCTION
Latin-​American crops, such as quinoa, amaranth, kañiwa, tarwi, maize,
chia, sacha inchi and black turtle beans have an excellent nutritional value
and are widely cultivated and used in the region. Quinoa and tarwi contain
antinutrients (saponins and alkaloids, respectively) which must be removed
before consumption. Among these crops, kañiwa is the least known and
has enormous potential as a very nutritious ingredient in the food industry.
Amaranth is an ancient food crop traditionally consumed by people from
Latin America. Maize is cultivated and used all over Latin America and has
several traditional and modern food applications. Sacha inchi is a crop with
an exceptionally high oil content with a very healthy fatty acid composition.
Due to the physical–​chemical and nutraceutical properties of chia seeds, they
have interesting technological properties and potential uses for different
food applications. Whole seed flours can be obtained from all of these crops,
and they are promising ingredients for different kinds of products, including
gluten-​free applications.

5.2 ANDEAN GRAINS
The four native Andean grains, quinoa (Chenopodium quinoa), kañiwa
(Chenopodium pallidicaule), kiwicha (Amaranthus caudatus) and tarwi (Lupinus
mutabilis) have an exceptional nutritional composition. They can be used as
ingredients in conventional and gluten-​free bakery products to improve the
5.2 Andean Grains 163

nutritional value of these products. They have interesting techno-​functional


characteristics which could enhance the sensorial properties of the final
products.

5.2.1 Processing of Quinoa
Quinoa contains macro-​and micronutrients (starch, protein, fiber, fat, minerals
and vitamins), which are found in different parts of the seed. The quinoa germ or
embryo, represents approximately 30% of the whole seed (in the case of common
cereals it represents only 5% of the seed and extends through the starchy endo-
sperm), surrounds the perisperm like a ring covering the seed. It is part of the
bran fraction, which is relatively rich in fats and proteins (Alonso-​Miravalles
and O’Mahony, 2018; García Solaesa et al., 2020; Mufari et al., 2018). The starch
granules occupy the cells of the perisperm, which constitutes 60% of the grain,
while most of the lipids and proteins are found in the germ (Ando et al., 2002;
Prego et al., 1998).
Saponins, located in the outer layers of quinoa, generate a bitter taste and
have anti-​nutritional effects. The removal of saponins is easily accomplished by
rinsing the seed in cold alkaline water (wet method) or mechanical abrasion
(dry method). The wet method is generally used, generating a large amount of
water waste and causing pollution in rivers and lakes. Abrasive de-​hulling is an
alternative method to eliminate the saponins. Han et al. (2019) investigated the
effect of degree of milling (DOM) on the content of saponins and free and bound
phenolics, and their antioxidant activity The saponin content decreased from
15.50 to 9.02 mg oleanolic acid equivalents/​g dry weight (OAE/​g DW) after using
a disc mill device. These results indicate that the milling process could become
an effective method for removing the saponins in quinoa.
Ando et al. (2002) tested abrasive grinding for the fractionation of quinoa
and obtained a bran fraction (8.2%) and an embryo fraction (30.1%) that were
significantly different in their composition; in addition, they were successfully
separated from the perisperm (58.8%) and the proportion of protein fraction
was not affected by abrasive grinding. Hemalatha et al. (2016) investigated
sequential abrasive grinding and produced four fractions: husk (10–​12%),
de-​hulled grain, bran (12–​15%) and ground grain (mainly white perisperm).
The most abundant fraction was grain flour at 75–​77% of the total grain
weight. Ruiz et al. (2016) explored combined dry and wet processing and
obtained, through impact grinding, a coarse fraction rich in perisperm and a
fine fraction mainly composed of the embryo, which had a different protein
content.
164 Latin-American Crops in Gluten-Free Applications

D’Amico et al. (2019) investigated abrasive milling with commercial-​quality


seeds; 10 g of quinoa was sequentially ground for 8 min at 1-​min intervals and
sieved to separate it into the following components: de-​hulled kernel, coarse
( fraction), fine ( fraction) and embryo fractions. They reported that in the first
2 min the coarse fraction was mainly separated (shell and bran), in 3 to 5 min
the embryo was broken and after only 5 min the fine fraction was obtained,
which contained parts of the perisperm. The approach of grinding the kernels
by abrasive milling in layers allowed them to obtain very detailed information
about the nutritional composition of the selected seed tissues. The outer layers
of the kernels were rich in minerals and fiber (arabinose and galactose-​rich
polysaccharides). After 2 min of milling, there was no longer any coarse fraction
obtained and removal of the embryo was boosted, which resulted in very high
protein content of the abraded layers. After 8 min of milling, more than 65% of
the original seed mass was removed and the polished kernels were low in pro-
tein, ash and fat, whereas the amount of glucose increased. Although quinoa
has a completely different structure with respect to cereals, some similarities
were found. The high concentrations of ash and dietary fiber were comparable
to those of cereals, whereas the protein composition in the milling course was
more affected in true cereals.
Comparison of fat localization showed a completely diverse picture, because
in cereals the germ is located close to the bran, whereas in quinoa the embryo
surrounds the perisperm. The ash content can be used to estimate the DOM
yield. Thus, classification of quinoa flours can be performed as for cereals. To
know the protein composition of the quinoa grain, extractions were carried out
(based on the solubility of the proteins) with pure water and 2% sodium chloride,
obtaining two fractions, albumins (43.3%) and globulins (27.9%) that corres-
pond to Osbourne fractions. In a third extraction step, 1% sodium dodecyl sul-
fate (SDS) was used instead of 60–​70% ethanol to extract the prolamines that
represented 11.1%; glutelins (insoluble residues) are calculated as the mass diffe-
rence of all soluble protein fractions with respect to total protein content and
accounted for 21.8% (D’Amico et al., 2019).
The high proportion of albumins and globulins (more than 70% of the total
proteins) and the lower abundance of prolamines and glutelins correspond to
a typical dicotyledonous plant. The predominance of albumins was confirmed
from the study by Ruiz et al. (2016), who found 40.3% water-​soluble proteins in
quinoa. These results demonstrate that the protein composition within quinoa
grains is different from that of common cereals. Gluten proteins, composed of
prolamines and glutelins, are predominantly found in the endosperm, while
bran and especially the aleurone layer are rich in albumins and globulins. In
summary, proportions of protein fractions are much more affected by husking
in wheat than in quinoa.
5.2 Andean Grains 165

Ballester-​Sánchez et al. (2019) isolated starch from quinoa using a wet


milling procedure. They investigated the effect of duration (1, 5 and 9 h) and
temperature (30, 40 and 50°C) of maceration in an SO2 solution with lactic acid
on the recovery of starch and its quality. The effect of the steeping conditions
on the starch was evaluated in terms of whiteness, proteins, lipids, amylase and
the content of damaged starch. The highest level of recovery and the best quality
of starch were obtained after 6.5 h of maceration at 30°C.

5.2.2 Processing of Kañiwa
Traditionally, kañiwa seeds are eaten roasted and ground into flour (kañiwaco).
They are used to make soup, breakfast drinks, bread, cookies and cakes.
Compared to quinoa, kañiwa has the advantage of not containing detectable
amounts of saponins, so it does not require washing prior to consumption
(Bustos et al., 2019). The grinding of grains can be carried out in various types
of mills; flour improves the nutritional characteristics of bakery, baked or pasta
products; however, the percentage used is limited by the absence of gluten,
which affects the composition of the starch–​gluten matrix, of utmost import-
ance for textural properties.
Bustos et al. (2019) investigated the production of pasta prepared with wheat
flour enriched with kañiwa wholemeal flour at replacement percentages of
10%, 20% and 30%. Wholemeal flour was prepared in a hammer mill, reaching
granulometry of less than 250 μm. The partial replacement of wheat flour with
kañiwa flour is a viable option for improving the nutritional value of breads,
with acceptable dough yields. Rosell et al. (2009) evaluated the production of
bread with 10% to 100% replacement of wheat flour with kañiwa flour obtained
through a cyclone mill, concluding that up to 25% substitution could produce
doughs with acceptable thermo-​mechanical patterns and breads with good sen-
sory acceptability, but variable in color.
Kañiwa flour is also a source of protein extraction. For example, Betalleluz-​
Pallardel et al. (2017) established optimal conditions for the extraction of kañiwa
protein from flour with a granulometry of less than 500 μm; the optimal protein
extraction conditions correspond to a temperature of 21°C, extraction time of 5
min and solvent (water)/​flour ratio of 37:1 (v/​w) at pH 10, resulting in a protein
yield of 80.4%.

5.2.3 Processing of Kiwicha
Kiwicha is recognized for its excellent nutritional content, mainly for its good
quality proteins. Kiwicha does not have detectable amounts of saponins, so it
does not require pre-​treatment washing, and it is gluten-​free, which makes it an
166 Latin-American Crops in Gluten-Free Applications

interesting ingredient for products aimed at celiac consumers (Álvarez-​Jubete


et al., 2009, 2010a; Calderón de la Barca et al., 2010).
Although numerous studies of milling kiwicha grains have been carried
out at laboratory level, industrial use of these methodologies is still very
complicated. This is due to factors such as poor development of grain
cultivars suitable for mechanical harvesting and the present low yields
associated with grain size, so that studies which allow optimization of these
processes are necessary (Roa et al., 2014). Calzetta et al. (2006) developed a
wet milling methodology with an optimal starch yield of 45.0%. This method-
ology begins with a previous soaking of seeds in an aqueous solution; subse-
quently, the wet materials are ground and then sieved to retain the fiber. The
remaining suspension is centrifuged, the supernatant discarded, the muci-
laginous layer separated and the precipitated starch suspended in distilled
water. Finally, the starch is dried in a forced-​air oven. Roa et al. (2014) used
wet milling methodology, obtaining a starch with 98.9% carbohydrates and
0.6% proteins.
By dry milling, due to the structure and morphology of kiwicha grains, it
is possible to separate the anatomical fractions to obtain flours of different
composition. Tosi et al. (2000) developed a dry differential milling technique to
produce three flour fractions: one rich in protein (40%), another rich in dietary
fiber and a fraction containing approximately 79% starch. They determined that
the grains should be dried before grinding to improve the protein yield in the first
fraction; thus, the optimal conditions for grain drying ( for an initial humidity
of 11.5%) included treating with air at 90°C for 3 min, with a flour extraction
yield of 36.9%; however, as a consequence of the drying effect, the available lysine
decreased by 38.5%.
Roa et al. (2014) used abrasive milling, with a Suzuki MT mill that automatic-
ally separates pearl amaranth and bran, to obtain a starch-​rich grinding fraction
(88%) and a lipid-​protein-​rich fraction with 23% lipid and 46% protein contents,
respectively. Next, they carried out a planetary ball mill grinding. They found
that abrasive milling reduces the degree of crystallinity of the starch due to its
amortization during these processes.
Sindhuja et al. (2005) developed biscuits with amaranth flour, finding that
25% substitution of wheat flour for amaranth flour was optimal. Also, due to
its high starch content, amaranth is a promising material for edible film pro-
duction (Roa et al., 2014). On the other hand, Basilio-​Atencio et al. (2020) used
whole kiwicha flour, obtained in a hammer mill and with a 0.1 mm mesh sieve,
to obtain healthy extruded kiwicha snacks, managing to determine that the
optimal extrusion conditions were at a temperature of 190°C and 14% initial
moisture in the flour.
5.2 Andean Grains 167

5.2.4 Processing of Tarwi
Tarwi seeds are recognized for their exceptionally high protein and oil content;
their consumption is limited due to the high content of anti-​nutrients such as
bitter alkaloids, phytic acid, tannins, nitrates and trypsin inhibitors (Villacrés
et al., 2020). For safe consumption, the seeds must be subjected to a treatment
to considerably reduce the amount of anti-​nutrients. This specific treatment,
­traditionally known as debittering, is therefore a critical step in tarwi processing.
Debittering consists of an aqueous thermal treatment, which is very efficient
in reducing unwanted components of the seed, managing to reduce levels of
alkaloids to average values below 0.2 g/​kg of dry matter. Cortés-​Avendaño et al.
(2020) carried out a study to identify alkaloids after the debittering process in ten
tarwi ecotypes from different regions of Peru. They managed to identify residue
of only two types of alkaloid: lupanin (average 0.0012 g/​100 g of dry matter) and
sparteine (average 0.0014 g/​100 g of dry matter). However, this methodology
favors a reduction in the antioxidant capacity and of some of the main nutrients
of the grain such as proteins, carbohydrates, ascorbic acid, phenolic compounds
and carotenoids (Villacrés et al., 2020).
The debittering process of tarwi seeds requires large amounts of water and
takes a long time. The water used for the elimination of alkaloids is ­traditionally
used as a biocide for pests. Recent studies have been conducted with emer-
ging procedures through biotechnological processes such as fungal fermenta-
tion, which allows a reduction of alkaloids and tannins, with the advantage of
maintaining a higher content of other nutrients such as carotenoids and phen-
olic compounds (Romero-​Espinoza et al., 2020; Villacrés et al., 2020).
Various types of mill can be used for grinding. Cortés-​Avendaño et al. (2020)
used a hammer mill to obtain flours with particle sizes between 100 and 500
μm. Wang et al. (2016) carried out two-​stage milling: first by grinding with a pin
mill and later in an impact mill, obtaining flours with 8 to 10% humidity and an
average particle size of 10 to 100 μm. The particle size obtained from grinding is
important for subsequent processes. They also managed to obtain flour with a
high protein content using an electrostatic separation method, for flours with a
smaller average particle size.
Rosell et al. (2009) used tarwi flour, obtained by milling in a laboratory cyc-
lonic mill, in combination with wheat flour in bread formulations; however,
they verified that breads with up to a minimum content of 12.5% of tarwi flour
presented poor sensory acceptability.

5.2.5 Gluten-​Free Products
The gluten-​free products market has increased throughout Latin America, due
to the rise in patients diagnosed with coeliac disease and consumers seeking
168 Latin-American Crops in Gluten-Free Applications

healthier alternatives. However, these are not necessarily healthier, because the
gluten-​free bakery products on the market are made from starches or white rice
flour, which are lacking in high-​quality proteins, dietary fiber and important
micronutrients. Andean grains such as quinoa, kiwicha, kañiwa and tarwi do
not contain peptides like wheat gluten; therefore, they are appropriate raw
materials for consumption by people with coeliac disease. Quinoa, kañiwa and
kiwicha, the Andean amaranth, are sources of starches (more than 70% of their
composition) (Ramos Diaz et al., 2013; Rosell et al., 2009; Schöenlechner, 2017a),
which are necessary to create the structure of bread (Benavent-​Gil and Rosell,
2018). On the other hand, tarwi is a legume with very high protein and oil con-
tent (almost 50% and around 20%, respectively). Tarwi oil could function as a
natural emulsifier to retain the gas produced during fermentation of gluten-​
free breads (Vidaurre-​Ruiz et al., 2019). Quinoa is appreciated because of its
high protein quality, having a balanced essential amino acid composition, and
it is also considered an excellent source of dietary fiber and minerals (Repo-​
Carrasco-​Valencia et al., 2010). Kiwicha is a very good source of iron, calcium
and zinc. It contains more zinc and iron than conventional corn and common
beans (Burgos et al., 2018). The inclusion of these grains in the formulation of
gluten-​free products is promising.
The quality of wheat bread is the result of special properties of gluten proteins
(gliadins and glutenins). They are responsible for the high water absorption cap-
acity and give dough cohesiveness, viscosity, elasticity and gas-​holding ability
and a final product with a high volume and porous crumb. All other cereal flours
produce breads of poor quality with low volume and a small-​pored, inelastic
crumb when baked under standardized conditions (Koehler et al., 2014). It is
very difficult to replace gluten in bakery products and it requires the employ-
ment of a mix of permitted flours, proteins, hydrocolloids and special tech-
nologies in an attempt to replace the numerous functions of gluten (Arendt
et al., 2008). Generally, basic materials in the production of gluten-​free bread
are starch-​containing flours or starches from gluten-​free sources (e.g., corn,
rice, potatoes). To achieve an appropriate water absorption capacity and dough
viscosity, several hydrocolloids can be used. Hydroxypropylmethyl-​cellulose,
carboxymethylcellulose, carrageenan, xanthan gum, guar gum and sodium
alginate are examples of compounds used in the production of gluten-​free bread
(Koehler et al., 2014).
Dairy substitutes, like soy milk, can be produced from quinoa flour by a hot
mashing process, or by drum drying (Schöenlechner, 2017b). These beverages
provide novel alternatives to current dairy substitute products, having a rela-
tively high protein content and low glycemic index (Pineli et al., 2015).
Quinoa can be used in bakery products such as cookies without the incorp-
oration of wheat flours, despite the lack of gluten (Schöenlechner et al., 2006).
5.2 Andean Grains 169

For biscuit and cookie making, a strong gluten network is not needed, making
pseudo-​cereals suitable for this purpose. Studies by Inglett et al. (2015) showed
that amaranth-​ based cookies have good sensorial and textural attributes.
Popped amaranth flour and wholegrain popped amaranth have been employed,
resulting in cookies with good sensory properties (Calderón de la Barca et al.,
2010). Malted amaranth flour cookies have shown superior properties to those
of conventional wheat cookies (Chauhan et al., 2015).
The baking properties of amaranth, quinoa and buckwheat were studied by
Álvarez-​Jubete et al. (2010b). Their study demonstrated that pseudo-​cereal flour
can be introduced into gluten-​free bread formulation with the aim of enhancing
crumb softness and cohesiveness without adversely affecting the other sensory
properties of the bread. The use of quinoa and lupin flour in the production of
gluten-​free and egg-​free pasta was studied by Linares et al. (2019). They tested
the effect of vegetable proteins and the oxidizing enzyme POx on the quality of
noodles. After these trials, the final recipe containing lupin four, pea protein and
POx had satisfactory gluten-​free noodle quality and possessed a valuable nutri-
tional composition with high protein and dietary fiber content.
Schöenlechner et al. (2010), in a study on the use of amaranth flour in bread
making, found that water content had the greatest effect on bread properties,
whereas variation in fat content did not have any significant influence. They
also detected that the combined addition of fat and egg albumin resulted in the
best overall sensory acceptance. Amaranth flour improves the freeze stability
of gluten-​free dough but reduces its shelf life (Leray et al., 2010). Calderón de la
Barca et al. (2010) produced gluten-​free bread with a uniform crumb and higher
specific volume compared to other gluten-​free breads using 60–​70% popped
and 30–​40% raw amaranth flours. Sourdough fermentation could be a prom-
ising strategy in improving the quality of gluten-​free bread. Jekle et al. (2010)
demonstrated that amaranth sourdough properties could be enhanced by using
different lactic acid bacteria. Coda et al. (2010) used Lactobacillus plantarum
C48 and L. lactis subsp. lactis PU1 in sourdough with the aim of producing
gluten-​free bread using a blend of buckwheat, amaranth, chickpea and quinoa
flours. Textural analysis showed that sourdough fermentation enhanced several
characteristics of the gluten-​free bread, thus approaching the features of wheat
flour bread. Sensory analysis showed that sourdough fermentation allowed the
production of breads with good palatability and overall taste appreciation.
Lupin protein concentrate has been used by Horstmann et al. (2017) and
Ziobro et al. (2013) in gluten-​free baking. Vidaurre-​Ruiz (2020) reported that the
inclusion of flour of the Andean lupin, tarwi, in a gluten-​free bread formulation
with quinoa flour and potato starch gives the final product the following nutri-
tional content: 10.1% protein, 18.6% dietary fiber and 3.3% minerals. Similar
results were obtained using tarwi flour in a kiwicha–​potato starch formulation.
170 Latin-American Crops in Gluten-Free Applications

The protein content of these breads is between two to three times higher than
commercial gluten-​free breads.
Kañiwa is the least known of all Andean native grains. There is scarce infor-
mation regarding the use of this nutritious grain in gluten-​free products. Luna-​
Mercado and Repo-​Carrasco-​Valencia (2020) evaluated the techno-​functional,
thermo-​mechanical and physico-​chemical properties of two kañiwa varieties
(Cupi and Illpa Inia), in order to assess their potential as ingredients in gluten-​
free bakery products. The characterization of these flours of two Peruvian kañiwa
varieties provided new information on the suitability of kañiwa for use in gluten-​
free baking. Kañiwa is an interesting and novel ingredient for use in gluten-​free
products, enhancing their nutritional value.
Commercial products based on quinoa, amaranth, kañiwa and tarwi are
shown in Figures 5.1–​5.4.

5.3 MAIZE
Maize (Zea mays L.) is one of the cereals that celiac patients can consume, if suit-
able conditions are provided during production. Whole maize or its flours are
widely used in the preparation of typical South America foods. Nevertheless, the
absence of prolamins that form a viscoelastic network like gluten represents a
challenge in the elaboration of pasta or baked products. Maize flour is not rich in
essential amino acids like lysine and tryptophan; other gluten-​free raw materials
can be added to nutritionally enhance different gluten-​free products (Figure 5.5).

5.3.1 Processing of Maize
5.3.1.1 Nixtamalization
Alkaline cooking of maize (nixtamalization) is a process used in the preparation
of Mexican-​style foods. During nixtamalization, maize is first cooked in water
in the presence of lime, steeped overnight for several hours, and then washed
to produce nixtamal, which is then stone-​ground to form dough. Alternatively,
nixtamalized maize dough can be dried and dry-​milled to produce maize flour
(Cravioto et al., 1945; Bressani, 1990). In this process, the pericarp is hydrolyzed
mostly into arabinoxylans; the starch of the endosperm undergoes a partial gel-
atinization; some lipids are saponified, and proteins are partially solubilized and
polymerized. Another important change is the calcium absorption including
other minerals from lime (Santiago-​Ramos et al., 2018). The extent of these
changes and interactions among some components produces nixtamalized
maize dough with different viscoelastic behavior (Vazquez-​Carrillo et al., 2015).
The nixtamalization process increases the calcium content, and other minerals,
5.3 Maize 171

A B

C D

Figure 5.1 Commercial gluten free foods with amaranth: A. Puffed amaranth compared
to amaranth seeds; B. Amaranth bars with probiotics; C. Amaranth tortillas; D. “Alegría”,
E. Gluten free “alfajores” with amaranth. Source: Authors’ own work.
Notes: A) Two small round plates next to each other, one of them with amaranth and the
other with a puffier and therefore slightly bigger amaranth.
172 Latin-American Crops in Gluten-Free Applications

Figure 5.1 (continued)


B) Medium round plate with three different flavored popped amaranth (peanuts, choc-
olate and raisins), rectangular prism snacks (in their respective plastic bags) parallel to
each other.
C) Medium round plate with a round amaranth tortilla; the seeds are embedded on
the surface.
D) Five popped amaranth, short cylinder snacks in their respective plastic bags in a
small round bowl.
E) Two alfajores next to each other. The alfajor is formed by two small, round amaranth
cookies one at the top and the other at the bottom with a dairy, sweet ingredient called
“dulce de leche’‘ in the middle.

in maize flours, improves the bioavailability of niacin and iron and decreases
the content of mycotoxins and phytic acid. However, during alkaline cooking,
phenolic compounds and other important antioxidants can be lost (Cravioto
et al., 1945; Bressani et al., 2002, de la Parra et al., 2007, Gutiérrez-​Dorado
et al., 2008).
In nixtamalization, the wet milling of nixtamal is an important operation.
Two cylindrical stones are used to mill the maize limed samples; one of the
stones is fixed, while the other rotates (Serna-​Saldívar and Chuck-Hernández,
2019). The distance between the stones is adjusted to obtain a certain particle
size distribution; at a higher compression force, the dough is soft, and is used
to produce tortillas, but if the compression force decreases, the resulting dough
turns into coarse particles, and is used in different products. Alternatively, steel
plate dough mills are used. These mills are more efficient and produce dough at
a lower temperature (Villada et al., 2017).
Nixtamalized maize flours are generally industrially obtained and have spe-
cific characteristics for the production of a particular food. Some companies
offer more than 25 different nixtamalized flours formulated to meet certain
requirements for color, pH, granulation, water absorption, viscosity and sen-
sorial characteristics (Serna Saldívar and Chuck-​Hernández, 2019). Interestingly,
from a home-​made limed maize in the Mayan and Aztec cultures, before the
conquistadores arrived, it has now reached a wide international market in all
continents of the world (Paredes-​López et al., 2012).

5.3.1.2 Milling
According to Abarza and Schimpf (2009), different products are obtained
from the grinding process of native maize from the Andean zone of Jujuy,
Argentina, such as frangollo (broken maize grains that were peeled with lime),
semolina and flour. The last two are mainly used to make soups (calapi, caldo
majado and tulpo) or to elaborate bread known as bollo mixed with wheat flour
5.3 Maize 173

A B C

D E F

G H I

Figure 5.2 Commercial gluten free foods with quinoa: A. White and red quinoa ready-​
to-​eat; B. Bread with quinoa and millet; C. Crispbread with sunflower seed and quinoa; D.
Mini corn with quinoa cakes; E. Quinoa chips; F. Quinoa cookies; G. Crispy quinoa toast;
H. Quinoa and rice crackbread; I. Whole rice and quinoa cakes; J. Penne, spaghetti or
fusilli quinoa and rice pasta. Source: Authors’ own work.
Notes: A) Cooked, ready-​to-​eat red and white quinoa on a small round plate.
174 Latin-American Crops in Gluten-Free Applications

Figure 5.2 (continued)


B) Five small slices of quinoa and millet bread in a small round bowl.
C) Three rectangular shaped crispbread with sunflower seed, quinoa and other grains
on a medium size round plate.
D) Many circular mini corn crackers with quinoa in a small round bowl.
E) Many quinoa chips in a corrugated rectangular shape in a small round bowl.
F) Five half-​moon shaped quinoa cookies, with a rough surface on a medium sized
plate.
G) Three pieces of thin and rectangular crispy quinoa toast in a fan-​like position in a
cup.
H) Five thin and rectangular quinoa and rice crackerbreads in a fan-​like position on a
medium size plate.
I) Four rectangles with rounded corners, whole rice and quinoa cakes in a fan-​like pos-
ition on a medium size round plate.
J) Rectangular plate positioned horizontally with three types of pasta. Spaghetti is
found horizontally across the middle of the plate, penne on the upper part of the plate
and fusilli on the lower side of the plate.

(Cámara-​Hernández and Arancibia-​Cabezas, 2007). In this region, hydraulic


mills brought by Spaniards were incorporated as a technological change; but
this did not imply a cultural change since grain milling was a practice already
known and used. On the contrary, grinding in stone mortars, domestic and
community, is an important feature of many Andean societies. These mortars –​
pecans, cutanas –​continued to be used together with hydraulic mills, gener-
ally differentiating uses of the flours that were produced in each. Currently,
electric mills for community use have also been incorporated (Bugallo and
Mamani, 2014). Semolina and flours obtained from maize from the Andean
region of Argentina generally come from the whole grinding of the grain that,
depending on culinary use, can be subjected to successive sieves to eliminate
part of the bran.

5.3.1.3 Extrusion
Extrusion technology is considered one of the most economic processes for
cereal and snack food industries. This process can produce dough with similar
characteristics to that by using nixtamalization traditional processes (Corrales-​
Bañuelos et al., 2016). Extrusion is a suitable technology for obtaining gluten-​
free products such as pasta, even using whole grain maize flours. This process
can also be applied to improve the stability of maize whole flour, reaching an
adequate temperature to the inactivation of lipases and lipooxygenases respon-
sible for rancidity, avoiding undesirable sensorial changes and increasing ran-
cidity time (Paesani et al., 2020). According to Kljak et al. (2015), the different
5.3 Maize 175

Figure 5.3 Commercial products of kañiwa: A. Kañiwa and kiwicha bar; B. Kañiwa pop;
C. Kañiwa flakes; D. Kañiwaco. Source: Authors‘ own work.
Notes: A) Five rectangular kañiwa and kiwicha bars on an oval shape plate. Three of them
horizontally in the middle of the plate and two vertically at either side.
B) Kañiwa popped seeds on a round medium size plate.
C) Thin and small kañiwa flakes on a medium size plate.
D) Toasted kañiwa flour on a medium size plate.

genotypes and different physical and chemical characteristics of maize lead to


different nutritional values and different responses in the extrusion process.
In this regard, Merayo et al. (2011) showed that the quality of maize spaghetti
depended on extrusion conditions and endosperm hardness, finding that soft
endosperm allowed for obtaining better quality spaghetti. Application of the
176 Latin-American Crops in Gluten-Free Applications

Figure 5.4 Commercial products of tarwi: A. Quinoa and tarwi cookies; B. Tarwi cheese;
C. Tarwi and lucuma flour; D. Snack of tarwi (lupine) with spices.
Notes: A) Quinoa and tarwi heart-​shaped cookies with a circle in the middle of each
cookie, on a small round plate. Source: Authors’ own work.
B) Molded cylindrical tarwi cheese on a medium size plate.
C) Tarwi and lucuma flour on a large plate.
D) Plate with many white, oval-​shaped tarwi seeds treated with spices.

extrusion process in different Andean maize varieties for spaghetti’s manufac-


ture allowed the obtention of products of adequate quality and sensory
acceptance for celiac consumers (Giménez et al., 2014). The textural properties
of maize-​spaghetti and their behavior during overcooking depend on the degree
of gelatinization reached during the extrusion-​cooking process and the forma-
tion of the amylose-​lipid complex (Merayo et al., 2011, Giménez et al., 2013).
5.3 Maize 177

A B

C D

E F

Figure 5.5 Gluten free foods with Andean maize: A. Roasted yellow corn flavored with
spices; B. Gluten free “alfajores” with capia maize; C. Spaghettis with culli, capia or yellow
maize; D. Pasta made with Andean maize augmented with quinoa, beans or amaranth; E.
Pudding made with Andean maize culli and Andean maize capia; F. Toasted yellow corn
flour in preparation for an Andean drink “Chilcan”. Source: Authors’ own work.
Notes: A) Roasted yellow corn flavored with spices in a squared plastic bag.
B) Two alfajores, formed by two small, round capia maize cookies, one at the top and
the other at the bottom with a dairy, sweet ingredient called “dulce de leche” in the middle
of them.
178 Latin-American Crops in Gluten-Free Applications

Figure 5.5 (continued)


C) Three different types of Andean maize spaghetti, front view (the top one is made
with culli, middle one with capia and the bottom one with yellow maize).
D) Three plastic bags of Andean maize spaghetti positioned vertically, with quinoa,
beans or amaranth.
E) Two large rectangular pudding cakes (one beige culli and the other dark brown
capia), and beside each one there is a slide of them.
F) Toasted yellow corn flour in a rectangular plastic bag.

5.3.1.4 Toasting
Toasting is a poorly studied process of maize products, and its importance relies
on the fact that they are widely consumed by the rural population of Mexico and
South America in different typical preparations. Toasted grains exhibit improved
digestibility, color, shelf life, flavor and texture, greater volume and crispiness
and reduced anti-​nutrient factors (Sandhu et al., 2017). The toasting process is
carried out at dry high temperatures (250–​270°C). The grain structure protects
the starch granule during heat treatment, preventing thermal fractionation and
a large melting degree. Therefore, it may be considered as an alternative heat
treatment to obtain starch with some desirable properties, such as greater crys-
tallinity (Carreras et al., 2015).

5.3.2 Gluten-​Free Products
The health benefits of consuming whole grains and their flours are clear
(Shepherd et al., 2012). However, a larger scale of production requires a thermal
treatment for the destruction of lipolytic enzymes and lipoxygenases in order
to prevent spoilage. On the other hand, the presence of bran is a critical factor
in whole flours due to a negative effect on the quality of some products such as
pastas and bread (Liu et al., 2016).
Despite the nutritional interest in the use of whole grain flours in gluten-​
free products, there are few studies on the use of whole-​grain maize. Paesani
et al. (2020) showed that the presence of whole-​grain maize flours in gluten-​free
sugar-​snap doughs and cookies increased the particle size and water absorp-
tion values while the elastic and viscoelastic modulus of the doughs decreased,
allowing a greater expansion in the oven and a lower hardness. Segundo et al.
(2019) formulated a gluten-​free layer and sponge cakes with whole maize flour
replaced by green banana flour, achieving nutritionally improved products with
acceptable texture and sensory characteristics. According to Carmelo-​Méndez
et al. (2017) and Žilić et al. (2012) whole blue maize flour can be used to develop
5.3 Maize 179

gluten-​free pasta and cookies with high dietary fiber contents, increasing
the total phenolic compound content and antioxidant capacity as well as a
decreased starch hydrolysis due to the increase in resistant starch content. Red
and blue maize flours were used for functional cookie preparation. The effect
of the baking conditions, such as baking time and temperature, affected the
content of phenolic compounds, antioxidant capacity and color of cookies.
According to Žilić et al. (2016), the addition of citric acid to pigmented maize
dough increased the phenolic compounds in maize cookies by anthocyanins
stabilization and improved their accessibility. In addition, the Maillard reac-
tion had a crucial influence on the increment of antioxidant capacity in cookies
made from anthocyanin-​rich maize flour. Flour derived from milled conven-
tional maize (Zea mays) presents limitations in breadmaking due to its flavor
and distinctive yellow color. Maize breads are characterized by a strong yellow
color, a low specific volume, and a dense and firm crumb (Capriles and Areas,
2014). Because of its color, white maize flour is an interesting alternative for the
production of gluten free bread. According to de la Hera et al. (2013), coarser
particle size maize flours are best suited for making gluten free bread, since
they give products with more volume and less firmness than bread made with
finer flour.
Giménez et al. (2014) studied the acceptability of spaghetti of two races of
Andean maize in celiac and non-​celiac consumer groups. In this work, celiac
consumers assigned high acceptability scores to the samples and described it
as tasty, smooth, tender, novel, with a pleasant flavor and good quality. Also,
they characterized maize spaghetti as a product that can be consumed every
day and for the whole family. However, this product should be reformulated
for non-​celiac consumers. The combination of maize flour with other gluten-​
free flour of pseudo-​cereals, legumes and vegetables among others provided
an increase in the nutritional value of maize gluten free products such as
pastas and bread (Padalino et al., 2013; Dib et al., 2018). The gluten free spa-
ghetti and alfajores made with flours of maize, quinoa (Chenopodium quinoa),
amaranth (Amaranthus caudatus) and broad bean (Vicia faba) remarkably
increased the content of dietary fiber, unsaturated fatty acids, iron and zinc.
These nutritionally improved spaghettis had a higher “cooking loss” than spa-
ghetti made with only maize, and adequate sensorial properties (Giménez
et al., 2013, 2016).
Other gluten free products such as toasted maize grains with salt and other
local condiments and Chilcan (toasted ground maize flour, with spices and
sugar to prepare a drink) are commonly consumed in the Andean zone of Jujuy,
Argentine as an energetic food.
180 Latin-American Crops in Gluten-Free Applications

5.4 SACHA INCHI AND BLACK TURTLE BEAN


Plukenetia volubilis (L.) and Phaseolus vulgaris (L.) species are plants from the oil-
seed and legume family. Around the world, they are known by their common name
sacha inchi and black turtle bean, respectively; some cultivars of Phaseolus vulgaris
(L.) are also known as common beans. However, in this chapter only black turtle
beans are included. Their nutritional composition and the presence of bioactive
compounds make these products a good alternative to extend their consumption,
transforming them into raw ingredients such as flours and using them to develop
new foods with added value. Sacha inchi and this legume seeds are novel foods that
can be used as unconventional flours in some formulations. Their consumption
could reduce malnutrition, guarantee food security and promote healthy eating,
helping to ensure the longevity of the population by improving quality of life.
Lipids are the major compound present in sacha inchi seeds (35–​60%), higher
than values reported for chia and soybean seeds (Chirinos et al., 2013). Their
lipid profile, rich in unsaturated (6.8–​9.1%) and saturated fatty acids, is a good
source of α-​linolenic acid, linoleic acid and oleic acid (Wang et al., 2018) with a
high content of tocopherols (Štěrbová et al., 2017). Furthermore, seeds present
elevated protein content (27%) with relatively high levels of cysteine, tyrosine,
threonine and tryptophan in their amino-​acid profile. Castaño et al. (2019) and
de la Rosa-​Millán et al. (2019) reported that powder from these seeds contained
more than 50% of protein.
Black turtle bean is a good source of carbohydrates (50–​60%) being starch,
the main component. These pulses present dietary fiber, ash (> 4%), high pro-
tein (20–​25%) (Rocchetti et al., 2019) and low fat content (Ikezu et al., 2015).
Bioactive compounds (i.e., saponins, flavonoids) and phenolics are also present
(de la Rosa-​Millán et al., 2019).
In this chapter various terms are used to indicate the particle size of solids
like flour (particle size: 100–​5000 µm) and powder (particle size: 50–​200 µm)
following the criteria of Bhandari (2013).
González et al. (2018) evaluated the nitrogen balance in blood samples of
30 healthy adults after consuming meals with 30g of sacha inchi and soy flour,
respectively. They showed that nitrogen balance was similar in both cases and
recommended these sources of vegetable protein as a good healthy option to be
consumed. Some researchers highlighted their antioxidative, antihypertensive
(Chirinos et al., 2020), antibacterial, anticancer and antidyslipidemic cap-
acity (Chirinos et al., 2013; González-​Aspajo et al., 2015; Štěrbová et al., 2017).
Moreover, Alayón et al. (2018) evaluated the acute postprandial effect of sacha
inchi oil on changes in carbohydrate metabolism in clinically healthy subjects.
The authors showed that consumption of sacha inchi oil improved carbohydrate
metabolism, increasing insulin sensitivity.
5.4 Sacha Inchi and Black Turtle Bean 181

On the other hand, several authors have shown the anticancer activity
(Kumar et al., 2017) of black turtle beans related to phytochemical compounds
(Ikezu et al., 2015). However, there is scientific evidence regarding the low digest-
ibility of pulse protein mainly due to anti-​nutritional phytic acid, tannins and
cell walls that are present in the seeds among other components (Sozer et al.,
2017). Other authors have shown that a low-​pH incubation reduces the aller-
genicity of black turtle bean lectin to enhance the safety of legume consumption
(Audu et al., 2013).

5.4.1 Methods of Processing and Industrial Applications


Sacha inchi seeds are usually employed for oil extraction taking advantage of
their high content (Wang et al., 2018) and their residues have been used for
protein extraction (59%) (Chirinos et al., 2017). Black turtle beans are boiled in
water to elaborate black bean stew or are blended with rice to prepare a typical
dish consumed in Cuba named “arroz congris” (Figure 5.6).

Figure 5.6 Traditional Cuban dish “arroz congris or moros y cristianos” rice with black
turtle beans. Source: Authors’ own work.
Note: The image on the left shows a large frying pan (“paellero”) with rice and turtle black
beans on a wooden table in a terrace with views of a city and vegetation at the sea coast.
The image on the right shows an up-​close view of the Cuban dish seen from the top, with
whole turtle black beans, vegetables and rice.
182 Latin-American Crops in Gluten-Free Applications

In order to extend their industrial application, sacha inchi seed and black
beans’ potential must be explored. Analyzing technological characteristics shows
that sacha inchi seeds have a strong flavor and contain heat-​labile substances
with a bitter taste (Bueno-​Borges et al., 2018; Chirinos et al., 2013) that may nega-
tively impact the overall sensory characteristics of processed foods.
Roasting, baking, steaming, boiling, extrusion, milling and drying are the con-
ventional methods of processing seeds to obtain several raw materials and to
process them into edible food. Roasted seeds had been employed to produce
snacks (Wang et al., 2018). The authors recommend optimizing roasting
conditions since this parameter may influence the flavor, antinutrients (Bueno-​
Borges et al., 2018), nutritional composition (Kim and Joo, 2019) and healthy
properties. In order to obtain raw materials without anti-​nutritional factors
from sacha inchi seeds, several authors have employed different alternatives.
Suwanangul et al. (2021) showed that the extrusion and autoclaving process
causes the destruction of tannins, phytates, and trypsin inhibitors. A similar
conclusion was reported by some authors who showed that an optimal roasting
process (temperatures and time) influences the nutritional and phytochemical
composition (Chirinos et al., 2013; Kim and Joo, 2019), increasing the antioxidant
capacity and total phenolic compounds (Štěrbová et al., 2017). A similar study
conducted by Bueno-​Borges et al. (2018), demonstrated that roasting seeds for
15 min at 160ºC decreased all assessed antinutrients except saponins.
In general, seeds can be aggregated into foods or be transformed into powder;
they could also be added using the main by-​product obtained after oil extrac-
tion, as pressed cake. Pressed cakes are the main by-​product obtained from
sacha-​inchi oil extraction, contenting bioactive compounds with high antioxi-
dant activities (Rawdkuen et al., 2016). Additionally, this product presents good
physical properties (low solubility: 7.96%, high volume: 3.92 mL/​g and low water
adsorption capacity: 2.16 g/​g ) (Alcívar et al., 2020). Residual seed cake could
be transformed into a powder in order to be used in the food industry. In the
past, sacha inchi pressed-​cake flour was mainly recommended in animal feeding
(Alcívar et al., 2020; Viamonte-​Garcés et al., 2020). Recently, several authors
have proposed this material as nonconventional sources of isolated vegetable
proteins from agro-​industry since the protein content from the remaining cake
was more than 50% (Chirinos et al., 2017; Mercado et al., 2015). In this context,
some authors have shown the in vitro antioxidant and antihypertensive poten-
tial of bioactive peptides from protein hydrolysate (Chirinos et al., 2020).
Concerning flour from black turtle bean, some authors emphasized that
different processing methods (boiling, cooking, roasting, sprouting and
fermenting) significantly influenced their functional and technological
parameters, likewise, their anti-​nutritional factors (Audu and Aremu, 2015;
Audu et al. 2013; Gobbetti et al., 2020). For example, during milling some crit-
ical parameters influence the quality of flour like particle size. In this sense,
5.4 Sacha Inchi and Black Turtle Bean 183

Balandrán-​Quintana et al. (1998) suggested grounding the seeds from pinto


beans to pass through a 40-​mesh screen in order to develop extruded whole pinto
bean flour. While Gularte et al. (2011) employed powder from chickpeas, peas,
lentils and beans with particle size lower than 210 μm. Chompoorat et al. (2018)
sieved seeds to obtain powders with a particle size less than 250 μm in order to
develop gluten-​free cakes. Regarding the fermentation process, Gobbetti et al.
(2020) explained that sourdough is a powerful process for exploiting the poten-
tial of legumes in the baking goods industry.

5.4.2 Baked Goods Application


Concerning food application, Lee (2019) elaborated a pound cake with 20%
of sacha inchi powder and evaluated its technological properties, confirming
good results in quality and sensory characteristics. Vázquez-​Osorio et al. (2017)
developed cookies including 25, 50 and 100% of sacha inchi flour. The best results
were obtained at 50% inclusion and recommended evaluating the influence of
100% studying different strategies in order to obtain cookies with good techno-
logical and sensory properties. Aylas et al. (2015) substituted 10% of wheat flour
with sacha inchi residual cake, obtaining a sweet bread with acceptable sensory
parameters. Betancur et al. (2016) confirmed a high nutritional composition
and sensory acceptance of the snack including sacha inchi seed flour as a novel
ingredient.
Cappa et al. (2020) evaluated the behavior of 25 edible dry bean powder during
the baking process in order to develop healthy cookies with higher protein
content, lower rapid digestible starch and higher resistant starch. The authors
concluded that bean powder with particle size similar to wheat flour (particle
size ≤ 0.5 mm) showed better results in terms of cookie diameter and hardness.

5.4.3 Gluten-​Free Products
Due to their high nutritional value and safety for human consumption sacha
inchi and black turtle bean seeds are good alternatives for use in gluten free
baked goods (Rawdkuen et al., 2016). Castaño et al. (2019) employed 100%
residual cake seeds from sacha inchi to develop a gluten-​free brownie; first, the
powder was treated for 3h at 80°C to eliminate typical astringency.
To extend the use of the black turtle bean several authors have studied inclu-
sion of its powder into formulations of gluten-​free baked goods (Chompoorat
et al., 2018; Gularte et al., 2011) showing good results in order to improve the
technological properties and nutritional value. For example, Chompoorat
et al. (2018) evaluated the influence of precooked red kidney beans on the visco-
elastic properties of gluten-​free cake batter and cupcake. The authors showed
that different thermal processes improved batter viscoelasticity and sensory
184 Latin-American Crops in Gluten-Free Applications

characteristics. They also concluded that boiling red kidney beans for 20 min
with 3-​h drying at 80ºC is sufficient to obtain an acceptable legume powder
for gluten free cupcake production. Therefore, Medina et al. (2018) obtained
brownies from legume flour with added value, since this flour increases the anti-
oxidant content and promotes inhibition of α-​glucosidase.
Studies mentioned previously showed the powerful potential of seeds and
common pulses to be included in baked goods. Processing methods and types of
technologies employed to develop baked goods from sacha inchi and black turtle
bean are the main parameters that must be controlled in order to develop value-​
added products. Some strategies like the addition of hydrocolloids, enzymes
and starch blended with powders could improve the overall quality of products.
Besides that, more scientific studies are needed in these fields because there is
scarce information about the use of sacha inchi and black turtle bean powder in
gluten-​free products.
Commercial gluten-​free products with Sacha inchi in Cuban markets can be
seen in Figure 5.7.

B C

D E

Figure 5.7 Gluten free goods from Cuban bakeries with Sacha inchi; A: burger bread, B:
loaf bread, C: muffins, D: brownie, E: brioche bread. Source: Authors’ own work.
Notes: A) Four burger breads made with sacha inchi viewed from the front.
B) Four loaf breads, seen sideways, positioned horizontally and parallel to each other.
C) Muffins made with sacha inchi in their paper muffin cups (of different sizes) viewed
from above.
D) Square shaped brownies made with sacha inchi viewed from above.
E) Many brioche breads made with sacha inchi viewed from above.
5.5 Chia 185

5.5 CHIA
Due to the technological importance of gluten, today the production of gluten-​
free food products is still a challenge for the industry. Indeed, large quantities
of fat, sugars, structuring agents, and flavor enhancers are added to gluten free
formulations to make textural and sensorial characteristics comparable to
conventional products, resulting in nutritional and caloric intake imbalances
(Montemurro et al., 2021). Chia seeds do not contain gluten (Ali et al., 2012) and
due to their physical-​chemical properties, these seeds have different techno-
logical capabilities and potential utilities for food application (Figure 5.8). Due to
its nutraceutical and physicochemical properties, chia has been widely used as
a whole seed, flour, seed mucilage, gel and oil to develop various enriched food
products, such as bread, pasta, cakes, cookies, chips, cheese, yoghurt, meat, fish
and poultry (Kaur and Bains, 2019).
One of the most important aspects of this seed is its high fiber content with
a high proportion of mucilage, capable of absorbing large quantities of water
(Reyes-​Caudillo et al., 2008) (Figure 5.9). The addition of whole or ground chia
had a favorable effect on bread crumb texture, as chia swells intensely with
water and can be used as a hydrocolloid replacement in gluten free bread or as
a fat replacer and thickener (Zettel and Hitzmann, 2018; De Lamo and Gómez,
2018; Ziemichod et al., 2019).
Another notable aspect of chia is its high content in omega-​3-​rich oil that
is appropriately balanced with omega-​6 (Ayerza, 2010). Therefore, chia seeds
and co-​products are excellent food ingredients with which to enrich foods by
offering good technological benefit due to their mucilage properties, oil and
proteins (Iglesias-​Puig and Haros, 2013). Nonetheless, an in-​depth study about
baked foods is necessary because ingredients rich in omega-​3 are oxidized
while being baked (Miranda-​Ramos et al., 2020), and those ingredients rich
in proteins could favor acrylamide formation (Mesías et al., 2016) especially
in foods with low water activity. Acrylamide is mainly produced at elevated
temperatures and medium to low moisture content as a consequence of aspara-
gine degradation produced by the Maillard reaction initiated by carbohydrates
(Zamora and Hidalgo, 2011). However, lipid carbonyls can also play a role in
these reactions. Asparagine degradation in the presence of lipid carbonyls is
a two-​step reaction in which the amino acid is first decarboxylated produ-
cing the corresponding biogenic amine (3-​aminopro-​pionamide) and then
converted to acrylamide by elimination of ammonia by deamination (Zamora
et al., 2009). The European Food Safety Authority (EFSA) has announced that
the presence of acrylamide in food is a public health problem and that efforts
must still be made to reduce its exposure (EFSA, 2015). Therefore, such aspects
need to be thoroughly investigated. In this sense, chia seeds are considered
a “novel food” in the European Union (EU) and have restrictions despite the
186 Latin-American Crops in Gluten-Free Applications

A B

C D

Figure 5.8 Gluten free chia foods: A. Lemon chia cookies; B. Soy lecithin with chia; C.
nachos with chia; D. tortillas with chia. Source: Authors’ own work.
Notes: A) Three round lemon cookies with whole chia seeds, with a rough surface, on a
medium round plate.
B) Granules of soy lecithin with black chia seeds in a small cup, viewed from above.
C) “Nachos” with embedded chia seeds in a small bowl.
D) Medium round plate with a round amaranth chia; the seeds are embedded on the
surface.

huge interest due to their technological, nutritional and functional poten-


tial. According to Regulation 258/​97, new foods and new food ingredients are
defined:

as food that has not been consumed to a significant degree by humans in


the EU prior to 1997, when the first Regulation on “novel food” came into
force. Novel food can be newly developed innovative food or food produced
using new technologies and production processes as well as food tradition-
ally eaten outside of the EU.
OJEU, 1997
5.5 Chia 187

Figure 5.9 Lyophilized hydrated chia. Source: Authors’ own work.

There are not many records about the intake of chia in Europe before 1997. So
after considering extensive studies about the nutritional characteristics of chia,
the EFSA issued a report in favor of its commercialization as “novel food” with
restrictions (EFSA, 2009). Whole chia seeds could be used in baked products
and breakfast cereals up to 10%; ground chia seeds up to 5% in bread; whole
chia seeds up to 5% in sterilized ready-​to-​eat meals based on cereal/​pseudo-​
cereal grains and/​or pulses; pre-​packaged chia seeds as such, and fruit/​nut/​
seed mixes; and chia in confectionery products and chocolates; edible ices; fruit
and vegetable products; non-​alcoholic beverages and puddings (<120ºC in their
preparation) without limitations (EFSA, 2020). Recently, the use of two partially
defatted powders of chia enriched with proteins or fibers was authorized as food
supplements for the adult population (up to 7.5 and 12 g/​day, respectively), or as
188 Latin-American Crops in Gluten-Free Applications

nutritional ingredients in a variety of foods (yogurt, vegetable beverages, energy


drinks, chocolate, fruit and pasta) at a level of 0.7–​10% (Turck et al., 2019).
Nowadays, technical and administrative formalities are accomplished by com-
panies to allow the commercialization and use of other ingredients of interest in
Europe. For this purpose, the nutritional potential of these ingredients will have
to be validated and their innocuousness has to be guaranteed, as set out in the
legislation currently in force. Outside the EU, whole chia and its seed fractions
(whole chia flours, fractions rich in proteins, fractions rich in fibers, chia oil,
chia partially defatted, etc.) have no restrictions regarding their utility in food
preparation.

5.5.1 Processing Chia Seeds


Alternatives to the direct consumption of chia seeds could be recommended.
Some processes such as soaking, sprouting or fermentation could increase
its digestibility. From these alternatives, sprouting has gained popularity as it
is a simple method that improves the nutritional value of the seeds, including
increased phenols and mineral availability (Calvo-​Lerma et al., 2020). Similarly,
the milling of seeds has also been proven to enhance nutrient digestibility
because of the matrix disruption and ease of nutrient release from the interior of
plant cells (Grundy et al., 2016).
Individual chia components cannot be obtained by either wet or dry milling,
as in the case of cereals/​pseudo-​cereals. The main chia components are extracted
from whole seeds, or wholemeal flour, by milling or crushing whole seeds. The
most popular procedure for particle size reduction and homogenization is
cryogenic grinding (Gouveia et al., 2002). The temperature rise during grinding
can be reduced with the use of a cryogenic fluid mainly to avoid thermo-​labile
compounds such as unsaturated fatty acids (Gouveia et al., 2002). Cryogenic
fluids must be sufficiently inert not to give up components to the foods, without
affecting the composition of the food or altering their sensorial characteristics.
Besides, the low temperature keeps the material brittle during grinding and
helps to achieve a finer particle size.
Chia oil, which is rich in omega-​3, can be extracted by cold-​pressing, solid-​
liquid hexane extraction or by supercritical CO2 extraction (Ixtaína et al., 2010).
The oil extraction residue has a high content of proteins (19–​23%) and fiber
(~ 37%); moreover, it has high potential to be applied to human food and animal
feed. On the other hand, chia offers a higher fiber content than cereals/​pseudo-​
cereals, and a high proportion of soluble fiber. Mucilage is a compound that
exhibits high viscosity in water (Lin et al., 1994) and is extracted by seed hydra-
tion, followed by drying and sieving, and offers excellent techno-​functional
properties (Muñoz et al., 2009, 2012). The residual flour from the oil extraction
5.5 Chia 189

process and the fiber fraction (mucilage) are rich in protein with a high bio-
logical value. All of these flours from chia processing are potential ingredients
for developing gluten free products with a high nutritional/​functional value and
excellent techno-​functional quality.

5.5.2 Gluten-​Free Foods with Chia


5.5.2.1 Bread Products
The inclusion of chia or other oilseeds modifies the rheology of the dough,
depending on the way in which it is made ( flour or seeds, prehydration or not)
and the percentage used (de Lamo and Gómez, 2018). To ensure commercial
success of these inclusions, it is necessary to consider the acceptability of con-
sumers, which may vary depending on the type of inclusion and its percentage
(de Lamo and Gómez, 2018). In addition, as stated above, due to the muci-
lage release, chia and its by-​products can also be regarded as a technological
improvement for gluten-​free breads (Zettel and Hitzmann, 2018).
Rice breads were elaborated with the addition of 15g of chia flour or seeds,
either dry or pre-​hydrated, per 100g of rice flour. In general, the addition of
chia reduced bread specific volume, increased the crumb firmness with a
darker crust/​crumb color compared to the control sample, and the effect was
more evident with the flour than with the seed (Steffolani et al., 2014). The pre-
vious hydration significantly modified the dough rheology but it seems did not
produce significant differences in the final product formulations (Steffolani
et al., 2014). On the other hand, Montemurro et al. (2021) developed a formula-
tion of the novel “clean-​label” gluten free bread including a commonly used mix-
ture of maize and rice flour (ratio 1:1); it was fortified with quinoa and naturally
hydrocolloid-​containing flours, such as psyllium, flaxseed or chia, as structuring
agents for their mucilage content. The use of chia at a 2.5% concentration could
possibly replace gum in gluten-​free bread formulations (Huerta et al., 2016).
In another study, rice flour bread was augmented with chia seeds and chia
flour which resulted in products with good technological characteristics. The
addition of chia seeds and flour in gluten-​free bread provided an increase
in the nutritional value of the products, mainly in the protein content; all the
breads tested were well accepted by consumers, obtaining global acceptance
values above 70%, with no significant difference between them (da Costa Borges
et al., 2021). The development of premixes based on buckwheat and chia flour
represents an alternative with high nutritional value that can be applied to
the formulation of gluten-​free breads, with adequate technological properties
(Costantini et al., 2014; Coronel et al., 2021).
Some chia by-​ products such as partially defatted chia flour,
hydroxypropylmethylcellulose gum (HPMC) and xanthan gum were used, in
190 Latin-American Crops in Gluten-Free Applications

order to develop gluten-​free breads with high nutritional and sensory quality
(Ewerling et al., 2020). Even chia seeds’ by-​products of oil pressing can be
regarded as valuable technological and functional food additives in gluten-​free
bakery products (Zdybel et al., 2019). On the other hand, rice breads with added
chia protein hydrolysate resulted in good technological characteristics and anti-
oxidant properties, and breads made with 3 mg of chia hydrolysate/​g flour had
good sensory characteristics, which were not lower than those of the control
breads (Madruga et al., 2020).
The use of chia sourdoughs fermented with selected lactic acid bacteria
strains as ingredients for gluten-​free bread making, could have an impact on the
functionality and healthiness of baked foods (Maidana et al., 2020a). The poten-
tial use of Lactobacillus sanfranciscensis W2 for the fermentation of chia flour to
produce gluten-​free maize/​rice bread was investigated. The strain adapted to
the environment of the chia flour, which decreased pH, specific volume, firmness
and rate of bread staling and increased bread porosity compared to bread only
made with chia seed flour (Jagelaviciute and Cizeikiene, 2021). Sorghum-​based
laboratory breads manufactured with different chia percentages or flaxseed
sourdoughs fermented by Weissella cibaria CH28 together with L. plantarum
FUA3171 and L. fermentum FUA3165 from fermented sorghum, significantly
improved specific volume and visual appearance compared to 100% sorghum
breads, with higher acceptability by panelists (Maidana et al., 2020a). The devel-
opment of lactic acid bacteria during spontaneous chia flour fermentation (sour-
dough) was investigated by Maidana et al. (2020b). Lactic acid bacteria species
from genus Enterococcus (E. faecium, E. mundtii) were the most abundant; they
were isolated species from Lactococcus (Lc. laths), Lactobacillus (L. rhamnosus)
and Weissella (W. cibaria). They were the dominant species in the final propaga-
tion stages while Bacillus and Clostridium were mostly present during the first
fermentation stages.

5.5.2.2 Biscuits, Cookies, Cakes and Snacks


In addition to the development of breads, some formulations of gluten-​free biscuits,
cookies, cakes and snacks that include chia have been described in the literature.
The substitution of buckwheat flour, millet flour, and chia seeds for wheat flour can
be considered a suitable alternative for the preparation of gluten-​free cookies. The
optimum formulation was defined as that with the closest characteristics (diam-
eter, expansion factor, thickness and hardness) to the control cookie elaborated
with wheat, which was with 7.5% chia, 40% millet and 52.5% buckwheat (Ferreira
Brites et al., 2019). In another investigation, an optimized gluten-​free biscuit formu-
lation with chia seeds and curcuma was developed with the same statically sensory
characteristics as the control biscuits (Laczkowski et al., 2021).
5.6 Conclusions 191

The effect of using chia gel as a gluten substitute in cakes was also investigated
which showed better acceptance by consumers than the control formulation
made with rice flour (Hargreaves et al., 2018). Substituting rice flour with 10%
prehydrated chia seed flour can achieve an acceptable piece center height and
volume index of gluten-​free layer cakes with a similar overall acceptability, tex-
ture, flavor and odor scores compared to those of gluten-​free layer cake made
with 100% rice flour and layer cake made with 100% wheat flour (Sung et al,
2020). Chia and azuki flour mixes were also investigated to develop a gluten-​
free chocolate cake with a higher amount of polyunsaturated fatty acids, mainly
alpha-​linolenic acid, which showed that the inclusion of chia can improve lipid
composition (Gohara et al., 2016).
Chia was also utilized to prepare sweet snacks suitable for celiac people using
popped amaranth, dried peaches, textured soy, corn flakes and whole sesame
seeds agglutinated with glucose syrup, honey and soy lecithin and flavored with
cinnamon in order to develop a highly nutritious product with a shelf life of
around 4 months (Malka et al., 2020).

5.5.2.3 Pasta Products
Some investigators developed pasta products with chia. This seed is generally
used in gluten-​free pasta formulations for its nutritional value and thickening
properties (Levent, 2017; Menga et al., 2017). Formulation of noodles with
20% of chia seed flour with DATEM (diacetyl tartaric acid ester of mono-​and
diglycerides) can be used with acceptable sensory attributes of raw and cooked
samples, to improve the nutritional quality of noodles (Levent, 2017). On the
other hand, partially defatted chia flour, high in protein content, dietary fiber
and phenolic compounds was investigated as a gluten-​free ingredient to develop
pasta with an improved antioxidant capacity (Aranibar et al., 2018).

5.6 CONCLUSIONS
Quinoa, kañiwa, amaranth, tarwi, maize, chia, black turtle bean and sacha inchi
are nutritious crops cultivated in Latin America. These grains are traditionally
consumed by local inhabitants directly, without further processing and trans-
formation. They also offer excellent ingredients for the modern food industry.
They are all gluten free and can be used to produce bakery and pasta products
without using wheat. The protein quality of quinoa, kiwicha and kañiwa is
excellent; they can all substantially improve the nutritional value of traditional
gluten-​free products. Tarwi has an exceptionally high protein content, making
it an interesting ingredient for the preparation of protein concentrates. The
nixtamalization process used for maize in Mexico and Central America increases
192 Latin-American Crops in Gluten-Free Applications

the calcium content in maize grains, improving the bioavailability of niacin and
iron, and decreasing the content of mycotoxins and phytic acid. Sacha inchi has
a very high oil content, being an excellent source of α-​linolenic, linoleic and oleic
acids and of tocopherols. It also has an elevated protein content. The flour of
black turtle beans can be included in gluten-​free baked goods, improving the
technological properties and nutritional values of the final product. Due to the
physical–​chemical properties of chia seeds, they have interesting technological
properties and potential uses for food applications. Chia has been widely used as
a whole seed, flour, seed mucilage, gel and oil to develop various enriched food
products, such as bread, pasta, cakes, cookies, chips, cheese, yoghurt, meat, fish
and poultry.
These crops are staple foods for millions of people living in areas of Latin
America where nutritional problems are common. They have been part of
their traditional diets and are still important for food security. Additionally,
these seeds are also a source of interesting ingredients for novel and specialty
products beyond their traditional cultivation area, as consumers worldwide
are increasingly conscious of their nutritional, healthy and sustainable foods.
Unfortunately, there are important sectors of society in most geographical areas
of the north and south, where excess weight and obesity have become endemic
diseases. Thus, research is also needed to develop tasty and nutritious products
but simultaneously with a low glycemic index for those consumers; crops like
those analyzed in this chapter are called on to play a key role in the food strategy
of the future which has already arrived.

ACKNOWLEDGMENTS
Authors would like to express their gratitude to the CYTED Program for its
kind support through Project119RT0567, Food4ImNut Food4ImNut PID2019-
107650RB-C21 funded by MCIN/AEI/10.13039/501100011033, Spain.

REFERENCES
Abarza, S. del V., and J. H. Schimpf. 2009. Los Maíces indígenas en la cultura alimentaria
andina. Revista Cientifica Ciencia 4(9): 159–​170.
Alayón, A. N., J. G Ortega Ávila, and I. Echeverri Jiménez. 2018. Carbohydrate metabolism
and gene expression of sirtuin 1 in healthy subjects after Sacha inchi oil supplemen-
tation: A randomized trial. Food and Function 00: 1–​9. doi:10.1039/​C7FO01956D.
Alcívar, J. L., M. Martínez Pérez, P. Lezcano, I. Scull, and A. Valverde. 2020. Technical note
on physical-​chemical composition of Sacha inchi (Plukenetia volubilis) cake. Cuban
Journal of Agricultural Science 54(1): 19–​23.
References 193

Ali, Norlaily Mohd, Swee Keong Yeap, Wan Yong Ho, Boon Kee Beh, Sheau Wei Tan, and
Soon Guan Tan. 2012. The promising future of chia, salvia hispanica L. Journal of
Biomedicine and Biotechnology. https://​doi.org/​10.1155/​2012/​171​956.
Alonso-​ Miravalles, L., and J. A. O’Mahony. 2018. Composition, protein profile and
rheological properties of pseudocereal-​ based protein-​ rich ingredients. Foods
7(73): 1–​17.
Álvarez-​Jubete, L., M. Holse, A. Hansen, E. Arendt, and E. Gallagher. 2009. Impact of baking
on vitamin E content of pseudocereals amaranth, quinoa, and buckwheat. Cereal
Chemistry 86(5): 511–​515.
Álvarez-​Jubete, L., M. Auty, E. Arendt, and E. Gallagher. 2010a. Baking properties and
microstructure of pseudocereal flours in gluten-​free bread formulations. European
Food Research and Technology 230: 437–​445.
Álvarez-​Jubete, L., H. Wijngaard, E. Arendt, and E. Gallagher. 2010b. Polyphenol com-
position and in-​vitro antioxidant activity of amaranth, quinoa and buckwheat as
affected by sprouting and bread baking. Food Chemistry 119: 770–​778.
Ando, H., Y. C. Chen, H. Tang, M. Shimizu, K. Watanabe, and T. Mitsunaga. 2002. Food
components in fractions of quinoa seed. Food Science and Technology Research
8(1): 80–​84.
Aranibar, C., N. B. Pigni, M. Martinez, A. Aguirre, P. Ribotta, D. Wunderlin, and R. Borneo.
2018. Utilization of a partially-​deoiled chia flour to improve the nutritional
and antioxidant properties of wheat pasta. LWT –​ Food Science and Technology
89: 381–​387.
Arendt, E. K., A. Morrissey, M. M. Moore, and F. dal Bello. 2008. Gluten-​free breads. In
Gluten-​Free Cereal Products and Beverages, edited by E. K. Arendt and F. dal Bello,
289–​319. New York: Academic Press.
Audu, S. S., and M. O. Aremu. 2015. Effect of domestic processing on the levels of some
functional parameters in black turtle bean (Phaseolus vulgaris L.). Food Science and
Quality Management 38: 55–​59.
Audu, S. S., M. O. Aremu, and L. Lajide. 2013. Effects of processing on physicochemical and
antinutritional properties of black turtle bean (Phaseolus vulgaris L.) seeds flour.
Oriental Journal of Chemistry 29(03): 979–​989.
Ayerza, R. 2010. Effects of seed colour and growing locations on fatty acid content and
composition of two chia (Salvia hispanica L.) genotypes. Journal of the American Oil
Chemists’ Society, 87: 1161–​1165. http://​dx.doi.org/​10.1007/​s11​746-​010-​1597-​7.
Aylas, T., M. Pingus, and D. Alva. 2015. Effect of the partial replacement of wheat flour by
cake of sacha inchi (Plukenelia volubilis L.) on the rheological properties of dough in
sweet bread. Revista Ciencia & Desarrollo 2015: 16–​21.
Balandrán-​Quintana, R. R., G. V. Barbosa-​Cánovas, J. J. Zazueta-​Morales, A. Anzaldúa-​
Morales, and A. Quintero-​Ramos. 1998. Functional and nutritional properties of
extruded whole pinto bean meal (Phaseolus Vulgaris L.). Journal of Food Science
63(1): 113–​116.
Ballester-​Sánchez, J., J. V. Gil, M. T. Fernández-​Espinar, and C. M. Haros. 2019. Quinoa
wet-​milling: Effect of steeping conditions on starch recovery and quality. Food
Hydrocolloids 89: 837–​843.
Bhandari, B. 2013. Introduction to food powders. In Handbook of Food Powders. Processes
and Properties, edited by Bhesh Bhandari, Nidhi Bansal, Min Zhang and Pierre
Schuck, 1–​25. Cambridge, UK: Woodhead Publishing.
194 Latin-American Crops in Gluten-Free Applications

Basilio-​Atencio, J., L. Condezo-​Hoyos, and R. Repo-​Carrasco-​Valencia. 2020. Effect of extru-


sion cooking on the physical-​chemical properties of whole kiwicha (Amaranthus
caudatus L) flour variety Centenario: Process optimization. LWT –​ Food Science and
Technology 128: 109426.
Benavent-​Gil, Y., and C. M. Rosell. 2018. Technological and Nutritional Applications of
Starches in Gluten-​Free Products. Starches for Food Application. London: Elsevier.
Betalleluz-​Pallardel, I., M. Inga, L. Mera, R. Pedreschi, D. Campos, and R. Chirinos. 2017.
Optimization of extraction conditions and thermal properties of protein from the
Andean pseudocereal canihua (Chenopodium pallidicaule Aellen). International
Journal of Food Science and Technology 52: 1026–​1034.
Betancur, Edwin, Amparo Urango, Marchena Luz, Luis Fernando Restrepo Betancur.
2016. Effect of adding sacha inchi (Plukenetia volubilis L.) seeds to a proto-
type of convenience food draft, on the nutritional composition and sensory
acceptance. Journal of Medicinal Plant Research 10(29):435–​441, DOI: 10.5897/​
JMPR2016.6064.
Bressani, R. 1990. Chemistry, technology, and nutritive value of maize tortillas. Food
Reviews International 6(2): 225–​264.
Bressani, R., J, C. Turcios, and A. S. C. de Ruiz. 2002. Nixtamalization effects on the contents
of phytic acid, calcium, iron and zinc in the whole grain, endosperm and germ of
maize. Food Science and Technology International 8(2): 81–​86.
Bueno-​Borges, Larissa Braga, Marco Aurélio Sartim, Claudia Carreño Gil, Suely Vilela
Sampaio, Paulo Hercílio Viegas Rodrigues and Marisa Aparecida Bismara Regitano-​
d’Arce. 2018. Sacha inchi seeds from sub-​tropical cultivation: Effects of roasting on
antinutrients, antioxidant capacity and oxidative stability. Journal of Food Science
and Technology 55: 4159–​4166).
Bugallo, L. and L. Mamani. 2014. Molinos en la quebrada de Humahuaca: lugares de
encuentro de gentes y caminos. La región molinera del norte jujeño, 1940-​1980. In
Espacialidades Altoandinas. Avances de investigación desde el noroeste argentino,
edited by A. Benedetti and J. Tomasi, 63–​118. Editorial de la Facultad de Filosofía y
Letras, Universidad de Buenos Aires.
Burgos, V. E., M. J. Binaghi, P. A. Ronayne de Ferrer, and M. Armada. 2018. Effect of
precooking on antinutritional factors and mineral bioaccessibility in kiwicha grains.
Journal of Cereal Science 80: 9–​15.
Bustos, M. C., M. I. Ramos, G. T. Pérez, and A. E. León. 2019. Utilization of kañawa
(Chenopodium pallidicaule Aellen) flour in pasta making. Journal of Chemistry
2019: 4385045.
Calderón de la Barca, A. M., M. E. Rojas-​Martínez, A. R. Islas-​Rubio, and F. Cabrera-​Chávez.
2010. Gluten-​free breads and cookies of raw and popped amaranth flours with
attractive technological and nutritional qualities. Plant Foods for Human Nutrition
65: 241–​246.
Calvo-​Lerma, J., C. P. Yépez, A. A. Grau, A. Heredia, and A. Andrés. 2020. Impact of pro-
cessing and intestinal conditions on in vitro digestion of chia (Salvia hispanica)
seeds and derivatives. Foods 9(3): 290.
Calzetta, A. N., R. J. Aguerre, and C. Suárez. 2006. Hydration kinetics of amaranth grain.
Journal of Food Engineering 72: 247–​253.
Cámara-​Hernández, J., and D. Arancibia-​Cabezas. 2007. Maíces Andinos y sus Usos en la
Quebrada de Humahuaca y Regiones Vecinas (Argentina). Buenos Aires: Editorial
Facultad de Agronomía, Universidad de Buenos Aires.
References 195

Camelo-​Méndez, G., P. Flores, E. Agama-​Acevedo, J. Tovar, and L. Bello-​Pérez. 2017.


Incorporation of whole blue maize flour increases antioxidant capacity and reduces
in vitro starch digestibility of gluten-​free pasta. Starch –​Stärke 70: 1700126.
Cappa, C., J. D. Kelly, and P. K. W. Ng. 2020. Baking performance of 25 edible dry bean
powders: Correlation between cookie quality and rapid test indices. Food Chemistry
302: 125338.
Capriles, V., and J. Areas. 2014. Novel approaches in gluten-​free breadmaking: Interface
between food science, nutrition, and health. Comprehensive Reviews in Food Science
and Food Safety 13: 871–​890.
Carrera, Y., R. Utrilla Coello, A. Bello-​Pérez, J. Ávarez Ramírez, and E. Vernon Carter. 2015.
In vitro digestibility, crystallinity, rheological, thermal, particle size and morpho-
logical characteristics of pinole, a traditional energy food obtained from toasted
ground maize. Carbohydrate Polymers 123: 246–​255.
Castaño, L. A. C., M. M. P. Valderrama, H. J. P. Díaz, and S. M. Montesino. 2019.
Physicochemical and microbiological evaluation of flour obtained from the residual
cake of sacha inchi (Plukenetia volubilis l.) for its potential use in the agri-​food sector.
Italian Journal of Food Science : 31(6SI): 69–​78.
Chauhan, A., D. C. Saxena, and S. Singh. 2015. Total dietary fibre and antioxidant activity
of gluten free cookies made from raw and germinated amaranth (Amaranthus spp.)
flour. LWT –​ Food Science and Technology 63: 939–​945.
Chirinos, R., G. Zuloeta, R. Pedreschi, E. Mignolet, Y. Larondelle, and D. Campos. 2013.
Sacha inchi (Plukenetia volubilis): a seed source of polyunsaturated fatty acids,
tocopherols, phytosterols, phenolic compounds and antioxidant capacity. Food
Chemistry 141(3): 1732–​1739.
Chirinos, R., M. Aquino, R. Pedreschi, and D. Campos. 2017. Optimized methodology for
alkaline and enzyme-​assisted extraction of protein from sacha inchi (Plukenetia
volubilis) kernel cake. Journal of Food Process Engineering 40(2): e12412.
Chirinos, R., R. Pedreschi, and D. Campos. 2020. Enzyme‐assisted hydrolysates from sacha
inchi (Plukenetia volubilis) protein with in vitro antioxidant and antihypertensive
properties. Journal of Food Processing and Preservation 44: e14969.
Chompoorat, P., P. Rayas-​Duarte, Z. J. Hernández-​Estrada, C. Phetcharat, and Y. Khamsee.
2018. Effect of heat treatment on rheological properties of red kidney bean gluten
free cake batter and its relationship with cupcake quality. Journal of Food Science and
Technology 55(12): 4937–​4944.
Coda, R., C. G. Rizzello, and M. Gobbetti. 2010. Use of sourdough fermentation and
pseudo-​cereals and leguminous flours for the making of a functional bread enriched
of γ-​aminobutyric acid (GABA). International Journal of Food Microbiology 137(2–​
3): 236–​245.
Coronel, E. B., E. N. Guiotto, M. C. Aspiroz, M. C. Tomás, S. M. Nolasco, and M. I. Capitani. 2021.
Development of gluten-​free premixes with buckwheat and chia flours: Application in
a bread product. LWT –​ Food Science and Technology 141: 110916.
Corrales-​Bañuelos, A. B., E. O. Cuevas-​Rodríguez, J. A. Gutierrez-​Uribe, E. M. Milán-​Noris,
C. Reyes-​Moreno, J. Milán-​Carrillo, and S. Mora-​Rochín. 2016. Carotenoid compos-
ition and antioxidant activity of tortillas elaborated from pigmented maize landrace
by traditional nixtamalization or lime cooking extrusion process. Journal of Cereal
Science 69: 64–​70.
196 Latin-American Crops in Gluten-Free Applications

Cortés-​Avendaño, P., M. Tarvainen, J. P. Suomela, P. Glorio-​Paulet, B. Yang, and R. Repo-​


Carrasco-​Valencia. 2020. Profile and content of residual alkaloids in ten ecotypes of
Lupinus mutabilis Sweet after aqueous debittering process. Plant Foods for Human
Nutrition 75: 184–​191.
Costantini, L., L. Luksic, R. Molinari, I. Kreft, G. Bonafaccia, L. Manzi, and N. Merendino.
2014. Development of gluten-​free bread using tartary buckwheat and chia flour rich
in flavonoids and omega-​3 fatty acids as ingredients. Food Chemistry 165: 232–​240.
Cravioto, R. O., R. K. Anderson, E. E. Lockhart, F. P. Miranda, and R. S. Harris. 1945. Nutritive
value of the Mexican tortilla. Science 45(102): 91–​93.
da Costa Borges, V., S. Santos Fernandes, E. da Rosa Zavareze, C. M. Haros, C. Prentice
Hernandez, A. R. Guerra Días, and M. de las Mercedes Salas-​ Mellado. 2021.
Production of gluten free bread with flour and chia seeds (Salvia hispanica L). Food
Bioscience 43: 101294.
D’Amico, S., S. Jungkunz, G. Balasz, M. Foeste, M. Jekle, S. Tömösköszi, and R. Schöenlechner.
2019. Abrasive milling of quinoa: Study on the distribution of selected nutrients and
proteins within the quinoa seed kernel. Journal of Cereal Science 86: 132–​138.
de la Hera E, M. Talegon, P. Caballero, and M. Gomez. 2013. Influence of maize flour par-
ticle size on gluten-​free breadmaking. Journal of the Science of Food and Agriculture
93(4): 924–​32.
de la Parra, C., S. O. Serna Saldivar and R. H Liu. 2007. Effect of processing on the phyto-
chemical profiles and antioxidant activity of corn for production of masa, tortillas,
and tortilla chips. Journal of Agricultural and Food Chemistry 55(10): 4177–​4183.
de la Rosa-​Millán, J., E. Heredia-​Olea, E. Pérez-​Carrillo, D. Guajardo-​Flores, and S. R. O.
Serna-​Saldívar. 2019. Effect of decortication, germination and extrusion on physico-
chemical and in vitro protein and starch digestion characteristics of black beans
(Phaseolus vulgaris L.). LWT –​Food Science and Technology 102: 330–​337.
de Lamo, B., and M. Gómez, M. 2018. Bread enrichment with oilseeds. A Review. Foods
7(11): 191.
Dib, A., A. Wójtowicz, L. Benatallah, A. Bouasla, and M. Zidoune. 2018. Effect of hydro-
thermal treated corn flour addition on the quality of corn-​field bean gluten-​free
pasta. BIO Web of Conferences. Contemporary Research Trends in Agricultural
Engineering 10: 02003.
EFSA (European Food Safety Authority). 2009. Scientific opinion of the panel on dietetic
products nutrition and allergies on a request from the European Commission on the
safety of ‘Chia seed (Salvia hispanica) and ground whole Chia seed’ as a food ingre-
dient. The EFSA Journal 996: 1–​26.
EFSA (European Food Safety Authority). 2015. Scientific opinion on acrylamide in food.
EFSA Panel on Contaminants in the Food Chain (CONTAM). EFSA Journal 13(6):4104.
EFSA (European Food Safety Authority). 2020. Commission Implementing Regulation (EU)
2020/​24 of 13 January 2020 authorising an extension of use of chia seeds (Salvia
hispanica) as a novel food and the change of the conditions of use and the specific label-
ling requirements of chia seeds (Salvia hispanica) under Regulation (EU) 2015/​2283 of
the European Parliament and of the Council and amending Commission Implementing
Regulation (EU) 2017/​2470. Official Journal of the European Union L 8/​12.
EU (European Union). 2020. Authorising an extension of use of chia seeds (Salvia
hispanica) as a novel food and the change of the conditions of use and the specific
labelling requirements of chia seeds (Salvia hispanica) under Regulation (EU) 2015/​
References 197

2283 of the European Parliament and of the Council and Amending Commission
Implementing Regulation (EU) 2017/​2470. Oficial Journal of the European Union:
Brussels: 12–​16.
Ewerling, M., N. C. Steinmacher, M. R. dos Santos, D. L. Kalschne, N. E. de Souza, F. M.
Arcanjo, A. H. P. de Souza, and A. C. Rodrigues. 2020. Defatted chia flour improves
gluten-​free bread nutritional aspects: A model approach. Food Science and
Technology 40: 68–​75.
Ferreira Brites L. T. G., F. Ortolan, D. W. da Silva, F. R. Bueno, T. de Souza Rocha, Y. K. Chang,
and C. Joy Steel. 2019. Gluten-​free cookies elaborated with buckwheat flour, millet
flour and chia seeds. Food Science and Technology 39(2): 458–​466.
García Solaesa, A., M. Villanueva, A.J. Vela, and F. Ronda. 2020. Protein and lipid enrich-
ment of quinoa (cv.Titicaca) by dry fractionation. Techno-​functional, thermal and
rheological properties of milling fractions. Food Hydrocolloids 105: 1–​9.
Giménez, M. A., R. J. González, J. Wagner, R.Torres, M. O. Lobo, and N. C. Sammán.
2013. Effect of extrusion conditions on physicochemical and sensorial prop-
erties of corn-​broad beans (Vicia faba) spaghetti type pasta. Food Chemistry
136(2): 538–​545.
Giménez, M. A., A. Gámbaro, M. Miraballes, A. Roascio, M. Amarillo, N. Sammán, and
M. Lobo. 2014. Sensory evaluation and acceptability of gluten-​free Andean corn
spaghettis. Journal of the Science of Food and Agriculture 95: 186–​192.
Giménez, M. A., S. Drago, M. Bassett, M. Lobo, and N. Samman. 2016. Nutritional
improvement of corn pasta-​like product with broad bean (Vicia faba) and quinoa
(Chenopodium quinoa). Food Chemistry 199: 150–​156.
Gobbetti, M., M. De Angelis, R. Di Cagno, A. Polo, and C. G. Rizzello. 2020. The sourdough
fermentation is the powerful process to exploit the potential of legumes, pseudo-​
cereals and milling by-​products in baking industry. Critical Reviews in Food Science
and Nutrition 60(13): 2158–​2173.
Gohara, A. K., A. H. P. Souza, E. M. Rotta, G. L. Stroher, S. T. M. Gomes, J. V. Visentainer, N. E.
Souza, and Makoto Matsushita. 2016. Application of multivariate analysis to assess
the incorporation of omega-​3 fatty acid in gluten-​free cakes. Journal of the Brazilian
Chemical Society 27(1): 62–​69.
Gonzales, G. F., J. Tello, A. Zevallos-​Concha, L. Baquerizo, and L. Caballero, L. 2018.
Nitrogen balance after a single oral consumption of sacha inchi (Plukenetia volubilis
L.) protein compared to soy protein: A randomized study in humans. Toxicology
Mechanisms and Methods 28(2): 140–​147.
González-​Aspajo, G., H. Belkhelfa, L. Haddioui-​Hbabi, G. Bourdy, and E. Deharo. 2015.
Sacha inchi oil (Plukenetia volubilis L.), effect on adherence of Staphylococus aureus
to human skin explant and keratinocytes in vitro. Journal of Ethnopharmacology
171: 330–​334.
Gouveia, S. T., G. S. Lopes, O. Fatibello-​Filho, A. R. A. Nogueira, and J. A. Nóbrega. 2002.
Homogenization of breakfast cereals using cryogenic grinding. Journal of Food
Engineering 51: 59–​63.
Grundy, M. M. L., K. Lapsley, and P. R. Ellis. 2016. A review of the impact of processing
on nutrient bioaccessibility and digestion of almonds. International Journal of Food
Science & Technology 51: 1937–​1946.
Gularte, M. A., M. Gómez, and C. M. Rosell. 2011. Impact of legume flours on quality and in
vitro digestibility of starch and protein from gluten-​free cakes. Food and Bioprocess
Technology 5(8): 3142–​3150.
198 Latin-American Crops in Gluten-Free Applications

Gutiérrez-​Dorado, R., A. E. Ayala-​Rodríguez, J. Milán-​Carrillo, J. López-​Cervantes, J. A.


Garzón-​Tiznado, J. A. López-​Valenzuela, O. Paredes-​López, and C. Reyes-​Moreno,
2008. Technological and nutritional properties of flours and tortillas from
nixtamalized and extruded quality protein maize (Zea mays L.). Cereal Chemistry
85: 808–​816.
Han, Y., J. Chi, M. Zhang, R. Zhang, S. Fan, L. Dong, F. Huang, and L. Liu. 2019. Changes in
saponins, phenolics and antioxidant activity of quinoa (Chenopodium quinoa Willd)
during milling process. Food Science and Technology 114: 1–​7.
Hargreaves, S. M., R. P. Zandonadi, and R. P. Zandonadi. 2018. Flaxseed and chia seed
gel on characteristics of gluten-​free cake. Journal of Culinary Science & Technology
16(4): 378–​388.
Hemalatha, P., D. P. Bomzan, B. V. S. Rao, and Y. N. Sreerama. 2016. Distribution of phenolic
antioxidants in whole and milled fractions of quinoa and their inhibitory effects on
α-​amylase and α-​glucosidase activities. Food Chemistry 199: 330–​338.
Horstmann, S., M. Foschia, and E. Arendt. 2017. Correlation analysis of protein quality
characteristics with gluten-​free bread properties. Food and Function 8(7): 2465–​2474.
Huerta, K. d M., J. d S. Alves, A. F. Coelho da Silva, E. H. Kubota, and C. Severo da Rosa. 2016.
Sensory response and physical characteristics of gluten-​free and gum-​free bread
with chia flour. Food Science and Technology 36: 15–​18.
Iglesias-​puig, E. and Haros, M. 2013. Evaluation of performance of dough and bread
incorporating chia (Salvia hispanica L.). European Food Research and Technology
237: 865–​874.
Ikezu, U. J. M., I. P. Udeozo, and D. E. Egbe. 2015. Phytochemical and proximate analysis of
black turtle beans (Phaseolus vulgaris). African Journal of Basic & Applied Sciences
7(2): 88–​90.
Inglett, G. E., D. Chen, and S. X. Liu. 2015. Physical properties of gluten-​free sugar cookies
made from amaranth–​ oat composites. LWT –​ Food Science and Technology
63: 214–​220.
Ixtaína, V. Y., A. Vega, S. M. Nolasco et al. 2010. Supercritical carbon dioxide extraction of
oil from mexican chia seed (Salvia hispanica L.): characterization and process opti-
mization. The Journal of Supercritical Fluids 55(1): 192–199.
Jagelaviciute J., and D. Cizeikiene. 2021. The influence of non-​traditional sourdough made
with quinoa, hemp and chia flour on the characteristics of gluten-​free maize/​rice
bread. LWT –​ Food Science and Technology 137: 110457.
Jekle, M., A. Houben, M. Mitzscherling, and T. Becker. 2010. Effects of selected lactic acid
bacteria on the characteristics of amaranth sourdough. Journal of the Science of Food
and Agriculture 90: 2326–​2332.
Kaur, S. and K. Bains. 2019. Chia (Salvia hispanica L.) -​A rediscovered ancient grain, from
Aztecs to food laboratories. A review. Nutrition & Food Science 50(3): 463–​479.
Kim, D.-​S., and N. Joo. 2019. Nutritional composition of Sacha inchi (Plukenetia volubilis
L.) as affected by different cooking methods. International Journal of Food Properties
22(1): 1235–​1241.
Kljak, K., E. Šárka, P. Dostálek, P. Smrčková, and D. Grbeša. 2015. Influence of physico-
chemical properties of Croatian maize hybrids on quality of extrusion cooking. Food
Science and Technology 60(1): 472–​477.
Koehler, P., H. Wieser, and K. Konitzer. 2014. Gluten-​free products. In Celiac Disease and
Gluten. Multidisciplinary Challenges and Opportunities, chap. 4, 174–​214. London/​
Waltham, MA/​San Diego, CA: Academic Press/​Elsevier Science.
References 199

Kumar, S., V. K. Sharma, S. Yadav, and S. Dey. 2017. Antiproliferative and apoptotic effects
of black turtle bean extracts on human breast cancer cell line through extrinsic and
intrinsic pathway. Chemistry Central Journal 11(1): 56.
Laczkowski, M. S., M. R. Baqueta, V. M. A. T. de Oliveira, T. R.Goncalves, S. T. M. Gómes, P. H.
Marco, M. Matsushita, and P. Valderrama. 2021. Application of chemometric tools
in the development and sensory evaluation of gluten-​free cracknel biscuits with the
addition of chia seeds and turmeric powder. Journal of Food Science and Technology –​
Mysore DOI: 10.1007/​s13197-​020-​04874-​9.
Lee, M.-​H. 2019. Study on the quality characteristics and functional analysis of pound cakes
containing sacha inchi (Plukenetia volubilis L.) flour. Culinary Science & Hospitality
Research 25(8): 28–​40.
Leray, G., B. Oliete, S. Mezaize, S. Chevallier, and M. de Lamballerie. 2010. Effects of freezing
and frozen storage conditions on the rheological properties of different formulations
of non-​yeasted wheat and gluten-​free bread dough. Journal of Food Engineering
100: 70–​76.
Levent, H. 2017. Effect of partial substitution of gluten-​free flour mixtures with chia (Salvia
hispanica L.) flour on quality of gluten-​free noodles. Journal of Food Science and
Technology–​ Mysore 54(7): 1971–​1978.
Linares, L., R. Repo‑Carrasco‑Valencia, P. Glorio Paulet, and R. Schöenlechner. 2019.
Development of gluten‑free and egg‑free pasta based on quinoa (Chenopdium quinoa
Willd) with addition of lupine flour, vegetable proteins and the oxidizing enzyme
POx. European Food Research and Technology 245: 2147–​2156.
Liu, T., G. G. Hou, B. Lee, L. Marquart, and A. Dubat. 2016. Effects of particle size on the
quality attributes of reconstituted whole-​wheat flour and tortillas made from it.
Journal of Cereal Science 71: 145–​152.
Luna-​Mercado, G. I., and R. Repo-​ Carrasco-​ Valencia. 2020. Gluten-​ free bread
applications: Thermo-​ mechanical and techno-​ functional characterization of
Kañiwa flour. Cereal Chemistry 98(3): 474–​481.
Madruga, K., M. da Rocha, S. Santos Fernándes, M. de las M. Salas-​Mellado. 2020. Properties
of wheat and rice breads added with chia (Salvia hispanica L.) protein hydrolyzate.
Food Science and Technology 40: 596–​603.
Maidana, S. D., S. Finch, M. Garro, G. Savoy, M. Ganzle, and G. Vignolo. 2020a
Development of gluten-​free breads started with chia and flaxseed sourdoughs
fermented by selected lactic acid bacteria. LWT –​ Food Science and Technology 125:
109189.
Maidana, S. D., C. A. Ficoseco, D. Bassi, P. S. Cocconcelli, E. Puglisi, G. Savoy, G. Vignolo,
and C. Fontana. 2020b. Biodiversity and technological-​functional potential of lactic
acid bacteria isolated from spontaneously fermented chia sourdough. International
Journal of Food Microbiology 316: 108425.
Malka, T., R. Bomben, L. Balmaceda, J. Leporatti, T. Batlle, and S. Zaniolo. 2020. Gluten-​
free amaranth-​based sweet snack formulation. Avances en Ciencias e Ingenieria
11(2): 21–​29.
Medina, J. J. R., K. Ramírez, J. Rangel-​Peraza, and G. J. Aguayo-​Rojas. 2018. Incremento del
valor nutrimental, actividad antioxidante y potencial inhibitorio de α-​glucosida
en brownies a base de leguminosas cocidas. Archivos Latinoamericanos de
Nutrición 68(2).
200 Latin-American Crops in Gluten-Free Applications

Menga, V., M. Amato, T. D. Phillips, D. Angelino, F. Morreale, and C. Fares. 2017. Gluten-​free
pasta incorporating chia (Salvia hispanica L.) as thickening agent: An approach to
naturally improve the nutritional profile and the in vitro carbohydrate digestibility.
Food Chemistry 221: 1954–​1961.
Merayo, Y., R. González, S. Drago, R. Torres, and D. Greef. 2011. Extrusion conditions and
Zea mays endosperm hardness affecting gluten-​free spaghetti quality. International
Journal of Food Science & Technology 46(11): 2321–​2328.
Mercado, R. J. L., P. C. C. A. Elías, and C. G. L. Pascual. 2015. Obtención de un aislado proteico
de torta de sacha inchi (Plukenetia volubilis L.) y evaluación de sus propiedades
tecno-​funcionales. Anales Científicos 76(1): 160–​167.
Mesías, M., F. Holgado, G. Márquez-​Ruiz, and F. J. Morales. 2016. Risk/​benefit consider-
ations of a new formulation of wheat-​based biscuit supplemented with different
amounts of chia flour. LWT –​ Food Science and Technology 7: 528–​535.
Miranda-​ Ramos, K., M.C. Millán-​ Linares, and C.M. Haros. 2020. Effect of chia as
breadmaking ingredient on nutritional quality, mineral availability, and glycemic
index of bread. Foods 9(5): 663.https://​doi.org/​10.3390/​foods​9050​663.
Montemurro, M., E. Pontonio, and C. G. Rizzello. 2021. Design of a “Clean-​Label” gluten-​
free bread to meet consumer demand. Foods 10(2): 462.
Mufari, J. R., P. P. Miranda-​Villa, and E. L. Calandri. 2018. Quinoa germ and starch separation
by wet milling, performance and characterization of the fractions. Lebensmittel-​
Wissenschaft & Technologie 96: 527–​534.
OJEU (Official Journal of the European Union). 1997. Regulation (EC) Nº 258/​97 of the
European Parliament and of the Council of 27 January 1997, concerning novel foods
and novel food ingredients. Official Journal of the European Communities No L 43/​1,
14. 2. 97.
Padalino, L., M. Mastromatteo, L. Lecce, F. Cozzolino, and M. A. Nobile. 2013. Manufacture
and characterization of gluten-​free spaghetti enriched with vegetable flour. Journal
of Cereal Science 57: 333–​342.
Paesani C., A. Bravo-​Núñez, and M. Gómez. 2020. Effect of extrusion of whole-​grain maize
flour on the characteristics of gluten-​free cookies. Food Science and Technology
132: 109931.
Paredes López, O., F. Guevara Lara, and L. A. Bello Pérez. 2012. Los Alimentos Mágicos de las
Culturas Indígenas Mesoamericanas. México D.F:Fondo de Cultura Económica. ISBN
978-​968-​16-​7567-​7.
Pineli, L. de L. de O., R. B. A. Botelho, R. P. Zandonadi, J.L. J. L. Solorzano, G. T. de
Oliveira, C. E. G. Reis, and D. da S. Teixeira. 2015. Low glycemic index and increased
protein content in a novel quinoa milk. LWT –​Food Science and Technology
63: 1261–​1267.
Prego, I., S. Maldonado, and M. Otegui. 1998. Seed structure and localization of reserves in
Chenopodium quinoa. Annals of Botany 82(4): 481–​488.
Ramos-​Díaz, J.M., S. Kirjoranta, S. Tenitz, P. A. Penttilä, R. Serimaa, A.-​M. Lampi, and K.
Jouppila. 2013. Use of amaranth, quinoa and kañiwa in extruded corn-​based snacks.
Journal of Cereal Science 58: 59–​67.
Rawdkuen, S., D. Murdayanti, S. Ketnawa, and S. Phongthai. 2016. Chemical properties and
nutritional factors of pressed-​cake from tea and sacha inchi seeds. Food Bioscience
15: 64–​71.
References 201

Repo-​Carrasco-​Valencia, R. A. M., C. R. Encina, M. J. Binaghi, C. B. Greco, and P. A. Ronayne


de Ferrer. 2010. Effects of roasting and boiling of quinoa, kiwicha and kañiwa on
composition and availability of minerals in vitro. Journal of the Science of Food and
Agriculture 90: 2068–​2073.
Reyes-​Caudillo E., A. Tecante, M.A. Valdivia-​López. 2008. Dietary fibre content and antioxi-
dant activity of phenolic compounds present in Mexican chia (Salvia hispanica L.)
seeds. Food Chemistry 107(2): 656–​663.
Roa, D. F., P. R. Santagapita, M. P. Buera, and M. P. Tolaba. 2014. Amaranth milling strategies
and fraction characterization by FT-​IR. Food Bioprocess Technology 7: 711–​718.
Rocchetti, G., L. Lucini, J. M. L. Rodríguez, F. J. Barba, and G. Giuberti. 2019. Gluten-​free
flours from cereals, pseudocereals and legumes: Phenolic fingerprints and in vitro
antioxidant properties. Food Chemistry 271: 157–​164.
Romero-​ Espinoza, A. M., S. O. Serna-​ Saldívar, M. C. Vintimilla-​Álvarez, M. Briones-​
García, and M. A. Lazo-​Vélez. 2020. Effects of fermentation with probiotics on anti-​
nutritional factors and proximate composition of lupin (Lupinus mutabilis Sweet).
LWT –​ Food Science and Technology 130: 109658.
Rosell, C. M., G. Cortez, and R. Repo-​Carrasco. 2009. Breadmaking use of Andean crops
quinoa, kañiwa, kiwicha, and tarwi. Cereal Chemistry 86(4): 386–​392.
Ruiz, G.A., A. Arts, M. Minor, and M. Schutyser. 2016. A hybrid dry and aqueous fraction-
ation method to obtain protein-​rich fractions from quinoa (Chenopodium quinoa
Willd). Food Bioprocess Technology 9: 1502–​1510.
Sandhu, K. S., P. Godara, M. Kaur, and S. Punia. 2017. Effect of toasting on physical, func-
tional and antioxidant properties of flour from oat (Avena sativa L.) cultivars. Journal
of the Saudi Society of Agricultural Sciences 16: 197–​203.
Santiago-​Ramos, D., J. D. Figueroa-​Cárdenas, and J. J. Véles-​Medina. 2018. Viscoelastic
behaviour of masa from corn flours obtained by nixtamalization with different cal-
cium sources. Food Chemistry 248: 21–​28.
Schöenlechner, R. 2017a. Pseudocereals in gluten-​free products. In Pseudocereals: Chemistry
and Technology, 193–​216. Chichester, UK: John Wiley & Sons.
Schöenlechner, R. 2017b. Quinoa: Its unique nutritional and health-​promoting attributes.
In Gluten-​Free Ancient Grains. Cereals, Pseudocereals, and Legumes: Sustainable,
Nutritious, and Health-​Promoting Foods for the 21st Century, edited by J. R. N. Taylor
and J. M. Awika, 105–​130. Woodhead Publishing Series in Food Science, Technology
and Nutrition. Duxford, UK/​Cambridge, MA/​Kidlington, UK: Woodhead Publishing/​
Elsevier.
Schöenlechner, R., G. Linsberger, L. Kaczyk, and E. Berghofer. 2006. Production of gluten-​
free short dough biscuits from the pseudocereals amaranth, quinoa and buckwheat
with common bean. Ernaehrung/​Nutrition 30: 101–​107.
Schöenlechner, R., I. Mandala, A. Kiskini, A. Kostaropoulos, and A. Berghofer. 2010. Effect
of water, albumen and fat on the quality of gluten-​free bread containing amaranth.
International Journal of Food Science and Technology 45: 661–​669.
Segundo, C., A. Giménez, M. Lobo, L. Iturriaga, and N. Sammán. 2019. Formulation and
attributes of gluten-​free cakes of Andean corn improved with green banana flour.
Food Science and Technology International 26(2): 95–​104.
Serna-​Saldívar, S. O., and C. Chuck-​Hernández. 2019. Food uses of lime-​cooked corn with
emphasis in tortillas and snacks. In Corn: Chemistry and Technology, edited by S. O.
Serna-​Saldívar, 469–​500. Washington, DC: AACC International Press.
202 Latin-American Crops in Gluten-Free Applications

Shepherd, R., M. Dean, P. Lampila, A. Arvola, A. Saba, M. Vassallo, E. Claupein, M.


Winkelmann, and L. Lähteenmäki. 2012. Communicating the benefits of wholegrain
and functional grain products to European consumers. Trends in Food Science &
Technology 25: 63–​69.
Sindhuja, A., M. Sudha, and A. Rahim. 2005. Effect of incorporation of amaranth flour on
the quality of cookies. European Food Research and Technology 221: 597–​601.
Sozer, N., U. Holopainen-​Mantila, and K. Poutanen. 2017. Traditional and new food uses of
pulses. Cereal Chemistry Journal 94(1): 66–​73.
Steffolani, E., E. de la Hera, G. Pérez, and M. Gómez. 2014. Effect of chia (Salvia hispanica L)
addition on the quality of gluten-​free bread. Journal of Food Quality 37(5): 309–​317.
Štěrbová, L., P. Hlásná Čepková, I. Viehmannová, and D. C. Huansi. 2017. Effect of thermal
processing on phenolic content, tocopherols and antioxidant activity of sacha inchi
kernels. Journal of Food Processing and Preservation 41(2): e12848.
Sung, W. C., E. T. Chiu, A. Sun, and H. I. Hsiao. 2020. Incorporation of chia seed flour into
gluten-​free rice layer cake: Effects on nutritional quality and physicochemical prop-
erties. Journal of Food Science 85(3): 545–​555.
Suwanangul, S., N. Jittrepotch, and K. Ruttarattanamongko. 2021. Effects of thermal
treatments od physico-​chemical properties and antinutritional factor of sacha inchi
(Plukenetia volubilis L.). Naresuan University Journal: Science and Technology 29(3).
Tosi, E. A., E. D. Re, H. Lucero, and R. Masciarelli. 2000. Amaranth (Amaranthus spp.) grain
conditioning to obtain hyperproteic flour by differential milling. Food Science and
Technology International 6(6): 433–​438.
Turck, D., J. Castenmiller, S. de Henauw, K. Hirsch-​ Ernst, J. Kearney, A. Maciuk,
I. Mangelsdorf, H. J. McArdle, A. Naska, C. Peláez, K. Pentieva, A. Siani, F. Thies, S,
Tsabouri, M. Vinceti, F. Cubadda, K.-​H. Engel, T. Frenzel, M. Heinonen, R. Marchelli,
M. Neuhäuser-​Berthold, A. Pöting, M. Poulsen,Y. Sanz, J. R. Schlatter, H. van Loveren,
L. Matijević, and H. K. Knutsen. EFSA Panel on Nutrition, Novel Foods and Food
Allergens (NDA). 2019. Safety of chia seeds (Salvia hispanica L.) powders, as novel
foods, pursuant to Regulation (EU) 2015/​2283. EFSA Journal 17: e05716.
Vázquez-​Carrillo, M. G., J. D. Santiago-​Ramos, M. Gaytán-​Martínez, E. Morales-​Sánchez
and M. Guerrero-​Herrera. 2015. High oil content maize: Physical, thermal and
rheological properties of grain, masa, and tortillas. Food Science and Technology
60(1): 156–​161.
Vázquez-​Osorio, D. C., J. D. Jaramillo Ramírez, G. A. Hincapié Llanos, and L. M. Vélez
Acosta. 2017. Development of cookies with sacha inchi (Plukenetia volubilis L.) flour
coming from residual cake. UGCiencia 23: 101–​113.
Viamonte-​Garcés, M. I., J. M. Sánchez-​Campuzano, A. Ramírez-​Sánchez, A. Tapuy Cabrera,
and V. C. Andrade-​Yucailla. 2020. Chemical characterization and fatty acid profile of
sacha inchi flour (Plukenetia volubilis) as raw material, in the elaboration of diets for
animal use. Mol2Net 2020: 6. DOI:10.3390/​mol2net-​06-​xxxx.
Vidaurre-​Ruiz, J. M. 2020. Desarrollo de panes libres de gluten con harinas de granos
andinos. PhD Dissertation. Lima: Universidad Nacional Agraria La Molina.
Vidaurre-​Ruiz, J. M., W. F. Salas-​Valerio, and R. Repo-​Carrasco-​Valencia. 2019. Propiedades
de pasta y texturales de las mezclas de harinas de quinua (Chenopodium quinoa),
kiwicha (Amaranthus caudatus) y tarwi (Lupinus mutabilis) en un sistema acuoso.
Revista de Investigaciones Altoandinas 21(1): 5–​14.
References 203

Villacrés, E., M. Quelal, E. Fernández, G. García, G. Cueva, and C. Rosell. 2020. Impact
of debittering and fermentation processes on the antinutritional and antioxi-
dant compounds in Lupinus mutabilis Sweet. LWT –​ Food Science and Technology
131: 109745.
Villada, J. A., F. Sánchez-​ Sinencio, O. Zelaya-​ Ángel, E. Gutiérrez-​ Cortez, and M. E.
Rodríguez-​ García. 2017. Study of the morphological, structural, thermal, and
pasting corn transformation during the traditional nixtamalization process: From
corn to tortilla. Journal of Food Engineering 212: 242–​251.
Wang, J., J. Zhao, M. de Wit, R. M. Boom, and M. A. I. Schutyser. 2016. Lupine protein enrich-
ment by milling and electrostatic separation. Innovative Food Science and Emerging
Technologies 33: 596–​602.
Wang, S., F. Zhu, and Y. Kakuda. 2018. Sacha inchi (Plukenetia volubilis L.): Nutritional com-
position, biological activity, and uses. Food Chemistry 265: 316–​328.
Zamora, R., and F. J. Hidalgo. 2011. The Maillard reaction and lipid oxidation. Lipid
Technology 23: 59–​62.
Zamora, R., R. M. Delgado, and F. J. Hidalgo.2009. Conversion of 3-​aminopropionamide and
3-​alkylaminopropionamides into acrylamide in model systems. Molecular Nutrition
& Food Research 53(12): 1512–​1520.
Zdybel, B., R. Rozylo, and S. Sagan. 2019. Use of a waste product from the pressing of chia
seed oil in wheat and gluten-​free bread processing. Journal of Food Processing and
Preservation 43(8): e14002.
Zettel, V., and B. Hitzmann. 2018. Applications of chia (Salvia hispanica L.) in food products.
Trends in Food Science & Technology 80: 43–​50.
Ziemichod, A., M. Wojcik, and R. Rozylo. 2019. Ocimum tenuiflorum seeds and Salvia
hispanica seeds: Mineral and amino acid composition, physical properties, and use
in gluten-​free bread. CYTA-​Journal of Food 17(1): 804–​813.
Žilić, S., A. Serpen, G. Akıllıoğlu∥, V. Gökmen, and J. Vančetović. 2012. Phenolic compounds,
carotenoids, anthocyanins, and antioxidant capacity of colored maize (Zea mays L.)
kernels. Journal of Agricultural and Food Chemistry 60: 1224−1231.
Žilić, S., T. Tolgahan, J. Vančetović, and V. Gökmen. 2016. Effects of baking conditions and
dough formulations on phenolic compound stability, antioxidant capacity and color
of cookies made from anthocyanin-​rich corn flour. Lebensmittel-​Wissenschaft und-​
Technologie 65: 597–​603.
Ziobro, R., L. Juszczak, M. Witczak, and J. Korus. 2016. Non-​gluten proteins as struc-
ture forming agents in gluten free bread. Journal of Food Science and Technology
53(1): 571–​580.
Chapter 6

Fractionation of Seeds or Grains


Marianela Capitani1, Adriana Scilingo2, Edgardo Calandri3,
María Alejandra Giménez4, Marcela Lilian Martínez5,
Vanesa Ixtaína2, Nancy Chasquibol Silva6,
M. Carmen Pérez Camino7, Natalia Bassett4,
Víctor Delgado-​Soriano8, Ritva Ann-​Mari Repo-​Carrasco-​
Valencia8, Claudia M. Haros9, and Mabel Tomás2
1TECSE, Olavarría, Buenos Aires, Argentina –​

CONICET, CCT Tandil, Buenos Aires, Argentina


2Universidad Nacional de La Plata (UNLP),

La Plata, Buenos Aires, Argentina


3Instituto de Ciencias y Tecnología de los Alimentos -​ICTA.

Córdoba, Argentina/​Instituto de Ciencias y Tecnología de los


Alimentos Córdoba –​ICyTAC (CONICET), Córdoba, Argentina
4Universidad Nacional de Jujuy, San Salvador de Jujuy, Argentina
5Instituto Multidisciplinario de Biología Vegetal (IMBIV),

(CCT Córdoba) –​CONICET, Córdoba, Argentina


6Universidad de Lima, Peru
7Instituto de la Grasa (IG-​CSIC), Seville, Spain
8Universidad Nacional Agraria La Molina (UNALM), Lima, Peru
9Instituto de Agroquímica y Tecnología de

Alimentos (IATA-​CSIC), Valencia, Spain)

DOI: 10.1201/9781003088424-6 205


206 Fractionation of Seeds or Grains

CONTENTS
6.1 Quinoa and Kañiwa 208
6.1.1 Milling 208
6.1.1.1 Dry Milling 208
6.1.1.2 Wet Milling 208
6.1.2 Oil Extraction 208
6.1.2.1 Solvent Extraction 209
6.1.2.2 Cold Extraction 209
6.1.3 Protein and Fiber Isolation 209
6.1.3.1 Protein Isolation Methods 209
6.1.3.2 Fiber Extraction Methods 210
6.1.3.3 Techno-​functional Properties of Proteins 210
6.1.4 Physicochemical, Thermal and Rheological Properties of
Starch 211
6.2 Maize 216
6.2.1 Grain Characteristics 216
6.2.2 Starch and Lipid Extraction 216
6.2.3 Protein Fractions 217
6.2.4 Fiber 217
6.2.5 Techno-​Functional properties of Flours, Proteins,
Starch and Fiber 217
6.2.5.1 Physico-​Chemical, Thermal and Rheological
Properties of Maize Starch 218
6.3 Amaranth 219
6.3.1 Milling 219
6.3.2 Oil Extraction 220
6.3.3 Fiber Extractions 220
6.3.4 Protein Isolation 221
6.3.5 Techno-​Functional Properties of Amaranth Flour
and Proteins 221
6.3.6 Physico-​Chemical, Thermal and Rheological
Properties of Amaranth Starches 223
6.3.6.1 Modification of Amaranth Starches 224
6.4 Chia 224
6.4.1 Seed Conditioning and Oil Extraction 224
6.4.1.1 Conventional Processes: Cold Pressing and
Solvent Extraction 224
Fractionation of Seeds or Grains 207

6.4.1.2 Non-​Conventional Processes: Supercritical


Carbon Dioxide Extraction (CO2-​SE) 225
6.4.1.3 Residual Flours 228
6.4.2 Protein and Fiber Isolation 228
6.4.2.1 Protein Rich Fraction 228
6.4.2.2 Protein Isolates 229
6.4.3 Fiber Rich Fraction 229
6.4.3.1 Mucilage: Extraction and Characterization 229
6.4.4 Techno-​Functional Properties of Residual Flours, Proteins
and Soluble Fiber (Mucilage) 230
6.4.4.1 Water-​holding Capacity, Absorption and
Adsorption Capacity 230
6.4.4.2 Oil Holding, Absorption and Adsorption Capacities,
and Absorption Capacity of Organic Molecules 232
6.4.4.3 Emulsifying Properties, Foaming Capacity and
Foam Stability 232
6.5 Sacha Inchi 232
6.5.1 Decapsulation and Seed Conditioning 232
6.5.2 Oil Extraction Processes 233
6.5.2.1 Conventional Processes 233
6.5.2.2 Non-​conventional Processes 234
6.5.2.3 Residual Flour 234
6.5.3 Isolates and Protein Hydrolysates 234
6.5.4 Techno-​Functional Properties of Sacha Inchi Cake Flour 235
6.6 Tarwi 235
6.6.1 Alkaloid Extraction 235
6.6.2 Oil Extraction 236
6.6.2.1 Characterization of Residual Flour 236
6.7 Black Turtle Bean 238
6.7.1 Milling Process 238
6.7.2 Protein Concentrates and Isolates 240
6.7.3 Oil Extraction 241
6.7.4 Techno-​Functional Properties of Flours, Protein
Concentrate, Oil and Fiber 242
6.8 General Conclusions 242
Acknowledgments 243
References 243
208 Fractionation of Seeds or Grains

6.1 QUINOA AND KAÑIWA


6.1.1 Milling
6.1.1.1 Dry Milling
Quinoa white flour was obtained at laboratory scale by roller milling; authors
found enhanced bread quality, due to the reduction in bran content and the
parallel increase in diastasic capacity of quinoa flour (Elgeti et al. 2015). The
process, essentially the same as that (Chirinos et al. 2018) used in the wheat
flour industry, opens up the possibility of scaling up production of quinoa white
flour without the need of new machinery. In another study D’Amico et al. (2019)
tried the abrasive milling of quinoa; the protein, fat and ash of the polished
seeds decreased with processing time, while starch rose up strongly. At 8 min
milling, 65% of seed mass is lost and with it, most of its nutritional quality. As
abrasion was the main procedure used for saponin removal, the results should
inspire a radical evaluation of nutritional loss during quinoa industrial pro-
cessing. Kañiwa (Chenopodium pallidicaule Aellen) belongs to the same genus
as quinoa, but the grain is smaller. A traditional way to ground kañiwa is the
kcona milling, using stones as grinding device. Flour can be obtained by disc
or hammer milling, from whole or toasted grains; this last option is known as
Cañihuaco.

6.1.1.2 Wet Milling
Together with the well-​known qualities of quinoa protein (Abugoch James 2009)
its starch shows interesting functional properties (Li et al. 2016). Quinoa oil has
also shown valuable characteristics (Mufari et al. 2020). While protein and oil
are mainly found in the embryo, starch granules reside mostly in the perisperm
(Prego et al. 1998). When wet enough, the embryo may easily be detached from
the perisperm; this enables separation of the components. We showed that
this is possible when seed humidity and roller milling conditions are carefully
controlled (Mufari et al. 2018a). In a similar manner, Ballester-​Sánchez (2019)
obtained good yield and quality in quinoa starch separation, employing a plate
mill. Dry milling and sieving of quinoa grain also yields enriched fractions, but it
is not as effective as wet milling (Solaesa et al. 2020).

6.1.2 Oil Extraction
Quinoa and kañiwa are not considered oily grains; their content is around 6–​8%
(Bazile et al. 2013; Torrejón et al. 2016), so at the moment there is no industrial
interest in it. Nevertheless, quinoa oil quality is well documented (Mufari et al.
2020) and future interest in it as a profitable by-​product must not be discarded;
6.1 Quinoa and Kañiwa 209

also kañiwa shows a good fatty acids profile (Torrejón et al. 2016). Traditionally,
quinoa oil extraction consisted in boiling seeds in iron pots covered with soap-
stone lids. The oil that condensed against the lids was then collected; it was
sold mainly in Baltimore (USA), hence it is known as “Baltimore oil” (Thoufeek
Ahamed et al. 1998).

6.1.2.1 Solvent Extraction
This is the most common way to remove oil from different natural raw materials.
Usually, non-​polar petroleum derivatives are used, such as hexane or pet-
roleum ether. Small scale processes involve soxhlet type devices (AOAC 1999).
For industrial production, the whole flour or grain fractions could be extracted
in a continuous process, similar to those employed in obtaining the oil from
cereals (Willis and Marangoni 2017). In all cases, the raw material must be
finely grounded, due to the extraction efficiency, which strongly depends on the
particles’ size. Supercritical extraction was also assayed for quinoa; the extracted
oil showed an elevated presence of antioxidants and tocopherols but the yields
were low (Benito-​Román et al. 2018).

6.1.2.2 Cold Extraction
According to Eckey (1954) it is possible to extract quinoa oil with screw mills
or twin screw extruders; nevertheless these processes are designed to maxi-
mize oil yield, so the high temperatures and pressure involved severely affect
the nature of both proteins and oil. Today the solvent extraction method is more
recommendable (Willis and Marangoni 2017). As we have shown before (Mufari
et al. 2018a), it is possible to separate an oil enriched germ fraction from quinoa
seeds, so that mild alternatives like those developed for cereal germs could also
be applied to quinoa germ oil extraction enabling the recovery of a good quality
protein (Stamenkovic et al. 2020). The process essentially involves a solid-​liquid
exchange where the oil moves from the germ to the continuous phase (solvent),
on a countercurrent basis.

6.1.3 Protein and Fiber Isolation


6.1.3.1 Protein Isolation Methods
Quinoa is a well-​valued food mainly due to the protein’s nutritional quality;
because of that it has been largely studied and characterized by several authors
all round the world (Abugoch et al. 2008). The most common method for
obtaining quinoa protein isolates is isoelectric precipitation, taking advantage
of the fact that most proteins are insoluble at the isoelectric point (pI) (Mir et al.
2021). It requires two steps: solubilization at a proper pH, normally far from the
210 Fractionation of Seeds or Grains

pI value, and second, that the pH is gradually changed until it approaches the pI;
at this moment the proteins get the lowest solubility and most of them aggre-
gate and precipitate. Guerreo-​Ochoa et al. (2015) found that optimal quinoa
protein extraction conditions depend not only on pH but also on temperature,
solvent/​meal relationship and extraction time. On the other hand, Ruiz et al.
(2016) found a correlation between isolated protein yield and the extraction pH.
Kañiwa proteins concentrate was also obtained with a procedure that resembles
the one used in quinoa (Betalleluz-​Pallardel et al. 2017). Quinoa and kañiwa
being both from Chenopodium genus it is not surprising that they show similar
composition in their protein fractions, as can be seen in Table 6.1.

6.1.3.2 Fiber Extraction Methods


Quinoa seeds present an interesting content in fibers, around 3–​4% (d.b.) by
weight (Jancurová et al. 2009). In kañiwa, fiber content is widely variable but
generally superior to quinoa (Repo-​Carrasco-​Valencia et al. 2010). Cereals fibers
are in the outer shell, but in pseudo-​cereals such as quinoa and kañiwa the shell
is very thin; when an abrasive method is applied this narrow cover is immedi-
ately removed; if the action persists other components of the seed will become
detached and nutritional quality can be affected (D’Amico et al. 2019). Ballester-​
Sánchez et al. (2019) found that wet milling of red quinoa produced a higher
yield, recovery and purity than the dry milling procedure, but this last one gave
breads with better technological qualities, attributed to the characteristics of
its fibers. Quinoa soluble fiber showed a compact structure, higher reducing
power and digestive enzyme inhibitory activity (Chen et al. 2021) in comparison
with wheat.

6.1.3.3 Techno-​functional Properties of Proteins


Mostly related with properties other than nutritional, techno-​ functional
examples involve the capacity to modify the texture or sensory properties of food.

TABLE 6.1 PROTEIN COMPOSITION IN QUINOA AND KAÑIWA


Protein fraction kañiwa (% d.b.)a quinoa (% d.b.)b
Albumin 15.4–​15.8 24.9 +​8.2
Globulin 24.1–​26.7 23.5 +​5.6
Prolamin 9.6–​9.9 5.1 +​0.4
Glutelin 22.9–​21.5 25.3 +​2.1
Insoluble fraction 28.0–​26.1 21.2 +​13.0
Source: aMoscoso-​Mujica et al. (2017); b .Mufari et al. (2018b).
Notes: d.b. dry basis.
6.1 Quinoa and Kañiwa 211

Also, they can influence the physical behavior of food or ingredients during their
processing or storing. Hydration properties are related to the capacity of proteins
to interact with a solvent, particularly water; protein–​protein interaction mainly
affects food structures and surface properties which are closely related to their
ability to stabilize bubbles or emulsions. In this respect, emulsifying properties
mainly depend on the ability of proteins to lower surface tension, and prevent
coalescence of droplets and also to increase surface hydrophobicity. Quinoa
protein isolates present a low emulsifying ability compared to bovine serum
albumin (BSA) but they show good emulsion stability, close to BSA and higher
than pearl millet, wheat and soybean. Enzymatic hydrolysis, saponins removal
and protein-​linked polysaccharides can modify these properties (Dakhili et al.
2019). Foaming capacity for quinoa proteins stay between soy protein and egg
albumin and makes stable foam that could be used in food prototypes (Abugoch
et al. 2008). The gelling properties of quinoa proteins depend on the pH of the
extraction. If it is lower than 9 a semi-​solid gel is obtained. When extraction pH
overtakes 10 the gel is not formed; and according to Hermansson (1986), it could
be used in liquid food. Quinoa proteins showed a water absorption capacity
close to that of soy protein, but the oil absorption capacity was lower (Ashraf
et al. 2012). Probably, the difference between oil and water absorption can reside
in the conformational characteristics of quinoa albumins and globulins (Sathe
and Salunkhe 1981). Quinoa proteins and soy proteins showed similar water
holding capacities (Abugoch et al. 2008); additionally, this property of quinoa
proteins is expected to confer improved texture on different food applications
(Tang et al. 2015). It must be pointed out that the functional properties of quinoa
protein can be affected by drying conditions (Mufari 2017). High temperatures
may definitively affect functional properties of quinoa protein, as shown by Mir
et al. (2021), who suggest that milder treatments would yield better results.
Finally, interaction with other components in food can modify the quinoa protein
behavior, as was observed by Roa-​Acosta et al. (2020). These authors tried
with quinoa flour which had received different abrasive milling treatments that
changed protein conformation, thus affecting both pasting and water absorption
behaviors.

6.1.4 Physicochemical, Thermal and Rheological


Properties of Starch
The starch granules of quinoa are polygonal/​irregular, of very small diameter
(0.46–​2.53 µm) and which have considerable variability in their amylose content
(3.5–​26.5%) (Table 6.2). In general, the high amount of amylopectin conducted
its gelatinization at relatively lower temperatures (44.5–​77.2°C) and exhibited a
high pasting viscosity (4150.8 Cp) compared to normal maize starch (61.7–​89.0°C
and 3058.0 Cp, respectively). Due to these large differences in amylose content
212 Fractionation of Seeds or Grains

and physicochemical properties among quinoa starches, they have a wide var-
iety of food and non-​food applications such as in cosmetics, in biodegradable
films and coatings, candy dusting, flavor carriers, in fat replacement and to
produce creamy/​smooth textures (Ahamed et al. 1996; Repo-​Carrasco-​Valencia
and Valdez Arana 2017). Furthermore, the good freeze–​thaw stability of quinoa
starch suggests application as a thickener in frozen food products; it shows
resistance to retrogradation, which also suggests possible uses in emulsions
such as salad dressings, as well as in sauces, cream soups and pie fillings (Repo-​
Carrasco-​Valencia and Valdez Arana 2017).
Kañiwa starches showed a similar amylose content to quinoa’s (6.48–​17.44%),
both have lower contents than cereals such as maize (22.0–​25.4%) or wheat (25–​
28%) starches (Table 6.2; JianYa-​Qian 1999; Jian et al. 2020; Lindeboom 2005).
Their granule size is very small with an average of around 1.45 µm. Additionally,
their granule shape is irregular and polygonal with similar onset and peak gel-
atinization temperatures compared to quinoa starches, but presented higher
values of firmness and lower syneresis than quinoa starches, which could be used
in foods where consistency and firmness are required (Steffolani et al. 2013).
The high viscosity value obtained during the heating process of quinoa
starches suggests a higher water absorption capacity, whereas the breakdown
parameter (PV–​HPV, where peak viscosity [PV] and hot paste viscosity [HPV])
provide information about stability during heating. This parameter is lower in
quinoa, kañiwa and amaranth starches compared to conventional maize starches
(Table 6.2), displaying a higher structural fragility during cooking than the last
item in the table (Haros et al. 2006). During cooling, an important parameter to
consider is the setback, which is the tendency to restructuration/​retrogradation
(Haros et al. 2006). High setback viscosities indicate low resistance to retrograd-
ation mainly attributed to amylose, with a larger flexible structure than amylo-
pectin. However, the lack of differences in setback values between maize and
pseudo-​cereals starches, between samples with different amounts of amylose,
indicates that another phenomenon should be implicated as was suggested by
Repo-​Carrasco-​Valencia and Valdez Arana (2017).
The gelatinization enthalpy (ΔHG) varied from 1.66 to 14.9 J/​g for quinoa
starch and from 1.84 to 9.32 J/​g for kañiwa starch, whereas the maize starch
showed values between 10 and 19.0 J/​g (with the exception of INIFAP-​Mexico
hybrids), demonstrating that, in general, higher energy was required to disrupt
the crystalline structure of the last item in the table (maize starch) (Table 6.2).
Low values in starch gelatinization and/​or pasting processes of some varieties of
quinoa and kañiwa might suggest a less crystalline structure, which could result
in higher enzymatic/​digestion susceptibility (Srichuwong et al. 2017). The rela-
tionship between the physicochemical and structural properties of both quinoa
and kañiwa starches, is as yet uncertain (Jiang et al. 2020).
newgenrtpdf
TABLE 6.2 PHYSICOCHEMICAL AND TECHNO-​FUNCTIONAL PROPERTIES OF NATIVE STARCHES
Physicochemical Latin-​American
characteristics Units Maize Quinoa Kañiwa Amaranth
Amylose (%) 20 ; 22.0
(x) ;
(h)(w) 3.5–​19.6 ; 7 ; 7.1 ;
(e) (h) (a) 6.48 ;
(l) 0–​14(q); 0.1–​19.2(r); 0.1–​22.3(t);
22.58(i); 25.4®; 27(y) 8.22–​9.30( f); 9.43–​ 10.70–​17.44( f) 2.5(c); 7.8(b); 13.6(s)
10.90(i); 12.1(j); 12.2(b);
19.2–​26.5(g); 22.5(c)
Granule Size (µm) 2.4–​33.8(n)(p); 3-​ 0.46–​5.56(i); 0.5–​3.0(d); 1.45/​< 2.0( f) 0.75–​1.25(c);1–​2(b)(d); 0.5-​2(j)
20(w)(x)(y); 5–​20(j); 0.6–​1.5(c); 1(a); 1–​2(b);
3.16–​ 30.22(i); 2.53( f)
6-​30˃(u)
Cristallinity (%) 22(y); 26(x); 32(w); 21.00–​29.67(i); 35.0(a); 34.0–​35.6( f) 24.5–​45.5®(r); 27.9(q)
36.05(i) 35.4(b); 36.3–​39.6( f)
Shape –​ Irregular/​ Polygonal(b); Polygonal(l); Polygo®(b)(q)(r); Spherical(c)
Polyhedric(j); Spherical(c); Irregular/​ Irregular/​ Irregular/​ Polygonal(j)
Irregular(u); Pol®nal(e)( f)(i) Polygonal( f)
Spherical(w)(x)(y)
Pasting Properties (RVA)

6.1
Pasting (ºC) 71.1–​72.6(p); 75.5(h); 62.7( f); 63.0–​64.0(e); 71.1–​74.®) 70–​75.7(r); 71.7(b); 73–​78(q)
Temperature 75.93(i); 91.0(e) 64.50–​68.44(g); 66.5(h);

Quinoa and Kañiwa


66.8(b); 72.6(i)
Peak time (min) 4.47–​4.73(® 5.1(h); 4.97–​5.17( f); 5.1–​6.9(e); 4.67–​5.43( f) N/​A
5.4(e) 6.72 –​7.00(g); 7.0(h)
(continued)

213
newgenrtpdf
214
Fractionation of Seeds or Grains
TABLE 6.2 (CONTINUED)
Physicochemical Latin-​American
characteristics Units Maize Quinoa Kañiwa Amaranth
Peak Viscosity (Cp) 1515–​3058 ; (p) 43.2–​309.6 ; 2983–​
(e) 237.0–​498.0 ;(k)* 216–​2064(q); 276–​3456(r);
(PV) 367.2(e); 2906(h); 3551(i); 3322–​4306( f); 1907–​2657( f) 1596–​2556(t); 1662.0(b)
2669(i) 3064–​3898 (g); 3380(h);
4150.8(b)
Hot Paste (Cp) 310–​1196(p); 1794(i); 2231–​3430(g); N/​A 792.0(b)
Viscosity (HPV) 1852(h) 2250–​3205(i); 2653(h);
3253.2(b)
Breakdown (Cp) 72.0(e); 295.2(e); 4.8–​1056.0(e); 34.8–​ 5.04–​28.0(k)*; 108–​1020(q); 120–​1572(r); 324–​
(PV – HPV) 875(i); 1052(h); 214.8(e); 313–​733(i); 234–​673( f) 696(s); 870.5(b)
1205–​1882(p) 897.6(b); 439–​1161( f);
728(h)
Final viscosity (Cp) 345.6(e); 491–​ 50.4–​346.8(e); 2692–​ 269.04–​810.00(k)*; 1296.0(b)
(FV) 2133(p); 2966(i); 3705(i); 3725–​4109( f); 2787–​3332( f)
3089(h) 3557–​4340 (g); 3867(h);
4994.4(b)
Setback (Cp) 64.8(e); 181–​957(p); 15.6–​153.6(e); 442–​ 60.0-​316.8(k)*; 0–​588(q); 84–​2292(r); 368.4–​
(PV –​TV) 1172(i); 1237(h) 730(i); 676–​963( f); 1041–​1332( f) 1404(s); 504.0(b)
1215(h); 1741.2(b)
Thermal properties (DSC conditions: starch:water, 1:3 at 10ºC/​min)
Onset (ºC) 61.7(e); 64.7(e); 44.6–50.6(e); 50.7(h); 53.6(l); 62–66.1(q); 62.3–69.3(r); 66.3(b);
Temperature (To) 64.9(h); 65.8(x); 51.2–55.4(g); 55.3–57.1(k)*; 66.8–68.0(t); 69.5(s)
65.9(w); 67.41(i); 54.25–55.72( f); 54.5(a); 58.3–58.9(m)*;
68.8–72.0(n); 57.89–61.76(i); 59.9(b) 59.06–58.39( f)
69.1–70.4(0)
newgenrtpdf
Peak Temperature (ºC) 69.1(e); 69.6(h); 50.5–​61.7(e); 58.6–​ 60.0(l); 68.8–​73.9(q); 69.2–​78.0(r);
(Tp) 71.00(i); 72.8(w); 61.1(g); 58.7(h): 61.66–​ 61.0–​61.8(k)*; 73.4(s); 73.4–​78.8(t); 74.5(b)
73.1(x); 73.2–​75.6(o); 63.01( f); 62.6(a); 62,8–​64.8(m)*;
73.3(y); 73.4-​77.1(v); 63.77–​67.44(i); 64.5(b) 66.18–​66.68( f)
73.6–​76.2(n);
Final Temperature (ºC) 76.0(h); 78.47(i); 68.2–​70.1(g); 69.3(h); 68.1(l); 77.7–​89.6(r); 78.8–​88.8(q); 81.2–​
(Tf ) 79.7(w); 79.9(x); 71.0(b); 71.3(a); 68.8–​69.0(k)*; 89.1(t); 86.9(b); 89(s)
80.2(y); 82.0–​84.3(o); 71.86–​77.24(i) 70.9–​72.9(m)*
85.1–​89.0(n)
Gelatinization (J/​g ) 1.9-​4.7(v); 10.0–​ 1.66(b); 7.79–​11.76(i); 1.84–​2.57(k)*; 3.5–​ 2.58(b); 7.1–​18.4(r); 13–​15.8(q);
Enthalphy(ΔHG) 13.7(o); 10.5–​13.9(n); 8.45–​ 9.00( f); 10.3(® 4.8(m)*; 7.49–​9.32( f); 13.5(s)
12.5(h); 13.1(e); 10.4(h); 12.4–​13.5(g); 8.00(l)
13.30(i); 18.5(y); 12.8–​14.9(e)
18.9(w); 19.0(x)
Notes: N/​A, not available; *flour; RVA: Rapid Visco-​Analyzer; DSC: Differential Scanning Calorimetry; Cp: centipoise.
Source: (a)Tang et al. (2002); (b)Qian and Kuhn (1999), Amaranthus cruentus; (c) Tari et al. (2003), Amaranthus paniculatus; (d)Lindeboom
et al. (2005), (e)2005); ( f)Steffolani et al. (2013); (g)Ballester-​Sanchez et al. (2019); (h)Selma-​Gracia et al. (2020); (i)Jian et al. (2020);
(j)Jane et al. (1994); (k)Bustos et al. (2019); (l)Luna-​Mercado and Repo-​Carrasco-​Valencia (2020); (m)Salas-​Valero et al. (2015);
(n)Pérez et al. (2003), semi-​dent maize, Zea mays c.v. Tilcara; (o)Haros et al. (2003), flint and dent maize; (p)Haros et al. (2004),

semi-​dent maize, Zea mays c.v. Tilcara; Zhu (2017), (q)A. hypochondriacus, (r)A. cruentus, (s)A. caudatus, (t)A. hybridus; (u)Méndez-​

6.1
Montealvo et al. (2008) INIFAP-​Mexico (H515); (v)Méndez-​Montealvo et al. (2005) INIFAP-​Mexico; Agama-​Acevedo et al.
(2005) pigmented maize (w)black, (x)blue and (z)white from Mexico.

Quinoa and Kañiwa


215
216 Fractionation of Seeds or Grains

6.2 MAIZE
6.2.1 Grain Characteristics
The word “corn” has been used over time with different meanings; however,
after the discovery of America, Europeans called it “Indian corn” and named
it “mahis” in the native American language, which was later called “maize”
(Salvador-​Reyes and Pedrosa Silva Clerici 2020). In this chapter, the name
maize will be used to indicate that it refers to the native races with production
aimed at local markets.
Vitreousness-​and hardness-​associated properties are significantly correlated
with the end usage of maize. The ratio of vitreous or hard to floury endosperm
influences post-​harvest resistance to insects and fungi, milling yield, dust for-
mation during processing, rate of starch digestibility and textural properties of
maize-​based foods (Dombrink-​Kurtzman and Bietz 1993; Córdova-​Noboa et al.
2020; Osorio-​Diaz et al. 2011; Cruz-​Vázquez et al. 2019).
Currently there is great interest in the extraction of colorants from pigmented
races (Jing and Giusti 2007). Purple and blue corn, characterized by high con-
tent of anthocyanin, are a good source for production of natural colorants for
the food industry (Salinas et al. 1999). For a long time, its extracts were used as
coloring agents for food and beverages, such as chicha Morada and mazamorra
Morada (Saldaña et al. 2018). Most blue varieties of maize are planted in the high-
lands of Mexico and Peru and are still landraces inherited by past generations
of local types. Major drawbacks of blue maize varieties are that plants yield
lower amounts of grain, and that kernels are soft-​textured. Due to the increased
interest in colored maize, plant breeders have developed high-​yielding hybrids
adapted to other ecosystems such as subtropical regions. These hybrids with
excellent yield potential produce harder kernels with higher concentrations
of anthocyanins (Serna-​ Saldivar and Pérez-​ Carrillo 2019). Recently, several
authors studied different techniques to obtain concentrated extracts of phenolic
compounds (Monroy et al. 2020). However, there is little research on the milling
properties of these maize varieties, so it is necessary to verify their suitability for
conventional maize processing.

6.2.2 Starch and Lipid Extraction


The wet milling process is designed to efficiently take maize apart and purify its
constituents (starch, oil, protein, and fiber), making them suitable for human
and animal food ingredients use, industrial products, or as feedstocks for
converting into other products with higher added value (Anderson and Watson
1982). Conventional wet milling is generally applied to hybrid maize as they have
a high extractability of starch and oil.
6.2 Maize 217

Despite the great diversity of native maize in Mesoamerica and South America,
little is known about their wet milling characteristics, so the separation of starch
and oil from them is generally not practiced. In this regard, Uriarte-​Aceves et al.
(2015) showed that despite the wide variability found in the physical and chem-
ical characteristics of different blue maize from northeast Mexico, they present
great potential for wet milling with high yields and starch recovery and with an
exceptional quality of whiteness.
Triglycerides from maize oil have a high level of linoleic acid (c.18:2), which
is an essential polyunsaturated fatty acid for humans (Weber 1979). According
to Giménez et al. (2020) and Ortíz-​Prudencio (2006) different races of native
maize from Mesoamerica and South America showed a kernel oil concentration
of about 4–​5% on dry weight basis. Development of maize cultivars with high oil
without yield loss is a challenge in breeding programs (Rajendran et al. 2018).
Specialty high oil maize contains 6% or more oil on dry weight basis.

6.2.3 Protein Fractions
The protein content in Latin-​American maize races ranges from 5.7–​12.5% of
their dry weight (Giménez et al. 2020). Zeins are the most abundant in the grain;
they are located in the endosperm. They have very low content of lysine and spe-
cially of tryptophan, which is why zeins are considered of very poor nutritional
quality (Zarkadas et al. 1995; Rascón-​Cruz et al., 2004).

6.2.4 Fiber
The most important sources of dietary fiber in maize are the bran and the tip tap
(Boyer and Shannon 2003). The dietary fiber content in native maize races varies
between 8.47 and 19.5% (Ortíz-​Prudencio 2006). The high dietary fiber content
of some Peruvian Andean maize can be related to the greater thickness and
number of their pericarp layers compared with other maize cultivars (Salvador-​
Reyes et al. 2021). It is generally separated during milling in the shell and used for
animal feed; however, it could be used in other formulated products.

6.2.5 Techno-​Functional properties of Flours,


Proteins, Starch and Fiber
Great variability in techno-​functional properties has been observed in Andean
and Mesoamerican maize races. Knowledge of the starch gelatinization pro-
cess provides important information about functionality, energy requirements
and end use. Some maize races from the north of Argentina were differentiated
mainly due to the amylose content, thermal properties and firmness, and stability
218 Fractionation of Seeds or Grains

of their gels. In these races there were no significant differences in enthalpy of


starch gelatinization; however, the floury kernels showed a lower gelatinization
range than races with the highest endosperm hardness (Giménez et al. 2020).
According to Narváez-​González et al. (2007), the size of the starch granule and
endosperm hardness greatly influenced the thermal and pasting properties
of the flours of different races of Latin-​American maize. The hard or vitreous
grains presented small starch granules, while in the floury grains they were large.
Kernels with the highest endosperm hardness tended to develop low viscosities,
and enthalpies presented a long gelatinization range. The opposite was found
for floury kernels. According to Salvador-​Reyes et al. (2021) the differences in
the chemical composition and endosperm structure between Peruvian Andean
maize samples influenced their pasting properties. The floury races had a lower
pasting temperature, high viscosity pastes and did not retrograde easily, making
it an excellent option for improving the texture of soups, sauces, food mixes or
ice creams where it is required to maintain the texture and avoid syneresis.
According to the broad techno-​functional behaviors, the Mesoamerican
and Andean maize races could be used in different technological processes to
produce gluten-​free foods. Their proteins, mainly zeins, provide a cohesive and
extensible mass above its glass transition temperature, approximately 35°C,
with moisture content greater than 20%, so they can be a suitable ingredient for
this type of products (Taylor et al. 2018). Fiber of hybrid maize has been used
to produce functional soluble dietary fiber. The product, known as corn fiber
gum, is primarily composed of arabinoxylan and possesses a strong emulsifying
property (Ai and Jane 2016). The majority of maize bran is almost completely
insoluble dietary fiber (Boyer and Shannon 2003). Unfortunately, this can lead to
undesirable changes in product quality (Rose et al., 2010).

6.2.5.1 Physico-​Chemical, Thermal and Rheological


Properties of Maize Starch
Unfortunately, little information was found with regard to Latin-​American maize
starches. The parameters described in Table 6.2 correspond to native maize
from Argentina (types flint, semi-​dent and dent), pigmented maize from Mexico
(white, black and blue) or varieties/​hybrids of maize (INIFAP [National Institute
of Agricultural and Livestock Forestry Research]), Mexico).
Salvador-​Reyes et al. (2021) have shown that some maize races from Peru
presented starch granules with spherical shape at the center of their endo-
sperm, while in other irregular and polyhedral shapes, most of the races studied
by these authors presented only large starch granules (>10 µm), while a bimodal
distribution presenting large granules (>10 μm) and small granules (<10 μm) was
also observed. The presence of small granules could be related to the extreme
conditions of growth (Tester 1997). In these races, the starch granules presented
6.3 Amaranth 219

both smooth and porous surfaces. Although how pore formation occurs is still
unknown; it has been shown that pore formation is directly related to the α-​
amylase activity of grain (Prompiputtanapon et al. 2020). The porous starches
influence the mechanical and textural properties of food products (Sarifudin
et al. 2020).

6.3 AMARANTH
6.3.1 Milling
Amaranth flour can be obtained by grinding the whole grain or by using strategies
that enrich the product in any of its macro-​components, starch, fiber, proteins
and/​or lipids (Roa et al. 2013; Roa-​Acosta et al. 2017; Tosi et al. 2000). The grains
directly ground in a planetary ball mill provided flour with 16.8, 7.7 and 73%
weight for weight (w/​w) for proteins (Nx6.25), lipids and carbohydrates (starch+​
fiber) respectively (Roa et al. 2013). When an abrasive mill step prior to grinding
was performed, two different products were obtained: one rich in starch and
another rich in lipids and proteins (1.5, 5 and 88% w/​w and 23, 47 and 10% w/​w
for lipids, proteins and carbohydrates, respectively). The products also differ in
lipid oxidative stability during storage time, and in the starch’s degree of crystal-
linity, which could affect its technological properties (Roa et al. 2014; Roa-​Acosta
et al. 2017). A successful differential milling technique to obtain Amaranthus
caudatus high protein flours was described by Tosi et al. (2000): intense friction
of specially conditioned dried grains in fluidized bed equipment was performed
to maintain moisture within the seed and reduce it on the surface. Differential
milling allowed a selective detachment of the different anatomical parts of the
grain and minimized the loss of lysine during drying. After sieving and carrying
out a pneumatic classification high protein and high starch flour was obtained.
The protein percentage of this product was much higher than that of whole flour
(near to 40% vs. 21% w/​w respectively) with a yield close to 40%, while the differ-
ential grinding of unconditioned grains exhibited a lower yield (10%).
Amaranth wet grinding has been used as a way of obtaining starch. Seeds must
be soaking in order to soften the grains. This phase, which is critical in the overall
process, can be carried out in an aqueous medium (Pérez et al. 1993), in the
presence of alkalis (Vázquez Batti et al. 2006), acids (Calzetta-​Resio et al. 2006) or
even detergents such as SDS (sodium dodecyl sulfate; Joaqui et al. 2020). Process
conditions, use of additives and their concentration, soaking time, temperature,
separation conditions, among others, have been studied in several publications
(Pérez et al. 1993; Calzetta-​Resio et al. 2006; Vázquez-​Batti et al. 2006). Fractions
rich in proteins and fiber can be separated out as by-​products during starch pro-
duction. Their characteristics and techno-​functional properties strongly depend
220 Fractionation of Seeds or Grains

on the process conditions. It is possible to adjust the conditions to minimize the


presence of proteins in the starch, or in the fraction rich in fiber depending on
what it is desired (Calzetta-​Resio et al. 2006). Joaqui et al. (2020) have recently
studied the structural modifications of seed proteins when the milling process
is carried out in the presence of sodium hydroxide (NaOH) and SDS using the
Fourier-​transform infrared spectroscopy (FT-​IR). Important changes occur in
the gelling capacity associated with modifications in the proteins’ secondary
and tertiary structure.
An alternative way implies using wet milling of the grains to prepare disper-
sion instead of obtaining starch. The grain in water grinding using a colloidal
mill (model AD 35-​R-​ColMil) has allowed the development of a beverage amar-
anth like vegetable milk with a protein content similar to cowx’ milk (Guzmán-​
Maldonado and Paredes-​Lopez 1998; Manassero et al. 2020).

6.3.2 Oil Extraction
Amaranth lipids in the seed represent between 6 and 8.5% w/​w of its weight
with some variations among the different species (León-​Camacho et al. 2001;
Martirosyan et al. 2007; Gamel et al. 2007). Triacylglycerols contribute approxi-
mately 80% w/​w, with various unsaturated and saturated fatty acids in a 3:1
ratio; linoleic acid is the most important fatty acid, followed by oleic and pal-
mitic acids (Krist 2020). Phospholipids and squalene represent 10% and 5% of
the oil, respectively (Gamel et al. 2007), making amaranth oil an interesting
squalene source. Several sterols have been found, such as cholesterol and Δ7-​
campesterol, with the peculiarity that most of them are esterified. The main
tocopherol is β-​tocopherol, doubling that of α-​tocopherol (Krist 2020). Given
that β-​tocopherol does not possess antioxidant activity, amaranth oil is much
more resistant to oxidation than sunflower oil, preserving its characteristics for
more than 9 months (Gamel et al. 2007). Amaranth oil with a high content of
squalene and polyunsaturated fatty acids lowers total cholesterol, triglycerides,
low density lipoproteins (LDL) and very LDL, diminishing overall risk of heart
and circulatory diseases (Martyrosan et al. 2007). The presence of phytosterols,
tocopherols and tocotrienols, along with squalene, provide antioxidant capacity.
Although it is edible oil, its applications are not limited to the food area; they
are extended to cosmetics and health. Squalene is a skin moisturizer, analgesic,
healing and anti-​inflammatory (Krist 2020).

6.3.3 Fiber Extractions
Amaranth seeds contain between 8 and 16% w/​w of dietary fiber (DF) composed
mainly of xyloglycans and pectic polysaccharides. This type of fiber is similar
6.3 Amaranth 221

to fruits, vegetable and legume fiber, rather than to that of cereals (Lamothe
et al. 2015). Insoluble DF (78% w/​w) contains mainly galacturonic acid, ara-
binose, galactose, xylose and glucose, while soluble DF (22% w/​w) contains
glucose, galacturonic acid, and arabinose (Velarde Salcedo et al. 2019; Repo-​
Carrasco et al. 2009). Tosi et al. (2001) obtained a product rich in fiber by differ-
ential milling on conditioned grains and pneumatic classification, containing
approximately 64 and 7% w/​w of insoluble and soluble amaranth DF, respect-
ively. Calzetta-​Resio et al. (2006) have described obtaining a fraction very rich in
fiber as a by-​product of starch isolation. These authors used grain wet milling in
the presence of sodium metabisulfite in order to obtain amaranth starch as the
main product.

6.3.4 Protein Isolation
An amaranth protein isolate can be obtained by dispersing previously defatted
flour in an aqueous medium, solubilizing the proteins at a pH˃7 and then pre-
cipitating them at their isoelectric point, near to pH 4.5 (Martínez and Añón
1996; Salcedo Chávez et al. 2002; Cordero de los Santos et al. 2005). Amaranth
seeds contain different types of protein: albumins, present in small amounts,
globulins and glutelins which constitute the majority of reserve proteins.
Amaranth globulins belong to the 7S and 11S families, along with a third type
of globulin, P-​glb, with great tendency to polymerization (Castellani et al. 2000),
in which the precursor of 11S family A–​B subunits is not cleaved (Martínez et al.
1997; Quiroga et al. 2007). The pH used for protein solubilization in the isolation
process modifies the color and protein content of the isolate, in addition to the
polypeptide composition and the conformation of the proteins, increasing the
proportion of glutelins and the degree of denaturation when pH is close to 11
(Abugoch et al. 2010). Another technique for obtaining protein isolates is by the
micellization process, in which flour proteins are solubilized at neutral pH in the
presence of 0.8M NaCl, and then precipitated after ultrafiltration (membranes of
10 kDa nominal molecular weight limit) by diluting the solution in cold distilled
water and separating them by centrifugation. Even though the proteins obtained
by micellization showed a lower degree of denaturation and some improvement
in functional behavior, the yield was lower than that achieved by isoelectric pre-
cipitation (Cordero de los Santos et al. 2005).

6.3.5 Techno-​Functional Properties of Amaranth


Flour and Proteins
The information available for evaluating the techno-​functional quality of amar-
anth flours is scarce and is a field that needs to be further explored to promote
222 Fractionation of Seeds or Grains

the use of amaranth flours in diverse products. Tests designed for the study of
weak wheat flours have been applied to characterize the functional behavior
of amaranth flour (protein content 16% w/​w, Nx5.85). Water imbibing capacity
(WIC) and solvent retention capacity (SRC) determination showed that amar-
anth flour presented SRC values that were almost double those of wheat flour,
similar to values found in a high percentage of damaged starch wheat. WIC
values of amaranth and wheat flour were similar, while the initial hydration rate
was four times lower in amaranth flour, showing that it is necessary to hydrate
it before using it in cookies preparation to reach appropriate texture properties
(Sabbione et al. 2019).
Amaranth seed major protein fractions are good candidates due to their
variety in physicochemical and structural properties (Segura-​Nieto et al. 1994;
Martínez et al. 1997). Amaranth protein isolates provide important functional
versatility due to the structural diversity in their composition, but exhibit a rather
low solubility in aqueous solvents, limiting their use in the food industry (Segura-​
Nieto et al. 1994). Enzymatic hydrolysis is a chemical approach used to alter pro-
tein functionality improving their solubility, and the conditions can be selected
according to the desired application in food formulation. Several works docu-
ment this strategy by using plant proteases such as cucurbita or papain (Scilingo
et al. 2002) or trypsin (Condés et al. 2009). The proteins obtained significantly
improved their foaming properties and solubility even when heat treatments
were applied (Condés et al. 2009).The pH strongly influences the functional
behavior of proteins. Amaranth protein isolates showed the best foaming prop-
erties at acid pH (Bolontrade et al. 2013, 2016). Under these conditions, not only
does denaturation occur due to pH effect (close to 2), but also the activity of an
aspartic protease present in the isolate produces partial proteolysis modifying
the size of the polypeptides, resulting in the improvement of foaming properties
(Ventureira et al. 2012a). Recently, Malgor et al. (2020) described the application
of this functional behavior in an amaranth lemon sorbet, using amaranth isolate
as the protein source of the product. The best condition for emulsion formation
and stabilization of amaranth protein isolates was also the acidic one (Ventureira
et al. 2012b), even though isolates were able to form stable emulsions at pH close
to neutrality. Under acidic pH conditions, there is higher soluble protein con-
tent in the system. In addition, when denaturation of these proteins occurs, the
aspartic protease partially hydrolysates them, generating a greater exposure of
hydrophobic sites capable of anchoring in the interface, hence forming a visco-
elastic interfacial film more resistant to destabilization processes (Suárez and
Añón 2018). Protein dispersion obtained at pH close to neutrality was also suit-
able for obtaining stable emulsions at high protein concentration due to the
formation of flocs that considerably increased the viscosity of the continuous
6.3 Amaranth 223

phase, indicating that the presence of insoluble protein in the continuous phase
is not always detrimental to the stability of the emulsion. Amaranth protein iso-
late can also be used to form gels (Avanza et al. 2006) at concentrations higher
than 7% weight in volume (w/​v), increasing the water holding capacity of the
matrices when protein concentration increases (up to 20% w/​v). In addition,
increasing the heating temperature (70 to 95°C) provides high luminosity and
low water holding capacity gels. The main interactions involved in the gel net-
work are disulfide bonds between 11S globulin and P-​glb.
The search for foods that contribute to consumers’ health discovers in amar-
anth a very interesting ingredient. Different authors have described biological
activities derived from peptides encrypted in their proteins, which are released
by digestion or by specifically designed enzymatic treatments. Among the prop-
erties studied it is worth mentioning their action on cholesterol metabolism,
antioxidant, anticancer, anticoagulant and antihypertensive activities, among
others (Caselato-​Sousa and Amaya-​Farfán 2012; Silva-​Sánchez et al. 2008). Some
of these bioactivities have been proven using in vivo studies as well as in vitro, in
sílico and/​or ex vivo tests (Nardo et al. 2020; Sabbione et al. 2016a, 2016b; Sisti
et al. 2019).

6.3.6 Physico-​Chemical, Thermal and Rheological


Properties of Amaranth Starches
Amaranth starch shows outstanding properties from a techno-​ functional
point of view (Table 6.2). In general, the small granule size (0.75-​2 µm) and in
some cases low amylose content provide unique techno-​functional properties
of amaranth starch for food and non-​food applications such as fat replacers,
development of biomaterials as carrier or encapsulation, food thickeners, paper
coatings, laundry starch, dusting powers and cosmetics, among others (López
et al. 1994; Repo-​Carrasco-​Valencia and Valdez Arana 2017; Zhu et al. 2017). The
gelatinization temperature of the Amaranth spp. starch has a wide range, from
62.0–​69.5 and 77.7–​89.6°C. The high starch fraction from A. cruentus obtained by
differential milling of the amaranth grain can be considered an interesting raw
material for the production of precooked amaranth high-​starch flours having a
wide range of hydration properties (González et al. 2007). Amaranth starch had
better stability after freezing and thawing cycles, in the presence or not of sugars
or salt (Baker and Rayas-​Duarte 1998a) and a slower retrogradation rate than
maize, wheat and rice starches (Baker and Rayas-​Duarte 1998b) –​both proper-
ties important in food technology. In general, amaranth starches present lower
viscosity and a higher temperature for gelatinizing than quinoa and kañiwa
(Table 6.2).
224 Fractionation of Seeds or Grains

6.3.6.1 Modification of Amaranth Starches


Amaranth native starch has been chemically and/​or physically modified mainly
into three species: A. cruentus, A. hypocondriacus and A. paniculatus. The chemical
modifications were made by cross linking, acid treatment, oxidization, substitution
of hydroxyl groups by hydroxypropylation, succinylation, octenyl succinylation,
carboxymethylation or phosphorylation. Whereas physical modifications were
done by heat-​moisture treatment, annealing was performed by high hydrostatic
pressure or γ-​irradiation (Zhu 2017). The acid hydrolyzed amaranth starch could
be used in the candy industry, as sausage or dressing stabilizer or in parenteral food
and instant beverages (Repo-​Carrasco-​Valencia and Valdez Arana (2017). Starch
derivatives (succinylated, acetylated and phosphorylated) have proved effective
in encapsulating agents when used in the spray drying of flavors, pigments and
probiotics (Repo-​Carrasco-​Valencia and Valdez Arana (2017). Besides, amaranth
starch could be an excellent substrate for producing cyclodextrins because of its
high dispersibility, high starch-​granule susceptibility to amylases and its exception-
ally high amylopectin content (Urban et al. 2012).

6.4 CHIA
6.4.1 Seed Conditioning and Oil Extraction
An essential step in processing oilseeds is preparation of the seeds, which
includes seed cleaning, moisture conditioning and milling according to the pro-
cess employed for oil extraction.

6.4.1.1 Conventional Processes: Cold Pressing and


Solvent Extraction
Chia has been identified as a species with potential oil production because
of its nutritional quality. The seed contains a high level of oil, 25–​39%, rich in
polyunsaturated fatty acids, omega-​3 (α-​linolenic acid 54 –​67%) and omega-​6
(linoleic acid 12–​21%) (Ayerza and Coates 2011; Ixtaína et al. 2011a; Martínez
et al. 2012; Coorey et al. 2014). Many extraction techniques can be used to this
purpose; however, variation in the methods and the conditions used produce
differences in the yield and in the final chemical quality of the oil (Özcan et al.
2019; Fernandes et al. 2019). Conventional chia oil extraction processes include
cold pressing and solid-​liquid extraction with organic solvents. The oil yield,
fatty acid profile and the physicochemical and quality characteristics of chia
seed oils obtained by these methods have been studied by many researchers.
Oil recovery under pressing is smaller than by solvent extraction (Ali et al. 2012;
Dąbrowski et al. 2017). Martínez et al. (2012) and Dabrowski et al. (2017) could
6.4 Chia 225

maximize chia oil seed extraction using a Komet screw press (Model CA 59 G,
IBG Monforts, Germany). Martínez et al. (2012) hydrated chia seeds up to 10.1%
moisture content and used a restriction die of 6 mm, a screw press speed of
20 rpm and a barrel temperature of 30°C to achieve an oil yield value of 26.1%
(82.2% of the available oil). However, Dabrowski et al. (2017), who only reported
chia seed moistening up to 9%, obtained oil yields, for both hot (110°C) and cold
(room temperature) pressing conditions, of 26.3 and 24.1% (76.4 and 70.1% of the
available oil), respectively. These conditions allowed obtention of the remaining
oils with acceptable values for chemical quality parameters: peroxide (meq.
O2/​kg) and acid (mg KOH/​g ) values below 2.5 and 1.6, respectively; specific
extinction coefficients, k232 and k270, below 1.43 and 0.22, respectively. Regarding
minor compounds, total tocopherol content was between 498 and 616 mg/​kg;
phytosterol content of 4,800 mg/​kg and phenolic compounds (mg D-catechin/​
kg), carotenoids (mg lutein/​kg), squalene (mg/​kg) around 6.5, 5.3 and 71.1,
respectively. The oil oxidative stability was ~2 h with no significant differences
observed in the fatty acids profile. Classical solvent extraction technique requires
grinding the seed before extraction and enables the recovery of almost 95% of
the available lipid fraction. Silva et al. (2016) observed that the increase in the
amount of solvent had little influence on oil yield, and Kentish and Ashokkumar
(2011) reported that pretreating milled seeds in an ultrasonic bath constituted
an alternative to increasing the amount of extracted oil. The main solvents used
in chia oil extraction have included hexane (Ixtaína et al. 2011a; Silva et al. 2016;
de Souza et al. 2017; Dąbrowski et al. 2017), petroleum ether (Timilsena et al.
2017), acetone (Dąbrowski et al. 2017) and isopropanol and ethyl acetate (Silva
et al. 2016). It is important to highlight that these solvents permit obtention of
oils with acceptable values for chemical quality parameters, mainly acid and per-
oxide values; but great variability associated with minor compounds compos-
ition and oxidative stability was recorded. Dąbrowski et al. (2017) reported that
the higher polarity of the organic solvent increases both the extraction efficiency
of total lipids and bioactive components, especially phenolic compounds and
carotenoids, producing oils with higher oxidative stability. Regarding the fatty
acid profile of oils, the solvent used and seed to solvent ratio seem not to affect
this characteristic. However, the main disadvantages in using these solvents are
the damage caused to the environment and that the by-​products, which may
contain traces of solvent, cannot be used in the food industry due to a risk to
human health.

6.4.1.2 Non-​Conventional Processes: Supercritical


Carbon Dioxide Extraction (CO2-​SE)
The supercritical CO2 extraction (CO2-​SE) has emerged as an alternative to the
processes conventionally used for oil extraction previously mentioned. CO2-​SE
226 Fractionation of Seeds or Grains

is an environmentally friendly method for obtaining high-​quality products, free


of toxic solvent residues (Jokić et al. 2014). CO2 is used above its critical point (Tc
31.1°C, Pc 73.8 bar), presenting similar densities to those of the liquids and trans-
port properties that most closely approximate those of gases. Little modification
of the operating conditions close to the critical point affects either the yield of
extractable compounds or their composition, promoting selectivity within the
process (Ixtaína 2010).
Several authors carried out the extraction of chia seed oil with CO2-​SE using
different operating conditions, both on a laboratory and pilot scale (Table 6.3).
The extraction consists of two periods: an initial linear fast period and a second
slow linear period. Most of the extraction of chia seed oil mainly occurred in
the first extraction period (Ixtaína et al. 2010). Pressure and the time of extrac-
tion have a high positive impact on oil yield whereas the temperature presents
a lesser influence (Ixtaína et al. 2010, 2011b; Rocha Uribe et al. 2011; Dąbrowski
et al. 2018; Ishak et al. 2021). Thus, by optimizing the operating conditions it is
possible to obtain yields and fatty acid compositions similar to those of conven-
tional solvent extraction (Ixtaína et al. 2010, 2011a; Rocha Uribe 2011; Villanueva
et al. 2019). As can be seen in Table 6.3, oil recovery achieved was >87% in all
cases, and the maximum level was obtained at the highest pressures studied.
Regarding fatty acid composition, α-​linolenic and linoleic acids were the most
abundant, about 65 and 20% of total fatty acids, respectively (Ixtaína et al. 2010,
2011b; Rocha Uribe et al. 2011; Villanueva et al. 2019; Ishak et al. 2021).
The addition of low levels of co-​solvents such as water, ethanol, acetone, iso-
propanol and others, can modify the polarity of the CO2-​SE and consequently the
solubility of different compounds. Dąbrowski et al. (2017, 2018) studied the CO2-​
SE of chia seed oil using diverse conditions and acetone as a co-​solvent. These
authors obtained oils with similar fatty acid composition but with different total
phytochemical content, highlighting the most abundant in squalene, sterols and
tocopherols (1 h, pure CO2-​SE) and the richest in polyphenols and carotenoids
(1 h, 10% acetone as co-​solvent). These differences are due to the rise in polarity
of solvent maximizing the content of amphiphilic compounds (polyphenols),
whereas shorter extraction time increased concentration of the most apolar
phytochemicals (tocopherols, squalene and phytosterols) (Dąbrowski et al. 2018).
A recent study of the CO2-​SE from edible and discarded chia seeds, which were
rejected during post-​harvest due to their low weight, small size and presence of
broken pericarp, among others, reported that it is possible to use these seeds
for the extraction of oil with similar characteristics to those of the edible seeds
(Villanueva et al. 2020). Thus, this environmentally friendly technology has great
potential to be used by the food industry to extract chia oil of high quality with
bioactive compounds (antioxidants and phytosterols), depending on the oper-
ating conditions applied.
newgenrtpdf
TABLE 6.3 SUPERCRITICAL CARBON DIOXIDE EXTRACTION (CO2-​SE) OF CHIA SEED OIL AS NON-​CONVENTIONAL
PROCESSES AT DIFFERING OPERATING CONDITIONS
Operating conditions
Maximum oil
CO2 flow Co-​ Maximum oil yield1 recovery 1,2
Scale P (bar) T (°C) t (min) rate solvents (g oil/​100 g seeds) (%) Reference
Lab 250/​350/​ 40/​60/​80 60/​150/​ 30 g/​min -​ 29.0 88 Ixtaina et al.
450 240 (450 bar, 80°C, 240 min) (450 bar, 80°C, 240 min) (2010)
Pilot 250/​450 40/​60 135–​423 133 g/​min -​ 33.0 97 Ixtaina et al.
(450 bar, 60°C, 138 min) (450 bar, 60°C, 138 min) (2011)
Lab 136/​272/​ 40/​60/​80 40 1.8 g/​min -​ 7.2 n.a. Rocha-​Uribe
408 (408 bar, 80°C, 40 min) et al. (2011)
Lab 280 70/​90 300 8 mL/​min -​ 29.9 87 Dąbrowski
(280 bar, 70°C, 300 min) (280 bar, 70°C, 300 min) et al. (2017)
Lab 280 70 60/​300 8 mL/​min 2/​6/​10 33.7 97 Dąbrowski
% acetone (280 bar, 70°C, 300 min, (280 bar, 70°C, 300 min, et al. (2018)
10% acetone) 10% acetone)
Pilot 250/​450 40/​60 240 40 g/​min -​ 18.6 90.3 Villanueva
(450 bar, 60°C, 240min) (450 bar, 60°C, 240min) Bermejo
et al. (2019)
Pilot 250/​450 40/​60 240 40 g/​min -​ 18.6 (DCS) 93.5 (DCS) Villanueva
25.2 (ECS) 89.9 (ECS) Bermejo
(450, 60°C, 240 min) (450, 60°C, 240 min) et al. (2020)
Lab 220/​280/​ 40/​60/​80 300 4 mL/​min -​ 30.7 87 Ishak et al.

6.4
340 (335 bar, 45°C, 300 min) (335 bar, 45°C, 300 min) (2021)
Notes: P, extraction pressure; T, extraction temperature; t: extraction time; DCS, discarded chia seeds; ECS, edible chia seeds.

Chia
1 Operating conditions for the maximum oil yield/​recovery obtained are indicated in parentheses.
2 Values calculated with respect to the total oil content analyzed by Soxhlet or Folch procedures.
n.a. not available.

227
228 Fractionation of Seeds or Grains

6.4.1.3 Residual Flours

6.4.1.3.1  Characterization of residual flours from different oil


extraction processes
The residual flours from the oil-​extracting processes of chia seeds are a good
source of total dietary fiber (>40%, d.b.), proteins and minerals –​mainly phos-
phorus, calcium and magnesium –​polyphenolic compounds (chlorogenic and
caffeic acids, quercetin, myricetin and kaempferol), and tocopherols with high
antioxidant activity (Reyes-​Caudillo et al. 2008; Capitani et al. 2012). Chia flour
from the pressing process exhibits residual oil content 7.06–​11.39% (d.b.), with
high omega-​3 and omega-​6 fatty acids concentration (6.85 and 2.16%, respect-
ively) (Capitani et al. 2012; Aranibar et al. 2018).

6.4.1.3.2  Characterization of residual flour from seeds after


removal of mucilage
Capitani et al. (2013a) performed a comparative evaluation of the physico-
chemical and functional properties of chia flours obtained from seeds with and
without mucilage. The flour obtained from seeds with previous mucilage extrac-
tion exhibited a higher content of insoluble dietary fiber (45.62% d.b.) than the
others (41.13% d.b.), with a significant decrease in their soluble dietary fiber
content (1.51% d.b.). The flour without mucilage exhibited a lower absorption
and water holding capacity than those of the whole flour. This behavior can be
associated with the mucilage, soluble dietary fiber, capable of holding water
inside its matrix. Both types of chia flours presented a low absorption of organic
molecules and oil-​holding capacity, being significantly higher in chia flour
without mucilage. These properties turn those flours into important ingredients
in the manufacture of fried products due to their low fatty mouth-​feel contri-
bution. Moreover, they show the potential use of a by-​product of the mucilage
extraction with favorable nutritional characteristics.

6.4.2 Protein and Fiber Isolation


6.4.2.1 Protein Rich Fraction
The fractionation procedure by dry sieving to obtain protein rich fraction (PRF)
from chia defatted flour consists of sifting the flour using a mesh through which
the high protein content fraction passes, and the fiber-​enriched fraction is
retained (Vázquez-​Ovando et al. (2010). A high globulin proportion (64.86%) and
a minor glutenin level (20.21%) were present, containing the PRF 44.62% of crude
protein. Capitani et al. (2013b) obtained PRF from chia flours from different oil
extraction processes. After sieving both meals, the fractions that passed through
the sieve mesh exhibited an increase in protein content (45.6 and 63.5% fraction
cold-​pressing and solvent, respectively). Julio et al. (2019) determined the amino
6.4 Chia 229

acid composition of the chia PRF, indicating an important content of essential


sulfur amino acids but limited lysine and tryptophan. These authors obtained
differing protein fractions (albumins, globulins, glutelins and prolamins) from
PRF by solubility gradient, with globulins being the predominant fraction
(64.86%). Also, oil in water (O/​W) emulsions with PRF and their fractions at
different pH levels in their native and denatured state were obtained. Diverse
authors studied the physicochemical, functional and biological properties of the
PRF (Segura-​Campos et al. 2013a, 2013b; Vázquez-​Ovando et al. 2013).

6.4.2.2 Protein Isolates
Chia seed is a source of vegetable protein with biological activities. The extrac-
tion of protein fractions from seeds with previous mucilage removal was carried
out by Olivos-​Lugo et al. (2010), Sandoval-​Oliveros and Paredes-​López (2013),
Timilsena et al. (2016a), López et al. (2018), and Urbizo-​Reyes et al. (2019),
followed by lipid extraction using solvents or a cold screw-​press. Other authors
generated protein isolates from defatted seeds but without mucilage removal
(Cárdenas et al. 2018; Julio et al. 2019). Chia protein isolates showed good water-​
holding capacity (4.06 g/​g ) and excellent oil-​retention capacities (4.04 g/​g ),
making it an interesting additive in bakery products and food emulsions (Olivos-​
Lugo et al. 2010). The denaturation temperatures of crude albumins, globulins,
prolamins and glutelins were 103, 105, 86, and 91°C, respectively, indicating
excellent thermal stability for albumins and globulins and suggesting that they
are suitable for food products undergoing heat treatment (Sandoval-​Oliveros
and Paredes-López 2013; López et al. 2018). Chia protein hydrolysates obtained
by the sequential hydrolysis with microwave treatment showed high in vitro
antioxidant activity (Urbizo-​Reyes et al. 2019).

6.4.3 Fiber Rich Fraction


Chia fiber rich fraction (FRF) was produced by dry processing of defatted flour
(Vázquez-​Ovando et al. 2009; Capitani et al. 2012), and pressing extraction
(Capitani et al. 2012; Oliveira-​Alves et al. 2017). The generated fraction presented
a high content of total dietary fiber (51.98–​56.46% d.b.) consisting mostly of
insoluble dietary fiber (47.65–​53.45% d.b.) with low soluble dietary fiber content
(3.01–​4.33% d.b.). Moreover, the chia FRF exhibited high antioxidant activity
associated with the polyphenolic compounds, mainly caffeic acid and rosmarinic
and salvianolic acids, and the presence of tocopherols in the pressing fraction
(Vázquez-​Ovando et al. 2009; Capitani et al. 2012; Oliveira-​Alves et al. 2017).

6.4.3.1 Mucilage: Extraction and Characterization


Chia mucilage (CM) is an anionic heteropolysaccharide consisting mainly of the
sugars xylose and glucose in a 2:1 ratio, with significant amounts of glucuronic
230 Fractionation of Seeds or Grains

and galacturonic acids (Timilsena “al. 2016b”). The complex monosaccharide


composition in terms of xylose as building backbone, is similar to other muci-
lage, such as those extracted from Plantago psyllium seed, basil seed and Chinese
quince (Liu et al. 2021). CM is localized in the fruit exocarp and exuded from
the polygonal cells of the epidermis coat of chia seeds when they are placed in
an aqueous solution, being composed mainly of carbohydrates fiber. Upon full
hydration, filaments became apparent and conformed to a transparent “capsule”
attached to the seed (Muñoz et al. 2012; Capitani et al. 2013c; Salgado-​Cruz et
al. 2013). Different methods have been developed for CM extraction obtaining a
wide range of yields, influenced by the applied operating conditions (Table 6.4).
In a recent study by Muñoz-​Tebar et al. (2021), CM was extracted from defatted
flour chia obtaining a yield of 15.3%, relating this high yield with the compounds,
such as proteins, released after grinding the seeds. CM contains different protein
levels after extraction (2.6–​11.95%), which could be related to the methods used
to separate the seeds from the liquid mucilage, soaking conditions, and proteins
covalently linked to the polysaccharide (Capitani et al. 2015; Timilsena et al.
2016b); Ferreira Ignácio Câmara et al. 2020). The CM fatty acid profile includes
significant levels of linolenic acid (56.16%) and linoleic acid (20.64%) (Ferreira
Ignácio Câmara et al. 2020). CM solubility increases with temperature, with pH
being maximized at 60ºC (87%) (Capitani et al. 2013c) and alkaline pH (97%)
(Timilsena et al. 2016b). Timilsena et al. (2016b) indicate that CM resists pyro-
lytic decomposition (T >250°C). CM dispersed in water presents irregular shape,
fibrous microgel particles and average size D4,3 ∼700 μm (Goh et al. 2016). García-​
Salcedo et al. (2018) studied its pasting properties indicating that its relative vis-
cosity did not exhibit dependency on temperature. Rheological measurements
reported a weak viscoelastic gel with shear-​thinning behavior dependent prop-
erties even at low levels (Capitani et al. 2015; Timilsena et al. 2015; Goh et al.
2016; García-​Salcedo et al. 2018; Punia et al. 2019). Velázquez-​Gutiérrez et al.
(2015) determined the sorption isotherms (25 and 40°C), and some thermo-
dynamic properties indicating a sigmoidal shape for the sorption isotherms of
freeze-​dried CM, characteristic of glassy biopolymer isotherms. The pore radius
of CM varied from 0.87 to 6.44 nm and maximum stability 7.56–​7.63 kg H2O/​100
kg of dry solids (aw 0.34–​0.53) in the studied range. The glass transition Tg of the
mucilage was found to be between 42.93 and 57.93°C.

6.4.4 Techno-​Functional Properties of Residual Flours,


Proteins and Soluble Fiber (Mucilage)
6.4.4.1 Water-​holding Capacity, Absorption and
Adsorption Capacity
By-​products from oil chia seed extraction ( flours, proteins and fiber fraction)
present high WHC for the mucilaginous fraction which has excellent
newgenrtpdf
TABLE 6.4 CHIA MUCILAGE EXTRACTION YIELD AT DIFFERING OPERATING CONDITIONS
Operating conditions
Mucilage
seed: water yield
T (°C) ratio t (h) pH solubilization separation (% d.b.) Reference
25 1:20 1 -​ magnetic stirring vacuum-​assisted filtration/​ 7.78 Marin Flores et al.
drying (2008)
25 1:20 2 -​ sonication vacuum-​assisted filtration/​ 10.79 Marin Flores et al.
drying (2008)
25 1:20 2 -​ magnetic stirring vacuum-​assisted filtration/​ 15.1 Marin Flores et al.
and sonication drying (2008)
20/​40/​80 1:20/​1:30/​ 2 4/​6/​8 magnetic stirring drying oven/​rubbing 6.97 Muñoz et al. (2012)
1:40
25 1:10 -​ without stirring freeze-​drying/​rubbing 3.8 Capitani et al. (2013c)
50 1:20 0. 45 -​ magnetic stirring centrifugation/​freeze-​drying 10.9 Segura-​Campos et al.
(2014)
30-​80 1:10–​1:30 2-​4 -​ magnetic stirring drying oven with forced air 4.95 Campos et al. (2016)
circulation/​rubbing
25 1.20 4 -​ magnetic stirring centrifugation/​ethanol 1.20 Goh et al. (2016)
precipitation/​ freeze-​drying
-​ 1:20 2 -​ magnetic stirring freeze-​drying/​rubbing 5.6 Timilsena et al. (2016b)
15-​85 1:12–​1:40.8 ½/​3 -​ magnetic stirring freeze-​drying/​rubbing 11.6 Orifici et al. (2018)
27 1:10–​1:40 2 8 magnetic stirring cold pressing/​freeze-​drying 8.46 Tavares et al. (2018)

6.4
refrigeration 1:20 24 -​ pre-​heat (55°C) vacuum-​assisted filtration/​ 7.65 Urbizo-​Reyes et al.
and sonication drying 40°C, 12h (2019)

Chia
Notes: T, extraction temperature; t: extraction time; d.b.: dry basis.

231
232 Fractionation of Seeds or Grains

water-​holding properties (Vázquez-​Ovando et al. 2009, Capitani et al. 2013b).


By-​products from solvent extraction exhibit better functional properties than
others, by their lower level of residual lipids and the higher protein content
(Capitani et al. 2013b). In contrast, chia by-​products present low water adsorp-
tion capacity(WAdC).

6.4.4.2 Oil Holding, Absorption and Adsorption Capacities,


and Absorption Capacity of Organic Molecules
Vázquez-​Ovando et al. (2009) and Capitani et al. (2013b) reported low oil-​
holding capacity (OHC) (1.09–​2.06 g/​g ) in chia flours, fiber and protein fraction,
suggesting that these by-​products are potential ingredients in fried products
since they would provide a non-​greasy sensation. Chia by-​products could effi-
ciently interact with fats, biliary acids, cholesterol, drugs and toxic compounds
at the intestinal level, due to the absorption capacity of organic molecules (0.79–​
1.94 g/​g ). The FRF exhibits the highest capacity to link to organic molecules
(Vázquez-​Ovando et al. 2009; Capitani et al. 2013b).

6.4.4.3 Emulsifying Properties, Foaming Capacity and


Foam Stability
Chia by-​products present high emulsifying activity and emulsion stability
(44.33–​56.00, 34.33–​94.84 mL/​100 mL, respectively), being at the higher levels
of these parameters in the solvent extracted by-​products than others because
of their low level of residual lipids and higher protein content (Vázquez-​
Ovando et al. 2009; Capitani et al. 2013b). In chia PRF, Vázquez-​O vando et al.
(2013) reported emulsion activity of 50–​56%, independent of pH and emulsion
stability of 92% at pH 8 and 10. Capitani et al. (2012) reported fairly stable
emulsions with chia flours and fibrous fraction from solvent by-​products due
to their high proportion of proteins and fibers as emulsifying and stabilizing
agents. Capitani et al. (2016) and Guiotto et al. (2016) obtained stable O/​W
emulsions with ≥0.75% of chia mucilage (CHM) during refrigerated storage
by a reduction in the mobility of oil particles in a tridimensional network.
Foaming capacity (FC) and foam stability (FS) of chia PRF indicated that it
is not a good foaming agent but presents a foam stability of 80% (Vázquez-​
Ovando et al. 2013).

6.5 SACHA INCHI
6.5.1 Decapsulation and Seed Conditioning
Sacha inchi seeds, and the most widespread species, Plukenetia volubilis and
Plukenetia huayllabambana are relatively large almonds (1.3–​2.1cm) with a
6.5 Sacha Inchi 233

high amount of oil content (49–​54 wt%), followed by the proteins (24–​25 wt%).
The seeds also contain minor quantities of fiber (11–​15 wt%), water (5–​7 wt%),
carbohydrates (3–​7 wt%) and ashes (3–​4 wt%). The oils are formed in high pro-
portion by essential fatty acids, ω3-​α-​linolenic (>48%) and linoleic (>26%). Before
extracting the oil, one of the post-​harvest essential stages of decapsulation are
implemented –​drying (artificially, or naturally under the sun) –​as well as the
subsequent stage, dehulling. For all these stages, there are currently different
technological possibilities for releasing the sacha inchi seeds from the capsule
and shell that contain them, with the least possible damage. To facilitate oil
extraction using any of the chosen methods, it is sometimes necessary to reduce
the seed size by grinding.

6.5.2 Oil Extraction Processes


In general terms, four kinds of extraction methods are implemented for sacha
inchi oils, namely, pressing extraction, aqueous extraction, solvent extraction
and supercritical fluid extraction.

6.5.2.1 Conventional Processes
Extraction of sacha inchi oil from seeds by pressing is cost‐efficient and the
most extended method in the industry for obtaining genuine extra virgin oil,
and it is the shortest process. The pressure extraction equipment in most cases
consists of a perforated cylinder subjected to high pressure. It is quite easy
to control, needs low investment and has a low operating cost. Furthermore,
few parameters besides the speed of rotation and the geometry of the press
cylinder need to be controlled. In some cases, to guarantee a steady oil tem-
perature, a cooling system is installed at the outlet. With the pressing process,
usually 75–​95 wt% of the oil is recovered from the seeds. The effects of the
drying (60, 75, 90°C) and press head (60, 75, 90°C), temperature on oil yields
from sacha inchi (P. volubilis L.) seeds using a single screw press machine
have been studied (Muangrat et al. 2018). As no temperature or solvent is
used, most of the health-​beneficial components remain in the oil, such as
tocopherols and phytosterols. It is the eco-​friendliest extracting system,
together with extraction by steam drag. Steam dragging is a technique that
uses a stream of water vapor for the extraction; after some maceration time,
the plant tissue is broken, releasing oil. Oil extraction is facilitated by incorp-
orating several classes of proteases and cellulase enzymes. Nevertheless, even
using optimal conditions of enzyme loading, water-​to-​sample ratio, extrac-
tion time and temperature, oil yield is lower than 30 wt%. Contrary to the
solvent-​free processes mentioned above, there are procedures which use
solvents for oil extraction. Soxhlet (AOAC 1990), with petroleum ether, is the
234 Fractionation of Seeds or Grains

most widespread and preferred solvent extraction method for many seed oils.
However, in the case of sacha inchi, this method is preferred for extracting the
oil remaining after the first pressure extraction, which is usually in the range
of 5–​25 wt%.

6.5.2.2 Non-​conventional Processes
The supercritical extraction using carbon dioxide (CO2) is the most widespread
non-​conventional procedure for extracting sacha inchi oils. Studies of the influ-
ence of temperature on the system, pressure ranges and CO2 flow were tracked
in lab scale and at pilot plant scale (Triana-​Maldonado et al. 2017). Maximum
recovery was 60 wt%, achieved at the CO2-​SE pilot scale. Other different solvents
such as n-​propane were also assayed (Zanqui et al. 2016).

6.5.2.3 Residual Flour
After the oil is obtained, the sacha inchi cake is ground in a hammer mill to
obtain ∼500μm particles. Then it is exhaustively defatted, usually by an organic
solvent method. The flour obtained from the sacha inchi cake has a very high
percentage of protein (>50 wt%) and can be used for human consumption in
the meat, dairy, bakery and confectionery industries. Sacha inchi protein has
been reported to have high content of cysteine, tyrosine, threonine and trypto-
phan and low content of phenylalanine (Ruiz et al. 2013). In order to eliminate
its astringent taste, an extrusion process is necessary, and this is conducted at
a moderate temperature of 60–​90°C for a very short time, 5–​7 sec. This process
improves the biological quality and inactivates tannins, improving digestibility
of protein and other nutrients.

6.5.3 Isolates and Protein Hydrolysates


The production of sacha inchi protein isolates is of growing interest to the
food industry because of its increasing application in food markets, nutra-
ceutical products, and functional foods. A water-​soluble albumin fraction
comprising ∼25 wt% of sacha inchi cake has been obtained (Sathe et al. 2002;
Quinteros et al. 2016). This is composed of two glycosylated polypeptides,
with estimated molecular weights (MW) of 32.8 and 34.8 kDa, respectively. It
contains all the essential amino acids in adequate amounts when compared to
the FAO/​WHO (FAO 2007) recommended pattern for a human adult. Protein
from the sacha inchi cake can also be removed by optimized enzyme‐assisted
extraction (Chirinos et al. 2016), or by an alkaline process using isoelectric
precipitation.
6.6 Tarwi 235

6.5.4 Techno-​Functional Properties of Sacha


Inchi Cake Flour
The values of solubility, oil absorption capacity, foaming capacity, foam stability
and emulsifying capacity are techno-​functional properties that influence sen-
sorial characteristics, and are essential in different food systems. In the case
of sacha inchi cake flour, those properties show values of higher magnitudes
compared to other protein isolates and can be particularly useful for the pro-
duction of sausages, mayonnaise and ice cream, among other food products
(Mercado et al. 2015).

6.6 TARWI
Tarwi or Andean lupin (Lupinus mutabilis) is a leguminous crop with a remark-
able chemical composition and an elevated content of both protein and
oil. Tarwi thus has a twofold advantage: it is a valuable food crop in the fight
against malnutrition, as well as a promising cash crop for edible oil produc-
tion. It has more protein than soybean and even more than other lupine species.
According to Carvajal-​Larenas et al. (2016), the average seed protein content
(g/​100 g db) reported in the literature for different Lupinus species is: L. albus,
38.2; L. angustifolius, 33.9; L. luteus, 42.2; and L. mutabilis, 43.3. Tarwi oil is of
nutritionally good quality, its main fatty acids being oleic and linoleic acids
(Repo-​Carrasco-​Valencia 2020). The low linolenic acid content gives tarwi
oil good stability against oxidation. The oil does not contain any erucic acid,
which is known to be present in the oil of other lupines such as L. angustifolius
(Hatzold et al. 1983). The main carbohydrates in tarwi are oligosaccharides. The
stachyose content in tarwi is relatively high. Starch is present only in very small
amounts, therefore products of high viscosity cannot be produced, although
this is compensated by the relatively high pentosan content (Beirão da Costa
1989). The main hindrance to the wider use of tarwi as food and animal feed
has been its content of toxic, bitter quinolizidine alkaloids. Lupanin, spartein,
4-​hydroxylupanin and 13-​hydroxylupanin are the principal alkaloids found in
L. mutabilis. The alkaloid content in most common varieties of tarwi is in average
5 g/​100 g (Cortés-​Avendaño et al. (2020).

6.6.1 Alkaloid Extraction
The debittering of tarwi seeds is necessary for the removal of toxic alkaloids.
This process is traditionally carried out as follows: The lupin grains are hydrated
236 Fractionation of Seeds or Grains

for 12 h at room temperature, followed by cooking for 1 h. The cooking water is


discarded and the seeds are washed with running water for 5 days. Simultaneous
extraction of fat and alkaloids may be applicable for tarwi. The efficiency of the
extraction procedure was studied using different solvents by Beirão da Costa
(1989). The solvent which produced the best results was ethanol 35%. Concerning
alkaloid extraction by water, some trials showed that this may be improved
100% if the lupin material is primarily subjected to enzymatic treatment with
cellulases.
Debittered tarwi can be consumed directly as a snack (Villacrés et al. 2010)
but it can also be used as an ingredient in many different products such as salads,
traditional dishes, bread, cookies, noodles, extruded drinks, protein isolate and
concentrate, and refined oil (Hatzold et al. 1983; Córdova-​Ramos et al. 2020) as
shown in Figure 6.1.
Lupine flour can be obtained after the oil extraction process either ground
through a roller mill (Koivunen et al. 2016) or hammer mill or grinder (Zaworska
et al. 2017). The flour can be used in different products, such as breads, muffins
and pasta (Wandersleben et al. 2018; Nasar-​Abbas and Jayasema 2011; Martínez-​
Villaluenga et al. 2010)

6.6.2 Oil Extraction
Lupine oil can be obtained by pressing or by the conventional method using a
solvent. To perform the extraction of the oil, Hatzold et al. (1983) carried out
a process which included the following steps: cleaning, peeling and cooking
at 80°C to finally extract the oil using hexane as a solvent. The crude oil was
degummed and debittered by washing with diluted acids, then neutralized,
blanched and deodorized. The alkaloids are concentrated in the defatted oil
cake. Crude oil, however, contains only small amounts of alkaloids. At the end
of the refining process the alkaloid content of the edible oil had decreased to the
very low level of 5 ppm.

6.6.2.1 Characterization of Residual Flour


Residual lupin flour and its extracted components, such as protein or fiber, can
be added to foods to enhance their nutritional quality (Villarino et al. 2016).
Lupin protein isolates are extracted by solubilization of protein from wet-​milled
kernels or lupin flour (defatted or non-​defatted) at high alkaline pH (e.g., 9),
removal of the insoluble portion (dietary fiber) through centrifugation, followed
by acid precipitation of the major globulin proteins (i.e., α-​ and β-​conglutins) at
pH 4.5 (Johnson et al. 2017). The insoluble fiber residue, acid-​precipitated pro-
tein, and acid-​soluble “whey fraction” can then be dried for use as staple food
ingredients against deteriorative processes.
6.6 Tarwi 237

Tarwi

Salad
Debittering Traditional dishes

Drying
Bread
Cookies
Coarse grinding Fine grinding Noodles
Extruded
Drinks

Oil extraction Protein isolate


Pressing cake Protein concentrate
Solvent

Animal feeding

Refinement Refined oil

Figure 6.1 Products obtained from tarwi.


Notes: Tarwi can undergo successive processes in which different products/​ingredients
can be obtained at each step. The processes are: debittering, in which the seeds could be
used to prepare salad and traditional dishes; followed by drying and then coarse grinding
or fine grinding (in which the flour can be used to prepare: bread, cookies, noodles,
extruded products and drinks); after that oil obtention by pressing and/​or by solvent
extraction, followed by the refinement step to obtain the refined oil; with the remaining
cake (after extraction) the isolate or concentrate protein is obtained, or the cake can be
used for animal nutrition. Source: Authors’ own work.

Sironi et al. (2005) studied the techno-​functional properties of lupin pro-


tein isolate obtained by isoelectric precipitation. They found that this fraction
has excellent emulsifying properties but low viscosity and reduced gel-​forming
properties. The acid-​soluble whey fraction obtained in isoelectric precipita-
tion also contains valuable protein components, such as the bioactive peptide
γ-​conglutin; it has stable foam-​forming properties (Wong et al. 2013). Due to
their functional properties, which are important in food manufacture, lupin pro-
tein fractions have been categorized as emulsifying fraction (α-​, β-​conglutin)
and foaming fraction (γ-​conglutin-​rich) (Sironi et al. 2005).
238 Fractionation of Seeds or Grains

Only limited studies have been reported on the use of lupin fiber fractions
as ingredients for the development of high-​fiber foods. Archer et al. (2004) used
lupin kernel fiber as an effective fat replacement in sausage patties. Roberfroid
(2007) studied acid-​soluble whey from lupin fractionation and discovered that
the oligosaccharides present could have prebiotic activity which stimulates the
growth of beneficial gut bacteria.

6.7 BLACK TURTLE BEAN


6.7.1 Milling Process
Dry bean milling has attracted interest due to the increasing need for non-​wheat
food ingredients. Some studies have provided useful information on the poten-
tial of black turtle bean (Phaseolus vulgaris) flour as a functional ingredient in
the food system (Audu and Aremu 2015; He et al. 2018).
Milling is a process where the reduction of grains to meal or flour occurs and
includes grinding, sieving and purifying (Thakur et al. 2019). For many years,
wheat and corn milling has been studied, and advances in knowledge have led
to an understanding of the different milling procedures used today. Little infor-
mation is available concerning dry bean milling on different flours. Nowadays,
study of the milling of dry beans is increasing due to the beans’ rich nutritional
profile.
In general, for the production of black turtle bean flour, seeds that were previ-
ously (or not) subjected to some treatment processes, are now dried in air forced
ovens or in sun-​dryers and finally they are ground. The ground seeds are sieved
through a single mesh to obtain flour (Figure 6.2).
To obtain defatted black bean flour, the full-​fat flour is soaked in a solvent;
after being allowed to stand, the mixture is removed, dried and finally pulverized
(Figure 6.3). The particle size of milled dry beans is an important variable on
flour quality. For size reduction, a combination of forces (e.g., abrasion, shearing,
compression and impact) might be used. Dry beans are often milled using a
one-​pass system that utilizes a plate mill (disk mill, burr mill) or a hammer mill
(Raigar and Mishra 2018). The hammer mill is an impact-​type mill where particle
size reduction depends on hammer design. A disadvantage of hammer milling
is that it is less uniform in the particle size distribution and less energy efficient
compared to a roller mill. A burr mill, also known as a plate mill, consists of a
set of knives, which use cutting, shearing and crushing actions for particle size
reduction. It has two circular plates and the material is fed between them. It
utilizes a multiple-​stage approach in a series of rolls for fine particle size reduc-
tion. Carter and Manthey 2019 have demonstrated the effectiveness of using a
newgenrtpdf
6.7 Black Turtle Bean
Figure 6.2 Flow chart for the production of fatted black turtle bean flours.
Notes: Flow chart of whole black turtle bean seeds, describing the six different types of processes for obtaining the full black turtle flour.
These processes start with the raw seeds or soaked seeds (which are manually de-​hulled and then dried) or sprout seeds (which are de-​
hulled and dried) or de-​hulled raw seeds (which can be roasted and then cleaned, cooked or boiled and dried or cooked and dried) or
fermented seeds (which are blanched and dried); all of the above are then followed by milling and sieving. Source: Authors’ own work.

239
240 Fractionation of Seeds or Grains

Figure 6.3 Flow chart for the production of defatted black turtle bean flours.
Source: Authors’ own work.
Notes: Flow chart of the production of defatted black turtle bean flours, which undergo
the following processes: soaking the flours, solvent extraction (by hexane or petroleum
ether) of crude oil followed by cake desolventizing, defatted slurry, dried and ground to
obtain defatted black turtle flour.

centrifugal mill for milling black beans. A centrifugal mill consists of an impeller
that spins at high velocity and is located inside a cutting or an abrading screen.

6.7.2 Protein Concentrates and Isolates


Black turtle bean flour can be further processed by fractionation to isolate its
components (e.g., protein, starch and dietary fiber) for using in food applications.
Protein isolates and concentrates have been extensively studied as food and
dietary supplements for several oilseeds, but there is very little information on
their production procedure and corresponding functionality. The easiest method
of separating protein from the cellular structure of the seed is by extracting with
an aqueous solvent for successful utilization of food products. Figure 6.4 is a dia-
gram showing how to obtain black turtle bean protein isolate and concentrate.
In general, the defatted flour with alkaline solution is mixed by stirring for
few minutes. The slurry is centrifuged and the supernatants are treated with
95% (v/​v) ethanol; the pH is adjusted by stirring with an acid solution and the
precipitated proteins are recovered by filtration. The protein concentrate is
dried or lyophilized in a forced-​air oven. To obtain the isolate, the defatted bean
flour is suspended in water and the pH adjusted with an alkaline solution to
extract the protein while stirring. Then, the supernatants’ pH is adjusted to 4
with an acid solution and the precipitate is obtained by centrifugation. The pre-
cipitate is washed with distilled water and after the pH has been readjusted, it
is finally dried in an oven. Recently, more protein researchers were concerned
about the conformational changes of the food protein molecules by adjusting
the pH to neutral after a low acidic condition treatment (i.e., pH-​shifting) (Jiang
et al. 2017). According to He et al. (2020), low pH changes in treatment during the
protein extraction process could have great potential to produce improvements
in emulsifying, foaming and fat-​holding properties, and lead to the reduced
6.7 Black Turtle Bean 241

Figure 6.4 Flow chart for the production of protein concentrates and protein isolates of
black turtle bean flours. Source: Authors’ own work.
Notes: Flow chart for the production of protein concentrate and isolate from defatted
black turtle beans flour. To produce the protein concentrate, defatted black turtle flour
has to be processed to carbohydrate-​free flour (by adjusting the pH and the extraction of
ethanol) and then has to be dried. On the other hand, to produce a protein isolate, the
defatted black turtle flour has to be soaked in water to produce a protein extract, then the
obtention of protein slurry (by adjusting the pH and centrifugation) and finally it under-
goes drying, milling and sieving. The whey is discarded. After the first soaking, the plant
residues undergo a second protein extraction and the residue is discarded.

solubility and water-​holding capacity of the isolate of black turtle bean protein
with a high level of consumer safety.

6.7.3 Oil Extraction
Edible oils from plant sources are receiving growing interest in various food
applications and industries, due to their high concentration of bioactive lipid
components, such as polyunsaturated fatty acids and phytosterols that have
shown various health benefits (Aremu et al. 2015). Oil can be extracted from
the black turtle bean seeds using traditional methods (small-​scale), mech-
anical (hydraulic and screw) presses that can be manual, semi-​automated or
automated, and extraction with solvents ( for example, hexane, fluid carbon
dioxide), or a combination of two of these methods (Aremu et al. 2015). Values of
the physicochemical parameters reported by Audu et al. (2013) have shown that
242 Fractionation of Seeds or Grains

black turtle bean oils might be useful as edible oils due to their stability as frying
oils. Unsaturated fatty acids predominate in raw and processed black turtle bean
flours with an adequate amount of essential fatty acids (Padhi et al. 2017; Audu
et al. 2013).

6.7.4 Techno-​Functional Properties of Flours, Protein


Concentrate, Oil and Fiber
Several studies have reported that the type of mill used, as well as conditions
pertaining during pretreatment and milling, can influence the physical, chem-
ical and pasting properties of the flour (Audu et al. 2013; Audu and Aremu 2015;
Carter and Manthey 2019; Okafor et al. 2015). These changes in the external and
internal structure of black bean seeds affected their milling and flour proper-
ties. Pretreatment processing and milling significantly affected the bulk density
of black turtle bean flours by increasing it. Both soaking and cooking affected
the final flour color by decreasing the lightness due to the high leaching of seed
coat pigments into the cooking water. All the processing methods are shown in
Figure 6.2, which enhanced water absorption (WAC) and foaming capacities of
these flours. Boiling, cooking, roasting, sprouting and fermentation affected the
oil absorption capacities (OAC) and lowest gelation concentration or least gel-
atinous capacity (LGC) by increasing them. All the processing methods reduced
foaming stability. Emulsion capacity for the raw black turtle bean flour was lower
(Audu and Aremu 2015). Cooking, roasting and sprouting increased emulsion
capacity while boiling and fermenting reduced it. The relatively low emulsion cap-
acity of the fatted and defatted black turtle bean flours could be due to the nature
and type of protein. This indicates that black turtle bean flour might be useful
in the production of sausages, soups and cakes. Boiling, cooking and sprouting
increased emulsion stability. The swelling index of full-​fat and defatted flour
was lower, probably due to presence of fiber content. The swelling index would
have affected the texture of food prepared from such flours (Okafor et al. 2015).
Wang et al. (2010) reported that cooking black turtle beans in water significantly
increased the content of protein, starch, soluble dietary fiber (SDF), insoluble
dietary fiber (IDF), total dietary fiber (TDF), Mn and P (in dry weight), while it
reduced the ash content, K, Mg, trypsin inhibitory activity (TIA), tannin content,
sucrose and oligosaccharides (raffinose, stachyose and verbascose) content.

6.8 GENERAL CONCLUSIONS
Processing is key to the fractionation of seeds and grains and has been applied
on diverse Latin-​American crops like quinoa, kañiwa, maize, amaranth, chia,
References 243

sacha inchi and legumes. An overview of the different types of milling and their
impact on the characteristics of their by-​products was provided, mainly for
quinoa, kañiwa, amaranth and black turtle bean.
The protein isolation and dietary fiber extraction methods were described
characterizing the main components of each fraction from the different crops
studied. Also, the techno-​functional properties of proteins and the physico-
chemical, thermal and rheological properties of the starches, flours and muci-
lage were extensively summarized.
For oil seeds such as chia, sacha inchi and others, the corresponding extrac-
tion through conventional (cold pressing and solvent extraction) or non-​
conventional processes (supercritical carbon dioxide extraction) were also
described, in terms of the influence of the operating conditions on the oil
yield, the fatty acid composition and minor compounds (tocopherols, phen-
olic compounds).
This chapter brings to the fore knowledge and useful information about the
fractions of these ancient crops and their valuable potential for food and other
outstanding applications

ACKNOWLEDGMENTS
Authors would like to express their gratitude to the CYTED (Spain) for its kind
support through Project la ValSe-​Food 119RT0567, Food4ImNut Food4ImNut
PID2019-107650RB-C21 funded by MCIN/AEI/10.13039/501100011033, Spain.

REFERENCES
Abugoch James, L. E. 2009. Quinoa (Chenopodium quinoa Willd.): Composition, chemistry,
nutritional, and functional properties. Advances in Food and Nutrition Research
58: 1–​31.
Abugoch James, L. E., N. Romero, C. A. Tapia, J. Silva, and M. Rivera. 2008. Study of some
physicochemical and functional properties of quinoa (Chenopodium quinoa Willd)
protein isolates. Journal of Agricultural and Food Chemistry 56(12): 4745–​4750.
Abugoch James, L. E., E. N. Martínez, and M. C. Añón. 2010. Influence of pH on structure
and function of amaranth (A. hypochondriacus) protein isolates. Cereal Chemistry
87: 448–​453.
Agama-​Acevedo, E., M.-​A. Ottenhof, I. A. Farhat, O. Paredes-​López, J. Ortíz-​Cereceres, and
L. A. Bello-​Pérez. 2005. Isolation and characterization of starch from pigmented
maizes. Agrociencia 39(4): 419–​429.
Ahamed, N., R. Singhal, P. Kulkarni, and M. Pal. 1996. Physicochemical and functional
properties of Chenopodium quinoa starch. Carbohydrate Polymers 31: 99–​103.
Ai, Y. and J. Jane. 2016. Macronutrients in corn and human nutrition. Comprehensive
Reviews in Food Science and Food Safety 15(3): 581–​598.
244 Fractionation of Seeds or Grains

Anderson, R.A., and S. A. Watson. 1982. The corn milling industry. In CRC Handbook of
Processing and Utilization in Agriculture, edited by I. A. Wolf, Vol. 2, Part 1, 31–​61.
Boca Ratón, FL CRC Press.
AOAC (Association of Official Agricultural Chemists). 1990. Official Methods of Analysis of
AOAC 925.10:1990, 15th ed.. Washington, DC: AOAC International.
AOAC (Association of Official Agricultural Chemists). 1999. Official Methods of Analysis of
AOAC. Washington, DC: AOAC. Ce 2–​66.
Aranibar, C., N. B. Pigni, M. Martínez et al. 2018. Utilization of a partially-​deoiled chia
flour to improve the nutritional and antioxidant properties of wheat pasta. Journal
of Food Science and Technology 89: 381–​387.
Archer, B. J., S. K. Johnson, H. M. Devereux, and A. L. Baxter. 2004. Effect of fat replace-
ment by inulin or lupin-​kernel fibre on sausage patty acceptability, post-​meal
perceptions of satiety and food intake in men. British Journal of Nutrition
91: 591–​599.
Aremu, M., H. Ibrahim, and T. Bamidele. 2015. Physicochemical characteristics of the
oils extracted from some Nigerian plant foods –​A review. Chemical and Process
Engineering Research 32: 36–​52.
Ashraf, S., S. M. G. Saeed, S. A. Sayeed, and R. Ali. 2012. Impact of microwave treatment
on the functionality of cereals and legumes. International Journal of Agriculture &
Biology 14: 356–​370.
Audu, S., and M. Aremu. 2015. Effect of domestic processing on the levels of some func-
tional parameters in black turtle bean (Phaseolus vulgaris L.). Food Science and
Quality Management 38: 55–​59.
Audu, S., M. Aremu, and L. Lajide. 2013. Effects of processing on physicochemical and
antinutritional properties of black turtle bean (Phaseolus vulgaris L.) seeds flour.
Oriental Journal of Chemistry 29(3): 979–​989.
Avanza, M. V., M. C. Puppo, and M. C. Añón. 2006. Structural characterization of amaranth
protein gels. Journal of Food Science 70: E223–​E229.
Ayerza, R., and W. Coates. 2011. Protein content, oil content and fatty acid profiles as
potential criteria to determine the origin of commercially grown chia (Salvia
hispanica L.). Industrial Crops and Products 34: 1366–​1371.
Baker, L., and P. Rayas-​Duarte 1998a. Freeze-​thaw stability of amaranth starch and the
effects of salt and sugars. Cereal Chemistry 75(3): 301–​307.
Baker, L., and P. Rayas-​Duarte 1998b. Retrogradation of amaranth starch at different
storage temperatures and the effects of salt and sugars. Cereal Chemistry
75(3): 308–​314.
Ballester-​Sánchez, J., J. V. Gil, M. T. Fernández-​Espinar, and C. M. Haros. 2019. Quinoa
wet-​milling: Effect of steeping conditions on starch recovery and quality. Food
Hydrocolloids 89: 837–​843.
Bazile, D., D. Bertero, and C. Nieto. 2013. State of the art report on quinoa around the
world in 2013. Retrieved from www.fao.org/​3/​conte​nts/​ca682​370-​10f8-​40c2-​b084-​
95a8f​704f​44d/​i4042​e00.htm.
Beirão da Costa, M.L. 1989. Aspects of lupin composition as food. In Proceedings of the
Joint CEC–​NCRD Workshop held in Israel (Ginozar Kibbutz). “Lupin production and
bio-​processing for feed, food and other by-​products”: 94–​105.
References 245

Benito-​Román, O., M. Rodríguez-​Perrino, M. T. Sanz, R. Melgosa, and S. Beltrán. 2018.


Supercritical carbon dioxide extraction of quinoa oil: Study of the influence of
process parameters on the extraction yield and oil quality. Journal of Supercritical
Fluids 139: 62–​71.
Betalleluz-​Pallardel, I., M. Inga, L. Mera et al. 2017. Optimisation of extraction conditions
and thermal properties of protein from the Andean pseudocereal cañihua
(Chenopodium pallidicaule Aellen). International Journal of Food Science Technology
52(4): 1026–​1034.
Bolontrade, A. J., A. A. Scilingo, and M. C. Añón. 2013. Amaranth proteins foaming prop-
erties: Adsorption kinetics and foam formation-​ Part 1. Colloids and Surfaces
B: Biointerfaces 105: 319–​327.
Bolontrade, A. J., A. A. Scilingo, and M. C. Añón. 2016. Amaranth proteins foaming prop-
erties: Film rheology and foam stability-​Part 2. Colloids and Surfaces B: Biointerfaces
141: 643–​650.
Boyer, C. D. and C. J. Shannon. 2003. Carbohydrates of the kernel. In Corn: Chemistry
and Technology, 2nd ed., edited by P. J. White and L. A. Johnson, 289–​311. St Paul,
MN: AACC International.
Bustos, M. C., M. I. Ramos, G.T. Pérez, and A. E. León. 2019. Utilization of kañiwa
(Chenopodium pallidicaule Aellen) flour in pasta making. Journal of Chemistry
2019: 4385045.
Calzetta-​Resio, A. N., M. P. Tolaba, and C. Suárez. 2006. Effects of steeping conditions
on wet-​milling attributes of amaranth. International Journal of Food Science and
Technology 41: 70–​76.
Campos, B. E., T. Dias Ruivo, M. R. da Silva Scapim, G. Scaramal Madrona, and R.
Bergamasco. 2016. Optimization of the mucilage extraction process from chia
seeds and application in ice cream as a stabilizer and emulsifier. LWT –​Food Science
and Technology 65: 874−883.
Capitani, M. I., V. Spotorno, S. M. Nolasco, and M. C. Tomás. 2012. Physicochemical and
functional characterization of by-​products from chia (Salvia hispanica L.) seeds of
Argentina. LWT –​Food Science and Technology 45 (1): 94–​102.
Capitani, M. I., S. M. Nolasco, and M. C. Tomás. 2013a. Effect of mucilage extraction on
the functional properties of chia meals. In Food Industry, edited by I. Muzzalupo,
Chap. 19, 421–​437. Rijeka: InTech.
Capitani, M. I., S. M. Nolasco, and M. C. Tomás. 2013b. Characterization and function-
ality of by-​products from chia (Salvia hispanica L.) seeds of Argentina. In Dietary
Fibre: Sources, Properties and their Relationship to Health, edited by D. Betancur-​
Ancona, L. Chel-​ Guerrero, and M. Segura-​ Campos, 141–​ 158. New York: Nova
Science Publishers.
Capitani, M. I., V. Y Ixtaína, S. M Nolasco, and M. C. Tomás. 2013c. Microstructure, chem-
ical composition and mucilage exudation of chia (Salvia hispanica L.) nutlets from
Argentina. Journal of the Science of Food and Agriculture 93 (15): 3856–​3862.
Capitani, M. I., L. J. Corzo-​Ríos, L. A Chel-​Guerrero et al. 2015. Rheological proper-
ties of aqueous dispersions of chia (Salvia hispanica L.) mucilage. Journal of Food
Engineering 149: 70–​77.
Capitani, M. I., S.M Nolasco, and M. C Tomás. 2016. Stability of oil-​in-​water (O/​W)
emulsions with chia (Salvia hispanica L.) mucilage. Food Hydrocolloids 61:
537–​546.
246 Fractionation of Seeds or Grains

Cárdenas, M., C. Carpio, J. Welbaum, E. Vilcacundo, and W. Carrillo. 2018. Chia protein
concentrate (Salvia hispanica L.) anti-​inflammatory and antioxidant activity. Asian
Journal of Pharmaceutical and Clinical Research 11 (2): 382–​386.
Carter, C. and F. Manthey. 2019. Seed treatments affect milling properties and flour quality
of black beans (Phaseolus vulgaris L.). Cereal Chemistry 96(4): 689–​697.
Carvajal-​Larenas, F. E., A. R. Linnemann, M. J. R. Nout, M. Koziol, and M. A. J. S. van Boekel.
2016. Lupinus mutabilis: composition, uses, toxicology, and debittering. Critical
Reviews in Food Science and Nutrition 56: 1454–​1487.
Caselato-​Sousa, V. M. and J. Amaya-​Farfán. 2012. State of knowledge on amaranth grain: A
comprehensive review. Journal of Food Science 77: 93–​10.
Castellani, O. F., E. N. Martínez, and M. C. Añón. 2000. Globulin-​P structure modifications
induced by enzymatic proteolysis. Journal of Agricultural and Food Chemistry
48: 5624–​5629.
Chen H., M. Xiong, T. Bai, D. Chen, Q, Zhang, D. Lin, Y. Liu, A. Liu, Z. Huang, and W. Qin.
2021. Comparative study on the structure, physicochemical, and functional prop-
erties of dietary fiber extracts from quinoa and wheat. LWT –​ Food Science and
Technology 149: 111816.
Chirinos, R., M. Aquino, R. Pedreschi, and D. Campos. 2016. Optimized methodology for
alkaline and enzyme-​assisted extraction of protein from sacha inchi (Plukenetia
volubilis) kernel cake. Journal of Food Process Engineering 40(2): ee12412.
Chirinos, R., K. Ochoa, A. Aguilar-​Galvez et al. 2018. Obtaining of peptides with in
vitro antioxidant and angiotensin I converting enzyme inhibitory activities from
cañihua protein (Chenopodium pallidicaule Aellen). Journal of Cereal Science
83: 139–​146.
Condés, M. C., A. Scilingo, and M. C. Añón. 2009. Characterization of amaranth proteins
modified by trypsin proteolysis. Structural and functional changes. LWT –​ Food
Science and Technology 42: 963–​970.
Coorey, R., A. Tjoe, and V. Jayasena. 2014. Gelling properties of chia seed and flour. Journal
of Food Science 79: 859–​866.
Cordero-​de-​los-​Santos, M. Y., J. A. Osuna-​Castro, A. Borodanenko, and O. Paredes-​López.
2005. Physicochemical and functional characterisation of amaranth (Amaranthus
hypochondriacus) protein isolates obtained by isoelectric precipitation and
micellisation. Food Science and Technology International 11(4): 269–​280.
Córdova-​Noboa, H. A., E. O. Oviedo-​Rondon, A. Ortíz et al. 2020. Effects of corn kernel
hardness and grain drying temperature on particle size and pellet durability when
grinding using a roller mill or hammermill. Animal Feed Science and Technology
271: 114715.
Córdova-​Ramos, J., P. Glorio-​Paulet, A. Hidalgo, and F. Camarena. 2020. Efecto del proceso
tecnológico sobre la capacidad antioxidante y compuestos fenólicos totales del
lupino (Lupinus mutabilis Sweet) andino. Scientia Agropecuaria 11(2): 157–​165.
Cortés-​Avendaño, P., M. Tarvainen, J, P. Suomela et al. 2020. Profile and content of residual
alkaloids in ten ecotypes of Lupinus mutabilis Sweet after aqueous debittering pro-
cess. Plant Foods for Human Nutrition 75: 184–​191.
Cruz-​Vázquez, C., A. Villanueva-​Carvajal, G. Estrada-​Campuzan, and A. Dominguez-​
Lopez. 2019. Tamales texture properties as a function of corn endosperm type.
International Journal of Gastronomy and Food Science 16: 100153.
References 247

Curti, C., R. Curti, N. Bonini, and A. Ramón. 2018. Changes in the fatty acid compos-
ition in bitter Lupinus species depend on the debittering process. Food Chemistry
263: 151–​154.
D’Amico, S., S. Jungkunz, G. Balasz et al. 2019. Abrasive milling of quinoa: Study on the dis-
tribution of selected nutrients and proteins within the quinoa seed kernel. Journal
of Cereal Science 86: 132–​138.
Dąbrowski, G., I. Konopka, S. Czaplicki, and M. Tańska, M. (2017). Composition and oxida-
tive stability of oil from Salvia hispanica L. seeds in relation to extraction method.
European Journal of Lipid Science and Technology 119(5): 1600209.
Dąbrowski, G., I. Konopka, and S. Czaplicki. 2018. Supercritical CO2 extraction in chia oils
production: impact of process duration and co-​solvent addition. Food Science and
Biotechnology 27(3): 677–​686.
Dakhili, S., L. Abdolalizadeh, S. M. Hosseini, S. Shojaee-​Aliabadi, and L. Mirmoghtadaie.
2019. Quinoa protein: Composition, structure and functional properties. Food
Chemistry 299: 125–​161.
de Souza, A. L., F. P. Martínez, S. B. Ferreira, and C. R. Kaiser. 2017. A complete evalu-
ation of thermal and oxidative stability of chia oil. Journal of Thermal Analysis and
Calorimetry 130: 1307–​1315.
Dombrink-​Kurtzman, M.A., and J. A Bietz. 1993. Zein composition in hard and soft endo-
sperm of maize. Cereal Chemistry 70: 105–​108.
Eckey, E. W. 1954. Vegetable Fats and Oils. New York: Reinhold, 75–​79.
Elgeti, D., M. Jekle, and T. Becker. 2015. Strategies for the aeration of gluten-​free bread
-​A review. Trends in Food Science and Technology 46(1): 75–​84.
FAO (Food and Agriculture Organization). 2007. Protein and Amino Acid Requirements
in Human Nutrition. Report of a joint WHO/​FAO/​UNU expert consultation. WHO
Technical Report Series No. 935. Geneva: FAO.
Fernandes, S. S., D. Tonato, M. A. Mazutti et al. 2019. Yield and quality of chia oil extracted
via different methods. Journal of Food Engineering, 262: 200–​208.
Ferreira Ignácio Câmara A. K., P. Kiyomi Okurob, R. Lopes da Cunhab, A. M. Herreroc, C.
Ruiz-​Capillasc, and M. A. Rodrigues Pollonio. 2020. Chia (Salvia hispanica L.) muci-
lage as a new fat substitute in emulsified meat products: Technological, physico-
chemical, and rheological characterization. LWT –​ Food Science and Technology
125: 109193.
Gamel, T. H., A. S. Mesallam, A. A. Damir, L. A. Shekib, and J. P. Linssen. 2007.
Characterization of amaranth seed oils. Journal of Food Lipids 14: 323–​334.
García-​Salcedo, A. J., O. L. Torres-​Vargas, A. del Real, B. Contreras-​Jiménez, and M. E
Rodríguez-​García. 2018. Pasting, viscoelastic, and physicochemical properties of
chia (Salvia hispanica L.) flour and mucilage. Food Structure 16: 59–​66.
Giménez, M. A., C. N. Segundo, M. O. Lobo, and N. C. Samman. 2020. Physicochemical
and techno-​functional characterization of native corn reintroduced in the Andean
Zone of Jujuy, Argentina. Proccedings II Congreso Internacional de Ia ValSe-​Food
Network 53(1): 7.
Goh, K. K. T., L. Matia-​Merino, J. Hong Chiang, R. Quek, S. Jun Bing Soh, and R. G. Lentle.
2016. The physico-​chemical properties of chia seed polysaccharide and its microgel
dispersion rheology. Carbohydrate Polymer 149: 297−307.
248 Fractionation of Seeds or Grains

González, R., E. Tosi, E. Re, M. C. Añon, A. Pilosof, and K. Martinez K 2007. Amaranth
starch-​ rich fraction properties modified by high-​ temperature heating. Food
Chemistry 103: 927–​934.
Guerreo-​Ochoa, M. R., R. Pedreschi, and R. Chirinos. 2015. Optimised methodology for
the extraction of protein from quinoa (Chenopodium quinoa Willd). International
Journal of Food Science and Technology 50(8): 1815–​1822.
Guiotto, E. N., M. I. Capitani, S. M. Nolasco, and M. C. Tomás 2016. Stability of oil-​in-​water
(O/​W) emulsions with sunflower (Helianthus annuus L.) and chia (Salvia hispanica
L.) by-​products. Journal of the American Oil Chemists’ Society 93: 133–​143.
Guzmán-​Maldonado, S. H. and O. Paredes-​López. 1998. Production of high-​protein flours
as milk substitutes. In Functional Properties of Food Components, edited by J. R.
Whitaker and A. Lopez, 66–​79. Washington, DC: American Chemical Society.
Haros, M., M. P. Tolaba, and C. Suarez. 2003. Influence of corn drying on its quality for the
wet-​milling process. Journal of Food Engineering 60: 177–​184.
Haros, M., O. P. Pérez, and C. M. Rosell. 2004. Effect of steeping corn with lactic acid on
starch properties. Cereal Chemistry 81: 10–​14.
Haros, M., W. Blaszczak, O. E. Pérez, J. Sadowska, and C. M. Rosell. 2006. Effect of ground
corn steeping on starch properties. European Food Research and Technology
222: 194–​200.
Hatzold, T., I. Elmadfa, and R. Gross. 1983. Edible oil and protein concentrate from
Lupinus mutabilis. Plant Foods for Human Nutrition 32: 125–​132.
He, S., B. K. Simpson, H. Sun et al. 2018. Phaseolus vulgaris lectins: A systematic review
of characteristics and health implications. Critical Reviews in Food Science and
Nutrition 58 (1): 70–​83.
He, S., J. Zhao, X. Cao et al. 2020. Low pH-​shifting treatment would improve func-
tional properties of black turtle bean (Phaseolus vulgaris L.) protein isolate with
immunoreactivity reduction. Food Chemistry 330: 127217.
Hermansson, A. M. 1986. Soy protein gelation. Journal of the American Oil Chemists’ Society
63(5): 658–​666.
Ishak, I., N. Hussain, R. Coorey, and M. Abd Ghani. 2021. Optimization and character-
ization of chia seed (Salvia hispanica L.) oil extraction using supercritical carbon
dioxide. Journal of CO2 Utilization 45: 101430.
Ixtaína, V. Y. 2010. Caracterización de la semilla y el aceite de chía (Salvia hispanica L.).
Aplicaciones en tecnología de alimentos. Doctoral Thesis, Facultad de Ciencias
Exactas, Universidad Nacional de La Plata (FCE-​UNLP).
Ixtaína, V. Y., A. Vega, S. M. Nolasco et al. 2010. Supercritical carbon dioxide extraction of
oil from mexican chia seed (Salvia hispanica L.): characterization and process opti-
mization. The Journal of Supercritical Fluids 55(1): 192–​199.
Ixtaína, V.Y., M. L. Martínez, V. Spotorno et al. 2011a. Characterization of chia seed oils
obtained by pressing and solvent extraction. Journal of Food Composition and
Analysis 24: 166–​174.
Ixtaína, V. Y., F. Mattea, D. A. Cardarelli et al. 2011b. Supercritical carbon dioxide extrac-
tion and characterization of Argentinean chia seed oil. Journal of the American Oil
Chemists’ Society 88(2): 289–​298.
Jancurová, M., L. Minarovičová, and A. Dandár. 2009. Quinoa –​a review. Czech Journal of
Food Science 27(2): 71–​79.
References 249

Jane, J. L., T. Kasemsuwan, S. Leas et al. 1994. Anthology of starch granule morphology by
scanning electron microscopy. Starch/​Stärke 46: 121–​129.
Jiang, F., Ch. Du, Y. Guo et al. 2020. Physicochemical and structural properties of starches
isolated from quinoa varieties. Food Hydrocolloids 101: 105515.
Jiang, S., J. Ding, J. Andrade et al. 2017. Modifying the physicochemical properties of
pea protein by pH-​ shifting and ultrasound combined treatments. Ultrasonics
Sonochemistry 38: 835–​842.
Jing, P., M. M. Giusti. 2007. Effects of extraction conditions on improving the yield and
quality of an anthocyanin-​rich purple corn (Zea mays L.) color extract. Journal of
Food Science 72: C363–​367.
Joaqui, B. A., A. Bolaños-​Monilla, J. E. Bravo-​Gomez, J. F. Solanilla-​Duque, and D. F. Roa-​
Acosta. 2020. Wet milling of the amaranth grain: Relationship between the sec-
ondary structure of the protein and its ability to form gel. Sylwan 164: 31–​48.
Johnson, S., J. Clements, C. Villarino, and R. Coorey. 2017. Lupins: Their unique nutri-
tional and health-​promoting attributes. In Gluten-​Free Ancient Grains Cereals,
Pseudocereals, and Legumes: Sustainable, Nutritious, and Health-​ Promoting
Foods for the 21st Century, edited by J. R. N. Taylor and J. M. Awika, Chap. 8, 179–​
222. Cambridge: Woodhead Publishing Series in Food Science, Technology and
Nutrition.
Jokić, J., P. Vidović, and K. Aladić. 2014. Supercritical fluid extraction of edible oils. In
Handbook on Supercritical Fluids: Fundamentals, Properties and Applications, edited
by J. Osborne, 205–​228. New York: Nova Science Publishers.
Julio, L. M., J. C. Ruiz-​Ruiz, M. C. Tomás, and M. R. Segura-​Campos. 2019. Chia (Salvia
hispanica) protein fractions: Characterization and emulsifying properties. Journal
of Food Measurement and Characterization 13: 3318–​3328.
Kentish, S. and M. Ashokkumar. 2011. The physical and chemical effects of ultrasound. In
Ultrasound Technologies for Food and Bioprocessing, edited by H. Feng, G. Barbosa-​
Cánovas, and J. Weiss, 1–​12. New York: Food Engineering Series Springer.
Koivunen, E., K. Partanen, S. Perttilä et al. 2016. Digestibility and energy value of
pea (Pisum sativum L.), faba bean (Vicia faba L.) and blue lupin (narrow-​leaf )
(Lupinus angustifolius) seeds in broilers. Animal Feed Science and Technology 218:
120–​127.
Krist, S. 2020. Vegetable Fats and Oils. Cham: Springer.
Lamothe, L. M., S. Srichuwong, B. L. Reuhs, and B. R. Hamaker. 2015. Quinoa (Chenopodium
quinoa W.) and amaranth (Amaranthus caudatus L.) provide dietary fibres high in
pectic substances and xyloglucans. Food Chemistry 167: 490–​496.
León-​Camacho, M., D. L. García-González, and R. Aparicio. 2001. A detailed and compre-
hensive study of amaranth (Amaranthus cruentus L.) oil fatty profile. European Food
Research & Technology 213: 349–​355.
Li, G., S. Wang, and F. Zhu. 2016. Physicochemical properties of quinoa starch.
Carbohydrate Polymers 137: 328–​338.
Lindeboom, N., P. Chang, K. Falk, and R. Tyler . 2005. Characteristics of starch from eight
quinoa lines. Cereal Chemistry 82(2): 216–​222.
Liu, Y., Z. Liu, X. Zhu, X. Hu, H. Zhang, Q. Guo, R. Y. Yada and S. W. Cui. 2021. Seed coat
mucilages: Structural, functional/​bioactive properties, and genetic information.
Comprehensive Reviews in Food Science and Food Safety 20(3): 2534–​2559.
250 Fractionation of Seeds or Grains

López, D. N., R. Ingrassia, P. Busti et al. 2018. Structural characterization of protein isolates
obtained from chia (Salvia hispanica L.) seeds. LWT –​ Food Science and Technology
90: 396–​402.
López, M. G., L.A. Bello-​Pérez, and O. Paredes-López, O. 1994. Amaranth carbohydrates.
In Amaranth –​Biology, Chemistry and Technology, edited by O. Paredes-​Lopez, 107–​
131. Boca Raton, FL:CRC Press.
Luna-​Mercado, G. I. and R. Repo-​Carrasco-​Valencia. 2020. Gluten-​free bread applications:
Thermo-​ mechanical and techno-​ functional characterization of Kañiwa flour.
Cereal Chemistry 00: 1–​8.
Malgor, M., A. C. Sabbione, and A. Scilingo. 2020. Amaranth lemon sorbet, elaboration of a
potential functional food. Plant Foods for Human Nutrition 75: 404–​412.
Manassero, C. A., M. C. Añón, and F. Speroni. 2020. Development of a high protein bev-
erage based on amaranth. Plant Foods for Human Nutrition 75: 599–​607.
Marín Flores, F. M., M. J. Acevedo, R. M. Tamez, M. Nevero, and A. L. Garay. 2008. WO/​
2008/​0044908 Method for obtaining mucilage from Salvia hispanica L. Paris: UN
Word Internacional Property Organization.
Martínez, M. L., M. A. Marín, C. M. Salgado Faller et al. 2012. Chia (Salvia hispanica L.)
oil extraction: Study of processing parameters. LWT –​ Food Science and Technology
47: 78–​82.
Martínez, N. E. and M. C. Añón. 1996. Composition and structural characterization of
amaranth protein isolates: an electrophoretic and calorimetric study. Journal of
Agricultural and Food Chemistry 44: 2523–​2530.
Martínez, N. E., O. F. Castellani, and M. C. Añón. 1997. Common molecular features
among amaranth storage proteins. Journal of Agricultural and Food Chemistry
45: 3832–​3839.
Martínez-​Villaluenga, C., A. Torres, J. Frias, and C. Vidal-​Valverde. 2010. Semolina supple-
mentation with processed lupin and pigeon pea flours improve protein quality of
pasta. LWT –​ Food Science and Technology 43: 617–​622.
Martirosyan, D. M., L. A. Miroshnichenko, S. N. Kulakova, A. V. Pogojeva, and V. I. Zoloedov.
2007. Amaranth oil application for coronary heart disease and hypertension. Lipids
in Health and Disease 6: 1–​12.
Méndez-​Montealvo, G., J. Solorza-​Feria, M. Velázquez del Valle, N. Gómez-​Montiel, O.
Paredes-​López, and L. A. Bello-​Pérez. 2005. Chemical composition and calorimetric
characterization of hybrids and varieties of maize cultivated in Mexico. Agrociencia
39(3): 267–​274.
Méndez-​Montealvo, G., F. J. García-​Suárez, O. Paredes-​López, and L. A. Bello-​Pérez. 2008.
Effect of nixtamalization on morphological and rheological characteristics of
maize starch. Journal of Cereal Science 48: 420–​425.
Mercado, J., C. Elías, and G. Pascual. 2015. Protein isolated from cake of sacha inchi
(Plukenetia volubilis L.) and evaluation of its techno-​functional properties. Anales
Científicos 76 (1): 160–​167.
Mir N. A., S. Riar Ch, and S. Singh. 2021. Improvement in the functional properties of
quinoa (Chenopodium quinoa) protein isolates after the application of controlled
heat-​treatment: Effect on structural properties. Food Structure 28: 100189.
Monroy, Y. M., R. A. F. Rodrigues, A. Sartoratto, and F. Cabral. 2020. Purple corn (Zea
mays L.) pericarp hydroalcoholic extracts obtained by conventional processes
References 251

at atmospheric pressure and by processes at high pressure. Brazilian Journal of


Chemical Engineering 37: 237–​248.
Moscoso-​ Mujica, G., A. Zavaleta, Á. Mujica, M. Santos, and R. Calixto. 2017.
Fraccionamiento y caracterización electroforética de las proteínas de la semilla de
kañihua (Chenopodium pallidicaule Aellen) Fractionation and electrophoretic char-
acterization of (Chenopodium pallidicaule Aellen) kanihua seed proteins. Revista
Chilena de Nutrición 44(2): 144–​152.
Muangrat, R., P. Veeraphong, and N. Chantee. 2018. Screw press extraction of Sacha inchi
seeds: Oil yield and its chemical composition and antioxidant properties. Journal of
Food Processing and Preservation 42(6): ee13635.
Mufari, J. R., P. P. Miranda-​Villa, and E. L. Calandri. 2018a. Quinoa germ and starch sep-
aration by wet milling, performance and characterization of the fractions. LWT –​
Food Science and Technology (96): 527–​534.
Mufari, J., H. A. Gorostegui, P. P. Miranda-​Villa, A. E. Bergesse, and E. L. Calandri. 2020.
Oxidative stability and characterization of quinoa oil extracted from whole meal
and germ flours. Journal of the American Oil Chemists’ Society 97(1): 57–​66.
Mufari, J. R. 2017. Elaboración de subproductos derivados de la quinoa (Chenopodium
quinoa Willd.) a partir de distintos procesos extractivos. Doctoral Thesis, Universidad
Nacional de Córdoba.
Mufari, J., P. Miranda-​Villa, A. Bergesse, N. Cervilla, E. Calandri. 2018b. Physico-​chemical
analysis and protein fraction compositions of different quinoa cultivars. Acta
Alimentaria 47(4): 462–​469.
Muñoz, L., A. Cobos, O. Díaz, and J. M. Aguilera. 2012. Chia seeds: Microstructure, muci-
lage extraction and A hydration. Journal of Food Engineering 108: 216–​224.
Muñoz-​Tebar, N., A. Molina, M. Carmona, and M. I. Berruga. 2021. Use of chia by-​products
obtained from the extraction of seeds oil for the development of new biodegradable
films for the agri-​food industry. Foods 10: 620–​629.
Nardo, A. E., S. Suárez, A. V. Quiroga, and M. C. Añón. 2020. Amaranth as a source of
antihypertensive peptides. Frontiers in Plant Science 11: 1–​15.
Narváez-​González, E. D., J. de Dios Figueroa-​Cárdenas, S. Taba, E. Castaño Tostado, and R.
Martínez Peniche. 2007. Efecto del tamaño del gránulo de almidón de maíz en sus
propiedades térmicas y de pastificado. Revista Fitotecnia Mexicana 30(3): 269–​277.
Nasar-​Abbas, S. and V. Jayasena. 2011. Effect of lupin flour incorporation on the physical
and sensory properties of muffins. Quality Assurance and Safety of Crops and Foods
4: 41–​49.
Okafor, D., R. Enwereuzoh, J. Ibeabuchi et al. 2015. Production of flour types from black
bean (Phaseolus vulgaris) and effect of pH and temperature on functional physico-​
chemical properties of the flours. European Journal of Food Science and Technology
3(2): 64–​84.
Oliveira-​Alves, S. C., D. Barbosa Vendramini-​Costa, C. Bau Betim Cazarin et al. 2017.
Characterization of phenolic compounds in chia (Salvia hispanica L.) seeds, fiber
flour and oil. Food Chemistry 232: 295–​305.
Olivos-​Lugo, B. L., M. A. Valdivia-​López, and A. Tecante. 2010. Thermal and physico-
chemical properties and nutritional value of the protein fraction of Mexican
chia seed (Salvia hispanica L.). Food Science and Technology International 16 (1):
89–​96.
252 Fractionation of Seeds or Grains

Orifici, S. C., M. I. Capitani, M. C. Tomás, and S. M. Nolasco. 2018. Optimization of muci-


lage extraction from chia seeds (Salvia hispanica L) using response surface method-
ology. Journal of the Science of Food and Agriculture 98(12): 4495–​4500.
Ortíz-​Prudencio, S. A. 2006. Determinación de la composición química proximal y fibra
dietaria de 43 variedades criollas de maíz de 7 municipios del Sureste del Estado de
Hidalgo. Tesis de Licenciado en Nutrición. México. Universidad Autónoma, Pachuca
de Soto, México.
Osorio-​Díaz, P., E. Agama-​Acevedo, L. A. Bello-​Pérez et al. 2011. Effect of endosperm type
on texture and in vitro starch digestibility of maize tortillas. LWT –​ Food Science and
Technology 44: 611–​615.
Özcan, M. M., F. Y. Al-​Juhaimi, I. A. Mohamed Ahmed, M. A. Osman, and M. A. Gassem.
2019. Effect of different microwave power setting on quality of chia seed oil obtained
in a cold press. Food Chemistry 278(25): 190–​196.
Padhi, E., R. Liu, M. Hernandez, R. Tsao, and D. Ramdath. 2017. Total polyphenol con-
tent, carotenoid, tocopherol and fatty acid composition of commonly consumed
Canadian pulses and their contribution to antioxidant activity. Journal of Functional
Foods 38: 602–​611.
Pérez, E., Y. A. Bahnassey, and W. M. Breene. 1993. A simple laboratory scale method for
isolation of amaranth starch. Starch/​Stärke 45: 211–​214.
Pérez, O. E., M. Haros, C. Suárez, and C. M. Rosell. 2003. Effect of steeping time on the
starch properties from ground whole corn. Journal of Food Engineering 60: 281–​287.
Prego, I., S. Maldonado, M. Otegui. 1998. Seed structure and localization of reserves in
Chenopodium quinoa. Annals of Botany 82: 481–​488.
Prompiputtanapon, K., W. Sorndech, and S. Tongta. 2020. Surface modification of tapioca
starch by using the chemical and enzymatic method. Starch/​Stärke 72: 1900133.
Punia, S. and S. B. Dhull. 2019. Chia seed (Salvia hispanica L.) mucilage (a
heteropolysaccharide): Functional, thermal, rheological behaviour and its utiliza-
tion. International Journal of Biological Macromolecules 140(1): 1084–​1090.
Qian, J. Y. and M. Kuhn. 1999. Characterization of Amaranthus cruentus and Chenopodium
quinoa starch. Starch/​Stärke 51: 116–​120.
Quinteros, M. F., R. Vilcacundo, C. Carpio, and W. Carrillo. 2016. Isolation of proteins from
sacha inchi Plukenetia volubilis L. in presence of water and salt. Asian Journal of
Pharmaceutical and Clinical Research 9(3): 193–​196.
Quiroga, A., E. N. Martínez, and M. C. Añón. 2007. Amaranth globulin polypeptide hetero-
geneity. Protein Journal 26: 327–​333.
Raigar, R. and H. Mishra. 2018. Grinding characteristics, physical, and flow specific
properties of roasted maize and soybean flour. Journal of Food Processing and
Preservation 42(1): 1–​9.
Rajendran, A., D. Chaudhary, and N. Singh. 2018. Prospecting high oil in corn (Zea mays
L.) germplasm for better quality breeding. Maydica 63(1): 1–​5.
Rascón Cruz, Q., S. Singawa García, J. Osuna Castro, N. Bohorova, and O. Paredes López,
O. 2004. Accumulation, assembly, and digestibility of amarantin expressed in trans-
genic tropical maize. Theoretical and Applied Genetics 108(2): 335–​342.
Repo-​Carrasco-​Valencia, R. 2020. Nutritional value and bioactive compounds in andean
ancient grains. Proceedings II International Conference of La ValSe-​Food Network
“Development of Food Ingredients from Iberoamerican Ancestral Crops” 53: 1–​5.
References 253

Repo-​Carrasco-​Valencia, R., J. Peña, H. Callio, and S. Salminen. 2009. Dietary fiber and
other functional components of two varieties of crude and extruded kiwicha
(Amaranthus caudatus). Journal of Cereal Science 49: 219–​224.
Repo-​Carrasco-​Valencia, R., J. K. Hellström, J. M. Pihlava, and P. H. Mattila. 2010. Flavonoids
and other phenolic compounds in Andean indigenous grains: Quinoa (Chenopodium
quinoa), kañiwa (Chenopodium pallidicaule) and kiwicha (Amaranthus caudatus).
Food Chemistry 120(1): 128–​133.
Repo-​Carrasco-​Valencia, R. and J. Valdez Arana. 2017. Chapter 3: Carbohydrates of
kernels. In Pseudocereals: Chemistry and Technology, edited by C. M. Haros and R.
Schönlechner, chap. 3, 49–​70. Chichester: John Wiley and Sons.
Reyes-​Caudillo, E., A. Tecante, and M. A. Valdivia-​López. 2008. Dietary fibre content
and antioxidant activity of phenolic compounds present in Mexican chia (Salvia
hispanica L.) seeds. Food Chemistry 107(2): 656–​663.
Roa, D. F., P. R. Santagapita, M. P. Buera, and M. P. Tolaba. 2013. Amaranth milling strat-
egies and fraction characterization by FT-​IR. Food and Bioprocess Technology
7: 711–​718.
Roa, D., P. Santagapita, M. Buera, and M. Tolaba. 2014. Ball milling of amaranth starch-​
enriched fraction. Changes on particle size, starch crystallinity, and functionality as
a function of milling energy. Food and Bioprocess Technology 7: 2723–​2731.
Roa-​Acosta, D., C. González-​Callejas, and Y. Calderón-​Yonda. 2017. Control of abra-
sive grinding of amaranth grain to obtain two fractions with industrial potential.
Biotecnología en el sector agropecuario y agroindustrial Special edition 1: 59–​66.
Roa-​Acosta, D. F., J. E. Bravo-​Gómez, M. A. García-​Parra, R. Rodríguez-​Herrera, and
J. F. Solanilla-​ Duque. 2020. Hyper-​ protein quinoa flour (Chenopodium quinoa
Wild): Monitoring and study of structural and rheological properties. LWT –​ Food
Science and Technology 121(4): 108952.
Roberfroid, M. 2007. Prebiotics: The concept revisited. Journal of Nutrition 137: 830S–​837S.
Rocha Uribe, J. A., J. Y. Novelo Pérez, H. Castillo Kauil, G. Rosado Rubio, and C. G. Alcocer.
2011. Extraction of oil from chia seeds with supercritical CO2. The Journal of
Supercritical Fluids 56(2): 174–​178.
Rose, D.J., G. E. Inglett, and S. X. Liu. 2010. Utilization of corn (Zea mays) bran and corn
fiber in the production of food components. Journal of the Science of Food and
Agriculture 90: 915–​924.
Ruiz, C., C. Díaz, J. Anaya, and R. Rojas. 2013. Proximate analysis, antinutrients, fatty acids
and amino acids profiles of seeds and cakes from 2 species of sacha inchi: Plukenetia
volubilis and Plukenetia huayllabambana. Revista de la Sociedad Química del Perú
79(1): 29–​36.
Ruiz, G. A., W. Xiao, M. Van Boekel, M. Mino, and M. Stieger. 2016. Effect of extraction pH
on heat-​induced aggregation, gelation and microstructure of protein isolate from
quinoa (Chenopodium quinoa Willd). Food Chemistry 209: 203–​210.
Sabbione, A. C., A. E. Nardo, M. C. Añón, and A. A. Scilingo. 2016a. Amaranth peptides with
antithrombotic activity released by simulated gastrointestinal digestion. Journal of
Functional Foods 20: 204–​214.
Sabbione, A. C., G. Rinaldi, M. C. Añón, and A. A. Scilingo. 2016b. Antithrombotic effects
of Amaranthus hypochondriacus proteins in rats. Plant Foods for Human Nutrition
71: 19–​27.
254 Fractionation of Seeds or Grains

Sabbione, A. C., S. E. Suárez, M. C. Añón, and A. A. Scilingo. 2019. Amaranth func-


tional cookies exert potential antithrombotic and antihypertensive activities.
International Journal of Food Science and Technology 54: 1506–​1513.
Salas-​Valero, L. M., D. R. Tapia-​Blácido, and F. C. Menegalli. 2015. Biofilms based on
canihua flour (Chenopodium pallidicaule): design and characterization. Química
Nova 38(1): 14–​21.
Salcedo-​Chávez, B., J. A. Osuna-​Castro, F. Guevara-​Lara, J. Domínguez-​Domínguez, and
O. Paredes-​López. 2002. Optimization of the isoelectric precipitation method to
obtain protein isolates from amaranth (Amaranthus cruentus) seeds. Journal of
Agricultural and Food Chemistry 50: 6515–​6520.
Saldaña, E., J. Ríos-​Mera, H. Arteaga et al. 2018. How does starch affect the sensory
characteristics of mazamorra morada? A study with a dessert widely consumed by
Peruvians. International Journal of Gastronomy and Food Science 12: 22–​30.
Salgado-​Cruz, M. P., G. Calderón-​Domínguez, J. Chanona-​Pérez et al. 2013. Chia (Salvia
hispanica L.) seed mucilage release characterisation. A microstructural and image
analysis study. Industrial Crops and Products 51: 453–​462.
Salinas, M. Y., M. Soto, F. Martínez, V. González, and R. Ortega. 1999. Análisis de
antocianinas en maíces de grano azul y rojo provenientes de cuatro razas. Revista
Fitotecnica. Mexicana 22: 161–​174.
Salvador-​Reyes, R. and M. T. Silva Pedrosa Clerici. 2020. Peruvian Andean maize: General
characteristics, nutritional properties, bioactive compounds, and culinary uses.
Food Research International 130: 108934.
Salvador-​Reyes, R., A. P. Rebellato, J. Azevedo Lima Pallone, R. A. Ferrari, and M. T. Pedrosa
Silva Clerici. 2021. Kernel characterization and starch morphology in five varieties
of Peruvian Andean maize. Food Research International 140: 110044.
Sandoval-​Oliveros, M. R. and O. Paredes-​López. 2013. Isolation and characterization
of proteins from chia seeds (Salvia hispanica). Journal of Agricultural and Food
Chemistry 61: 193–​201.
Sarifudin, A., T. Keeratiburana, S. Soontaranon, C. Tangsathitkulchai, and S. Tongta.
2020. Pore characteristics and structural properties of ethanol-​treated starch
in relation to water absorption capacity. LWT –​ Food Science and Technology
129: 109555.
Sathe, S. K., and D. K. Salunkhe. 1981. Functional properties of the great northern bean
(Phaseolus vulgaris L.) proteins: Emulsion, foaming, viscosity, and gelation proper-
ties. Journal of Food Science 46(1): 71–​81.
Sathe, S. K., B. R. Hamaker, K. Wai, C. Sze-​Tao, and M. Venkatachalam. 2002. Isolation,
purification and biochemical characterization of a novel water-​soluble protein
from Inca peanut (Plukenetia volubilis). Journal of Agricultural and Food Chemistry
50(17): 4906–​4908.
Scilingo, A., S. Molina Ortíz, E. N. Martínez, and M. C. Añón. 2002. Amaranth protein
isolates modified by hydrolytic and thermal treatments. Relationship between
structure and solubility. Food Research International 35: 855–​865.
Segura-​Campos, M. R., I. M. Salazar-​Vega, L. A. Chel-​Guerrero, and D. A. Betancur-​Ancona.
2013a. Biological potential of chia (Salvia hispanica L.) protein hydrolysates and
their incorporation into functional foods. LWT –​ Food Science and Technology
50(2): 723–​731.
References 255

Segura-​Campos, M. R., F. Peralta-​González, L. A. Chel-​Guerrero, and D. A. Betancur-​


Ancona. 2013b. Angiotensin I-​converting enzyme inhibitory peptides of chia (Salvia
hispanica) produced by enzymatic hydrolysis. International Journal of Food Science
2013: 158482.
Segura-​Campos, M. R., Ciau-​Solís N., Rosado-​Rubio G., Chel-​Guerrero L. A. and Betancur-​
Ancona D. A. 2014. Chemical and functional properties of chia seed (Salvia hispanica
L.) gum. International Journal of Food Science 2014: 241053.
Segura-​Nieto, M., A. P. Barba de la Rosa, and O. Paredes-​López. 1994. Biochemistry of
amaranth proteins. In Amaranth: Biology, Chemistry and Technology, edited by O.
Paredes-​López, 75–​106. Boca Ratón, FL: CRC Press.
Selma-Gracia, R., J. M. Laparra, and C. M. Haros. 2020. Potential beneficial effect of hydro-
thermal treatment of starches from various sources on in vitro digestion. Food
Hydrocolloids 103: 105687.
Serna-​Saldivar, S. O. and C. Chuck-Hernández. 2019. Food uses of lime-​cooked corn with
emphasis in tortillas and snacks. In Corn: Chemistry and Technology, edited by S. O.
Serna-​Saldivar, 469–​500. Washington, DC: AACC International Press.
Serna-​Saldivar, S. O., and C. E. Pérez Carrillo. 2019. Food uses of whole corn and dry-​
milled fractions. In Corn: Chemistry and Technology, edited by S. O. Serna-​Saldivar,
435–​467. Washington, DC: AACC International Press.
Silva, C., V. A. dos Santos García, and C. M. Zanette. 2016. Chia (Salvia hispanica L.) oil
extraction using different organic solvents: oil yield, fatty acids profile and techno-
logical analysis of defatted meal. International Food Research Journal 23(3): 998–​1004.
Silva-​Sánchez, C., A. P. Barba de la Rosa, M. F. León-​Galván et al. 2008. Bioactive peptides
in amaranth (Amaranthus hypochondriacus) seed. Journal of Agricultural and Food
Chemistry 56: 1233–​1240.
Singh, B., J. P. Singh, A. Kaur, and N. Singh. 2017. Phenolic composition and antioxidant
potential of grain legume seeds: A review. Food Research International 101: 1–​16.
Sironi, E., F. Sessa, and M. Duranti. 2005. A simple procedure of lupin seed protein frac-
tionation for selective food applications. European Food Research and Technology
221: 145–​150.
Sisti, M. S., A. Scilingo, and M. C. Añón. 2019. Effect of the incorporation of amaranth
(Amaranthus mantegazzianus) into fat-​and cholesterol-​rich diets for Wistar rats.
Journal of Food Science 82: 3075–​3082.
Solaesa, Á. G., M. Villanueva, A. J. Vela, and F. Ronda. 2020. Protein and lipid enrichment
of quinoa (cv.Titicaca) by dry fractionation. Techno-​functional, thermal and rheo-
logical properties of milling fractions. Food Hydrocolloids 105: 1–​9.
Srichuwong, S., D. Curti, S. Austin et al. 2017. Physicochemical properties and starch
digestibility of whole grain sorghums, millet, quinoa and amaranth flours, as
affected by starch and non-​starch constituents. Food Chemistry 233: 1–​10.
Stamenković, O. S., M. D. Kostić, M. B.Tasić et al. 2020. Kinetic, thermodynamic and
optimization study of the corn germ oil extraction process. Food and Bioproducts
Processing 120: 91–​103.
Steffolani, M. E, A. E. León, and G. A. Pérez. 2013. Study of the physicochemical and func-
tional characterization of quinoa and kañiwa starches. Starch/​Stärke 65: 976–​983.
Suárez, S. E. and M. C. Añón. 2018. Comparative behaviour of solutions and dispersions of
amaranth proteins on their emulsifying properties. Food Hydrocolloids 74: 115–​123.
256 Fractionation of Seeds or Grains

Tang, H., K. Watanabe, and T. Mitsunaga, T. 2002. Characterization of storage starches


from quinoa, barley and adzuki seeds. Carbohydrate Polymers 49: 13–​22.
Tang, Y., X. Li, B. Zhang, P. Chen, R. Liu, and R. Tsao. 2015. Characterisation of phenolics,
betanins and antioxidant activities in seeds of three Chenopodium quinoa Willd.
genotypes. Food Chemistry 166: 380–​388.
Tari, T., U. Annapure, R. Singhal, and P. Kulkarni. 2003. Starch-​ based spherical
aggregates: Screening of small granule sized starches for entrapment of a model
flavouring compound, vanillin. Carbohydrate Polymers 53: 45–​51.
Tavares, L.S., L.A. Junqueira, I. C. de Oliveira Guimarães, and J. Vilela de Resende. 2018.
Cold extraction method of chia seed mucilage (Salvia hispanica L.): effect on yield
and rheological behavior. Journal of Food Science and Technology 55: 457–​466.
Taylor, J., J. O. Anyango, P. J. Muhiwa, S. L. Oguntoyinbo, and J.R. Taylor. 2018. Comparison
of formation of visco-​elastic masses and their properties between zeins and kafirins.
Food Chemistry 245: 178–​188.
Tester, R.F. 1997. Influence of growth conditions on barley starch properties. International
Journal of Biological Macromolecules 21: 37–​45.
Thakur, S., M. G. Scanlon, R. T. Tyler, A. Milani, and J. Paliwal. 2019. Pulse flour
characteristics from a wheat flour miller’s perspective: A comprehensive review.
Comprehensive Reviews in Food Science and Food Safety 18 (3): 775–​97.
Thoufeek Ahamed, N., R. S. P. R. Singhai, and M. Kulkarni Pal. 1998. A lesser-​known grain,
Chenopodium quinoa: Review of the chemical composition of its edible parts. Food
and Nutrition Bulletin 19(1): 61–​70.
Timilsena, Y. P., R. Adhikari, S. Kasapis, and B. Adhikari. 2015. Rheological and micro-
structural properties of the chia seed polysaccharide. International Journal of
Biological Macromolecules 81: 991–​999.
Timilsena, Y. P., B. Wang, R. Adhikari, and B. Adhikari. 2016a. Preparation and charac-
terization of chia seed protein isolate–​chia seed gum complex coacervates. Food
Hydrocolloids 52: 554–​563.
Timilsena, Y. P., R. Adhikari, S. Kasapis, and B. Adhikari. 2016b. Molecular and functional
characteristics of purified gum from Australian chia seeds. Carbohydrate Polymers
136: 128−136.
Timilsena, Y. P., J. Vongsvivut, R. Adhikari, and B. Adhikari. 2017. Physicochemical and
thermal characteristics of Australian chia seed oil. Food Chemistry 228: 394–​402.
Torrejón, I., B. L. Martín, T. B. de la Puente, J. R. Nasser, and R. Rizzi. 2016. La Kañiwa: nueva
alternativa alimentaria para la prevencion de la desnutricion y las enfermedades
cardiovasculares. Revista de Salud Pública 20(2): 17.
Tosi, E., E. Re, H. Lucero, and R. Masciarelli. 2000. Amaranth (Amaranthus spp.) grain
conditioning to obtain hyperproteic flour by differential milling. Food Science and
Technology International 5: 60–​63.
Tosi, E. A., E. Ré, H. Lucero, and R. Masciarelli. 2001. Dietary fiber obtained from
amaranth (Amaranthus cruentus) grain by differential milling. Food Chemistry 73:
441–​443.
Triana-​Maldonado, D. M., S. Torijano-​ Gutiérrez, and C. Giraldo-​ Estrada. 2017.
Supercritical CO2 extraction of oil and omega-​3 concentrate from Sacha inchi
(Plukenetia volubilis L.) from Antioquia, Colombia. Grasas y Aceites 68(1): e172.
References 257

Urban, M., M. Beran, L. Adamek et al. 2012. Cyclodextrin production from amaranth
starch by cyclodextrin glycosyltransferase produced by Paenibacillus macerans
CCM. Czech Journal of Food Science 30: 15–​20.
Urbizo-​Reyes, U., M. F. San Martin-​González, J. García-​Bravo, A. López Malo Vigil, and
A.M. Liceaga. 2019. Physicochemical characteristics of chia seed (Salvia hispanica)
protein hydrolysates produced using ultrasonication followed by microwave-​
assisted hydrolysis. Food Hydrocolloids 97: 105–​187.
Uriarte-​Aceves, P., E. Cuevas-​Rodriguez, R. Gutiérrez Dorado et al. 2015. Physical, com-
positional, and wet-​milling characteristics of Mexican blue maize (Zea mays L.)
landrace. Cereal Chemistry Journal 92: 491–​496.
Vázquez Batti, A. I., A. C. Resio, and M. P.Tolaba. 2006. Optimization of amaranth wet
milling in alkaline medium. Proceedings XXII Interamerican Congress of Chemical
Engineering, CIIQ 2006 and V Argentinian Congress of Chemical Engineering, CAIQ
2006 –​Innovation and Management for Sustainable Development. October, Buenos
Aires, pp. 811–​816.
Vázquez-​Ovando, A., G. Rosado-​Rubio, L. A. Chel-​Guerrero, and D. A. Betancur-​Ancona.
2009. Physicochemical properties of a fibrous fraction from chia (Salvia hispanica
L.). LWT –​ Journal of Food Science and Technology 42: 168–​173.
Vázquez-​Ovando, A., G. Rosado-​Rubio, L. A. Chel-​Guerrero, and D. A. Betancur-​Ancona.
2010. Dry processing of chia (Salvia hispanica L.) flour: Chemical characterization
of fiber and protein. CyTA Journal of Food 8(2): 117–​127.
Vázquez-​ Ovando, A., D. A. Betancur-​ Ancona, and L. A. Chel-​ Guerrero. 2013.
Physicochemical and functional properties of a protein-​rich fraction produced
by dry fractionation of chia seeds (Salvia hispanica L.). CyTA Journal of Food
11(1): 75–​80.
Velarde-​Salcedo, A. J., E. Bojórquez-​Velázquez, and A. P. B. de la Rosa. 2019. Amaranth. In
Whole Grains and their Bioactives: Composition and Health, edited by J. Johnson and
T. C. Wallace, 209–​250. Hoboken, NJ: John Wiley & Sons.
Velázquez-​Gutiérrez, S. K., A. C. Figueira, M. E. Rodríguez-​Huezo, A. Román-​Guerrero, H.
Carrillo-​Navas, C. Pérez-​Alonso. 2015. Sorption isotherms, thermodynamic proper-
ties and glass transition temperature of mucilage extracted from chia seeds (Salvia
hispanica L.). Carbohydrate Polymers 121: 411–​419.
Ventureira, J., E. N. Martínez, and M. C. Añón. 2012a. Effect of acid treatment on struc-
tural and foaming properties of soy amaranth protein mixtures. Food Hydrocolloids
29: 272–​279.
Ventureira, J. L., A. J. Bolontrade, F. Speroni et al. 2012b. Interfacial and emulsifying prop-
erties of amaranth (Amaranthus hypochondriacus) protein isolates under different
conditions of pH. LWT –​ Food Science and Technology 45: 1–​7.
Villacrés, E., M. Navarrete, O. Lucero, S. Espín, and E. Peralta. 2010. Evaluación del
rendimiento, características físico-​químicas y nutraceúticas del aceite de chocho
(Lupinus mutabilis sweet). Revista Tecnológica ESPOL –​RTE 23: 57–​62.
Villanueva-​Bermejo, D., M. V. Calvo, P. Castro-​Gómez, T. Fornari, and J. Fontecha. 2019.
Production of omega 3-​ rich oils from underutilized chia seeds. Comparison
between supercritical fluid and pressurized liquid extraction methods. Food
Research International 115: 400–​407.
258 Fractionation of Seeds or Grains

Villanueva-​Bermejo, D., T. Fornari, M. V. Calvo et al. 2020. Application of a novel


approach to modelling the supercritical extraction kinetics of oil from two sets of
chia seeds. Journal of Industrial and Engineering Chemistry 82: 317–​323.
Villarino, C. B. J., V. Jayasena, R. Coorey, S. Chakrabarti-​Bell, and S. K. Johnson. 2016.
Nutritional, health and technological functionality of lupin flour addition to bread
and other baked products: Benefits and challenges. Critical Reviews in Food Science
and Nutrition 26: 835–​857.
Wandersleben, T., E. Morales, C. Burgos-​Díaz et al. 2018. Enhancement of functional and
nutritional properties of bread using a mix of natural ingredients from novel var-
ieties of flaxseed and lupine. LWT –​ Food Science and Technology 91: 48–​54.
Wang, N., D. Hatcher, R. Tyler, R. Toews, and E. Gawalko. 2010. Effect of cooking on the
composition of beans (Phaseolus vulgaris L.) and chickpeas (Cicer arietinum L.).
Food Research International 43(2): 589–​594.
Weber, E. J. 1979. The lipids of corn germ and endosperm. Journal of the American Oil
Chemists’ Society 56: 637–​641.
Willis, W. M. and A. G. Marangoni. 2017. Food lipids. Chemistry, nutrition, and biotech-
nology. In Food Lipids: Chemistry, Nutrition, and Biotechnology, edited by C. Akoh,
899–​939. Boca Ratón, FL: CRC Press.
Wong, A., K. Pitts, V. Jayasena, and S. Johnson. 2013. Isolation and foaming functionality of
acidsoluble protein from lupin (Lupinus angustifolius) kernels. Journal of Science of
Food and Agriculture 93: 3755–​3762.
Zanqui, A. B., C. Marques da Silva, D. Rodrigues de Morais et al. 2016. Sacha inchi
(Plukenetia volubilis L.) oil composition varies with changes in temperature and
pressure in subcritical extraction with n-​propane. Industrial Crops and Products
87: 64–​70.
Zarkadas, G. G., Z. R. Yu, R. I. Hamilton, P. L. Pattison, and N. G. W. Rose. 1995. Comparison
between the protein-​quality of northern adapted cultivars of common maize and
quality protein maize. Journal of Agricultural and Food Chemistry 43: 84–​93.
Zaworska, A., M. Kasprowicz-​Potocka, A. Frankiewicz, Z. Zduńczyk, and J. Juśkiewicz.
2017. Effects of fermentation of narrow-​leafed lupine (L. angustifolius) seeds on
their chemical composition and physiological parameters in rats. Journal of Animal
and Feed Science 25: 326–​334.
Zhu, F. 2017. Structures, physicochemical properties, and applications of amaranth
starch. Critical Reviews in Food Science and Nutrition 57(2): 313–​325.
Chapter 7

Food Uses of Selected


Ancient Grains
Claudia M. Haros1, Marcela Lilian Martínez2,
Bernabé Vázquez Agostini3, and Loreto A. Muñoz4
1Instituto de Agroquímica y Tecnología de

Alimentos (IATA), Paterna, Valencia, Spain


2Instituto Multidisciplinario de Biología Vegetal (IMBIV,

CONICET-​UNC), Córdoba, Argentina


3Universidad Nacional de Jujuy and Consejo

Nacional de Investigaciones Científicas y Técnicas


(CONICET), San Salvador de Jujuy, Argentina
4Universidad Central de Chile, Santiago de Chile

CONTENTS
7.1 Introduction 260
7.2 Bakery and Bread Products 261
7.3 Biscuits, Cookies and Cakes 269
7.4 Pasta Products 273
7.5 Snacks and Breakfast Mixtures 276
7.6 Drinks and Fermented Beverages 279
7.7 Infant Foods 281
7.8 Traditional Foods 283
7.9 Haute Cuisine 287
7.10 Conclusions 287
Acknowledgments 289
References 289

DOI: 10.1201/9781003088424-7 259


260 Food Uses of Selected Ancient Grains

7.1 INTRODUCTION
The consumption of grains, cereals and pseudo-​cereals such as Andean maize
(Zea mays), amaranth (Amaranthus spp), quinoa (Chenopodium quinoa Willd),
kañiwa (Chenopodium pallidicaule Aellen), chia (Salvia hispanica L.), sacha inchi
(Plukenetia volubilis) and legumes as black turtle bean (Phaseolus vulgaris) and
tarwi (Lupinus mutabilis) date backs to ancient times. Andean crops have been
cultivated for thousands of years (since pre-​Colombian times) including pseudo-​
cereals such as quinoa, kañiwa and amaranth, as well as the Andean lupin or
tarwi and several pigmented varieties/​races of corns. Purple corn may well be
the most representative and consumed among this type of cereal (Ayala, 1998;
Ranilla et al.; Apostolidis et al., 2009). On the other hand, chia is originally from
the central valley of Mexico and northern Guatemala (Peláez et al., 2019); it
began to be used in human food around 3500 BC and acquired importance as
a staple crop in Central Mexico between 1500 and 900 BC (Ayerza and Coates,
2005; Cahill, 2003). The seed was one of the main crops of pre-​Columbian soci-
eties, consumed by Aztecs and Mayans in many food preparations; it was also
used in medicine and paintings –​surpassed only by corn and beans (Muñoz
et al., 2012). Finally, black beans are native to some regions of the American con-
tinent and were domesticated from the wild legume distributed from northern
Mexico to northeastern Argentina; nowadays black beans are popular around
the world and consumed in various preparations. An example would be black
bean-​chili in Bolivia, “feijoada” in Brazil, “gallo pinto” in Costa Rica, mixed with
rice called “Moros y Cristianos” in Cuba, in soup called “sopa de frijoles negros”
in Puerto Rico and in a thick corn tortilla (pupusa) stuffed with black beans or
in many dishes in Mexico like intact or refried (boiled, mashed) with rice and
tortillas, in stews, in soups, in mixed dishes or in casseroles (The Bean Institute,
2014; Wright, 2008).
All these crops have played an important role in human nutrition as sources
of nutrients. There are multiple and diverse food uses of these grains, depending
on the region or country where they are consumed; for example, in the quo-
tidian Andean diet, plant foods were usually boiled into soups or stews in
large pots and commonly consumed or in bowls or on plates (Hastorf, 2015).
In ancient times, grains were consumed whole and/​or ground in millstone or
small stone mortar, to later be incorporated into different foods such as bread,
stews and diverse traditional dishes. In the last two decades, these ancestral
grains have been transformed from denigrated foods (produced and consumed
almost exclusively by people from rural locations) into new ingredients thanks
to the knowledge we have about them today, but also thanks to the chefs who
have played a leading role in the revalorization and rediscovering of these
7.2 Bakery and Bread Products 261

ancestral Latin-​American crops (Ayerza and Coates, 2005; Izquierdo and Roca,
1998; McDonell, 2019).
This chapter is focused on the food uses of ancient Latin-​American grains and
how they have been used in food consumption throughout history, and how they
are an important component in our daily or common diets.

7.2 BAKERY AND BREAD PRODUCTS


A growing number of investigations have studied the use of alternative ingredients
to wheat in the production of nutrient-​rich bakery products. For the inclusion of
high levels of whole flours/​seeds of alternative cereals/​pseudo-​cereals/​oilseeds/​
legumes in bread, it would be necessary to make modifications to traditional
technological breadmaking procedures, which could enable the development of
a range of new baking products with enhanced nutritive and sensory value. The
addition of these unconventional flours, particularly in high amounts, results in
technological challenges such as: increases in dough yield, resulting in moist and
shorter dough; decreases in fermentation tolerance; lower volume, tense and non-​
elastic crumb; and various flavor changes, depending on ingredients and bread
type, due to the higher amount of fiber and dilution of gluten (Sanz-​Penella et al.,
2013; Sanz-​Penella and Haros, 2017). In this sense, pseudo-​cereals are an excellent
alternative, being more frequently used due to consumer demand. In particular,
amaranth is the pseudo-​cereal that has been mostly studied for breadmaking
purposes (Table 7.1).
The use of amaranth in the formulation gradually and significantly increases
protein, lipid, minerals and dietary fiber content compared to white bread
(Table 7.1). The inclusion of whole amaranth flour also shows a slight tendency
to decrease specific volume and increases crumb firmness, whereas crumb
structure sometimes exhibits no significant changes (Sanz-​Penella et al., 2013;
Miranda-Ramos et al., 2019). Color tristimulus values were significantly affected
when whole amaranth flour was used in both crumb and crust. Despite bread
made with amaranth not achieving greater acceptability than wheat bread, par-
ticularly with a high percentage of substitution, consumers concluded they pre-
ferred amaranth bread due to its more nutritious condition even though its taste
and aroma were different from traditional bread.
Popped amaranth was also evaluated as a bread supplementation (Bodroza-​
Solarov et al., 2008), which had a similar effect on bread quality than whole
amaranth flours. If the purpose of pseudo-​cereal inclusion in bread making
is to increase protein amount and quality, protein isolates could be used in
order to avoid quality deterioration of the end product (Tömösközi et al., 2011).
newgenrtpdf
262
Food Uses of Selected Ancient Grains
TABLE 7.1 CHARACTERISTICS OF BREAD MADE WITH AMARANTH AND OTHER INGREDIENTS
Level of
Substitution
Amaranthus Ingredient in flour basis Formulation Benefits Loss of quality References
A. caudatus Whole Flour up to 100% Wheat ↑Nutritional value ↑Crumb Firmness, Rosell et al.,
↓Specific Volume, 2009
↓Acceptability
Whole Flour up to 40% Wheat, Chia, ↑Nutritional value, ↑Crumb Firmness, Miranda-
and/​or Quinoa ↑Minerals (Ca, Fe, Zn) ↓Specific Volume, Ramos, 2020
↑Phytates
Whole Flour up to 40% Wheat ↑Nutritional value, ↑Antinutritional Zula et al.,
↓Microbial load Components, 2020
↓Acceptability
A. cruentus Popped Wheat ↑Crumb quality, taste, N.D Simurina
Grains odour and freshness et al., 2005

Popped up to 20% Wheat ↑Nutritional value, ↑Crumb Firmness, Bodroza-​


grains ↑Squalene, ↑color ↓Crumb Elasticity, Solarov et al.,
and flavor, ↑Uniform ↑Crumb Density 2008
porosity
Whole flour up to 40% Wheat, phytase ↑Nutritional Value, ↓Specific Volume, Sanz-​Penella
↑Fe Availability, ↑Crumb Firmness et al., 2012
↓Phytates
Whole flour up to 50% Wheat ↑Nutritional Value ↓Specific Volume, Sanz-​Penella
↑Crumb Firmness, et al., 2013
↑Colour
newgenrtpdf
Whole flour up to 50% Wheat, ↑Nutritional Value, ↓Specific Volume, Garcia-​
Bifidobacterium ↓Phytates, ↑Mineral ↑Crumb Firmness Mantrana
phytases Availability (Ca, Fe, Zn) et al., 2014

Whole flour 25% Wheat ↑improved the Hb N.D Laparra &


concentrations, ↓GI Haros, 2016

A. hypochondriacus Whole flour up to 50% Chapatti ↑Nutritional Value, N.D. Banerji et al.,
Indean Flat ↑Fe, Ca, Mg, ↑Lysine, 2018
Bread ↑Protein Digestibility
Whole flour, up to 25% Wheat ↑Specific Volume, ↓Whitness, ↓Shape Mykolenko
Flakes and ↑Porosity, ↑Nutritional stability et al., 2020
croats Value ↓Consumer
characteristics
Whole flour up to 50% Wheat ↑Nutritional Value ↓Crumb Staling, Miranda-
↑Phytates, Ramos et al.,
↓Acceptability 2019

Whole flour 25% Wheat ↓GI in vivo N.D. Laparra &

7.2
Haros, 2018

Bakery and Bread Products


Amaranthus dubius Whole flour up to 20% Wheat ↑Nutritional Value, N.D. Montero-​
Mart. ex Thell ↑Digestibility Quintero
et al., 2015
A. spinosus Whole flour up to 50% Wheat ↑Nutritional Value ↓Crumb Staling, Miranda-
↑Phytates, Ramos et al.,
↓Acceptability 2019

Source: Authors’ own work.

263
264 Food Uses of Selected Ancient Grains

Another option would be to use defatted hyperproteic amaranth flours as an


alternative ingredient to supplement breads (Tosi et al., 2002). Traditional
Brazilian bread made with cheese was also studied with the addition of amar-
anth flour (Lemos et al., 2012) or as a raw material for sourdough fermentation
(Jekle et al., 2010; Houben et al., 2010). Amaranth is a suitable ingredient for
fermentation by various species of Lactobacillus, and sourdough fermentation
was able to produce doughs with viscosity and elasticity similar to those found
in pure wheat flours (Houben et al., 2010). However, more balanced fermen-
tation quotient values should encourage improvement in bread flavor (Jekle
et al., 2010).
Quinoa (Chenopodium quinoa) was also investigated for the development
of lactic ferment as an alternative to biopreservation of packed bread. It is an
optimal substrate for growth and production of improved amounts of antifungal
compounds by Lactobacillus plantarum isolated from sourdough (phenyllactic
and hydroxyphenyllactic acids), allowing a quantity reduction of the chemical
preservative calcium propionate commonly added to bread (Dallagnol et al.,
2015). Strains of Lactobacillus plantarum and Lactococcus lactis subsp. lactis,
previously selected for biosynthesis of γ-​aminobutyric acid (GABA), were used
for sourdough fermentation of cereal, pseudo-​cereal and leguminous flours,
where pseudo-​cereals and chickpea flours were the most suitable for enrich-
ment with GABA (Coda et al., 2010). In general, whole quinoa flour could be
a good replacement for wheat flour (Figure 7.1.A) in bread formulations,
increasing the product’s nutritional value in terms of dietary fiber, minerals,
proteins and healthy fats, with only a small depreciation in bread quality (Stikic
et al., 2012; Iglesias-​Puig et al., 2015; Ballester et al., 2019a, 2019b, 2019c) taking
into account that a high level of substitution did not show the small depreci-
ation in bread quality (Rosell et al., 2009; Wang et al., 2015). The same trend was
observed through the replacement of wheat flour by 25% kañiwa (Chenopodium
pallidicaule) with a good sensory acceptability but variable in color (Rosell
et al., 2009).
Compared to wheat breads, the products resulting from composed flours
such as quinoa, amaranth or kañiwa/​wheat, affect their quality in terms of reduc-
tion of specific volume, increment of density, firmness and chewiness of crumb
bread, its darkness, redness and/​or yellowness (depending on the color of the
raw material) with acceptable marks by consumers (Rosell et al., 2009; Iglesias-​
Puig et al., 2015; Ballester et al., 2019a, 2019b; Miranda-Ramos et al., 2019).
Despite the change of color, pseudo-​cereal breads were well accepted, showing
good commercial potential among consumers, with most of those interviewed
stating that they would buy the product mainly for the health benefits it could
bring (Calderelli et al., 2010; Bilgiçli and İbanoğlu, 2015; Iglesias-​Puig et al., 2015;
Miranda-Ramos et al., 2019; Ballester et al., 2019b; Miranda-​Ramos and Haros,
7.2 Bakery and Bread Products 265

Figure 7.1 Foods with quinoa: A. Bread with whole black quinoa flour, 25%; B. Quinoa
balls with cocoa; C. Chocolate with quinoa; D. Quinoa and buckwheat cookies; E. Quinoa,
cinnamon and lemon cookies; F. Quinoa cookies with cocoa; G. Crispy chocolate quinoa
toast; H. Puffy quinoa, compared with quinoa seeds. Source: Authors’ own work.
Notes: A) A slice of bread with 25% of whole black quinoa flour bread.
B) Four, thin rectangular quinoa and buckwheat crackers, in a cup; one side of the
cracker is covered with chocolate.
C) Spherical, popped cereals with quinoa on a small round plate.
D) Three squashed, ball-​like quinoa and buckwheat cookies, on a small round plate.
E) Three rectangular quinoa, cinnamon and lemon cookies; on the surface there is a
cookie mold drawing of four parallel lines down the middle. Placed on a small round plate.
F) Three circular quinoa cookies with cocoa, with a rough surface, on a small
round plate.
G) Three squared chocolate with quinoa, on a small round plate.
H) Puffier white quinoa compared to raw white quinoa seeds, each on a small
round plate.
266 Food Uses of Selected Ancient Grains

2020). Ballester et al. (2019a, 2019b, 2019c) developed bakery products with 25%
replacement of wheat by Real quinoa of different colors in order to evaluate its
functionality as a bread-​making ingredient, with excellent results in sensory
evaluation, nutritional/​functional value but lower technological bread quality
than the counterpart without quinoa. In this sense, the fiber fraction with high
amounts of antioxidant compounds was isolated, by wet and dry milling, and
included in bread formulation at 5% to get the same functional value as the
samples with 25% of quinoa but with equal bread quality than the control sample
(wheat) (Ballester et al., 2020).
In general, the pseudo-​cereals improved mineral content but also the phytate
content of breads (Bilgiçli and İbanoğlu, 2015; García-​Mantrana et al., 2014;
Iglesias-​Puig et al., 2015; Miranda-​Ramos et al., 2019; Miranda-​Ramos and
Haros, 2020). High phytate contents were easily avoided by the use of exogenous
phytases, the use of flours from germinated seeds or the use of sourdough or
acidified sponges (Park and Morita, 2005; García-​Mantrana et al., 2015; Dallagnol
et al., 2013; Iglesias-​Puig et al., 2015).
In regard to chia, the first bread products described in the literature with
this whole grain were developed for women, rich in fibers and proteins, forti-
fied with iron and folic acid, with or without soy and/​or linseed, which showed
high acceptability by consumers (Bautista Justo et al., 2007). Subsequently,
Iglesias-​Puig and Haros (2013), following the regulations in force in Europe
at that time (EFSA, 2009; OJEU, 2009), bread products with up to 5% substi-
tution levels in flour bases were developed (Figure 7.2.A). The objective of
this investigation was to develop new cereal-​based products with increased
nutritional quality by using chia ingredients such as its seeds, whole flour,
semi-​defatted flour and low-​fat flour, in order to evaluate its potential as a
bread-​making ingredient. Later, because of its high nutritional properties,
consumption of chia has spread widely within the European Community
(EC); and in the EU list of novel foods, the European Food Safety Authority
(EFSA) has authorized the use of chia seeds and/​or other chia ingredients
to higher levels and/​or their inclusion in other foods different than bread
(Turck et al., 2019a).
Whole chia seeds may be marketed in the EC as a food ingredient and
used in baked products and breakfast cereals up to 10%; ground chia seeds
up to 5% in bread; whole chia seeds up to 5% in sterilized ready-​to-​eat meals
based in cereal/​pseudo-​cereal grains and/​or pulses; prepackaged chia seed
as such and fruit/​nut/​seed mixes; and chia in confectionery products and
chocolates; edible ices; fruit and vegetable products; non-​alcoholic beverages
and puddings (<120ºC in their preparation) without limit, according to the EC
(EU, 2020). Recently, the use of two partially defatted powders of chia enriched
7.2 Bakery and Bread Products 267

Figure 7.2 Foods with chia: A. Bread with chia seeds, 5%; B. Fresh pasta with chia
seeds, 10%; C. Whole rice with chia, quinoa, spelt and flaxseeds; D. Digestive chia cookies;
E. Biscuits with curcuma and chia; F. Chia and spelt roscos; G. Crackers of chia and quinoa;
H. Chia and quinoa sticks; I. Sea biscuits with chia. Source: Authors’ own work.
Notes: A) Five slices of bread with chia seeds on a medium-​size round plate.
B) Close-​up of white tagliatelle spaghettis with black chia seeds.
C) Whole rice with ready-​to-​eat chia, quinoa, spelt and flax seeds on a small round plate.
D) Seven circular digestive cookies with chia and a rough surface, in a small round bowl.
E) Three yellow rectangular cookies with curcuma and chia; on the surface there is a
cookie mold drawing of four parallel lines down the middle; placed on a small round plate.
F) Seven chia and spelt doughnut-​shaped cookies in a small bowl.
G) Many crackers with chia and quinoa of irregular quadrilateral shape in a small bowl.
H) Many popped sticks with chia and quinoa in a small bowl.
I) Many circular sea biscuits with chia in a small bowl.
268 Food Uses of Selected Ancient Grains

with proteins or fibers was authorized as food supplements for the adult popu-
lation (up to 7.5 and 12 g/​day, respectively), or as nutritional ingredients in a
variety of foods (yogurt, vegetable beverages, energy drinks, chocolate, fruit
and pasta) at a level of 0.7–​10% (Turck et al., 2019b). Restrictions in the use
of chia/​chia ingredients –​principally in bakery products –​are only valid in
the EU; the rest of the world does not restrict their uses. The main reason for
these restrictions are the high amount of asparagine (and consequent forma-
tion of acrylamide in baked products) in chia and its ingredients (Galluzzo
et al., 2021), especially these that have higher protein concentrations than
the original seeds. The partial replacement of wheat by chia seeds, whole chia
flour and defatted chia flour in bread (up to a level of 5–​6%) obtained high
consumer acceptance in earlier studies, and it could extend the shelf-​life of
bread, since it inhibits the kinetics of retrogradation of amylopectin during
storage (Iglesias-​Puig and Haros, 2013; Sayed-​Ahmad et al., 2018). However,
the bread-​making process may affect the stability and/​or bioavailability of
nutrients/​bioactive compounds owing to various chemical and enzymatic
reactions during kneading, fermentation and baking, and accordingly this
enriched bread would provide a greater or lower health benefit (Benitez et al.,
2018; Miranda-​Ramos et al., 2020; Miranda-​Ramos and Haros, 2020). In gen-
eral, formulations with chia ingredients significantly increased the levels of
proteins with high biological value, healthy lipids, ashes/​minerals and soluble
dietary fiber, with similar or better technological quality to the control bread.
This is mainly due to the increase in specific bread volume, decrease in crumb
firmness, with change in crumb color and higher overall acceptability by con-
sumers (Miranda-​Ramos et al., 2020; Miranda-​Ramos and Haros, 2020; Guiotto
et al., 2020; Ahmed et al., 2020).
Only a few scientific investigations have been found that make use of
Andean maize (Zea mays), black turtle bean (Phaseolus vulgaris), tarwi (Lupinus
mutabilis) or sacha inchi (Plukenetia volubilis) in bakery product formulations.
There is more information on gluten-​free products, which were discussed in
Chapter 6. Regarding the case of maize flours from crops grown in the Andean
region, which some authors have termed as Andean maize, these can be used
in bread, pasta, cookies and cakes, providing color and flavor to the processed
product (Salvador-​Reyes and Pedrosa Silva Clerici, 2020). This type of maize
represents a potential ingredient for the development of new products: it can
provide a variety of color, flavor and texture, contributing to the reduction of the
use of additives in the food industry, besides being ancient grains, which have
survived naturally against extreme adverse growing conditions (Salvador-​Reyes
and Pedrosa Silva Clerici, 2020). In the case of black turtle bean, there are no
scientific publications in terms of its use in the bread-​making process. In gen-
eral, legumes and in particular lupins can be used to fortify the protein content
7.3 Biscuits, Cookies and Cakes 269

of pasta, bread, biscuits, salads, hamburgers and sausages, and can substitute
for milk and soya beans (Carvajal-​Larenas et al., 2016). Replacement of wheat
flour by <12.5% of tarwi produced breads with good sensory acceptability but
variable color and doughs with acceptable thermomechanical patterns (Rosell
et al., 2009). Bread with 10% of tarwi flour had a protein efficiency ratio (76%)
higher than the control (28%) and similar acceptability (Carvajal-​Larenas,
2019). However, it was an effort to evaluate the decrease or elimination of non-​
nutritional compounds, and changes in color in tarwi derivatives (protein
concentrated or isolated) to apply in three types of bread products (loaf, bun
and sweet) (Güemes-​Vera et al., 2008).
In the case of sacha inchi as a bread ingredient, a bread loaf enriched with
extruded sacha inchi cake has been developed (Rodríguez et al., 2018). The
incorporation of this bakery ingredient significantly increased the levels of ash,
fiber, fat and protein, and decreased the carbohydrate content. The best rheo-
logical properties were presented at 6.3% incorporation level, which highlights a
retention of 3.33% of alfa-​linolenic fatty acid (omega-​3), increase of texture and
crumb darkening. There was no significant difference between the control and
the bread with sacha inchi in terms of the attributes of color, appearance, aroma,
flavor and texture. The elaboration of bread with the incorporation of coconut
pulp flour and sacha inchi nibs was also investigated (Ordoñez et al., 2019). The
best results in the sensory evaluation (aroma, texture and volume-​symmetry),
physicochemical and farinographic tests were found in breads made with 7.5%
sacha inchi nibs with coconut flour (12.5%). The effect of partial replacement
of wheat flour by cake of sacha inchi on the rheological properties of dough in
sweet bread was also studied (Toralva Aylas et al., 2015).
The Latin-​American crops are known to be rich in some bioactive compounds,
such as flavonoids, phenolic acids, trace elements, dietary fibers, essential fatty
acids and vitamins with known effects on human health. In this sense, partial
substitution of wheat flour by whole flours from these seeds/​fruit or a fraction of
them (single or combined) constitutes a viable option to improve the nutritional
value of bakery products, with acceptable technological performance of dough
blends and composite products which provides a wide gamma of products with
a positive perception by consumers.

7.3 BISCUITS, COOKIES AND CAKES


Cookies/​biscuits/​cakes are other cereal products in the bakery industry in which
efforts have been made to improve the composition of the product by the inclu-
sion of ancient grains. They typically have: low fiber and water levels, high fat
and sugar and air cells varying in size embedded within the protein–​starch–​lipid
270 Food Uses of Selected Ancient Grains

matrix (Brennan and Samyue, 2004; Poutanen et al., 2014). Some examples are
shown in Figures 7.1–​7.4.
The incorporation of amaranth flour in sugar snap cookies improved the
cookie’s surface cracking as a consequence of the reduction in their breaking
strength. As a result of increasing the percentage of amaranth ( from 5 to 35%)
in the formulation, an increase in thickness was also observed (Sindhuja et al.,
2005). The sensory evaluation revealed that cookies containing 25% of amar-
anth scored the best marks due to their golden-​brown color, larger size and uni-
formity, malty-​sweet flavor and tenderness.
In another study, sensory evaluation of biscuits prepared with wholemeal
barley or wheat flour supplemented with amaranth showed that addition at 20%
and 30% levels to barley and wheat flour, respectively, gave the best color, and at
levels of 10–​20% gave the best biscuit taste (Sandak and El-​Hofi, 2000).
Wheat flour and corn starch could be successfully replaced by up to 20% amar-
anth flour in conventional and up to 30% in reduced fat pound cakes without
negatively affecting sensory quality in fresh products (Dias-​Capriles et al., 2008).
Biscuits were developed with amaranth –​and other foods –​to increase their
nutritional value with high acceptance by consumers (Nidhi and Indira, 2012;
Zanwar and Pawar, 2015). With the help of sucralose and soluble fibers, it is pos-
sible to develop a sweet biscuit with amaranth flour partially reduced in sugars
and fats with low caloric content and high nutritional value (Torres Palacios
et al., 2019). However, the diameter of the cookies proportionally decreased with
the increasing levels of supplementation with amaranth/​oat/​sorghum, whereas
hardness of cookies increased, obtaining highest overall acceptability scores
when the level of substitution is at 10% (Raihan et al., 2017).
The performance of quinoa–​wheat flour blends was also evaluated in cakes
biscuits and cookies, providing products with various technological and sen-
sory characteristics (Lorenz and Coulter, 1991; Wang et al., 2015; Watanabe
et al., 2014). Cake quality was acceptable with 5–​10% of quinoa flour, taste
improved and cake grain became more open and the texture less silky as the
level of quinoa substitution increased. On the other hand, cookie spread and
top grain scores decreased with increasing levels of quinoa flour blended
with high-​spread cookie flour. Flavor improved with an incorporation of up
to 20% quinoa flour in the blend, whereas cookie spread and appearance
improved with a quinoa/​low-​spread flour blend by using 2% lecithin (Lorenz
and Coulter, 1991). The cookies containing quinoa flour had greater oxidative
stability and nutritional quality, and were rich in dietary fiber, essential amino
acids, linolenic acid and minerals, with good sensory acceptability (Wang
et al., 2015; Watanabe et al., 2014). Quinoa-​tempeh –​fermented with Rhizopus
oligosporus –​powder is more suitable than quinoa powder as an ingredient for
biscuits, and it may be added to flour in amounts of up to 20% according to the
7.3 Biscuits, Cookies and Cakes 271

Figure 7.3 Foods with amaranth: A. Amaranth chocolate chip cookies; B. Amaranth
multigrain crispbread; C. Amaranth with chocolate; D. Amaranth snacks: moka flavor;
with raisins, peanuts and pumpkin seeds; with cranberries; chocolate flavor; with
peanuts and probiotics; with chocolate and probiotics; with cranberries and probiotics.
Probiotic: Bacillus coagulans GBI-​30. Source: Authors’ own work.
Notes: A) Eight round amaranth chip cookies with a rough surface created by the seeds,
on a medium size plate.
B) Five rectangular amaranth multigrain crispbread with cracks and holes on the sur-
face, in a small bowl.
C) Seven cubes, of three different types and colors (white, pink and brown), of amar-
anth with chocolate on a small plate.
D) Seven different snacks of popped amaranth, in their respective transparent plastic
bags, on a large rectangular plate. Distribution is two cubes at the top (moka and raisins,
peanuts and pumpkin seeds), followed by two other cubes (with cranberries and one
with chocolate flavor), and then three rectangular shaped snacks (all with probiotics)
positioned horizontally, parallel to one another (with peanuts, chocolate and cranberries,
respectively).

biscuit quality and sensory analysis (Matsuo, 2006). Quinoa affected the scores
of sensory properties of cookie samples, in terms of color, taste, crispness and
overall acceptability, with the exception of odor scores (Demir et al., 2017).
Even wheat flour substitution with quinoa flour severely decreased biscuit
acrylamide, and almost maintained biscuit browning index in comparison to
the control sample (Sazesh et al., 2020). There is no information in the litera-
ture concerning kañiwa as a raw material to develop cookies/​biscuits or cakes,
272 Food Uses of Selected Ancient Grains

but there are many recipes and commercial products with kañiwa, with and
without other crops.
Concerning chia, there are many investigations which describe the use
of its seeds and whole flour to develop cookies, biscuit or cake formulations
with the objective of improving their nutritional/​functional quality, and also
to take advantage of by-​products of the chia industry such as defatted chia
flour, mucilage fractions, its oils and/​or enriched protein flours. Usually, the
addition of chia seeds and chia flour in cookies/​biscuits/​cakes improves their
nutritional value and sensory acceptance of the products (Goyat et al., 2018;
Alcántara Brandao et al., 2019). Supplementation of cookies with 10% defatted
chia flour could be recommended to improve antioxidant quality without redu-
cing technological or sensorial properties (Mas et al., 2020). In another study, it
was found that the mixture of canola-​chia oils and gelatin enhances the nutri-
tional and sensory quality of the pound cake reduced in margarine without
changing final cake quality (Sánchez-​Paz et al., 2020). In this sense, to avoid
lipid oxidation, partial substitution of margarine by microencapsulated chia oil
in carnauba wax solid lipid microparticles was investigated, to obtain cookies
containing omega-​3 and omega-​6 (Carvalho de Almeida et al., 2018; Hernándes
Venturini et al., 2019). On the other hand, chia mucilage also proved to be a
new alternative for replacing fats or eggs in cake products, preserving quality
attributes while making them healthier foods (Borneo et al., 2010; Ferrari
Felisberto et al., 2015; Santos Fernándes and Salas Mellado, 2017). Sensory ana-
lysis showed that chia gel stored in different conditions could replace egg in
chocolate cake without impairing acceptability. It is possible that cold storage
improves chia gel functionality, especially when it is stored frozen (Rodrígues
dos Reis Gallo et al., 2020). In cookie formulations with roasted chia, the baking
loss, hardness and brightness were inversely proportional to roasting time,
affecting the texture and sweetness scores in a consumer preference test (Song
et al., 2019). It is important to remark that although nutritional properties are
improved by the incorporation of chia/​chia by-​products into biscuit/​cookie
formulations, the increase in content of process contaminants –​such as acryl-
amide, hydroxymethyl-​furfural and/​or furfural, furan and/​or methyl furans –​
and the extent of the lipid oxidation should be carefully considered (Mesías
et al., 2016; Myrisis et al., 2022).
As was previously mentioned, Andean maize flour can be used in cookies
and cakes, as a potential ingredient for the development of new and innovative
products in terms of natural colors and new textures and flavors (Salvador-​Reyes
and Pedrosa Silva Clerici, 2020). On the other hand, fruits and extracts from
sacha inchi were studied such as integral sacha inchi with linseed sticks (Yovera
Méndez and Ruddy Gianny, 2018); cookies and cakes with residual sacha inchi
cake flour after oil extraction (Toralva Aylas and Rodas Pingus, 2015; Vásquez
7.4 Pasta Products 273

Osorio et al., 2017). On websites several recipes are described with this fruit or
its derivates.
In general, lupin flour is widely considered an excellent raw material for
supplementing different food products owing to its high protein content
(Kohajdová et al., 2011) and is largely used as an egg and/​or butter substitute; for
example, in cakes, pancakes, biscuits, brioche, and/​or croissants (Tronc, 1999).
The optimal formulation of cookies made with tarwi oil and chia was at 12% and
3% level, respectively, with an acceptability of 6.58 on a 7-​point scale (Salvatierra-​
Pajuelo et al., 2019).

7.4 PASTA PRODUCTS
The production of pasta has been favored with the inclusion of ancient grains and
their by-​products. Fresh and dry gluten-​free spaghetti were developed varying
the proportions of maize, quinoa and soy flours between 20.00 and 31.25%, 7.50
and 30.00% and 2.50 and 10.00%, respectively (Mastromatteo et al., 2011). These
authors evidenced that the spaghetti samples with highest maize and quinoa
content presented the best extensibility values, and the sensory quality did not
significantly vary among formulations. Moreover, Giménez et al. (2015) made
spaghetti from Andean corn, capia and cully varieties from northern Argentina
and compared the sensory profile of these types of spaghetti with those made
with rice and wheat flours. These authors observed that celiac consumers
assigned high acceptability scores to the corn products. Then, Giménez et al.
(2016) determined the nutritional quality of spaghetti made with maize flour
enriched with 30% broad bean flour and 20% quinoa flour. These gluten-​free
spaghetti showed a significant increase in net protein utilization and decreased
digestibility compared with a non-​enriched control sample. One portion of this
product supplied the 10–​20% of recommended daily intake of fiber. Addition of
quinoa flour had a positive effect on the Fe supply as did broad bean flour on
Zn. Meanwhile, Deepa et al. (2017) made three types of pasta: unmicronized
maize flour pasta with guar gum, micronized maize flour pasta with guar gum
and micronized maize flour pasta with guar and xanthan gum. Pasta samples
prepared with micronized flour showed increased firmness, improved color
index and overall acceptability, compared with those of unmicronized flour.
Many pasta formulations have been developed based on amaranth flour
and in general, the results demonstrated that this raw material affects texture
firmness, stickiness and cooking time (Rayas-​Duarte et al., 1996; Schöenlechner
et al., 2010; Fiorda et al., 2013). Rayas-​Duarte et al. (1996) showed that an ideal
level of substituted grain would optimize nutritional quality without destroying
functional properties. These authors reasoned that a multigrain pasta with better
274 Food Uses of Selected Ancient Grains

nutritional quality, acceptable cooking properties and good sensory attributes


can be produced, using buckwheat, amaranth or lupin flour without exceeding
the following substitution levels: 30% for light buckwheat, 15% for dark buck-
wheat, 25% for amaranth, and 15% for lupin. Cárdenas-​Hernández et al. (2016)
reported that a pasta based on amaranth flour (25.5–​50.0%) and semolina (51.0–​
21.0%) with 15.0% of egg presented acceptable technological quality compared
to a control pasta (85.0% semolina and 15.0% egg). Even though the incorpor-
ation of amaranth produced a reduction in pasta luminosity values and sensory
acceptance, all the formulations exhibited a significantly higher content of pro-
tein (17.47%), crude fiber (2.64%), ash (2.03%), fat (4.40%), total phenol content
(2.25 mg Ferulic acid equivalents/​g ) compared to the control pasta (15.13; 0.72;
0.90; 2.12; 0.98 respectively).
Pantoja Tirador and Prieto Rosales (2014) developed pasta made with 80%
wheat, 10% tarwi and 10% quinoa flours with an acceptable sensorial quality and
high nutritional value providing 19.06% protein; 3.23% fat; 0.65% ash and 2.16%
fiber. The incorporation of 20% (w/​w) of quinoa flour to semolina improved
the nutritional profile without affecting the technological and sensory quality
of pasta (Lorusso et al., 2017). Furthermore, these authors demonstrated that
the quinoa flour fermentation with lactic acid bacteria can also enhance the
favorable outcomes of quinoa. Moreover, Demir and Bilgicli (2020) found that
the replacement of wheat semolina with 10, 20 and 30% of raw and germinated
quinoa flours in pasta formulations significantly improved the nutritional quality
of the product, keeping technological and sensory qualities within acceptable
limits. These authors communicated that as the raw and germinated quinoa
flour ratios increased in pasta, ash (0.81 to 1.35%), crude fat (0.49 to 1.39%) and
protein (12.56 to 16.65%), total phenol content (0.54 to 1.22 mg GAE/​g ), anti-
oxidant activity (12.39 to 27.69%) and mineral matter (Ca, Fe, K, Mg, P and Zn)
amounts significantly increased.
Kañiwa flour can be used to replace partial wheat flour in pasta to increase
its nutritional value. Bustos et al. (2019) developed enriched pasta with
wheat and kañiwa blends. These authors studied three levels of kañiwa flour
substitutions: 10, 20 and 30% and observed that this flour replaced negatively
affected pasta cooking properties, such as water absorption and cooking loss.
With respect to texture parameters, Kañiwa pasta firmness and chewiness
generally decreased when kañiwa content increased. A replacement of 20% of
kañiwa flour improved the nutritional quality of pasta, increasing the dietary
fiber (57.0 to 97.2 g/​kg pasta) and protein quality (130.5 to 136.4 g/​kg pasta),
obtaining a functional pasta with satisfactory cooking quality.
Ground chia seed and defatted chia flour have been widely used for the devel-
opment of these types of products. Home style noodles were made with 10%
substitution of durum flour with ground chia seeds. The final product changed
7.4 Pasta Products 275

from “yellow” to “yellowish-​gray” preserving acceptable taste properties with


an increase in macroelements (calcium by 2.2, magnesium by 2 and phosphor
by 1.4, times respectively) and of microelements’ concentration (copper by 2,
zinc by 1.5 and iron by 3.5, times respectively). Moreover, the microstructure of
the product, vitamin value, and physicochemical indicators of quality of pasta
products were not affected during their lifetime (Naumova et al., 2017). Levent
et al. (2017), using chia flour in pasta based on rice and corn flour, obtained a
healthier product with acceptable sensory results, with 20% of chia flour incorp-
oration in combination with diacetyl tartaric esters of mono (and di-​)glycerides.
In addition, Aranibar et al. (2018) evaluated the nutritional, sensory and techno-
logical quality of pasta supplemented with partially de-​oiled chia flour (PDCF)
at different proportions (2.5%, 5.0% and 10.0%).These authors concluded that
incorporation of PDCF into wheat pasta allows significant improvement to nutri-
tional properties (total dietary fiber, 2.86 to 9.08%, d.b; ω-​3/​ω-​6 ratio, 0 to 2.14;
total phenolic content, 10 to 35 mg GAE/​100g pasta and antioxidant capacity)
preserving in general the color, sensory, texture and cooking characteristics of
pasta samples compared to non-​supplemented pasta.
Legume flours were used in combination with precooked rice in order to
produce gluten free spaghetti (Bouasla et al. 2017). The precooked rice pasta
was enriched with different levels (10, 20 and 30%) of legume flours (yellow pea,
chickpea and lentil). The products obtained presented an improved chemical
profile compared with control pasta (rice pasta). In fact, the content of proteins,
fat, ash, insoluble and soluble fiber was among 9.68–​13.95, 0.08–​0.38, 0.69–​1.29,
2.66–​5.15 and 1.09–​1.70% d.b., respectively; while the control sample presented
the following composition 8.25, 0.09, 0.50, 2.52, 0.69% d.b., respectively. These
authors maintained that incorporation of legume flour reduces lightness and
expansion ratios, and increases yellowness, adhesiveness and firmness, without
affecting cooking time. On the other hand, Sudha et al. (2012), Nielsen et al.
(1980), and Bahnassey and Khan (1986) who studied the fortification of pasta
with legume flours recognized significant structural changes in the protein net-
work, which were reflected in greater cooking loss. Similar trends were observed
by Zhao et al. (2005) who maintained that the cooked weight of pasta was not sig-
nificantly affected, but cooking loss increased with the rise in content of legume
flours. In general, the addition of legume flours decreased pasta firmness, so the
incorporation of additives like gluten, glycerol mono-​stearate or sodium stearoyl
lactylate, in combination, improved the quality characteristics of pasta.
In the literature there are several applications of tarwi flour for the develop-
ment of pasta. Pepe Guato (2011) made wheat flour pasta with a degree of substi-
tution with tarwi flour of 0, 15, 20, 25 and 30%. The best treatment was pasta with
20% replacement tarwi flour, as it presented a higher than average acceptability,
lower stickiness and greater firmness; and shorter cooking time. Regarding the
276 Food Uses of Selected Ancient Grains

chemical composition, it presented 9.41% humidity; carbohydrates 69.94%; pro-


tein 22.56%; fiber 2.81%. Ponce et al. (2018) who studied a partial substitution
of wheat flour with tarwi flour in the production of long pasta concluded that
substitution of 25% tarwi flour and 18% egg in the mixture was the best formula-
tion to obtain a more nutritious and good quality pasta. In addition, Albuja and
Yépez (2017) developed gluten-​free rice pasta with partial substitution of tarwi
flour, egg and guar gum addition. The best pasta was achieved with 20% tarwi
flour, 30% egg and 0.15% guar gum. This pasta compared with a control of 100%
rice presented an increase in protein 12.06 to 20.64% and fiber 0.32 to 0.75% d.b.
Finally, Pinares Huamani (2019) formulated pasta with a partial substitution of
tarwi flour (between 8 and 22%) and adding powdered egg shell (between 0.8 and
2.2%). The replacement of 10% tarwi and 1% egg shell powder was the best alter-
native regarding general acceptability and cooking properties. The pasta showed
a high nutritional value: 12.33% protein; 0.27% fat; 73.58% carbohydrates; 139.9
mg /​100 g of calcium and 7.25 μg /​100 g of vitamin D.

7.5 SNACKS AND BREAKFAST MIXTURES


Ritva Repo-​Carrasco-​Valencia et al. (2011) designed a nutritious extruded snack
product using corn and three varieties of quinoa. The optimal mix was 70%
quinoa and 30% corn. This product offers a nutritious alternative to traditional
snacks. In addition, Figueiredo de Sousa et al. (2019) developed snacks bars (SB)
with different corn bran (CB) proportions (0, 10, 25, 40 and 55%). The SB with the
highest proportions of CB had the highest levels of dietary fiber. Formulations
with up to 40% CB were well-​accepted by the judges, compared with the for-
mulation without any CB. On the other hand, Carvajal et al. (2019) developed
a snack based on corn, beans and sweet potato with the following proportion
(w/​w): 80/​10/​10 and 70/​15/​15. This last formulation presented the higher nutri-
tional content: protein (12.48%), ash (1.36%) and fiber (1.28%) in relation to the
commercial control (100% corn) that presented protein (8.32%), ash (0.57%) and
fiber (0.64%) content.
Chávez-​Jauregui et al. (2000) prepared an extrudate based on amaranth
flour with a fat, protein and carbohydrate content of 0.18%, 15.82% and 80.77%,
respectively. Moreover, Gearhart and Rosentrater (2014) made an extrudate
snack with 100% of amaranth containing 6 to 8%, 13 to 18% and 63% of fat, pro-
tein and carbohydrate, respectively. In addition, Ramos Díaz (2015) produced
expanded CB snacks containing up to 50% amaranth, quinoa and kañiwa and at
most 20% lupine maintaining textural and sensorial properties as well as added
nutritional value.
7.5 Snacks and Breakfast Mixtures 277

The development of cereal bar formulations has been investigated by authors


such as Delgado and Barraza (2014) who produced bars based on quinoa, sacha
inchi and amaranth with a maximum inclusion of 24.9% of quinoa in the for-
mulation. The general acceptance of the bar with the highest quinoa content
presented the lowest score, while the formulation containing 21.8% was the most
accepted. Regarding the proximal analysis, it was reported that the bars have a
protein content between 2.67 and 2.25 g/​g d.b. Moreover, Calisto-​Guzmán (2009)
formulated quinoa bars with red beans and honey that obtained good sensory
acceptance. Another alternative are the quinoa flakes which have a protein con-
tent of 8.5%, fiber 3.8% and minerals such as calcium 114 mg, phosphorus 60 mg
and iron 4.7 mg /100 g (González Rojas and Moya Valenzuela, 2004; Rojas et al.,
2010). According to a report by Bergesse et al. (2015), the nutritional difference
between quinoa flakes and corn flakes is their essential amino acid content.
Repo-​Carrasco-​Valencia et al. (2003) formulated two dietary mixtures: quinoa-​
kañiwa-​beans and quinoa-​amaranth-​beans, with high nutritional value. At the
same time, Repo-​Carrasco-​Valencia et al. (2009) produced expanded snacks
based only on kañiwa with two different varieties. The products presented on
average the following composition: proteins, 14.13%; fat, 5.63%; crude fiber,
4.56%; ash, 4.23%; carbohydrates, 71.45%; total dietary fiber, 19.52%; phenolic
compounds, 2.48 mg gallic acid equivalent/​g d.b.; antioxidant activity, 4125 µg
trolox eq./​g d.b.
Coorey et al. (2012) produced chips with rice, potato and chia flours. The
chips elaborated with 0, 5, 10, 12 and 15% of chia flour substitution presented the
following nutritional parameters: protein, 3.87, 5.10, 5.97, 6.59, 7.14%; fat, 3.88,
4.46, 5.30, 3.78, 3.32%; dietary fiber, 2.79, 2.91, 8.20, 10.43, 11.72%; ash, 5.89, 6.29,
7.11, 6.54, 7.33%; calcium, 0.00, 39.75, 94.85, 110.41, 126.21%; and antioxidant
activity, 1.32, 5.55, 25.66, 35.01, 44.04%, respectively. There were no significant
differences in sensory acceptability between a commercial chip sample and the
5% chia chips. Chemical analysis indicated that all four trial chips were excellent
sources of omega-​3.
Delgado and Jáuregui (2014) evaluated the effect of the proportion of quinoa
(21.8%, 24.9% and 15.7%), amaranthus (21.8%, 16.9% and 22%) and sacha inchi
(5.3%, 7.1% and 11.2%) on the acceptability and nutritional profile of an energy
bar. It was determined that the proportion of quinoa, amaranth and sacha inchi
did not influence the general acceptability, flavor and texture of an energy bar
unlike the proximate composition, where significant influence was observed,
reporting that the formulation with 21.8% of quinoa, 21.8% amaranth and 5.3%
saccha inchi presented higher protein (2.85%) and fiber (1.21%) and lower fat
(4.37%) content than the rest of the combinations. Moreover, Verduga Verdezoto
(2019) designed an energy bar with 40.00% sacha inchi, 30.00% honey, 6.66% of
amaranth, quinoa, oat and 10.00% dehydrated pineapple with good sensorial
278 Food Uses of Selected Ancient Grains

acceptance. The product contained 9.74% humidity, 23.93% fat, 14.21% protein,
1.82% ashes, 13.74% fiber and 50.30% carbohydrates.
Tiwari et al. (2011) developed deep-​fried snacks with nutritional benefits and
an acceptable quality of cereal (rice) and legumes. The protein and fat content
ranged from 13.5% to 17.5% and 23.2% to 30.0%, respectively. Patil et al (2016)
prepared wheat-​based extrudates using four different legume flours: lentil,
chickpea, green pea, and yellow pea flour added at different levels (0%, 5%, 10%,

A B C

D E

Figure 7.4 A. Natural sachi inchi snack, B. Sacha inchi snack with salt, C. Wasabi-​flavored
snack with sacha inchi, pistachio nuts, raisins and maize, D: Wheat pancakes with sacha
inchi (10%), E: Vanilla flavored hydrolyzed collagen with sacha inchi. Source: Authors’
own work.
Notes: A) Oval shaped, toasted sacha inchi seed snacks covered in salt on a small plate.
B) Toasted, salty wasabi-​flavored snacks with sacha inchi seeds, pistachio nuts, raisins
and maize grains, on a small plate.
C) Two large, rectangular wheat pudding cakes with sacha inchi (10%), positioned on a
plate and parallel to each other.
D) Dust-​like vanilla flavored hydrolyzed collagen with sacha inchi in a small bowl.
7.6 Drinks and Fermented Beverages 279

and 15%). The snack with legumes caused a slight increase in protein content
but the extrusion technique increased protein digestibility by 37%–​62% w/​v.
Tarazona and Bocanegra Reyes (2018) developed an extruded snack based on
rice (90%) and tarwi (10%) flours. The product presented great sensory accept-
ability and nutritional properties (13.11% of protein and 4.03% of fiber).
Some commercial snacks made with quinoa, chia, amaranth and sacha inchi
are shown in Figures 7.1–​7.4.

7.6 DRINKS AND FERMENTED BEVERAGES


One of the most important traditional beverages in Andean cultures is the
so-​called “chicha”, which is a fermented or unfermented drink made from
maize (Nicholson, 1960). This beverage has been produced since pre-​Hispanic
times in the north-​west regions of Argentina, and Andean regions of Bolivia,
Colombia, Ecuador and Peru (Elizaquível et al., 2015). “Chicha” has played a
very important role in Andean culture as an essential part of political, social,
ceremonial, and religious transactions. In addition, an ethnohistorical informa-
tion report declared that fermented beverages were highly esteemed and crit-
ical for ceremonial gatherings (Hastorf, 2015; Hastorf and Johannessen, 1993).
Cereal fermented products, those derived from maize, have been important
in Latin America; in this case chicha is produced by handmade ancestral
procedures. Chicha has been defined by Lorence-​Quiñones et al. (1999) as a
clear, yellowish, effervescent, alcoholic beverage prepared from maize with
flavor similar to that of cider; the alcoholic content varies from 2 to 12% (v/​v).
When pigmented maize varieties are used, their color varies from red to purple,
for example, purple maize is used to produce the traditional Peruvian drink
and dessert called chicha morada and mazamorra morada (Salvador-​Reyes and
Clerici, 2020).
On the other hand, amaranth has also been used to produce different kinds
of beverages due to its characteristics as a raw material with nutritional and
therapeutic properties in the prevention and treatment of different diseases
(Aderibigbe et al., 2020). In recent years, there has been an interest in func-
tional food and beverages with beneficial physiological properties; in this sense,
extruded amaranth grains have been used to develop an instant nutraceutical
beverage with high antioxidant capacity, rich in dietary fiber and proteins. These
can contribute to the prevention of cardiovascular disease, type-​2 diabetes,
among others (Milán-​Carrillo et al., 2012); amaranth proteins have been used to
formulate a functional high protein beverage suitable for vegans, celiac patients,
and lactose intolerants (Manassero et al., 2020). An amaranth instant beverage
was developed by mixing amaranth, rice and corn flour with milk whey and
280 Food Uses of Selected Ancient Grains

powder milk by spray dryer to obtain an instant beverage with nutritional and
high protein properties (Arcila and Mendoza, 2006).
In addition, some mix of malted cereals and pseudo-​cereals such as sorghum,
buckwheat, quinoa and amaranth were used to produce gluten-​free beer; the
results showed high fermentability of malted amaranth suggesting that it is
possible to produce gluten-​free beer as an alternative for celiac people (de Meo
et al., 2011).
Quinoa has also been used to produce different beverages, fermented and
unfermented, gluten free and with functional properties. A beverage rich in phen-
olic components, high protein and high antioxidant capacity with antidiabetic
and antihypertensive potential was developed by Kaur and Tanwar (2016) with
germinated and malted quinoa seeds. The use of different techniques such as
soaking, germination and malting can improve bio-​accessibility by enzyme
action, improve nutrient content and decrease the antinutrients in grains
(Owusu-​Mensah et al., 2011). In the same context, quinoa has also been used to
produce a fermented beverage with probiotic cultures including Bifidobacterium
sp., Lactobacillus acidophilus, and Streptococcus thermophilus, obtaining a sig-
nificant increase in proteins and phenolic compounds, and producing an
end-​product which can increase health benefits due to the content and bio-​
accessibility of its bioactive compounds (Karovičová et al., 2020). A fermented
quinoa-​based beverage was developed by Ludena Urquizo et al. (2017) who used
two quinoa varieties (Rosada de Huancayo and Pasankalla) fermented by three
strains of Lactobacillus to produce a high protein, fiber, vitamins and mineral
content and low saponins beverage. Results showed that both varieties can be
used for this kind of beverage.
Tarwi has also been used to produce beverages; its proteins exhibit excel-
lent techno-​functional properties such as solubility and emulsification, which
make them an attractive raw material producing plant-​based beverages (Nawaz
et al., 2020). In this context, tarwi has been used to produce spray-​dried milk
as an alternative to cows’ milk for the institutional and vegetarian market in
Chile; this vegetable milk showed a protein, a similar fat and ash content to
cows’ milk, while minerals such as calcium and phosphorus were slightly lower
(Camacho et al., 1988). In addition, lupin-​based milk alternatives and different
exo-​polysaccharides which produce lactic acid bacteria were investigated by
Hickisch et al. (2016) to produce a yogurt substitute for vegetarians or vegan
consumers. For this purpose, protein isolate was used to produce the milk which
was thermally treated with two temperatures (80° and 140°C) and two strains of
Lactobacillus, and one Pediococcus were used to produce the exo-​polysaccharide.
The more intensive thermal treatment resulted in better rheological and textural
properties and lower syneresis tendency, which revealed that lupin proteins are
appropriate for producing plant-​based yogurt.
7.7 Infant Foods 281

On the other hand, Sacha inchi is recognized as a good source of unsatur-


ated lipids and bioactive compounds; in this sense, the seed has been used to
produce an enriched yogurt with a significant increase in protein content, n-​
3 and n-​6 fatty acids, tocopherols, phytosterols, phenolic compounds, dietary
fiber and essential amino acid, making it a valuable resource for high value
compounds to be used in food and beverages (Wang et al., 2018). Chia seed has
been used to elaborate beverages since the pre-​1600 period, for example, an
Aztec beverage called “Chinatoles” was made with roast and ground seeds; the
whole seeds were also used to prepare a refreshing drink “chia fresca” which has
recently gained great popularity in Mexico (Cahill, 2003). Today, the whole chia
seed and protein fractions are used in many commercial beverages mixed with
fruit juices. In this context, the Company “Mamma Chia” has developed a great
variety of beverages using organic seeds in the development of organic chia
beverages, milks and energy beverages, as well as different kinds of squeeze
pouches as protein smoothies which have been produced using protein
fractions. A known company developed chia beverages with functional prop-
erties and rich in nutrients. Moreover, another company produced fruit juices
with chia seeds highlighting omega-​3 content. In addition, in 2009 a patent was
created to develop a chia seed beverage with enhanced gastrointestinal regu-
larity and heart health effects (Minatelli et al., 2009).
Finally, Andean and ancestral grains offer a significant nutritional compos-
ition which allows development of fermented and non-​fermented beverages
with distinctive sensorial properties and providing a significant contribution to
the prevention of different diseases.

7.7 INFANT FOODS
According to UNICEF (2019), at least one in three children under the age of five
shows signs of chronic malnutrition –​undernourished or overweight –​and one
in two suffers from hidden hunger, impairing their ability to develop their full
potential. Proteins are especially important in childhood growth and their main
source is in food of animal origin. Moreover, meat and derivatives are not eco-
nomically accessible for the entire population, hence the importance of looking
for other sources of cheaper proteins with a good amino acid profile. In this con-
text, native crops from Latin America play a key role as an economic, sustain-
able, nutritious and versatile opportunity to be used in various infant dishes and
preparations.
In 2011, the Food and Agriculture Organization of the United Nations
(FAO, 2011) considered quinoa as a strategic crop for food security due to its
characteristics: nutritional quality –​represented by its essential amino acid
282 Food Uses of Selected Ancient Grains

composition; wide genetic variability to develop varieties that improve pro-


duction performance; and its adaptation to adverse climate and soil conditions
(Cerezal Mezquita et al., 2012). In the same sense, kiwicha and kañiwa are foods
also recognized for their high protein quality and content of calcium, phos-
phorus, iron, potassium, zinc, vitamin E and B complex (Ayala, 2004; Mujica and
Jacobsen, 2006). Therefore, some combination of these could be an effective sub-
stitute for animal protein (Bhat and Karim, 2009).
A research work regarding the frequency of use of Andean grains in Lima –​
the Peruvian capital –​showed that the majority of mothers (44%) stated that
they used quinoa once or twice per month, 91% of them never used kiwicha
and 100% of them never used kañiwa in their children’s preschool diet. These
interesting results show that in this case local crops are scarcely consumed; des-
pite the necessity of improving children’s diets and to provide a wide variety of
dishes (Baltazar-​Ñahui, 2015).
Repo-​Carrasco-​Valencia and Hoyos (1993) developed an infant dish for
children from 5 to 24 months of age using grains of quinoa, kañiwa, kiwicha,
castilla bean (Vigna sinensis), broad bean (Vicia faba) and rice (Oriza sativa).
The protein content ranged from 11.35 to 15.46% with a good balance of essen-
tial amino acids and good acceptability. Cerezal Mezquita et al. (2011) used
quinoa and lupine (Lupinus albus L.) mixtures with traditional cereals such
as corn and rice, obtaining gluten-​free products with higher protein quality
for children under 24 months of age who suffer from celiac disease. In Peru,
an instant protein mixture was developed with yellow maize, red quinoa,
kiwicha and lentils, as an easy-​to-​prepare option to include in infant diets
(Huamaní-​Condori, 2019). Cerezal Mezquita et al. (2007) developed a similar
food, where quinoa and tarwi were combined with the aim of obtaining a
high protein formulation with additives and chicken flavoring. In this case, a
powder is reconstituted by adding water to obtain porridge for children from
2 to 5 years.
In Argentina, Jiménez et al. (2020) studied quinoa and amaranth flours from
germinated and non-​ germinated grains to develop baby purees, obtaining
products with good sensory acceptance. Furthermore, the nutritional and rheo-
logical properties achieved improved protein digestibility and an appropriate
texture for swallowing, both characteristics which may be taken into consider-
ation during infant food development.
Biscuits are snacks massively consumed by the school population, usually
made with refined wheat flour. In Ecuador, Gaibor-​Monar al. (2016) developed
biscuits based on quinoa and amaranth flours –​to replace wheat flour. The
utilization of these two regional crops improved their nutritional quality, since
these Andean grains have around 16% of protein with an excellent amino acids
balance and some minerals such as iron and calcium. In addition, due to their
7.8 Traditional Foods 283

versatility, they can be used in products such as desserts, popcorn and other
sweets with minimal loss of nutrients during cooking.
Improving food quality in vulnerable sectors, such as childhood, involves a
great challenge because of the high cost of healthy food, lack of correct nutri-
tional information and constant exposure of many children, adolescents and
families to ultra-​processed food advertising.

7.8 TRADITIONAL FOODS
Food consumption is one of the most important acts in human life. Beyond its
nutritional message, food is framed by social, symbolic and material aspects of
human culture. According to UNESCO (2003), traditional gastronomy constitutes
an element of identity, social cohesion and cultural distinction. Regional trad-
itional cuisines constitute part of the intangible heritage of communities, and
set the background where those foods and culinary knowledge and practices
are transmitted to following generations, reinforcing their cultural identity
(Meléndez Torres and Cañez de la Fuente, 2009).
Besides the cultural role of Latin-​American crops, their contribution in
daily nutrition and the possible economic impact in the region are funda-
mental factors. There is enough scientific evidence that favors the total or par-
tial replacement of crops such as rice and wheat or its complement with these
crops (Campos et al., 2018). Therefore, talking about and vindicating dishes with
native crops is of the utmost importance to maintaining traditional culture and
improving nutrition and local economies.
Maize –​with its wide spectrum of colors and shapes –​is one of the crops
with the greatest gastronomic and cultural preponderance, occupying a rele-
vant place in food, as well as in ancestral ritual (Abarza and Schimpf, 2009).
Probably one of the most important cultural roles of dishes prepared with maize
is as an offering, presented in front of the divinity (Mazzetto, 2013). Maize is a
staple food in Mexican and Central-​American diets, and has moved throughout
South America, where it is consumed in a wide range of traditional dishes
such as “tamales” and “arepas” in Colombia and Venezuela, “indio Viejo” and
“guirilas” in Nicaragua, “humitas” in Bolivia and Chile, and “locro” in Argentina
(Tanumihardjo et al., 2020).
In pre-​Hispanic times, amaranth was used in religious ceremonies. According
to Mazzetto (2013), to celebrate the feast of Xiuhtecuhtli –​God of Fire –​“tamales”
were prepared with amaranth leaves, since it was conceived of as the body of
the gods. Its cultivation and consumption were almost eradicated during the
viceroyalty due to the evangelization process (Guerrero-​Jacinto and Viesca-​
González, 2015).
284 Food Uses of Selected Ancient Grains

Quinoa was cultivated and used by pre-​Inca civilizations and its expansion
was consolidated with the Inca Empire, extending from Colombia to Chile and
Argentina, and was replaced by cereals upon the arrival of the Spanish, despite
constituting a staple food of the population at that time. In more recent decades,
quinoa has achieved worldwide renaming, but its regional production is mainly
exported. Local populations –​especially autochthonous –​usually prefer to sell
their quinoa production and buy cheaper commodities, such as wheat or rice.
Therefore, its local consumption is reserved for special occasions and dishes,
such as those mentioned in Table 7.2.

TABLE 7.2 LATIN-​AMERICAN CROPS AND THEIR TYPICAL DISHES


Consumption/​
Crop Typical dish History/​origin preparation
Maize Tortilla Mexico. Its origin is It is consumed filled with
debated, probably from meats, vegetables, with
the Teotihuacan valley cheese. Multiple variants.
1500 BC–​AQ 500.
Maize Tamal o Mesoamerican origin. Stuffed with vegetables,
flour hallaca From Mexico to meat, seasonings. It is
northern Argentina. wrapped in the same corn
Used in rituals and or banana leaves.
tombs as an offering.
Maize Tlacoyos Pre-​Hispanic origin, Thick oval dough filled
dough from Nahuatl tlaoyo with beans, cheese, chili
Mexico and onion.
Whole Pozole Mexico. Pre-​Hispanic In ancient times it was
maize origin, prepared by the made with human flesh
Mexican to venerate the and blood.
god Xipe Tótec.
Maize Humitas From south of Dough made from crushed
dough Colombia to Argentina maize kernels, with goat’s
From quechua: cheese and seasonings,
humint’a wrapped in its green husk
Maize Maize cake From south of Similar to humita, with
dough Colombia to Argentina egg, sugar and cooked in
oven
Maize Tulpo Argentinean northwest Thick soup of cornmeal,
flour and Bolivia sheep fat, vegetables and
jerky (ground in a mortar).
7.8 Traditional Foods 285

TABLE 7.2 (CONTINUED)

Consumption/​
Crop Typical dish History/​origin preparation
Whole Locro Argentinean, Chilean Stew with grated corn,
maize and Bolivian norwest meat, vegetables and
seasonings.
Kernels of Tijtincha Argentinean norwest Cob boiled overnight until
different and Bolivia. Consumed grains are soft, then added
colors especially on August to a beef, lamb, or chicken
1st, Pachamama Day broth, with vegetables
Maize Calapurca Argentina, Chile and Corn stew with mote
Bolivia and beef or llama that is
From Aymara qala cooked on hot stones
phurk’a, qala, ‘stone’,
and phurk’a, ‘roast’, ‘on
the embers.
White or Mazamorra Colombia, Argentina, Dessert with milk, sugar
purple Peru. Widely consumed and vanilla or cinnamon,
maize in Argentina during the cloves, raisins, plums and
(also with Viceroyalty period, until fruits such as pineapple
quinoa/​ the beginning of the and peach.
cañihua) 20th century.
In the Caribbean
region, bananas are
added.
Amaranth Alegría Mexico. Dish declared Honey, peanuts and
(Figure 7.3.D) intangible heritage of raisins.
the city of Santiago de
Tulyehualco.
Red and Olli (cakes) Mexico. They were They were made with
black made to venerate different types of
amaranth Xochipilli God of the amaranth that gave it a
game. different color.
Amaranth Gachas Mexico, Peru With cinnamon, yogurt or
milk for breakfast.
Quinoa Quínoa puree Peru. It is a traditional garnish
with potatoes
Quinoa Chaulafan Ecuador. Local Rice is replaced by quinoa.
variation of Chinese
fried rice. The Peruvian
equivalent is known as
chaufa.

(continued)
286 Food Uses of Selected Ancient Grains

TABLE 7.2 (CONTINUED)

Consumption/​
Crop Typical dish History/​origin preparation
Kañihua Pito de Bolivia, Peru Drink consumed for
flour kañihua breakfast. Sweetened with
honey, cinnamon, and
cloves.
Tarwi Ají de tarwi Ecuador Spicy preparation to dress
locro, stews, empanadas,
meats.
Tarwi Ceviche de Ecuador. Typical dish of It is mixed with fish and
tarwi the coastal area garnished with citrus
fruits such as lime, orange.
Chía Chianpinolli Mexico, Guatemala, El Dense flour that is later
seeds Salvador used as a raw material for
breads, dough, etc.
Chía flour Chapatas Mexico, Guatemala, El Thin bread
Salvador
Chía flour Pinole Mexico, Guatemala, El To prepare energy drinks,
Salvador combined with cinnamon
and honey
Toasted Atole Mexico Hot and thick drink made
chía seeds from corn and chia seeds.
In some places, fruits and
cinnamon were added.
Source: Authors’ own work.

Kañiwa is a less extended crop –​not spread beyond Bolivia and Peru –​with
qualities that make it a nutritional support when others fail. On the other hand,
tarwi is used in a wide variety of dishes ranging from purees, sauces and soups
to main dishes such as stews and a version of ceviche, as well as in desserts, and
tarwi flour can even be used in popular drinks (Mikuy and Mikuy, 2013).
Chia crops have been cultivated in Mexico and Guatemala since pre-​
Columbian times as part of the basic diet along with corn and beans. Aztec
warriors used to be rewarded with bags of chia seeds for their military
exploits and for their widows for their loss on the battlefields. Traditionally,
they were added to drinks, as a thickener and due to their fiber content
(McQuown, 1958).
7.10 Conclusions 287

7.9 
HAUTE CUISINE
Not all ancestral crops have earned a position in haute cuisine restaurants, only
some of them have achieved that recognition, but that is exclusively due the
“celebrity” chefs who have seen the great potential that these grains have. Since
ancestral seeds have been recognized for their contribution to nutrition, in add-
ition to their important characteristics in terms of disease prevention, many of
them have begun to be used in culinary preparations and are beginning to be
desired by many guests in the most prestigious restaurants around the world.
An example would be Ferran Adriá, the Spanish chef with a three-​star restaurant
“elBulli” who has used quinoa in his preparations called Quinoa gelada de foie-​
gras d´anec amb consomé which is frozen duck foie gras quinoa with consommé
(Svejenova et al., 2007). On the other hand, elite chefs from Perú have constructed
the haute Peruvian cuisine redefining the cuisine and positioning themselves as
national and global culinary leaders; examples of these are Gaston Acurio and
Virgilio Martinez who use different grains from Andean highlands in their dishes
(McDonell, 2019).
Haute cuisine is also used in hospital food services which is significantly
correlated with overall patient satisfaction (Cox, 2006). In this context, chefs
are using ancestral seeds to create interesting dishes, for example bulgur wheat
salad with berries, quinoa with roasted chicken and multigrain pancake. On the
other hand, the magazine Time Out from Mexico highlights some haute cuisine
restaurants that are using ancestral seeds to produce elaborate dishes such as
“amaranth tamal coated in salsa verde and quelites” (Time Out Magazine, 2016).
In Seville, Spain, the recognized restaurant Abantal (one Michelin star), uses
chia seed, amarant and quinoa for its preparations (Figure 7.5).

7.10 CONCLUSIONS
Latin-​American crops and especially grains, such as cereals and pseudo-​cereals
are known to be a source of nutrients and bioactive compounds that contribute
to the prevention of many diseases; these outstanding food sources had been
part of the diet of Latin-​American cultures throughout history. In recent decades
and thanks to their recognized nutritional and functional contribution, they
have been reintroduced and today are cultivated and used as a source of food
and ingredients for the elaboration of countless products not only in the Latin-​
American region but around the world. This chapter describes food components,
ingredients, and products with the aim of opening, and reopening, additional
alternatives for these crops, which may be classified as superfoods.
288 Food Uses of Selected Ancient Grains

Figure 7.5 A. Quinoa dish; B. Amaranth-​based preparation with haute cuisine.


Notes: A) Quinoa dish composed of a round, soft orangey, flat-​disk base with cooked
quinoa inside the disk. On top of the quinoa, larger (than quinoa) orange jelly-​like balls
are found and then on top of them a white foam-​like larger ball. Decorated with green
leaves. Source: Courtesy of Julio Fernández, ABANTAL Restaurant, Sevilla, Spain.
B) Amaranth dish with a foam-​like, flat-​disk base, acting as a bowl, where cooked
amaranth is inside. On top of the quinoa rests a small square shaped type of fish.
References 289

ACKNOWLEDGMENTS
Authors would like to express their gratitude to the CYTED Program through
Project119RT0567, FONDECYT Project 1201489 from the National Agency
for Research and Development (ANID, Chile) and Food4ImNut Food4ImNut
PID2019-107650RB-C21 funded by MCIN/AEI/10.13039/501100011033, Spain
for its kind support.

REFERENCES
Abarza, S., and J.H. Schimpf. 2009. Los maíces indígenas en la Cultura alimenticia Andina.
Ciencia, Universidad de Jujuy 4(9): 159–​170.
Aderibigbe, O. R., O.O. Ezekiel, S.O. Owolade, J.K. Korese, B. Sturm, and O. Hensel. 2020.
Exploring the potentials of underutilized grain amaranth (Amaranthus spp.) along
the value chain for food and nutrition security: A review. Critical Reviews in Food
Science and Nutrition 62(3): 656–​669.
Ahmed, I.B.H., A. Hannachi, and C.M. Haros. 2020. Combined effect of Chia flour and soya
lecithin incorporation on nutritional and technological quality of fresh bread and
during staling. Foods 9(4): 446.
Albuja Vaca, D. E., and C.A. Yépez Vizuete. 2017. Desarrollo de pasta de arroz libre de gluten
con sustitución parcial por harina de chocho (Lupinus Mutabilis Sweet) a través de
un diseño experimental de proceso-​mezclas. Bachelor’s thesis, Quito: USFQ. http://​
repo​sito​rio.usfq.edu.ec/​han​dle/​23000/​6408.
Alcantara Brandao N., M. Boges de Lima Dutra, A.L. Andrade Gaspardi, and M.R. Segura-​
Campos. 2019. Chia (Salvia hispanica L.) cookies: Physicochemical/​microbiological
attributes, nutrimental value and sensory analysis. Journal of Food Measurement
and Characterization 13(2): 1100–​1110.
Aranda Tarazona, J. J., and G.I. Bocanegra Reyes. 2018. Evaluación de parámetros durante
la extrusión de una mezcla de harinas de tarwi (Lupinus mutabilis) y arroz (Orysa
sativa) para la producción de un snack. http://​repo​sito​rio.uns.edu.pe/​han​dle/​
UNS/​3052.
Aranibar, C., N. B. Pigni, M. Martínez, A. Aguirre, P. Ribotta, D. Wunderlin, and R. Borneo.
2018. Utilization of a partially-​deoiled chia flour to improve the nutritional and anti-
oxidant properties of wheat pasta. LWT –​ Journal of Food Science and Technology
89: 381–​387.
Arcila, N., and Y. Mendoza. 2006. Elaboration of an instant beverage of amaranth seeds
(Amaranthus cruentus) and its potential use in the human diet. Revista de la
Facultad de Agronomía 23(1): 114–​124.
Ayala, G. 1998. Aporte de los cultivos andinos a la nutrición humana. In Raíces Andinas:
Contribuciones al conocimiento y a la capacitación, 101–​112). Lima: Universidad
Nacional Mayor de San Marcos.
Ayala, G. 2004. Aporte de los cultivos andinos a la nutrición humana. In Raíces
Andinas: Contribuciones al conocimiento y a la capacitación, 2nd edn., 101–​112).
Lima: Universidad Nacional Mayor de San Marcos.
290 Food Uses of Selected Ancient Grains

Ayerza, R., and W. Coates. 2005. Chia: Rediscovering a Forgotten Crop of the Aztecs. Tucson,
AZ: University of Arizona Press.
Bahnassey, Y., and K. Khan. 1986. Fortification of spaghetti with edible legumes.
II. Rheological, processing, and quality evaluation studies. Cereal Chemistry
63(3): 216–​219.
Ballester-​Sánchez, J., J.V. Gil, C.M. Haros, and M.T. Fernández-​Espinar. 2019a. Effect of
incorporating white, red or black quinoa flours on the total polyphenol content,
antioxidant activity and colour of bread. Plant Food for Human Nutrition 89: 837–​843.
Ballester-​ Sánchez, J., E. Yalcin, M.T. Fernández-​ Espinar, and C.M. Haros. 2019b.
Rheological and thermal properties of royal quinoa and wheat flour blends for
breadmaking. European Food Research and Technology 254: 1571–​1582.
Ballester-​Sánchez J., M.C. Millán-​Linares, M.T. Fernández-​Espinar, and C.M. Haros. 2019c.
Development of healthy, nutritional bakery products by incorporation of quinoa.
Foods 8(9): pii: E379.
Ballester-​Sánchez J., M.T. Fernández-​Espinar, and C.M. Haros. 2020. Isolation of red
quinoa fiber by wet-​and dry-​milling and application as a potential functional
bakery ingredient. Food Hydrocolloids 101: 105513.
Baltazar-​Ñahui, R. 2015. Conocimientos, actitudes y prácticas sobre uso de granos andinos
en la alimentación del preescolar de madres en una institución educativa. Thesis,
Universidad Nacional Mayor de San Marcos, Lima, Perú. Retrieved from https://​cyb​
erte​sis.unmsm.edu.pe/​bitstr​eam/​han​dle/​20.500.12672/​5443/​Baltaz​ar_​%c3%b1r.
pdf ?seque​nce=​1&isAllo​wed=​y.
Banerji, A., L. Ananthanarayan, and S. Lele. 2018. Rheological and nutritional studies of
amaranth enriched wheat chapatti (Indian flat bread). Journal of Food Processing
and Preservation 42(1): e13361.
Bautista-​Justo M., A.D. Castro Alfaro, E. Camarena Aguilar, K. Wrobel, K. Wrobel, G.A.
Guzmán, Z. Gamiño Sierra, and V. da Mota Zanella. 2007. Integral bread develop-
ment with soybean, chia, linseed, and folic acid as a functional food for women.
Archivos Latinoamericanos de Nutrición 57(1): 78–​84.
The Bean Institute. 2014. Dry beans commonly consumed (by nationality). Retrieved from
https://​beanin​stit​ute.com/​.
Benítez V., R.M. Esteban, E. Moniza, N. Casado, Y. Aguilera, and E. Molla. 2018. Breads for-
tified with wholegrain cereals and seeds as source of antioxidant dietary fibre and
other bioactive compounds. Journal of Cereal Science 82: 113–​120.
Bergesse, A. E., P.N. Boiocchi, E.L. Calandri, N.S. Cervilla, V. Gianna, C.A. Guzmán, P.
Miranda, P. Montoya, and J.R. Mufari. 2015. Aprovechamiento integral del grano de
Quinoa. Aspectos tecnológicos, fisioquímicos, nutricionales y sensoriales. http://​
hdl.han​dle.net/​11086/​1846.
Bhat, R., and A.A. Karim. 2009. Exploring the nutritional potential of wild and underutilized
legumes. Comprehensive Reviews in Food Science and Food Safety 8(4): 305–​331.
Bilgiçli N., and S. İbanoğlu. 2015. Effect of pseudo cereal flours on some physical, chem-
ical and sensory properties of bread. Journal of Food Science and Technology
52: 7525–​7529.
Bodroza-​Solarov M., B. Filipcev, Z. Kevresan, A. Mandic, and O. Simurina. 2008. Quality
of bread supplemented with popped Amaranthus cruentus grain. Journal of Food
Process Engineering 31: 602–​618.
References 291

Borneo R., A. Aguirre, and A.E. León. 2010. Chia (Salvia hispanica L) gel can be used as
egg or oil replacer in cake formulations. Journal of the American Dietetic Association
110(6): 946–​949.
Bouasla, A., A. Wójtowicz, and M.N. Zidoune. 2017. Gluten-​free precooked rice pasta
enriched with legumes flours: Physical properties, texture, sensory attributes and
microstructure. LWT –​ Journal of Food Science and Technology 75: 569–​577.
Brennan C. S., and E. Samyue. 2004. Evaluation of starch degradation and textural
characteristics of dietary fibre enriched biscuits. International Journal of Food
Properties 7: 647–​657.
Bustos, M. C., M.I. Ramos, G.T. Pérez, and A.E. Leon. 2019. Utilization of Kañawa
(Chenopodium pallidicaule Aellen) flour in pasta making. Journal of Chemistry
2019: 4385045, 1–​8. https://​doi.org/​10.1155/​2019/​4385​045.
Cahill, J. P. 2003. Ethnobotany of chia, Salvia hispanica L. (Lamiaceae). Economic Botany
57(4): 604–​618.
Calderelli A. V. S., M. de Toledo Benassi, J.V. Visentainer, and G. Matioli. 2010. Quinoa and
flaxseed: Potential ingredients in the production of bread with functional quality.
Brazilian Archives of Biology and Technology –​An International Journal 53: 981–​986.
Calisto Guzmán, L. A. 2009. Desarrollo de producto snack a base de materias primias
no convencionales: poroto (Phaseolus vulgaris L.) y quinua (Chenopodium quinoa
Wild). www.repo​sito​rio.uch​ile.cl/​han​dle/​2250/​105​325.
Camacho, L., M. Vasquez, M. Leiva, and E. Vargas. 1988. Effect of processing and methio-
nine addition on the sensory quality and nutritive value of spray-​dried lupin milk.
International Journal of Food Science & Technology 23(3): 233–​240.
Campos, D., R. Chirinos, L. Gálvez Ranilla, and R. Pedreschi. 2018. Chapter Eight
-​Bioactive potential of andean fruits, seeds, and tubers. In F. Toldrá (Ed.),
Advances in Food and Nutrition Research, edited by F. Toldrá (Vol. 84, 287–​343).
New York: Academic Press.
Cárdenas-​Hernández, A., T. Beta, G. Loarca-​Piña, E. Castaño-​Tostado, J.O. Nieto-​Barrera,
and S. Mendoza. 2016. Improved functional properties of pasta: Enrichment with
amaranth seed flour and dried amaranth leaves. Journal of Cereal Science 72: 84–​90.
Carvajal-​Larenas, F.E. 2019. Nutritional, rheological and sensory evaluation of Lupinus
mutabilis food products -​A review. Czech Journal of Food Sciences 37(5): 301–​331.
Carvajal-​Larenas F.E., A.R. Linnemann, M.J.R. Nout, M. Koziol, and M.A.J.S. Van Boekel.
2016. Lupinus mutabilis: Composition, uses, toxicology, and debittering. Critical
Reviews in Food Science and Nutrition 56:1454–​1487.
Carvalho de Almeida M. M., C.R.F. Lopes, A. de Oliveira, S.S. de Campos, A.P. Bilck, R.
Hernandez Barros Fuchs, O. Hess Gonçalves, P. Velderrama, A. Kamal Genena,
and F.V. Leimann. 2018. Textural, color, hygroscopic, lipid oxidation, and sensory
properties of cookies containing free and microencapsulated chia oil. Food and
Bioprocess Technology 11(5): 926–​939.
Cerezal Mezquita, P., A. Carrasco Verdejo, K. Pinto Tapia, N. Romero Palacios, and R.Arcos
Zavala. 2007. Suplemento alimenticio de alto contenido proteico para niños de 2-​
5 años: Desarrollo de la formulación y aceptabilidad. Interciencia 32: 857–​864.
Retrieved from http://​ve.sci​elo.org/​sci​elo.php?scr​ipt=​sci_​artt​ext&pid=​S0378-​
184420​0700​1200​013&nrm=​iso.
292 Food Uses of Selected Ancient Grains

Cerezal Mezquita, P., V. Urtuvia Gatica, V. Ramírez Quintanilla, and R. Arcos Zavala.
2011. Desarrollo de producto sobre la base de harinas de cereales y leguminosa
para niños celíacos entre 6 y 24 meses; II: Propiedades de las mezclas. Nutrición
Hospitalaria 26: 61–​169. Retrieved from http://​sci​elo.isc​iii.es/​sci​elo.php?scr​ipt=​
sci_​artt​ext&pid=​S0212-​161120​1100​0100​019&nrm=​iso.
Cerezal Mezquita, P., E. Acosta Barrientos, G. Rojas Valdivia, N. Romero Palacios, and R.
Arcos Zavala. 2012. Desarrollo de una bebida de alto contenido proteico a partir
de algarrobo, lupino y quinoa para la dieta de preescolares. Nutrición Hospitalaria
27: 232–​ 243. Retrieved from http://​sci​elo.isc​iii.es/​sci​elo.php?scr​ipt=​sci_​artt​
ext&pid=​S0212-​161120​1200​0100​030&nrm=​iso.
Chávez‐Jáuregui, R. N., M.E.M.P. Silva, and J.A. Arěas. 2000. Extrusion cooking process for
amaranth (Amaranthus caudatus L.). Journal of Food Science 65(6): 1009–​1015.
Coda R., C.G. Rizzello, and M. Gobbetti. 2010. Use of sourdough fermentation
and pseudo-​ cereals and leguminous flours for the making of a functional
bread enriched of γ-​aminobutyric acid (GABA). International Journal of Food
Microbiology 137: 236–​245.
Coorey, R., A. Grant, and V. Jayasena. 2012. Effect of chia flour incorporation on the nutri-
tive quality and consumer acceptance of chips. Journal of Food Research 1: 85–​95.
Cox, S.A. 2006. Improving hospital foodservice. Food Technology Magazine 60(6). www.
ift.org/​news-​and-​publi​cati​ons/​food-​tec​hnol​ogy-​magaz​ine/​iss​ues/​2006/​june/​featu​
res/​improv​ing-​hospi​tal-​food​serv​ice.
Dallagnol A.M., M. Pescuma, G.F. de Valdez, and G. Rollán. 2013. Fermentation of quinoa
and wheat slurries by Lactobacillus plantarum CRL 778: proteolytic activity. Applied
Microbiology and Biotechnology 97: 3129–​40.
Dallagnol A.M., M. Pescuma, G. Rollán, M.I. Torino, and G.F. de Valdéz. 2015. Optimization
of lactic ferment with quinoa flour as bio-​preservative alternative for packed bread.
Applied Microbiology and Biotechnology 99: 3839–​3849.
de Meo, B., G. Freeman, O. Marconi, C. Booer, G. Perretti, and P. Fantozzi. 2011. Behaviour
of malted cereals and pseudo-​cereals for gluten-​free beer production. Journal of the
Institute of Brewing 117(4): 541–​546.
de Sousa, M. F., R.M. Guimarães, M. de Oliveira Araújo, K.R. Barcelos, N.S. Carneiro, D.S.
Lima, ... and M.B. Egea. 2019. Characterization of corn (Zea mays L.) bran as a new
food ingredient for snack bars. Food Science and Technology 101: 812–​818.
Deepa, C., S. Sarabhai, P. Prabhasankar, and H.U. Hebbar. 2017. Effect of micronization of
maize on quality characteristics of pasta. Cereal Chemistry 94(5): 840–​846.
Delgado, L., and G. Barraza. 2014. Efecto de la proporción de Chenopodium quinoa
(quinua), Amaranthus caudatus (kiwicha) y Plukenetia volubilis l. (sacha inchi) en
la aceptabilidad general y el análisis proximal de una barra energética. Cientifi-​k
2(2): 56–​70.
Demir, B., and N. Bilgiçli. 2020. Changes in chemical and anti-​nutritional properties of
pasta enriched with raw and germinated quinoa (Chenopodium quinoa Willd.)
flours. Journal of Food Science and Technology 57: 3884–​3892.
Demir, M.K., and M. Kilinc. 2017. Utilization of quinoa flour in cookie production.
International Food Research Journal 24(6): 2394–​2401.
Dias-​Capriles V., E. Lopes-​Almeida, R.E. Ferreira, J.A. Gomes-​Arêas, C. Joy-​Steel, and Y.
Kil Chang. 2008. Physical and sensory properties of regular and reduced-​fat pound
cakes with added amaranth flour. Cereal Chemistry 85: 614–​618.
References 293

EFSA (European Food Safety Authority). 2009. Opinion on the safety of ‘Chia seeds (Salvia
hispanica L.) and ground whole Chia seeds’ as a food ingredient. EFSA Journal
996: 1–​26.
Elizaquível, P., A. Pérez-​Cataluña, A. Yépez, C. Aristimuño, E. Jiménez, P. S. Cocconcelli,
... and R. Aznar. 2015. Pyrosequencing vs. culture-​dependent approaches to ana-
lyze lactic acid bacteria associated to chicha, a traditional maize-​based fermented
beverage from Northwestern Argentina. International Journal of Food Microbiology
198: 9–​18.
EU (European Union). 2020. Authorising an Extension of Use of Chia Seeds (Salvia
hispanica) as a Novel Food and the Change of the Conditions of Use and the
Specific Labelling Requirements of Chia Seeds (Salvia hispanica) under Regulation
(EU) 2015/​2283 of the European Parliament and of the Council and Amending
Commission Implementing Regulation (EU) 2017/​ 2470; Official Journal of the
European Union 12–​16, Brussels.
FAO (Food and Agriculture Organization/​Oficina Regional para America Latina y el
Caribe). 2011. La quinoa: cultivo milenario para contribuir a la seguirdad alimentaria
mundial. Rome: FAO.
Ferrari Felisberto M. H., A.L. Wahanik, C. Rodrigues Gomes-​Ruffi, M.T. Pedrosa Silva
Clerici, Y. Kil Chang, and C. Joy Steel. 2015. Use of chia (Salvia hispanica L.)
mucilage gel to reduce fat in pound cakes. LWT –​ Food Science and Technology
63(2): 1049–​1055.
Fiorda, F. A., M.S. Soares Jr, F.A. da Silva, M.V. Grosmann, and L.R. Souto. 2013.
Microestructure, texture and colour of gluten-​free pasta made with amaranth
flour, cassava starch and cassava bagasse. LWT –​ Food Science and Technology
54(1): 132–​138.
Gaibor-​Monar, F. M., J.P. Torres-​Cadena, and L.V. Yépez-​Martínez. 2016. Valor nutricional
de las galletas a base de amaranto y quinua asociado a la aceptabilidad
microbiológica. Revista Caribeña de Ciencias Sociales 12(8): 1–​20.
Galluzzo F.G., G. Cammilleri, L. Pantano, G. Lo Cascio, A. Pulvirenti, A. Macaluso, A. Vella,
and V. Ferrantelli. 2021. Acrylamide assessment of wheat bread incorporating
chia seeds (Salvia hispanica L.) by LC-​LM/​MS. Food additives and contaminants
Part A -​Chemistry analysis control exposure & risk assessment. DOI: 10.1080/​
19440049.2020.1853823.
García-​Mantrana I., V. Monedero, and M. Haros. 2014. Application of phytases isolated
from bifidobacteria in the development of cereal-​based products with amaranth.
European Food Research and Technology 238: 853–​862.
García-​Mantrana I., V. Monedero, and M. Haros. 2015. Myo-​inositol hexakisphosphate
degradation by Bifidobacterium pseudocatenulatum ATCC 27919 improves min-
eral availability of high fibre rye-​wheat sour bread. Food Chemistry 178: 267–​275.
doi: 10.1016/​j.foodchem.2015.01.099.
Gearhart, C., and K.A. Rosentrater. 2014. Extrusion processing of amaranth and quinoa.
Paper No. 1 41912019, presented at ASABE and CSBE/​SCGAB Annual International
Meeting, Montreal, PQ, Canada July 13–​July 16, 2014, 1–​36. St. Joseph, MI: American
Society of Agricultural and Biological Engineers.
Giménez, M. A., A. Gámbaro, M. Miraballes, A. Roascio, M. Amarillo, N. Sammán, and M.
Lobo. 2015. Sensory evaluation and acceptability of gluten‐free Andean corn spa-
ghetti. Journal of the Science of Food and Agriculture 95(1): 186–​192.
294 Food Uses of Selected Ancient Grains

Giménez, M. A., S.R. Drago, M.N. Bassett, M.O. Lobo, and N.C. Sammán. 2016. Nutritional
improvement of corn pasta-​like product with broad bean (Vicia faba) and quinoa
(Chenopodium quinoa). Food Chemistry 199: 150–​156.
González Rojas, M. J., and A.C. Moya Valenzuela. 2004. Producción y comercialización en
Bogotá de hojuelas de quinua empacadas. http://​hdl.han​dle.net/​10554/​7089.
Goyat Y., S.J. Passi, S. Suri, and H. Dutta 2018. Development of chia (Salvia hispanica, L.)
and quinoa (Chenopodium quinoa, L.) seed flour substituted cookies -​physico-
chemical, nutritional and storage studies. Current Research in Nutrition and Food
Science 6(3): 757–​769.
Güemes-​ Vera N., R.J. Peña-​ Bautista, C. Jiménez-​ Martínez, G. Dávila-​ Ortiz, and G.
Calderón-​Domínguez. 2008. Effective detoxification and decoloration of Lupinus
mutabilis seed derivatives, and effect of these derivatives on bread quality and
acceptance. Journal of the Science of Food and Agriculture 88: 1135–​1143.
Guerrero Jacinto, M. and F. Viesca González. 2016. Recetario de Santiago Tulyehualco, El
Amaranto Como Patrimonio Gastronómico. Benito Juárez, Mexico: Editorial Época.
Guiotto E.N., M.C. Tomás, and C.M. Haros. 2020. Development of high nutritional breads
with by-​products of chia (Salvia hispanica L.) seeds. Foods 9(6): 819.
Hastorf, C. A. 2015. Andean luxury foods: Special food for the ancestors, deities and the
élite. Antiquity 77(297): 545–​554.
Hastorf, C. A., and S. Johannessen. 1993. Pre-​Hispanic political change and the role of
maize in the Central Andes of Peru. American Anthropologist 95(1): 115–​138.
Herndandes Venturini, L., T. Fernades Moya Moreira, T. Barlati Vieira da Silva, M.M.
Carvalho de Almeida, F.C.R. Lopes, A. de Oliveira, S. Silva de Campos, A.P. Bilck,
R. de Souza Leone, A.A. Coelho Tanamati, O. Hess Gonçalves, F.V. Leimann 2019.
Partial substitution of margarine by microencapsulated chia seeds oil in the formu-
lation of cookies. Food and Bioprocess Technology 12(1): 77–​87.
Hickisch, A., R. Beer, R.F. Vogel, and S. Toelstede. 2016. Influence of lupin-​based milk
alternative heat treatment and exopolysaccharide-​producing lactic acid bacteria
on the physical characteristics of lupin-​based yogurt alternatives. Food Research
International 84: 180–​188.
Houben A., H. Götz, M. Mitzscherling, and T. Becker. 2010. Modification of the rheological
behavior of amaranth (Amaranthus hypochondriacus) dough. Journal of Cereal
Science 51: 350–​356.
Huamaní-​Condori, J. 2019. Desarrollo de una mezcla instantánea proteica a partir de maíz
amarillo (Zea mays indurata St.), quinua roja (Chenopodium quinoa Willd), kiwicha
(Amaranthus caudatus) y lenteja (Lens culinaris). Engineering Thesis. Universidad
Peruana Unión, Lima, Perú. Retrieved from http://​repo​sito​rio.upeu.edu.pe/​han​dle/​
UPEU/​2034.
Iglesias-​Puig, E., and Haros, M. 2013. Evaluation of performance of dough and bread
incorporating chia (Salvia hispanica L.). European Food Research and Technology
237(6): 865–​874.
Iglesias-​Puig E., V. Monedero, and M. Haros. 2015. Bread with whole quinoa flour and
bifidobacterial phytases improve contribution to dietary mineral intake and their
bioavailability without substantial loss of bread quality. LWT –​ Food Science and
Technology 60: 71–​77.
References 295

Izquierdo, J., and W. Roca. 1998. Under-​Utilized Andean Food Crops: Status and Prospects
of Plant Biotechnology for the Conservation and Sustainable Agricultural Use of
Genetic Resources. Paper presented at the Symposium on Plant Biotechnology
as A Tool for the Exploitation of Mountain Lands, Leuven, Belgium. www.fao.
org/​temp​ref/​GI/​Reser​ved/​F TP​_​Fao​Rlc/​old/​prior/​rec​nat/​recur​sos/​bio​div/​andi​
nos.pdf.
Jekle M., A. Houben, M. Mitzscherling, and T. Becker. 2010. Effects of selected lactic acid
bacteria on the characteristics of amaranth sourdough. Journal of the Science of
Food and Agriculture 90: 2326–​2332.
Jiménez, D., M. Miraballes, A. Gámbaro, M. Lobo, and N. Sammán. 2020. Baby purees
elaborated with andean crops. Influence of germination and oils in physico-​chemical
and sensory characteristics. LWT –​ Food Science and Technology 124: 108901.
Karovičová, J., Z. Kohajdová, M. Lauková, L. Minarovičová, M. Greifová, J. Hojerová,
and G. Greif. 2020. Utilisation of quinoa for development of fermented beverages.
Potravinarstvo Slovak Journal of Food Sciences 14: 465–​472.
Kaur, I., and B. Tanwar. 2016. Quinoa beverages: Formulation, processing and potential
health benefits. Romanian Journal of Diabetes Nutrition and Metabolic Diseases
23(2): 215–​225.
Kohajdová Z., J. arovičováand, and Š. Schmidt. 2011. Lupin composition and possible use
in bakery –​A review. Czech Journal of Food Science 29(3): 203–​211.
Laparra J. M., and M. Haros. 2016. Inclusion of ancient Latin-​American crops in bread
formulation improves intestinal iron absorption and modulates inflammatory
markers. Food & Function 7(2): 1096–​1102.
Laparra J. M., and M. Haros. 2018. Inclusion of whole flour from Latin-​American crops
into bread formulations as substitute of wheat delays glucose release and uptake.
Plant Foods for Human Nutrition 73(1): 13–​17.
Lemos A. R., V.D. Capriles, M.E. Machado Pinto e Silva, and J.A. Gomes Arêas. 2012. Effect
of incorporation of amaranth on the physical properties and nutritional value of
cheese bread. Ciencia e Tecnología de Alimentos 32: 427–​431.
Levent, H. 2017. Effect of partial substitution of gluten-​free flour mixtures with chia
(Salvia hispanica L.) flour on quality of gluten-​free noodles. Journal of Food Science
and Technology 54(7): 1971–​1978.
Lorence-​ Quiñones, A., C. Wacher-​ Rodarte, and R. Quintero-​ Ramírez. 1999. Cereal
fermentations in Latin American countries. In Fermented Cereals. A Global
Perspective, 99–​114). Rome: FAO Agricultural Services Bulletin.
Lorenz K., and L. Coulter. 1991. Quinoa flour in baked products. Plant Foods Human
Nutrition 41: 213–​224.
Lorusso, A., M. Verni, M. Montemurro, R. Coda, M. Gobbetti, and C.G. Rizzello. 2017. Use
of fermented quinoa flour for pasta making and evaluation of the technological and
nutritional features. Food Science and Technology 78: 215–​221.
Ludena Urquizo, F. E., S.M. García Torres, T. Tolonen, M. Jaakkola, M.G. Pena-​Niebuhr, A.
von Wright, … and C. Plumed-​Ferrer. 2017. Development of a fermented quinoa-​
based beverage. Food Science & Nutrition 5(3): 602–​608.
McDonell, E. 2019. Reproducing “Indian Food”: Race, Value, and Development in Peru’s
Quinoa Boom-​Bust. Bloomington, IN: Indiana University Press.
296 Food Uses of Selected Ancient Grains

McQuown, N. A. 1958. The general history of the things of New Spain, by Bernardino de
Sahagún. The Hispanic American Historical Review 38(2): 235–​238.
Manassero, C. A., M.C. Añón, and F. Speroni. 2020. Development of a high protein bev-
erage based on amaranth. Plant Foods for Human Nutrition 75(4): 599–​607.
Mas A. L., F.I. Brigante, E. Salvucci, N.B. Pigni, M.L. Martinez, P. Ribotta, D.A. Wunderlin,
and M.V. Baroni. 2020. Defatted chia flour as functional ingredient in sweet cookies.
How do processing, simulated gastrointestinal digestion and colonic fermentation
affect its antioxidant properties? Food Chemistry 316: 126279.
Mastromatteo, M., S. Chillo, M. Iannetti, V. Civica, and M.A. Del Nobile. 2011. Formulation
optimisation of gluten‐free functional spaghetti based on quinoa, maize and soy
flours. International Journal of Food Science & Technology 46(6): 1201–​1208.
Matsuo M. 2006. Suitability of quinoa fermented with Rhizopus oligosporus as an ingre-
dient of biscuit. Journal of the Japanese Society for Food Science and Technology
53(1): 62–​69.
Mazzetto, E. 2013. La comida ritual en las fiestas de las veintenas mexicas: un acercamiento
a su tipología y simbolismo. Amérique Latine Histoire et Mémoire 25. doi:https://​doi.
org/​10.4000/​alhim.4461
Meléndez Torres, J. M., and G.M. Cañez de la Fuente. 2009. La cocina tradicional regional
como un elemento de identidad y desarrollo local: el caso de San Pedro El Saucito,
Sonora, México. Estudios sociales (Hermosillo, Son.) 17: 181–​204. Retrieved from
www.sci​elo.org.mx/​sci​elo.php?scr​ipt=​sci_​artt ​ext&pid=​S0188-​455720​0900​0300​
008&nrm=​iso.
Mesías M., F. Holgado, G. Márquez Ruiz, and F.J. Morales. 2016. Risk/​benefit consider-
ations of a new formulation of wheat-​based biscuit supplemented with different
amounts of chia flour. LWT –​ Food Science and Technology 73: 528–​535.
Mikuy, A., and S. Mikuy. 2013. Traditional High Andean Cuisine. Santiago: Food and
Agriculture Organization of the United Nations. Regional Office for Latin America
and the Caribbean (FAO/​RLC).
Milán-​Carrillo, J., A. Montoya-​Rodríguez, R. Gutiérrez-​Dorado, X. Perales-​Sánchez, and
C. Reyes-​Moreno. 2012. Optimization of extrusion process for producing high anti-
oxidant instant amaranth (Amaranthus hypochondriacus L.) flour using response
surface methodology. Applied Mathematics 3(104): 1516–​1525.
Minatelli, J. A., H. Stephen, R. Moerck, and U. Nguyen. 2009. United States Patent No.
US20090181114A1. U. Nutraceuticals.
Miranda-​Ramos K., and C.M. Haros. 2020. Combined effect of chia, quinoa and amaranth
incorporation on the physico-​chemical, nutritional and functional quality of fresh
bread. Foods 9(1859): 1–​22.
Miranda-​Ramos, K., N. Sanz-​Ponce, and C.M Haros. 2019. Evaluation of technological and
nutritional quality of bread enriched with amaranth flour. LWT –​ Food Science and
Technology 114: 108418.
Miranda-​Ramos K., M.C. Millán-​Linares, and C.M. Haros. 2020. Effect of chia by-​products
as breadmaking ingredients on nutritional quality, mineral availability and gly-
caemic index of bread. Foods 9(5): 663.
Montero-​Quintero, Keyla Carolina, Rafael Moreno-​Rojas, Edgar Ali Molina, Máximo
Segundo Colina, and Adriana Beatriz Sánchez-​Urdaneta. 2015. Evaluation of amar-
anth enriched bread for dietary regimes. Interciencia 40(7): 473–​478.
References 297

Mujica, A., and S. Jacobsen. 2006. La quinua (Chenopodium quinoa Willd.) y sus
parientes silvestres. Botánica Económica de los Andes Centrales. In Botánica
Económica de los Andes Centrales, edited by M. Moraes, B. Øllgaard, L.P. Kvist,
F. Borchsenius, and H. Balslev, 449–​457. La Paz, Bolivia: Universidad Mayor de
San Andrés.
Muñoz, L. A., A. Cobos, O. Díaz, and J.M. Aguilera. 2012. Chia seeds: Microstructure, muci-
lage extraction and hydration. Journal of Food Engineering 108(1): 216–​224.
Mykolenko S., D. Zhygunov, and T. Rudenko. 2020. Baking properties of different amar-
anth flours as wheat bread ingredients. Journal of Food Science and Technology-​
Ukraine 14(4). DOI:10.15673/​fst.v14i4.1896
Myrisis, G., S. Aja and C. M. Haros. 2022. Substitution of critical ingredients of cookie
products to increase nutritional value. Biology and Life Sciences Forum 17(1): 15.
https://​doi.org/​10.3390/​bls​f202​2017​015.
Naumova, N., A. Lukin, and V. Erlikh. 2017. Quality and nutritional value of pasta
products with added ground chia seeds. Bulgarian Journal of Agricultural Science
23: 860–​865.
Nawaz, M. A., M. Tan, S. Øiseth, and R. Buckow. 2020. An emerging segment of
functional legume-​based beverages: A review. Food Reviews International 38(5):
1064–​1101.
Nicholson, G. E. 1960. Chicha maize types and chicha manufacture in Peru. Economic
Botany 14(4): 290–​299.
Nidhi B., and V. Indira. 2012. Organoleptic and nutritive scoring of value added cereal
products from grain amaranth (Amaranthus spp.). Indian Journal of Nutrition and
Dietetics 49: 31–​37.
Nielsen, M. A., A.K. Sumner, and L.L. Whalley. 1980. Fortification of pasta with pea flour
and air-​classified pea protein concentrate. Cereal Chemistry 57(3): 203–​206.
OJEU (Official Journal of the European Union). 2009. Authorizing the placing on the
market of Chia seed (Salvia hispanica) as novel food ingredient under regulation
(EC) No 258/​97 of the European Parliament and of the Council 294: 14–​15.
Ordoñez E. S., K.A. Castillo, D. Reátegui, and V.E. Condori. 2019. Elaboration of bread with
the incorporation of coconut pulp flour and sacha inchi (Plukenetia volubilis L.)
nibs. Agroindustrial Science 9(2): 189–​198.
Owusu-​Mensah, E., I. Oduro, and K. J. Sarfo. 2011. Steeping: A way of improving the
malting of rice grain. Journal of Food Biochemistry 35(1): 80–​91.
Pantoja Tirado, L. R., and G.P. Prieto Rosales. 2014. Evaluación tecnológica y sensorial
de pastas alimenticias enriquecidas con harina de quinua (Chenopodium quinua
wild.) y tarwi (Lupinus mutabilis sweet). Masters Thesis. Nuevo Chimbote –​
Perú: Universidad Nacional del Santa.
Park S.H., and N. Morita. 2005. Dough and breadmaking properties of wheat flour
substituted by 10% with germinated quinoa flour. Food Science and Technology
International 11: 471–​476.
Patil, S. S., M.A. Brennan, S.L. Mason, and C.S. Brennan. 2016. The effects of fortification
of legumes and extrusion on the protein digestibility of wheat based snack. Foods
5(2): 26.
298 Food Uses of Selected Ancient Grains

Peláez, P., D. Orona-​Tamayo, S. Montes-​Hernandez, M.E. Valverde, O. Paredes-​López, and


A. Cibrian-​Jaramillo. 2019. Comparative transcriptome analysis of cultivated and
wild seeds of Salvia hispanica (chia). Scientific Reports 9: 9761.
Pepe Guato, M. F. 2011. Comparación de las Mezclas de Harina de Trigo (Triticum spp) y
Chocho (Lupinus mutabilis) en la Evaluación Sensorial de Pastas. Bachelor’s Thesis).
http://​repo​sito​rio.uta.edu.ec/​han​dle/​123456​789/​3055.
Pinares Huamaní, C. 2019. Formulación de pasta alimenticia (Tallarín de Casa) con
sustitución parcial de Tarwi (Lupinus mutabilis Sweet) y adición de cáscara de
huevo en polvo. http://​repo​sito​rio.una​mba.edu.pe/​bitstr​eam/​han​dle/​UNA​MBA/​
818/​T_​0​509.pdf ?seque​nce=​1&isAllo​wed=​y.
Poleth Carvajal Basantes, S. 2019. Efecto de los parámetros de extrusión sobre la
composición nutricional de un snack a base de maíz, fréjol y camote. http://​repo​
sito​rio.utn.edu.ec/​bitstr​eam/​123456​789/​8827/​2/​ARTIC​ULO%20.pdf
Ponce, M., D. Navarrete, and M.G. Vernaza. 2018. Sustitución parcial de harina de trigo
por harina de lupino (lupinus mutabilis sweet) en la producción de pasta larga.
Información Tecnológica 29(2): 195–​204.
Poutanen K., N. Sozer, and G. Della Valle. 2014. How can technology help to deliver
more of grain in cereal foods for a healthy diet? Journal of Cereal Science 59:
327–​336.
Raihan M., and C.S. Saini. 2017. Evaluation of various properties of composite flour
from oats, sorghum, amaranth and wheat flour and production of cookies thereof.
International Food Research Journal 24(6): 2278–​2284.
Ramos Diaz, J. M. 2015. Use of amaranth, quinoa, kañiwa and lupine for the development
of gluten-​free extruded snacks. http://​urn.fi/​URN:ISBN:978-​951-​51-​1657-​4, http://​
hdl.han​dle.net/​10138/​157​282.
Ranilla, L. G., E. Apostolidis, M.I. Genovese, F.M. Lajolo, and K. Shetty. 2009. Evaluation
of indigenous grains from the Peruvian Andean Region for antidiabetes and
antihypertension potential using in vitro methods. Journal of Medicinal Food
12(4): 704–​713.
Rayas-​Duarte, P., C.M. Mock, and L.D. Satterlee. 1996. Quality of spaghetti
containing buckwheat, amaranth, and lupin flours. Cereal Chemistry 73(3):
381–​387.
Repo-​Carrasco-​Valencia, R., and N.L. Hoyos. 1993. Elaboración y evaluación de alimentos
infantiles con base en cultivos andinos. Archivos Latinoamericanos de Nutrición
43(2): 168–​175.
Repo-​Carrasco-​Valencia, R., C. Espinoza, and S.E. Jacobsen. 2003. Nutritional value and
use of the Andean crops quinoa (Chenopodium quinoa) and kañiwa (Chenopodium
pallidicaule). Food Reviews International 19(1–​2): 179–​189.
Repo-​Carrasco-​Valencia, R., A.A. de la Cruz, J.C.I. Alvarez, and H. Kallio. 2009. Chemical
and functional characterization of kañiwa (Chenopodium pallidicaule) grain,
extrudate and bran. Plant Foods for Human Nutrition 64(2): 94–​101.
Repo-​Carrasco-​ Valencia, R., J.J. Pilco, and C.R. Encina-​ Zelada. 2011. Desarrollo y
elaboración de un snack extruido a partir de quinua (Chenopodium quinoa Willd.) y
maíz (Zea mays L.). Ingeniería Industrial 29: 209–​224.
References 299

Rodrigues dos Reis Gallo, L., R. Braz Assunção Botelho, V. Cortez Ginani, L. de Lacerda
de Oliveira, R. Figueiredo Resende Riquette, and E. dos Santos Leandro. 2020. Chia
(Salvia hispanica L.) Gel as egg replacer in chocolate cakes: Applicability and micro-
bial and sensory qualities after storage. Journal of Culinary Science & Technology
18(1): 29–​39.
Rodríguez G., S. Avellaneda, R. Pardo, E. Villanueva, and E. Aguirre. 2018. Bread
leaf enriched with extruded cake from sacha inchi (Plukenetia volubilis
L.): Chemistry, rheology, texture and acceptability. Scientia Agropecuaria 9(2):
199–​208.
Rojas, W., J.L. Soto, M. Pinto, M. Jäger, and S. Padulosi. 2010. Granos andinos: avances,
logros y experiencias desarrolladas en quinua, cañahua y amaranto en Bolivia.
www.biov​ e rsi​ tyin​ t ern​ a tio​ n al.org/​ f ilead​ m in/​ _ ​ m igra​ t ed/​ u plo​ a ds/​ t x_​ n ​ e ws/​
Granos_​andinos_​_​avances_​_​logros_​y_​experiencias_​desa​rrol​lada​s_​en​_​qui​nua_​_​
ca%c3%b1ahua​_​y_​a​mara​nto_​en_​B​oliv​ia_​1​413.pdf.
Rosell, C. M., G. Cortez, and R. Repo-​Carrasco. 2009. Breadmaking use of Andean crops
quinoa, kañiwa, kiwicha, and tarwi. Cereal Chemistry 86: 386–​392.
Salvador-​Reyes, R., and M.T. Pedrosa Silva Clerici. 2020. Peruvian Andean maize: General
characteristics, nutritional properties, bioactive compounds, and culinary uses.
Food Research International 130: 108934.
Salvatierra-​Pajuelo Y.M., M.E. Azorza-​ Richarte, and L.M. Paucar-​ Menacho. 2019.
Optimization of the nutritional, textural and sensorial characteristics of cookies
enriched with chia (Salvia hispanica) and oil extracted from tarwi (Lupinus
Mutabilis). Scientia Agropecuaria 10(1): 7–​17.
Sánchez-​Paz L.A., C. Pérez-​Alonso, O. Dublan-​García, J.C. Arteaga‐Arcos, M. Mayorga‐
Rojas, L. Romero‐Salazar, and M. Díaz‐Ramírez 2020. Effect of a mixture of canola-​
chia oils and gelatin addition on a pound cake reduced in margarine. Journal of Food
Processing and Preservation 44(1): e14298.
Sandak R.N., and A.A.A. El-​Hofi. 2000. Utilization of barley and Amaranthus seeds in bis-
cuit processing. Egyptian Journal of Agricultural Research 78: 1241–​1252.
Santos Fernandes S., and M.M. Salas-​Mellado. 2017. Addition of chia seed mucilage for
reduction of fat content in bread and cakes. Food Chemistry 227: 237–​244.
Sanz-​Penella J.M., and C.M. Haros. 2017. Food uses of whole pseudocereals.
Chapter 8 In: Pseudocereals: Chemistry and Technology, edited by C.M. Haros
and R. Schönlechner, chap. 8, 163–​ 192. Chichester: John Wiley & Sons.
ISBN: 978-​1-​118-​93828-​7.
Sanz-​Penella J.M., J.M. Laparra, Y. Sanz, and M. Haros. 2012. Bread supplemented with
amaranth (Amaranthus cruentus): effect of phytates on in vitro iron absorption.
Plant Foods for Human Nutrition 67: 50–​56.
Sanz-​Penella J.M., M. Wronkowska, M. Soral-​Śmietana, and M. Haros. 2013. Effect of
whole amaranth flour on bread properties and nutritive value. LWT –​ Food Science
and Technology 50: 679–​685.
Sayed-​Ahmad B., T. Talou, E. Straumite, M. Sabovics, Z. Kruma, Z. Saad, A. Hijazi, and O.
Merah. 2018. Evaluation of nutritional and technological attributes of whole wheat
based bread fortified with chia flour. Foods 7: 35.
300 Food Uses of Selected Ancient Grains

Sazesh, B. and M. Goli. 2020. Quinoa as a wheat substitute to improve the textural prop-
erties and minimize the carcinogenic acrylamide content of the biscuit. Journal of
Food Processing and Preservation 44(1): 14563. DOI:10.1111/​jfpp.14563.
Schöenlechner, R., J. Drausinger, V. Ottenschlaeger, K. Jurackova and E. Berghofer. 2010.
Functional properties of gluten-​free pasta produced from amaranth, quinoa and
buckwheat. Plant Foods for Human Nutrition 65(4): 339–​349.
Simurina O., M. Bodrom-​Solarov, and B. Filipcev. 2005. Expanded Seed of Amaranthus
Cruentus as a Component in Speciality Bread Formulations. Paper presemted at the
2nd Central European Meeting/​5th Croatian Congress of the Food Technologists,
Biotechnologists and Nutritionists, Opatija, Croatia, October 17–​20, 2004.
Sindhuja A., M.L. Sudha, and A. Rahim. 2005. Effect of incorporation of amaranth flour on
the quality of cookies. European Food Research and Technology 221: 597–​601.
Song, K.Y., K.Y. Joung, S.Y. Shin, and Y.S. Kim. 2019. Effects of chia (Salvia hispanica L.)
seed roasting conditions on quality of cookies. Italian Journal of Food Science
31(1): 54–​66.
Stikic R., D. Glamoclija, M. Demin, B. Vucelic-​Radovic, Z. Jovanovic, D. Milojkovic-​
Opsenica, S.E. Jacobsen, and M. Milovanovic. 2012. Agronomical and nutritional
evaluation of quinoa seeds (Chenopodium quinoa Willd.) as an ingredient in bread
formulations. Journal of Cereal Science 55: 132–​138.
Sudha, M. L., and K. Leelavathi. 2012. Effect of blends of dehydrated green pea flour and
amaranth seed flour on the rheological, microstructure and pasta making quality.
Journal of Food Science and Technology 49(6): 713–​720.
Svejenova, S., C. Mazza, and M. Planellas. 2007. Cooking up change in haute
cuisine: Ferran Adrià as an institutional entrepreneur. Journal of Organizational
Behavior 28(5): 539–​561.
Tanumihardjo, S. A., L. McCulley, R. Roh, S. López-​Ridaura, N. Palacios-​Rojas, and N.S.
Gunaratna. 2020. Maize agro-​food systems to ensure food and nutrition security in
reference to the Sustainable Development Goals. Global Food Security 25: 100327.
Time Out Magazine. 2016. Haute cuisine in Mexico City. Time Out. www.timeou​tmex​ico.
mx/​ciu​dad-​de-​mex​ico/​resta​uran​tes/​haute-​cuis​ine-​in-​mex​ico-​city.
Tiwari, U., M. Gunasekaran, R. Jaganmohan, K. Alagusundaram, and B.K. Tiwari. 2011.
Quality characteristic and shelf life studies of deep-​fried snack prepared from rice
brokens and legumes by-​product. Food and Bioprocess Technology 4(7): 1172–​1178.
Tömösközi S., L. Gyenge, Á. Pelcéder, T. Abonyi, R. Schönlechner, and R. Lásztity. 2011.
Effects of flour and protein preparations from amaranth and quinoa seeds on the
rheological properties of wheat-​flour dough and bread crumb. Czech Journal of
Food Science 29: 109–​116.
Toralva Aylas, A. D., and M. Rodas Pingus. 2015. Efecto de la sustitución parcial de la
harina de trigo por torta de sacha inchi (Plukenetia volubilis L) sobre las propiedades
reológicas y sensoriales en el bizcocho. Peru: Universidad Nacional de Callao,
Facultad de Ingeniería Pesquera y de Alimentos, Escuela Profesional de Ingeniaría
de Alimentos.
Toralva Aylas A. D., M. Rodas Pingas, and D.M. Guerrero Alva. 2015. Effect of the par-
tial replacement of wheat flour by cake of sacha inchi (Plukenelia volubilis L.) on
the rheological properties of dough in sweet bread. Revista Ciencia & Desarrollo
10: 16–​21.
References 301

Torres Palacios, L. M., I.F. Pallares, and M.P. Tarazona Diaz. 2019. Development of a sweet
biscuit reduced in fat and sugar, fortified with amaranth flour. Nutricion Clinica y
Dietetica Hospitalaria 39(2): 135–​140.
Tosi, E. A., E.D. Re, R. Masciarelli, H. Sanchez, C. Osella, and M.A. de la Torre. 2002. Whole
and defatted hyperproteic amaranth flours tested as wheat flour supplementation
in mold breads. LWT –​ Food Science and Technology 35: 472–​475.
Tronc, E. 1999. Lupin flour: A new ingredient for human food. Grains Legumes 25: 3, 24.
Turck D., J. Castenmiller, S. de Henauw, K. Hirsch-​ Ernst, J. Kearney, A. Maciuk, I.
Mangelsdorf, H. McArdle, A. Naska. C. Pelaez et al. 2019a. Safety of chia seeds
(Salvia hispanica L.) as a novel food for extended uses pursuant to Regulation (EU)
2015/​2283. European Food Safety Authority Journal 17: e05657.
Turck D., J. Castenmiller, S. de Henauw, K. Hirsch-​ Ernst, J. Kearney, A. Maciuk, I.
Mangelsdorf, H. McArdle, A. Naska. C. Pelaez et al. 2019b. Safety of chia seeds
(Salvia hispanica L.) powders, as novel foods, pursuant to Regulation (EU) 2015/​
2283. European Food Safety Authority Journal 17: e05716.
UNESCO (Producer). (2003). Convención para la Salvaguardia del Patrimonio Cultural
Inmaterial. Instrumentos Normativoswe. https://​es.une​sco.org/​about-​us/​legal-​affa​
irs/​con​venc​ion-​salva​guar​dia-​del-​pat​rimo​nio-​cultu​ral-​inm​ater​ial.
UNICEF. (2019). Estado Mundial de la Infancia 2019 –​Niños, alimentos y nutrición: Crecer
bien en un mundo en transformación. Retrieved from www.uni​cef.org/​es/​infor​
mes/​est​ado-​mund​ial-​de-​la-​infan​cia-​2019.
Vásquez Osorio D. C., J.D. Jaramillo Ramírez, G.A. Hincapié Llanos, and L.M. Vélez Acosta.
2017. Development of cookies with sacha inchi (Plukenetia volubilis l.) flour coming
from residual cake. UGCiencia, 23: 101–​113.
Verduga Verdezoto, K. A. 2019. Elaboración de una barra energética a partir de Sacha Inchi.
Bachelor’s Thesis. Quito: UCE). www.dsp​ace.uce.edu.ec/​han​dle/​25000/​19677.
Wang S., A. Opassathavorn, and F. Zhu. 2015. Influence of quinoa flour on quality
characteristics of cookie, bread and Chinese steamed bread. Journal of Texture
Studies 46: 281–​292.
Wang, S., F. Zhu, and Y. Kakuda. 2018. Sacha inchi (Plukenetia volubilis L.): Nutritional
composition, biological activity, and uses. Food Chemistry 265: 316–​328.
Watanabe, K., M. Kawanishi-​Asaoa, C. Myojin, S. Awata, K. Ofusa, and K. Kodama. 2014.
Amino acid composition, oxidative stability, and consumer acceptance of cookies
made with quinoa flour. Food Science and Technology Research 20: 687–​691.
Wright, E. M. 2008. Mapping QTL for agronomic and canning quality traits in black
bean (Phaseulus vulagaris L.). Masters Thesis. East Lansing, MI: Michigan State
University.
Yovera Méndez S. L., and L.M. Ruddy Gianny. 2018. Palitos integrales de semilla de sacha
inchi (Plukenetia volubilis) y linaza (Linum usitatissimum) y efecto dietético en
escolares con sobrepeso. Masters Thesis. Peru: Universidad Nacional Jose Faustino
Sánchez Carrión Facultad de Bromatología y Nutrición Escuela Profesional de
Bromatologia y Nutriciónhuacho.
Zanwar S.R., and V.S. Pawar. 2015. Preparation of biscuit from amaranth, sago and
Echinochloa colonum (Bhagar). Bioinfolet -​A Quarterly Journal of Life Sciences
12: 254–​255.
302 Food Uses of Selected Ancient Grains

Zhao, Y. H., F.A. Manthey, S.K. Chang, H.J. Hou, and S.H. Yuan. 2005. Quality characteristics
of spaghetti as affected by green and yellow pea, lentil, and chickpea flours. Journal
of Food Science 70(6): 371–​376.
Zula A. T., D.A. Ayele, and W.A. Egigayhu. 2020. Proximate, antinutritional, microbial,
and sensory acceptability of bread formulated from wheat (Triticum aestivum)
and amaranth (Amaranthus caudatus). International Journal of Food Science
9429584.
Chapter 8

Nutritional Composition, Bioactive


and Anti-​Nutritional Compounds
of Latin-​American Crop Grains
Norma Sammán1, María C. Rossi1, Sonia Calliope1,
and Ritva Ann-​Mari Repo-​Carrasco-​Valencia2
1Universidad Nacional de Jujuy and Consejo

Nacional de Investigaciones Científicas y Técnicas


(CONICET), San Salvador de Jujuy, Argentina
2Universidad Nacional Agraria La Molina, Peru

CONTENTS
8.1 Nutritional Composition of Latin-​American Grains of Interest 304
8.2 Macronutrient Content 304
8.3 Fatty Acids Profile 309
8.4 Amino Acids Profile 311
8.5 Vitamins and Minerals 314
8.6 Bioactive Compounds 316
8.7 Anti-​Nutritional Compounds In Latin-​American
Grains of Interest 319
8.8 Effects of Different Technological and Culinary
Procedures on Nutritional Composition 321
8.9 Changes in Nutritional Value and Modifications of Bioactive
Compounds Content Through Processing 322
8.10 Conclusions 328
Acknowledgments 330
References 330

DOI: 10.1201/9781003088424-8 303


304 Compounds of Latin-American Crop Grains

8.1 NUTRITIONAL COMPOSITION OF LATIN-​AMERICAN


GRAINS OF INTEREST
Latin America produces food crops which have great importance worldwide.
Among them are maize (Zea maize), quinoa (Chenopodium quinoa), amaranth
(Amaranthus spp.), chia (Salvia hispanica L.), beans such as black turtle bean
(Phaseolus vulgaris), tarwi (Lupinus mutabilis), and also some lesser-​known ones,
such as kañiwa (Chenopodium pallidicaule) and sacha inchi (Plukenetia spp.).
Latin-​American grains are an important source of macronutrients, vitamins,
minerals, and bioactive compounds. Consumption of these grains, especially
quinoa has increased mainly due to their nutritional profile, in particular, their
high protein value and mineral content (Motta et al., 2016a, 2016b).
Suitable nutrition is a critical element in the prevention of numerous
diseases related to current lifestyle conditions. The leading chronic diseases
due to unhealthy food habits nowadays are cancer, diabetes, and other non-​
communicable diseases. With growing global health awareness due to the
increasing number of non-​communicable diseases, demand for healthier food
options that prevent and control these diseases has increased. Furthermore, the
main challenge today, and even in future, will be providing sufficient safe and
healthy foods –​including the crops already mentioned –​in populations’ diets to
help meet these challenges.
The word corn has been used over time and it has been given different
meanings; after the discovery of America, Europeans called it Indian corn
and named it “mahis” in native American language. Later it was called maize
(Galinat, 1971; Salvador-​ Reyes and Pedrosa Silva Clerici, 2020). So in this
chapter, the name “maize” will refer to the native races, with production aimed
mainly at local markets. The terms “grains” and “seeds” are used along the text
indiscriminately.

8.2 MACRONUTRIENT CONTENT
Maize is a very old crop and originated through the domestication process
carried out by ancient inhabitants of Mesoamerica, from teosinte, which used to
grow naturally in Mexico. Due to its great genetic diversity as well as the influ-
ence of different agroecological conditions where it is grown, its nutritional
composition presents great variability. Maize has been used in all areas since
ancient times as a food (Bañuelos Pineda et al., 2018).
The protein content in maize races varies between 5.7–​12.5 g/​100 g dry weight
(db) (Bressani, 1991; Bello Perez et al., 2016), and is mainly distributed in the
germ and endosperm. Carbohydrates are their major component, specifically
8.2 Macronutrient Content 305

starch, which represents 63–​73 g/​100 g db; it is found mainly in the endosperm
and is an energy reserve for the embryo. Digestibility of maize starch decreases
with the increasing amylose content and the length of the branched-​chain of
amylopectin (Rincon Londoño et al., 2016).
Maize also contains, although to a lesser extent, other important components
for the diet such as lipids and fiber (Gimenez et al., 2020). The grains of native
maize have lipid content from 2.70–​6.14 g/​100 g. According to Navarro-​Cortez
et al. (2016), it depends on the races as well as climatic factors, such as drought.
Lipids are found mainly in the germ (76–​83%), then in the aleuronic layer (13–​
15%) and pericarp and endosperm (1–​2%). Regarding fiber content, several
studies have shown lower postprandial glycemic and insulin responses after the
ingestion of high-​fiber corn-​based foods (Lv et al., 2017; Colín-​Chávez, 2020).
The macronutrient content of maize reported in several studies are shown in
Table 8.1.
Amaranth is considered to be one of the oldest crops in Latin America; there
is evidence that it was used by the Mayas, Aztecs, and Incas approximately
10,000 years BC. Amaranth grains have significant amounts of proteins and
starch; they are also an excellent source of insoluble fiber, principally lignin and
cellulose. According to Bhosale et al. (2021) and Bressani (2018), the total fiber
content in amaranth is higher compared to widespread cereals. The lipid fraction
of amaranth grain represents from 6–​10%; it is composed of triacylglycerides,
fatty acids, phospholipids, squalene, phytosterols, and tocopherols. Amaranth,
unlike quinoa, contains few bitter substances such as saponins that must be
eliminated prior to consumption (Coelho, 2018). The nutritional composition of
the grains described are presented in Table 8.1.
Quinoa (Chenopodium quinoa) is a native species of the South American Andes
and its center of origin would be the surroundings of Lake Titicaca. This species
has been cultivated in the Andes for the past 5,000 to 7,000 years, and during that
time has spread to altitudes around 4,000 meters above sea level and to sea level
in southern Chile. This distribution reveals its high adaptability, which, together
with the nutritional quality of its protein, has led to its reevaluation as a new
alternative crop that has extended its cultivation to other regions outside Latin
America (Abugoch-​James, 2009). The protein content ranges from 12.1 to 20.8
g/​100 g) and in particular its soluble protein value is remarkable; albumins and
globulins (31 and 37 g/​100 g protein respectively) were identified as the major
fractions. Another important aspect is that quinoa grains do not contain gluten,
which allows their inclusion in foods for celiac or gluten intolerant patients. In
addition, quinoa contains 49 g/​100 g of carbohydrates; quinoa, like other Andean
cereals, has higher sugar content compared to common cereals (6.2 g/​100 g db).
Starch represents most of the carbohydrate content (approximately 55 g/​100 g
newgenrtpdf
306
Compounds of Latin-American Crop Grains
TABLE 8.1 PROXIMATE COMPOSITION OF LATIN-​AMERICAN GRAINS*
Grain Moisture Proteins* Lipids* CH* Fiber* Ashes* References
Maize 8–​19 5.7–​12.5 2.70–​6.14 63–​73 8.47–​19.5 1.13–​1.84 Cazares Sánchez et al. (2015)
Amaranth 10.50 13.4 6.43 55.30 11.30 2.89 Coelho et al. (2018)
Quinoa 11.30 12.10 6.31 57.20 10.40 2.01 Abugoch-​James (2009);
Gargiulo et al. 2019
Chia 5.82 25.32 30.22 34.57 37.50 4.07 Ullah et al. (2016)
Kaniwa 11.1 15.4 7.2 66.5 24 4,4 Repo-​Carrasco-​Valencia et al.
(2009, 2019)
Tarwi 7.7 44.3 16.5 28.2 7.1 3.3 Huaccho-​Human and
Lope-​Surichaqui (2007)
Turtle bean 5–​8.2 25.66–​25.93 1.52–​1.59 67.83–​68.09 5.4–​7.6 4.71–​4.65 Hayat et al. (2014)
Sacha Inchi 3.3–​8.32 24.2–​27.0 33.4–​54.3 13.4–​30.9 72.4 2.7–​6.46 Wang et al. (2018)
Notes: CH: Carbohydrates; *Values expressed as g/​100 g dry basis.
8.2 Macronutrient Content 307

db) and is located in the perisperm. The starches are rich in amylopectin that
gelatinizes at low temperatures (Haros and Schöenlechner, 2017). The lipid con-
tent of quinoa is two to three times higher compared to common cereals such
as buckwheat and wheat, reaching 4 g/​100 g db (Vilcacundo and Hernández-​
Ledesma, 2017).
Kañiwa (Chenopodium pallidicaule) is a little-​known Andean native grain. It is
very closely related to quinoa (Chenopodium quinoa Willd) and was considered
as a variety of quinoa until it was classified as a species of its own (Gade, 1970).
Kañiwa does not contain detectable quantities of saponins. It is one of the most
resilient crops cultivated in the Andes. For highland farmers, kañiwa is very
important because it is the only crop that can resist frost. The most intensive
production of kañiwa occurs in the southern Andes of Peru and Bolivia in the
surroundings of Lake Titicaca. The grains of kañiwa are much smaller than
quinoa´s as can be seen in Figure 8.1.
The protein content of kañiwa seeds is between 14.4 and 16.9 g/​100 g.
Bioactive peptides in Andean grains have been studied by Chirinos et al. (2018),
and these peptides show antidiabetic, antioxidant, and antihypertensive prop-
erties in vitro. The fat content of kañiwa is between 5.7 and 8.9 g/​100 g and is
mainly located in the embryo. Its oil content is considerably higher compared
to common cereal grains. Kañiwa is an important source of dietary fiber, which
is mainly insoluble, as is common in all grains.
All these crops, but specifically maize, quinoa, amaranth, and kañiwa are
included in recipes that are part of the diet of native people in America. After
the conquest process, foods were mixed to form new dishes that up until today
continue to be part of gastronomic cultural identity (García, 2015). Maize is
used in foods such as tortillas, arepas, snacks, and breakfast cereals, as well
as being a source of starch to use as main products or ingredients in the food
industry. Amaranth can be consumed in the form of sprouts, young leaves in
salad, or ground to be served in soups, among many other preparations. In the
form of flour, quinoa and amaranth are used as raw materials to make tortillas,
bread, cookies, cakes, smoothies, strains, energy bars, cakes, and countless other
products.
Chia (Salvia hispanica L.) seeds are brown with dark brown and white
irregular spots (Alcocer Villacís, 2018); they provide significant amounts of
lipids, proteins, carbohydrates, and dietary fiber. On the other hand, chia’s high
fiber content makes its inclusion in the diet meet the recommended daily intake
of this component (Ullah et al., 2016). The human health benefits of chia seeds
have drawn much attention from researchers and consumers because of the
essential fatty acids, high protein levels, and other components such as phenolic
compounds (Zettel and Hitzmann, 2018; Katunzi-​Kilewela et al., 2021).
308 Compounds of Latin-American Crop Grains

Figure 8.1 Quinoa grain Amarilla Sacaca variety: (a) raw; (b) washed. Kañiwa grain
Cupi variety: (c) raw; (d) washed Source: Repo-​Carrasco-​Valencia (unpublished).

Tarwi or chocho (Lupinus mutabilis) is a legume of high nutritional value that


is distinguished by its protein content and agronomic characteristics such as
rusticity, its ability in the uptake and fixation of atmospheric nitrogen in the soil
benefiting crops, and its adaptability to dry ecological environments, located
between 2,800 and 3,600 meters above sea level. Also, tarwi has a high lipid con-
tent rich in mono-​and polyunsaturated fatty acids. Regarding starch, it is usually
absent, and the carbohydrates are mainly oligosaccharides (e.g., stachyose and
raffinose) and the cell wall stores polysaccharides. The macronutrient content
of tarwi grains are shown in Table 8.1. Since chronic diseases such as obesity
and diabetes are increasing, sugar and certain carbohydrates are considered
the main culprits of bad health, while unsaturated fats and proteins are greatly
needed as part of a balanced diet. Tarwi’s nutritional composition makes it a very
suitable ingredient of healthier foods; it is used mainly as flour to enrich bread,
cereal bars, and specific foods for different age groups (Suca, 2015; Salvatierra-​
Pajuelo et al., 2019).
Sacha inchi (Plukenetia spp.), also known as Inca peanut, wild peanut, sacha
peanut, or mountain peanut, is an oilseed plant that grows in the lowlands of
the Peruvian Amazon. It has been cultivated and consumed for centuries, as
demonstrated by the clay pots in the shape of its fruit, which were found in
8.3 Fatty Acids Profile 309

pre-​Inca cultures, traditionally part of various indigenous dishes in the Peruvian


Amazon. The proximal composition shows that the seeds have remarkable
high-​fat content, between 34.23 and 49.28% in different varieties (Arévalo et al.,
2019). Sacha inchi seeds spark attention as an important and alternative source
of omega-​3, which shows health benefits in lifestyle-​related diseases (Chirinos
et al., 2013).
Common beans (Phaseolus vulgaris L.) are economically important edible
seeds cultivated and consumed worldwide. P. vulgaris stands out as a source
of protein, starch, and dietary fiber. There is information on the slower digest-
ibility of bean starch compared to the same cereal macromolecules. This desir-
able behavior of bean starch is associated with its amylose content, the length
of the amylopectin branched chain, and its tendency to retrograde (Los et al,
2018). Starches that are digested slowly show a low glycemic index which has
shown health benefits like improving glucose tolerance in both healthy people
and diabetics. Other studies have shown that a diet characterized by a low gly-
cemic index reduces risk factors for diabetes and dyslipidemia (Bello-​Pérez
et al., 2020). The proximate composition of this legume can be observed in
Table 8.1.

8.3 FATTY ACIDS PROFILE


The plant triacylglycerol provides essential fatty acids (EFA); humans cannot
synthesize them and they are required for good health. There are two families of
EFA, and the fatty acids that give rise to them are linoleic (ω-​3) and α-​linolenic
(ω-​6). They must be provided by foods.
The total content of unsaturated fatty acids in maize oils is greater than 80%
which could provide positive health effects (Serna-​Saldivar et al. 2017). Palmitic
(11.4–​15.0%), stearic (2.2–​3.5%), oleic (31.4–​46.6%), and linoleic (39.0-​53.0%)
fatty acids are mainly found in native maize oils (Guzman et al. 2015). The fatty
acid composition of this grain is shown in Table 8.2.
The lipids of amaranth grains are mainly found in the germ, most of which
are polyunsaturated; the fatty acid composition varies according to the cultivar,
growing conditions, and genetic traits of the seed (D’Amico et al., 2017). Ranges
of the main fatty acids in amaranth are shown in Table 8.2.
Regarding quinoa, the predominant fatty acid groups are polyunsaturated
and monounsaturated. Fatty acid values in the grain are 8.1, 52.3, and 23 g/​100 g
of linolenic, linoleic, and oleic fatty acids, respectively.
Chia oil contains mainly α-​linolenic, linoleic, oleic, palmitic, and stearic acids,
with a predominant amount of α-​linolenic acid. Chia oil shows higher quan-
tities of α-​linolenic acid compared to flaxseed oil (Khattab and Zeitoun, 2013).
310 Compounds of Latin-American Crop Grains

TABLE 8.2 FATTY ACIDS (FA) CONTENT IN LATIN-​AMERICAN


GRAINS (G/​100 G FA)*
Grain/​
seed/​fruit 16:0 18:0 18:1 18:2 18:3 Reference
Maize 0.56 0.07 1.25 2.1 0.06 Sánchez-​
Ortega
(2014)
Amaranth 17.0–​21.35 2.75–4​ .11 20.2–​32.1 33.5–​43.9 NA Cuevas-​
Espinal et al.
(2017)
Quinoa 9.6–​10 0.84–0​ .94 25.8–​29.12 46.6–​49.5 7.34–​8.51 Vera (2014)
Chia 6.39–​7.72 2.36–3​ .74 6.59–​9.12 16.9–​22.5 56.9–​64.7 Coates
(2011)
Kaniwa 17.9 0.4 23.5 42.6 6 Repo-​
Carrasco-​
Valencia
(2011)
Tarwi 9.48 6.66 36.8 39.6 3.91 Ejigui et al.
(2005)
Turtle bean 8.37–​10 4.36–5​ .14 25–​32.6 41.2–​45.7 8.7–​10.4 Ejigui et al.
(2005)
Sacha Inchi 1.7–​2.1 1.1–​1.3 3.5–​4.7 12.4–​14.1 12.8–​16 Chirinos
(2013)
Notes: NA=​Not available. *Values expressed as g/​100 g dry basis.

The ω6/​ω3fatty acid ratio of Chilean chia oil (0.29) is similar to the chia seed
oil of Argentina (Ayerza, 2010) and markedly lower compared to most vegetable
oils, such as maize (76.57), rapeseed (2.26), soybean (6.68), sunflower (30.77),
and olive (17.86). The use of chia seed oil in the human daily diet is beneficial
since the use of vegetable oils with high content of polyunsaturated fatty acids
(PUFAs) have been shown to provide several health benefits (Villanueva et al.,
2017). Low ω6/​ω3 (normal ratio ranges are from 1:1 to 4:1) fatty acid ratio has
also been associated with the reduction of cardiovascular disease risk (Ganesan
et al., 2014). Costantini et al. (2014) substituted 10% of wheat flour with chia
flour to make nutritionally improved bread; authors found an increase in the
content of fat, dietary fiber, and total phenolic compounds, and a decrease in the
content of carbohydrates.
8.4 Amino Acids Profile 311

Tarwi seeds are rich in lipids, with a range from 12 to 24 g/​100 g (dry basis).
In decreasing order, fatty acids per 100 grams of seed are oleic, linoleic, palmitic,
and stearic (Elena, 2011; Suca, 2015).
Sacha inchi seeds show large amounts of unsaturated fatty acids, with high
concentrations of linolenic and linoleic acids. Ranges in different varieties are
between 39 and 47% and 34 to 41 g/​100 g (dry basis) respectively (Maurer et al.,
2012; Chirinos, 2013; Chirinos et al. 2016). The oil also contains other fatty acids
such as oleic, palmitic, and stearic to a lower extent. Despite its composition
which is rich in fatty acids susceptible to oxidation, sacha inchi oil presented
better oxidative stability than linseed and sesame oils, allowing it to preserve its
sensorial characteristics. This fact could be attributed to the rich presence of nat-
ural antioxidant substances such as tocopherols, carotenoids, and polyphenols
(Sánchez-​Sánchez, 2012).
Black turtle beans are also considered a source of unsaturated fatty acids. This
bean has a predominance of PUFAs (0.30 g/​100 g wb), followed by saturated fatty
acids (0.2 g/​100 g wb), and finally monounsaturated (0.07 g/​100 g wb). Among
the individual fatty acids are mainly found linoleic, alpha-​linolenic, and palmitic
acid (Wang et al., 2010).

8.4 AMINO ACIDS PROFILE


Proteins provide essential (eaa) and non-​essential (neaa) amino acids for
human metabolism, not only for the growth of infants and children but also
for the constant replacement and turnover of body proteins in adults. Of the
total amino acids that make up food proteins, eight are considered eaa for
adults and nine for children (histidine, isoleucine, leucine, lysine, methionine,
phenylalanine, threonine, tryptophan, and valine), because the human body
cannot synthesize them and they must be provided through diet. High-​quality
protein is one that contains all the eaa in the right proportions. A useful
method to evaluate the protein quality of foods is based on the proportional
content of eaa with the age group requirements (amino acid scoring system)
(FAO FINUT, 2017).
Maize protein is deficient in lysine and tryptophan (limiting amino acids)
with a score of approximately 0.6 and 0.84 respectively according to the FAO/​
WHO (FAO, 1985; WHO, 2007) standard requirement.
One of the most important characteristics of amaranth grain is that its
protein content is higher and better balanced in essential amino acids than
most cereals. Amaranth proteins are relatively rich in tryptophan, histidine,
valine, phenylalanine, lysine, and threonine and poor in methionine and leu-
cine with respect to values reported by the FAO/​WHO amino acid pattern
312 Compounds of Latin-American Crop Grains

requirements for 2–​5 year-​old children (FAO, 1985; WHO, 2007). Lysine con-
tent of A. hypochondriacus (5.95 g/​16 g N) is higher compared to 2.90 g/​16 g N in
wheat flour. A suitable content of both lysine and tryptophan with a low content
of leucine makes it a high-​quality supplement for maize, which is rich in leu-
cine but poor in lysine and tryptophan. The amino acid profile and the chemical
score confirm the suitability of amaranth to mix with other cereals and achieve
proteins of higher nutritional quality (Motta et al., 2019).
Quinoa provides all the eaa; and compared to similar commodities, it has
high protein content and a balanced eaa profile with the exception of trypto-
phan (Filho et al., 2017).
The amino acid profile of quinoa shows significant differences between
cultivars and geographical origins due to the different agroecological conditions
that are cultivated. Essential amino acids, i.e., methionine, leucine, lysine,
phenylalanine, tyrosine, isoleucine, threonine, tryptophan, and valine were pre-
sent in several cultivars of quinoa, but their distribution patterns were different
between them. Table 8.3 shows the amino acids content of different crops
(Valenzuela Zamudio and Segura Campos, 2020). Compared to the requirement
for children WHO (2007), the quinoa amino acid is adequately balanced, with a
chemical score of 0.82 for tryptophan.
The kañiwa protein has an excellent biological quality because of the balanced
eaa composition, especially because of its high lysine content. The lysine content
of kañiwa is twice that of wheat (Repo-​Carrasco-​Valencia et al., 2003). Kañiwa
proteins meet the FAO/​WHO/​UNU (UN University) protein reference pattern for
children (FAO, 1985; WHO, 2007).
Coates (2011) found that chia seeds from lower altitude ecosystems
showed 61% more protein content than those from higher cultivation areas.
Nevertheless, differences in the profile of individual eaa among the seeds were
not statistically significant (P <0.05). Percentages of amino acids in chia are
shown in Table 8.3.
The amino acid composition of tarwi seed proteins indicates that it is defi-
cient in sulfur amino acids and rich in lysine (Laurente-​Flores, 2016).
Like other legumes, black turtle bean proteins also contain greater amounts
of eaa including lysine that is deficient in cereal grains. Therefore, beans and
cereal proteins are nutritionally complementary in eaa and the combined con-
sumption of beans and cereals may alleviate these mutual deficiencies ensuring
a balanced diet (Audu et al., 2013).
The main amino acids present in sacha inchi are arginine, valine, threonine,
methionine, histidine, isoleusine, leucine, phenylalanine, and tryptophan and
low amounts of lysine (Gutiérrez et al., 2011; Ruiz et al., 2013). The amino acid
content for different grains is shown in Table 8.3.
newgenrtpdf
TABLE 8.3 AMINO ACIDS PROFILE OF LATIN-​AMERICAN GRAINS (MG/​G PROTEIN)
Histidine Isoleucine Leucine Lysine Threonine Tryptophan Valine Methionine References
Maize 30.7 37.6 125.2 34 38.4 5.9 53.9 ND Nuss and
Tanumihrdjo
2010
Amaranth Olpir 20 36 62 80 49 ND 48 20 Písaříková
2005
Koniz 17 38 69 80 45 ND 53 23
Elbrus 17 34 59 59 47 ND 47 20
Quinoa Amilda 34.8 29.5 73.7 64.1 44.3 7.5 41.1 18.6 González et al.
2012
Robura 24.4 23.2 58.2 50.9 35.4 9 31.1 15
Sayana 26 23.7 55.3 48.4 33.8 8.3 30.7 30.7
Chia 22.1 28.1 51.2 40.3 29.2 0.1 46.1 ND Ayerza 2013
Kaniwa 27 34 61 56 34 11 42 31 Repo-​
Carrasco-​
Valencia et al.
2009

8.4 Amino Acids Profile


Tarwi ND 48 79 59 36 7 45 4 Suca 2015
Turtle 32 41 77 65 29 12 47 ND Audu
bean et al.,2013
Sacha 19 30 52 31 40 12 52 13 Ruiz et al. 2013
Inchi
Note: ND =​Not determined.

313
314 Compounds of Latin-American Crop Grains

8.5 VITAMINS AND MINERALS


In maize, phosphorus and potassium are the most abundant minerals. It also
contains two fat-​soluble vitamins, vitamin E and provitamin A (carotenoids).
Most of the carotenoids are located in the hard endosperm of the grain and in
small quantities in the germ. Maize β-​carotene and cryptoxanthin content are
20% and 51% respectively of the total carotenoids present in the grain; vitamin
E is mainly located in the germ (Shah et al., 2016). It is important to mention
that among the different varieties of maize, several of the carotenes present in
each one of them and their proportions vary. The water-​soluble vitamins are
more abundant in the aleurone layer of the grain; part is lost when these grains
are subject to the wet cooking process (Siyuan et al., 2018). Table 8.4 shows the
vitamin content of these Latin-​American grains.
Amaranth’s high content of calcium, iron, and magnesium is very note-
worthy; the contribution of riboflavin and folates is higher compared to other
cereals of widespread consumption such as maize and rice. The niacin con-
tent is the lowest. The mineral contents are of great importance throughout
all biological stages of humans making this Mesoamerican and Andean grain
an important ingredient in the preparation of different foods like bread,
cookies, drinks, and infant formulas, among others (Suárez et al., 2013;
Hernández, 2018).
Regarding quinoa, it can be observed that both in minerals and in the content
of other nutrients, there is great variability between the mean values reported
by different authors; this variability is due to environmental and genetic factors
(Bravo et al., 2013; Miranda et al., 2012) (Table 8.4). It is worth mentioning
that amaranth and quinoa are rich in iron, copper, manganese, and zinc.
Kañiwa is a particularly good source of tocopherols (Repo-​Carrasco-​Valencia
et al., 2003). Tocopherols are compounds with high antioxidant capacity and
other important physiological functions; some have the function of vitamin
E. Tocopherols exist as four different isomers with antioxidant power, that is,
in decreasing order: δ> γ> β> α. Kañiwa has α-​tocopherols and γ-​tocopherols,
γ-​tocopherols being the main compounds. The tocopherol content in kañiwa
is superior to that in common cereals. Results from previous in vitro and in
vivo studies suggest that γ-​ and δ-​tocopherols could be additional potential
anticancer agents compared to α-​tocopherol (Das Gupta and Suh, 2016; Yang
and Suh, 2013). Due to its broad cancer-​preventive activity and availability, the
naturally occurring tocopherol mixture, rich in γ-​tocopherol or similar tocoph-
erol mixtures may have a significant role in practical application (Das Gupta
and Suh, 2016). In this respect, kañiwa is an outstanding source of these valu-
able phytochemicals.
newgenrtpdf
TABLE 8.4 VITAMINS AND MINERALS IN IN LATIN-​AMERICAN GRAINS (MG/​100 G)**
Nutrient Maize Amaranth Quinoa Chia Kaniwa Tarwi Turtle Bean Sacha Inchi
Pro Vitamin A 17.93–​266.2 -​ -​ 13.2* -​ -​ -​ -​
Vitamin C -​ -​ 16.4 -​ -​ -​ -​ -​
Vitamin E 0.07 -​ -​ 8.23 -​ 57.4 -​ 78.6–​137
Niacin 2.6 -​ -​ -​ 1.34 1.17 -​ -​
Thiamin (B1) 0.16 -​ 0.4 0.18 -​ 0.47 0.34 -​
Riboflavin (B2) 0.23 0.21 -​ 0.04 0.55 0.24 0.19 -​
Pyridoxine (B6) 0.47 0.22 -​ -​ -​ -​ -​ -​
Manganese -​ -​ -​ 2.3 -​ -​ -​ -​
Iron 2.3 5.3 4.4 16.4 15 1.08 6.22 10.3
Zinc -​ -​ -​ 3.7 -​ -​ 2.68 -​
Magnesium 147 344 -​ 390 -​ -​ -​ -​

8.5 Vitamins and Minerals


Calcium 158 303 47 714 110 72.8 179.1 240
References Nuss and Coelho Nascimento Ayerza Repo-​ Suca Ejigui Chirinos et al.
Tanumihardjo et al. et al. (2014) (2013) Carrasco-​ (2015) et al. (2005) (2013)
(2010) (2018) Valencia
(2011)
Notes: *µg; **Values expressed as g/​100g db.

315
316 Compounds of Latin-American Crop Grains

Regarding minerals, kañiwa has a relatively high iron, calcium, and zinc con-
tent. Compared to unenriched wheat flour (iron, 0.68 mg/​100 g; zinc, 0.98 mg/​
100 g; and calcium, 18.46 mg/​100 g db; concentrations of these minerals are con-
siderably higher in Andean grains (Dyner et al., 2007; Repo-​Carrasco-​Valencia
et al., 2010). Iron content in kañiwa is higher than in rice (1.32 mg/100 g) and
finger millet (2.13 mg/​100 g db) (Hemalatha et al., 2007). Pachon et al. (2009)
analyzed iron and zinc content in conventional and nutritionally enhanced
beans and maize. Andean grains contain more zinc and iron than conventional
maize and beans.
Chia seeds have been characterized as a good source of B vitamins such as
riboflavin, thiamine, vitamin E, and carotenes with provitamin A activity and
minerals like calcium, phosphorus, magnesium, potassium, iron, zinc, and
copper. It is also noted that chia’s sodium content is low. All of these minerals
and vitamins make chia a great addition to foods for all ages (Muñoz et al., 2013;
da Silva, 2017). They are used in different foods from cookies, cereal bars, and
more elaborate products (Zettel and Hitzmann, 2018; Coelho et al., 2014).
Tarwi has been designated by the FAO (2016) as an underutilized food during
the Year of Legumes since tarwi is native to the Andes and its cultivation is
maintained in different production systems, from Ecuador to Chile and north-
eastern Argentina. After the Spanish conquest its production declined, despite
providing not only proteins, but also vitamins and minerals. The main min-
eral contents of tarwi are phosphorus, calcium, and iron. Vitamins present in
tarwi are C, niacin, thiamine, and riboflavin in decreasing order according to
their amount per 100 grams of this food (Elena et al., 2011; Fernández-​Mejía and
Guivar-​Delgado, 2020).
As with other Latin-​American crops, information on the vitamin and mineral
composition of sacha inchi is very scarce. For this reason, the data is very focused
on its macronutrients. Some authors refer to Vitamin E as the most important
vitamin in this crop, in the form of tocopherols and tocotrienols (de Souza et al.,
2013). This, along with its richness in essential fatty acids, makes sacha inchi a
great potential for food preparation and inclusion in the usual diet (Clavijo et al.,
2015; Paucar-​Menacho et al., 2015).
In all the crops mentioned, an important contribution of both vitamins and
minerals is observed. This shows the relevance of highlighting their production
and promoting consumption by all population groups across the world.

8.6 BIOACTIVE COMPOUNDS
Current nutritional approaches are starting to reflect a fundamental change in
the understanding of the diet–​health relationship. Increasing knowledge about
8.6 Bioactive Compounds 317

the impact of diet on regulation at a genetic and molecular level is changing the
way the role of nutrition is considered, resulting in new dietary strategies.
Nowadays it is considered that diet must not only provide adequate
nutrients to meet metabolic requirements, but consequently can also con-
tribute to improving human health through these and other components. These
compounds with health-​promoting properties are called bioactives. They are
essential and nonessential compounds (e.g., vitamins or polyphenols) that occur
in nature and are part of the food chain, and can be shown to have beneficial
effects on human health. Consequently, plant extracts or individual bioactive
compounds must be identified and developed for the food market to supplement
a balanced diet.
Polyphenols are widely distributed as secondary metabolites from plants.
Many researchers have found health properties in phenolic compounds,
especially in flavonoids. Several studies report a relationship between higher
intakes of flavonoids and lower prevalence of some types of cancer and car-
diovascular diseases (Zhang and Tsao, 2016; Belščak-​Cvitanović et al., 2018;
Bever et al. 2021). Accumulated evidence supports the notion that diets
rich in specific flavonoids may play a role not only in promoting health but
also in preventing and mitigating the consequences of a range of chronic
degenerative diseases. A popular term used to describe the action of phen-
olic compounds is antioxidant. This term refers to polyphenols used on the
assumption that antioxidant activity in vitro translates to a similar function
in vivo (Rossi et al., 2018).
Many authors question if the health benefits from phenolic compound come
from their antioxidant activity (AA), since it is not proved that in vitro AA has the
same impact as in vivo (Harnly, 2015). This does not mean that these compounds
are not health-​promoting, but that they have antioxident action and are also
involved in broader physiological mechanisms.
Among all the varieties of native maize, those that contain the greatest
amounts of bioactive compounds are those with colors, such as blue, red, or
purple. Anthocyanins are the main water-​soluble pigment in colored maize. In
grain, the presence of anthocyanins has been reported mainly in the pericarp
(Guillén Sánchez et al., 2014). Although there are several scientific researches
on the biological properties of purple maize anthocyanins and their antioxi-
dant power in vitro, there is a lack of in vivo evidence (Smorowska et al., 2021;
Mazewski et al., 2017). Therefore, a comprehensive knowledge of the bioavail-
ability and metabolism of anthocyanins is essential to understanding their
health effects.
Repo-​Carrasco-​Valencia et al. (2019) studied the content of total phenolic
compounds and the antioxidant capacity of kañiwa, quinoa, and kiwicha.
The kañiwa sample had the highest value compared with the other grains
318 Compounds of Latin-American Crop Grains

(57.87 mg gallic acid/​100 g of sample). This same tendency was found in


the values of the antioxidant capacity of the studied grains. Flavonoids are
phenolic compounds with many health promoting properties, for example,
antioxidant activity, anti-​inflammatory, and anticarcinogenic properties.
In kañiwa the principal flavonoids are quercetin and isorhamnetin (Repo-​
Carrasco-​Valencia, 2011). Individual phenolic compounds present in amar-
anth and quinoa are 4-​dihydroxybensoic acid, ferulic acid, rutin, ellagic
acid, kaemferol-​3-​rutinoside and quercetin (Stikić et al., 2020). Berries have
been considered an excellent source of flavonoids, especially quercetin and
myricetin. For example, lingonberry contains 10 mg/​100 g wb of quercetin
and cranberry contains 10.4 and 6.9 mg/​100 g wb quercetin and myricetin,
respectively (Mattila et al., 2000). The levels in these flavonoid-​rich berries
are 5–​10 times lower than those found in Chenopodium seed samples.
When compared on a dry weight basis, the flavonoid content in berries and
Chenopodium samples are of the same magnitude. Abderrahim et al. (2012)
studied the effect of germination on total phenolic compounds, antioxidant
capacity, Maillard reaction products, and oxidative stress markers in kañiwa.
They observed an increase in the content of phenolic compounds and anti-
oxidant capacity after 72 h of germination.
Kañiwa is exceptionally rich in resorcinol compounds, not very common
in plants (Peñarrieta et al., 2008). Of the major cereals, resorcinols have been
reported to be present in high levels in wheat, rye, and triticale and in low
amounts in barley, millet, and maize. Cereal alkylresorcinols (ARs) have been
reported to have anticancer and antimicrobial effects, as well as the ability to
inhibit some metabolic enzymes in vitro. ARs have also been reported to have
antioxidant activity (Ross et al., 2003).
Phytosterols are natural components of plant cell membranes and
are abundant in vegetable oils, seeds, and grains. Phytosterols have
various biological effects, such as anti-​inflammatory, antioxidative, and
anticarcinogenic activity, as well as cholesterol-​lowering capacity (Abugoch
James, 2009). Repo-​Carrasco-​Valencia et al. (2019) studied the phytosterol
content of Andean grain oils and found that total phytosterol content in
kañiwa was 74.74 mg/​100 g dry weight. Besides campesterol, stigmasterol,
and β-​sitosterol, which are the most common phytosterols, ergost-​7-​en-​3-​ol,
gramisterol, ∆7-​sitosterol, ∆7-​avenasterol, 24-​m ethylenecycloartanol, and
citrostadienol were identified.
Regarding chia, there are higher values of total polyphenols in its black seeds
than in white seeds 295 and 185 mg/​GAE (gallic acid equivalent), respectively. Of
these polyphenols, the majority are flavonoids, most of which are anthocyanidins
and proanthocyanidins (Muller Tito, 2015).
8.7 Anti-Nutritional Compounds In Latin-American Grains of Interest 319

The total content of phenol in the 16 varieties of the sacha inchi almonds
studied is between 64.6-​80.0 mg GAE/​100 g; this is higher compared to other ole-
aginous fruits such as almonds and hazelnuts (Chirinos et al., 2013).
Pérez-​Hernández et al. (2020) reported that in P. vulgaris seeds, there was
a wide array of polyphenols including hydroxycinnamic acids, flavonoids,
condensed tannins, and anthocyanins. The major anthocyanins in black beans
are delphinidin-​3-​O-​β-​D-​glucoside, petunidin-​O-​β-​D-​glucoside, and malvidin-​
3-​O-​β-​D-​glucoside. Catechins were the predominant forms of anthocyanidins
(Hall et al., 2017).
In natural tarwi grains, the total polyphenols were found in ranges between
518.9 to 533.9 mg GAE/​100 g (wb). Within the polyphenol groups, tarwi contains
flavonoids and phenolic acids (Sabelino Francia, 2020).
Amaranth and quinoa have a series of antinutrients such as saponins,
oxalic acid, tannins, trypsin inhibitors, and phytic acid. Phytic acid is found in
large quantities in amaranth and has the property of forming complexes with
metal cations such as Mg+​2, Fe+​2, Fe+​3, Ca+​2, Zn+​2, Co+​2, and Cu+​2 causing a loss of
intestinal bioavailability of these components. However, some authors have
also described beneficial functions of phytic acid, such as preventing chronic
and coronary disease, among others (Ranilla, 2019). Saponins are characterized
by the formation of foam, having detergent and surfactant properties; they
largely determine the bitter taste in some seeds such as quinoa and sacha
inchi. Consumption of these products without de-​ saponification or when
consumed raw can cause disorganization in the cell membranes, and in excess
its symptoms begin in the gastrointestinal tract with nausea, vomiting, and diar-
rhea. A method of eliminating saponins from quinoa is scarification (Mackay
et al., 1995) followed by five washes; this leaves a concentration of less than 0.1%
and eliminates its bitter taste (Calliope et al., 2015).

8.7 ANTI-​NUTRITIONAL COMPOUNDS IN
LATIN-​AMERICAN GRAINS OF INTEREST
Oboh et al. (2010) found that the roasting process produced a significant
decrease in phytic acid, in yellow maize (26.1%) and white maize (36.81%); it
also caused a significant decrease of flavonoids in both types of maize. Grundy
et al. (2020) demonstrated that fermentation, in addition to improving the
nutritional value and digestibility of amaranth grain proteins, degrades
phytates and oxalates. Cooking amaranth in excess of water, even without
bursting the grain, as well as other thermal processes (sterilization by auto-
claving and blanching) and germination reduce the content of phytic acid,
oxalate, and tannins.
320 Compounds of Latin-American Crop Grains

Legumes contain anti-​nutritional factors such as trypsin inhibitors, phytic


acid, tannins, and oligosaccharides that limit the use of proteins and minerals.
Heat treatment significantly improves protein quality in legumes by destroying
or inactivating heat-​labile anti-​nutritional factors. Cooking produces a signifi-
cant reduction in the activity of the trypsin inhibitor, from 9.9 to 1.1 mg/​g of dry
matter; it also reduces tannins from 13 to 3.2 g/​kg of dry matter but has little
effect on phytic acid (Wang et al., 2010).
He et al. (2015) highlighted that excessive use of P. vulgaris beans could cause
toxic and allergenic reactions. Besides the beneficial roles of lectins for the
plant, they can from an agronomic viewpoint bind to the surface of epithelial
cells in the digestive system due to their high affinity with carbohydrates and can
result in toxic reactions with changes in intestinal permeability (Menard et al.
2010). Like other anti-​nutritional factors, lectins are heat-​stable proteins; they
can be inactivated after special heating, such as cooking (around 100ºC), with
an improvement in the accessibility of proteins to enzymatic attack. Thermal
inactivation of lectin from black turtle beans can be achieved by heating at 90ºC
for 5 min, demonstrating the importance of high temperature and short time
(HTST) treatment in the packaged product. In an autoclave, it can be completely
inactivated at 121ºC for 5 min.
Valles Ramírez (2012) used a chemical peeling by immersion in NaOH for
3 min to separate the tegument and precooking at 100°C for 20 min to inacti-
vate anti-​nutritional factors from the sacha inchi seed. To inactivate the fixed
coumarins and alkaloids in sacha inchi, a pre-​toast at temperatures between 103
and 110°C was also used.
Curti et al. (2018) reported that grains of Lupinus species contain between 1
and 4 g/​100 g dw of toxic alkaloids that must be eliminated before consumption.
The traditional debittering process applied in L. mutabilis includes a soaking
stage of seeds for 18 h, followed by a cooking step (0.5–​6 h). The cooking stage
is essential to inactivate the germination capacity of the seeds, their enzymes
(lipase, lipoxygenase), to eliminate occurring microorganisms for food safety,
to reduce the loss of proteins through their coagulation, and to facilitate
the leaching of the alkaloids by increasing cell wall permeability. The process
is completed with a wash for several days in which most of the alkaloids are
removed. The debittering process is a necessary step required by Lupinus species
to ensure safe human consumption.
Once the anti-​nutritional compounds are eliminated, these grains are good
ingredients to be included in foods since they provide beneficial bio-​functional
compounds and dietary fiber for the prevention of non-​transmissible chronic
diseases.
8.8 Effects of Different Technological and Culinary Procedures 321

8.8 EFFECTS OF DIFFERENT TECHNOLOGICAL AND


CULINARY PROCEDURES ON NUTRITIONAL
COMPOSITION
Food composition tables (FCT) generally include the nutritional composition of
foods in their natural state, even when consumed after industrial or culinary
processes, heating, or storage. These treatments can affect the nutritional
quality and content of nutrients and bioactive compounds of raw foods because
they can cause degradation or leaching of their components after processing or
treatments. Therefore, these changes must be analyzed to know the true amount
of nutrients. Few studies have been carried out on this subject. This is done by
calculating the nutrient retention factors (NRF%) that correspond to the per-
centage of preservation of nutrients, especially vitamins and minerals. To calcu-
late NRF%, it is necessary to know the yield factor as well the nutrient content of
raw and cooked or processed foods.
It is important to know how the different proteins (and their breakdown during
digestion) are affected by treatments including baking, extrusion, cooking, etc. It
is well-​known that the physicochemical and biological characteristics, including
antigenicity, of proteins, are changed after heating and also that the extent of
their degradation during digestion can be either increased or lowered after heat
processing (Davis, 1998). All of these modifications can affect the characteristics
of the digestion products of proteins and, consequently, their effects on human
physiology (Passini et al., 2001).
Yield factor (YF) is the term used to indicate the change in the food or dish
weight after preparation or processing.

weight of cooked or processed food


YF = Ec.8.1
weight of raw food
nutrient content per g of cooked food
NRF % = x YF x100 Ec.8.2
nutrient content per g of raw food

When the NRFs are known for each processed food under standardized
conditions, it is possible to apply them to other similar products processed
under identical conditions. Some food composition tables (FCT) include NRF%
for the most commonly consumed processed foods, but there is a lack of infor-
mation on the grains discussed in this chapter.
322 Compounds of Latin-American Crop Grains

8.9 CHANGES IN NUTRITIONAL VALUE AND


MODIFICATIONS OF BIOACTIVE COMPOUNDS
CONTENT THROUGH PROCESSING
Maize is prepared and consumed in various ways; it can be dried in the sun,
cooked, fermented, roasted, mashed, or crushed to use in different preparations,
depending on the region or ethnic group. Within these processes, toasting is a
traditional practice. It is used to increase nutritional properties, improve flavor,
and reduce anti-​nutritional factors. Oboh et al. (2010) studied this process in two
races of yellow and white maize. After roasting, they found a decrease in protein
content caused by denaturation and the participation of some amino acids in
the Maillard reaction; there was an 18% increase in fat content in yellow maize
and 20% in white maize after roasting; there were no significant differences in
ash content between raw and roasted maize. Nutrient content of both yellow
and white maize is shown in Table 8.5.
Modifications in the content of nutrients in the final product are variable due
to the fact that two factors cause them: (i) modification of the total weight of the
final product; (ii) the loss/​gain of nutrients (leaching, transformation, reaction
with other components). In this case the first factor has not been considered.
Nixtamalization is nowadays international technology to process corn and
produce dough for the preparation of tortillas and other foods. It was used by
the Mayans and Aztecs and continues to be used today with some modifications
(Cravioto et al., 1945). Briefly, maize undergoes aqueous cooking with lime, is
left to rest for about 12 h and then washed to produce the nixtamal; finally it
is ground to form dough (Bressani, 1990). The process produces chemical and
nutritional changes; the nixtamal produced after alkaline cooking depends on

TABLE 8.5 PROXIMAL COMPOSITION OF MAIZE ACCORDING TO


TREATMENT (G/​100 G DRYBASE)
Composition

Maize Treatment Protein Crude fat Ash Moisture


Yellow Raw 8.45a ± 1.5 6.21b ± 1.6 1.98d ± 1.2 13.67a ± 1.5
Roasted 7.85 ± 1.7
b 7.34 ± 2.0
c 1.85 ± 0.3
e 10.93b ± 1.0
White Raw 12.97c ± 2.1 5.32d ± 1.2 1.93d ± 0.7 16.92c ± 2.0
Roasted 10.86 ± 1.6
d 6.39 ± 0.9
b 2.00 ± 0.1
d 14.31f ± 1.2
Source: Oboh et al. (2010).
Notes: Data are presented as mean±standard deviation; n =​3. Values with the same
letter per column are not significantly different.
8.9 Changes in Nutritional Value and Modifications 323

the processing parameters used during cooking and steeping, as well as the
physicochemical properties of the maize used. Nixtamalization alters the struc-
ture and solubility of maize proteins; this together with the cooking process to
make tortillas reduces the solubility of albumins, globulins, and prolamines, and
induces the appearance of high molecular weight protein fractions as glutelins;
digestibility of the protein decreases slightly in both the nixtamal and the tortilla
but improves the overall protein quality. Paredes López et al. (2009) reported
that when yellow maize underwent nixtamalization, the lipids decreased, up to
3.4% in the yellow maize tortilla and 2.5% in the white maize. These losses are
due to the elimination of the pericarp, aleurone, and the germ –​which can be
partial or total –​where most of the lipids are located. The losses in vitamins that
alkaline cooking and tortilla production caused are variable; in nixtamalized
yellow maize losses of 15–​28% of carotenes and thiamine and riboflavin were
reduced by up to 60–​70%. However, during this process, the soluble dietary fiber
goes from 0.9% in the grain to 1.3% in the dough and 1.7% in the tortilla. The
alkaline cooking process improves the bioavailability of niacin because this
vitamin is present in the maize kernel attached to other components and there-
fore not then available to be released as nicotinic acid, an active component of
niacin. During alkaline cooking, calcium content in the dough increases. The
phosphorus in maize is mainly found as part of phytic acid, a chemical com-
pound that strongly interferes with the absorption of various bivalent elements,
including calcium, and whose content decreases from 1% in the maize grain to
0.4% in the tortilla, which leads to the relationship between calcium and phos-
phorus from 1:20 in maize grains become 1:1 in the tortilla. López-​Martínez
et al. (2011) found that the nixtamalization process reduced phenolic and antho-
cyanin compounds, and total antioxidant activities of white, blue, red, and purple
maize. Also significantly reduced were aflatoxin (up to 90%) and fumonisins
(up to 80%). Overall, there was no significant difference among the sensory
properties prepared from nixtamal and non-​nixtamal maize flour. The study
findings suggest that nixtamalization is a promising and affordable processing
technology for reducing mycotoxin levels in maize and enhances the nutrient
profile of maize products, simultaneously increasing consumer acceptability
(Bressani, 1990; Paredes López et al., 2009; Maureen et al., 2020). Additionally,
the nixtamalization process significantly increases calcium availability, which
represents an outstanding nutritional fact. Many alternative nixtamalization
technologies have emerged, mostly based on the use of extrusion, with scien-
tific evidence in improving the nutrimental/​nutraceutical properties, reduction
of contaminant effluents, processing time and water use, although none of them
has been widely adopted by the tortilla industry; the reaction of consumers to
the change is mainly explained by the sensorial properties provided by trad-
itional nixtamalization processes (Escalante-​Aburto et al., 2020).
324 Compounds of Latin-American Crop Grains

Salvador-​Reyes et al. (2020) highlight chicha Morada and mazamorra Morada


as highly consumed products in Peru. In this regard, Bhornchai et al. (2014)
reported that the boiling process to obtain these products leads to a 60% reduc-
tion in the content of anthocyanins and other phenolic compounds in purple
waxy maize. Jing and Giusti (2007) maintained that the loss of anthocyanins in
the cooking water of purple maize during the process of obtaining sweets and
beverages was due to the denaturation of proteins at 100°C, associated with their
complexation with anthocyanins and tannins. De la Parra et al. (2007) associated
the decrease in bioactive compounds to the indicated treatment (total phenols,
anthocyanins, ferulic acid, carotenoids) with making pasta, tortillas, and chips
using white, yellow, blue, and red maize.
According to Ramos Diaz et al. (2013) the first step toward the industrial-
ization of Andean grains was the adoption of low-​cost technological processes,
such as extrusion cooking, which is characterized by its ability to increase the
digestibility of starch and proteins. The extrusion of starchy products causes gel-
atinization of starch, denaturation of proteins, and the formation of starch-​lipids
and protein-​lipids complexes. These authors studied the effects of extrusion on
lipid stability in extruded products made with mixtures of maize with quinoa,
amaranth, and kañiwa flours with different humidity conditions. Depending on
the type of interaction between amylose and fatty acids in solid matrices, oxi-
dation of lipids can lead to the formation of volatile compounds during storage.
A tendency to lipid oxidation was observed in extrudates containing quinoa
and kañiwa exposed to 11% humidity. However, the lipid stability of extrudates
containing amaranth or just pure maize did not appear to be affected by exposure
to different moisture conditions and physical modification. It was also observed
that extrudates with higher fiber content (lower hardness) were more prone
to oxidation. This could be because plasticizers (and possibly fiber) could have
limited chain-​to-​chain interaction, thus reducing the ability of macromolecules
(amylose and amylopectin) to form a protective layer around lipids (entrap-
ment). The effect of this process on micronutrients and bioactive compounds
remains a subject of scientific research.
Phytate content of kañiwa was studied by Repo-​Carrasco-​Valencia et al. (2009)
and was found to be 8.0 mg/​g, which is higher than that in amaranth, 3.4–​6.1
mg/​g (Guzman-​Maldonado and Paredes-​Lopez, 1998), but lower than common
cereals. Gualberto et al. (1997) found 14.2, 43.2, and 52.7 mg/​g of phytate in oat,
rice and wheat bran, respectively. Phytic acid has long been considered as an
anti-​nutrient because it chelates minerals and trace elements. However, its anti-
oxidant potential is now recognized (Fardet et al., 2008). Repo-​Carrasco-​Valencia
et al. (2010) studied the effect of common processing methods, such as roasting
and boiling, on the content of three micronutrients (Fe, Ca, Zn) in Andean grains.
There was a significant decrease in iron content during the boiling process of
8.9 Changes in Nutritional Value and Modifications 325

kañiwa. Wet processing procedures in general cause loss of dry matter and iron.
Boiling reduced the content of zinc in quinoa and kañiwa, but not in amaranth.
Roasting negatively affected the content of calcium in quinoa but not in kañiwa
and amaranth. Boiling enhanced the iron, zinc, and calcium dialyzability in
kañiwa. To increase the potential contribution of minerals from Andean crops,
and prevent losses, it is important to study the effect on their nutritional com-
position of the different processes to which they are subjected.
Zare et al. (2019) indicated that soaking chia seeds increases the extraction
capacity of ω3 and ω6 fatty acids compared to seeds without soaking.
Currently, interest in sacha inchi is growing because its oil is a new source of
polyunsaturated fatty acids with a favorable ratio of ω6/​ω3, and also because it
contains other functional ingredients (Wang et al., 2018). The residual cake of
sacha inchi seeds, after the oil extraction process by cold pressing, preserves a
high percentage of proteins (56.6 g/​100 g cake wb), oil (4.1 g/​100 g), and mois-
ture (2.6 g/​100 g sample wb) so it would be a good source for protein extrac-
tion as a way to add value to the crop. The pressed cakes contain 4.2% sucrose,
2.2% starch, and 11.1% dietary fiber. The lipid composition shows saturated and
unsaturated fatty acids, and minerals such as potassium, which was the highest
element followed by phosphorus, magnesium, calcium, and, to a lesser extent,
zinc, sodium, manganese, and copper; it also contains 0.51 mg GAE/​g. All these
components suggest that sacha inchi seed press cake could be a useful ingredient
for the formulation of food for human consumption (Rawdkuen et al., 2016).
Curti et al. (2018) studied the effect of the debittering process on the fatty acid
composition of two species of bitter lupine (L. albus and L. mutabilis). They found
that after applying the process, there were higher changes in the fatty acid com-
position of L. mutabilis than L. albus; the content of linoleic, palmitic, and ste-
aric acid increased in L. mutabilis, while in L. albus PUFA/​SFA, behenic, linolenic,
and arachidic acids increased. Changes in the fatty acid composition between
the processed Lupinus species were explained by the differences in their initial
composition. Carvajal-​Larenas et al. (2016) found that the debittering process
can induce decreasing the content of soluble carbohydrates, fiber, and fat and
increasing the protein content in L. mutabilis.
Tannins form complexes with proteins, which leads to a decrease in their
digestibility. Wang et al. (2010) reported that heat treatment significantly
improves legume protein quality by destroying or inactivating heat-​labile anti-​
nutritional factors such as phytic acid and tannins. The boiling method affects
the chemical composition of legumes. Wang et al. (2010) evaluated the effect of
boiling on the nutritional and anti-​nutritional components of black turtle beans.
Proteins and starch apparently increased since the loss of soluble solids during
cooking and the decrease in ash due to leaching of some minerals in the cooking
water modify the total weight of the cooked product. The calcium, copper, iron,
326 Compounds of Latin-American Crop Grains

and zinc in turtle bean were not significantly modified while potassium and
magnesium decreased significantly, from 1415–​1070 mg/​100 g (db) and 200–​156
mg/​100 g (db) respectively, and manganese and phosphorus increased from 1.5–​
1.7 mg/​100 g (db) and 370–​443 mg/​100 g (db) respectively (Wang et al., 2010).
All of the above examples indicate that it is of great importance to choose
suitable processing/​ cooking methods to achieve maximum preservation
of nutrients and bioactive compounds and also to maintain the natural/
typical flavor of all these Latin-​ American crops after processing (Guillén-​
Sánchez et al., 2014).
On the other hand, Motta et al. (2016b) provided information on the effects
of cooking (boiling and steaming) on NRF% in quinoa and amaranth. These
processes produced a significant increase in grain moisture, probably because
heating affected the water absorption capacity of the starch. These processes had
effects on mineral content and NRF%. In Table 8.6 mineral content in raw and
processed grains and NRF% are shown. These results indicate that, in amaranth,
both types of culinary method produce similar losses, while in quinoa the steam
process produces lower losses for all minerals. Different effects on minerals can
be explained by their cellular location in the seed matrix.
The term folic acid or folates is applied to a family of vitamers with equivalent
biological activity. It is a water-​soluble vitamin whose function resides mainly in
its ability to donate and capture carbon units. Folic acid is an essential nutrient
for cell life; its deficiency leads to the development of numerous pathologies.
Several studies reported the negative effects of cooking processes on folate con-
tent in foods, especially legumes and vegetables (Delchier et al., 2013, Stea et al.,
2007). Quinoa and amaranth are good sources of folates.
Motta et al. (2017) studied the folate content and the effect of steaming,
boiling, and malting methods applied to quinoa and amaranth grains. The boiled
amaranth samples showed a significant decrease in total folate content, while in
the malted it increased significantly with respect to the raw grains (Figure 8.2).
This increase in amaranth after the malting process was also previously reported
in legumes and wheat by other authors (Hefni and Witthöft, 2011; Koehler
et al., 2007; Shohag et al., 2012). This is possibly because during the germination
phase and before the grain drying process, folate production increases (Nelson
et al., 2013).
The boiled and steamed quinoa did not show significant differences on folate
content and were higher than the raw grain and lower than boiled samples
(Figure 8.2). The increase in total folate content in quinoa after cooking was
reported by other authors (Stea et al., 2007). As folates occur intracellularly,
they are released from the matrix when the cell/​seed structure is destroyed.
According to Motta et al. (2017), amaranth seeds are smaller and have a higher
surface area than quinoa. This may explain how boiling and steaming increases
the extracted folates from amaranth more than from quinoa.
8.9 Changes in Nutritional Value and Modifications 327

TABLE 8.6 MOISTURE AND MINERAL CONTENT IN RAW AND PROCESSED


AMARANTH AND QUINOA AND NUTRIENT RETENTION FACTORS (NRF)
Amaranth**

Cooking % NRF % NRF


method Raw Boiled Boiled Steamed Steamed
Moisture (g/​ 12.2a±0.09 73.6b±0.84 NA 52.4c±0.13 NA
100 g wb)
Cu (mg/​100 0.572a±0.02 0.652b±0.02 98.29a 0.601a±0.02 91.09a ±1.03
g-​db) ±4.67
Mn (mg/​100 4.42a±0.34 4.56a±0.06 89.30a 4.38a±0.09 86.21a ±7.01
g-​db) ±7.57
Fe (mg/​100 7.35a±0.46 7.42a±0.07 87.32a 7.25a±0.28 85.70a ±6.13
g-​db) ±5.86
Zn (mg/​100 4.55a±0.17 4.70a±0.12 89.05a 4.68a±0.15 89.31a ±6.16
g-​db) ±4.05
Mg (mg/​100 328a±9.2 307b±4.7 80.72a 302b±8.7 79.96a ±4.60
g-​db) ±2.63
Ca (mg/​100 200a±7.2 207a±1.5 89.06a 205a±2.1 88.98a ±3.94
g-​db) ±3.87
P (mg/​100 663a±13 655a±8.3 85.23a 647a±9.1 84.69a ±2.77
g-​db) ±2.14
K (mg/​100 552a±10 538b±2.8 84.05a 535b±5.1 84.07a ±2.37
g-​db) ±1.83
Na (mg/​100 <LoQ <LoQ NA <LoQ NA
g-​db)
Quinoa**
Cooking Raw Boiled %NRF Steamed %NRF
method Boiled Steamed
Moisture (g/​ 11.7a±0.18 66.6b±0.18 NA 62.8c±0.02 NA
100 g wb)
Cu (mg/​100 0.502a±0.005 0.461b±0.02 81.65a 0.465b±0.02 89.7b ±3.77
g-​db) ±2.71
Mn (mg/​100 1.89a±0.030 1.91a±0.16 96a 1.99a±0.24 109.93a
g-​db) ±7.99 ±10.97
Fe (mg/​100 4.29a±0.07 4.44a±0.29 92.98a 4.34a±0.12 98.46a ±3.24
g-​db) ±4.03
(continued)
328 Compounds of Latin-American Crop Grains

TABLE 8.6 (CONTINUED)

Amaranth**

Cooking % NRF % NRF


method Raw Boiled Boiled Steamed Steamed
Zn (mg/​100 2.97a±0.05 2.93a±0.24 84.42a 2.88a±0.22 90.37a ±6.27
g-​db) ±7.14
Mg (mg/​100 196a±3.1 192a±7.9 89.76a 189a±8.7 96.20a ±3.51
g-​db) ±4.83
Ca (mg/​100 77.6 a±2.1 74.2a±10 77.58a 71.4a±7.4 85.06a ±9.09
g-​db) ±10.37
P (mg/​100 436a±4.7 447ab±10 92.68a 451b±10 100.04b ±2.88
g-​db) ±1.76
K (mg/​100 559b±7.5 536a±7.8 87.18a 520c±9.7 90.70a ±2.94
g-​db) ±1.62
Na (mg/​100 <LoQ <LoQ NA <LoQ NA
g-​db)
Source: Motta et al. (2016a).
Notes: LoQ (limit of quantification) =​2.5 mg/​100 g; * dry base; **mean ± standard
deviation (n =​3). Different letters in each file indicate significant differences.
NA: Not applicable.

Table 8.7 shows the NRF% of folates in quinoa and amaranth subjected to the
three mentioned processes. It can be seen that these were dependent on both
the food matrix and the method of processing (boiling, steaming, and malting).
Retention values provided information on nutrient gains and losses, facilitating
the selection of the most suitable cooking/​processing method.
When incorporating processed grains or products derived from them into a
recipe, the nutrient content must be adjusted by NRF% to create the nutritional
profile of the final product.

8.10 CONCLUSIONS
The information compiled in this chapter, regarding the content of nutrients
and bioactive compounds of the following grains of Latin-​American origin, corn,
quinoa, amaranth, chia, and beans such as black turtle bean, tarwi, and some
8.10 Conclusions 329

Figure 8.2 Effect of cooking methods and malting on total folate content (μg/​100 g dry
weight basis) in quinoa and amaranth. Different letters in the same grain indicate signifi-
cant differences. Source: Motta et al. (2017).
Notes: Bar graph depicting effect of cooking methods and malting on total folate con-
tent (μg/​100 g dry weight basis) in quinoa and amaranth. Gray bars indicate raw, boiled,
steamed, and malted treatments.

TABLE 8.7 FOLATE NUTRIENT RETENTION FACTORS


FOR PROCESSED QUINOA AND AMARANTH
NRF (%)

Process

Grain Boiled Steamed Malted


Amaranth 41 78 121
Quinoa 118 110 92
Source: Motta et al. (2017).

lesser known examples, such as kañiwa and sacha inchi, allow us to affirm that
they are excellent foods to be included in the diet. Some, such as quinoa, amar-
anth, and kañiwa have been shown to be mainly a source of good biological
quality protein; others such as chia, sacha inchi, and turtle bean provide essen-
tial fatty acids; all contain different minerals and vitamins required for a healthy
diet. Corns, mainly colored ones, contain high amounts of bioactive compounds
330 Compounds of Latin-American Crop Grains

with antioxidant activity, which can contribute to the prevention of chronic non-​
communicable diseases and can also be used as natural colorants.
Some of these grains contain anti-​ nutritional factors, which by simple
methods and also by culinary processes can be eliminated; this is necessary and
useful information for the consumer and for those who have never previously
consumed them.
The by-​products of all these crops –​depending on the process they are
subjected to –​can potentially be raw material rich in nutritional and functional
components to be incorporated into formulated foods. An example of this is the
residual cakes from the extraction of chia and sacha inchi, sources of essential
amino acids, unsaturated fatty acids, dietary fiber, and minerals.
The technological processes applied for the transformation of these grains
and also the culinary examples produce changes in the components, nutrients,
and non-​nutrients, of the food which can cause losses in different proportions; to
safeguard this, it is necessary that FCTs include information on the composition
corrected by NRFs. When incorporating processed grains or products derived
from them into a recipe, the nutrient content for calculating intakes must be
adjusted to create the nutritional profile of the final product.

ACKNOWLEDGMENTS
This work was supported by grant laValSe-​ Food-​
CYTED (Ref. 119RT0567);
Consejo Nacional de Investigaciones Científicas y Técnicas (CONICET) and
Universidad Nacional de Jujuy, Argentina.

REFERENCES
Abderrahim, F., E. Huanatico, R. Repo-​ Carrasco-​Valencia, S. M.Arribas, M. C.
Gonzalez, and L. Condezo-​Hoyos. 2012. Effect of germination on total phenolic
compounds, total antioxidant capacity, Maillard reaction products and oxi-
dative stress markers in canihua (Chenopodium pallidicaule). Journal of Cereal
Science 56(2): 410–​417.
Abugoch-​James, L. E. 2009. Quinoa (Chenopodium quinoa Willd.): Composition, chem-
istry, nutritional, and functional properties. Advances in Food and Nutrition
Research 58: 1–​31.
Alcocer-​Villacís, I. R. 2018. Determinación del contenido de azúcares reductores, totales,
almidón y de los elementos minerales calcio (Ca), hierro (Fe) y magnesio (Mg) en
harinas de Amaranto variedad Alegría (Amaranthus Caudatus L.), Chía (Salvia
Hispanica L.), papa Puca-​ Shungo (Solanum Tuberosum), y zanahoria blanca
(ArracachaXxanthorrhiza Bancroft) Bachelor’s Thesis, Universidad Técnica de
Ambato. Facultad de Ciencia e Ingeniería en Alimentos. Carrera de Ingeniería en
Alimentos.
References 331

Arévalo, J. B., C. M. Zegarra, B. C. Amado, A. M. R. Del-​Castillo, and G. Vargas-​Arana 2019.


Composición nutricional y capacidad antioxidante de tres especies de sacha inchi
plukenetia spp. de la Amazonía Peruana. Folia Amazónica 28(1): 65–​74.
Audu, S. S., M. O. Aremu., and L. Lajide 2013. Effects of processing on physicochemical and
antinutritional properties of black turtle bean (Phaseolus vulgaris L.) seeds flour.
Oriental Journal of Chemistry 29(3): 979–​989.
Ayerza, R. 2010. Effects of seed color and growing locations on fatty acid content and
composition of two chia (Salvia hispanica L.) genotypes. Journal of the American Oil
Chemists’ Society 87(10): 1161–​1165.
Ayerza, R. 2013. Seed composition of two chia (Salvia hispanica L.) genotypes which differ
in seed color. Emirates Journal of Food and Agriculture 25(7): 495–​500.
Bañuelos-​Pineda, J., C. C. Gómez-​Rodiles, R. Cuéllar-​José, and O. Aguirre López. 2018.
The Maize Contribution in Human Health. Corn: Production and Human Health in
Changing Climate, 29. www.int​echo​pen.com/​books/​corn-​pro​duct​ion-​and-​human-​
hea​lth-​in-​chang​ing-​clim​ate/​the-​maize-​contr​ibut​ion-​in-​the-​human-​hea​lth
Bello-​Pérez, L. A., G. A. Camelo-​Mendez, E. Agama-​Acevedo, and R. G. Utrilla-​Coello.
2016. Aspectos nutracéuticos de los maíces pigmentados: Digestibilidad de los
carbohidratos y antocianinas. Agrociencia 50(8): 1041–​1063.
Bello-​Perez, L. A., P. C. Flores-​Silva, E. Agama-​Acevedo, and J. Tovar. 2020. Starch digest-
ibility: Past, present, and future. Journal of the Science of Food and Agriculture
100(14): 5009–​5016.
Belščak-​Cvitanović, A., K. Durgo, A. Huđek, V. Bačun-​Družina, and D. Komes. 2018.
Overview of polyphenols and their properties. In Polyphenols: Properties, recovery,
and Applications, 3–​44. Cambridge, UK: Woodhead Publishing.
Bever, A. M., A. Cassidy, E. B. Rimm, M. J. Stampfer, and D. J. Cote. 2021. A prospective
study of dietary flavonoid intake and risk of glioma in US men and women. The
American Journal of Clinical Nutrition 114(4): 1314–​1327.
Bhornchai, H., S. Bhalang, T. Ratchada, M. Scott, and L. Kamol. 2014. Anthocyanin,
phenolics and antioxidant activity changes in purple waxy corn as affected by trad-
itional cooking. Food Chemistry 164: 510–​517.
Bhosale, S. S., B. S. Agarkar, R. B. Kshirsagar, and B. M. Patil. 2021. Studies on physico-
chemical properties of cereals (Rice, sorghum, finger millet, amaranth) and pulses
(Green gram, black gram and chickpea). Pharma Innovation 10(4): 110–​114.
Bravo, M., J. Reyna, and M. Huapaya, 2013. Estudio químico y nutricional de granos
andinos germinados de quinua (Chenopodium Quinoa) y Kiwicha (Amarantus
Caudatus). Revista Peruana de Química e Ingeniería Química 16(1): 54–​60.
Bressani, R. 1990. Chemistry, technology, and nutritive value of maize tortillas. Food
Reviews International 6(2): 225–​264.
Bressani, R. 1991. Protein quality of high-​lysine maize for humans. Cereal Foods World
36(9): 806–​811.
Bressani, R. (2018). Composition and nutritional properties of amaranth. In Amaranth
Biology, Chemistry, and Technology, edited by Paredes Lopez, 185–​205. London:
CRC Press.
Calliope, S. R., M. O. Lobo, and N. C. Sammán. 2015. Proceso de elaboración de hojuelas
cocidas de quínoa (Chenopodium quinoa Willd). Archivos Latinoamericanos de
Nutrición 65(4): 234–​242.
332 Compounds of Latin-American Crop Grains

Carvajal-​Larenas, F. E., A. R. Linnemann, M. J. R. Nout, M. Koziol, and M. A. J. S. van


Boekel. 2016. Lupinus mutabilis: Composition, uses, toxicology, and debittering.
CriticalReviews in Food Science and Nutrition 56(9): 1454–​1487.
Cázares-​Sánchez, E., J. L. Chávez-​Servia, Y., Salinas-​Moreno, F. Castillo-​González, and P.
Ramírez-​Vallejo. 2015. Variación en la composición del grano entre poblaciones de
maíz (Zea mays l.) nativas de Yucatán, México. Agrociencia 49(1): 15–​30.
Chirinos, R., G. Zuloeta, R. Pedreschi, E. Mignolet, Y. Larondelle, and D. Campos. 2013.
Sacha inchi (Plukenetia volubilis): A seed source of polyunsaturated fatty acids,
tocopherols, phytosterols, phenolic compounds and antioxidant capacity. Food
Chemistry 141(3): 1732–​1739.
Chirinos, R., O. Necochea, R. Pedreschi, and D. Campos. 2016. Sacha Inchi
(Plukenetiavolubilis L.) shell: An alternative source of phenolic compounds and
antioxidants. International Journal of Food Science & Technology 51: 986–​993.
Chirinos R., K. Ochoa, A. Aguilar-​Galvez et al. 2018. Obtaining of peptides with in vitro
antioxidant and angiotensin I converting enzyme inhibitory activities from cañihua
protein (Chenopodium pallidicaule Aellen). Journal of Cereal Science 83: 139–​146.
Clavijo, D. B., F. V. Rodríguez, and J. E. C. Estupiñán. 2015. Utilización de plukenetia volubilis
(sacha inchi) para mejorar los componentes nutricionales de la hamburguesa.
Enfoque UTE 6(2): 59–​76.
Coates, W. 2011. Protein content, oil content and fatty acid profiles as potential criteria
to determine the origin of commercially grown chia (Salvia hispanica L.). Industrial
Crops and Products 34(2): 1366–​1371.
Coelho, M. S., and M. D. L. M. Salas-​Mellado. 2014. Chemical characterization of chia
(Salvia hispanica L.) for use in food products. Journal of Food and Nutrition Research
2(5): 263–​269.
Coelho, L. M., P. M. Silva, J. T. Martins, A. C. Pinheiro, and A. A. Vicente. 2018. Emerging
opportunities in exploring the nutritional/​functional value of amaranth. Food and
Function 9(11): 5499–​5512.
Colín-​Chávez, C., J. J. Virgen-​Ortiz, L. E. Serrano-​Rubio, M. A. Martínez-​Téllez, and M.
Astier. 2020. Comparison of nutritional properties and bioactive compounds
between industrial and artisan fresh tortillas from maize landraces. Current
Research in Food Science 3: 189–​194.
Costantini, L., L. Lukšič, R. Molinari, I. Kreft, G. Bonafaccia, L. Manzi, and N. Merendino.
2014. Development of gluten-​free bread using tartary buckwheat and chia flour rich
in flavonoids and omega-​3 fatty acids as ingredients. Food Chemistry 165: 232–​240.
Cravioto, R. O., R. K. Anderson, E. E. Lockhart, F. de P. Miranda, and R. S. Harris.1945.
Nutritive value of the Mexican tortilla. Science 102: 91–​93.
Cuevas Espinal, M. M., and N. R. Lozano Julián 2017. Actividad antioxidante y
composición de ácidos grasos, tocoferoles y tocotrienoles en tres variedades de
Chenopodium quinoa Willdenow y elaboración de una crema dermocosmética
antienvejecimiento. Thesis de grado. Lima: Universidad Nacional Mayor de San
Marcos. https://​cyb​erte​sis.unmsm.edu.pe/​bitstr​eam/​han​dle/​20.500.12672/​7113/​
Cuevas​_​em.pdf ?seque​nce=​1.
Curti, C. A., R. N. Curti, N. Bonini, and A. N. Ramón. 2018. Changes in the fatty acid
composition in bitter Lupinus species depend on the debittering process. Food
Chemistry 263: 151–​154.
References 333

Da Silva, B. P., P. C. Anunciação, J. C. da Silva Matyelka, C. M. Della Lucia, H. S. D. Martino,


and H. M. Pinheiro-​Sant’Ana. 2017. Chemical composition of Brazilian chia seeds
grown in different places. Food Chemistry 221: 1709–​1716.
D’Amico, S., and R. Schoenlechner. 2017. Amaranth: Its unique nutritional and health-​
promoting attributes. In Gluten-​ Free Ancient Grains, 131–​ 159. Cambridge,
UK: Woodhead Publishing.
Das Gupta, S., and N. Suh. 2016. Tocopherols in cancer: An update. Molecular Nutrition
and Food Research 60: 1354–​1363.
Davis, P. J., and S. C. Williams (1998). Protein modification by thermal processing. Allergy
53: 102–​105.
De la Parra, C., S. S. O. Serna, and L. R. Hai. 2007. Effect of processing on the
phytochemical profiles and antioxidant activity of corn for production of masa,
tortillas, and tortilla chips. Journal of Agriculture and Food Chemistry 55(10):
4177–​4183.
De Souza, A. H. P., A. K. Gohara, Â. C. Rodrigues, N. E. De Souza, J. V. Visentainer, and
M. Matsushita. 2013. Sacha inchi as potential source of essential fatty acids and
tocopherols: Multivariate study of nut and shell. Acta Scientiarum Technology
35(4): 757–​763.
Delchier, N., C. Ringling, J. Le Grandois, D. Aoudé-​Werner, R. Galland, S. Georgé, M.
Rychlik, and C. M. Renard. 2013. Effects of industrial processing on folate content in
green vegetables. Food Chemistry 139(1–​4): 815–​824.
Dyner, L., S. Drago, A. Pineiro, H. Sanchez, R. Gonzalez, and M. E. Valencia. 2007.
Composición y aporte potencial de hierro, calcio y zinc de panes y fideos
elaborados con harinas de trigo y amaranto. Archivos Latinoamericanos de
Nutrición 57: 69–​77.
Ejigui, I., L. Savoie, I. Marin, and T. Desrosiers. 2005. Antinutritional factors of red peanuts
(Arachis hypogea) and small red kidney beans (Phaseolus vulgaris). Journal of
Biological Sciences 5(5): 597–​605.
Elena, M., L. Marroú, V. González, S. Elizabeth, and P. Flores 2011. Composición química
de oca (Oxalis tuberosa), arracacha (Arracaccia xanthorriza) y tarwi (Lupinus
mutabilis). Formulación de una mezcla base para productos alimenticios. Revista
Venezolana de Ciencia y Tecnología de Alimentos 2(2): 239–​252.
Escalante-​Aburto, A., R. M. Mariscal-​Moreno, D. Santiago-​Ramos, and N. Ponce-​García.
2020. An update of different nixtamalization technologies, and its effects on chem-
ical composition and nutritional value of corn tortillas. Food Reviews International
36(5): 456–​498.
FAO (Food and Agriculture Organization). 1985. Joint FAO/​WHO/​UNU Expert Consultation
on Protein-​Energy Requirements. Geneva: WHO.
FAO (Food and Agriculture Organization). 2016. Legume consumption and production
has lost strength in Latin America and the Caribbean compared to more commer-
cial crops. www.fao.org/​ameri​cas/​notic​ias/​ver/​es/​c/​455​947/​.
FAO (Food and Agriculture Organization) and FINUT (Fundación Iberoamericana
de Nutrición). 2017. Evaluación de la Calidad de la proteína de la dieta humana.
Consulta de Expertos. Granada: Publicado por Organización de las Naciones
Unidas para la Alimentación y a Agricultura (FAO) y la Fundación Iberoamericana
de Nutrición.
334 Compounds of Latin-American Crop Grains

Fardet, A., E. Rock, and C. Remesy. 2008. Is the in vitro antioxidant potential of whole-​
grain cereals and cereal products well reflected in vivo? Journal of Cereal Science
48(2): 258–​276.
Fernández Mejía, J. L., and C. L. Guivar Delgado. 2020. Formulación de harina proteica
y extruida a base de harina de: arveja (Pisum sativum), kiwicha (Amaranthus
caudatus) y tarwi (Lupinus Mutabilis). https://​ali​cia.concy​tec.gob.pe/​vuf​i nd/​Rec​
ord/​UPRG_​dbc98​eb5d​2d40​3153​f75c​bfcd​8b29​b7b/​Deta​ils.
Filho, A. M. M., M. R. Pirozi, J. T. D. S. Borges, H. M. Pinheiro Sant’Ana, J. B. P. Chaves, and J.
S. D. R. Coimbra. 2017. Quinoa: nutritional, functional, and anti-​nutritional aspects.
Critical Reviews in Food Science and Nutrition 57(8): 1618–​1630.
Gade, D. W. 1970. Ethnobotany of cañihua (Chenopodium pallidicaule), rustic seed crop of
the Altiplano. Econonic Botany 24(1): 55–​61.
Galinat, W. C. 1971. The origin of maize. Annual Review of Genetics 5(1): 447–​478.
Ganesan, B., C. Brothersen, and D. J. McMahon. 2014. Fortification of foods with omega-​
3 polyunsaturated fatty acids. Critical Reviews in Food Science and Nutrition
54(1): 98–​114.
García, D. C., A. Mateo del Pino, and N. Pascual Soler 2015. Comidas bastardas.
Gastronomía, tradición e identidad en América Latina. Revista de Filología y
Lingüística de la Universidad de Costa Rica 41(2): 198–​202.
Gargiulo, L., Å Grimberg, R. Repo-​Carrasco-​Valencia, A. S. Carlsson, and G. Mele. 2019.
Morpho-​ densitometric traits for quinoa (Chenopodium quinoa Willd.) seed
phenotyping by two X-​ray micro-​CT scanning approaches. Journal of Cereal Science
90: 102829.
Giménez, M. A., C. N. Segundo, M. O. Lobo, and N. Samman. 2020. Physicochemical and
techno-​functional characterization of native corn reintroduced in the Andean
Zone of Jujuy, Argentina. Procceding la ValSe-​Food 53(1): 7.
González, J. A., Y. Konishi, M. Bruno, M. Valoy, and F. E. Prado. 2012. Interrelationships
among seed yield, total protein and amino acid composition of ten quinoa
(Chenopodium quinoa) cultivars from two different agroecological regions. Journal
of the Science of Food and Agriculture 92(6): 1222–​1229.
Grundy, M. M., D. K. Momanyi, C. Holland, F. Kawaka, S. Tan, M. Salim, and W. O. Owino.
2020. Effects of grain source and processing methods on the nutritional profile and
digestibility of grain amaranth. Journal of Functional Foods 72: 104065.
Gualberto, D., C. Bergman, M. Kazemzadeh, and C. Weber. 1997. Effect of extrusion pro-
cessing on the soluble and insoluble fiber and phytic acid contents of cereal brands.
Plant Foods for Human Nutrition 51: 187–​198.
Guillén-​Sánchez, J., S. Mori-​Arismendi, and L. M. Paucar-​Menacho. 2014. Características
y propiedades funcionales del maíz morado (Zea mays L.) var. subnigroviolaceo.
Scientia Agropecuaria 5(4): 211–​217.
Gutiérrez, L. F., L. M. Rosada, and A. Jiménez. 2011. Chemical composition of Sacha Inchi
(Plukenetia volubilis L.) seeds and characteristics of their lipid fraction. Grasas y
aceites 62(1): 76–​83.
Guzman-​Maldonado, S., and O. Paredes-​Lopez. 1998. Functional products of plants indi-
genous to Latin America. Amaranth and quinoa, common beans and botanicals. In
Functional Foods. Biochemical and Processing Aspects, edited by G. Mazza, 293–​328.
Lancaster: Technomic Publishing Company.
References 335

Guzman-​Maldonado S. H., M. G. Vazquez-​Carrillo, J. A. Aguirre-​Gómez, and I. Serrano-​


Fujarte. 2015. Contenido de ácidos grasos, compuestos fenólicos y calidad indus-
trial de maíces nativos de Guanajuato. Revista Fitotecnia Mexicana 8(2): 213–​222.
Hall, C., C. Hillen, and J. G. Robinson. 2017. Composition, nutritional value, and health
benefits of pulses. Cereal Chemistry 94(1): 11–​31.
Harnly, J. 2015. Antioxidant methods. Journal of Food Composition and Analysis
64(2): 145–​146
Haros, C. M., and R. Schöenlechner (Eds.). 2017. Pseudocereals: Chemistry and Technology.
Hoboken, NJ: John Wiley and Sons.
Hayat, I., A. Ahmad, T. Masud, A. Ahmed, and S. Bashir. (2014). Nutritional and health
perspectives of beans (Phaseolus vulgaris L.): An overview. Critical Reviews in Food
Science and Nutrition 54(5): 580–​592.
He, S., B. K. Simpson, H. Sun, M. O. Ngadi, Y. Ma, and T. Huang. 2015. Phaseolus vulgaris
lectins: A systematic review of characteristics and health implications. Critical
Reviews in Food Science and Nutrition 58(1): 70–​83.
Hefni, M., and C. M. Witthöft. 2011. Increasing the folate content in Egyptian baladi
bread using germinated wheat flour. LWT –​ Food Science and Technology 44:
706–​712.
Hemalatha, S., K. Platel, and K. Srinivasan. 2007. Zinc and iron contents and their
bioaccessibility in cereals and pulses consumed in India. Food Chemistry
102: 1328–​1336.
Hernández, A. D. S. C. 2018. Composición química, características de calidad y actividad
antioxidante de pasta enriquecida con harina de amaranto y hoja de amaranto
deshidratada. http://ri-ng.uaq.mx/​han​dle/​123456​789/​921.
Huaccho-​Human, C. V., and M. Lope-​Surichaqui. 2007. Elaboración de una mezcla
alimenticia a base de tarwi (lupinus mutabilis sweet), Quinua (chenopodium quinoa
willd), Maca (lepidium peruvianum chacón), y Lúcuma (pouteria lucuma) mediante
extrusión. Doctoral dissertation, Universidad Nacional del Centro de Peru, Junin,
Peru. http://​hdl.han​dle.net/​20.500.12894/​1872.
Jing, P., and M. Giusti. 2007. Effects of extraction conditions on improving the yield and
quality of an anthocyanin-​rich purple corn (Zea mays L.) color extract. Journal of
Food Science 72(2): 366–​368.
Katunzi-​Kilewela, A., L. D. Kaale,O. Kibazohi, and L. M. Rweyemamu. 2021. Nutritional,
health benefits and usage of chia seeds (Salvia hispanica): A review. African Journal
of Food Science 15(2): 48–​59.
Khattab, R. Y., and M. A. Zeitoun. 2013. Quality of flaxseed oil obtained by different extrac-
tion techniques. LWT –​Food Science and Technology 53: 338–​345.
Koehler, P., G. Hartmann, H. Wieser and M. Rychlik. 2007. Changes of folates, dietary fiber,
and proteins in wheat as affected by germination. Journal of Agriculture and Food
Chemistry 55: 4678–​4683.
Laurente Flores, Y. R. 2016. Obtención del concentrado proteico y determinación del perfil
de aminoácidos de dos variedades de Tarwi (Lupinus mutabilis Sweet). Doctoral dis-
sertation. Universidad Nacional del Altiplano, Puno, Peru. https://​repos​itor​iosl​atin​
oame​rica​nos.uch​ile.cl/​han​dle/​2250/​3275​226.
López-​Martinez, L., K. L. Parkin, and H. S. García. 2011. Phase II –​inducing, polyphenols
content and antioxidant capacity of corn (Zea mays L.) from phenotypes of white,
336 Compounds of Latin-American Crop Grains

blue, red and purple colors processed into masa and tortillas. Plant Foods for
Human Nutrition 66(1): 41–​47.
Los, F. G. B., A. A. F. Zielinski, J. P. Wojeicchowski, A. Nogueira and I. M. Demiate 2018.
Beans (Phaseolus vulgaris L.): Whole seeds with complex chemical composition.
Current Opinion in Food Science 19: 63–​71.
Lv, J. S., X. Y. Liu, X. P. Zhang and L. S. Wang. 2017. Chemical composition and functional
characteristics of dietary fiber-​rich powder obtained from core of maize straw.
Food Chemistry 227: 383–​389.
Mackay, W., T. Davis, and D.Sankhla. (1995). Influence of scarification and temperature
treatments on seed germination of Lupinus havardii. Seed Science and Technology
23(3): 815–​821.
Mattila, P., J. Astola, and J. Kumpulainen. 2000. Determination of flavonoids in plant
material by HPLC with diode-​ array and electro-​array detections. Journal of
Agricultural and Food Chemistry 48(12): 5834–​5841.
Maureen, M., A. N. Kaaya, J. Kauffman, C. Narrod and A. Atukwase. 2020. Enhancing
nutritional benefits and reducing mycotoxin contamination of maize through
nixtamalization. Journal of Biological Sciences 20: 153–​162.
Maurer, N. E., B. Hatta-​ Sakoda, G. Pascual-​ Chagman, and L. E Rodriguez-​ Saona
2012. Characterization and authentication of a novel vegetable source of
omega-​3 fatty acids, sacha inchi (Plukenetia volubilis L.) oil. Food Chemistry 134(2):
1173–​1180.
Mazewski, C., K. Liang, and E. G. de Mejia. 2017. Inhibitory potential of anthocyanin-​rich
purple and red corn extracts on human colorectal cancer cell proliferation in vitro.
Journal of Functional Foods 34: 254–​265.
Mburu, M. W. 2016. Properties of a Complementary Food based on Grain Amaranth
(Amaranthus cruentus). Doctoral dissertation. Food Science and Technology,
JKUAT) http://​journ​als.jkuat.ac.ke/​index.php/​pgthe​sis_​abs/​arti​cle/​view/​408.
Menard, S., N. Cerf-​Bensussan, and M. Heyman. 2010. Multiple facets of intestinal per-
meability and epithelial handling of dietary antigens. Mucosal Immunology
3(3): 247–​259.
Miranda, M., A. Vega-​Gálvez, I. Quispe-​Fuentes, M. J. Rodríguez, H. Maureira, and E. A.
Martínez. 2012. Nutritional aspects of six quinoa (Chenopodium quinoa Willd.)
ecotypes from three geographical areas of Chile. Chilean Journal of Agricultural
Research 72(2): 175.
Motta, C., A. C. Nascimento, M. Santos, I. Delgado, I. Coelho, A. Rego, A. S Matos, D.
Torres, and I. Castanheira. 2016a. The effect of cooking methods on the min-
eral content of quinoa (Chenopodium quinoa), amaranth (Amaranthus sp.) and
buckwheat (Fagopyrum esculentum). Journal of Food Composition and Analysis
49: 57–​64.
Motta, C., M. Santos, R. Mauro, N. Samman, A. S. Matos, D. Torres, and I. Castanheira
.2016b. Protein content and amino acids profile of pseudocereals. Food Chemistry
193: 55–​61.
Motta, C., I. Delgado, A. S. Matos, G. B. Gonzales, D. Torres, M. Santos, V. Chandra-​Hioe,
J. Arcot, and I. Castanheira. 2017. Folates in quinoa (Chenopodium quinoa), amar-
anth (Amaranthus sp.) and buckwheat (Fagopyrum esculentum): Influence of
cooking and malting. Journal of Food Composition and Analysis 64: 181–​187.
References 337

Motta, C., I. Castanheira, G. B. Gonzales, I. Delgado, D. Torres, M. Santos, and A. S. Matos.


2019. Impact of cooking methods and malting on amino acids content in amaranth,
buckwheat and quinoa. Journal of Food Composition and Analysis 76: 58–​65.
Muller-​Tito, K. 2015. Capacidad antioxidante y contenido de flavonoides entre las semillas
de chia negra (Salvia Nativa) y chia blanca (Salvia hispánica L.). Bachelor disserta-
tion. Universidad Nacional del Altiplano, Puno, Peru.
Muñoz, L. A., A. Cobos, O. Diaz, and J. M. Aguilera. 2013. Chia seed (Salvia hispanica): an
ancient grain and a new functional food. Food Reviews International 29(4): 394–​408.
Nascimento, A. C., C. Mota, I. Coelho, S. Gueifão, M. Santos, A. S. Matos, A. Gimenez, M.
Lobo, N. Samman, and I. Castanheira. 2014. Characterization of nutrient profile of
quinoa (Chenopodium quinoa), amaranth (Amaranthus caudatus), and purple corn
(Zea mays L.) consumed in the north of Argentina: Proximate, minerals and trace
elements. Food Chemistry 148: 420–​426.
Navarro Cortez, R. O., C. A. Gómez-​Aldapa, E. Aguilar-​Palazuelos, E. Delgado-​Licon,
J. Castro Rosas, J. Hernández-​ Ávila, A. Solís-​
Soto, L. A. Ochoa-​ Martínez, and
H. Medrano-​Roldán. 2016. Blue corn (Zea mays L.) with added orange (Citrus
sinensis) fruit bagasse: Novel ingredients for extruded snacks. CyTA Journal of Food
14(2): 349–​358.
Nelson, K., L. Stojanovska, T. Vasiljevic, and M. Mathai. 2013. Germinated grains: A superior
whole grain functional food? Canadian Journal of Physiology and Pharmacology
91(6): 429–​441.
Nuss, E. T., and S. A. Tanumihardjo. 2010. Maize: A paramount staple crop in the con-
text of global nutrition. Comprehensive Reviews in Food Science and Food Safety
9(4): 417–​436.
Oboh, G., A. O. Ademiluyi, and A. A. Akindahunsi. 2010. The effect of roasting on the nutri-
tional and antioxidant properties of yellow and white maize varieties. International
Journal of Food Science and Technology 45(6): 1236–​1242.
Pachon, H., D. A. Ortiz, C. Araujo, M. W. Blair, and J. Restrepo. 2009. Iron, zinc, and pro-
tein bioavailability proxy measures of meals prepared with nutritionally enhanced
beans and maize. Journal of Food Science 74: 147–​154.
Paredes López, O., F. Guevara Lara, and L. A. Bello Pérez. 2009. La nixtamalización y el
valor nutritivo del maíz. Ciencias 92: 60–​70. www.revi​stac​ienc​iasu​nam.com/​es/​
compon​ent/​cont​ent/​arti​cle/​41-​revis​tas/​revi​sta-​cienc​ias-​92-​93/​205-​la-​nixt​amal​
izac​ion-​y-​el-​valor-​nutrit​ivo-​del-​maiz-​05.html.
Pasini, G., B. Simonato, M. Giannattasio, A. D. Peruffo, and A. Curioni (2001). Modifications
of wheat flour proteins during in vitro digestion of bread dough, crumb, and
crust: An electrophoretic and immunological study. Journal of Agricultural and Food
Chemistry 49(5): 2254–​2261.
Paucar-​Menacho, L. M., R. Salvador-​Reyes, J. Guillén-​Sánchez, J. Capa-​Robles, and C.
Moreno-​Rojo. 2015. Estudio comparativo de las características físico-​químicas
del aceite de sacha inchi (Plukenetia volubilis L.), aceite de oliva (Olea europaea) y
aceite crudo de pescado. Scientia Agropecuaria 6(4): 279–​290.
Peñarrieta, J. M., J. A. Alvarado, B. Åkesson, and B. Bergenståhl. 2008. Total antioxidant
capacity and content of flavonoids and other phenolic compounds in canihua
(Chenopodium pallidicaule): an Andean pseudocereal. Molecular Nutrition and Food
Research 52: 708–​717.
338 Compounds of Latin-American Crop Grains

Perez-​Hernandez, L. M., K. Nugraheni, M. Benohoud, W. Sun, A. J. Hernández-​Álvarez,


M. R. A. Morgan, C. Boesch, and C. Orfila. 2020. Starch digestion enhances
bioaccessibility of anti-​inflammatory polyphenols from borlotti beans (Phaseolus
vulgaris). Nutrients 12(2): 295.
Písaříková, B., S. Kráčmar, and I. Herzig. 2005. Amino acid contents and biological
value of protein in various amaranth species. Czech Journal of Animal Science
50(4): 169–​174.
Ramos Diaz, J. M., S. Kirjoranta, S. Tenitz, P. A. Penttilä, R. Serimaa, A. M. Lampi, and K.
Jouppila. 2013. Use of amaranth, quinoa and kañiwa in extruded corn-​based snacks.
Journal of Cereal Science 58(1): 59–​67.
Ranilla, L. G. 2019. Bioactive ingredients from corn and lactic acid bacterial biotransform-
ation. In Functional Foods and Biotechnology, 19–​45). Boca Raton, FL: CRC Press.
Rawdkuen, S., D. Murdayanti, S. Ketnawa, and S. Phongthai. 2016. Chemical proper-
ties and nutritional factors of pressed-​cake from tea and sacha inchi seeds. Food
Bioscience 15: 64–​71.
Repo-​Carrasco-​Valencia, R. 2011. Andean Indigenous Food Crops: Nutritional Value and
Bioactive Compounds. Doctoral dissertation on Food Chemistry, Finland: University
of Turku.
Repo-​Carrasco-​Valencia, R., C. Espinoza, and S. E. Jacobsen. 2003. Nutritional value and
use of the Andean crop’s quinoa (Chenopodium quinoa) and kañiwa (Chenopodium
pallidicaule). Food Reviews International 19: 179–​189.
Repo-​Carrasco-​Valencia, R., A. de la Cruz, J. C. Alvarez, and H. Kallio. 2009. Chemical and
functional characterization of kañiwa (Chenopodium pallidicaule) grain, extrudate
and bran. Plant Foods for Human Nutrition 64: 94–​101.
Repo-​Carrasco-​Valencia, R., C. R. Encina, C. Binaghi, M. Greco, and P. Ronayne de Ferrer.
2010. Effects of roasting and boiling of quinoa, kiwicha and kaniwa on compos-
ition and availability of minerals in vitro. Journal of Science of Food and Agriculture
90: 2068–​2073.
Repo-​Carrasco-​Valencia R., S. Melgarejo-​Cabello, and J. M. Pihlava. 2019. Nutritional
value and bioactive compounds in quinoa (Chenopodium quinoa Willd.), kañiwa
(Chenopodium pallidicaule Aellen) and kiwicha (Amaranthus caudatus L.). In
Quinoa: Cultivation, Nutritional Properties and Effects on Health, edited by P. Peiretti
and F. Gai, 83–​113. New York: Nova Science Publishers.
Rincon-​Londoño, N., L. J. Vega-​Roja, M. Contreras-​Padilla, A. A. Acosta-​Osorio, and M. E.
Rodrıguez-​Garcıa. 2016. Analysis of the pasting profile in corn starch: Structural,
morphological, and thermal transformations. Part I. International Journal of
Biological Macromolecules 91: 106–​114.
Ross, A.B., M. J. Shepherd, and M. Schüpphaus. 2003. Alkylresorcinols in cereals and cereal
products. Journal of Agriculture and Food Chemistry 51: 4111–​4118.
Rossi, M. C., M. N. Bassett, and N. C. Samman. 2018. Dietary nutritional profile and
phenolic compounds consumption in school children of highlands of Argentine
Northwest. Food Chemistry 238: 111–​116.
Ruiz, C., C. Diaz, J. Anaya, and R. Rojas. 2013. Análisis proximal, antinutrientes, perfil de
ácidos grasos y de aminoácidos de semillas y tortas de 2 especies de Sacha inchi
(Plukenetia volubilis y Plukenetia huayllabambana). Revista de la Sociedad Química
del Perú 79(1): 29–​36.
References 339

Sabelino-​Francia, Z. D. P. 2020. Modelos de calibración del contenido de proteína y


fenólicos totales usando espectroscopia del infrarrojo medio en Tarwi (Lupinus
mutabilis). Doctoral dissertation, Universidad Nacional Agraria La Molina, Lima,
Peru. https://​repo​sito​rio.lamol​ina.edu.pe/​han​dle/​20.500.12996/​4428.
Salvador-​Reyes, R., and M. T. Pedrosa-​Silva-​Clerici. 2020. Peruvian Andean maize: General
characteristics, nutritional properties, bioactive compounds, and culinary uses.
Food Research International 130: 108934.
Salvatierra-​Pajuelo, Y. M., M. E. Azorza-​Richarte, and L. M. Paucar-​Menacho. 2019.
Optimización de las características nutricionales, texturales y sensoriales de
cookies enriquecidas con chía (Salvia hispánica) y aceite extraído de tarwi (Lupinus
mutabilis). Scientia Agropecuaria 10(1): 7–​17.
Sánchez Ortega, I., and E. Pérez-​Urria-​Carril. 2014. Maíz I (Zea mays). REDUCA Biologia
7(2): 151–​171.
Sánchez-​Sánchez G. L. 2012. Caracterización y cuantificación de los ácidos grasos omega
3 y omega 6 presentes en el aceite de sacha inchi (Plukenetia volubilis L). Escuela de
Química. Masters thesis. Bogotá: Universidad Nacional de Colombia. https://​repo​
sito​rio.unal.edu.co/​han​dle/​unal/​11657.
Serna-​Saldivar, S. O., and C. Chuck-​Hernandez. 2019. Food uses of lime-​cooked corn with
emphasis in tortillas and snacks. In Corn (3rd edn). Chemical and Technology, Chap
17, 469–​500. https://​doi.org/​10.1016/​B978-​0-​12-​811​971-​6.00017-​.
Shah, Tajamul Rouf, K. Prasad, P. Kumar, and F. Yildiz. 2016. Maize. A potential source
of human nutrition and health: A review. Cogent Food and Agriculture 2(1):
1166995.
Shohag, M. J. I., Y. Wei, and X. Yang. 2012. Changes of folate and other potential health-​
promoting phytochemicals in legume seeds as affected by germination. Journal of
Agriculture and Food Chemistry 60: 9137–​9143.
Sierra-​Macías, M., P. Andrés-​Meza, A. Palafox-​Caballero, and I. Meneses-​Márquez.
2016. Diversidad genética, clasificación y distribución racial del maíz nativo
en el estado de Puebla, México. Revista de Ciencias Naturales y Agropecuarias
3(9): 12–​21.
Siyuan-​Sheng, L. T., and H. Liu-​Rui. 2018. Corn phytochemicals and their health benefits.
Food Science and Human Wellness 7: 185–​195.
Smorowska, A. J., A.K. Żołnierczyk, A. Nawirska-​Olszańska, J. Sowiński, and A. Szumny.
2021. Nutritional properties and in vitro antidiabetic activities of blue and yellow
corn extracts: A comparative study. Journal of Food Quality https://​doi.org/​10.1155/​
2021/​8813​613.
Stea, T. H., M. Johansson, M. Jägerstad, and W. Frølich. 2007. Retention of folates in
cooked, stored and reheated peas, broccoli and potatoes for use in modern large-​
scale service systems. Food Chemistry 101: 1095–​1107.
Stikić, R. I., D. D. Milinčić, A. Ž. Kostić, Z. B. Jovanović, U. M. Gašić, Ž. L.Tešić, N. Z.
Djordjević, S. K. Savić, B. G. Czekus, and M. B. Pešić. 2020. Polyphenolic profiles,
antioxidant, and in vitro anticancer activities of the seeds of Puno and Titicaca
quinoa cultivars. Cereal Chemistry 97(3): 626–​633.
Suárez, P. A., J. G. Martínez, and J. R. Hernández. 2013. Amaranto: Efectos en la nutrición
y la salud. Tlatemoani: Revista Académica de Investigación 12: 1. https://​eco​npap​ers.
repec.org/​arti​cle/​ervtla​tem/​y_​3​a201​3_​3a​i_​3a​12_​3​a14.htm.
340 Compounds of Latin-American Crop Grains

Suca, G. R. 2015. Potencial del tarwi (Lupinus mutabilis Sweet) como futura fuente
proteínica y avances de su desarrollo agroindustrial. Revista Peruana de Química e
Ingeniería Química 18(2): 55–​71.
Ullah, R., M. Nadeem, A. Khalique, M. Imran, S. Mehmood, A. Javid, and J. Hussain. 2016.
Nutritional and therapeutic perspectives of Chia (Salvia hispanica L.): A review.
Journal of Food Science and Technology 53(4): 1750–​1758.
Valenzuela-​Zamudio, F., and M. R. Segura-​Campos. 2020. Amaranth, quinoa and chia bio-
active peptides: A comprehensive review on three ancient grains and their poten-
tial role in management and prevention of Type 2 diabetes. Critical Reviews in Food
Science and Nutrition: 62(10): 2707–​2721.
Valles Ramírez, S. M. 2012. Obtención de Leche de Sacha Inchi (Plukenetia Volubilis
Linneo). Tesis Ingeniería. Tarapoto: Universidad Nacional de San Martín, Facultad
de Ingeniería Agroindustrial. https://​repo​sito​rio.unsm.edu.pe/​han​dle/​11458/​2271.
Vera, J. L. 2014. Perfil de ácidos grasos en granos tres cultivares de quinua (Chenopodium
quinoa Willd.) sometidos a tres tipos de procesamiento. Revista Investigaciones
Altoandinas 16(1): 13–​20.
Vilcacundo, R., and B. Hernández-​Ledesma. 2017. Nutritional and biological value of
quinoa (Chenopodium quinoa Willd.). Current Opinion in Food Science 14: 1–​6.
Villanueva, E., G. Rodríguez, E. Aguirre, and V. Castro. 2017. Influence of antioxidants
on oxidative stability of the oil chia (Salvia hispanica L.) by rancimat. Scientia
Agropecuaria 8(1): 19–​27.
Wang, N., D. W. Hatcher, R. T. Tyler, R. Toews, and E. J. Gawalko. 2010. Effect of cooking on
the composition of beans (Phaseolus vulgaris L.) and chickpeas (Cicer arietinum L.).
Food Research International 43(2): 589–​594.
Wang, S., F. Zhu, and Y. Kakuda. 2018. Sacha inchi (Plukenetia volubilis L.): Nutritional
composition, biological activity, and uses. Food Chemistry 265: 316–​328
WHO (World Health Organization) and United Nations University. 2007. Protein and
Amino Acid Requirements in Human Nutrition (Vol. 935). Geneva: World Health
Organization.
Yang, C., and Suh, N. 2013. Cancer prevention by different forms of tocopherols. Topics in
Current Chemistry 329: 21–​33.
Zare, T., T. W. Rupasinghe, B. A. Boughton, and U. Roessner. 2019. The changes in the
release level of polyunsaturated fatty acids (ω-​3 and ω-​6) and lipids in the untreated
and water-​soaked chia seed. Food Research International 126: 108665.
Zettel, V., and B. Hitzmann.2018. Applications of chia (Salvia hispanica L.) in food
products. Trends in Food Science & Technology 80: 43–​50.
Zhang, H., and R. Tsao. 2016. Dietary polyphenols, oxidative stress and antioxidant and
anti-​inflammatory effects. Current Opinion in Food Science 8: 33–​42.
Chapter 9

Contributions from Latin-​American


Grains to Nutrition and Health
Carla Motta1, Norma Sammán2, and Isabel Castanheira1
1National Institute of Health Doutor

Ricardo Jorge, Lisbon, Portugal


2Universidad Nacional de Jujuy and Consejo

Nacional de Investigaciones Científicas y Técnicas


(CONICET), San Salvador de Jujuy, Argentina

CONTENTS
9.1 Introduction 341
9.2 Noncommunicable Diseases: Diabetes Mellitus, Cardiovascular
Risk Factors and Obesity 342
9.2.1 Diabetes Mellitus 342
9.2.2 Cardiovascular Risk Factors 344
9.2.3 Obesity 345
9.3 Cancer 346
9.4 Celiac Disease 351
9.5 Prebiotic Effect and Microbiota Modulation 356
9.5 Conclusions 363
Acknowledgments 364
References 364

9.1 INTRODUCTION
Health effects from Andean crops have been a matter of discussion in recent years.
In this chapter, we are addressing the health benefits, and evidence regarding

DOI: 10.1201/9781003088424-9 341


342 Contributions to Nutrition and Health

the influence of Latin-​American grains such as quinoa (Chenopodium quinoa),


kiwicha (Amaranthus caudatus), kañiwa (Chenopodium pallidicaule), chia
(Salvia hispanica), and legumes such as black turtle bean (Phaseolus vulgaris)
and tarwi (Lupinus mutabilis) on the improvement and maintenance of different
health-​related aspects. The impact of these grains has shown promising results
on disorders like cancer, type-​2 diabetes, allergic conditions, celiac disease, and
on the modulation of cardiovascular risk factors (through hypocholesterolemic
properties) either in in vitro or in vivo approaches or clinical studies with bio-
logical evaluations. Some of the grains, like quinoa and amaranth, have been
reported as gluten-​free foods, revealing their importance to celiac disease or
non-​celiac gluten sensitivity patients, either in the maintenance of a gluten-​free
diet or in the reduction of nutritional deficiencies, respectively (Bergamo et al.,
2011; Zevallos et al., 2014). Due to the high content of bioactive peptides and
antioxidant molecules (such as phenolic compounds) and micronutrient avail-
ability of minerals and vitamins of these grains, different impacts on immuno-
logical and cellular activity can be observed (Ayyash et al., 2019; Lee and Joo,
2018; Tang and Tsao, 2017). Some process methods (e.g., fermentation, germin-
ation) have recently been reported to increase the bioaccessibility of some of
these nutrients. These studies also suggest that prebiotic and probiotic effects
on microbiota modulation can be associated with specific health benefits. Fast-​
growing in vitro and in vivo studies evidence that Andean grain consumption,
like quinoa and amaranth, can modify specific intestinal bacteria (Gullón et al.,
2016). Intestinal microbiota can be associated, for instance, with the decrease
of obesity levels and inflammation-​mediated chronic disorders, working as
a modulator for several chronic disease biomarkers (Ugural and Akyol, 2020).
Compared with other grains or legumes used in western diets, Latin-​American
grains are a good alternative for a balanced diet to prevent noncommunicable
diseases, the major causes of disability, ill-​health, health-​related retirement, and
premature death.

9.2 NONCOMMUNICABLE DISEASES: DIABETES


MELLITUS, CARDIOVASCULAR RISK FACTORS
AND OBESITY
9.2.1 Diabetes Mellitus
Diabetes mellitus is a chronic and metabolic disease characterized by high levels
of blood glucose (or blood sugar). High glucose levels lead over time to severe
health conditions: damage to the kidneys, heart, blood vessels, eyes, and nerves.
Type 2 diabetes is common in adults and is characterized by insulin resistance
9.2 Diabetes Mellitus, Cardiovascular Risk Factors and Obesity 343

or low insulin production. Diabetes is a global health problem, presenting


high prevalence and an incidence that rises every day. In 2016, WHO reported
422 million people worldwide with diabetes, especially in low-​and middle-​
income countries (WHO, 2021).
Several research publications are available concerning the association
between dietary factors and the increase in the incidence of type 2 diabetes.
Comprehensive systematic reviews and meta-​analyses have been published in
recent years showing evidence of association between dietary behaviors or diet
quality and this pathology (Neuenschwander et al., 2019). Some studies have also
demonstrated strong evidence of the correlation between higher consumption
of whole grains, beans, nuts, or grains, and dietary fiber and the decrease of the
incidence of type 2 diabetes (Bellou et al., 2018; Micha et al., 2017; Schwingshackl
et al., 2017).
As already described throughout the book, almost all Latin-​American crops,
like quinoa, amaranth, kañiwa, or black turtle beans, are rich in dietary fiber and
should be used as a healthy base diet to prevent and control type 2 diabetes and
obesity.
Indeed, amaranth’s capacity to modulate glycemic and insulinemic response
was studied by Guerra-​Matias and Arêas (2005), showing that the presence of
dietary fiber is correlated with a decrease in glycemic response. The content of
pectins, the main fiber fraction not digested and not absorbed in the small intes-
tine, can be responsible for the decrease of rises in glucose after meals and for a
satiety effect. The same conclusion was addressed in a review study conducted
by Marventano et al. (2017), which resumes several clinical trials that evaluate
and demonstrate the effects in healthy subjects, with high consumption of whole
grain diets rich in fiber, the improvement of postprandial glucose and insulin
response.
Graf et al. (2014), in their studies using mice observed that the hypergly-
cemic condition improved with quinoa consumption. Biologically active
phytoecdysteroids, flavonoids, and proteins present at high levels in quinoa
grains lead to lower fasting blood glucose in obese, hyperglycemic mice.
The same conclusion was reported by the Fornasini team in two different
human studies addressing lupin consumption (Baldeón et al., 2012; Fornasini
et al., 2012). Purple corn extracts are characterized by a rich composition of
anthocyanins and functional phenolics (Baldeón et al., 2012). Anthocyanins
have been reported to exhibit antiangiogenic, antidiabetic, and anticarcinogenic
properties (Burton-​Freeman et al., 2019; Sancho and Pastore, 2012). In in vitro
studies, the anthocyanins present in purple corn extracts can attenuate high
glucose levels, retarding diabetes-​associated renal fibrosis (Li et al., 2012).
Recently, other authors in an in silico essay refer to the role and importance of
purple maize, due to its phenolic compounds, in controlling obesity, diabetes,
344 Contributions to Nutrition and Health

and anti-​inflammatory action (Zhang et al., 2019; Damián-​Medina et al., 2020).


In vitro and clinical studies address the importance of essential nutrients like
amino acids, phytochemicals (phenolic acids and flavonoids), fatty acids, espe-
cially the polyunsaturated fatty acids (PUFAs), carotenoids, and tocopherols,
present in quinoa and amaranth as crucial in lowering the risk of type 2 dia-
betes by assessing antihyperglycemic and antihypertension activity (Tang and
Tsao, 2017).

9.2.2 Cardiovascular Risk Factors


Together, diabetes, obesity, and cardiovascular diseases (CVD), where hyper-
tension is included, are responsible for more deaths than any other conditions.
Approximately 10 million people die each year because of hypertension (Frieden
and Jaffe, 2018). Reduction of sodium from food and increasing legume con-
sumption, like beans (Bazzano et al., 2011; Jiang et al., 2020; Mudryj et al., 2014)
and diets that include Andean grains such as quinoa and amaranth can reduce
the risk factors in the affected individuals (Chmelík, et al., 2019; Jiang et al., 2020;
Tang and Tsao, 2017).
Quinoa and amaranth provide all the essential amino acids (Motta et al.,
2019), which have been identified as modulators of several systemic pathways
due to their high nutraceutical potential demonstrating benefits for human
health (Orona-​Tamayo et al., 2019). Grain amino acids and peptides produced
by protein hydrolysis are associated with a wide range of beneficial effects
such as modulation of blood pressure and lipid metabolism; with anticancer,
immunomodulatory, anti-​ thrombotic, anti-​ atherosclerotic, and anti-​
inflammatory potential as well (Orona-​Tamayo et al., 2019; Millán-​Linares
et al., 2014). Latin-​American ancient grains such as amaranth, quinoa, maize,
common bean, and chia grains have been identified as important sources of
these bioactive peptides, with potential anti-​inflammatory activity, by several
authors (López et al., 2019; Orona-​Tamayo et al., 2019; Valenzuela et al., 2020).
The anti-​inflammatory activity was tested in cell cultures, demonstrating an
essential role in preventing chronic diseases, like diabetes and CVD associated
with chronic inflammation (Millán-​Linares et al., 2014). Lupins, due to their
protein profile content, also contributed to the plasma low density lipopro-
tein (LDL) cholesterol modulation. Bähr et al. (2013) reported that lupin pro-
tein could positively impact cardiovascular risk factors, especially in higher
hypercholesterolemic subjects. Besides bioactive peptides, the antioxidant
activity also demonstrated by these grains provides important molecules that
can modulate immune response. The role of bioactive molecules and the anti-
oxidant activity provided by the consumption of ancient grains may have a posi-
tive impact on several functions in the chronic disease pathways with health
9.2 Diabetes Mellitus, Cardiovascular Risk Factors and Obesity 345

benefits that have been identified in several studies in recent years including
illness prevention (López et al., 2019; Orona-​Tamayo et al., 2019).

9.2.3 Obesity
With the advance of knowledge of Andean grains’ chemical composition, their
use as an anti-​obese food is growing at the same rate at which obesity has become
a serious disease. World Health Organization (WHO, 2021) reports more than
two billion overweight people in the world and several comorbidities associated
with obesity such as diabetes, hypertension, cardiovascular diseases; and
recently, their correlation with the increase of death risk because of SARS-​CoV-​2
infection. Furthermore, levels of hormones such as leptin and ghrelin, with influ-
ence on metabolic homeostasis, are increased in obese subjects. Other biochem-
ical compounds, such as cytokines and interleukins released by adipose tissues,
seem to be linked to insulin resistance and atherosclerosis. Since obesity is a dis-
ease related to genomics, diets, lifestyle, and environmental factors, interest in
Andean grains as anti-​obesity activity foods is growing mainly due to their chem-
ical components such as fiber and flavonoids. An intervention study reported
that patients under Salba-​chia (Salvia hispanica L.) treatment lose more weight
than do the control patients. Participants on this diet maintained glycemic con-
trol and a reduction of risk factors (Vuksan et al., 2017). Reductions in C-​reactive
protein (hs-​CRP) and increased adiponectin concentrations were observed. At
present, mechanisms underlying the observed effects remain unknown, and
more studies are needed. Studies on chemical composition reveal that Salba-​
chia is lignin-​free with low available carbohydrate fraction and a rich source of
magnesium, calcium, and iron; and rich in antioxidants. Information on anti-​
nutrients is lacking. It can also be observed that not all literature studies agree.
Mithila and Khanum (2015) evaluated the effect of amaranth and quinoa in
rats. In this study, food intake and satiety were evaluated by monitoring bio-
chemical parameters. Leptin, produced by adipose cells, and ghrelin segregated
by cells of the gastrointestinal tract, are regulated by each other. In this work, a
decrease in ghrelin was observed, and an improvement in leptin and cholecysto-
kinin was detected, resulting in hunger and satiety changes. A faster response
to glucose blood level was observed in rats submitted to Andean grain diets.
The studies also linked the richness of the amino acids profile and the fiber of
Andean crops as quinoa with the observed results on reducing obesity (Mithila
and Khanum, 2015).
Recent research demonstrated the anti-​obesity properties of quinoa in obese
mice. Animals receiving a quinoa diet were revealed to have a reduced level of
plasma cholesterol, interleukin, and hepatic steatosis (Martínez-​Villaluenga
et al., 2020). Other research has focused on the mechanism underlying the
346 Contributions to Nutrition and Health

effects of this grain on increasing satiety with a concomitant decrease in leptin


(Tang and Tsao, 2017).
A clinical trial was conducted among Australian participants with different
body mass indexes (25 < BMI < 40), where subjects were segregated into over-
weight: obese type I; obese type II; and obese type III groups. Reduction in
triglycerides in serum and metabolic syndrome was correlated with the quinoa
dose. For the subjects that consumed 50 g/​day of quinoa, better results were
obtained, with a decrease of 70% in the prevalence of metabolic syndrome, than
those who received 25 g/​day (Navarro-​Perez et al., 2017).
Chaiittianan and Sutthanut (2017) demonstrated in vivo experiments that
purple corn has anti-​obesity effects, with action in the adipose cycle through
inhibition of adipogenesis and induction on lipolysis and apoptosis in adipocytes
(Chaiittianan et al., 2017). Several works suggest that polyphenols including
anthocyanins, quercetin, and phenolic acids and derivatives are the main ones
responsible for these effects. However, since weak evidence has been observed
in in vitro studies, authors recommend more in vivo investigations and clinical
trials to support the stimulating impact of polyphenols as anti-​obesity agents.
The inhibition of the adipose cycle has been demonstrated with quinoa (Teng
et al., 2020).
Berti et al. (2005) studied the effect on satiety of oat bread, buckwheat oat-
meal pasta, and quinoa compared to their wheat counterparts; they worked with
healthy young male volunteers. The results showed that the satiety efficiency
indices (SEI) for alternative crop foods were higher in contrast to traditional
cereal foods, suggesting that they can be exploited for their potential impact on
eating behavior, especially considering that they have good nutritional value and
are good sources of functional compounds (Berti et al., 2005).
In brief, it is clear that all studies enhance the need for future investigation
to clarify the gaps observed in present levels of knowledge. In spite of important
advances, more research is needed, either in vivo or in clinical trials, with quinoa
and other Andean grains.
Table 9.1 shows some clinical and preclinical studies related to the nutritional
effects and health benefits of Andean grains in non-​communicable diseases.

9.3 CANCER
Together with cancer, chronic diseases from several etiologies increase the risk
of cancer itself and death (Shu-​Ju et al., 2018). Cancer is a group of diseases that
can affect almost any organ or tissue of the body. It can be caused by different
risk factors and constitutes the second highest cause of death in many countries.
Lung, colorectum, liver, stomach, and breast cancers contribute to about 50% of
cancer deaths worldwide (Sung et al., 2021).
newgenrtpdf
TABLE 9.1 CLINICAL AND PRECLINICAL STUDIES ON THE NUTRITIONAL EFFECTS AND HEALTH BENEFITS
REGARDING ANDEAN GRAINS’ IMPACT ON NONCOMMUNICABLE DISEASES
Andean crop Dietary Compound Type of Trial Main Biological Effect Reference
Amaranth Starch In vivo; Human; female ↓ Plasma glucose; ↑Insulin Guerra-​Matias
(22–​34 years); BMI: 22 response and Arêas (2005)
Phytosterols, tocopherols, In vivo, animal studies Improve cardiovascular Chmelík,
tocotrienols risk profiles by modifying Šnejdrlová, and
cardiovascular risk factors Vrablík (2019)
such as cholesterol, diabetes
and hypertension
Amaranth; Phytochemicals phenolics, vitro enzyme assays and Lowering risk of type 2 Tang and Tsao
Quinoa betacyanins and lipophilic, the in vivo anti-​obesity diabetes by assessing the (2017)
fatty acids, tocopherols, and effect by using obese, antihyperglycemia and
carotenoids hyperglycemic mice model antihypertension
Amaranth, Chia Bioactive peptides In vitro and in vivo in Regulation of blood glucose Valenzuela
and Quinoa diabetic mice level, increase insulin Zamudio and
production or enhanced Segura Campos
insulin sensitivity (2020)
Beans; Whole Dietary fiber Meta-​analyses of Causal cardiometabolic Micha et al.
grains prospective studies or effects, CVD and diabetes (2017)
randomized clinical
trials

9.3
(continued)

Cancer
347
newgenrtpdf
348
Contributions to Nutrition and Health
TABLE 9.1 (CONTINUED)

Andean crop Dietary Compound Type of Trial Main Biological Effect Reference
Blue corn (Zea Phenolic compounds; In silico study Modulate the activity of Damián-​Medina
mays L.) and anthocyanins proteins involved in the et al. (2020)
black bean main pathways of type 2
(Phaseolus diabetes mellitus such as
vulgaris L.) insulin secretion, insulin
resistance and carbohydrate
absorption
Chia Benefits of Salba-​chia In vivo human 6-​month Beneficial role of Chia ↑ Vuksan et al.
(S. hispanica) consumption randomized, double-​blind, weight loss ↓ obesity related (2017)
parallel design study with risk factors, maintain
77 eligible participants glycemic control
Corn; Maize; Carbohydrates/​starch Clinical trials Improve postprandial Marventano
Sorghum; glucose and insulin response et al. (2017)
Whole Grains
Lupinus Protein hydrolysates In vitro -​ THP-​1-​derived ↓ proinflammatory Millán-​Linares
macrophage model cytokines; ↑ expression of et al. (2014)
anti-​inflammatory marker
genes
Lupin consuption; alkaloids In vivo; human; n=​30, males ↓ Plasma glucose; ↑Insulin Baldeón et al.
and females response (2012)
Lupin consuption, Case control study (Young Young healthy no changes; Fornasini et al.
conglutin-​γ and alkaloids healthy vs high glucose high glucose level volunteers (2012)
level volunteers) ↓ Plasma glucose; ↑Insulin
response
newgenrtpdf
Bioactive peptides In vivo human, ↓ plasma LDL cholesterol Bähr et al. (2013)
randomized, controlled significantly reduced after
crossover study in 33 4 wks
hypercholesterolemic
subjects
Purple corn Phenolic compounds; In vitro -​cell culture Retarded diabetes-​ Li et al. (2012)
(Z. mays L) anthocyanins HRMCs associated renal fibrosis and
mesangial inflammation
Phenolic composition; In silico study anti-​inflammatory and Zhang et al.
Anthocyanin anti-​diabetic properties, (2019)
improving insulin sensitivity
in insulin-​resistant
adipocytes
Quinoa Ecdysterone flavonoids In vivo -​diet-​induced in Significantly lowered fasting Graf et al. (2014)
obese, hyperglycemic mice blood glucose
(metabolic syndrome)
Gluten free In vivo, human healthy Effect on appetite control Berti et al. (2005)
voluntears no celiacs

9.3
Cancer
349
350 Contributions to Nutrition and Health

Throughout the years, research works have reported the influence of phys-
ical activity and food habits on the maintenance of health and the decrease of
chronic diseases and cancer (Sung et al,. 2021). Several in vitro studies show
that grains can be used as an anti-​inflammatory and antiproliferative tool for
controlling the disease, especially for colorectum cancers. In particular, quinoa,
amaranth, purple and red corn, and chia were associated with the decrease in
oxidative stress (Gawlik-​Dziki et al., 2013; Lee and Joo, 2018; Sabbione et al.,
2019), cancer cell viability and the inhibition of tumor cell growth factors with
an antiproliferative activity, cell necrosis, and apoptosis (Mazewski et al., 2017;
Sabbione et al., 2019; Srdić et al., 2020; Vilcacundo et al., 2018). Research results
have assigned these properties to those grains and plant products containing
bioactive peptides and antioxidant compounds, total phenols including
flavonoids, and anthocyanins, whose properties may be mainly responsible for
the described benefits in cells used for experimentation. Apart from intestinal
cancer cells, some research works also evidence the importance of quinoa, chia,
and lupinus for treating individuals with breast cancer and in subjects with a
multidrug resistance phenotype. The use of chia seed mucilage in cell culture
studies gives promising results on improving the treatment in these particular
individuals. The results found in cell culture studies evidence that the chia
oligosaccharides from mucilage and the antioxidant from quinoa, chia, and
lupinus constitute the main bioactive compounds responsible for decreasing
cell proliferation (Ayyash et al., 2019; Rosas-​Ramírez et al., 2017).
Purple corn, chia, and quinoa evidence a promissory impact on liver cancer.
In vivo studies focused on transgenic animal models evidence that Andean
cereals can influence carcinogenic mechanisms due to their chemical compos-
ition. Bioactive compounds, including peptides, antioxidants, and fatty acids,
have a beneficial immunometabolic effect, chemopreventive potential, and pro-
mote resistance to hepatocarcinogenesis under pro-​tumorigenic inflammation
(Laparra and Haros, 2019; Llopis et al., 2020; Yokohira et al., 2008). The protease
inhibitory activity was achieved by studying two groups of rats after receiving
the extracts enriched with serine-​type protease inhibitors (STPIs) from quinoa
and chia grains. Only STPIs from quinoa and chia seemed to promote the pro-
duction of inflammatory mediators contributing to the antitumoral macro-
phage phenotype activity. This functionality of STPIs can be used to control
hepatocellular cancer aggressiveness (Laparra and Haros, 2019). Due to its
content in anthocyanins and bioactive peptide composition, purple corn was
employed as a chemopreventive agent for prostate and mammary tumors from
non-​transgenic rat models. Consumption of purple corn seemed to induce apop-
tosis in mammary tumors by suppressing the Ras protein that controls signaling
pathways to malignant transformation of cells (Fukamachi et al., 2008). Purple
corn also evidences an inhibitory activity on prostate carcinogenesis through
9.4 Celiac Disease 351

the action of cyanidin-​3-​glucoside and pelargonidin-​3-​glucoside present as an


active compound (Long et al., 2013).
A randomized clinical trial on 75 patients diagnosed with duodenal ulcers
caused by Helicobacter pylori was conducted to evidence the antimicro-
bial activity and the inhibitory effect of anthocyanins, present in purple corn
and amaranth oil, on a Helicobacter pylori bacterium and its toxin control.
Helicobacter has been described as a risk factor to developing gastric cancer
by induction of oxidative stress in gastric mucosa. For the experimentation, two
groups were created: apart from the standard treatment, supplementation with
amaranth oil was given in the test group. Compared with the control group,
the duodenal peptic ulcer patients significantly reduced the accumulation of
the oxidative stress markers in gastric mucosa and improved the histological
parameters (Cherkas et al., 2018).
Table 9.2 shows the biological effect of Andean grains on tumor cell prolif-
eration and presents the in vitro and in vivo studies that evidence the impact of
several bioactive compounds on immunometabolism and as chemoprotective
agents.

9.4 CELIAC DISEASE
Celiac disease (CD) is an autoimmune digestive disorder triggered by gluten
ingestion in genetically predisposed individuals (Lebwohl et al., 2018). This
pathology involves activation of an inflammatory process in the digestive tract,
particularly the intestine, which produces hyperplasia of the crypts and villous
atrophy in the small intestine (Caio et al., 2019). CD presents various clinical
pictures that range from tangible symptoms such as diarrhea, weight loss, and
osteoporosis to more vague symptoms such as iron and folic acid deficiency,
arthralgia, fatigue, and abdominal discomfort that can manifest in childhood
and adulthood (Mezquita Cerezal et al., 2011). It is considered one of the most
common gluten-​related disorders in western society. The highest CD preva-
lence is found in people with family predisposition and is associated with auto-
immune diseases. CD is common worldwide, affecting 1 in 100 people in the
world’s population (Butterworth and Los, 2019). Destruction of enterocytes
caused by the immune system with atrophic intestinal epithelium results in
a decrease in the surface area for mineral absorption, which leads to vitamin
and protein deficiencies, consequently to general malnutrition, triggering
reduced body mass index. Although CD is a well characterized disease, it is
highly underdiagnosed, despite the serious consequences of prolonged gluten
ingestion, such as increased autoimmunity, refractory CD, and intestinal T-​
cell lymphoma (Mirijello et al., 2019). The main treatment for CD is lifetime
newgenrtpdf
352
Contributions to Nutrition and Health
TABLE 9.2 ANDEAN GRAINS’ BEHAVIOR IN DIFFERENT TUMOR CELLS
Andean crop Dietary Compound Type of Trial Main Biological Effect Reference
Amaranth Amaranth Oil In vivo (75 patients with Reduced HNE (oxidative stress marker) Cherkas et al.
Duodenal Ulcera by in gastric mucosa. Improvement of (2018)
H. pylori) histological parameters in Duodenal
Peptid Ulcera patients, reduced
manifestations of chronic oxidative stress
Antioxidant; total In vitro CT-​26 cell Amaranth antioxidation activity inhibits Lee and Joo
phenol, and total growth of CT-​26 cell Rectal cancer cells (2018)
flavonoid
Bioactive peptides In vitro HT-​29 colon Amaranth peptides reveals a potential Sabbione et al.
tumor cells antiproliferative activity over HT-​29 colon (2019)
tumor cells by induction of cell necrosis
and apoptosis
Chia Oligosaccharides In vitro -​colon (HCT-​15 Chia oligosaccharides from mucilage Rosas-​Ramírez
(S. hispanica) and HCT-​116), cervix modulate compounds of the et al. (2017)
(HeLa), breast (MCF7 multidrug resistance phenotype to
and MDA-​MB-​231) chemotherapeutic in breast cancer cells
carcinoma cell lines
Lupinus, Antioxidant activitie In Vitro -​Caco-​2 and Antiproliferative activities Fermented Ayyash et al.
Quinoa MCF-​7 cancer cell Quinoa had greater proliferative (2019)
lines of intestinal and inhibition against the breast cancer cell
mammary origin, line (MCF-​7) when compared with the
respectively colon cancer cell line
Purple and red Anthocyanin -​ In vitro human colon ↓ Colon cancer cell viability in a dose-​ Mazewski
corn Antioxidant activitie cancer cells, HT-​29 and dependent manner. Inhibition of colon et al. (2017)
HCT-​116 cancer cell proliferation
newgenrtpdf
Purple Corn Antioxidant; In vivo; 100 male F344 Liver carcinogenesis effect; Yokohira et al.
Anthocyanin rats chemopreventive potential against (2008)
liver preneoplastic lesion development.
Evidence of antioxidant power in vivo
Bioactive peptides In vivo heterozygous Prostate cancer chemoprevention; inhibit Long et al.
male transgenic rats for prostate carcinogenesis by cyanidin-​3-​ (2013)
adeno-​carcinoma of glucoside and pelarg-​ onidin-​3-​glucoside
prostate as active compounds from corn
Anthocyanin In vivo female c-​Ha-​ras Induces apoptosis in mammary tumors Fukamachi
-​ Antioxidant transgenic (Hras128, by decreasing ras protein levels; can be et al. (2008)
activities Tg) and non-​transgenic used for screening for chemopreventive
(non-​Tg) rats agents that act via suppressing the Ras
signaling pathway
Quinoa Chenopodium quinoa In vitro bioaccessibility Inhibitory effect on lipoxygenase Gawlik-​Dziki
Leaves Extract; and bioavailability activity, antioxidative, antiradical and et al. (2013)
phenolic content study; Rat prostate reducing power. Chemopreventive and
cancer AT-​2 and MAT-​ anticarcinogenic effect on oxidative
LyLu cells (the Dunning stress; prevention of Reactive oxygen
rat model), HTB-​140 species related deseases
and normal mouse 3T3
fibroblasts

9.4 Celiac Disease


Bioactive peptides In vitro -​Human ↓ oxidative stress-​associated diseases, Vilcacundo
colorectal cancer cell including cancer et al. (2018)
lines (Caco-​2, HT-​29,
and HCT-​116)

(continued)

353
newgenrtpdf
354
Contributions to Nutrition and Health
TABLE 9.2 (CONTINUED)

Andean crop Dietary Compound Type of Trial Main Biological Effect Reference
Quinoa Fatty acids and iron In vivo mice C57BL/​ Beneficial immunometabolic effects Llopis et al.
and Chia 6 hepatocarcinoma-​ of protease (serine-​type) from quinoa (2020)
(C. quinoa and developing mice and chia promote resistance to
S. hispanica) hepatocarcinogenesis
Bioactive peptides In vitro -​ Human-​like Immunometabolic effects of bioaccesible Srdić et al.
macrophages Cells polypeptides; prevention of diet-​ (2020)
(HB-​8902 associated innate immune imbalances
Serine-​type protease In vivo C57Bl/​6 mice STPIs from C. quinoa and S. hispanica Laparra and
inhibitors (STPIs) can be used to control Hepatocarcinoma Haros (2019)
aggressiveness. S. hispanica showed
positive effects, ↑ F4/​80+​ cells
normalizing the expression (mRNA) of
CD36 and the innate immune receptors.
9.4 Celiac Disease 355

adherence to a gluten-​free diet (GFD) from the time it is diagnosed. Gluten is a


protein found naturally in cereals such as wheat, barley, rye, oats, kamut, spelt,
and hybrid varieties, and products derived from these cereals such as flours and
starches or semolina. In general, GFD is low in nutrients, vitamins, minerals,
and dietary fiber (Saturni et al., 2010). Calcium and vitamin D, among other
components of importance for normal development of the human being, gener-
ally have poor absorption, which leads to increased hypocalcemia and vitamin
D deficiency (Krupa-​Kozak, 2014). This suggests that more emphasis should be
placed on nutritional quality in GFDs; the raw materials used to make these
dietary products are mainly cereals such as corn, rice, sorghum, millet, quinoa,
buckwheat, and amaranth. In general, the quality and availability of GFD
products on the market have continuously improved over the past decades.
However, many CD patients are still dissatisfied with these products, especially
due to low palatability, texture, mouthfeel, and nutritional quality compared to
their gluten-​containing counterparts. At present, achieving high production in
GF quality baked goods is a great challenge due to the absence of the techno-
logical properties that gluten confers on the dough (Rai et al., 2018). In this area,
Jagelaviciute and Cizeikiene (2021) evaluated the potential use of Lactobacillus
sanfranciscensis for the fermentation of chia, quinoa, and hemp flour in the
production of gluten-​free bread. The application of unfermented chia and hemp
flour increased the firmness and ageing rate of the bread, while the use of non-​
traditional hemp and quinoa sourdough reduced the bread’s ageing rate. In
many cases, chia, hemp, and quinoa flour increased the acceptability of gluten-​
free corn/​rice bread (Jagelaviciute and Cizeikiene, 2021).
Food scientists, nutritionists, biochemists, and the food industry actually
study the technological and nutritional properties of new grains to make GF
products as substitutes for wheat (Rai et al., 2018). At present, there is consider-
able interest in the consumption of alternative crops, such as quinoa, amaranth,
and kañiwua as potential ingredients for healthy food production and special
dietary use. These are the most consumed Andean grains in Latin America; they
possess high nutritional value and are formidable food alternatives for celiac
patients and/​or those suffering from gluten-​sensitivity (Jnawali et al., 2016).
Mezquita Cerezal et al. (2011) found that the mixtures of quinoa (Chenopodium
quinoa Willd) and lupine (Lupinus albus L) flours, with two traditional cereals,
corn (Zea mays L.) and rice (Oryza sativa L.), were suitable for the elaboration of
cakes and other sweet foods as a good alternative and a supplement for the nutri-
tion of children between 6 and 24 months who suffer from celiac disease. The
gluten-​free premix developed by Coronel et al. (2021) was based on buckwheat
flour supplemented with chia flour obtained as a by-​product of cold-​pressed oil
extraction, with and without the addition of xanthan gum, for the production of
gluten-​free bread. It had better nutritional characteristics (higher protein and
356 Contributions to Nutrition and Health

crude fiber content), higher antioxidant activity, and a significantly high con-
tent of essential polyunsaturated fatty acids when compared with the control
bread (commercial premix). Other work also found that the inclusion of quinoa
in alternative grain-​based products significantly increased the nutrient profile of
GF dietary products in protein, iron, calcium, and fiber content (Lee et al., 2009).
In vitro research works using quinoa and amaranth demonstrate that both
grains can be used as GF alternatives for celiac patients. The results evidenced
low content of protein cross-​reacting with anti-​gliadin antibodies either with
animals or serum immunoglobulins from celiac patients, demonstrating that
amaranth and quinoa are safe (Ballabio et al., 2011; Bergamo et al., 2011; Peñas
et al., 2014). Other studies also evaluate quinoa compliance with the Codex
Alimentarius conditions of GF products (gluten <20 mg/​Kg). Zevallos et al.
(2012) evaluated 15 quinoa cultivars from different origins and found that four
cultivars had quantifiable concentrations of celiac-​toxic epitopes but were
below the maximum allowed for GF foods. However, two cultivars stimulated
T-​cell lines to levels similar to gliadin and caused cytokine secretion from biopsy
samples cultured at levels comparable to those of gliadin. They concluded that
most quinoa is potentially suitable for CD patients but that further investigation
with in vivo studies is required.
In other studies, the addition of quinoa to the GFD of celiac patients was well
tolerated. There was a positive trend towards improvement in histological and
serological parameters, particularly a mild hypocholesterolemic effect. Overall,
these are the first clinical data suggesting that celiac patients can safely tolerate
50 g of quinoa daily for 6 weeks. However, more research is needed to deter-
mine the long-​term effects of quinoa consumption (Zevallos et al., 2012; Zevallos
et al., 2014). The evaluation of amaranth as a GF alternative for children was
tested by Bavykina et al. (2017), who found that it was well tolerated, and that
89.2% of parents of the 37 evaluated children did not notice allergic or dyspeptic
reactions.
Table 9.3 summarizes some in vivo and in vitro studies on the effect of
Andean grains on gluten related diseases.

9.5 PREBIOTIC EFFECT AND MICROBIOTA MODULATION


There is a consensus within the literature that many health aspects are influenced
by gut microbiota despite information dealing with the mechanism of action still
missing. Presently diet is considered one of the most important modifiers of gut
microbiota (Danneskiold-​Samsøe et al., 2019; Ferrario et al., 2017). Food pattern
as a direct mediator of gut microbiota offers a tremendous opportunity to pre-
vent noncommunicable disease (Spencer et al., 2019).
9.5 Prebiotic Effect and Microbiota Modulation 357

TABLE 9.3 THE EFFECT OF ANDEAN GRAINS ON GLUTEN RELATED


DISEASES
Andean
crop Type of Trial Main Biological Effect Reference
Amaranth In vitro reaction with Amaranth is safe for celiac; Ballabio
Serum from Celiac ↓ content of proteins cross-​ et al.
Subjects reacting with anti-​gliadin (2011)
antibodies
In vivo study, in 37 Amaranth tested in the study Bavykina
children from was well tolerated, allergic et al.
1–​17 years long term and dyspeptic reactions were (2017)
gluten free diet not noted. 89.2% of parents
commented positively on the
new gluten-​free amaranth
products
Amaranth; In vitro –​intestinal Amaranth and quinoa did Bergamo
Quinoa T-​cell lines, cultures not show any immune et al.
of duodenal explants cross-​reactivity confirming (2011)
from HLA-​DQ21 CD their safety in the diet of CD
patients and HLA-​DQ8 patients
transgenic mice for
signs of activation
Quinoa In vitro culture of Quinoa cultivars do not Zevallos
celiac duodenal biopsy present quantifiable amounts et al.
samples and exame of celiac-​toxic epitopes; Well (2012)
unknown cultivars tolerated among individuals
in human inume with CD
response in T cells
19 treated celiac Addition of quinoa to Zevallos
patients, with GFD of celiac patients, is et al.
evaluation of diet, well tolerated, improuve (2014)
serology, and histological and serological
gastrointestinal parameters; mild
parameters hypocholesterolemic effect
In vitro reaction with 11 quinoa demonstrate Peñas et al.
serum from Celiac safe for celiac; absence of (2014)
subjects gliadin proteins and no
binding affinity to animal
IgG or serum IgA from celiac
patients
358 Contributions to Nutrition and Health

The western diet, characterized by monosaccharides (glucose, galactose, and


fructose) and disaccharides (saccharose) consumption, causes microbiota and
health impairment changes. Scientific evidence suggests that regular sugar
and sugar substitute consumption, as well as other nocuous food ingredients,
can cause dysbiosis, intestinal inflammation, and even noncommunicable
diseases (Ferrario et al., 2017). The composition of gut microbiota has been
associated with dysregulation of blood pressure, chronic kidney disease, cardio-
vascular disease, age-​related health impairment, immunomodulation, allergy, or
asthma (Danneskiold-​Samsøe et al., 2019).
In contrast, the consumption of fiber, polyphenolic rich foods, saponins, and
prebiotics and probiotics has a positive effect on gut microbiota due to their
balance being restored (Del Hierro et al., 2020; Ferrario et al., 2017).
Studies on the characterization of Andean grains’ nutrient profile reveal a
high content of dietary fiber, polyunsaturated fatty acids, and a high-​quality
protein source. Furthermore, these grains contain an abundance of anti-​
inflammatory phytochemicals (Liu et al., 2018). These substances have shown a
range of immunomodulation effects due to the enhancement of Bifidobacterium
spp. and Lactobacillus spp., which positively regulate a wide range of physio-
logical functions. Andean grains (amaranth and quinoa) present a profile of fiber
polysaccharides, more similar to those found in fruits and vegetables than in
cereals (Zhu, 2020). In amaranth and quinoa grains, pectins, when compared with
other whole grains are quantitatively predominant, as recently demonstrated
(Ciudad-​Mulero et al., 2019; Martínez-​Villaluenga et al., 2020). Pectin is a complex
group of polysaccharides formed by D-​galacturonic acid monomers, linked by α-​
(1–​4) glycosidic bonds and branched regions primarily formed by various types
of neutral monosaccharides (mainly rhamnose, xylose, mannose, and arabinose)
(Ciudad-​Mulero et al., 2019). Soluble dietary fiber polysaccharides promote bac-
teria growth, increasing the production of short-​chain fatty acids (SCFAs) in the
gut. They improve immunity while decreasing excessively stimulated immune
responses through regulating gut microbiota (Zhu, 2020). Among carbohydrates,
resistant starch cannot be digested and absorbed in the small intestine reaching
the colon, and here it is slowly fermented by microorganisms to produce short-​
chain fatty acids (Lehmann and Robin, 2007). Andean grains present more than
20% of resistant starch, considering European Food Safety Authority (EFSA)
dietary guidelines (EFSA Panel on Dietetic Products, 2011) that impose regula-
tion of intestinal microbiota resulting in health benefits (Martínez-​Villaluenga
et al., 2020).
Polyphenols are the other category of bioactive compounds with a positive
effect on the modulation of gut microbiota. The profile of quinoa polyphenols
has been characterized by Tang et al. (2015). In quinoa grains, 23 different
phenolic compounds can be identified on red, white, and dark cultivars, where
9.5 Prebiotic Effect and Microbiota Modulation 359

vanillic and ferulic acids and their derivatives, and main flavonoids quercetin,
kaempferol, and their glycosides represent the main compounds. Betanin and
isobetanin are the predominant nitrogen-​containing phenols of betacyanins
(Tang et al., 2015).
Saponins are a class of glycosides whose aglycones can be either triterpenes or
helical spirostanes. Considered an anti-​nutrient, it reveals immunostimulatory,
hypocholesterolemic, antitumoral, anti-​inflammatory, antibacterial, antiviral,
antifungal, or antiparasitic activities (He et al., 2019). Saponins consist of a
hydrophobic aglycone backbone designated as sapogenin linked to the hydro-
philic sugar chain. Glycosides are poorly absorbed in the gastrointestinal tract
and transformed there into sapogenins through hydrolysis. Due to the chem-
ical properties of sapogenins caused by lack of a sugar chain, these compounds
exhibit high bioactive properties.
However, the contribution by the gastric passage to the digestion of saponins
remains unclear, and it is suggested that, at colon level, saponins are transformed
by the resident microbiota into sapogenins (Spencer et al., 2019). Recently,
Del Hierro et al. (2020) studied the transformation of saponin into sapogenins
using lentil and quinoa as triterpenoid and fenugreek as steroid saponins.
Quinoa reveals the highest production of sapogenins, namely oleanolic acid,
hederagenin, serjanic acid, and phytolaccagenic acid.
Bioactive components of Andean grains have increasingly interested the sci-
entific community due to their interaction with human microbiota and con-
sequent health benefits. These interactions determine the possible beneficial
effects of polysaccharides, polyphenols, and saponins provided by their pre-
biotic actions. Therefore, interaction between bioactive components of Andean
grains and human gut microbiota may impact human host health.
On the other hand, the effect of intestinal microbiota on the biotransform-
ation of these compounds into metabolites with health impact constitutes a
new topic of interest for scientific communities. Relevant reviews have been
published as well as research papers focusing on in vitro studies (Ugural and
Akyol, 2020; Valero-​Cases et al., 2020). However, our knowledge of the impact
of Andean grains on the human intestinal microbiota at different stages of life
is still limited. Among Andean grains, quinoa and amaranth received attention.
In Northern Europe, Australia, and New Zealand, interest in its cultivation has
grown since the 1990s. One of the principal reasons for interest in these grains
is their nutritional composition and the prebiotic effect of their components.
Table 9.4 presents the most relevant studies reporting recent findings on the
impact of Andean grains on the modulation of gut microbiota. The studies use
grains in several ways as foodstuffs where the prebiotic effect is evaluated in vivo
or in vitro in a starter culture where the microorganisms are inoculated with
Andean grains to enhance the prebiotic effect.
newgenrtpdf
360
Contributions to Nutrition and Health
TABLE 9.4 IMPACT OF DIETARY INTERVENTION OF ANDEAN GRAINS ON GUT MICROBIOTA
APPLYING IN VITRO AND IN VIVO STUDIES
Dietary Other Biological
Andean crop Compound Type of Trial Main Microbiota Effect Effects References
Amaranth; Prebiotic effect In vitro cultures ↑ Bifidobacterium and ↑ short-​chain fatty Gullón et al.
Quinoa with human faecal Atopobium or Bacteroides acids (SCFAs) (2016)
microbiota
Amaranth; Probiotic/​prebiotic In vitro and in ↓ pathogenic bacterias; ↓ inflammation Ugural and
Fermented effect vivo studies on ↑ Peptoclostridium, and colonic Akyol (2020)
quinoa fermented grains Prevotellaceae, Lactobacillus, damage clinical
Bifidobacterium, symptoms and
Enterococcus, and dysbiosis
Eubacteriaceae
Fermented Prebiotic ( fructo-​ In vitro simulator ↓ pathogenic bacterias No changes in Gullón et al.
Quinoa oligosaccharide); of the human Clostridium spp., Bacteroides short-​chain fatty (2016)
beverage Probiotic intestinal microbial spp., enterobacteria acids (SCFAs);
(Lactobacillus ecosystem and Enterococcus spp; ↑
casei Lc-​01); Lactobacillus spp. and
Synbiotic ( fructo-​ Bifidobacterium spp
oligosaccharide and
L. casei Lc-​01)
Quinoa Water soluble, In vitro human Preserve microbial diversity; ↑ short-​chain fatty Possomato-​
nondigestible fecal microbiome ↑ beneficial bacteria acids (SCFAs). Vieira et al.
polysaccharides fermentation and (2016)
and fibers in vivo studies in
obesity C57BL/​6J
mice.
newgenrtpdf
Serjanic acid; In vitro colonic ↑ Bifidobacterium spp. and Transformation Hierro et al.
saponins fermentation Lactobacillus spp of saponin-​rich (2020)
extracts of quinoa
by human gut
microbiota to
sapogenins
Polysaccharide In vivo, colitis ↓ abnormal expansion of Quinoa modified Liu et al.
induced by dextran phylum Proteobacteria, the dysbiosis (2018)
sodium sulfate, in and ↓ overgrowth of genera of intestinal
C57BL/​6 mice Escherichia/​Shigella and microbiota
Peptoclostridium and family
Lachnospiraceae

9.5 Prebiotic Effect and Microbiota Modulation


361
362 Contributions to Nutrition and Health

All works reinforce the need for more research to clarify the mechanisms
involved in gut microbiota modulation, both in microbiota species and in the
metabolites of bioactive compounds (Zhu, 2020).
Preclinical studies have illustrated that Andean crops can modulate gut
microbiota, suggesting quinoa and amaranth as possible alternatives to pharma-
ceutical drugs to improve dysbiosis (Ugural and Akyol, 2020).
Gullón et al. (2016) studied the prebiotic effect of quinoa and amaranth
through its incubation with fecal human inocula. The authors observed the
decrease of pH and production of SCFAs (acetate, propionate, and butyrate).
A significant difference was observed in these parameters. The presence of
propionate is very relevant. It is a glucogenerator that inhibits biosynthesis
of cholesterol and fatty acids in the liver and can reduce the risk of cardio-
vascular diseases. Furthermore, acetate is considered a lipogenic substance.
Regarding the microbiota profile, they observed changes in Bifidobacterium
spp., Lactobacillus-​Enterococcus, Atopobium, Bacteroides-​Prevotella, Clostridium
coccoides-​Eubacterium rectale, Faecalibacterium prausnitzii, and Roseburia
intestinalis. The authors concluded that Andean grains maintain a balance on
gut microbiota enhanced dysbiosis. All these bacterial groups affect the produc-
tion of SCFA and decrease pH in the gut. Furthermore, the study recommends
more preclinical intervention to elucidate the role of bioactive compounds,
both interaction of main nutrients as well as its bioavailability.
Recently Liu et al. (2018), in a comprehensive work, studied the effect of
quinoa to modulate gut microbiota in murine colitis model rats induced by dex-
tran sodium sulfate, a colitogen agent with anticoagulant properties. The robust
model was informative in demonstrating quinoa components’ effect on redu-
cing the disease index, epithelium damage, and microbiota profile. Escherichia/​
Shigella and Peptoclostridium expanded strongly due to colitis and dysbiosis.
On the other hand, in animals fed with quinoa, a decrease of pathogenic
family Lachnospiraceae was observed. Also, proteobacteria phylum inhibition,
designated as a microbial signature in dysbiosis, occurred in animals submitted
to a quinoa diet.
The in vitro assays presented by del Hierro et al. have studied the effect of
human fecal microbiota on saponin extract of quinoa, lentil, and fenugreek and
its biotransformation in sapogenin (del Hierro et al., 2020). The authors observed
an increase in Bifidobacterium spp and Lactobacillus spp after the addition of
quinoa extracts to the feces of the volunteers. This is promising since, as far as
we know, it is the first study using saponin quinoa extracts to detect a modular
effect in human gut microbiota. Furthermore, a decrease in Enterococcus spp
was observed. These findings contribute to understanding the communication
routes between quinoa and microbiota and their association with the impact on
health of Andean grains.
9.5 Conclusions 363

A study evaluated the effect of prebiotic, probiotic, and symbiotic beverage


formulations containing aqueous extracts of soy and quinoa, using a dynamic
model of human gut simulator of the human intestinal microbiota ecosystem
(Bianchi et al., 2014). This model allows for monitoring the effect of formulations
in different colon compartments. In the work, the effect of a beverage with
soy and quinoa water extracts on colon compartments –​ascending, trans-
verse, and descending –​with a combination of prebiotic and probiotic agents
acting synergistically was tested. The beverage conferred significant increase
of Lactobacillus spp. and Bifidobacterium spp., and reduced Clostridium spp.,
Bacteroides spp., enterobacteria, and Enterococcus spp. The bacterial modu-
lation promoted by this type of beverage also contributes to understanding
how to enhance gut health by expanding the beneficial bacteria and reducing
pathogenics.

9.5 CONCLUSIONS
This chapter has evaluated the research literature concerning the dietary intake
of crops of Latin-​American origin (corn, quinoa, amaranth, chia, black turtle
bean, tarwi, kañiwa, and sacha inchi) associated with several health effects.
Based on the data presented in scientific studies, especially in vivo or in vitro,
that correlates health with nutritional and/​or phytochemical compounds, it is
possible to conclude that these crops are revealed to be influential on human
health. Impact on noncommunicable diseases such as diabetes mellitus; on
cardiovascular risk factors such as on the reduction of cholesterol and arterial
hypertension; on obesity; on different types of cancer cells, as well as on the
fields of gastrointestinal health with a relevant prebiotic effect and microbiota
modulation and the reduction of the inflammatory response on celiac disease,
reveals the importance of these crops in improving human health.
While Andean and Mesoamerican crops exhibit various protective effects in
the context of cardiovascular diseases, many key questions remain unanswered
considering the limited number of randomized clinical trials or case-​control
studies, and the limited sample size of some studies. These facts show that
extensive prospective epidemiological studies need to be conducted. Future
studies should include a well-​defined population in order to address the dose-​
response relationship to assess more directly the association of each Andean and
Mesoamerican crop with their respective health impact, especially considering
increased consumer demands toward healthy foods.
364 Contributions to Nutrition and Health

ACKNOWLEDGMENTS
This work was supported by grants laValSe-​Food-​CYTED (Ref. 119RT0567);
Consejo Nacional de Investigaciones Científicas y Técnicas (CONICET), and
Universidad Nacional de Jujuy, Argentina.

REFERENCES
Ayyash, M., S.K. Johnson, S.Q. Liu, N. Mesmari, S. Dahmani, A.S. Al Dhaheri, and J.
Kizhakkayil. 2019. In vitro investigation of bioactivities of solid-​state fermented
lupin, quinoa and wheat using lactobacillus spp. Food Chemistry 275: 50–​58.
Bähr M, A. Fechner, J. Krämer, M. Kiehntopf, and G. Jahreis. 2013. Lupin protein
positively affects plasma LDL cholesterol and LDL:HDL cholesterol ratio in
hypercholesterolemic adults after four weeks of supplementation: A randomized,
controlled crossover study. Nutrition Journal 12: 107.
Baldeón, M.E., J. Castro, E. Villacrés, L. Narváez, and M. Fornasini. 2012. Efecto
hipoglicemiante de lupinus mutabilis cocinado y sus alcaloides en sujetos con dia-
betes tipo-​2. Nutricion hospitalaria 27(4): 1261–​1266.
Ballabio, C., F. Uberti, C. Di Lorenzo, A. Brandolini, E. Penas, and P. Restani. 2011.
Biochemical and immunochemical characterization of different varieties of amar-
anth (amaranthus l. ssp.) as a safe ingredient for gluten-​free products. Journal of
Agricultural and Food Chemistry 59(24): 12969–​12974.
Bavykina, I.A., A.A. Zvyagin, L.A. Miroshnichenko, K.Y. Gusev, and I.M. Zharkova. 2017.
Efficient products from amaranth in gluten-​free nutrition of children with gluten
intolerance. Voprosy Pitaniia 86(2): 91–​99.
Bazzano, L.A., A.M. Thompson, M.T. Tees, C.H. Nguyen, and D.M. Winham. 2011. Non-​
soy legume consumption lowers cholesterol levels: a meta-​analysis of randomized
controlled trials. Nutrition, Metabolism and Cardiovascular Diseases 21(2): 94–​103.
Bellou, V., L. Belbasis, I. Tzoulaki, and E. Evangelou. 2018. Risk factors for type 2 dia-
betes mellitus: An exposure-​wide umbrella review of meta-​analyses. PLoS ONE
13(3): 1–​27.
Bergamo, P., F. Maurano, G. Mazzarella, G. Iaquinto, I. Vocca, A.R. Rivelli, E. De Falco, C.
Gianfrani, and M. Rossi. 2011. Immunological evaluation of the alcohol-​soluble
protein fraction from gluten-​free grains in relation to celiac disease. Molecular
Nutrition and Food Research 55(8): 1266–​1270.
Berti, C., P. Riso, A. Brusamolino, and M. Porrini. 2005. Effect on appetite control of
minor cereal and pseudocereal products. British Journal of Nutrition 94(5): 850–​
858. www.cambri​dge.org/​core/​prod​uct/​ide​ntif​i er/​S00071​1450​5002​564/​type/​jour​
nal_​arti​cle.
Bianchi, F., E.A. Rossi, I.K. Sakamoto, M.A.T. Adorno, T. Van de Wiele, and K. Sivieri. 2014.
Beneficial effects of fermented vegetal beverages on human gastrointestinal micro-
bial ecosystem in a simulator. Food Research International 64: 43–​52.
Burton-​Freeman, B., M. Brzeziński, E. Park, A. Sandhu, D. Xiao, and I. Edirisinghe. 2019.
A selective role of dietary anthocyanins and flavan-​3-​ols in reducing the risk of type
2 diabetes mellitus: a review of recent evidence. Nutrients 11(4): 1–​16 .
References 365

Butterworth, J., and L. Los. 2019. Coeliac disease. Medicine 47(5): 314–​319. https://​lin​king​
hub.elsev​ier.com/​retri​eve/​pii/​S13573​0391​9300​43X.
Caio, G., U. Volta, A. Sapone, D.A. Leffler, R. De Giorgio, C. Catassi, and A. Fasano. 2019.
Celiac disease: A comprehensive current review. BMC Medicine 1(17): 1–​20.
Chaiittianan, R., K. Sutthanut, and A. Rattanathongkom. 2017. Purple corn silk: A poten-
tial anti-​obesity agent with inhibition on adipogenesis and induction on lipolysis
and apoptosis in adipocytes. Journal of Ethnopharmacology 201: 9–​16.
Cherkas, A., K. Zarkovic, A. Cipak Gasparovic, M. Jaganjac, L. Milkovic, O. Abrahamovych,
O. Yatskevych, G. Waeg, O. Yelisyeyeva, and N. Zarkovic. 2018. Amaranth oil
reduces accumulation of 4-​hydroxynonenal-​histidine adducts in gastric mucosa
and improves heart rate variability in duodenal peptic ulcer patients undergoing
helicobacter pylori eradication. Free Radical Research 52(2): 135–​149.
Chmelík, Z., M. Šnejdrlová, and M. Vrablík. 2019. Amaranth as a potential dietary adjunct
of lifestyle modification to improve cardiovascular risk profile. Nutrition Research
72: 36–​45.
Ciudad-​ Mulero, M., V. Fernández-​ Ruiz, M.C. Matallana-​ González, and P. Morales.
2019. Dietary fiber sources and human benefits: The case study of cereal and
pseudocereals. Advances in Food and Nutrition Research, 90: 83–​134. DOI:Org/​
10.1016/​bs.afnr.2019.02.002.
Coronel E.B., E.N. Guiotto, M.C. Aspiroz, M.C. Tomás, S. M Nolasco, and M. I.
Capitani. 2021. Development of gluten-​free premixes with buckwheat and chia
flours: Application in a bread product. Lebensmittel-​Wissenschaft Technologie
(LWT) 141: 110916.
Damián-​Medina, K., Y. Salinas-​Moreno, D. Milenkovic, L. Figueroa-​Yáñez, E. Marino-​
Marmolejo, I. Higuera-​Ciapara, A. Vallejo-​Cardona, and E. Lugo-​Cervantes. 2020. In
silico analysis of antidiabetic potential of phenolic compounds from blue corn (Zea
mays l.) and black bean (Phaseolus vulgaris l.). Heliyon 6(3): 1–​13.
Danneskiold-​Samsøe, N.B., H. Dias de Freitas Queiroz Barros, R. Santos, J.L. Bicas, C.B.B.
Cazarin, L. Madsen, K. Kristiansen, G.M. Pastore, S. Brix, and M.R. Maróstica Júnior.
2019. Interplay between food and gut microbiota in health and disease. Food
Research International 115: 23–​31. .
Del Hierro, J.N., C. Cueva, A. Tamargo, E. Núñez-​Gómez, M.V. Moreno-​Arribas, G. Reglero,
and D. Martin. 2020. In vitro colonic fermentation of saponin-​rich extracts from
quinoa, lentil, and fenugreek. Effect on sapogenins yield and human gut micro-
biota. Journal of Agricultural and Food Chemistry 68 (1): 106–​116.
EFSA (European Food Safety Authority). 2011. Panel on Dietetic Products, N. and
A. Scientific opinion on the substantiation of health claims related to resistant starch
and reduction of post-​prandial glycaemic responses (ID 681), “Digestive Health
Benefits” (ID 682) and “Favours a Normal Colon Metabolism” (ID 783) Pursuant to
Article 13. EFSA Journal 9(4): 1–​17.
Ferrario, C., R. Statello, L. Carnevali, L. Mancabelli, C. Milani, M. Mangifesta, and S.
Duranti. 2017. How to feed the mammalian gut microbiota: Bacterial and meta-
bolic modulation by dietary fibers. Frontiers in Microbiology 8: 1–​11.
Fornasini, M., J. Castro, E. Villacres, L. Narvaez, M.P. Villamar, and M.E. Baldeon. 2012.
Efecto hipoglicemiante de lupinus mutabilis en voluntarios sanos y sujetos con
disglicemia. Nutricion Hospitalaria 27(2): 425–​433.
366 Contributions to Nutrition and Health

Frieden, T.R., and M.G. Jaffe. 2018. Saving 100 million lives by improving global treatment
of hypertension and reducing cardiovascular disease risk factors. Journal of Clinical
Hypertension 20(2): 208–​211.
Fukamachi, K., T. Imada, Y. Ohshima, J. Xu, and H. Tsuda. 2008. Purple corn color
suppresses ras protein level and inhibits 7,12-​dimethylbenz[a]‌anthracene-​induced
mammary carcinogenesis in the rat. Cancer Science 99(9): 1841–​1846.
Gawlik-​Dziki, U., M. Świeca, M. Sułkowski, D. Dziki, B. Baraniak, and J. Czyz. 2013.
Antioxidant and anticancer activities of chenopodium quinoa leaves extracts -​in
vitro study. Food and Chemical Toxicology 57: 154–​160.
Graf L.B., A. Poulev, P. Kuhn, M.H. Grace, M.A. Lila, and I.Raskin. 2014. Quinoa seeds
leach phytoecdysteroids and other compounds with anti-​diabetic properties. Food
Chemistry 163: 178–​185.
Guerra-​Matias, A.C., and J.A.G. Arêas. 2005. Glycemic and insulinemic responses in
women consuming extruded amaranth (Amaranthus cruentus L). Nutrition Research
25(9): 815–​822.
Gullón, B., P. Gullón, F.K. Tavaria, and R. Yáñez. 2016. Assessment of the prebiotic effect
of quinoa and amaranth in the human intestinal ecosystem. Food & Function
7(9): 3782–​3788.
He, Y., Z. Hu, A. Li, Z. Zhu, N. Yang, Z. Ying, J. He, C. Wang, S. Yin, and S. Cheng. 2019.
Recent advances in biotransformation of saponins. Molecules 24 (13): 1–​23.
Jagelaviciute, J., and D. Cizeikiene. 2021. The influence of non-​traditional sourdough made
with quinoa, hemp and chia flour on the characteristics of gluten-​free maize/​rice
bread. LWT –​ Food Science and Technology 137: 110457. https://​lin​king​hub.elsev​ier.
com/​retri​eve/​pii/​S00236​4382​0314​456.
Jiang, Y.T., J.Y. Zhang, Y.S. Liu, Q. Chang, Y.H. Zhao, and Q.J. Wu. 2020. Relationship
between legume consumption and metabolic syndrome: A systematic review and
meta-​analysis of observational studies. Nutrition, Metabolism and Cardiovascular
Diseases 30(3): 384–​392.
Jnawali, P., V. Kumar, and T.B. Anwar. 2016. Celiac disease: Overview and considerations
for development of gluten-​free foods. Food Science and Human Wellness 5: 169–​176.
Krupa-​Kozak, U. 2014. Pathologic bone alterations in celiac disease: Etiology, epidemi-
ology, and treatment. Nutrition 30(1) (January): 16–​24. https://​lin​king​hub.elsev​ier.
com/​retri​eve/​pii/​S08999​0071​3002​89X.
Laparra, J.M., and C.M. Haros. 2019. Plant seed protease inhibitors differentially affect
innate immunity in a tumor microenvironment to control hepatocarcinoma. Food
and Function 10(7): 4210–​4219.
Lebwohl, B., D.S. Sanders, and P.H.R. Green. 2018. Coeliac disease. The Lancet
391(10115): 70–​81.
Lee, A.R., D.L. Ng, E. Dave, E.J. Ciaccio, and P.H.R. Green. 2009. The effect of substituting
alternative grains in the diet on the nutritional profile of the gluten-​free diet. Journal
of Human Nutrition and Dietetics 22(4): 359–​363.
Lee, H., and N. Joo. 2018. Antioxidative properties of amaranth cauline leaf and suppressive
effect against ct-​26 cell proliferation of the sausage containing the leaf. Korean
Journal for Food Science of Animal Resources 38(3): 570–​579 .
Lehmann, U., and F. Robin. 2007. Slowly digestible starch -​its structure and health
implications: A review. Trends in Food Science and Technology 18(7): 346–​355.
References 367

Li, J., S.S. Lim, J.Y. Lee, J.K. Kim, S.W. Kang, J.L. Kim, and Y.H. Kang. 2012. Purple corn
anthocyanins dampened high-​glucose-​induced mesangial fibrosis and inflamma-
tion: Possible renoprotective role in diabetic nephropathy. Journal of Nutritional
Biochemistry 23(4): 320–​331.
Liu, W., Y. Zhang, B. Qiu, S. Fan, H. Ding, and Z. Liu. 2018. Quinoa whole grain diet
compromises the changes of gut microbiota and colonic colitis induced by dextran
sulfate sodium in C57bl/​6 mice. Scientific Reports 8 (1): 1–​9.
Llopis, J.M.L., D. Brown, and B. Saiz. 2020. Chenopodium quinoa and salvia hispanica pro-
vide immunonutritional agonists to ameliorate hepatocarcinoma severity under a
high-​fat diet. Nutrients 12(7): 1–​15.
Long, N., S. Suzuki, S. Sato, A. Naiki-​Ito, K. Sakatani, T. Shirai, and S. Takahashi. 2013.
Purple corn color inhibition of prostate carcinogenesis by targeting cell growth
pathways. Cancer Science 104(3): 298–​303.
López, D.N., M. Galante, G. Raimundo, D. Spelzini, and V. Boeris. 2019. Functional prop-
erties of amaranth, quinoa and chia proteins and the biological activities of their
hydrolyzates. Food Research International 116: 419–​429.
Martínez-​ Villaluenga, C., E. Peñas, and B. Hernández-​ Ledesma. 2020. Pseudocereal
grains: Nutritional value, health benefits and current applications for the develop-
ment of gluten-​free foods. Food and Chemical Toxicology 137: 111178.
Marventano, S., C. Vetrani, M. Vitale, J. Godos, G. Riccardi, and G. Grosso. 2017. Whole
grain intake and glycaemic control in healthy subjects: A systematic review and
meta-​analysis of randomized controlled trials. Nutrients 9(7): 769.
Mazewski, C., K. Liang, and E. Gonzalez de Mejia. 2017. Inhibitory potential of anthocyanin-​
rich purple and red corn extracts on human colorectal cancer cell proliferation in
vitro. Journal of Functional Foods 34: 254–​265.
Mezquita Cerezal, P., V. Gatica Urtuvia, V. Quintanilla Ramírez, and R. Zavala Arcos.
2011. Desarrollo de producto sobre la base de harinas de cereales y leguminosa
para niños celíacos entre 6 y 24 meses; II: Propiedades de Las Mezclas. Nutricion
Hospitalaria 26(1): 161–​169.
Micha, R., M.L. Shulkin, J.L. Peñalvo, S. Khatibzadeh, G.M. Singh, M. Rao, S. Fahimi, J.
Powles, and D. Mozaffarian. 2017. Etiologic effects and optimal intakes of foods and
nutrients for risk of cardiovascular diseases and diabetes: Systematic reviews and
meta-​analyses from the nutrition and chronic diseases expert group (NutriCoDE).
PLoS ONE 12(4): 1–​25.
Millán-​Linares, M. del C., B. Bermúdez, M. del M. Yust, F. Millán, and J. Pedroche. 2014.
Anti-​inflammatory activity of lupine (Lupinus angustifolius l.) protein hydrolysates
in thp-​1-​derived macrophages. Journal of Functional Foods 8(1):224–​233.
Mirijello, A., C. D’Angelo, S. De Cosmo, A. Gasbarrini, and G. Addolorato. 2019.
Management of celiac disease in daily clinical practice: Do not forget depression!
European Journal of Internal Medicine 62(17).
Mithila M. V., and F. Khanum. 2015. Effectual comparison of quinoa and amaranth
supplemented diets in controlling appetite; A biochemical study in rats. Journal of
Food Sience and Technology 52(10): 6735–​6741.
Motta, C., I. Castanheira, G.B. Gonzales, I. Delgado, D. Torres, M. Santos, and A.S. Matos.
2019. Impact of cooking methods and malting on amino acids content in amaranth,
buckwheat and quinoa. Journal of Food Composition and Analysis 76: 58–​65.
368 Contributions to Nutrition and Health

Mudryj, A.N., N. Yu, and H.M. Aukema. 2014. Nutritional and health benefits of pulses.
Applied Physiology, Nutrition, and Metabolism 39(11): 1197–​1204.
Navarro-​Perez D., J. Radcliffe, A. Tierney, and M. Jois. 2017. Quinoa seed lowers serum
triglycerides in overweight and obese subjects: A dose-​response randomized con-
trolled clinical trial. Current Developments in Nutrition 1(9): 1–​9.
Neuenschwander, M., A. Ballon, K.S. Weber, T. Norat, D. Aune, L. Schwingshackl, and
S. Schlesinger. 2019. Role of diet in type 2 diabetes incidence: Umbrella review
of meta-​analyses of prospective observational studies. The BMJ Clinical Research
366: 1–​18.
Orona-​Tamayo, D., M.E. Valverde, and O. Paredes-​López. 2019. Bioactive peptides from
selected Latin-​American food crops –​A nutraceutical and molecular approach.
Critical Reviews in Food Science and Nutrition 59(12): 1949–​1975.
Peñas, E., F. Uberti, C. di Lorenzo, C. Ballabio, A. Brandolini, and P. Restani. 2014.
Biochemical and immunochemical evidences supporting the inclusion of quinoa
(Chenopodium quinoa Willd.) as a gluten-​free ingredient. Plant Foods for Human
Nutrition 69(4): 297–​303.
Possomato-​Vieira, José S. and R.A.K. Khalil. (2016). Prebiotics from acorn and sago pre-
vent high-​fat diet-​induced insulin resistance via microbiome-​gut-​brain axis modu-
lation. Physiology & Behavior 176(12): 139–​148.
Rai, S., A. Kaur, and C.S. Chopra. 2018. Gluten-​free products for celiac susceptible people.
Frontiers in Nutrition 5(116): 1–​23.
Rosas-​Ramírez, D.G., M. Fragoso-​Serrano, S. Escandón-​Rivera, A.L. Vargas-​Ramírez, J.P.
Reyes-​ Grajeda, and M. Soriano-​ García. 2017. Resistance-​ modifying activity in
vinblastine-​resistant human breast cancer cells by oligosaccharides obtained from
mucilage of chia seeds (Salvia Hispanica). Phytotherapy Research 31(6): 906–​914.
Sabbione, A.C., F.O. Ogutu, A. Scilingo, M. Zhang, M.C. Añón, and T.H. Mu. 2019.
Antiproliferative effect of amaranth proteins and peptides on ht-​29 human colon
tumor cell line. Plant Foods for Human Nutrition 74(1): 107–​114.
Sancho, R.A.S., and G.M. Pastore. 2012. Evaluation of the effects of anthocyanins in type 2
diabetes. Food Research International 46(1): 378–​386.
Saturni, L., G. Ferretti, and T. Bacchetti. 2010. The gluten-​free diet: Safety and nutritional
quality. Nutrients 2(1): 16–​34.
Schwingshackl, L., G. Hoffmann, A.M. Lampousi, S. Knüppel, K. Iqbal, C. Schwedhelm,
A. Bechthold, S. Schlesinger, and H. Boeing. 2017. Food groups and risk of type 2
diabetes mellitus: A systematic review and meta-​analysis of prospective studies.
European Journal of Epidemiology 32(5):362–​375.
Shu-​Ju T., W. Chih-​Wei, P. Kuang-​Tse, W. Yi-​Cheng, and W. Chen-​Te. 2018. Localized thin-​
section CT with radiomics feature extraction and machine learning to classify
early-​detected pulmonary nodules from lung cancer screening. Physics in Medicine
& Biology 63(6): 065005.
Spencer, S.P., G.K. Fragiadakis, and J.L. Sonnenburg. 2019. Pursuing human-​relevant gut
microbiota-​immune interactions. Immunity 51(2): 225–​239.
Srdić, M., I. Ovčina, B. Fotschki, C.M. Haros, and J.M. Laparra Llopis. 2020. C. Quinoa and
S. Hispanica L. seeds provide immunonutritional agonists to selectively polarize
macrophages. Cells 9(3): 593.
References 369

Sung, H., J. Ferlay, R.L. Siegel, M. Laversanne, I. Soerjomataram, A. Jemal, and F. Bray.
2021. Global cancer statistics 2020: GLOBOCAN estimates of incidence and mor-
tality worldwide for 36 cancers in 185 countries. CA: A Cancer Journal for Clinicians
71(3): 209–​249. DOI.org/​10.3322/​caac.21660.
Tang, Y., and R. Tsao. 2017. Phytochemicals in quinoa and amaranth grains and their
antioxidant, anti-​inflammatory, and potential health beneficial effects: A review.
Molecular Nutrition and Food Research 61(7): 1–​16.
Tang, Y., X. Li, B. Zhang, P.X. Chen, R. Liu, and R. Tsao. 2015. Characterization of phenolics,
betanins and antioxidant activities in seeds of three chenopodium quinoa willd.
genotypes. Food Chemistry 166: 380–​388.
Teng, C., Z. Shi, Y. Yao, and G. Ren. 2020. Structural characterization of quinoa poly-
saccharide and its inhibitory effects on 3t3-​l1 adipocyte differentiation. Foods
9(10): 1511.
Ugural, A., and A. Akyol. 2020. Can pseudocereals modulate microbiota by functioning as
probiotics or prebiotics? Critical Reviews in Food Science and Nutrition. https://​doi.
org/​10.1080/​10408​398.2020.1846​493.
Valenzuela Zamudio, F., and M.R. Segura Campos. 2020. Amaranth, quinoa and chia
bioactive peptides: A comprehensive review on three ancient grains and their
potential role in management and prevention of type 2 diabetes. Critical Reviews
in Food Science and Nutrition www.tand​fonl​ine.com/​doi/​full/​10.1080/​10408​
398.2020.1857​683.
Valero-​Cases, E., D. Cerdá-​Bernad, J.J. Pastor, and M.J. Frutos. 2020. Non-​dairy fermented
beverages as potential carriers to ensure probiotics, prebiotics, and bioactive
compounds arrival to the gut and their health benefits. Nutrients 12(6). https://​doi.
org/​10.3390/​nu1​2061​666.
Vilcacundo, R., B. Miralles, W. Carrillo, and B. Hernández-​ Ledesma. 2018. In vitro
chemopreventive properties of peptides released from quinoa (Chenopodium
quinoa Willd.) Protein under simulated gastrointestinal digestion. Food Research
International 105: 403–​411.
Vuksan, V., A.L. Jenkins, C. Brissette, L. Choleva, E. Jovanovski, A.L. Gibbs, R.P. Bazinet.
2017. Salba-​Chia (Salvia Hispanica L.) in the treatment of overweight and obese
patients with type 2 diabetes: A double-​blind randomized controlled trial. Nutrition,
Metabolism and Cardiovascular Diseases 27(2): 138–​146.
WHO (World Health Organization). 2021. Obesity and Overweight. www.who.int/​news-​
room/​fact-​she​ets/​det​ail/​obes​ity-​and-​ove​rwei​ght.
Yokohira, M., K. Yamakawa, K. Saoo, Y. Matsuda, K. Hosokawa, N. Hashimoto, T. Kuno,
and K. Imaida. 2008. Antioxidant effects of flavonoids used as food additives (purple
corn color, enzymatically modified isoquercitrin, and isoquercitrin) on liver car-
cinogenesis in a rat medium-​term bioassay. Journal of Food Science 73(7): 561–​568.
Zevallos, V.F., H.J. Ellis, T. Suligoj, L.I. Herencia, and P.J. Ciclitira. 2012. Variable activation
of immune response by quinoa (Chenopodium quinoa Willd .) prolamins in celiac
disease. The American Journal of Clinical Nutrition 96: 337–​344.
Zevallos, V.F., L.I. Herencia, and P.J. Ciclitira. 2014. Gastrointestinal effects of eating
quinoa (Chenopodium quinoa Willd.) in celiac patients. The American Journal of
Gastroenterology 109(2): 270–​278.
370 Contributions to Nutrition and Health

Zhang, Q., E. Gonzalez de Mejia, D. Luna-​Vital, T. Tao, S. Chandrasekaran, L. Chatham,


J. Juvik, V. Singh, and D. Kumar. 2019. Relationship of phenolic composition of
selected purple maize (Zea Mays L.) genotypes with their anti-​inflammatory, anti-​
adipogenic and anti-​diabetic potential. Food Chemistry 289: 739–​750.
Zhu, F. 2020. Dietary fiber polysaccharides of amaranth, buckwheat and quinoa
grains: A review of chemical structure, biological functions and food uses.
Carbohydrate Polymers 248: 116819. https://​lin​king​hub.elsev​ier.com/​retri​eve/​pii/​
S01448​6172​0309​929.
Chapter 10

Ingredients of High Nutritional


Value Obtained from Latin-​
American Crops through
Biotechnology
Manuel Oscar Lobo, Ana Laura Mosso, María Dolores Jiménez,
and Norma Sammán
Universidad Nacional de Jujuy and Consejo
Nacional de Investigaciones Científicas y Técnicas
(CONICET), San Salvador de Jujuy, Argentina

CONTENTS
10.1 Introduction 372
10.2 Fermentation 373
10.2.1 Lactic Fermentation of Latin-​American Grains 373
10.2.2 Production of Vitamin and Antioxidant Compounds
through Lactic Fermentation 375
10.2.3 Reducing Antinutritional Factors through Fermentation 376
10.2.4 Traditional Beverages through Maize Fermentation 377
10.2.5 Impact of Fermentation on Matrix Structural Changes 377
10.3 Germination 378
10.3.1 Germination Conditions 378
10.3.2 Changes in Nutritional Profile during Germination 378
10.3.2.1 Protein Content, Amino Acids Profile and
Protein Digestibility 378
10.3.2.2 Lipid Content and Fatty Acids’ Profile 381
10.3.2.3 Carbohydrates Profile 382

DOI: 10.1201/9781003088424-10 371


372 Nutritional Value Obtained from Latin-American Crops

10.3.3 Antioxidant Content and Antioxidant Capacity 382


10.3.4 Antinutrients and Mineral Bioavailability 383
10.3.5 Technological, Thermal and Sensory Changes during
Germination 384
10.3.5.1 Rheological and Textural Characteristics 384
10.3.5.2 Thermal Behaviors 385
10.3.5.3 Sensory Characteristics 385
10.4 Bioactive Peptide Production 386
10.4.1 Enzymatic Hydrolysis 386
10.4.2 Physiological Functions 388
10.4.2.1 Antioxidant Activity 388
10.4.2.2 Cytotoxicity 389
10.4.2.3 Anti-​Inflammatory Activity 389
10.4.2.4 Antihemolytic Activity 390
10.4.2.5 Antimicrobial Activity 390
10.4.2.6 Antiadipogenic Activity 391
10.4.2.7 Antidiabetic Activity 391
10.4.2.8 Antithrombotic Activity 391
10.4.2.9 Anticancer Activity 392
10.4.2.10 Immunomodulatory Activity 392
10.4.3 In Silico Approach and Molecular Docking 392
10.5 Conclusions 393
Acknowledgments 394
References 394

10.1 INTRODUCTION
Andean and Mesoamerican crops are characterized by a wide spectrum of
interesting nutritional and functional compounds, such as gluten-​free proteins
of good biological quality, micronutrients and antioxidants. For this reason,
quinoa, amaranth, chia, Andean corn, black bean, tarwi and sacha inchi, are
suitable options for reinforcing diets that are deficient in proteins, essential
nutrients and functional components. Furthermore, biotechnologies such as
fermentation, germination and protein hydrolysis can improve the nutritional
and functional properties of these crops, by implementing simple, economical
and easily adjustable processes, which may lead to highly nutritional value foods
and ingredients.
In this chapter, the terms seeds and grains will be used indiscriminately.
10.2 Fermentation 373

10.2 FERMENTATION
Fermentation is an economically and sustainable method of producing and
preserving foods applied since ancient times. The process is determined by
the metabolic action of microorganisms (bacteria, fungi or yeasts) on different
substrates, in order to obtain favorable transformations. In this sense, tradition-
ally, microorganisms naturally present in the source of origin to improve and
control natural processes have been used. Moreover, in recent decades, research
on microbial strains with specific characteristics has led to a wide spectrum of
interesting, fermented products.
Fermentation can improve nutritional value, sensory profile, textural and
rheological features, and functional properties of foods (Rizzello et al., 2017).
Some examples include yeast fermentation of the sugars present in wheat flour
for making bread or of malted barley sugars for making beer; also, the use of
symbiotic bacterial cultures for the fermentation of milk to produce cheese,
yogurt and other fermented milk derivatives and the fermentation of juices
derived from different fruits and grains to produce alcoholic beverages, among
others (Carrizo et al., 2020).
Advances in fermentation studies lead to the development of high-​value
composite products, especially with traditional crops –​such as Latin-​American
grains –​either fermented with specific lactic bacteria and yeast, or complex
microbial combinations, as in spontaneous and sourdough fermentation.

10.2.1 Lactic Fermentation of Latin-​American Grains


Lactic acid bacteria (LAB) have been at the core of microbial fermentation of
foods and beverages as far back as 6000 BC. LAB strains form a heterogeneous
group of microorganisms that reside in a wide variety of habitats and produce
organic acids –​mainly lactic acid –​as a consequence of substrate fermentation
(Carrizo et al., 2020).
Furthermore, the metabolic action of these bacteria leads to amino acids and
B-​group vitamins synthesis, anti-​nutrients degradation, shelf life improvement,
and in some cases functional features, like probiotics (Blandino et al., 2003).
Also, smaller molecule spreadables such as peptides, amino acids and other
nitrogenous compounds, which possess physiological functions, are formed
during fermentation via enzymatic degradation.
Currently, research is focused on the advantage of fermentation to increase
the bioactive and nutritional values of cereals and vegetables, including some
Latin and Mesoamerican grains. Several studies have evaluated quinoa fer-
mentation, from improving the quality of quinoa seeds to the development of
fermented vegan products, such as: quinoa-​based yogurt (Zannini et al., 2018),
374 Nutritional Value Obtained from Latin-American Crops

spoonable products (Väkeväinen et al., 2020), quinoa pasta (Carrizo et al., 2020)
and beverages (Canaviri Paz et al., 2020; Jeske et al., 2018).
Li et al. (2018) studied the fermentation of quinoa seeds with Lactobacillus
casei and found that proteins, free amino acids, carbohydrates and B1 and B2
vitamins were significantly higher in fermented quinoa seeds. However, fat and
dietary fiber content decreased by 52.05% and 45.87%, respectively. Regarding
fermented product development, Väkeväinen et al. (2020) studied the poten-
tial of two quinoa varieties –​Pasankalla and Rosada de Huancayo –​to obtain
spreadable vegan products fermented with a potentially probiotic Lactobacillus
plantarum strain. Products were flavored with date, wild berry and banana. The
strain showed high functionality, nevertheless, cereal products as replacements
for yogurt or desserts usually present sensorial difficulties. In the case of this
quinoa spreadable product, consumers indicated the presence of an unpleasant
aftertaste and sandy mouthfeel. Jeske et al. (2018) developed a milk substitute
with quinoa fermented with L. citreum TR116 and L. brevis TR055, which in com-
bination with exogenous amylolytic enzymes, enabled the reduction of glucose
by 40%. In this case, the use of mannitol-​producing LAB showed potential in a
nutritious and sugar-​reduced, plant-​based beverage.
In addition to quinoa, some authors studied the fermentation of less well-​
known crops, such as kiwicha, tarwi and sacha inchi. Aguilar et al. (2019) studied
the effect of adding the liquid portion of hydrolyzed kiwicha in the production of
probiotic drinks with tarwi juice using Lactobacillus paracasei, Bifidobacterium
longum and a culture of both microorganisms. Results showed an improve-
ment in LAB viability, especially when the amount of hydrolyzed kiwicha was
increased. Furthermore, the viability of B. longum was reduced with the decrease
of pH, but improved in the co-​culture with L. paracasei. Acidity levels enhance
the stability of fermented drinks against pathogen microorganisms. In addition,
the sensorial characteristics of the drinks were favorable.
Vanegas et al. (2018) studied the addition of sacha inchi seeds –​which
are rich sources of n-​3 and n-​6 fatty acids, tocopherols, phytosterols, phen-
olic compounds, protein of high nutritional value and dietary fiber –​ and
betaglucans in yogurt with the aim of improving nutritional and functional
characteristics. This incorporation did not affect fermentation kinetics but sig-
nificantly increased α-​linolenic (49.3%) and linoleic (32.2%) acids, whose values
were about 50-​and 25-​fold higher than those of the control yogurt. Despite tex-
tural parameters ( firmness, consistency, cohesiveness and index of viscosity)
of the enriched yogurts, they were significantly lower than those of the con-
trol samples during the whole storage period; all enriched yogurts showed a
sensorial acceptance higher than 70% by untrained panelists, reaching values
similar to control.
Furthermore, Micanquer et al. (2020) evaluated the formulation of an uncon-
ventional fermentation substrate from agro-​industrial waste of pineapple and
10.2 Fermentation 375

sacha inchi to produce biomass. Nowadays, the main use of sacha inchi is to
obtain oil. This process generates a sub-​product of 59% protein content (Wang
et al., 2018), which represents between 60 and 75% of the seed’s total weight. The
authors studied the use of this high protein and carbohydrate portion with pine-
apple peel to formulate a fermentation substrate, appropriate for the growth
of Weissella cibaria, an LAB with potentially multiple uses in the food industry.
Different homogenization and heating treatments were assessed to increase the
feasibility of reducing sugars and improving the efficiency of these wastes as fer-
mentation substrates.
Regarding chia fermentation, Dentice-​Maidana et al. (2020) isolated LAB
from chia sourdoughs to develop sorghum gluten-​free bread. Strains –​Weissella
cibaria, Lactobacillus plantarum and Lactobacillus fermentum –​were selected
based on techno-​functional and safety properties. Fermentation products pre-
sent in sourdoughs after 24 h showed a marked increase in lactate, xylose, ara-
binose, free amino acids and hydrogen peroxide while glucose was undetectable.
Bread manufactured with 30 and 40% of chia sourdoughs significantly improved
specific volume and visual appearance compared to 100% sorghum bread.

10.2.2 Production of Vitamin and Antioxidant


Compounds through Lactic Fermentation
Fermentation is a complex process where different metabolic products are
developed –​and some conditions favor the presence of specific ones of par-
ticular interest –​such as B group vitamins and antioxidants. The study of
strains, substrates and proper fermentation conditions may be a promising bio-​
enrichment strategy.
For instance, Silva-​Vieira et al. (2017) and Mosso et al. (2020) studied folate
enrichment through lactic fermentation of different substrates with the incorp-
oration of amaranth flour. In both cases, this Andean grain addition increased
folate production by B. longum subsp infantis BB-​0 in the first and L. sakei
CRL2210 in the latest. Results demonstrated that this vitamin production is not
only strain-​dependent but also influenced by the addition of different substrates
in the growth media.
High antioxidant activities have been found in many fermented cereals, such
as wheat, rice, oat, maize and sorghum, which were the basis for investigations
regarding fermentation of Latin-​American seeds and grains as an effective
approach for the preparation of bioactive substances. According to Rizzello
et al. (2016), the antioxidant properties of quinoa flour were improved via
fermentation with Lactobacillus plantarum. Regarding quinoa seeds, Li
et al. (2018) found that fermentation with L. casei increased the 2.2-​diphenyl-​
1-​
picrylhydrazyl (DPPH) radical scavenging activity, reducing ability and
Fe2+​-​chelating activity. Fermented quinoa seeds showed a higher total phenolic
376 Nutritional Value Obtained from Latin-American Crops

content (16.53 mg gallic acid equivalent (GAE)/​g extract, dry weight) than
unfermented seeds (13.85 mg GAE/​g ). Fermentation increased free phenolics
and decreased bound phenolics.
Antioxidant activity related to amino acid residues was studied by
Montemurro et al. (2019) who fermented quinoa flour with autochthonous
L. plantarum T0A10. Five peptides, released during fermentation and having
from 5–​9 amino acid residues, were purified and identified as responsible for
the increase in antioxidant activity. Such sequences, encrypted in native quinoa
proteins and released through the proteolytic activity of the LAB strain, showed
antioxidant activity on human keratinocytes NCTC 2544 artificially subjected to
oxidative stress.

10.2.3 Reducing Antinutritional Factors through


Fermentation
Mineral absorption in humans is reduced in the presence of anti-​nutrients, such
as alkaloids, in tarwi and phytic acid in quinoa, kañihua and amaranth.
Phytic acid binds to positively charged divalent cations such as iron, zinc, cal-
cium and proteins, forming phytate complexes that are stable at intestinal pH
(6–​7), thus inhibiting the absorption of minerals in the small intestine (Castro
Alba et al., 2019). Phytase, an enzyme present in plants and microorganisms,
including certain strains of LAB and yeasts, decreases phytate content and
improves mineral bioavailability (Carrizo et al., 2020).
For instance, Castro Alba et al. (2019) studied the phytic acid degradation
during spontaneous fermentation, specifically using Lactobacillus plantarum
299v in quinoa, kañiwua and amaranth grains and their flours. Higher phytic acid
degradation was found during the fermentation of flours (64–​93% from original
content) than with grains (12–​51%). The addition of Lactobacillus plantarum
299v increased the concentration of lactic acid, with a concomitant pH decrease,
which provided the conditions necessary for the activation of endogenous and
microbial phytase.
Tarwi consumption is limited due to the presence of bitter alkaloids and
other anti-​nutritional factors, such as phytic acid, tannins, nitrates and trypsin
inhibitors that have undesirable physiological effects. Nowadays, debittering
is performed through a conventional aqueous thermal treatment. Villacres
et al. (2020) assessed the efficacy of fermentation with Rhizopus oligosporus to
reduce these compounds. The fungal treatment decreased the following anti-​
nutrients from debittered grain by the following percentages: nitrates (94.6%),
tannins (82.1%), alkaloids (94.0%), urease activity (93.7%), phytic acid (70.0%)
and trypsin inhibitors (76.7%), and increased total carotenoids (165.4%), phen-
olic compounds (1055.5%) and antioxidant capacity (1515.6%).
10.2 Fermentation 377

10.2.4 Traditional Beverages through Maize Fermentation


Chicha is a traditional beverage prepared in several Andean countries from
ancient times –​the first reports of chicha production date back to 200 BC, before
the establishment of the Incas in the region. This type of clear, yellow and frothy
beer is prepared mainly from malted yellow maize grains; nevertheless, other
traditional maize strains are usually used, according to the region/​country, e.g.
white, black, chulpi and morocho maize (Barbosa Piló et al., 2018).
Grains are first germinated (left in water for a day) and then put onto straw
mats or plastic tarps under the sun for two days until they are completely
dried. After boiling, the mixture is strained and then placed in a container to
ferment. Usually, after two days of spontaneous fermentation by indigenous
microorganisms, the beverage is ready for consumption (Vallejo et al., 2013).
Though fermentation is mainly alcoholic, by S. cerevisiae and other yeasts, LAB
and other bacteria are also involved and play a key role in the flavor develop-
ment of chicha. This rich biodiversity has been studied by Bassi et al. (2020) and
Grijalva-​Vallejos et al. (2020) as a resource of biodiversity with potential techno-
logical and health-​beneficial properties.
Pozol is one of the most ancient Mesoamerican fermented products still
consumed in south Mexico and in several places of Central America. This
traditional beverage is used for alimentary and ceremonial purposes in many
ethnic groups of southeastern Mexico. Typical elaboration starts with little
nixtamalized maize balls wrapped in banana leaves and left to ferment with
endogenous microbiota for a certain time, which can be as short as a few hours
or as long as a month. After natural fermentation, the maize balls are mixed with
water and consumed as a refreshing beverage (Pérez-​Armendáriz et al., 2020).

10.2.5 Impact of Fermentation on Matrix


Structural Changes
In recent decades, research on gluten-​free products has increased. Dentice-​
Maidana et al. (2020) used sorghum to develop gluten-​free bread augmented
with different percentages of chia sourdoughs (0–​40%), and studied techno-
logical and sensorial features. Based on techno-​functional and safety proper-
ties, Weissella cibaria CH28 together with Lactobacillus plantarum FUA3171
and Lactobacillus fermentum FUA3165, were used as inoculants to ferment chia
sourdoughs used for bread making. Sensory evaluation showed that breads with
40% of fermented chia were the most accepted by panelists who were also able to
discern bread inoculated with fermented and unfermented chia dough.
Another trend in the food market is development of lactose-​free products that
mimic milk, cheese and yogurt characteristics. In this sense, fermentation also
378 Nutritional Value Obtained from Latin-American Crops

plays a key role. Zannini et al. (2018) developed a novel beverage fermented with
Weissella cibaria MG1 based on aqueous extracts of wholemeal quinoa flour to
simulate yogurt. This strain is characterized by exo-​polysaccharides (EPS) pro-
duction at the end of the fermentation, which improves water holding capacity
and viscosity. Microstructure observation indicated that the network structures
of EPS-​protein improve the texture of fermented quinoa milk. Overall, Weissella
cibaria MG1 showed satisfactory technological properties and great potential
for further possible application in the development of high viscosity fermented
grains.

10.3 GERMINATION
10.3.1 Germination Conditions
Germination is another traditional and economical, biotechnological process,
which can easily be performed with a wide variety of crops such as quinoa,
amaranth, kañiwa, chia, black beans, corn, tarwi and sacha inchi, among others.
Under controlled conditions of temperature, humidity, light and time, germin-
ation improves protein digestibility, increases soluble sugars, free fatty acids and
antioxidants, and decreases anti-​nutritional compounds.
Different antifungal procedures are recommended before seed germination.
In most cases, prior to germination, seeds are soaked to induce water imbibi-
tion. Germination occurs in nature when seeds find proper conditions and
start the first steps of plant development; in order to mimic that process, sev-
eral parameters must be studied specifically for each seed. Some germination
conditions are described in Table 10.1. After germination, seeds are dried to stop
the enzymatic activity (in a vacuum oven, forced air circulation oven or by lyoph-
ilization), usually milled, and finally stored until analysis.

10.3.2 Changes in Nutritional Profile during Germination


10.3.2.1 Protein Content, Amino Acids Profile and
Protein Digestibility
Germination increased the total protein content of amaranth, quinoa
(Padmashree et al., 2018), chia (Gómez-​Favela et al., 2017; Pająk et al., 2019) and
sacha inchi (Chandrasekaran and Liu, 2014), possibly due to protein synthesis
activation and/​or dry matter reduction, particularly through the oxidation
of carbohydrates and the loss of carbon dioxide, water and small amounts of
ethanol during respiration.
Changes in protein profiles have been found during germination. For instance,
after 18 h of amaranth germination, most of the essential amino acids increased,
newgenrtpdf
TABLE 10.1 GERMINATION CONDITIONS
Antifungal Germinative
Crop Treatment Germination Conditions Observations capacity Reference
Amaranth NaClO Soaking: 6 h Plastic boxes with For 18 h: 52% Guardianelli
(20% v/​v) 30°C –​Agitation absorbent paper (MSLa: 0.40 cm) et al. (2019)
20 min 18 and 24 h For 24 h: 73%
(MSLa: 0.55 cm)
Quinoa -​ Soaking: 6 h (RSWb 1:5) Plastic trays covered 90% (MSLa: 1.0–​ Jimenez et al.
22–​14°C –​ RHc 90% –​Dark with wet filter papers 1.5 cm) (2019)
24 h
Purple corn NaClO Soaking: 24 h at room Trays on wet filter -​ Paucar-​Menacho
(0.1% v/​v) temperature (RSWb 1:5) paper et al. (2017)
RSWb 1:5 12–​28°C (optimal 26°C) –​
30 min RHc 90% –​Dark
12–​72 h (optimal 63 h)
Black turtle -​ Soaking: 24 h (RSWb 1:3) -​ 90% (MSLa: 2.1 cm) Guajardo-​Flores
bean with aeration et al. (2012)
20°C –​RHc 92% -​Dark
5 days
-​ 25 °C –​Dark 6 days Into a semi-​automatic 75% (MSLa: 18.5 cm) Xue et al. (2016)
germination machine

10.3 Germination
Chia Etanol(96%) 22±2ºC –​12/​12 h day/​night Sterile stackable trays -​ Pajak et al.
1 min 7 days (2019)
Kañiwa NaClO Hydrated: 14 h at 20°C and -​ -​ Abderrahim
(0.3% v/​v) RHc 45% et al. (2012)
15 min 20°C –​Dark 48, 72 or 96 h
(continued)

379
newgenrtpdf
380
Nutritional Value Obtained from Latin-American Crops
TABLE 10.1 (CONTINUED)

Antifungal Germinative
Crop Treatment Germination Conditions Observations capacity Reference
Sacha inchi -​ 18°C Dark or 26 °C Light On sterile filter paper 90% Chandrasekaran
3, 6, 10, 20 and 30 days in Petri dish and Liu (2014)
Tarwi -​ Soaking: 24 h at 16°C with -​ -​ Villacrés et al.
stirring (2015)
16°C and 20°C –​RHc 45%
and 50%
2, 3 and 4 days
Notes: aMSL: Mean sprout length; bRSW: Ratio seeds:water; cRH: Relative humidity.
10.3 Germination 381

except for histidine and tyrosine levels that were maintained and sulfur amino
acids (methionine and cysteine) which significantly decreased. In addition, after
24 h of germination, the content of proline significantly increased, possibly due
to having an adaptive role in plant stress tolerance (Guardianelli et al., 2019).
Regarding the germination of black bean, amaranth and quinoa, a reduc-
tion in high-​molecular-​weight proteins and an increase in polypeptides was
observed, because part of the energy used for germination and plant develop-
ment comes from the mobilization of storage proteins that were hydrolyzed by
proteolytic enzymes and converted into soluble peptides and amino acids (Xue
et al., 2016; Guardianelli et al., 2019; Jimenez et al., 2019; Piñuel et al., 2019).
Protein digestibility of quinoa and amaranth significantly improved with ger-
mination (24% and 20%, respectively) (Chaparro et al., 2010; Jimenez et al., 2019),
possibly due to protein hydrolysis. Villacrés et al. (2015) observed an increase in
protein digestibility reaching values between 85 and 87% during germination of
tarwi; and Gómez-​Favela et al. (2017) also observed a significant increase in pro-
tein digestibility (5%) with chia germination.

10.3.2.2 Lipid Content and Fatty Acids’ Profile


During germination of white quinoa, the lipid content was kept the same
(Padmashree et al., 2018); however, a decrease in the amount of lipids was
observed after germination of red quinoa, amaranth (Jimenez et al., 2019), chia
(Pająk et al., 2019) and sacha inchi (Chandrasekaran and Liu, 2015). A decrease
in lipid content may be due to catabolic activity since, during germination, glu-
cose molecules can be obtained from fatty acids of lipids in the glyoxylate cycle.
Regarding amaranth germination, Guardianelli et al. (2019) observed a
decrease in the content of palmitic acid, stearic acid and oleic acid; while lino-
leic and α-​linolenic acids increased, probably due to the conversion of oleic acid
by desaturase enzymes. On the other hand, Jiménez et al. (2020a) found that
saturated fatty acids decreased with both quinoa and amaranth germination,
caused by the reduction in palmitic and behenic acids, which was possibly due
to the lipolytic activity and the decomposition of triglycerides and polar lipids
into simpler compounds.
Thin layer chromatography (TLC) of quinoa and amaranth grains showed
a decrease in triglycerides and an increase in free fatty acids, phospholipids,
diacylglycerols and sterols during germination (Jimenez et al., 2020a); the
pathway could be related to the breakdown of reserve triglycerides into simpler
compounds through hydrolytic degradation by the action of lipases (Jan et al.,
2018). During sacha inchi germination, a rapid decline in the triacylglycerol and
diacylglycerol content was observed after the early stages (3–​10 d after imbi-
bition) followed by a steady breakdown during the following stages of germin-
ation. Also, a rapid increase in free fatty acids content between 10 and 20 days
382 Nutritional Value Obtained from Latin-American Crops

of sacha inchi germination was observed, after which, free fatty acids decreased
during the late stage of seed germination, which could indicate the possible bio-
synthesis of fatty acids during the later stages of germination (Chandrasekaran
and Liu, 2015).

10.3.2.3 Carbohydrates Profile
The carbohydrate content decreased after 48 h of germination of both white
and red quinoa varieties (62–​52% and 58–​50%, respectively), possibly because
carbohydrates were utilized as an energy source for the growth of the embryo
during germination (Padmashree et al., 2018). The starch amount significantly
decreased, due to its degradation into low-​molecular-​weight compounds such
as glucose and fructose to provide energy for cell division. Villacrés et al. (2015)
observed a reduction of starch from 1.6 to 1.1% after 4 d of tarwi germination.
Jiménez et al. (2019) found a reduction in digestible starch of 52% and 64% during
the germination of amaranth and quinoa, respectively; however, resistant starch
content was kept constant, since it cannot be hydrolyzed by endogenous enzymes.
Besides, changes in starch conformation occur with germination. Apparently,
amylose content incremented with germination between 13 and 65% in quinoa
and over 100% in amaranth grains, possibly due to the hydrolysis of primary
amylopectin by the action of the amylolytic enzymes α-​amylase and β-​amylase,
which could release the linear branches of glucan chains and dextrin that react
with iodine such as amylose (Jimenez et al., 2019).
The effect on dietary fiber –​total, soluble and insoluble –​depends on ger-
mination conditions and the nature of the crop, due to the different structures
and compositions of cell walls (Omary et al., 2012). Repo-​Carrasco-​Valencia and
Serna (2011) and Jimenez et al. (2019) observed that total, soluble and insol-
uble fiber content did not change with germination of quinoa and amaranth.
Nevertheless, Padmashree et al. (2018), Dueñas et al. (2016) and Gómez-​Favela
et al. (2017) observed an increase in both insoluble and total fiber with the ger-
mination of quinoa, black bean and chia respectively, which may be related to
the hemicellulose, cellulose and pectic polysaccharides produced in advanced
stages of germination.
In the case of chia germination, soluble dietary fiber decreased 13.5%, pos-
sibly due to the loss of mucilage during the hydration of the seeds. However,
other authors discerned an increase in soluble fiber and a decrease in insoluble
fiber with corn germination (Gong et al., 2018).

10.3.3 Antioxidant Content and Antioxidant Capacity


During germination, the bonds of phenolic compounds with non-​ starch
polysaccharides localized in the cell walls are degraded by enzymatic action,
mainly esterase, which produces a consequent increase in free phenolic acids
10.3 Germination 383

( ferulic, p-​coumaric and synapic acids). On the other hand, secondary metab-
olism (specifically the phenylpropanoid metabolic pathway) could be initiated
by the activation of phenylalanine lyase (a key enzyme in phenolic biosynthesis),
causing the production of phenolic compounds (such as p-​coumaric, caffeic and
ferulic acids) from aromatic compounds such as amino acids (phenylalanine,
tyrosine and tryptophan) (Gómez-​Favela et al., 2017; Pajak et al., 2019). An
increase in total phenolic and flavonoid is correlated with enhanced antioxidant
activity during germination.
The effect of light in germination parameters was studied for some crops. For
instance, in amaranth, the light had no effect on gallic acid content but increased
rutin concentration; whereas, when germination was performed in the dark, a
higher content of isovitexin and vitexin was recorded (Nemzer et al., 2019).
An increase in phenolic compounds and γ-​aminobutyric acid (GABA) was
observed during chia germination under optimal conditions (Gómez-​Favela
et al., 2017). Nevertheless, in purple corn, a reduction in phenolic compounds,
anthocyanins and flavonoids content was found when performed in optimal
germination conditions to produce GABA (26°C, 63 h). This phenomenon may
be attributed to polyphenol-​oxidase activation –​an enzyme responsible for
phenolic compound degradation during germination –​or to the leaching of
these compounds during the soaking period. GABA increase may be related to
the activation of enzymes involved in its production and glutamate synthesis,
primarily by the decarboxylation of L-​glutamic acid and catalyzed by glutamate
decarboxylase during seed germination (Paucar-​Menacho et al., 2017).
Tocopherols are lipophilic antioxidants that are also modified during germin-
ation. Carciochi et al. (2016) observed that the total tocopherol content (TTC) of
quinoa significantly increased after 72 h of germination; however, Jiménez et al.
(2020a) found a significant decrease in TTC after 24 h-​quinoa germination, with
a decrease in β and γ-​tocopherols, despite the increase in α-​tocopherol. On the
contrary, after 48 h-​amaranth germination, TTC was raised due to the significant
increase of α-​ and β and γ-​tocopherols, despite the reduction in δ-​tocopherol.
Changes in TTC are related to the activation of metabolic pathways during
germination, such as cytosolic shikimate and plastid methylerythritol phos-
phate. Besides, biosynthesis and the increase or decrease in TTC depends on
stress conditions. Lushchak and Semchuk (2012) explained that in dicots (like
soybean, quinoa and amaranth) α-​tocopherol levels initially increased during
germination and then decreased.

10.3.4 Antinutrients and Mineral Bioavailability


An increase in mineral bioavailability is often observed during germination, due
to the hydrolysis of phytic acid by the increment in phytate enzyme activity (pos-
sibly myoinositol hexakisphosphate phosphohydrolase), which consequently
384 Nutritional Value Obtained from Latin-American Crops

leads to the release of minerals associated with phytic acid, mainly divalent
minerals such as calcium, magnesium, iron, copper, manganese and zinc (Świeca,
2016). An increase in the bioavailability of minerals was observed with the ger-
mination of chia and tarwi (Pająk et al., 2019; Villacrés et al., 2015). Padmashree
et al. (2018) also observed a reduction of phytic acid after 48 h of both white and
red quinoa germination (52.3–​42.7 mg/​100 g and 63.7–​44.9 mg/​100 g, respect-
ively), where phytic acid might have been broken down for the utilization of
phosphorus during germination. Saponin and tannin contents decreased in red
quinoa germination (0.6–​0.4 g/​100 g and 3.4–​2.9 g/​100 g, respectively), while no
significant variation was found after white quinoa germination.
Decrease in saponins may be due to leaching during washing and soaking
before germination. During the three days of black turtle bean germination,
the total saponin amount in the seed increased due to the higher concentra-
tion detected in sprouts and cotyledons, which represents 90.8% of the total seed
weight. This increase is possibly due to the synthesis and activation of different
enzyme systems with germination, which enhances the production of these sec-
ondary metabolites and the weakening of the seed structure, which facilitated
solvent extraction procedures. However, saponins of the seed coats decreased,
possibly due to leaching (Guajardo-​Flores et al., 2012).

10.3.5 Technological, Thermal and Sensory Changes


during Germination
Changes occurring during germination cause modifications in technological,
rheological, textural, thermal and sensory properties which may be considered,
particularly, when sprouts are used as ingredients in food formulation.

10.3.5.1 Rheological and Textural Characteristics


Several authors have studied the incorporation of germinated flours in bread,
purees and pasta, among other bakery products. For instance, in bread making,
the replacement of native quinoa flour with germinated quinoa flour (ratio 80:20
wheat to quinoa) produced an increase in dough water absorption and in the
softening degree. Nevertheless, stability decreased, suggesting a weakening of
the gluten network. Additionally, there was improved dough development and
gas production during leavening, which produced a high specific volume and
low crumb firmness, probably due to increased sugar content (maltose, sucrose
and glucose) (Suárez-​Estrella et al., 2020).
The germinated quinoa flours showed lower pasting properties than those
made with ungerminated grains. Viscosity decreased after 48 h of quinoa ger-
mination, from 177 rapid-​visco analyser units (RVU to 42 RVU). In addition, set-​
back viscosity also decreased with germination; thus, starch molecules have a
10.3 Germination 385

lower ability to disperse in hot paste and reassociate during cooling (Padmashree
et al., 2018).
Ujiroghene et al. (2019) made a functional yogurt from sprouted and non-​
sprouted quinoa milk. The use of sprouted quinoa milk increased the total con-
tent of phenols and flavonoids, and the antioxidant capacity. Furthermore, the
water-​holding capacity of “germinated quinoa yogurt” was significantly higher
than the non-​germinated one, which may be attributed to the production of
compounds such as soluble sugars with good water-​holding capacity.
Jiménez et al. (2019) formulated baby purees with quinoa and amaranth
flours and studied the changes produced by their replacement with sprouted
grain flours. Purees made with germinated grain flours showed a significant
decrease in the consistency coefficient, and in the storage (G’) and in the loss (G’’)
modules in both stress and frequency sweeps. In addition, the flow curves (shear
stress versus shear rate) showed downward displacement, indicating a weaker
food matrix and greater fluidity with the use of germinated grains flours, which
may be related to the enzymatic hydrolysis of starch and proteins (Cornejo et al.,
2019). Also, purees elaborated with germinated grain flours had significantly less
hardness and adhesiveness than those made with non-​germinated flours, due
to the lower number of polysaccharide chain networks within the food matrix
caused by lower starch content.

10.3.5.2 Thermal Behaviors
The thermal properties of germinated and non-​germinated amaranth flours were
determined by a differential scanning calorimetry (Guardianelli et al., 2019). The
gelatinization peak temperature increased with germination time (18 and 24 h)
because protein fractions were changed by germination. Besides, significantly
lower changes of enthalpy (ΔH) values were obtained in the germinated grain
flours (6.7 J/​g and 4.7 J/​g for 18 and 25 h of germination, respectively) as opposed
to the non-​germinated grain flours (9.6 J/​g ), which could be attributed to lower
starch content and changes in protein structure due to enzymatic hydrolysis
during germination. Jiménez et al. (2019) agreed with the reduction of ΔH with
germination of quinoa (24 h) and amaranth (48 h), but no changes in gelatiniza-
tion temperatures (initial, peak or final) were found.

10.3.5.3 Sensory Characteristics
In some foods, the use of sprouted grains showed no significant changes in sen-
sory evaluation; for instance, in the use of sprouted quinoa liquid to obtain a
functional yogurt-​like product, germination had little or no importance in the
overall organoleptic acceptability of the final product (Ujiroghene et al., 2019).
Nevertheless, replacement of raw quinoa flour with germinated quinoa flour
in pasta formulation significantly changed the taste (Demir and Bilgicli, 2020).
386 Nutritional Value Obtained from Latin-American Crops

Moreover, bread with sprouted quinoa flour was characterized by decreased


bitterness assessed by electronic tongue in comparison to bread with unsprouted
quinoa flour (Suárez-​Estrella et al., 2020).
In addition, the use of sprouted quinoa and amaranth flours for the elab-
oration of infant purees had a negative impact on sensory analysis, since the
attributes most frequently used to describe purees made with sprouted grain
flours were: ‘more liquid’, ‘slightly acidic’, ‘with a bitter taste’, ‘intense taste
with an aftertaste’. Therefore, it is necessary to continue studying ways to mask
unpleasant flavors in using some sprouts as food ingredients (Jimenez et al.,
2020b).

10.4 BIOACTIVE PEPTIDE PRODUCTION


According Morales et al. (2020), the Food and Agriculture Organization (FAO)
has classified plant proteins as highly important for the human diet, highlighting
their natural abundance and nutritional value. For these reasons, their biological
properties are currently being studied, especially as a source of biologically active
peptides when included in the diet. These peptides –​amino acid sequences
with no activity in the precursor protein –​may display physiological functions
in the body after their release by gastrointestinal digestion or in vitro digestion
by chemical or enzymatic hydrolysis, fermentation and/​or germination (Orona
Tamayo et al., 2019; Apostolopoulos et al., 2021).

10.4.1 Enzymatic Hydrolysis
Enzymatic hydrolysis has prevailed over chemical hydrolysis since the
hydrolysates and peptides led by the first technique show higher activity and
absorption, in addition to the easy regulation of the bioprocess, low cost and
simple implementation. The bioprocess can be easily implemented; furthermore,
peptides with better absorption and activity can be produced. Peptide length,
amino acid constitution, presence of polar and ionizable groups and hydropho-
bicity play a key role in the resulting functional and bioactive properties.
The enzyme catalase is the most used in protein hydrolysis studies of Latin-​
American crops, usually used at high temperatures (45°C) and alkaline pH
(8.5–​10). In addition, other enzymes have been studied, such as trypsin, α-​
chymotrypsin, protease, bromelain, papain, pepsin, pancreatin, flavourzyme,
neutrase, in in vitro gastrointestinal systems and even enzymatic extracts from
the stomach and intestines of different animals.
Studies on hydrolysates and peptides obtained by enzymatic hydrolysis of
proteins from Latin-​American crops are mainly focused on antioxidant properties
10.4 Bioactive Peptide Production 387

(through antioxidant assays of ABTS [2,2’-​azino-​bis(3-​ethylbenzothiazoline-​6-​


sulfonic acid)], DPPH and metal chelating capacity-​ferrozine), antihypertensive
properties (by angiotensin-​I converting enzyme –​ACE-​inhibitory activity) and
antimicrobial properties (well diffusion method).
Otherwise, there is little but promising information on the inhibitor activity of
adipogenesis, antihaemolytic activity in red blood cells, antihypertensive activity
in spontaneously hypertensive rats and renin inhibition activity, inhibitory
activity of colon cancer cell viability and antidiabetic and hypocholesterolemic
cholesterolemic activity, as well as immunomodulatory activity (Morales et al.,
2020; Moughan et al., 2014; Daliri et al., 2017).
Several enzymes and enzyme systems are used for protein hydrolysis of
different quinoa cultivars, where incubation intervals and the enzyme to sub-
strate ratio varies considerably. For instance, chymotrypsin showed higher effi-
ciency than bromelian and protease enzymes, reaching 85% degree of hydrolysis
in 6 h, for an enzyme to substrate ratio of 1:100. Glutelins were the most sus-
ceptible proteins to enzymatic attack and globulins showed the highest resist-
ance (Mudgil et al., 2019). In vitro simulation of gastrointestinal digestion was
also assessed, where pepsin and pancreatin were studied successively and sep-
arately; moreover, the influence of bile salts in the process was determined.
The most common enzyme to substrate ratio was 1:100, reaching over 40% of
the degree of hydrolysis in 2 h. Peptide fractions were separated by size (>/​<5
kilodalton (kDalton), even a separation was performed by reversed-​phase-​high
performance liquid chromatography (RP-​ HPLC), identifying peptides with
encrypted structures in the 11 S globulin protein (Vilcacundo et al., 2018a; Shi
et al., 2019). Papain was also used, in natural and refined extract, verifying a
high degree of hydrolysis on quinoa protein concentrates, using an enzyme
to substrate ratio of 2:100. In this case, the peptides fraction of less than 5
KDalton represented 50% of the hydrolysate, after 180 min of incubation at 50°C
(Nongonierma et al., 2015).
Furthermore, the enzymatic hydrolysis of amaranth proteins was extensively
studied, finding that in all cases the process was slower than quinoa proteins’
hydrolysis. Mudgil et al. (2019) determined that the hydrolysis degree of amar-
anth proteins with chymotrypsin was around 40% after 6 h. Ayala-​Niño et al.
(2019) obtained a similar hydrolysis degree in Amaranth hypochondriacus spp
proteins using alcalase (pH 10) and flavorzyme (pH 7), during 6 h, achieving a
greater efficiency with the use of both enzymes sequentially.
The protein isolates of black beans were studied by hydrolysis with a
simulated gastrointestinal system (pepsin and pancreatin) and with alcalase
(ratio 1:100) for the production of peptides with different functional properties,
especially hypoglycemic. The hydrolysates were separated into nine fractions
with molecular weights between 400 and 1200 Dalton. Peptides with the highest
388 Nutritional Value Obtained from Latin-American Crops

concentration were identified for their subsequent synthesis and study (Mojica
et al., 2017a, 2017b).
Aguilar-​Toalá et al. (2020) hydrolyzed chia proteins from defatted flours from
which their characteristic mucilage was removed. Chia seeds were hydrolyzed
using single (alcalase) or sequential (alcalase +​flavorzyme) enzymatic processes
at 2%, with the conventional water bath-​and microwave-​assisted process.
For the antimicrobial activity tests, they used the whole hydrolysate and the
peptide fractions of 3–​10 and <3 kDalton. On other hand, Martínez-​Leo and
Segura-​Campos (2020) hydrolyzed chia protein using a sequential pepsin (pH2)-​
pancreatin (pH7.5) system at 37°C. The protein hydrolysate was fractionated by
an ultrafiltration process to three peptide fractions < 1 kDalton, 1–​3 kDalton and
3–​5 kDalton in order to study pro-​inflammatory modulation of HMC3 micro-
glial cells.

10.4.2 Physiological Functions
10.4.2.1 Antioxidant Activity
The antioxidant activity of quinoa protein hydrolysates produced with chymo-
trypsin, protease and papain was studied. Generally, this activity increased
with the hydrolysis degree –​the smallest molecular size fractions (less than 5
kDalton) –​being the most efficient. Oxygen radical absorption capacity (ORAC)
was measured with ABTS and DPPH tests (Mudgil et al., 2019, Nongonierma
et al., 2015). ORAC of quinoa protein hydrolysates peptides obtained from
simulated in vitro digestion was measured by the fluorescein test. Results show
that successive digestion with pepsin and pancreatin is necessary to give the
hydrolysates a high antioxidant capacity, since gastric digestion alone did not
increase the antioxidant property in the same proportion as gastroduodenal
(Vilcacundo et al., 2018a).
Mudgil et al. (2019) determined that the antioxidant properties of amar-
anth protein hydrolysates were significantly higher than those determined in
native proteins and even quinoa hydrolysates, especially when hydrolysis was
performed with chymotrypsin. This capacity, determined by ABTS and DPPH
tests, increased with hydrolysis degree so that the enzymes with greater effi-
ciency for hydrolysis (chymotrypsin and protease) generated hydrolysates
with greater antioxidant capacity. On the other hand, Ayala-​Niño et al. (2019)
measured the antiradical (ABTS and DPPH) and Fe-​chelating (FRAP) activity
of Amaranth hypochondriacus protein hydrolysates produced with alcalase and
flavorzyme, finding an activity increase between 500 and 1000% compared to
native proteins.
The peptides obtained by in vitro gastrointestinal hydrolysis of black bean
proteins showed excellent antioxidant properties. The complete hydrolysates
10.4 Bioactive Peptide Production 389

were twice as effective as pure peptides in scavenging radicals, indicating


that the property was based on a joint action of various peptides or proteins
contained in the hydrolysates (Mojica et al., 2017b).
The chia peptide fraction of 1–​3 kDalton presented a protective effect after
the induction of tert-​Butyl hydroperoxide (TBHP) damage. HMC3 cell mortality
was significantly reduced. If peptides were used at concentrations of 25, 50 and
100 μg/​mL, a protective effect of 62.3%, 73.4% and 79.0% respectively was shown.
In addition, the generation of reactive oxygen species (ROS) such as H2O2, was
reduced by 50% (Martínez-​Leo and Segura-​Campos, 2020).

10.4.2.2 Cytotoxicity
Regarding cytotoxicity, several studies were performed using cells in order to
determine the functional activity of hydrolysates prepared with quinoa, amar-
anth and other Latin-​American crop proteins. In general, protein concentrates
and hydrolyzates, as well as the different peptide fractions, did not present cyto-
toxic activity against different cells (Caco-​2 and pre-​adipocyte cells, HMC3 cells)
up to concentrations closer to 3000 μg of hydrolyzate/​mL for quinoa and amar-
anth hydrolysates; and up to 400 μg/​mL for chia hydrolysates. This determin-
ation is essential to carrying out studies of different activities of ex vivo protein
hydrolysates.

10.4.2.3 Anti-​Inflammatory Activity
The most widely used in vitro tests to determine the anti-​inflammatory activity
of protein hydrolysates from Latin-​American cultures refer to inhibition of the
enzyme renin and ACE, which regulate blood pressure. The combined hydrolysis
with alcalase and flavourzyme on quinoa proteins caused an ACE inhibition of
up to 60% (Ayala-​Niño et al., 2019). On the other hand, there are studies refer-
ring to inhibition of the nitrous oxide production of phagocytic cells, caused
by quinoa protein hydrolysates produced with papain, pepsin and pancreatin
enzymes. Although they have a positive effect, no significant differences were
determined with respect to the native protein (Shi et al., 2019).
The anti-​inflammatory activity of hydrolyzed amaranth proteins and synthetic
peptides encrypted in amaranth proteins were studied, in order to assess the
inhibition of renin and ACE. Amaranth protein concentrates and hydrolysates
were tested together with seven synthetic peptides that regulated the kin-
etics of the enzyme, demonstrating that amaranth proteins, hydrolysates and
different encoded sequences were able to inhibit both of the enzymes studied.
In addition, a heptapeptide (FNLPILR) was identified as that with the highest
efficiency; also, probable sites of the renin enzyme in which the peptides can
associate to achieve their inhibition were determined (Nardo et al., 2020; Suárez-​
Estrella et al., 2020). A decrease of up to 40 mmHg in systolic blood pressure
390 Nutritional Value Obtained from Latin-American Crops

was also verified in hypertensive rats (renin-​ angiotensin system) through


treatments with concentrates, hydrolysates and synthetic peptides encrypted
in the Amaranth hypochondriacus proteins and even by the administration of
biscuits supplemented with amaranth hydrolysates. The antihypertensive cap-
acity of these foods showed a decrease comparable to that achieved with drugs
recommended for this purpose. The authors determined that the bioactive
amaranth peptides also behaved as vasorelaxant agents on the vascular system
(Suárez-​Estrella et al., 2020; Ontiveros et al., 2020).
In vitro anti-​inflammatory activity of black bean protein hydrolysates was
also verified, inhibiting the action of dipeptidyl peptidase IV (DPPIV) and ACE
enzymes with IC 50 of 0.14 and 0.29 mg/​mL (Mojica et al., 2017b).
Martínez-​Leo and Segura-​ Campos (2020) verified in vitro the anti-​
inflammatory action of different fractions of peptides, separated from chia pro-
tein hydrolysates, especially fraction 1-​3 kDalton. The hydrolysates (100 µg/​
mL) decreased the concentration of nitrous oxide and cytokines in lipopolysac-
charide induced HMC3 cells by up to 50%.

10.4.2.4 Antihemolytic Activity
The antihemolytic activity of quinoa hydrolysates increased with the degree
of hydrolysis and depended on the type of enzyme used. Quinoa hydrolysates
produced with protease increased this property five-​fold compared with protec-
tion of the native protein on human erythrocytes.
In amaranth, enzymatic hydrolysis enhanced the antihemolytic capacity of
proteins, especially when hydrolysis was carried out with protease. In addition,
a pronounced increase in capacity was concomitant with hydrolysis time, that is,
with a decrease in molecular size (Shi et al., 2019).

10.4.2.5 Antimicrobial Activity
Even though quinoa proteins lack antimicrobial activity, it was demonstrated
that their hydrolysates express this activity after a certain degree of hydrolysis.
Mudgil et al. (2019) determined antimicrobial activity of peptides produced
by the protease enzyme, contrasted with S. aureus, S. typhimurium, E. coli and
E. aerogenes. Amaranth protein concentrates and hydrolysates showed the
same behavior: native proteins lacked activity, nevertheless, the hydrolysates
presented antimicrobial activity, especially those produced with the enzyme
chymotrypsin and with a high degree of hydrolysis (Mudgil et al., 2019). Contrary
to what Mudgil et al. (2019) found, there is experimental evidence that anti-
microbial activity increases with the degree of hydrolysis, reaching a maximum
in which peptides have the correct molecular size and amino acid composition
to be able to interact with the phospholipids of the cell membrane, increasing
their thickness and creating pores that cause cell disruption (Mirzaei et al.,
2016), a situation that would be lost with the increasing degree of hydrolysis.
10.4 Bioactive Peptide Production 391

Chia seed protein hydrolysates reported antimicrobial activity against


Escherichia coli, Salmonella enterica and Listeria monocytogenes. Aguilar Toalá
et al. (2020) observed that the <3 kDalton peptide fraction had greater anti-
microbial activity than chia seed hydrolysate and the 3–​10 kDalton fraction;
also, peptides obtained by microwave-​assisted enzymatic hydrolysis had higher
efficiency than those obtained by the traditional method. Peptide fraction <3
kDalton caused modifications in the lag phase, maximum growth and growth
rate of the growth curves and promoted multiple indentations (transmembrane
tunnels), membrane wrinkling, and deformation in the integrity of the bacterial
cell membranes, against Gram-​positive and Gram-​negative strains. Besides,
different peptide sequences were determined in the <3 kDa fraction by liquid
chromatography mass spectrometry (LC-​MS/​MS), some of them with potential
antimicrobial scoring.

10.4.2.6 Antiadipogenic Activity
The adipogenesis inhibitory activity of quinoa and amaranth hydrolysates was
investigated. On insulin-​induced adipocytes, the hydrolysates showed positive
and significantly higher activity (IC50 786.58 μg/​mL) than the proteins without
hydrolyzing up to a concentration of 1,600 μg/​mL, especially in the hydrolysates
obtained with pepsin and with a higher degree of hydrolysis (Shi et al., 2019).

10.4.2.7 Antidiabetic Activity
The antidiabetic action of quinoa hydrolysates obtained with papain and those
processed by gastroduodenal digestions were assessed, measuring the inhib-
ition of the metabolic enzyme dipeptidyl peptidase IV (DPPIV), responsible for
the incretin hormones degradation to decrease insulin secretion in pancreatic
beta-​cells. The hydrolysates showed twice the inhibition of enzyme DPPIV with
respect to protein concentrates (Nongonierma et al., 2015; Vilcacundo et al.,
2018a).
Black bean hydrolysates were able to reduce glucose uptake in vitro using a
monolayer of Caco-​2 cells. These cells decreased glucose uptake by up to 20%
after being in contact with hydrolysate solutions for 30 min, and up to 60% when
using pure peptides (Mojica et al., 2017a). This property was also determined in
live tests, achieving a reduction of up to 20% of the postprandial glucose pre-
sent in the serum of animals treated with these hydrolysates. Furthermore, the
hydrolysates showed a 50% in vitro inhibition of alpha-​glucosidase enzyme.
For this reason, consumption of black bean protein hydrolysates may be
recommended for the treatment of type 2 diabetes (Mojica et al., 2017b).

10.4.2.8 Antithrombotic Activity
The antithrombotic activity of Amaranth hypochondriacus protein peptides
was tested by the inhibition of thrombin enzymes in vitro. These hydrolysates,
392 Nutritional Value Obtained from Latin-American Crops

produced with alcalase and flavourzyme, reached an inhibition of 90%, while the
amaranth proteins reached 10% inhibition (Ayala-​Niño et al., 2019).

10.4.2.9 Anticancer Activity
Recently Vilcacundo et al. (2018a, 2018b) reported that quinoa and amaranth
protein hydrolysates (gastroduodenal hydrolysis) showed antiproliferative
effects of colon cancer cells. Cytotoxic activity against human colorectal
cancer Caco-​2 cells and colon cancer HT-​29 and HCT-​116 cells was studied.
The hydrolysates and especially the high molecular weight peptide fractions
( fraction >5 kDalton) showed a growth inhibition (cytotoxic activity on cancer
cells) of up to 80%. Significant differences were observed between gastric and
gastroduodenal digestions indicating the importance of pancreatin in the gener-
ation of bioactive peptides.

10.4.2.10 Immunomodulatory Activity
The immunomodulatory properties of quinoa protein hydrolysates and 1–​3
kDalton fractions were evaluated in BALB/​c mice. The in vitro cytokine profile
and ex vivo macrophage activation were also evaluated. The quinoa hydrolysates
showed no cytotoxic effect on peritoneal macrophages at 32.5–​1000 µg/​mL and
induced peritoneal and spleen macrophage activation through the production
of cytokines INFγ and TNFα. Increased INFγ may serve as a stimulant, enhancing
phagocytic activity in peritoneal and spleen macrophages which is beneficial for
the immune surveillance system against pathogens. Macrophage activation was
down-​regulated by the co-​production of IL-​10, which guarantees a balanced
activation-​
regulation response for tissue homeostasis. Hence, hydrolysates
proved to be safe for oral administration and enhanced the phagocytic activity
of peritoneal and spleen macrophages by stimulating the innate immune system
(Rueda, 2020).

10.4.3 In Silico Approach and Molecular Docking


For the determination of bioactivity in peptides, computer tools are used more
and more frequently. There are databases such as PepBank, BIOPEP, RCSB PDB,
UniProtKB, in which the sequence of proteins, bioactive peptides, allergenic
proteins and sensory peptides are stored. With all this information, researchers
can compare the amino acid sequence of peptides found in different hydrolysis
assays with those of biofunctional peptides already identified. Therefore, a pep-
tide with biofunctional potential can be predicted by an in silico approach.
Consequently, it is not necessary to isolate or concentrate peptides from
hydrolysates to study their functional properties, since knowing their compos-
ition they can be synthesized and tested on peptides of 95% purity. Furthermore,
10.5 Conclusions 393

availability of the protein and peptide sequence, together with knowledge of


the specificity of the enzymes commonly used for protein hydrolysis, allows for
knowledge of the sequence of peptides encrypted in the proteins that could be
released in potential hydrolysis. In addition, the inhibitory or enhancing action
of these peptides on different enzymes could also be tested by a molecular
approach using a different software (SYBYL-​X 2.1.1-​Tripos a Certara Company,
Maestro 9.1 software –​Schrödinger Software Suite) in which the degree to which
the inhibitors bind to the enzyme can be identified, only with molecular models
(Zheng et al., 2019; Mojica et al., 2017a; Guo et al., 2020).

10.5 CONCLUSIONS
Fermentation, germination and bioactive peptides production are suitable
bioprocesses to apply on crops such as quinoa, amaranth, kañiwa, maize, black
turttle bean, tarwi, chia and sacha inchi, prior to their incorporation into food
formulation, since the bio-​and techno-​functional changes produced are prom-
ising. In particular, fermentation and germination are low-​cost and eco-​friendly
methods, which may improve the nutritional, technological and sensorial prop-
erties of products. Although they require specific controls; they are easy-​to-​make
processes, which can be implemented by small enterprises.
Fermentation is a traditional biotechnology that has been applied in the
manufacture and preservation of foods since ancient times. In order to obtain
favorable transformation of raw materials, a proper selection of microorganisms
is fundamental. In this way, fermentation can improve the nutritional value, sen-
sory profile, textural and rheological features and functional properties of foods.
Germination is considered a straightforward and economic bioprocess for
improving the nutritional value of the seeds and, in some cases, improve the
nutraceutical level. Germination –​by the action of proteolytic, lipolytic and
amylolytic enzymes, and activation of secondary metabolic ways –​increases
protein bioavailability, the content of free amino acids and free fatty acids; it
also improves the availability of minerals and vitamins, and increases phyto-
chemical content and, in some cases, decreases anti-​nutritional factors; unfor-
tunately, germination may, eventually, negatively affect some key components of
the substrate.
Quinoa and amaranth are the grains most studied for production of bioactive
peptides; however, currently, there are many scientific publications on other
crops such as chia, tarwi, kañiwa, different races of Andean maize and sacha
inchi, among others.
The latest results obtained indicate that the human gastrointestinal system
would produce peptides with functional properties from the consumption of
394 Nutritional Value Obtained from Latin-American Crops

these traditional foods, so the reevaluation of their production and consump-


tion would notably improve the intake of nutrients and especially functional
components in the region.
On the other hand, the generation of bioactive peptides by enzymatic
hydrolysis has great potential for industrial development since it involves simple
and cheap processes. With current technology, it is possible to achieve good regu-
lation of the degree of hydrolysis and therefore of molecular size, as well as the
composition of peptides to optimize their functional properties. Generally, the
best yields were observed in the 1–​3 and 3–​10 kDalton peptide fractions. A set of
tests in vivo, ex vivo, in vitro and even by computer, is currently available, able to
identify peptides and peptide fractions with functional action and quantifying
antioxidant, antihypertensive, antimicrobial, antiadipogenesis, antihaemolytic,
antidiabetic, anticancer, hypocholesterolemic and immunomodulatory activ-
ities, among others.
The nutritional benefits obtained by germinating most crops, and the fer-
mentation and protein hydrolysis of their flour and/​or concentrates make these
processes advisable in and applicable to the food industry. So, new products with
high nutritional and functional value can be offered to consumers, but care must
be taken so that technological changes are appropriate for food products and
do not affect sensory properties of them and they are acceptable to consumers.

ACKNOWLEDGMENTS
This work was supported by grant laValSe-​ Food-​
CYTED (Ref. 119RT0567);
Consejo Nacional de Investigaciones Científicas y Técnicas (CONICET) and
Universidad Nacional de Jujuy, Argentina.

REFERENCES
Abderrahim, F., E. Huanatico, R. Repo-​ Carrasco-​
Valencia and S.M. Arribas. 2012.
Effect of germination on total phenolic compounds, total antioxidant capacity,
Maillard reaction products and oxidative stress markers in canihua (Chenopodium
pallidicaule). Journal of Cereal Science 56(2):410–​417.
Aguilar, E., and E. Flores. 2019. Assessment of the use of the hydrolyzed liquid fraction
of the kiwicha grain in the fermentation process of probiotic drinks from tarwi
juice: microbiological, chemical and sensorial analysis. Food Science and Technology
39(3): 592–​598.
Aguilar-​Toalá, J.E., A.J. Deering, and A.M. Liceaga. 2020. New insights into the antimicro-
bial properties of hydrolysates and peptide fractions derived from chia seed (Salvia
hispanica L.). Probiotics and Antimicrobial Proteins 12: 1571–​1581.
References 395

Apostolopoulos, V., J. Bojarska, T.-​T. Chai, S. Elnagdy, K. Kaczmarek, J. Matsoukas,


R. New, K. Parang, O. Paredes-​López, O.H. Parhiz, C.O. Perera, M. Pickholz, M.
Remko, M. Saviano, M. Skwarczynski, Y. Tang, W.M. Wolf, T. Yoshiya, J. Zabrocki,
J. Zielenkiewicz, M. Alkhazindar, M. Barriga, K. Kelaidonis, E.M. Sarasia., and
I.Toth. 2021. A global review on short peptides: Frontiers and perspectives.
Molecules 26(2): 1–​45.
Ayala-​Niño, A., G.M. Rodríguez-​Serrano, L.G. González-​Olivares, E. Contreras-​López,
P. Regal-​López, and A. Cepeda-​Saez. 2019. Sequence identification of bioactive
peptides from amaranth seed proteins (Amaranthus hypochondriacus spp.)
Molecules 24(17): 3033.
Barbosa-​Piló, F., E. Carvajal-​Barriga, C. Guamán-​Burneo, P. Portero-​Barahona, A. Morato
Dias, L. Daher de Freitas, F. Oliveira-​Gomes, and C. Rosa. 2018. Saccharomyces
cerevisiae populations and other yeasts associated with indigenous beers (chicha)
of Ecuador. Brazilian Journal of Microbiology 49: 808–​815.
Bassi, D., L. Orrù, J. Cabanillas-​Vasquez, P. Cocconcelli, and C. Fontana. 2020. Peruvian
chicha: A focus on the microbial populations of this ancient maize-​based fermented
beverage. Microorganisms 8(1): 93.
Blandino A, M.E. Al-​Aseeri, S. Pandiella, D. Cantero, and C. Webb. 2003. Cereal-​based
fermented foods and beverages. Food Research International 36: 527–​543.
Canaviri-​Paz, P., R. Janny, and A. Hakansson. 2020. Safeguarding of quinoa beverage pro-
duction by fermentation with Lactobacillus plantarum DSM 9843. International
Journal of Food Microbiology 324: 108630.
Carciochi, R.A., L. Galván-​D’Alessandro, P. Vandendriessche, and S. Chollet. 2016. Effect
of germination and fermentation process on the antioxidant compounds of quinoa
seeds. Plant Foods for Human Nutrition 71(4): 361–​367.
Carrizo, S., A. de Moreno de LeBlanc, J. LeBlanc, and G. Rollán. 2020. Quinoa pasta
fermented with lactic acid bacteria prevents nutritional deficiencies in mice. Food
Research International 127: 108735.
Castro-​Alba, V., C. Lazarte, D. Perez-​Rea, N. Carlsson, A. Almgren, B. Bergenståhla, and Y.
Granfeld. 2019. Fermentation of pseudocereals quinoa, canihua, and amaranth to
improve mineral accessibility through degradation of phytate. Journal of the Science
of Food and Agriculture 99: 5239–​5248.
Chandrasekaran, U., and A. Liu. 2014. Stage-​specific metabolization of triacylglycerols
during seed germination of Sacha Inchi (Plukenetia volubilis L.). Journal of the
Science of Food and Agriculture 95(8): 1764–​1766.
Chaparro, D.C., P. Pismag, E. Correa, V. Quila, and C.A. Caicedo. 2010. Effect of germin-
ation on the protein content and digestibility in amaranth, quinoa, soybean and
grandul seeds. Biotecnología en el Sector Agropecuario y Agroindustrial 8(1): 35–​42.
Cornejo, F., G. Novillo, E. Villacrés, and C.M. Rosell. 2019. Evaluation of the physico-
chemical and nutritional changes in two amaranth species (Amaranthus quitensis
and Amaranthus caudatus) after germination. Food Research International
121: 933–​939.
Daliri, E., D.H. Oh, and B.H. Lee. 2017. Bioactive peptides. Foods 6(5): 32.
Demir B., and N. Bilgiçli. 2020. Changes in chemical and anti-​nutritional properties of
pasta enriched with raw and germinated quinoa (Chenopodium quinoa Willd.)
flours. Journal of Food Science and Technology 57: 3884–​3892.
396 Nutritional Value Obtained from Latin-American Crops

Dentice-​Maidana, S., S. Finch, M. Garro, G. Savoy, M. Gänzle, and G. Vignolo. 2020.


Development of gluten-​free breads started with chia and flaxseed sourdoughs
fermented by selected lactic acid bacteria. LWT –​ Food Science and Technology
125: 109189.
Dueñas, M., T. Sarmento, Y. Aguilera, V. Benitez, E. Mollá, R.M. Esteban, and M.A. Martín-​
Cabrejas. 2016. Impact of cooking and germination on phenolic composition
and dietary fibre fractions in dark beans (Phaseolus vulgaris L.) and lentils (Lens
culinaris L.). LWT –​ Food Science and Technology 66: 72–​78.
Gómez-​Favela, M.A., R. Gutiérrez-​Dorado, E.O. Cuevas-​Rodríguez, V.A. Canizalez-​Román,
C.R. León-​Sicairos, J. Milán-​Carrillo, and C. Reyes-​Moreno. 2017. Improvement
of chia seeds with antioxidant activity, GABA, essential amino acids, and dietary
fiber by controlled germination bioprocess. Plant Foods for Human Nutrition
72(4): 345–​352.
Gong, K., L. Chen, X. Li, L. Sun, and K. Liu. 2018. Effects of germination combined with
extrusion on the nutritional composition, functional properties and polyphenol
profile and related in vitro hypoglycemic effect of whole grain corn. Journal of Cereal
Science 83: 1–​8.
Grijalva-​Vallejos, N., A. Aranda, and E. Matallana. 2020. Evaluation of yeasts from
Ecuadorian chicha by their performance as starters for alcoholic fermentations in
the food industry. International Journal of Food Microbiology 317: 108462.
Guajardo-​ Flores, D., M. García-​ Patiño, D. Serna-​Guerrero, J.A. Gutiérrez-​ Uribe, and
S.O Serna-​Saldívar. 2012. Characterization and quantification of saponins and
flavonoids in sprouts, seed coats and cotyledons of germinated black beans. Food
Chemistry 134(3): 1312–​1319.
Guardianelli, L.M., M.V. Salinas, and M.C. Puppo. 2019. Chemical and thermal proper-
ties of flours from germinated amaranth seeds. Journal of Food Measurement and
Characterization 13: 1078–​1088.
Guo, H., A. Richel, Y. Hao, X. Fan, N. Everaert, X. Yang, and G. Ren. 2020. Novel
dipeptidyl peptidase-​ IV and angiotensin-​ I-​
converting enzyme inhibitory
peptides released from quinoa protein by in silico proteolysis. Food Science and
Nutrition 8: 1415–​1422.
Jan, R., D.C. Saxena, and S. Singh. 2018. Comparative study of raw and germinated
Chenopodium (Chenopodium album) flour on the basis of thermal, rheological
minerals, fatty acid profile and phytocomponents. Food Chemistry 269(15): 173–​180.
Jeske, S., E. Zannini, K. Lynch, A. Coffey, and E. Arendt. 2018. Polyol-​producing lactic acid
bacteria isolated from sourdough and their application to reduce sugar in a quinoa-​
based milk substitute. International Journal of Food Microbiology 286: 31–​36.
Jiménez, D., M. Lobo, B. Irigaray, M.A. Grompone, and N. Sammán. 2020a. Oxidative sta-
bility of baby dehydrated purees formulated with different oils and germinated grain
flours of quinoa and amaranth. LWT –​ Food Science and Technology 127: 109229.
Jimenez, D., M. Miraballes, A. Gámbaro, M. Lobo, and N. Samman. 2020b. Baby purees
elaborated with andean crops. Influence of germination and oils in physico-​chemical
and sensory characteristics. LWT –​ Food Science and Technology 124: 108901.
Jimenez, M.D., M. Lobo, and N. Sammán. 2019. 12th IFDC 2017 Special Issue Influence of
germination of quinoa (Chenopodium quinoa) and amaranth (Amaranthus) grains
on nutritional and techno-​functional properties of their flours. Journal of Food
Composition and Analysis 84: 103290.
References 397

Li, S., C. Chen, Y. Ji, J. Lin, X. Chen, and B. Qi. 2018. Improvement of nutritional value, bio-
activity and volatile constituents of quinoa seeds by fermentation with Lactobacillus
casei. Journal of Cereal Science 84: 83–​89.
Lushchak, V.I., and N.M. Semchuk. 2012. Tocopherol biosynthesis: Chemistry, regulation
and effects of environmental factors. Acta Physiologiae Plantarum 34: 1607–​1628.
Martínez-​Leo, E.E., and M.R. Segura-​Campos. 2020. Neuroprotective effect from Salvia
hispanica peptide fractions on pro‐inflammatory modulation of HMC3 microglial
cells. Journal of Food Biochemistry 44(6): e13207.
Micanquer, A., M. Cortes, and L. Serna-​Cock. 2020. Formulation of a fermentation sub-
strate from pineapple and sacha inchi wastes to grow Weissella cibaria. Heliyon
6: e03790.
Mirzaei, M., M. Aminlari, and E. Hosseini. 2016. Antioxidant, ACE-​inhibitory and anti-
microbial activities of Kluyveromyces marxianus protein hydrolysates and their
peptide fractions. Functional Foods in Health and Disease 6(7): 425–​439.
Mojica, L., E. Gonzalez de Mejia, M.A. Granados-​Silvestre, and M. Menjivar. 2017a.
Evaluation of the hypoglycemic potential of a black bean hydrolyzed protein iso-
late and its pure peptides using in silico, in vitro and in vivo approaches. Journal of
Functional Foods 31: 274–​286.
Mojica, L., D.A. Luna-​Vitala, and E. González de Mejía. 2017b. Characterization of
peptides from common bean protein isolates and their potential to inhibit markers
of type-​2 diabetes, hypertension and oxidative stress. Journal of the Science of Food
and Agriculture 97: 2401–​2410.
Montemurro, M., E. Pontonio, and C. Rizzello. 2019. Quinoa flour as an ingredient to
enhance the nutritional and functional features of cereal-​based foods. In Flour and
Breads and their Fortification in Health and Disease Prevention, edited by V. Preedy
and R. Watson, 453–​464. Cambridge, MA: Academic Press.
Morales, D., M. Miguel, and M. Garcés-​Rimón. 2020. Pseudocereals: A novel source of bio-
logically active peptides. Critical Reviews in Food Science and Nutrition ISSN: 1040–​
8398 (Print) 1549–​7852 (Online) Journal homepage: www.tand​fonl​ine.com/​loi/​
bfs​n20.
Mosso, A. L., J. LeBlanc, C. Motta, I. Castanheira, P. Ribotta, and N. Sammán. 2020. Effect
of fermentation in nutritional, textural and sensorial parameters of vegan-​spread
products using a probiotic folate-​producing Lactobacillus sakei strain. LWT –​ Food
Science and Technology 127: 109339.
Moughan, P.J., S.M. Rutherfurd, C.A. Montoya, and L.A. Dave. 2014. Food-​derived bioactive
peptides –​A new paradigm. Nutrition Research Reviews 27(1): 16–​20.
Mudgil, P., L.S. Omar, H. Kamal, B.P. Kilari, and S. Maqsood. 2019. Multi-​functional bio-
active properties of intact and enzymatically hydrolysed quinoa and amaranth
proteins. LWT –​ Food Science and Technology 110: 207–​213.
Nardo, A.E., M.C. Añón, and A.V. Quiroga. 2020. Identification of renin inhibitor peptides
from amaranth proteins by docking protocols. Journal of Functional Foods
64: 103683.
Nemzer, B., Y. Lin, and D. Huang. 2019. Chapter 3: Antioxidants in Sprouts of Grains.
Sprouted Grains –​Nutritional Value, Production, and Applications. Cambridge and
Washington, DC: Woodhead Publishing and AACC International Press.
Nongonierma, A.B., S.L. Maux, C. Dubrulle, C. Barre, and R.J. FitzGerald. 2015. Quinoa
(Chenopodium quinoa Willd.) protein hydrolysates with in vitro dipeptidyl
398 Nutritional Value Obtained from Latin-American Crops

peptidase IV (DPP-​IV) inhibitory and antioxidant properties. Journal of Cereal


Science 65: 112–​118.
Omary, M.B., C. Fong, J. Rothschild, and P. Finney. 2012. Effects of germination on the
nutritional profile of gluten-​ free cereals and pseudocereals: A review. Cereal
Chemistry 89(1): 1–​14.
Ontiveros, N., V. López-​ Teros, M.J. Vergara-​Jiménez, A.R. Islas-​Rubio, F.I. Cárdenas-​
Torres, E.O. Cuevas-​Rodríguez, C. Reyes-​Moreno, D.M. Granda-​Restrepo, S. Lopera-​
Cardona, G.I. Ramírez-​Torres, and F. Cabrera-​Chávez. 2020. Amaranth-​hydrolyzate
enriched cookies reduce the systolic blood pressure in spontaneously hypertensive
rats. Journal of Functional Foods 64: 103613.
Orona Tamayo, D., M.E.Valverde, and O. Paredes López. 2019. Bioactive peptides from
selected Latin American food crops -​A nutraceutical and molecular approach.
Critical Reviews in Food Science and Nutrician 59(12): 1949–​1975.
Padmashree, A., N. Negi, S. Handu, M. Khan, A. Semwal, and G. Sharma. 2018. Effect of ger-
mination on nutritional, antinutritional and rheological characteristics of quinoa
(Chenopodium quinoa). Defence Life Science Journal 4(1): 55–​60.
Pająk, P., R. Socha, J. Broniek, K. Królikowska, and T. Fortuna. 2019. Antioxidant proper-
ties, phenolic and mineral composition of germinated chia, golden flax, evening
primrose, phacelia and fenugreek. Food Chemistry 275: 69–​76.
Paucar-​Menacho, L.M., C. Martínez-​Villaluenga, M. Dueñas, J. Frias, and E. Peñas. 2017.
Optimization of germination time and temperature to maximize the content of
bioactive compounds and the antioxidant activity of purple corn (Zea mays L.) by
response surface methodology. LWT –​ Food Science and Technology 76: 236–​244.
Pérez-​Armendáriz, B., and G. Cardoso-​Ugarte. 2020. Traditional fermented beverages in
Mexico: Biotechnological, nutritional, and functional approaches. Food Research
International 136: 109307.
Piñuel, L., P. Boeri, F. Zubillaga, D.A. Barrio, J. Torreta, A. Cruz, G. Vásquez, A. Pinto,
and W. Carrillo. 2019. Production of white, red and black quinoa (Chenopodium
quinoa Willd Var. Real) protein isolates and its hydrolysates in germinated and
non-​ germinated quinoa samples and antioxidant activity evaluation. Plants
8(8): 257.
Repo-​Carrasco-​Valencia, R., and A.L. Serna. 2011. Quinoa (Chenopodium quinoa Willd.) as
a source of dietary fiber and other functional components. Ciência e Tecnologia de
Alimentos 31(1): 225–​230.
Rizzello, C.., A. Lorusso, M. Montemurro, and M. Gobbetti. 2016. Use of sourdough made
with quinoa (Chenopodium quinoa) flour and autochthonous selected lactic acid
bacteria for enhancing the nutritional, textural and sensory features of white bread.
Food Microbiology 56: 1–​13.
Rizzello, C., A. Lorusso, V. Russo, D. Pinto, B. Marzani, and M. Gobbetti. 2017. Improving
the antioxidant properties of quinoa flour through fermentation with selected
autochthonous lactic acid bacteria. International Journal of Food Microbiology
241: 252–​261.
Rueda, J. 2020. Actividad inmunomoduladora de hidrolizados proteicos de quínoa y su
aplicación a alimentos funcionales. Doctoral thesis. Faculty of Engineering of the
National University of Jujuy. Argentina.
References 399

Shi, Z., Y. Hao, C. Teng, Y. Yao, and G. Ren. 2019. Functional properties and adipogenesis
inhibitory activity of protein hydrolysates from quinoa (Chenopodium quinoa
Willd.) Food Science and Nutrition 7: 2103–​2112.
Silva-​Vieira, A.D., R. Bedani, M.A.C. Albuquerque, V. Biscola, and S.M. Isay-​Saad. 2017. The
impact of fruit and soybean by-​products and amaranth on the growth of probiotic
and starter microorganisms. Food Research International 97: 356–​363.
Suárez-​Estrella, D., G. Cardone, S. Buratti, M.A. Pagani, and A. Marti. 2020. Sprouting as
a pre-​processing for producing quinoa-​enriched bread. Journal of Cereal Science
96: 103111.
Świeca, M. 2016. Hydrogen peroxide treatment and the phenylpropanoid pathway
precursors feeding improve phenolics and antioxidant capacity of quinoa sprouts
via an induction of L-​Tyrosine and L-​Phenylalanine Ammonia-​Lyases activities.
Journal of Chemistry 2016(4): 1–​7.
Ujiroghene, O.J., L. Liu, S. Zhang, J. Lu, C. Zhang, J.L.X. Pang, and M. Zhang. 2019.
Antioxidant capacity of germinated quinoa-​based yoghurt and concomitant effect
of sprouting on its functional properties. LWT –​ Food Science and Technology
116: 108592.
Väkeväinen, K., F. Ludena-​Urquizo, E. Korkala, A. Lapveteläinen, S. Peräniemi, A. von
Wright, and C. Plumed-​ Ferrer. 2020. Potential of quinoa in the development
of fermented spoonable vegan products. LWT –​ Food Science and Technology
120: 108912.
Vallejo, J., P. Miranda, J. Flores Félix, F. Sánchez-​Juanes, J. Ageitos, J. González-​Buitrago, E.
Velázquez, and T. Villa. 2013. Atypical yeasts identified as Saccharomyces cerevisiae
by MALDI-​TOF MS and gene sequencing are the main responsible of fermentation
of chicha, a traditional beverage from Peru. Systematic and Applied Microbiology
36(8): 560–​564.
Vanegas, A., and L. Gutiérrez. 2018. Physicochemical and sensory properties of yogurts
containing sacha inchi (Plukenetia volubilis L.) seeds and β-​glucans from Ganoderma
lucidum. Journal of Dairy Science 101: 1020–​1033.
Vilcacundo, R., B. Miralles, W. Carrillo, and B. Hernandez-​Ledesma. 2018a. In vitro
chemopreventive properties of peptides released from quinoa (Chenopodium
quinoa Willd.) protein under simulated gastrointestinal digestion. Food Research
International 105: 403–​411.
Vilcacundo, R., C. Martínez-​Villalueng, B. Miralles, and B. Hernández-​Ledesma. 2018b.
Release of multifunctional peptides from kiwicha (Amaranthus caudatus) protein
under in vitro gastrointestinal digestion. Journal of Science of Food and Agricultural
99: 1225–​1232.
Villacrés, E., V. Allauca, E. Peralta, G. Insuasti, G., J. Álvarez, and M.B. Quelal. 2015.
Germination, an effective process to increase the nutritional value and reduce non-​
nutritive factors of Lupine Grain (Lupinus mutabilis Sweet). International Journal of
Food Science and Nutrition Engineering 5(4): 163–​168.
Villacrés, E., M.B. Quelai, E. Fernández, G. Garcia, G. Cueva, and C. Rosell. 2020. Impact
of debittering and fermentation processes on the antinutritional and antioxi-
dant compounds in Lupinus mutabilis sweet. LWT –​ Food Science and Technology
131: 109745.
400 Nutritional Value Obtained from Latin-American Crops

Wang, S., F. Zhu, and Y. Kakuda. 2018. Sacha inchi (Plukenetia volubilis L.): nutritional
composition, biological activity, and uses. Food Chemistry 265: 316–​328.
Xue, Z., C. Wang, L. Zhai, W. Yu, H. Chang, X. Kou, and F., Zhou. 2016. Bioactive compounds
and antioxidant activity of mung bean (Vigna radiata L.), soybean (Glycine max
L.) and black bean (Phaseolus vulgaris L.) during the germination process. Czech
Academy of Agricultural Sciences 34(1): 68–​78.
Zannini, E., S. Jeske, K. Lynch and E. Arendt. 2018. Development of novel quinoa-​based
yoghurt fermented with dextran producer Weissella cibaria MG1. International
Journal of Food Microbiology 268: 19–​26.
Zheng, Y., X. Wang, Y. Zhuang, Y. Li, H. Tian, P. Shi and G. Li 1. 2019. Isolation of novel
ACE-​inhibitory and antioxidant peptides from quinoa bran albumin assisted with
an in silico approach: Characterization, in vivo antihypertension, and molecular
docking. Molecules 24: 4562.
Chapter 11

Current Position of Legislation


on Latin-​American Grains and Its
Regional Socioeconomic Impact
María Dolores Jiménez1, Ana Laura Mosso1,
Claudia M. Haros2, and Norma Sammán1
1National University of Jujuy and National Council

for Scientific and Technical Research (CONICET),


San Salvador de Jujuy, Argentina
2Institute of Agrochemistry and Food Technology (IATA).

Spanish Council for Scientific Research (CSIC), Valencia, Spain

CONTENTS
11.1 Agro-​Economic Context in Latin and Mesoamerica 402
11.2 Production Systems and Farmers’ Situations 403
11.3 Diet Shifts and Food System Transformation 404
11.4 Access to Land and Gender Equity 405
11.5 Impact of Globalization on Economic Growth and the
Modernization of Production Chains: The Quinoa Boom 406
11.6 Nagoya Protocol 407
11.7 Food Legislation: The Global Situation 409
11.8 Crops’ Legal Framework in Latin and Mesoamerican Countries 410
11.9 Latin and Mesoamerican Crops’ Legal Framework in Importing
Countries 420
Conclusions 421
Acknowledgments 422
References 422

DOI: 10.1201/9781003088424-11 401


402 Legislation’s Impact on Latin-American Grains

11.1 AGRO-​ECONOMIC CONTEXT IN LATIN AND


MESOAMERICA
The Latin and Mesoamerican region is now experiencing a spectacular expan-
sion of its agricultural frontier. As a result, social, economic and ecological trans-
formation has occurred.
In the Central Andes of South America –​one of the eight most important
genocentres of plant species in the world –​about 8,000 years ago local
populations domesticated plants such as quinoa, amaranth, kiwicha and tarwi,
among others. These species reached our times due to the dedication and per-
severance of native communities, who preserved them as part of their cultural
heritage (Tapia, 1993). On the other hand, Mesoamerica, which encompasses the
lands from Mexico to Costa Rica, is also considered one of the most important
domestication centers of plant species in the world. Maize, tomato, cocoa, chia,
chili and common beans, among hundreds of others, are crops that originated in
this region (Pickersgill, 2007).
Traditionally, peasant communities maintained their own agricultural trad-
ition, taking care of the biodiversity of crops, ancestral technologies and the
proper conservation of environments for their production. In this sense, sev-
eral governmental research organizations, international research centers
and non-governmental organizations (NGOs) have collected, registered and
characterized the main species, and to a lesser extent those of restricted areas
and usage, despite their being of significant value for local economies and nutri-
tion. For instance, in Peru and Bolivia alone, estimates indicate that there are
more than 8,000 peasant communities –​heirs of the pre-​Hispanic Ayllus –​that
cultivate ancestral varieties despite technical assistance programs that seek to
impose cultivation of the so-​called improved varieties. There are crops which may
potentially be more productive, but in many cases they are not appropriate for
these agricultural systems or food sovereignty (Izquierdo and Roca, 1998).
Outside South and Mesoamerica, there are many underexploited crops that
are rapidly disappearing because of social unrest and environmental damage,
while others –​like quinoa and amaranth –​have crossed frontiers and are now-
adays found in Europe and the United States.
In recent decades, scientists have highlighted the benefits and new uses of
these crops to boost their global demand and generate economic opportunities
in regions where family incomes are far less than the minimum required for food
security.
The region is now one of the poorest in the world because of a decline in
farming, high rates of population growth, migration and misuse of natural
resources, despite once being the home of some of the most advanced cultures
11.2 Production Systems and Farmers’ Situations 403

at a global level (Izquierdo and Roca, 1998). According to a Food and Agriculture
Organization report, FAO (2020), the socioeconomic situation is critical and
hunger affects 42.5 million people in Latin America and the Caribbean (LAC),
6.5% of the regional population. The increase in hunger is closely related to
the general economic slowdown in the region. The fall of commodity prices
since 2011 has led to a deterioration in the public finances of many countries
dependent on the export of commodities.
In this sense, the member states of FAO in LAC agreed at the thirty-​sixth
regional conference on three major priorities to guide the actions of the organ-
ization during the 2020–​2021 biennium: (a) sustainable food systems to pro-
vide healthy diets for all; (b) hand-​in-​hand initiatives to achieve prosperous and
inclusive rural societies (c) sustainable and resilient agriculture. Furthermore,
the United Nations (UN) declared the period 2019–​2028 as the Decade of Family
Farming to bring a new perspective on what it means to be a family farmer in
a rapidly changing environment, highlighting its role in eradicating hunger
and building the future of food. Family farming provides 70% of the food in the
world -​reaching 97% in countries such as Peru –​and represents the basis of the
population’s food security.

11.2 PRODUCTION SYSTEMS AND FARMERS’


SITUATIONS
According to MINAGRI (2018), in Bolivia and Peru, most agricultural produ-
cers belong to family farming, about 40% are under the poverty line, 17% have
at least one unsatisfied basic need, 64% only reached first grade in elementary
school and 48% must work outside their farm since if they were dedicated only
to their plot, they would not have enough income to survive. This situation is
closely related to the fact that small-​scale producers have limited investment
capacity, low bargaining power and inadequate management of production
with regard to the demands of the market; and limited commercial knowledge
and low market opportunities. In addition, the scarce productive infrastructure
and road development, low coverage of technical assistance and training, poor
agricultural information services and few incentives for associativity aggravate
the situation.
In Mexico, small-​ scale producers form a very important agricultural
subsector. For instance, some crops produced in Puebla –​such as maize, chia,
and black common bean –​constitute a basic resource for the subsistence of rural
populations, generating jobs for communities and securing food for neighboring
404 Legislation’s Impact on Latin-American Grains

cities. About 74% of rural economic units are smallholdings of fewer than 5
hectares cultivated for subsistence or self-​consumption (Muñoz et al., 2019).
The practice of growing milpa (maize combined with common beans) is the
foundation of food security in many Guatemalan rural communities. Although
most peasants are aware of the potential to increase their returns from cash crops
or alternative economic activities, 99% of the households surveyed maintained
that the practice was important to their family food security. In this sense, the
contribution of milpa to food security represents much more than the nutrients
that it generates. It also guarantees that the basic sustenance needs of the family
are met (Altieri and Toledo, 2011).
Across all Latin and Mesoamerica, small-​scale farmers regularly integrate
rural trading systems, to sell or exchange their surplus production to acquire
other goods they do not produce. In these traditional systems, marketing
channels are not formally established and producers continually deal with sev-
eral problems, such as intermediaries, inadequate infrastructure for distributing
agricultural products and unequal terms between all the economic agents (pro-
ducers, collectors and merchants) who participate in the local, national and
international markets. The consequences of this informal trade are reflected
in the lack of official data on the volume of commercialization or on exchange
prices and profit margins in the value chains, and represent a disadvantage
mainly for primary producers; additionally it also carries phytosanitary risks
and difficulties with the maintenance of quality standards necessary for inter-
national trade (Ofstehage, 2011).

11.3 DIET SHIFTS AND FOOD SYSTEM


TRANSFORMATION
The transition from traditional value chains to modern food systems began in
the 1960s when governments set up grain and processed products parastatals
in wholesale, processing and retail. In the 1990s a phase of liberalization with
privatization and globalization of the food system started, which initiated rapid
investment by foreign firms. The improved infrastructure helped the prolifer-
ation of small and medium local enterprises (Pengue, 2004).
Five meta-​conditions have encouraged and facilitated food system changes
and diet shifts. These conditions are mutually dependent: income growth, policy
liberalization, infrastructure improvement, urbanization and the increase of
rural nonfarm employment. Regarding the last meta-​condition, nowadays in
Latin America 25% of the population lives in rural areas, and the majority of the
rural population lives near a city. Barbier and Hochard (2014) showed that less
than 10% of the rural population lives far from a town or a city, which implies
proximity to processed food stockists and large companies’ van networks.
11.4 Access to Land and Gender Equity 405

The shifts in dietary patterns in Latin and Mesoamerican communities are


the result of decades of changes in land tenure and local economies: liberal-
ization, rural-​urban migration and food aid programs, among other factors,
have affected the consumption of traditional foods. Over time, peasants have
transitioned from a diet based on what they produced on their land to one
that incorporates external foods from the market. Of course, diet is influenced
by socioeconomic status, however, the consumption of traditional crops is
influenced by history, migration, globalization and changing traditions. Dietary
habits are also shaped by desire for convenience and less labor-​intensive foods,
changing tastes, and the availability of other foods ( from imported products to
a greater diversity of cheaper local goods available for purchase). Once staples
for the region, legumes such as tarwi have declined as a critical part of the diet
over the past decade, representing less than 5 or 10% of the daily energy intake
(Popkin and Reardon, 2018).
In particular, in the Andes, diets of peasant families are also being influenced
by recent local and international efforts to promote quinoa products that are
changing the value and meaning of the ancestral crop in these communities.
Quinoa cultivated by Aymara farmers in Puno, Peru, has been traditionally
destined for home consumption, but peasants have been experiencing a shift
in their dietary habits in recent years (Bedoya-​Perales, 2018). Typical dishes
containing quinoa are still prepared (especially after harvest and for spe-
cial occasions), but the convenience and availability of other foods like rice,
wheat flour and pasta often make them more attractive, especially for younger
generations.

11.4 ACCESS TO LAND AND GENDER EQUITY


In the last decades, the lands of small peasants have decreased significantly. This
situation, according to an FAO report (2017), especially affects women, who only
own 8% of the land in Guatemala and 31% in Peru, properties that are usually
smaller and lower in quality. About 23% of Latin-​American lands are managed
or owned by indigenous people (Bose, 2017). From this situation, also in close
relation with traditional agroecology rooted in Andean culture, new small
farmer associations have emerged such as AOPEB (Asociación de Organizaciones
de Productores Ecológicos de Bolivia), founded in 1991 and composed of 75
organizations and about 70,000 families. In Peru, ANPE (Asociación de Productores
Ecológicos) has 12,000 members from 22 different regions of the country.
The revolution at the beginning of the 20th century in Mexico generated the
first agrarian reform on the continent, leaving a great part of the land, forests
and native germplasm in the hands of indigenous communities. Today, the
so-​called social property includes more than 100 million hectares. In addition, in
406 Legislation’s Impact on Latin-American Grains

Bolivia, a regulation of agro-​silvo-​pastoral production was promoted, in order to


protect indigenous rights on lands.
Gender equality is a pending debt and a continuous challenge in this region.
According to Katz (2002) a very small percentage of women (between 4 and
15%) in Chile, Colombia, Costa Rica, El Salvador, Honduras, Mexico, Nicaragua
and Peru have benefited from land distribution programs. An FAO report (2017)
indicated that the Gini coefficient, which measures gender inequality, applied
to the distribution of land in the region as a whole, reaching 0.79 in LAC, far
surpassing Europe (0.57), Africa (0.56) and Asia (0.55) (Bose, 2017).

11.5 IMPACT OF GLOBALIZATION ON ECONOMIC


GROWTH AND THE MODERNIZATION OF
PRODUCTION CHAINS: THE QUINOA BOOM
Scientists and development experts have especially taken an interest in the
potential of quinoa as a crop to feed the world, and have even promoted it as a
superfood throughout the western world (Bazile et al., 2015). Beyond its nutri-
tional profile, quinoa’s adaptability to different climates and growing conditions
makes it an important crop in respect of climate change. The UN declared 2013
as the International Year of Quinoa. This promoted cultivation areas mainly in
Bolivia, Peru and Ecuador, as well as increasing export volumes, which changed
the use of quinoa from a domestic food supply to being introduced to the global
market.
While quinoa prices rose, the majority of Andean producers depleted their
own consumption and sold their products, thus acquiring cheaper foods with a
lower nutritional value, such as noodles, wheat and rice (Kerssen, 2015).
The term quinoa boom refers to the globalization impact of this highly
nutritional grain that is expensively sold overseas to consumers interested in
their health. This phenomenon resulted in socioeconomic and environmental
impacts on indigenous and local communities. Initially, traditional cultivation
was conducted through manual labor, which was carefully carried out along
with nature in a sustainable way. When production of quinoa intensified, mech-
anical facilities were introduced with concomitant land degradation (Bedoya-​
Perales, 2018).
After 2013, commercial quinoa production started in Italy, India and China,
and intensified in the USA and Canada, where investment by market-​oriented
farmers with access to capital and technology increased. Moreover, quinoa is
cultivated in low-​altitude regions with conditions amenable to capital-​intensive
agricultural production. A similar situation occurred in the Andes when farmers
with no cultural link to quinoa but with capital to invest and productive farms
entered the market. On the other hand, Andean small farmers who had previously
11.6 Nagoya Protocol 407

monopolized quinoa production were then unable to compete in an even more


crowded market. Nevertheless, even with quinoa’s global popularity on the rise,
production overshot demand, and as a result, prices fell, and the economy of
many small farmers was crushed. The quinoa boom also created new challenges
in terms of environmental degradation, for example, manifesting as pesticide
contamination or drastic land-​use change from extensively used pastures to
intensively used agricultural land (Tschopp et al., 2018).
Equitable access to land is a human rights issue, and improving gender and
social inclusion in the land administration effectively enhances the sustainable
management of resources. Robinson et al. (2014) reported a particular case of
the collective land tenure system of indigenous women after the quinoa boom:
by Quechua women closer to Titicaca Lake, in Bolivia, who have an ancestral
tradition of land management including access to communal and family lands
which was affected after the quinoa boom. Traditionally, small-​scale community-​
based production was mainly used for household consumption. Since 2013, the
diversity of quinoa cultivation has been declining because the market demanded
white and red varieties. According to Bose (2017), Quechua women, who trad-
itionally bartered kañiwa and quinoa in the local market, started losing their
roles both as collective entrepreneurs and as individuals failing to barter quinoa
for household items. Furthermore, the quinoa boom has resulted in some forms
of land dispute, particularly communal land, which historically had been col-
lectively used for sowing. This was a practice of convenience to manage crops
rotationally and especially helped women support each other at times of vulner-
ability, due to drought or extreme frost. Nevertheless, after 2013, several changes
occurred in the way land was traditionally managed: the collective land has been
reclaimed and sold away as individual land to outsiders, and large-​scale white
and red quinoa cultivation has been introduced. Women continued supporting
the farming of quinoa crops, but decision making and marketing have been taken
over by men, mainly because men are the formal landowners, along with the fact
that most women only speak Quechuan and traders speak Spanish. Price vola-
tility of quinoa in the global market and individualization of land –​mostly by
men´s interest in land tenure –​were the main causes of vulnerability. Quechuan
women’s collective activities of protecting quinoa and kañiwa biodiversity have
been challenged due to an exclusively export-​driven orientation of mechaniza-
tion of agricultural production.

11.6 NAGOYA PROTOCOL
The Nagoya Protocol was adopted on October 29, 2010 in Nagoya, Japan, and
entered into force on October 12, 2014. Its objective is the fair and equitable
sharing of benefits arising from the utilization of genetic resources, thereby
408 Legislation’s Impact on Latin-American Grains

contributing to the conservation and sustainable use of biodiversity. The utiliza-


tion of genetic resources is understood as research and development
(R&D) into the genetic and biochemical composition of plant seeds, animals
or microorganisms. Access to genetic resources and the fair and equitable
sharing of benefits arising from their utilization to the Convention on Biological
Diversity (CBD, 2011) is an international agreement, which aims at sharing the
benefits arising from the utilization of genetic resources in a fair and equitable
way. This protocol helps to create greater legal certainty and transparency for

TABLE 11.1 STATUS OF RATIFICATION, ACCEPTANCE, APPROVAL OR


ACCESSION OF NAGOYA PROTOCOL IN IBERO-​AMERICAN COUNTRIES
Country Name Signed Ratification Party
Argentina 2011-​11-​15 2016-​12-​09 rtf 2017-​03-​09
Bolivia (Plurinational -​-​ 2016-​10-​06 acs 2017-​01-​04
State of)
Brazil 2011-​02-​02 2021-​03-​04 rtf 2021-​06-​02
Cuba -​-​ 2015-​09-​17 acs 2015-​12-​16
Dominican Republic 2011-​09-​20 2014-​11-​13 rtf 2015-​02-​11
Ecuador 2011-​04-​01 2017-​09-​20 rtf 2017-​12-​19
Guatemala 2011-​05-​11 2014-​06-​18 rtf 2014-​10-​12
Honduras 2012-​02-​01 2013-​08-​12 rtf 2014-​10-​12
Mexico 2011-​02-​24 2012-​05-​16 rtf 2014-​10-​12
Nicaragua -​-​ 2020-​06-​12 acs 2020-​09-​10
Panama 2011-​05-​03 2012-​12-​12 rtf 2014-​10-​12
Peru 2011-​05-​04 2014-​07-​08 rtf 2014-​10-​12
Portugal 2011-​09-​20 2017-​04-​11 apv 2017-​07-​10
Spain 2011-​07-​21 2014-​06-​03 rtf 2014-​10-​12
Uruguay 2011-​07-​19 2014-​07-​14 rtf 2014-​10-​12
Venezuela (Bolivarian -​-​ 2018-​10-​10 acs 2019-​01-​08
Republic of)
Colombia 2011-​02-​02 -​-​ -​-​ -​-​
Costa Rica 2011-​07-​06 -​-​ -​-​ -​-​
El Salvador 2012-​02-​01 -​-​ -​-​ -​-​
Source: CBD, 2020.
Notes: rtf: Ratification; acs: Accession; acp: Acceptance; apv: Approval; scs =​
Succession.
11.7 Food Legislation: The Global Situation 409

both providers and users of these resources. These establish more predictable
conditions for accessing genetic resources, ensuring benefit sharing when these
leave their providing country. The Nagoya Protocol also protects traditional
knowledge associated with the genetic resources covered by the CBD, and the
benefits arising from their utilization. Seed and plant collecting is still governed
by the laws of countries where the plants originate. However, the Nagoya
Protocol will ensure that anyone wishing to develop or research these plants
can access the relevant contracts to perform their labor. Not all countries have
signed the agreement. In this sense, as of December 2020, 128 parties have been
ratified, which includes 127 United Nations member states and the European
Union. Table 11.1 provides information on dates of signing and the status of
ratification, acceptance, approval or accession of the Nagoya Protocol in Ibero-​
American countries (CBD, 2020).

11.7 FOOD LEGISLATION: THE GLOBAL SITUATION


Legislation on crops and food products is necessary to ensuring the achievement
of nutritional quality and safety. Additionally, national and international
regulations and standards govern local and international trade, and guarantee
consumer protection (FAO/​W TO, 2018; OPS, 2020).
FAO and the World Trade Organization (WTO) provide governments with
the means to establish a framework that internationally facilitates agreements
on rules-​based food trade. Through the joint FAO/​WHO Codex Alimentarius
Commission, governments develop science-​ based food standards (FAO/​
WTO, 2018).
The Codex Alimentarius is the highest international body of food standards
where all 188 UN members and associated members have negotiated science-​
based recommendations in all areas related to food safety and quality (Codex
Alimentarius, 2020). The WTO considers the Codex Alimentarius the only
basis for harmonizing food safety measures in the context of international
food trade.
Furthermore, most countries –​or economic blocs, such as the EU or the
Southern Common Market (MERCOSUR), among others –​have established spe-
cific regulations based on Codex Alimentarius in accordance with the particular
situation of each country or bloc’s situation. In general, recommendations related
to hygiene during production, processing and food handling are included in
sanitary codes developed by health ministries. However, food standard codes are
generally formulated by independent government agencies specially designated
for this labor.
410 Legislation’s Impact on Latin-American Grains

11.8 CROPS’ LEGAL FRAMEWORK IN LATIN


AND MESOAMERICAN COUNTRIES
The development of novel foods with Latin and Mesoamerican seeds and grains
has recently achieved extraordinary levels. However, the lack of specific food
regulations for these crops hinders their production and commercialization.
Nevertheless, the regulatory framework for regional crops and their derivatives
has been improved in recent decades. An important advance in international
legislation on Andean crops was the approval of the first international standard
on quinoa in 2019 by the Codex Alimentarius Commission (CXS 333–​2019),
which was a decisive step toward promoting its production and consumption
throughout the world.
Regarding Latin-​American countries, several differences among laws and
regulations were found and are summarized in Table 11.2.
The Código Alimentario Argentino (CAA) [Argentinian Food Code] is an open-​
access technical regulation under constant update, which regulates all foods,
condiments, food additives and beverages, or their raw materials, that are
prepared, preserved and transported, as well as any person, commercial firm
or establishment that produces them. Therefore, it establishes the hygienic-​
sanitary, bromatological, quality and commercial identification standards.
CAA’s main objectives are to protect public health and to set legal frameworks
for commercial transactions (CAA, 2020). In recent years, CAA has incorporated
some regional crops such as amaranth, black turtle bean, chia, quinoa and chia
and quinoa flours.
Bolivia and Peru –​and especially since the quinoa boom –​have been
highlighted to have paid more attention due to the legislative framework
achieved regarding Andean crops. Bolivia has developed food standards
through the Instituto Boliviano de Normalización y Calidad (IBNORCA) [Bolivian
Institute of Standardization and Quality] –​a private non-​profit association.
All participants in the food supply chain, such as farmers, manufacturers or
retailers, can benefit from the guidelines and practices set out in the technical
standards, which include each link in the production chain, from food collection
harvesting to product packaging (IBNORCA, 2020). In Peru, the Normas Técnicas
Peruanas (NTP) [Peruvian Technical Standards], of the Instituto Nacional de
Calidad (INACAL) [National Institute of Quality], are documents that establish
test methods, sampling, packaging and labeling of foods, as well as the quality
specifications of products, processes and services (INACAL, 2020).
In Chile, Biblioteca del Congreso Nacional de Chile (BCN) [Library of the
National Congress of Chile] (BCN, 2020) and Reglamento Sanitario de los
Alimentos del Ministerio de Salud [Food Sanitary Regulations of the Ministry
of Health] (MINSAL, 2019) establishes the sanitary conditions that food must
meet during production, import, preparation, packaging, storage, distribution
newgenrtpdf
TABLE 11.2 LEGISLATION OF LATIN AND MESOAMERICAN CROPS IN LATIN-​AMERICAN COUNTRIES
Crops/​Products
Country Included Legislation Description Reference
Codex Alimentarius Quinua CXS 333-​2019 This standard defines quinoa (grain Codex

11.8
(International obtained from Chenopodium quinoa Alimentarius,
Food Standards) Willd.) and processed quinoa (quinoa 2020

Crops’ Legal Framework in Latin and Mesoamerican Countries


grains that have been subjected to
cleaning and removal of the pericarp
containing saponin). It defines the
requirements for human consumption
and includes the classification by color
and size.
Standard for CXS 153-​1985 This standard applies to maize (corn)
maize (corn) Adopted in 1985. Revised in for human consumption.
1995. Amended in 2019.
Whole maize CXS 154-​1985 This standard applies to whole
(corn) meal Adopted in 1985. Revised in maize (corn) meal for direct human
1995. Amended in 2019. consumption prepared from kernels of
common maize.
Argentina Amaranth. Chapter XI Amaranthus leaves for industrial use CAA, 2020
Leafy vegetables Vegetable foods. Article 822 that includes heat treatment and/​or
and Article 853 extrusion, having discarded the juices
produced in the process.
(continued)

411
newgenrtpdf
412
Legislation’s Impact on Latin-American Grains
TABLE 11.2 (CONTINUED)

Crops/​Products
Country Included Legislation Description Reference
Black turtle bean Chapter XI This classification includes fresh
Vegetable foods. Article 877 legumes of recent harvest and for
and Article 885 immediate consumption, and dried.
The name bean is understood as the
fresh or dried seed of the Phaseolus
vulgaris L. species, which includes the
black bean.
Chia seeds Chapter XI Denomination, requirements and
Vegetable foods. Article 917 classification of seeds for human
and Article 918 consumption.
Quinoa or Quinua Chapter XI
seeds Vegetable foods. Article 917
Amaranth flour Chapter IX Denomination and requirements
Farinaceous of seeds and seed flours for human
foods –​Cereals, flours and consumption.
derivatives. Article 660.
Quinoa or Quinua Chapter IX
seeds Farinaceous
foods –​Cereals, flours and
derivatives. Article 682.
Quinoa or Quinua Chapter IX
flour Farinaceous
foods –​Cereals, flours and
derivatives. Article 682 bis.
newgenrtpdf
Chia flour Chapter XIX Denomination and requirements
Flours, concentrates, of types of chia flour for human
isolates and derivatives consumption.
Protein. Article 1407 bis.
Chia oil Chapter VII Denomination and requirements of

11.8
Fatty foods, food oils. chia oil for industrial use.
Article 527 bis.

Crops’ Legal Framework in Latin and Mesoamerican Countries


Bolivia Amaranth grains Grain foods Definitions of amaranth grains. IBNORCA,
NB 336003:2005 2020
Amaranth grains Grain foods Classification and characteristics for
NB 336004:2006 amaranth to establish its class and
grade for marketing.
Cañahua grains Grain foods Definition of cañahua grains.
NB 336001:2004
Cañahua (or Grain foods Characteristics to establish the class
kañiwua) NB 336002:2005 and grade of cañahua grains for
Grains commercialization.
Chia seed Oil seed standards Classification and requirements for
NB 313025:2014 chia for human consumption.
Corn Cereals Classification and requirements of
NB 312008:2003 corn in grain for its commercialization
and industrialization.
Quinoa grains Cereal standard Energy determination of quinoa in
NB 312032:2006 grain.
(continued)

413
newgenrtpdf
414
Legislation’s Impact on Latin-American Grains
TABLE 11.2 (CONTINUED)

Crops/​Products
Country Included Legislation Description Reference
Raw corn flour Standard Flours and Characteristics and specifications
derivatives of raw corn flour for human
NB 583:1990 consumption.
Tarwi (chocho). Legumes Quality requirements and test
Bitter grain NB/​NA 0094:2011 methods for tarwi grain (chocho).
Tarwi (chocho). Leguminous standard Tarwi grain for marketing.
Debittered grain NB/​NA 0097:2011
Tarwi cookie Flour and its derivatives Standards for the development,
NB 39027:2009 characteristics and requirements of
tarwi products.
Tarwi queque Flour and derivatives
NB 39026:2009
Chile Chia seed oil Exempt Resolution 214 Authorizes the obtaining of edible oil BCN, 2020
from chia seed.
Quinoa seeds Exempt Resolution 9179 Establishes a specific standard for the
certification of quinoa seeds.
Sacha Inchi oil Exempt Resolution 528 Authorizes the obtaining of edible oil
from Plukenetia volubilis seeds.
Colombia Beans NTC 871:2005 Requirements for beans, including INCONTEC,
black bean, for human consumption. 2020
Corn NTC 2227:1986 Determination of moisture content in
corn grains, on milled grains and on
whole grains.
newgenrtpdf
Corn tortilla NTC 6173:2016 Quality requirements that must be
fulfilled and the test methods to
which the corn tortilla obtained
from nixtamalized corn for human
consumption must be submitted.

11.8
Quinoa flakes NTC 6071:2014 Requirements for quinoa flakes for
human consumption.

Crops’ Legal Framework in Latin and Mesoamerican Countries


Quinoa flour NTC 6069:2014 Requirements with which quinoa flour
must comply for human consumption.
Ecuador Amaranth grains NTE INEN 2646:2012 Quality requirements of amaranth INEN, 2020
grain for marketing and test methods
for the evaluation and verification of
these requirements.
Corn flour NTE INEN 1737:2016 Requirements for precooked cornmeal
precooked without germ.
without germ
Corn grains NTE INEN 187:2013 Requirements for grain corn, of
any variety, intended for human
consumption, zootechnical food and
industrial use.
Quinoa grains NTE INEN 1671:2013 Determination of infestation and
impurities levels.
Quinoa grains NTE INEN 1672:2013 Determination of saponins content
of quinoa grains by the foam method
(routine method).

(continued)

415
newgenrtpdf
416
Legislation’s Impact on Latin-American Grains
TABLE 11.2 (Continued)

Crops/​Products
Country Included Legislation Description Reference
Quinoa grains NTE INEN 1673:2013 Requirements for quinoa intended for
human consumption. It does not apply
to quinoa destined for seed.
Maize (milled NTE INEN-​ISO 6540 Maize. Determination of moisture
grains and on (identical translation to the content (on milled grains and on
whole grains) international standard ISO whole grains).
6540:1980)
Sacha inchi oil NTE INEN 2688:2014 Requirements that sacha inchi
oil must meet for human and/​or
industrial consumption.
Peru Amaranth grains NTP 205.054:2020 Good manufacturing practices for INACAL,
NTP 205.055:2017 processing plants and test methods. 2020
Amaranth NTP 011.461:2017 Good manufacturing practices for
expanded processing plants and test methods.
Cañihua grains NTP 011.452:2019 Requirements for obtaining products.
Quinua and NTP 011.453:2014
cañihua grains
Black turtle bean NTP 205.015:2016 Requirements for dry bean grain,
including black bean, for human
consumption.
newgenrtpdf
Cañihua flakes NTP 011.456:2015 Requirements for obtaining products.
Cañihua flour NTP 011.454:2015 Quality and safety requirements of oil
for direct human consumption.
Cañihua. Toasted NTP 011.455:2015
flour

11.8
Sacha inchi oil NTP 151.400:2019
Tarwi or chocho. NTP 205.090:2018 Requirements for debittering tarwi for

Crops’ Legal Framework in Latin and Mesoamerican Countries


Debittered grain direct human consumption or as raw
material for the food industry ( fresh
or dehydrated grain), and the test
method to determine alkaloids.
Uruguay Amaranth. Decreto N° 14/​013 Maximum limits of inorganic IMPO, 2020
Leafy vegetables contaminants in food.
Chia seeds Decreto N° 80/​019 Incorporation of chia in the list of raw
materials of vegetable origin used to
produce vegetable oils and fats.
Fatty acid composition for chia oil.
Corn Decreto N° 80/​019 Incorporation of seed germ of corn in
the list of raw materials of vegetable
origin used for the production of
vegetable oils and fats.
Fatty acid composition for corn oil.
Corn on the cob Norm no number Determination of phytosanitary
requirements for the introduction to
the country of corn on the cob (zea
mays).
Quinoa flour Decreto N° 315/​994 Description and specific provisions
for flour.

417
418 Legislation’s Impact on Latin-American Grains

and sale. This regulation includes the obtention and commercialization of chia,
quinoa, and sacha inchi oils. In addition, Resolution 5482 Exemption modifies
Resolution 2677, of 1999, which establishes import regulations for grains and
other products, intended for consumption and industrialization, including
amaranth imported from Mexico (BCN, 2020).
Normas Técnicas Colombianas (NTC) [Colombian Technical Standards]
of the Instituto Colombiano de Normas Técnicas y Certificación [INCONTEC]
(Colombian Institute of Technical Standards and Certification) –​a private
multinational organization –​offers free previews of the technical standards,
but full access must be paid for (INCONTEC, 2020). In Ecuador, Normas
Técnicas Ecuatorianas (NTE) [Ecuadorian Technical Standards] of the Servicio
Ecuatoriano de Normalización (INEN) [Ecuadorian Standardization Service] is
mandatory and with open access. Nevertheless, in some cases, standards are
adopted from other standardization bodies which are not available for free
download due to copyright (INEN, 2020). In recent years, in both Colombia
and Ecuador, some Latin and Mesoamerican crops (such as amaranth, beans,
corn and quinoa) and their by-​products (such as amaranth and quinoa flours,
precooked corn flour, quinoa flakes and sacha inchi oil) were included in the
regulations.
Uruguay has an open-​access Banco Electrónico de Datos Jurídicos Normativos
[Electronic Bank of Legal Normative Data] of the Dirección Nacional de
Impresiones y Publicaciones Oficiales (IMPO) [National Directorate of Prints and
Official Publications], which includes the national regulations and legal notices
(IMPO, 2020). The databank contains some requirements and specifications
for Latin and Mesoamerican crops and their products such as amaranth leafy
vegetables, chia seeds, corn and quinoa flour.
In Brasil, Agência Nacional de Vigilância Sanitária (ANVISA) [National Agency
of Health Surveillance] is a regulatory body linked to the Ministry of Health,
whose aim is to promote public health protection through sanitary control of the
production and commercialization of products and services subject to sanitary
regulation, including manufacturing, facilities and processes (ANVISA, 2020).
In Paraguay, Instituto Nacional de Alimentación y Nutrición (INAN) [National
Institute of Food and Nutrition] is technically responsible for the implementa-
tion and development of the National Food and Nutrition Plan that integrates
human, physical and administrative resources, in the areas of standardization,
laboratory control, medical care, food education and control of food and nutri-
tion outlets (INAN, 2020).
Paraguay also has the Instituto Nacional de Tecnología, Normalización y
Metrología (INTN) [National Institute of Technology, Standardization and
Metrology] with a catalog of standards available on the website (INTN, 2020).
Venezuela harmonized the national standards with Codex Alimentarius
standards. In addition, the Comisión Venezolana de Normas Industriales
11.8 Crops’ Legal Framework in Latin and Mesoamerican Countries 419

(COVENIN) made its transition to Fondo para la Normalización y Certificación


de la Calidad (FONDONORMA) which was created with the aim of supporting
the programs established by the Ministry of Public Works in the matter of
Standardization and Quality Certification. It was not possible to access the
online catalog of FONDONORMA (2020); and legislation of Andean products was
not found in the COVENIN database (COVENIN, 2020). However, even though
the database of regulations in these countries is quite complete, no regulations
were found for Andean crops or products. Possibly, their elaboration, control
and commercialization are legislated by non-​public regulations or are guided by
the Codex Alimentarius.
In Mexico, the Sistema Integral de Normas y Evaluación de la Conformidad
(SINEC) [Integral System of Standards and Conformity Assessment] contains the
Normas Mexicanas (NMX) [Mexican Standards], where test methods for some
Latin and Mesoamerican products are included; for example black turtle bean
(NMX-​FF-​038-​SCFI-​2013); chia seed oil (NMX-​F-​592-​SCFI-​2017); nixtamalized
corn flours (NMX-​F-​046-​SCFI-​2018); and popped amaranth grain (NMX-​FF-​116-​
SCFI-​2010) (Secretaría de Economía de Mexico, 2020).
In Panamá, Autoridad Panameña de Seguridad de Alimentos [Panamanian
Food Safety Authority] and Normas Nacionales para la Importación de Alimentos
([National Standards for Food Import] establishes an extension to present free
sale certificates and analyses of raw materials, ingredients and industrialized or
processed foods intended for human consumption (AUPSA, 2021). For example,
the standards establish requirements for the importation of amaranth from the
USA; chia from Argentina, Bolivia, Ecuador, Mexico, Nicaragua, Paraguay and
Peru; and quinoa from Bolivia, Colombia, Spain, the USA, Italy, Mexico, Paraguay
and Peru.
The Ministerio de Salud de Costa Rica [Ministry of Health of Costa Rica]
authorized Decree 27980, which sets the maximum level of aflatoxins in corn
and beans, among others (Ministerio de Salud de Costa Rica, 2021).
On the other hand, the Cámara Guatemalteca de Alimentos y Bebidas (CGAB)
[Guatemalan Chamber of Food and Beverages] was started by the action of
visionary and innovative companies to promote and enhance the union of the
country’s food and beverage industry; but nothing about Andean crops was
found in the database (CGAB, 2021). However, through Agreement 214–​2002 the
National Committee of the Codex Alimentarius of Guatemala was created, which
promotes the harmonization of national regulations on food safety and inter-
national trade with the standards, guidelines and recommendations defined
and established by the Codex Alimentarius. Likewise, Honduras and Nicaragua
adopted the Codex Alimentarius in March 1992 and through Decree 99–​2002,
respectively.
El Salvador and Belize have the Organismo Salvadoreño de Reglamentación
Técnica (OSARTEC) [Salvadoran Agency for Technical Regulation] and Belize
420 Legislation’s Impact on Latin-American Grains

Bureau of Standards as responsible for food legislation, respectively (OSARTEC,


2021; BBS, 2021) but nothing was found about Latin and Mesoamerican crops.

11.9 LATIN AND MESOAMERICAN CROPS’ LEGAL


FRAMEWORK IN IMPORTING COUNTRIES
The European Union (EU) is Latin America’s top investor and some of their
agreements include the Global Agreement with Mexico, Association Agreements
with Chile and the Central America countries, and Trade Agreement with Peru
and Colombia. Additionally, the EU maintains structured dialogues and regular
meetings with Latin-​American economic blocs (Andean Community, Mercosur,
Sistema de Integración Centroamericano [SICA]), which reflects the European
desire to support regional integration (EU, 2016).
The European Food Safely Authority (EFSA) was established as a source of sci-
entific advice and communication on risks associated with the food chain. EFSA
guidance documents are regularly updated, and the EFSA’s scientific advice
helps to protect consumers, animals and the environment from food-​related
risks (EFSA, 2020). The regulations and guidance documents in the nutrition
area include the regulatory framework on novel foods, in which some Andean
products are included, such as chia seeds, chia oil and sacha inchi oil.
Chenopodium quinoa and varieties of amaranth are mentioned in the EU
Novel Food Catalog (2021). Furthermore, Regulation EU 2015/​ 2283 of the
European Parliament and of the Council on novel foods has authorized the
commercialization of chia oil, establishing the labeling denomination of its
by-​products (such as fats and oils, puree and food supplements) (EU, 2017). In
addition, the resolution has set the labeling of chia-​contained products, such as
prepackaged chia seeds, bread, baked products, breakfast cereals, seed mixes,
fruit juices, and fruit/​vegetable blend beverages, fruit spreads and yogurt. In
addition, the EU Regulation 2020/​500, in 2020, authorized the placing on the
market of partially defatted chia seed powders as novel foods under Regulation
EU 2015/​2283 of the European Parliament and of the Council and amending EU
2017/​2470 (European Union, 2020). On the other hand, in the list of novel foods
of the EU 2017/​2470, sacha inchi oil was authorized to be placed on the market.
Food safety control in relation to quinoa as a food and food ingredient was
discussed and evaluated through an examination of the current statutory
provisions at UK and EU levels. Currently, neither the EU nor the UK requires
any specific legislation for quinoa if it complies with the provisions of the Food
Safety Act 1990 as amended and Regulation (EC) 178/​2002 on the general
principles and requirements of food law and procedures in matters of food safety
(Ojinnaka, 2016).
Conclusions 421

In the USA, the Guidance Industry 2018 –​Application of the Foreign Supplier
Verification Program Regulation to Importers of Grain Raw Agricultural
Commodities includes imported corn, amaranth and quinoa grains (Food and
Drug Administration (FDA), 2020).
The National Standard of China includes the following paid-​access legis-
lation: Milled quinoa (China Food Industry Standards LS/​T 32-​45-​2015) and
quality grading of the seeds (Amaranthus hypochondriacus L.) (China National
Standards GB/​T 26615–​2011) (China National Standards, 2020).
The Food Safety and Standards Authority of India (FSSAI) include definitions
and requirements for amaranth, chia seeds, and quinoa (FSSAI, 2020). On
the other hand, annexure of regulation 1-​ 1764/​FSSAI/​ Imports/​ 2018(Part-​
1) authorizes the importation of non-​transgenic food crops, including bean
and maize. Besides, a list of approved products/​ingredients under the Food
Safety and Standards (Approval for Non-​Specified Food and Food Ingredients
Regulations, 2017) mentions Inca Inchi (Plukenetia Volubilis) (23/​Std/​PA/​FSSAI/​
2018) of the Company Maxcure Nutravedics, a chia protein (26/​Std/​PA/​FSSAI/​
2019) and chia fiber powder (27/​Std/​PA/​FSSAI/​2019), both Company Amway
India Enterprises.
Food Standards Australia and New Zealand (FSANZ) has established that amar-
anth seeds meet the primary food authorization criteria and do not require fur-
ther evaluation, since no safety problems were identified after a preliminary risk
profiling exercise (Consultation Paper Proposed Future Regulation of Nutritive
Substances and Novel Foods in FSANZ, 2020). On the other hand, quinoa and
chia are mentioned in Schedule 20 (Proposal M1013 Schedule 20) which includes
maximum residue limits for agricultural and veterinary chemicals.

CONCLUSIONS
It is known that responsible trade promotes the development of regions and
favors socio-​economic growth. In particular, Latin and Mesoamerica are experi-
encing a spectacular expansion of its agricultural frontier. As a result, social, eco-
nomic and ecological transformations are occurring.
Because of this, it is necessary to find appropriate strategies that accompany
the development of the region in the framework of the modernization of pro-
ductive chains that the world is experiencing. Therefore, the experience of the
quinoa boom provides valuable information. In addition, the strategies must
consider the agricultural tradition of communities and what is appropriate for
local agricultural systems and food sovereignty.
Legislation is essential for proper product commercialization and, particu-
larly in regard to food, results vital to guarantee safety. Furthermore, treaties
422 Legislation’s Impact on Latin-American Grains

are necessary to ensure the sustainable use of genetic diversity and to protect
resources; many Ibero-​American countries adhered to the Nagoya Protocol, for
example, in pursuit of this objective.
This chapter presents the compilation of the legislation regarding Latin
American crops in different countries of Latin America and other regions of the
world. In recent years, advances in legislation allowed their commercialization
in international markets. Nevertheless, there is still a long way to go both for
the harmonization of existing laws and for the issuance of laws that accompany
the development of new foods made with Latin and Mesoamerican seeds and
grains.

ACKNOWLEDGMENTS
This work was supported by grant laValSe-​ Food-​
CYTED (Ref. 119RT0567);
Consejo Nacional de Investigaciones Científicas y Técnicas (CONICET) and
Universidad Nacional de Jujuy (Argentina) and Food4ImNut Food4ImNut
PID2019-107650RB-C21 funded by MCIN/AEI/10.13039/501100011033, Spain.

REFERENCES
Altieri, M., and V. M. Toledo. 2011. The agroecological revolution in Latin America: Rescuing
nature, ensuring food sovereignty and empowering peasants. Journal of Peasant
Studies 38(3): 587–​612.
ANVISA (Agência Nacional de Vigilância Sanitária). 2020. www.gov.br/​anv​isa/​pt-​br/​
setorr​egul​ado/​regula​riza​cao/​alimen​tos (accessed December 19, 2020).
AUPSA (Autoridad Panameña de Seguridad de Alimentos). 2021. www.aupsa.gob.pa/​
(accessed January 15, 2021).
Barbier, E., and J. Hochard. 2014. Poverty and the Spatial Distribution of Rural Population.
World Bank Policy Research Working Paper 7101. Washington, DC: World Bank.
Bazile, D., D. Bertero, and C. Nieto. 2015. State of the Art Report on Quinoa around the
World in 2013. Santiago de Chile: Montpelliere: Rome: FAO/​CIRAD. www.fao.org/​
583/​a-​i40​42e.pdf.
BBS (Belize Bureau of Standards). 2021. https://​bbs.gov.bz/​standa​rds-​catalo​gue/​
(accessed January 15, 2021).
BCN (Biblioteca del Congreso Nacional de Chile). 2020. www.bcn.cl/​leych​ile/​ (accessed
December 19, 2020).
Bedoya-​Perales, N., G. Pumi, M. Talamini, and A. Padula.2018. The quinoa boom in
Peru: Will land competition threaten sustainability in one of the cradles of agricul-
ture? Land Use Policy 79: 475–​480.
Bose, P. 2017. Land tenure and forest rights of rural and indigenous women in Latin
America: Empirical evidence. Women’s Studies International Forum 65: 1–​8.
References 423

CAA (Código Alimentario Argentino). 2020. Administración Nacional de Medicamentos,


Alimentos y Tecnología Médica (ANMAT) –​Ministerio de Salud. www.argent​ina.
gob.ar/​anmat/​codigo​alim​enta​rio (accessed December 19, 2020).
CBD (Convention on Biological Diversity). 2011. Nagoya Protocol on access to genetic
resources and the fair and equitable sharing of benefits arising from their utiliza-
tion to the convention on biological diversity, Secretariat of the Convention on
Biological Diversity, Montreal, Convention on Biological Diversity United Nations,
ISBN: 92-​9225-​306-​9, Quebec, Canada. www.cbd.int/​abs/​.
CBD (Convention on Biological Diversity). 2020. Parties to the Nagoya Protocol.
Convention on Biological Diversity. www.cbd.int/​abs/​nag​oya-​proto​col/​sign​ator​
ies/​, accessed December 10, 2020.
CGAB (Cámara Guatemalteca de Alimentos y Bebidas). 2021. http://​cgab.org.gt/​
(accessed January 15, 2021).
China National Standards. 2020. www.gbst​anda​rds.org/​ (accessed December 19, 2020).
Codex Alimentarius. 2020. www.fao.org/​fao-​who-​codexa​lime​ntar​ius/​es/​ (accessed
December 19, 2020).
COVENIN (Comisión Venezolana de Normas Industriales). 2020. www.senca​mer.gob.ve/​
senca​mer/​act​ion/​nor​mas-​find (accessed December 19, 2020).
EFSA (European Food Safety Authority). 2020. www.efsa.eur​opa.eu/​ (accessed December
19, 2020).
EU (European Union). 2016. EU trade relations with Latin America: Results and challenges
in implementing the EU-​Colombia/​Peru Trade Agreement. Doi: 10.2861/​056827.
EU (European Union). Official Journal of the European Union. 2017. https://​eur-​lex.eur​opa.
eu/​legal-​cont​ent/​EN/​TXT/​?uri=​celex%3A320​17R2​470.
EU (European Union). 2020. European Parliament and of the Council and amending
Commission Implementing Regulation EU 2017/​2470. https://​ec.eur​opa.eu/​food/​
saf​ety/​nov​el_​f​ood/​aut​hori​sati​ons/​union-​list-​novel-​foods​_​en (accessed December
19, 2020).
EU (European Union. 2021. Novel Food Catalogue. https://​ec.eur​opa.eu/​food/​saf​ety/​nov​
el_​f​ood/​catalo​gue/​sea​rch/​pub​lic/​index.cfm (accessed June 11, 2021).
FAO (Food and Agriculture Organization). 2020. The State of Food Security and Nutrition in
the World. www.fao.org/​3/​ca969​2en/​onl​ine/​ca969​2en.html.
FAO/​W TO (Food and Agriculture Organization)/​ (World Trade Organization). 2018.
Commerce and Standards Food. www.fao.org/​3/​i740​7es/​I740​7ES.pdf.
FDA (Food and Drug Administration). 2020. www.fda.gov/​food (accessed December
19, 2020).
FSANZ (Food Standards Australia New Zealand). 2020. www.foodst​anda​rds.gov.au/​
Pages/​defa​ult.aspx (accessed December 19, 2020).
FSSAI (Food Safety and Standards Authority of India). 2011. Fondo para la Normalización
y Certificación de la Calidad (FONDONORMA). www.fssai.gov.in/​upl​oad/​uplo​
adfi​les/​files/​Compendium_​Food_​Addit​ives​_​Reg​ulat​ions​_​08_​09_​2​020-​com​pres​
sed.pdf.
FSSAI (Food Safety and Standards Authority of India. 2020. www.fssai.gov.in/​cms/​about-​
fssai.php (accessed December 19, 2020).
IBNORCA (Instituto Boliviano de Normalización y Calidad). 2020. www.ibno​rca.org/​es
(accessed December 19, 2020).
424 Legislation’s Impact on Latin-American Grains

IMPO (Dirección Nacional de Impresiones y Publicaciones Oficiales). 2020. Banco


Electrónico de Datos Jurídicos Normativos [Electronic Bank of Normative Legal
Data]. www.impo.com.uy/​(accessed December 19, 2020).
INACAL (Instituto Nacional de Calidad). 2020. Normas Técnicas Peruanas (NTP)
[Peruvian Technical Standards]. www.ina​cal.gob.pe/​princi​pal/​catego​ria/​ntp
(accessed December 19, 2020).
INAN (Instituto Nacional de Alimentación y Nutrición). 2020. www.inan.gov.py/​site/​
(accessed December 19, 2020).
INCONTEC (Instituto Colombiano de Normas Técnicas y Certificación) [Colombian
Institute of Technical Standards and Certification]. 2020. Normas Técnicas
Colombianas (NTC) [Colombian Technical Standards]. https://​tie​nda.icon​tec.org/​
secto​res/​tec​nolo​gia-​de-​alimen​tos.html (accessed December 19, 2020).
INEN (Servicio Ecuatoriano de Normalización). 2020. Normas Técnicas Ecuatorianas
(NTE). http://​apps.normal​izac​ion.gob.ec/​desca​rga/​ (accessed December 19, 2020).
INTN (Instituto Nacional de Tecnología, Normalización y Metrología) [National Institute
of Technology, Standardization and Metrology]. 2020. http://​nor​mas.intn.gov.py/​
(accessed December 19, 2020).
Izquierdo, J., and W. M. Roca. 1998. Under-​utilized Andean food crops: Status and
prospects of plant biotechnology for the conservation and sustainable agricul-
tural use of genetic resources. ISHS Acta Horticulture 457: Symposium on Plant
Biotechnology as a Tool for the Exploitation of Mountain Lands.
Katz, E. 2002. La “feminización” de la economía rural en América Latina: evidencia, causas
y consecuencias. Rome: United Nations Food and Agriculture Organization (FAO)
Kerssen, T. 2015. Food Sovereignty: Convergence and contradictions condition and
challenges. Food sovereignty and the quinoa boom: Challenges to sustainable re-​
peasantisation in the southern Altiplano of Bolivia. Third World Quarterly 36(3).
MINAGRI (Ministerio de agricultura y riego). 2018. Viceministerio de políticas agrarias.
Dirección general de políticas agrarias. 2018. Manejo de granos andinos https://​
webca​che.google​user​cont​ent.com/​sea​rch?q=​cache:ITAXF​yRXG​a0J:https://​www.
mina​gri.gob.pe/​por ​tal/​anali ​sis-​econom​ico/​anali ​sis-​2019%3Fd​ownl​oad%3D14​
580:man ​ e jo-​ a gr​ o nom​ i co-​ d e-​ g ra​ n os-​ a ndi​ n os%26st​ a rt%3D20+​ & cd=​ 2 &hl=​ e s-​
419&ct=​clnk&gl=​ar.
MINSAL (Ministerio de Salud de la República de Chile). 2019. Reglamento sanitario de
los alimentos –​DTO. N° 977/​96. Publicado en el Diario Oficial de 13.05.97. https://​
dipol.min​sal.cl/​wp-​cont​ent/​uplo​ads/​2020/​11/​R SA-​DECR​ETO_​977-​96-​actu​aliz​
ado-​a-​noviem​bre-​2019.pdf,
Ministerio de Salud de Costa Rica. 2021. www.minist​erio​desa​lud.go.cr/​ (accessed January
15, 2021).
Muñoz, M. T., F. I. Ocampo, and I. F. Parra. 2019. Socioeconomic characterization of
the family production unit and the importance of the cultivation of chia (Salvia
hispanica L.) in the municipalities of Atzitzihuacán and Tochimilco, Puebla,
Mexico. Acta Universitaria 29: e2494.
Ofstehage, A. 2011. Nusta Juira’s gift of quinoa: Peasants, trademarks, and intermediaries
in the transformation of a Bolivian commodity economy. Anthropology of Work
Review 32(2): 103–​114.
References 425

Ojinnaka, D. 2016. Legislative control of quinoa in the United Kingdom and European
Union. Madridge Journal of Food Technology 1(1): 53–​ 57. doi: 10.18689/​
mjft.2016-​108.
OPS (Organización Panamericana de la Salud). 2020. Development of Food Legislation –​
Guidelines for the Development of Legislative and Executive Regulations in Food Control
Systems. www.paho.org/​hq/​index.php?opt​ion=​com_​cont​ent&view=​arti​cle&id=​
10708:2015-​des​arro​llo-​de-​la-​legi​slac​ion-​alimen​tos&Ite​mid=​41373&lang=​es.
OSARTEC (Organismo Salvadoreño de Reglamentación Técnica). 2021. www.osar​tec.gob.
sv/​(accessed January 15, 2021).
Pengue, W. 2004. https://​grain.org/​es/​arti​cle/​entr​ies/​413-​a-​short-​hist​ory-​of-​farm​ing-​in-​
latin-​amer​ica
Pickersgill, B. 2007. Domestication of plants in the Americas: Insights from mendelian
and molecular genetics. Annals of Botany 100(5): 925–​940.
Popkin, O., and T. Reardon. 2018. Obesity and food system transformation in Latin
America. Obesity Reviews 19(8): 1028–​1064.
Robinson, B., M. Holland and L. Naughton-​Treves. 2014. Does secure land tenure save
forests? A review of the relationship between land tenure and tropical deforest-
ation. Global Environmental Change 29: 281–​293.
Secretaría de Economía de Mexico. 2020. www.2006-​2012.econo​mia.gob.mx/​comuni​
dad-​negoc​ios/​normal​izac​ion/​catal​ogo-​mexic​ano-​de-​nor​mas (accessed December
19, 2020).
Tapia, M. 1993. Semillas andinas, el banco de oro. Lima: CONCYTEC.
Tschopp, M., S. Bieri, and S. Rist. 2018. Quinoa and production rules: How are cooperatives
contributing to governance of natural resources? International Journal of the
Commons 12(1): 402–​427.
Index

A oil extraction, 220–​221


in pasta processing, 99
Abiotic stress, 66 protein isolation, 221–​222
on crop yield, 58 proteins, techno-​functional properties
quinoa, 61 of, 222–​223
Above sea level (a.s.l.), 5 starches, 143, 223
Acid-​soluble whey fraction, 237 granules of, 143
Acrylamide, 185 modification, 224
Agência Nacional de Vigilância Sanitária physico-​chemical/​thermal/​
(ANVISA), 418 rheological properties of, 223
Agricultural production systems techno-​functional properties of,
diet shifts, 404–​405 222–​223
farmers’ situations, 403–​404 Amaranth foods, 271
food system transformation, 404–​405 Amaranth grains, 305
Agro-​biodiversity, 3 lipids of, 309
importance of, 2–​3 thin layer chromatography (TLC), 381
neglected and underutilized species Amaranth lipids, 220
(NUS), 3 Amaranth native starch, 224
Agro-​economic context, Latin and Amaranth oil, 220
Mesoamerican region, 402–​403 Amaranth protein, 389, 390
Albumins, 164, 323 Amaranth hypochondriacus protein
Alkaline cooking, 322 hydrolysates, 388
Alkaloid extraction, Andean lupine Amaranth hypochondriacus protein
(Lupinus mutabilis Sweet), 236 peptides, 391
Alkaloids, 17 Amaranth hypochondriacus proteins,
Amaranth (Amaranthus cruentus L./​A. 390
hypochondriacus L.), 22, 27, 55, emulsion formation/​stabilization of,
63–​65, 98–​100, 120, 260, 304 222
baking properties of, 169 Amaranth seed, 98, 222
commercial gluten free foods, 171 Amaranthus caudatus, 65, 219
cultivation, new regions of, 99–​100 Amaranthus tricolor, 63
present/​future uses, 98–​99 Amaranth wet grinding, 220
protein digestibility of, 381 Amino acids profile, 141
Amaranthaceae species, 11 anti-​nutritional compounds, 319–​320
Amaranth flour bioactive compounds, 317–​319
bread, characteristics of, 262–​263 modifications of, 322–​328
fiber extractions, 221 cooking methods/​malting, 329
milling, 219–​220 folate nutrient retention factors, 329

427
428 Index

nutritional composition, 311–​314 production, 28


technological/​culinary procedures, Andean soybean, 104
321 Andean weevil (Premnotrypes spp.), 17
vitamins/​minerals, 314–​316 Anthocyanidins, 318
γ-​aminobutyric acid (GABA), biosynthesis Anthocyanins, 317
of, 264 Anthracnose (Colletotrichum
Amylases, 224 gloeosporioides), 17
Amylose, in quinoa grains, 143, 213 Anti-​gliadin antibodies, 356
Andean crops, 362 Anti-​nutrient factors, 178
Andean grains, 163, 282, 316, 359, 375 Aranda Tarazona, J. J., 279
chemical composition, 345 Arginine, 314
effect of, 357 Arroz congris, 181
gluten-​free products, 167–​170 Asociación de Organizaciones de
kañiwa, processing of, 165–​166 Productores Ecológicos de Bolivia
nutrient profile, characterization of, (AOPEB), 405
358 Aztecs/​Mayans, in food preparations, 22,
quinoa, processing of, 163–​165 260
tarwi, processing of, 167
tumor cells, 352–​354 B
Andean lupine (Lupinus mutabilis Sweet),
14–​16, 74, 105, 137, 235, 236 Baked goods application, 183
alkaloids, 18 sacha inchi (Plukenetia volubilis L.), 183
extraction, 236 Banco Electrónico de Datos Jurídicos
botanical description, 16 Normativos, 418
demographic analysis, 15 Barley (Hordeum vulgare L.), 12
ecotypes, 74 Basal species (P. microcarpus, P. vulgaris),
environments, 16 123
future potential, 18–​19 Beans (Phaseolus vulgaris L.), 23, 26, 65,
growth of, 75 304
limitations, 17 Biblioteca del Congreso Nacional de Chile
nutritional point, 17 (BCN), 410
oil extraction, 236 Bifidobacterium sp., 280, 363
origin/​diversification/​domestication, BIOPEP database, 392
14–​15 Biscuits/​cookies/​cakes, 270–​273
residual flour, characterization of, Bitter lupine (L. albus and L. mutabilis),
236–​238 325
uses, 16–​17 Black turtle bean (Phaseolus vulgaris), 122,
Andean maize (Zea mays ssp. mays L.), 28, 143, 180–​182, 260, 304, 311, 342,
176, 218, 260 391
in bakery product formulations, 268 baked goods application, 183
future potential, 27–​28 in bakery product formulations, 268
genetic diversification/​cultivation, flours/​protein concentrate/​oil/​fiber,
26–​27 techno-​functional properties of,
gluten free foods, 177 238, 242
limitations, 27 gluten-​free products, 183–​185
origin/​distribution, 25–​26 lipids, 180
Index 429

milling process, 238–​240 nutrition/​health, Latin-​American


oil extraction, 241–​242 grains, 351–​356
processing methods/​industrial T-​cell lymphoma, 351
applications, 180–​181 Centro Internacional de Agricultura
production flow chart, 239–​241 Tropical (CIAT), 20
protein concentrates/​isolation, 240–​241 Cereal grains, 141
protein hydrolysates, in vitro anti-​ Cereals, 120
inflammatory activity of, 390 antioxidant activities, 375
protein isolates of, 387 zea mays, 142
proteins, 312 Chenopodium album, 9
Black turtle beans, 139 Chenopodium album L., 9
Black turtle starch, 144 Chenopodium genus, 9, 210
Blanco blandito, 120 Chenopodium pallidicaule, 58–​63
Blue maize, drawbacks of, 216 Chia (Salvia hispanica L.), 24, 55, 59,
Bolivia, quinoa, 6 70–​73, 100, 120, 122, 260, 304, 342
Bollo, 172 botanical description, 24
Bovine serum albumin (BSA), 211 cultivation, new regions of, 102–​103
Breeders, 102 future potential, 25
Broad bean (Vicia faba), 282 import value in 2020, 102
Buckwheat, baking properties of, 169 limitations, 24–​25
Mayan culture, 23
C origin/​distribution, 24
pH ranging, 102
Caco-​2 cells, 391 present/​future uses, 103–​104
Calose, 137 seeds, 22, 23
Cañahua (Chenopodium pallidicaule), 55, state of the art, 100–​101
58, 60 uses, 24
Cancer Chia crops, 22, 287
deaths worldwide, 346 Chia-​derived oil, 101
nutrition/​health, Latin-​American Chia flours, 103, 228, 275
grains, 346–​351 Chia, foods, 267
Cañihuaco, 208 Chia gel, 103
Carbohydrates, maize kernel, 142–​148 Chia grains, serine-​type protease
Carboxymethylation, 224 inhibitors (STPIs), 350
Cardinal-​McTeague, W. M., 30 Chia mucilage (CHM), 104, 232
Cardiovascular diseases (CVD), 25, extraction yield, 231
344 Chia nutlets, 22
nutrition/​health, Latin-​American Chia oil, 188
grains, 344–​345 Chia (Salvia hispanica L.) seeds, 102, 141,
Cas-​based systems (Clustered Regularly 185, 229, 307, 316, 388
Interspaced Short Palindromic conditioning/​oil extraction, 224
Repeats), 7 cold pressing/​solvent extraction,
Castilla bean (Vigna sinensis), 282 224–​226
Catechins, 319 flours (see Residual flours)
Celiac disease (CD) supercritical carbon dioxide
autoimmune digestive disorder, 351 extraction (CO2-​SE), 226–​228
430 Index

fiber rich fraction (FRF), 229–​231 production/​geographical distribution,


gluten, 185 21
gluten-​free foods, 186 Complementary School Food Program,
biscuits/​cookies/​cakes/​snacks, 13
190–​191 Consultative Group on International
bread products, 189–​190 Agricultural Research (CGIAR),
pasta products, 191 20
lyophilized hydrated, 187 Convention on Biological Diversity (CBD),
oil-​extracting processes of, 228 408
oil, use of, 310 Cookie formulations, with roasted chia,
pre-​packaged, 188 272
processing, 188–​189 Corn (Zea mays L.), 355
protein isolation, 229 Corn bran (CB), 276
protein rich fraction (PRF), 228–​229 Cotyledon, 137
Chia sourdoughs, 190 C-​reactive protein (hs-​CRP), 345
Chia, white and black-​spotted, 144 CRISP/​Cas9, 9
Chicha Morada, 216, 324 Crop development, biodiversity-​based
Chihaan, 23 strategy of, 2
Chinatoles, 281 Crop production systems, 5, 54
Chocho (Lupinus mutabilis), 308; see also Crude albumins, denaturation
Andean lupine temperatures of, 229
Chullpi maizes, 134, 140 Cytotoxicity, 57
Chymotrypsin, 387
Cianidine-​3-​Glucoside, 106 D
Cizeikiene, D., 355
Climate change, 56, 92 Dairy substitutes, 168
agricultural challenges, 54 Debittering, aqueous thermal treatment,
agriculture threatens, 54 167
seed quality, 54 Degree of milling (DOM), 163
Clostridium spp, 363 Delphinidin-​3-​O-​β-​D-​glucoside, 319
Cobalto-​60 drums, 13 Detoxifying, 109
Codex Alimentarius, 409, 410 D-​galacturonic acid monomers, 358
Código Alimentario Argentino (CAA), 410 Diabetes-​associated renal fibrosis, 343
CO2 exchange rates, 57 Diabetes mellitus
Comisión Venezolana de Normas metabolic disease, 342
Industriales (COVENIN), 419 nutrition/​health, Latin-​American
Common bean (Phaseolus Vulgaris), 20, grains, 342–​344
108–​109, 123, 309 type 2 diabetes, 342
agronomy, 21 Diacetyl tartaric acid ester of mono-​and
botanical description, 20 diglycerides (DATEM), 191
breeding, 22 Dietary fiber (DF), 221
germplasm collections, 19–​20 Diet shifts, agricultural production
limitations, 21 systems, 404–​405
nutrition, 21–​22 Dipeptidyl peptidase IV (DPPIV)
origin/​diversification/​domestication, hormones degradation, 391
19 metabolic enzyme, 391
Index 431

2,2-​Diphenyl-​2-​picryl-​hydrazyl (DPPH), Fermentation, 372, 373


62, 375 antinutritional factors, reducing, 376
DNA molecular markers, 66 lactic acid bacteria (LAB), 373–​375
Dough viscosity, 168 maize, traditional beverages, 377
Drought matrix structural changes, 377–​378
antioxidant capacity, 56 nutritional value, from Latin-​American
photosynthetic rates, 56 crops, 373
salinity treatments, 65 reducing antinutritional factors, 376
stress, 56, 64 Fermentation process, 183
water shortage, 69 Fermented beverages, 279–​281
Dry milling, 166, 237 Fiber extractions
amaranth flour, 221
quinoa/​kañiwa, 210–​211
E
Fiber rich fraction (FRF), 229
Economic growth, globalization, 406–​407 antioxidant activity, 229
Economic slowdown, 403 chia mucilage (CM), 230
Edible oils, from plant sources, 241 chia seeds, 229–​231
Egg-​free pasta, 169 Flavonoids, 60
Embryo, 133 Flowering genotypes, 72
Endosperm, 133 Foaming capacity (FC), 232
Enzymatic hydrolysis, 386 emulsifying properties, 232
Enzyme catalase, 386 for quinoa proteins, 211
Escherichia coli, 391 Foam stability (FS), 232
Escherichia/​Shigella, 362 Folates
Essential fatty acids (EFA), 309 Amaranth, 329
Euphorbiaceae (Plukenetia volubilis L.), cooking methods and malting, 329
30, 73 in quinoa, 328, 329
European Food Safety Authority (EFSA Fondo para la Normalización y
2020), 31, 186, 268, 420 Certificación de la Calidad
European Union (EU) (FONDONORMA), 419
seed oil, 31 Food and Agriculture Organization (FAO),
Trade Agreement, 420 9, 386
Exo-​polysaccharides (EPS) production, 378 Food and Agriculture Organization of the
Extrusion technology, 174 United Nations (FAOSTAT), 21, 92,
282
Food competition tables (FCT), 321, 330
F
Food legislation, on Latin-​American grains
Family Euphorbiaceae (Plukenetia volubilis agro-​economic context, in
L.), 73 Mesoamerica, 402–​403
FAO report (2017), 405 crops’ legal framework, 410–​420
Fat localization, 164 diet shifts/​food system transformation,
Fat-​soluble vitamins, 314 404–​405
Fatty acids (FA) economic growth, globalization on,
in Latin-​American grains, 310 406–​407
n-​3/​n-​6, 281, 374 farmers’ situations, 403–​404
nutritional composition, 309–​311 global situation, 409
432 Index

importing countries, 420–​421 Genetic diversity, 6


land/​gender equity, 405–​406 Genetic erosion, 13
Nagoya Protocol, 407–​408 Genotype and environment (GX E), 71
production chains, modernization of, Germination, 372, 382
406–​407 amino acids profile, 378–​381
production systems, 403–​404 antinutrients/​mineral bioavailability,
quinoa boom, 406–​407 383–​384
Food pattern, 356 antioxidant content/​capacity, 382–​383
Food protein molecules, 240 carbohydrates profile, 382
Food Safety and Standards Authority of conditions, 378–​380
India (FSSAI), 421 fatty acids’ profile, 381–​382
Food safety control, 420 germinated/​non-​germinated amaranth
Food security flours, thermal properties of, 385
ancient techniques, 5–​6 lipid content, 381–​382
human nutrition, 3 protein content, 378–​381
maize production, 27 protein digestibility, 378–​381
Food Standards Australia and New protein profiles, 378
Zealand (FSANZ), 421 rheological/​textural characteristics,
Food system transformation, agricultural 384–​385
production systems, 404–​405 sensory characteristics, 385–​386
Food uses sprouted quinoa and amaranth flours,
bakery/​bread products, 261–​270 386
biscuits, cookies and cakes, 270–​273 technological changes, 384
drinks/​fermented beverages, 279–​281 thermal behaviors, 385
haute cuisine, 287–​288 Germplasm, 74
infant foods, 281–​283 Germplasm Resources Information
overview of, 260–​261 Network (GRIN), 19
pasta products, 273–​276 Giant Cuzco, 134
snacks/​breakfast mixtures, 276–​279 Gini coefficient, 406
traditional foods, 283–​287 Global Crop Trust, 20
Fourier-​transform infrared spectroscopy Globalization, on economic growth,
(FT-​IR), 220 406–​407
Fruit, color changes, 135 Global quinoa production, 10
Globulins, 140, 164, 221, 323
G Glutelins, 140
Gluten-​free applications, Latin-​American
Gallic acid equivalent (GAE), 376 crops, 161–​192
Gamma radiation, 13 Gluten-​free diet (GFD), 355
Gelatinization, 177 Gluten-​free products, 179, 184
endosperm, 170 Andean grains, 167–​170
enthalpy, 212 with Andean maize, 177
lipids, 170 bakery products, 163
of starch, 324 black turtle bean, 183–​185
Gender equality, 406 bread making, 190
Genetically modified (GM) maize pollen chia foods, 186
fluxes, 28 cookies, 190
Index 433

Gluten free grains (Amaranthus Inchi (Plukenetia volubilis), 55


hypochondriacus, Amaranthus Indian food, 7
caudatus, Amaranthus cruentus), Infant foods, 281–​283
63 Inflammatory disorders, 25
maize (Zea mays L.), 178–​179 Insoluble dietary fiber (IDF), 242
sacha inchi (Plukenetia volubilis L.), Institute of Agricultural Science and
183–​185 Technology (ICTA), 20
spaghetti, 273 Instituto Colombiano de Normas Técnicas y
Gluten proteins, 164 Certificación (INCONTEC), 418
Glycemic index, 168 Instituto Nacional de Alimentación y
Glycosides, 359 Nutrición (INAN), 418
Grain filling, 69 Instituto Nacional de Calidad (INACAL),
Guatemala, 21 410
Guidance Industry 2018–​Application of Instituto Nacional de Tecnología,
the Foreign Supplier Verification Normalización y Metrología (INTN),
Program Regulation to Importers 418
of Grain Raw Agricultural Insulin-​induced adipocytes, 391
Commodities, 421 International Center of Tropical
Gut microbiota, andean grains dietary Agriculture (CIAT), 66
intervention, 360–​361 International Year of Quinoa, 406
Inter-​simple sequence repeat (ISSR), 16
H Intestinal microbiota, 359
Isoleusine, 314
Hard-​to-​cook (HTC) defect, 109 Isozymes, 66
Haute cuisine, 287
Helicobacter pylori, 351
K
Heptapeptide, 389
High temperature and short time (HTST) Kañawa (Chenopodium pallidicaule
treatment, 320 Aellen)
Histidine, 314 breeding needs, 13
Hot paste viscosity (HPV), 212 domestication/​cultivation history,
Hydration properties, 211 11–​12
Hydrolyzed amaranth proteins, anti-​ future potential, 13–​14
inflammatory activity of, 389 genetic improvement, 13
13-​Hydroxilupanine, 74 geographical distribution, 13–​14
4-​Hydroxylupanin, 235 Latin-​American crops, 10–​14
13-​Hydroxylupanin, 235 limitations, 13
Hydroxypropylation, 224 plants, 11
Hydroxypropylmethylcellulose gum plants, in field, 11
(HPMC), 190 uses, 13
Hypocalcemia, 355 Kañiwa (Chenopodium pallidicaule),
107–​108, 120, 163, 260, 266, 304,
I 307, 342
commercial products of, 175
Ibero-​American countries, Nagoya milling
Protocol, 408 dry, 208
Inca peanut, 309 wet, 209–​210
434 Index

oil extraction botanical classification of, 124–​125


cold, 209 edible plant species, 92
solvent, 209 food production, 93
protein composition, 210 gluten-​free applications, 161–​192
protein/​fiber isolation, 210 gross structural features, 126–​127
fiber extraction methods, 210–​211 legislation of Latin /​Mesoamerican
starch, physicochemical/​thermal/​ crops, 411–​417
rheological properties of, 212–​216 maize, races of, 121
techno-​functional properties, 211 nutritional composition/​bioactive/​
Kañiwa (Chenopodium pallidicaule anti-​nutritional compounds (see
Aellen), 208 Nutritional composition)
Kañiwa flours, 324 physical properties, 127–​133
phytate content, 324 proteins, 389
seeds, 165 pseudo-​cereals of, 122
Kañiwa starches, 143, 212 quinoa (see Quinoa (Chenopidium
Kernel shrinkage, 57 quinoa Willdenow))
Kiwicha (Amaranthus caudatus), 55, 59, structure/​composition, 119–​150
98–​100, 163, 342 typical dishes, 284–​286
cultivation, new regions of, 99–​100 World Population 2019: Wall Chart
present/​future uses, 98–​99 (ST/​ESA/​SER.A/​434), 93
Kiwicha (amaranth) cultivation, 99 Latin-​American grains
Kiwicha–​potato starch formulation, 169 amino acids profile of, 313
Kona Kona (Eurysacca quinoae P.), 17 proximate composition of, 306
vitamins and minerals, 315
L Latin-​American legumes, 144
Latin/​Mesoamerican seeds
Lactic acid, 165 crops’ legal framework, 410–​420
Lactic acid bacteria (LAB), 373–​375 importing countries, 420–​421
Lactobacillus acidophilus, 280 legislation of, 411–​417
Lactobacillus fermentum, 375 Least gelatinous capacity (LGC), 242
Lactobacillus fermentum FUA3165, 377 Lectins, 320
Lactobacillus plantarum, 264, 375, 376 Legumes, 312
T0A10, 376 anti-​nutritional factors, 320
Lactobacillus sanfranciscensis W2, 190 crop, gastronomic versatility/​
Lactobacillus spp., 363 nutritional qualities, 18
Lactococcus lactis subsp. lactis, 264 family (Leguminosae), 123
Land/​gender equity, 405–​406 flours, 275
Latin America Leguminous species (Lupinus mutabilis
Andean region, 3–​5 Sweet), 104
crops limitations, 7 Leptin, 345
crop species, 4 (see also Latin-​American Leucine, 314
crops) LIB222 genotype, 75
Seed quality, 53–​76 Lignin, 137
Latin America and the Caribbean (LAC), Linoleic acid, 73, 374
403 α-​Linolenic acid (ALA), 73, 149
Latin-​American crops, 59, 283 Lipid carbonyls, 185
Index 435

Lipids nixtamalization, 170–​172


black turtle bean, 180 toasting, 178
Latin-​American maize, 148 proteins
maize kernel, 148–​150 fractions, 217
sacha inchi (Plukenetia volubilis L.), 180 lysine, 311
sacha inchi seeds, 180 techno-​functional properties of, 218
Lipophilic antioxidants, 383 tryptophan, 311
Listeria monocytogenes, 391 proximal composition of, 322
Low density lipoprotein (LDL), 220 starch
amaranth oil, 220 lipid extraction, 217
cholesterol modulation, 344 physico-​chemical/​thermal/​
Lowest gelation concentration, 242 rheological properties, 218–​219
Lupanine, 74 techno-​functional properties of,
Lupine (Lupinus albus L), 282, 355 218
flour, 236 vitreousness-​and hardness-​associated
flour, use of, 169 properties, 216
sandy loam soils, 16 Maize β-​carotene, 314
seeds, 17 Maize kernel, 133
Lupino andino (Lupines mutabilis), 55 chemical composition of, 139
Lupin protein, 169 carbohydrates, 142–​148
Lupinus (L. albus, L. angustifolius and lipid, 148–​150
L. cosentinii), 142, 344 proteins, 140–​142
Lupinus ananeanus Ulbr., 15 structures, 133–​139
Lupinus mutabilis, 74–​75, 123 Malvidin-​3-​O-β-​D-​glucoside, 319
Lupunine, 17 Mazamorra Morada, 216, 324
Lyophilized hydrated chia, 187 Mesoamerican maize races, 218
Lysine, 64, 314 Metabolic fingerprinting, 72
A. hypochondriacus, 312 Metal chelating capacity-​ferrozine, 387
maize protein, 311 Methionine, 314
Microbiota modulation
M metabolites, 362
nutrition/​health, Latin-​American
Maize (Zea mays), 55, 59, 68–​70, 106–​107, grains, 356–​363
162, 170, 304 prebiotic and probiotic effects, 342,
alkaline cooking of, 170 356
breads, 179 Millet crops (Panicum miliaceum L.), 27
fiber, 217–​218 Milling process
techno-​functional properties of, 218 amaranth flour, 219–​220
flours, techno-​functional properties of, black turtle bean, 238–​240
218 quinoa/​kañiwa
gluten-​free products, 178–​179 dry, 208
grain characteristics, 216 wet, 209–​210
Latin-​American region, 121 types of, 167
processing of Milpa, 19, 403
extrusion technology, 174–​178 Mineral absorption, in humans, 376
milling, 172–​174 Ministerio de Salud de Costa Rica, 419
436 Index

Ministry of Agriculture and Irrigation of macronutrient content, 304–​309


Peru (MINAGRI), 100, 403 Nutritional value, from Latin-​American
Mint (Mentha L. sp.), 24 crops
Molecular weights (MW), 234 bioactive peptide production
Monounsaturated fatty acids (MUFAs), antiadipogenic activity, 391
149 anticancer activity, 392
Mucilage, 189 antidiabetic activity, 391
Mushrooms, 139 antihemolytic activity, 390
anti-​inflammatory activity, 389–​390
N antimicrobial activity, 390–​391
antioxidant activity, 388–​389
Nagoya Protocol, 407–​409 antithrombotic activity, 391–​392
genetic resources, 409 cytotoxicity, 389
in Ibero-​American countries, 408 enzymatic hydrolysis, 386–​388
National Botanical Research Institute immunomodulatory activity, 392
(NBRI), 64 in silico approach/​molecular
National Plant Germplasm System of the docking, 392–​393
United States of America (NPGS), fermentation, 373
19 maize, traditional beverages, 377
Native starches, physicochemical/​techno-​ matrix structural changes, 377–​378
functional properties of, 213–​215 reducing antinutritional factors, 376
Natural sachi inchi snack, 278 germination
Nazca culture, 14 amino acids profile, 378–​381
Neglected and underutilized species antinutrients/​mineral bioavailability,
(NUS), 3 383–​384
Nixtamalization processes, 26, 170, 172, antioxidant content/​capacity,
322, 323 382–​383
maize flours, 172 carbohydrates profile, 382
wet milling of, 172 conditions, 378
Noncommunicable diseases, clinical and fatty acids’ profile, 381–​382
preclinical studies, 347–​349 lipid content, 381–​382
Normas Técnicas Colombianas (NTC), 418 protein content, 378–​381
Normas Técnicas Ecuatorianas (NTE), 418 protein digestibility, 378–​381
Normas Técnicas Peruanas (NTP), 410 rheological/​textural characteristics,
Novel food, 186, 187 384–​385
European Community (EC), 268 sensory characteristics, 385–​386
nutritional ingredients, 268 technological changes, 384
Nuclear magnetic resonance (NMR) thermal behaviors, 385
spectroscopy, 72 overview of, 372
Nutlets, gelatinous mass of, 24 vitamin production/​antioxidant
Nutrient retention factors (NRF), 321 compounds, 375–​376
food composition tables, 321 Nutrition/​health, Latin-​American grains
moisture and mineral content, 327–​328 Andean crops, health effects, 342
Nutritional composition cancer, 346–​351
amino acids profile, 311–​314 cardiovascular risk factors, 344–​345
fatty acids profile, 309–​311 celiac disease (CD), 351–​356
Latin-​American grains, 304 diabetes mellitus, 342–​344
Index 437

microbiota modulation, 356–​363 antidiabetic activity, 391


obesity, 345–​346 antihemolytic activity, 390
prebiotic effect, 356–​363 anti-​inflammatory activity, 389–​390
antimicrobial activity, 390–​391
antioxidant activity, 388–​389
O antithrombotic activity, 391–​392
Oats (Avena sativa L.), 12 cytotoxicity, 389
Obesity enzymatic hydrolysis, 386–​388
Andean grains, 345 immunomodulatory activity, 392
cardiovascular risk factors, 342 in silico approach/​molecular docking,
noncommunicable diseases, 342 392–​393
nutrition/​health, Latin-​American Perisperm, 138
grains, 345–​346 Peruvian Andean maize, 217
Octenylsuccinate quinoa starch (OSQS), Peruvian-​Bolivian-​Chilean altiplano, 4
97 Petunidin-​O-​β-​D-​glucoside, 319
Octenyl succinylation, 224 PHA-​0683 accession, 67
Oil chia seed extraction, 232 Phaseolus vulgaris, 20, 108
Oil extraction processes, 105 Phaseolus vulgaris L. (beans), 65–​68
amaranth flour, 220–​221 Phenols, 60, 137
Andean lupine (Lupinus mutabilis Phenylalanine, 314
Sweet), 236 Phosphorylation, 224
black turtle bean, 241–​242 Phytic acid, 319, 376
quinoa/​kañiwa Phytosterols, 318, 374
cold, 209 Pineapple, agro-​industrial waste of,
solvent, 209 374
Sacha inchi (Plukenetia volubilis L.), 233 Plantago psyllium seed, 230
Oil-​holding capacity (OHC), 232 Plant seeds, research and development
Oil recovery, 225 (R&D), 408
Oilseeds, 189 Plukenetia huayllabambana, 233
Oleaginous seeds, 122, 141, 149 Plukenetia volubilis, 29–​31, 233
Omega-​6, 185 Polyphenols, 316, 358
Omega-​3 fatty acid, 71, 72, 104, 185 Polysaccharides, 137
Oregano (Origanum vulgare L.), 24 Polyunsaturated fatty acids (PUFAs), 71,
Organismo Salvadoreño de Reglamentación 149, 310, 344
Técnica (OSARTEC), 419 Poor man’s food, 7
ORURO chia seeds, 103 Potato (Solanum tuberosum L.), 27
Oxalic acid, 319 Pozol, 377
Prebiotic effect, nutrition/​health in
Latin-​American grains, 356–​363
P Pre-​Inca cultures, ethnic diversity of, 5
Pasta products, 273–​276 Pressed cakes, 182
Pea (Phaseolus vulgaris), 55 Proanthocyanidins, 318
Pectin, 137 Production chains, modernization of,
PepBank database, 392 406–​407
Peptide production Prolamines, 323
antiadipogenic activity, 391 Protein isolation methods
anticancer activity, 392 amaranth flour, 221–​222
438 Index

chia seeds, 229 Quinoa (Chenopidium quinoa Willdenow),


quinoa/​kañiwa, 210 94
Protein–​protein interaction, 211 overview of, 94
Protein rich fraction (PRF), 228, 229 present/​future uses, 95–​97
chia seeds, 228–​229 price of, 96
oil in water (O/​W) emulsions, 229 production/​market, 95
Proteins uses, 97
amaranth flour, techno-​functional worldwide worldwide, 97
properties of, 222–​223 Quinoa (Chenopodium quinoa), 13, 27, 55,
hydrolysis, 372 58–​63, 120–​122, 163, 260, 264, 304,
maize kernel, 140–​142 305, 307, 342, 355, 420
Provitamin A, 314 breeding, limitations, 9
Pseudo-​cereals, 140, 266 future potential, 9–​10
amaranth, structure of, 136 global demand, 10
inclusion, 264 Latin-​American crops, 7–​10
kernels, 148 origin of, 8–​9
starch percentage of, 143 Quinoa (Vikinga cv.), 8
Puebla, 403 Quinoa abiotic stress, 61
Pukara, 5 Quinoa boom, 406–​407, 410
Purple corn, 350 Quinoa dish, 288
Purple maize 2, 106–​107, 134, 140, 285, Quinoa flours
317, 323, 324, 344 classification of, 164
to semolina, 274
Q Quinoa, foods, 265
Quinoa gelada de foiegras d´anec amb
Quinoa consomé, 287
amino acid profile of, 312 Quinoa grains
baking properties of, 169 Amarilla Sacaca variety, 308
commercial gluten free foods, 173 germinated quinoa flours, 384
commercial production, 8 thin layer chromatography (TLC),
genetic diversity of, 61 381
milling Quinoa hydrolysates
dry, 208 antidiabetic action of, 391
wet, 209–​210 antihemolytic activity of, 390
oil extraction Quinoa production worldwide 2010–​2019,
cold, 209 97
solvent, 209 Quinoa proteins, 141, 211
protein composition, 210 antioxidant activity of, 388
protein digestibility of, 381 hydrolysates, immunomodulatory
protein/​fiber isolation, 210 properties of, 392
fiber extraction methods, 210–​211 Quinoa, starch granules of, 212
starch, physicochemical/​thermal/​ Quinoa-​tempeh–​fermented, with
rheological properties of, 212–​216 Rhizopus oligosporus–​powder, 272
techno-​functional properties, 211 Quinoa white flour, 208, 270
use of, 169 Quinoa worldwide production 2010–​2019,
in vitro research, 356 96
Index 439

Quinolizidine alkaloids (QAs), 18 in bakery product formulations, 268


Quinones, 137 decapsulation/​seed conditioning, 233
future potential, 32
germination (see Germination)
R
gluten-​free products, 183–​185
Rapid-​visco analyser units (RVU), 384 isolates/​protein hydrolysates, 234–​235
Ras protein, 350 limitations/​breeding opportunities,
RCSB PDB database, 392 31–​32
Regional socioeconomic impact, 402–​422 lipids, 180
Residual flours morphology/​phylogeny/​distribution,
Andean lupine (Lupinus mutabilis 29–​30
Sweet), 236–​238 oil extraction processes, 233
chia seeds, oil-​extracting processes of, conventional processes, 233–​234
228 non-​conventional processes, 234
mucilage, removal of, 228 residual flour, 234
proteins/​soluble fiber, techno-​ oilseed crop, 28
functional properties plant/​fruit, 29
absorption/​adsorption capacities, processing methods/​industrial
232 applications, 180–​181
emulsifying properties, 232 residual flour, 234
foaming capacity, 232 seeds, 143
foam stability, 232 techno-​functional properties, 235
oil holding capacity (OHC), 232 uses, 30–​31
organic molecules, absorption Sacsa, 134
capacity of, 232 Salba-​chia (Salvia hispanica L.) treatment,
water-​holding capacity, 232 345
sacha inchi (Plukenetia volubilis L.), 234 Salcedo-​INIA, 61
Reversed-​phase-​high performance liquid Salinity-​sensitive genotypes, 67
chromatography (RP-​HPLC), 387 Salmonella enterica, 391
Rhizopus oligosporus, 376 Salt resistance, 58
Rice Salvia hispanica, 4, 22, 24, 55, 59, 70, 100,
breads, 189 122, 138, 304, 307, 342, 345
Oriza sativa, 282 Saponins, 163, 319, 359, 384
Oryza sativa L., 355 Satiety efficiency indices (SEI), 346
ovary fertilization, 57 Seed coat, 133, 137
Roasted seeds, 182 Seed-​filling stage, photo-​thermal, 62
‘Romano’ cultivar, 67 Seed quality
Rosemary (Rosmarinus officinalis L. syn. abiotic stresses, 55
Salvia rosmarinus Schleid.), 24 climate change
drought, 56
high temperatures, 56–​57
S
salinity, 57–​58
Sacha inchi (Plukenetia volubilis L.), 59, crop production, under stress
105–​106, 120, 122, 138, 233, 260, conditions, 76
281, 304, 309, 314 crop productivity, genotype/​
agro-​industrial waste of, 375 environment interaction, 58
baked goods application, 183 Amaranthus ssp., 63–​65
440 Index

Chenopodium pallidicaule, 58–​63 Suzuki MT mill, 166


Chenopodium quinoa, 58–​63 SYBYL-​X 2.1.1-​Tripos, 393
Chia (Salvia hispanica L.), 70–​73 Symbiotic nitrogen fixation, 66
family Euphorbiaceae (Plukenetia
volubilis L.), 73 T
Lupinus mutabilis, 74–​75
maize (Zea mays), 68–​70 Tahuantinsuyo, 5
Phaseolus vulgaris L. (beans), 65–​68 Tannins, 319, 325
in Latin-​American crops, 53–​76 Targeting Induced Local Lesions IN
Seed ripening, non-​uniform, 60 Genomes (TILLING), 7
Seeds/​grains fractionation, 208 Tarwi (Lupinus mutabilis), 27, 55, 104–​105,
Serine-​type protease inhibitors (STPIs), 163, 260, 304, 308, 342; see also
350 Andean lupine (Lupinus mutabilis
Short-​chain fatty acids (SCFAs), 358 Sweet)
Simple sequence repeat (SSRs) marker, 72 in bakery product formulations, 268
Single nucleotide polymorphisms (SNPs), commercial products of, 176
70, 72 nutritional composition, 308
Single-​sequenced repeat (SSR), 16 wide genetic diversity of, 104
Sistema Integral de Normas y Evaluación de Tarwi flour, 276
la Conformidad (SINEC), 419 Tarwi oil, 168
Small-​seeded wild type (P.vulgaris var. Tarwi seeds, 60, 167, 236, 311
aborigineus (Burk.) Baudet), 126 debittering process of, 167
Snacks bars (SB), 276 products obtained, 237
Sodium dodecyl sulfate (SDS), 164 Techno-​functional properties, quinoa/​
Soil salinity, 14 kañiwa, 211
Soluble dietary fiber (SDF), 144, 242 Teosinte (Z. mays ssp.), 68
Solvent retention capacity (SRC), 222 Tert-​Butyl hydroperoxide (TBHP) damage,
Soxhlet type devices, 209, 234 389
Soya beans, 269 Tetrahydrorombifoline, 74
Soybeans (Glycine max (L.) Merr.), 14 Thin layer chromatography (TLC), 381
Sprouting seeds, 95 Threonine, 314
Sprouts, nutritional analysis of, 97 Titicaca, 61
Squash (Cucurbita pepo L.), 26 Titicaca Lake basin, 12
Starch Tiwanaku culture, 12
gluten matrix, 165 Tocopherols, 228, 314, 374, 383
granule susceptibility, 224 Total dietary fiber (TDF), 242
quinoa/​kañiwa, physicochemical/​ Trade Agreement, 420
thermal/​rheological properties of, Treaty Framework on Plant Genetic
212–​216 Resources for Food and
Streptococcus thermophilus, 280 Agriculture (TRFGAA), 20
Suberin, 137 Triglycerides, 220
Succinylation, 224 Trypsin, 222
Supercritical carbon dioxide extraction, Trypsin inhibitory activity (TIA), 242, 319
227 Tryptophan, 142, 314
Supercritical CO2 extraction (CO2-​SE), 226 beans, 66
Superfoods, 92, 94 maize protein, 311
Index 441

U Wheat
botanical classification, 124–​125
Underexploited Tropical Plants with
bread, quality of, 168
Promising Economic Value, 98
chemical composition of, 145–​147
UNESCO (2003), 283
flour, germinated/​non-​germinated
UNICEF, 281
seeds, 99
UniProtKB database, 392
Whey fraction, 237
United States Department of Agriculture
White maize, 319
(USDA), 64
White spots (Piscorunto), 126, 134
Unsaturated fatty acids, 242
Whole seed flours, 162
World Health Organization (WHO), 92,
V 345
World Trade Organization (WTO), 409
Valine, 314
Vegetable protein, 229
Vitamin D deficiency, 355 X
Vitamin E, 314, 316 Xanthan gum, 190

W Y
Water absorption, 242 Yellow maize, 319
Water imbibing capacity (WIC), 222 Yield factor (YF), 321
Water-​soluble vitamins, 314
Water stress, 75
Z
Weissella cibaria, 375
Weissella cibaria MG1, 378 Zea mays ssp. parviglumis, 26, 120
Wet processing procedures, 325 Zinc-​enriched maize program, 27

You might also like